Download View PDF - ADAN database

Document related concepts

Biochemical cascade wikipedia , lookup

Metabolism wikipedia , lookup

Enzyme wikipedia , lookup

Plant virus wikipedia , lookup

Artificial gene synthesis wikipedia , lookup

Plant nutrition wikipedia , lookup

Biosynthesis wikipedia , lookup

Amino acid synthesis wikipedia , lookup

Plant breeding wikipedia , lookup

Transcript
Progress in
Lipid Research
Progress in Lipid Research 44 (2005) 357–429
www.elsevier.com/locate/plipres
Review
Biogenesis, molecular regulation and function of plant
isoprenoids
Florence Bouvier, Alain Rahier, Bilal Camara
*
Institut de Biologie Moléculaire des Plantes du CNRS (UPR2357) et Université Louis Pasteur,
12 rue du Général Zimmer, 67084 Strasbourg Cedex, France
Abstract
Isoprenoids represent the oldest class of known low molecular-mass natural products synthesized by plants. Their biogenesis in plastids, mitochondria and the endoplasmic reticulum–cytosol proceed invariably from the C5 building blocks, isopentenyl diphosphate and/or dimethylallyl diphosphate according to complex and reiterated mechanisms. Compounds derived
from the pathway exhibit a diverse spectrum of biological functions. This review centers on advances obtained in the field
based on combined use of biochemical, molecular biology and genetic approaches. The function and evolutionary implications of this metabolism are discussed in relation with seminal informations gathered from distantly but related organisms.
2005 Elsevier Ltd. All rights reserved.
Keywords Isoprenoids; Carotenoid; Tocopherol; Phylloquinone; Plastoquinone; Chlorophyll; Phytosterol; Prenyl derivatives; Enzymes;
Genes; Cellular compartmentation
Contents
1.
2.
3.
4.
5.
*
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Biogenesis of IPP and DMAPP building blocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1. Plastid and cytosolic pathways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2. Plastid–cytosol interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3. Non-dispensability of IPP isomerase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Prenyltransferases as main trunk enzymes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1. General aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2. Seminal informations gathered from FPP synthase and GGPP synthase. . . . . . . . . . . . . . .
Polyprenyl polymer biogenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1. cis-Polyprenyltransferases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2. Rubber transferases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Lower terpenoid synthases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1. Isoprenoids branching from DMAPP and GPP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2. Semio-terpene synthases and traumato-terpene synthases . . . . . . . . . . . . . . . . . . . . . . . . .
Corresponding author.
E-mail address: [email protected] (B. Camara).
0163-7827/$ - see front matter 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.plipres.2005.09.003
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
358
358
358
363
366
367
367
370
372
372
372
372
372
374
358
6.
7.
8.
9.
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
Post-squalene sterol biogenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.1. Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.2. Pathway of phytosterol synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.3. Electrophilic processes in phytosterol synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.4. Membrane-bound non-heme iron oxygenases involved in phytosterol synthesis . . . . . . . . . . .
6.5. Kinetically favored pathway in the conversion of cycloartenol to D5-sterols . . . . . . . . . . . . .
Biogenesis of mitochondrial isoprenoids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.1. Ubiquinones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.2. Isoprenoid oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Biogenesis of plastid isoprenoids branching from GGPP. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.1. Diterpene biogenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.2. Carotenoid biogenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.3. Chlorophyll biogenesis and degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.4. Tocopherol and tocotrienol biogenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.5. Phylloquinone and plastoquinone biogenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
376
376
376
378
381
388
391
391
392
393
393
395
410
414
416
417
418
418
1. Introduction
It has been estimated that 15–25% of plant genes are dedicated to plant secondary metabolism pathways [1].
The latter comprise three major class of products including isoprenoids, alkaloids and flavonoids. When considering the presence of these products among different taxa, it is apparent that the isoprenoid pathway is the
oldest acquisition. In supporting this idea, one could note that plant dihydroflavonol reductase, the first specific enzyme of anthocyanin biosynthesis has extensive homology with isoprenoid dehydrogenases [2]. Isoprenoids are compounds built up of simple or multiple C5 units and are divided into different subgroups
synthesized in different plant cell compartments including plastids, mitochondria and the endoplasmic reticulum–cytosol (Fig. 1). Their synthesis is often tightly linked to the function, differentiation state and organization of the compartment with which they are associated.
Here, we review our current understanding of plant isoprenoid biogenesis integrating biochemical, molecular and functional data. The metabolic engineering will not be covered, as this was comprehensively addressed
[3]. For information concerning plant protein prenylation, readers are referred to previous reviews [4,5].
2. Biogenesis of IPP and DMAPP building blocks
2.1. Plastid and cytosolic pathways
It is now well established that in cyanobacteria and plants, isopentenyl diphosphate (IPP) is synthesized in
the plastids through 1-deoxy-D-xylulose-5-phosphate (DXP) pathway and in the cytosol through the mevalonic acid (MVA) pathway according to the pathways outlined in Fig. 2 (for a review, see [6–10]). It is worth
noting that alga belonging to the chlorophyta have apparently lost the mevalonic pathway and use exclusively
the DXP pathway [11]. Similarly, the malaria parasite Plasmodium and other apicomplexa synthesize their isoprenoids from the DXP pathway in the non-green plastid-type apicoplast [12]. All plant genes encoding
enzymes of the DXP and MVA pathways have been cloned and functionally analyzed. It has been observed
that the plant genes of the DXP pathway are more phylogenetically close to other bacteria than to cyanobacteria [13]. The DXP pathway generate IPP as well as dimethylallyl diphosphate (DMAPP), whereas only IPP is
released from the MVA pathways. One could note that the DXP pathway do not apparently involve phosphorylation of pathway intermediates as envisioned previously [14].
In Arabidopsis, the expression of 1-deoxy-D-xylulose-5-phosphate synthase (DXS), 1-deoxy-D-xylulose5-phosphate reductoisomerase (DXR), 2-C-methyl-D-erythritol-4-phosphate cytidyl transferase (MCT),
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
Nomenclature
AACT acetoacetyl-CoA thiolase
ABA abscisic acid
ADP adenosine 5 0 -diphosphate
ALAD 5-aminolevulinate dehydratase
AO
aldehyde oxidase
aPS
apophytoene synthase
ASC ascorbate
ASCH ascorbic acid
ATP adenosine 5 0 -triphosphate
BS
botryococcene synthase
CAO chlorophyll a oxygenase
CASE chlorophyllase
CBR chlorophyll b reductase
CCS
capsanthin-capsorubin synthase
CDP-ME 4-(cytidine 5 0 -diphospho)-2-C-methyl-D-erythritol
CDP-MEP 2-phospho-4-(cytidine 5 0 -diphospho)-2-C-methyl-D-erythritol
CHB carotenoid b-hydoxylase
CHE carotenoid e-hydoxylase
CLD chain-length determination domain
CMK 4-(cytidine 5 0 -diphospho)-2-C-methyl-D-erythritol kinase
CMP cytidine 5 0 -monophosphate
CPO coproporphyrinogen III oxidase
CPP
chrysanthemyl diphosphate
CPPS chrysanthemyl diphosphate synthase
CPS
ent-copalyl diphosphate synthase
CRTISO carotenoid isomerase
CS
chlorophyll synthase
CTP
cytidine 5 0 -triphosphate
cZ
cis-zeatin
cZR
cis-zeatin riboside
cZRMP cis-zeatin riboside monophosphate
DBMIB 2,5-dibromo-3-methyl-6-isopropyl-p-benzoquinone
DCMU 3-(3,4-dichlorophenyl)-1,1-dimethyl-urea
DCR divinyl protochlorophyllide reductase
DHA dehydroascorbate
DHAR dehydroascorbate reductase
DHNA 1,4-dihydroxy-2-naphtoate
DHNAPT DHNA phytyl transferase
DihydroGG-Chl dihydrogeranylgeranyl-chlorophyll
DihydroGGPP dihydrogeranylgeranyl diphosphate
DMAPP dimethylallyl diphosphate
DMGGBQ 2,3-dimethyl-5-geranylgeranyl benzoquinol
DMPBQ 2,3-dimethyl-5-phytyl benzoquinol
DMPQ demethylphylloquinone
DMPQMT demethylphylloquinone methyltransferase
DMNT (E)-4,8-dimethyl-1,3,7-nonatriene
DPMDC 5-diphosphomevalonate decarboxylase
DXP 1-deoxy-D-xylulose-5-phosphate
DXR 1-deoxy-D-xylulose-5-phosphate reductoisomerase
359
360
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
DXS 1-deoxy-D-xylulose-5-phosphate synthase
FARM first aspartate-rich domain
FCC fluorescent chlorophyll catabolite
FPP
farnesyl diphosphate
FPPS farnesyl diphosphate synthase
GA
gibberellin
GA12S GA12-aldehyde synthase
GA7OX GA 7-oxidase
GA13H GA 13-hydroxylase
GG-Chl geranylgeranyl-chlorophyll
GGPP geranylgeranyl diphosphate
GGPPS geranylgeranyl diphosphate synthase
GGR geranylgeranyl reductase
GPP geranyl diphosphate
GPPS geranyl diphosphate synthase
GR
glutathione reductase
GSAT glutamate-1-semialdehyde aminotransferase
GSH reduced glutathione
GSSG oxidized glutathione
GTR glutamyl-tRNA reductase
GTS glutamyl-tRNA synthetase
HBP450 heme-binding cytochrome P450
HCR hydroxychlorophyll a reductase
HDR 1-hydroxy-2-methyl-2-(E)-butenyl-4-diphosphate reductase
HDS 1-hydroxy-2-methyl-2-(E)-butenyl-4-diphosphate synthase
HGA homogentisic acid
HGGT homogentisate geranylgeranyl transferase
HMBPP 1-hydroxy-2-methyl-2-(E)-butenyl 4-phosphate
HMG-CoA 3-hydroxy-3-methylglutaryl-CoA
HMGR 3-hydroxy-3-methylglutaryl-CoA reductase
HMGS 3-hydroxy-3-methylglutaryl-CoA synthase
HPP p-hydroxyphenylpyruvate
HPPD p-hydroxyphenylpyruvate dioxygenase
HPT homogentisate-phytyl transferase
HST 3, homogentisate solanesyl transferase
iP
isopentenyladenine
IPP
isopentenyl diphosphate
IPPI
IPP-DMAPP isomerase
iPR
isopentenyladenine riboside
iPRMP isopentenyladenine riboside monophosphate
IPT
isopentenyl transferase
IS
isoprene synthase
KAO ent-kaurenoic acid oxidase
KET b-carotene ketolase
KO
ent-kaurene oxidase
KS
ent-kaurene synthase
LCB lycopene b-cyclase
LCE lycopene e-cyclase
LHC light-harvesting complex
LSU large subunit
MBS methylbutenol synthase
MCH magnesium chelatase
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
MCS 2-C-methyl-D-erythritol-2,4-cyclodiphosphate synthase
MCT 2-C-methyl-D-erythritol-4-phosphate cytidyl transferase
MDCH magnesium dechelatase
MECPP 2-C-methyl-D-erythritol-2,4-cyclodiphosphate
MEP 2-C-methyl-D-erythritol-4-phosphate
MGDG monogalactosyldiacylglycerol
MGGBQ 2-methyl-6-geranylgeranyl benzoquinol
MGGBQMT 2-methyl-6-geranylgeranyl benzoquinol methyltransferase
MK
mevalonate kinase
MPBQ 2-methyl-6-phytyl benzoquinol
MPBQMT 2-methyl-6-phytyl-1,4-benzoquinol methyltransferase
MPK mevalonate kinase
MSBQ 2-methyl-6-solaneyl-1,4-benzoquinol
MSBQMT 2-methyl-6-solaneyl-1,4-benzoquinol methyltransferase
MTC magnesium protoporphyrin IX monomethylester cyclase
MTF magnesium protoporphyrin IX methyltransferase
MVA mevalonate
NADP+ nicotinamide adenine dinucleotide phosphate, oxidized form
NADPH nicotinamide adenine dinucleotide phosphate, reduced form
NCCs non-fluorescent chlorophyll catabolites
NCED nine-cis-epoxycarotenoid dioxygenase
NHDIM non-heme-di-iron monooxygenase
NPQ non-photochemical quenching
NSY neoxanthin synthase
OSC (3S)-2,3-oxidosqualene cyclase
PAO pheophorbide a oxygenase
PBD porphobilinogen deaminase
PD
phytoene desaturase
PHB p-hydroxybenzoate
PHBPT p-hydroxybenzoate polyprenyl diphosphate transferase
Phytyl-Chl phytyl-chlorophyll
PhytylPP phytyl diphosphate
PMK 5-phosphomevalonate kinase
PPHB polyprenyl-p-hydroxybenzoate
PolyprenylPP polyprenyl diphosphate
POR NADPH-protochlorophyllide oxidoreductase
PP
diphosphate
PPPP prephytoene diphosphate
PPX
protoporphyrinogen IX oxidase
PQ
plastoquinone
PS
phytoene synthase
PSPP presqualene diphosphate
RCC red chlorophyll catabolite
RCCR RCC reductase
QTL quantitative trait loci
SAM S-adenosyl-L-methionine
SAM c-MT S-adenosyl-L-methionine c-tocopherol methyltransferase
SAM c-MT3 S-adenosyl-L-methionine c-tocotrienol methyltransferase
SARM second aspartate-rich domain
SC
squalene cyclase
SDR short-chain dehydrogenase/reductase
SMO sterol-4a-methyl-oxidase
361
362
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
SMT sterol methyltransferase
SPP
solanesyl diphosphate
SPPS solanesyl diphosphate synthase
SQS
squalene synthetase
SRPP small rubber particle protein
SSU
small subunit
TC
tocopherol cyclase
TetrahydroGG-Chl tetrahydrogeranylgeranyl-chlorophyll
TetrahydroGGPP tetrahydrogeranylgeranyl diphosphate
TMTT 4,8,12-trimethyltrideca-1,3,7,11-tetraene
TS
terpene synthase
tRNA transfer RNA
tZ
trans-zeatin
tZR
trans-zeatin riboside
tZRMP trans-zeatin riboside monophosphate
UQ
ubiquinone
UROD uroporphyrinogen III decarboxylase
UROS uroporphyrinogen III synthase
VDE violaxanthin de-epoxidase
WAF-1 11E,13E-k-11,13-diene-8a,15-diol
ZD
f-carotene desaturase
ZEP
zeaxanthin epoxidase
4-(cytidine 5 0 -diphospho)-2-C-methyl-D-erythritol kinase (CMK), 2-C-methyl-D-erythritol-2,4-cyclodiphosphate synthase (MCS), 1-hydroxy-2-methyl-2-(E)-butenyl-4-diphosphate synthase (HDS) is induced under
light conditions [15–17], while the expression of 1-hydroxy-2-methyl-2-(E)-butenyl-4-diphosphate reductase
(HDR) is constitutive [17]. Also, the expression of DXS, DXR, MCT and CMK is slightly induced in the dark
in the presence of sucrose [17]. Similarly, DXS is up regulated during chloroplast to chromoplast differentiation
in pepper fruits [18] and tomato fruits [19]. At the protein level, in Arabidopsis, MCT, DXS, HDS and DXR are
highly expressed in young photosynthetic tissues and in flower organs [16,20,21]. Interestingly, overexpression
of DXS in Arabidopsis [22], DXR in peppermint [23] or and HDR in Arabidopsis and tomato [24] led to
an increased accumulation of diverse plastid isoprenoids. Indirect evidences suggest that DXR, HDS and
HDR are potential targets of thioredoxin, and could be activated under light conditions [25,26]. In this context, it is interesting to note that Thermosynechococcus elongatus HDS has been characterized as a ferredoxindependent enzyme [27]. Thus the activity of HDS could be modulated via the light-dependent reduction of
ferredoxin.
In vivo feeding experiments pointed to the fact that the DXP pathway leading to the synthesis of isoprene is
feedback regulated at the level of DXS [28]. It has also been suggested that the accumulation of DXP pathway
enzymes was regulated by post-transcriptional mechanisms that could operate at the level of DXS [29]. It is
interesting to note that electrospray mass spectrometry combined with NMR showed that in vitro 2Cmethyl-D-erythritol-2,4-cyclodiphosphate synthase (MCS) could bind potential downstream products such
as IPP/DMAPP, geranyl diphosphate (GPP), and farnesyl diphosphate (FPP) in a ratio of 1:4:2 [30]. When
transposed to the in vivo situation these data are consonant to a possible metabolic retrocontrol of the
DXP pathway. Further testing this phenomenon could be achieved through the use of isolated chromoplasts
which have been used to analyze different sequences of the pathway [31].
Concerning the MVA pathway, previous and recent data available from Arabidopsis suggest that phytochrome and crytochrome photoreceptor could be implicated in its regulation [32,33]. In addition it has been
shown that in Arabidopsis cotyledons, 3-hydroxy-3-methylglutaryl-CoA reductase (HMGR) is not predominantly localized in the endoplasmic reticulum (ER), but within undefined spherical vesicles located in the cyto-
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
363
PRENYLATED FLAVONOIDS
CYTOSOL
MVA PATHWAY
DXP PATHWAY
ISOPRENE
PLASTID
METHYLBUTENOL
IPP
DMAPP
IPP
DMAPP
IRREGULAR TERPENES :
CHRYSANTHEMYL-PP
LAVANDULYL
CYTOKININS
HEMITERPENES
MONOTERPENES
GPP
SESQUITERPENES
FPP
GPP
MONOTERPENES
IRREGULAR TERPENES :
ANISOTOMENE
TRITERPENES
CHLOROPHYLLS
TOCOTRIENOLS
PHYTOL
GGPP
?
FARNESYL PROTEINS
GERANYLGERANYL
PROTEINS
BIOACTIVE
DITERPENES
CAROTENOIDS
GGPP
CHLOROPHYLLS
TOCOPHEROLS
PHYLLOQUINONES
SPP
?
POLYTERPENES
MVA PATHWAY
IPP
PLASTOQUINONES
UBIQUINONE
APOCAROTENOIDS
GIBBERELLINS
MITOCHONDRIA
Fig. 1. Overview of isoprenoid biosynthesis and compartmentation in plants. Abbreviations. DMAPP, dimethylallyl diphosphate, DXP,
1-deoxy-D-xylulose-5-phosphate; FPP, farnesyl diphosphate; GGPP, geranylgeranyl diphosphate; GPP, geranyl diphosphate;
IPP, isopentenyl diphosphate; MVA, mevalonate and SPP, solanesyl diphosphate.
plasm and within the central vacuole [34]. Finally, the physiological function of HMGR has been clarified
through the analysis of two hmgr mutants. Arabidopsis hmg2 mutant showed no visible phenotype while
hmg1 mutants exhibited dwarfism, early senescence and sterility [35].
2.2. Plastid–cytosol interaction
Cross-talk between the plastidial and the cytosolic pathways has been envisioned from two previous studies
using feeding experiments with stable isotopes. In one case, the biosynthesis of the diterpene ginkgolide was
analyzed using Ginkgo biloba seedlings [36]. These studies revealed that in the presence of 13C glucose, three
isoprene units were incorporated according to the cytosolic (MVA pathway) and the fourth residue via the
DXP pathway. In a similar vein, when Catharanthus roseus cell cultures were incubated with 1-deoxy-D-xylulose, it was found that part of the precursor was incorporated into sitosterol [37]. This behavior suggested a
potential exchange of isoprenoid between the plastid and the cytosol. At the other extreme, in vivo feeding of
glucose revealed that the contribution of mevalonate and the non-mevalonate pathways to the synthesis of
sitosterol was equal in green callus culture of Croton sublyratus [38]. When C. roseus cell cultures were grown
in the presence of glucose, it was shown that the plastid product phytol derives 60% from the DXP pathway
and 40% from the MVA pathway. Additional studies revealed that biosynthesis of abietane diterpenoids in
Torreya nucifera operates via the DXP and MVA [39]. Feeding experiments also indicated that in Arabidopsis
ent-kaurene and campesterol derived from 1-deoxy-D-xylulose (68–87%) and (3–7%), respectively, and from
MVA (5–8%) and (80%), respectively [40]. In Euglena the carotenoids diadinoxanthin and b-carotene were
synthesized from DXP as well as MVA, while the C20 phytyl side-chain derived exclusively from MVA
[41]. In liverworts and in hornworts the FPP portion of the phytyl chain derived from MVA, while the terminal IPP was from DXP [42–44]. Further evaluation of this cross-talk using 13C-labeled serine reveals that IPP
formed in the chloroplast was not used for the biosynthesis of cytosolic isoprenoids in the cultured cells of the
liverwort Heteroscyphus planus [44]. Also it has been shown that MVA was incorporated into the side-chain of
plastoquinone-9 (PQ-9) when tobacco BY2-cells were cultivated in the presence of fosmidomycin. On the
364
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
O
O
+
SCoA
Acetyl-CoA
CYTOSOL
SCoA
Acetyl-CoA
O
-
H
-
+
O P O
-
OH
AACT
COO
O
Glyceraldehyde
3-phosphate
O
PLASTID
O
O
Pyruvate
O
DXS
SCoA
CO2
Aceto-acetyl-CoA
OH
Acetyl-CoA
O
O
HMGS
-
O P O
HSCoA
-
OH
HOO C
DXR
NADP+
COSCoA
HMG-CoA
NADPH
O
DXP
NADPH
HO
HO
O
-
O P O
HMGR
NADP+
-
OH OH
MEP
CTP
HSCoA
HO
O
MCT
NH2
PPi
HOO C
CH2OH
N
HO
Mevalonate
O
O
O P O
AT P
-
OH OH
CDP-ME
MK
ADP
O
HOO C
O
CMK
NH2
ADP
O P O
CH2O
O
-
HOOH
ATP
HO
O
O
-
O
-
N
- P O
-
O
O
Mevalonate phosphate
O
PMK
O
O P O
-
OH OH
CDP-MEP
AT P
O
N
P O
O
N
P O
O
O
O
-
ADP
HOOH
MCS
CMP
HO
O
HOO C
O
P O P O
CH2O
O
-
-
-
O
O
-
O
O
P
O
Mevalonate diphosphate
AT P
CO2
HDS
O
O
O P O P O
O
IPP
-
O
OH OH
ME-CPP
DPMDC
ADP + Pi
O
P
O
-
-
O
O P O P O
IPPI
O
O
O
O
-
OH
HMBPP
O P O P O
-
O
MVA PATHWAY
O
O
-
HDR
O
O
-
-
-
O
IPP
O
-
O
-
O P O P O
-
DMAPP
O
-
-
O
O P O P O
IPPI
-
O
DMAPP
DXP PATHWAY
O
-
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
365
other hand, in the presence of mevilonin, the DXP contributed to the synthesis of the cytosolic sterols [45].
Using the same strategy, it was shown that the plastid–cytosol cooperation was relatively modest [46]. This
phenomenon also influences the biogenesis of lower terpenoids. In carrot leaves and roots, although monoterpenes were synthesized exclusively via the DXP pathway, the sesquiterpenes derived from the MVA as well as
DXP pathways [47]. In grape leaves, the DXP pathway was prominently used for the biogenesis of monoterpene and exceeded the MVA pathway by a factor around 100 [48]. In the epidermis of snapdragon flower, the
DXP pathway provided IPP used both for monoterpene and sesquiterpene synthesis [49].
In Arabidopsis, the albino phenotype observed in several mutants, dxr, mct, hds and hdr [9,15,17,20,50]
suggests that IPP derived from MVA could not efficiently compensate for the lack of the plastid derived
IPP. This was also illustrated by the inefficient rescue of dxs mutants [15,51] and in Nicotiana benthamiana
when the DXS-pathway specific genes were silenced [52]. Indeed, the silencing of DXR and HDR in N.
benthamiana is paralleled by chlorotic phenotypes and drastic loss of photosynthetic pigments (Fig. 3,
unpublished data). Data from isoprene-emitting plants revealed that the cross-talk between chloroplastic
and extrachloroplastic pathways of isoprenoid formation is marginal. The evidence is explained when
two facts are considered. First, isoprene synthesis consumes much more carbon precursors than other plastid isoprenoid synthesis [53]. Second, when the plastid pathway of isoprenoid was inhibited by fosmidomycin, the synthesis of isoprene remained extremely low [54–56]. Thus available data indicate that when the
plastid pathway is blocked, the amount IPP shuttled from the cytosol into the plastid could not compensate
the deficiency.
The mechanism regulating this cross-talk is unknown. It has been observed that when the cytosolic and the
plastid pathways of isoprenoid synthesis were alternatively blocked by lovastatin and fosmidomycin, there was
no correlation between the pattern of isoprenoid changes and gene expression. This led to suggest that posttranscriptional control may operate [46]. Also, because the labelling pattern varied according to the precursors
used and the nature of the products analyzed, it has been suggested that the cross-talk between plastid and the
cytosol required additional factors and could not be assimilated to a simple shuttle between the plastid and the
cytosol [57].
Besides the above unresolved questions, the question remains, how prenyl diphosphate is transported
between the plastid and the cytosol compartment. Uptake experiments using isolated plastids suggested that
the plastid envelope possesses a putative IPP transporter [58]. However, the protein responsible for this transport has not been identified. In more recent studies, prenyl diphosphate transport activities have been measured
employing liposomes containing chloroplast envelope proteins. This led to the suggestion that a unidirectional
proton symport gated by Ca2+ could be involved in the plastidial export of prenyl diphosphate to the cytosol
[59]. The transport efficiency was shown to decrease in the following order: isopentenyl diphosphate, geranyl
diphosphate, farnesyl diphosphate and dimethylallyl diphosphate. With this system, geranylgeranyl diphosphate (GGPP) was not transported [59]. The mechanism could involve a unidirectional proton symport system
[59]. In another study, in vitro experiments have been performed using isolated intact chloroplasts, liposomes
containing chloroplast envelope membrane proteins and several pathway intermediates including: DXP, 1hydroxy-2-methyl-2-(E)-butenyl 4-phosphate (HMBPP) and IPP. These studies revealed that HMBPP was
not exchanged between the cytosol and the chloroplasts while DXP was preferentially transported by the plastid
.
Fig. 2. Cytosolic MVA and plastidial DXP pathways of IPP and DMAPP synthesis in plants. [MVA (mevalonate) pathway] Enzyme
abbreviations. AACT, acetoacetyl-CoA thiolase; HMGS, 3-hydroxy-3-methylglutaryl-CoA synthase; HMGR, 3-hydroxy-3-methylglutaryl-CoA reductase; MK, mevalonate kinase; PMK, 5-phosphomevalonate kinase; DPMDC, 5-diphosphomevalonate decarboxylase;
IPPI, isopentenyl diphosphate-dimethylallyl diphosphate isomerase. Substrate abbreviations. DMAPP, dimethylallyl diphosphate; HMGCoA, 3-hydroxy-3-methylglutaryl-CoA; IPP, isopentenyl diphosphate. [DXP pathway] Enzyme abbreviations. DXS, 1-deoxy-D-xylulose-5phosphate synthase; DXR, 1-deoxy-D-xylulose-5-phosphate reductoisomerase; MCT, 2-C-methyl-D-erythritol-4-phosphate cytidyl
transferase; CMK, 4-(cytidine 5 0 -diphospho)-2-C-methyl-D-erythritol kinase; MCS, 2-C-methyl-D-erythritol-2,4-cyclodiphosphate synthase; HDS, 1-hydroxy-2-methyl-2-(E)-butenyl-4-diphosphate synthase; HDR, 1-hydroxy-2-methyl-2-(E)-butenyl-4-diphosphate reductase; IPPI, IPP-DMAPP isomerase. Substrate abbreviations. CDP-ME, 4-(cytidine 5 0 -diphospho)-2-C-methyl-D-erythritol; CDP-MEP,
2-phospho-4-(cytidine 5 0 -diphospho)-2-C-methyl-D-erythritol; DMAPP, dimethylallyl diphosphate; DXP, 1-deoxy-D-xylulose-5-phosphate;
HMBPP, 1-hydroxy-2-methyl-2-(E)-butenyl 4-phosphate; IPP, isopentenyl diphosphate; ME-CPP, 2-C-methyl-D-erythritol-2,4-cyclodiphosphate; MEP, 2-C-methyl-D-erythritol-4-phosphate.
366
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
Fig. 3. Effect of silencing DXR and HDR in Nicotiana benthamiana. Apparently the deficiency of plastid isoprenoids is not efficiently
compensated by cytosolic mevalonate precursors. A, control (untransfected plants); B, transfected plants using viral vector carrying a
fragment of 1-deoxy-D-xylulose-5-phosphate reductoisomerase (DXR) gene; C, transfected plants using viral vector carrying a fragment of
the 1-hydroxy-2-methyl-2-(E)-butenyl-4-diphosphate reductase (HDR) gene.
xylulose-5-phosphate translocator. On the other hand, IPP was not transported via already known plastid phosphate translocators. IPP was transported via a uniport mechanism that required the binding of phosphorylated
molecules on the opposite side of the membranes [60].
The complexity of the putative plastid–cytosol cooperation is paralleled in some plant tissues by an intercellular cooperation. It has been shown using Northern blot and in situ hybridization experiments, that
C. roseus, the transcripts of DXP-pathway genes, DXS, DXR MCS and geraniol 10-hydroxylase that catalyzes the first committed step in the formation of iridoid monoterpenoids were localized in the internal phloem
parenchyma. It has been hypothesized that under these conditions the intermediates are translocated sequentially from the internal phloem parenchyma to the epidermis and, finally, to laticifers and idioblasts during
monoterpenoid indole alkaloids [61].
2.3. Non-dispensability of IPP isomerase
Because the DXP pathway generates IPP as well as DMAPP (Fig. 2), the role of plastid IPP isomerase has
been questioned. Two types of isopentenyl diphosphate isomerase (IPPI) have been characterized in living
organisms. Type 1 is the enzyme that reversibly catalyzes the conversion of IPP into DMAPP in most eucaryotes and some bacteria [62]. This enzyme contains an essential atom of zinc [63]. More recently, type 2 IPPIs
have been characterized in archaea and some bacteria [64]. Type 2 is not homologous to type 1. Type 2 enzyme
requires redox co-enzymes, i.e., flavin mononucleotide (FMN) and NAD(P)H, for activity. During the catalysis, FMN is reduced by NAD(P)H [65].
In plant type organisms it has been shown previously that Synechocystis strain PCC 6803 lacks isomerase I
activity due to the absence of type-1 IPP isomerase gene [66]. In contrast ORFs homologous to type II IPP
isomerase are present in several cyanobacterial genomes and shown to be inactive [67]. However, in further
analysis the activity could be demonstrated [68]. These data revealed that IPPI is an essential gene. These
trends were supported by N. benthamiana plants in which IPPIs have been silenced. These plants had mottled
white-pale green leaves concomitantly to a drastic alteration of the plastid structure and photosynthetic pigments were decreased by 80%. Chloroplasts from silenced plants displayed a high IPP phosphatase activity
[52]. Thus there is no redundancy between the function of IPPI and HDR (Fig. 2) which independently catalyze the synthesis of DMAPP.
In tobacco, two IPPIs genes, IPPI1 and IPPI2 have been characterized [69]. IPPI1 encodes the previously
characterized plastid isoform, while IPPI2 encodes a cytosolic enzyme. The expression of IPPI1 was induced
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
367
by high-salt concentrations and high light stress. On the other hand, the expression of IPPI2 was induced
under high-salt concentrations and cold stress conditions [69]. In a similar vein, the presence of two differentially regulated IPPIs have been reported from Cinchona robusta. Suspension-cultured cells of C. robusta, the
source of pharmaceutically active alkaloids, produce large amount of anthraquinone alkaloids when challenged with phytopathogenic fungus Phytophtora cinnamomi [70]. Under these conditions, the activity of IPPI
is transiently induced. Two isoforms of IPPI (IPPI1 and IPPI2) have been characterized in C. robusta and
IPPI2 is specifically induced following the fungal elicitation. This phenomenon is paralleled by a decreased
FPP synthase (FPPS) activity [71]. These data suggest that IPPI2 is dedicated to the induced biosynthesis
of anthraquinone phytoalexins [72].
3. Prenyltransferases as main trunk enzymes
3.1. General aspects
The prenyltransferase reaction involves the condensation of an acceptor isoprenoid or non-isoprenoid molecule to an allylic diphosphate. Three types of reactions have been described [73]. The first and most studied
reaction involves a head-to-tail or 1 0 -4 condensation of the 5-carbon IPP, to DMAPP or to an elongating
allylic diphosphate chain. Depending on their specificities these prenyltransferases yield linear trans- or cisprenyl diphosphates. The second class of prenyltransferases reaction involves head-to-head or head-to-middle
condensations that lead to irregular terpenoids such as presqualene diphosphate, prephytoene diphosphate
and chrysanthemyl diphosphate which serve as pathway-specific intermediates. The third type of prenyltransferases reaction involves the alkylation of a non-isoprenoid acceptor as observed in common prenyllipids such as chlorophylls, tocopherols, ubiquinone and diverse secondary metabolites. Mechanistically, the
reaction catalyzed by prenyltransferases and terpenoid cyclases proceed according to the same mechanism.
In both cases, a new carbon–carbon bond is generated through the addition of a reactive electron-deficient
allylic carbocation to an electron-rich carbon–carbon double bond [74].
The synthesis linear polyprenyl chain is catalyzed by trans- and cis-prenyltransferases [75,76]. Prenyltransferases belonging to the trans serie are usually subdivided into two subgroups [75]. One subgroup leads
to GPP(C10), FPP(C15), GGPP(C20), and the other comprises octaprenyl diphosphate (C40), nonaprenyl
diphosphate also known as solanesyl diphosphate (SPP) (C45) and higher molecular weight homologues.
Although GPP, FPP and GGPP are the main precursors of monoterpenes, sesquiterpenes and diterpenes there
are exceptions peculiar to plants of the apiaceae, asteraceae and lamiaceae (Fig. 4). Plants belonging to these
families synthesize irregular monoterpenes, sesquiterpenes and diterpenes that derive from head-to-head coupling of DMAPP in the case of chrysanthemane and lavandulane-type (Fig. 4) [77] between DMAPP and GPP
in the case of allergenic sesquiterpene lactone, anthecotuloide [78] and between two GPPs in the case of anisotomenes (Fig. 4) [79].
In plants, FPPS and GGPP synthase (GGPPS) generate the key branchpoint intermediates of cytosolic and
plastid isoprenoids. Arabidopis FPPS [80] and pepper GGPPS [81,82] were characterized previously at the biochemical and molecular levels. Few inhibitors of these enzymes are availabale except phosphonate derivatives
or aza-GGPP [83] and the fungal product gerfelin [84]. Seven genes encoding GGPPS have been characterized
from Arabidopsis [85–87]. The in vitro activity of GGPPS isoenzymes (1,3–6) has been demonstrated [85–87].
Using green fluorescent protein (GFP) fusion protein, GGPPS1 and GGPPS3 have been localized in the chloroplasts. GGPPS2 and GGPPS4 were localized in the endoplasmic reticulum and GGPPS6 in the mitochondria [85]. GPP synthase (GPPS) has been isolated from Arabidopsis and shown to be potentially processed into
a plastid and a cytosolic isoform [88]. The cytosolic form is consonant with the fact, that in some plants monoterpene derivatives are synthesized from the MVA pathway [89]. Interestingly, when Perilla frutescens limonene synthase was routed either to the plastid or the cytoplasmic compartment of tobacco, limonene
synthesis was detectable in the resulting transgenic plants. The amount of plastid derived limonene was 3.5fold higher than that produced in the cytoplasm [90].
In peppermint, GPPS has been characterized as heterodimer composed of a large subunit (LSU) and a
small subunit (SSU) and both subunits were required for activity [91,92]. The LSU is highly homologous
and phylogenetically related to GGPPS (Fig. 5) and has no enzymatic activity per se. This scenario is
368
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
OH
OPP
MONOTERPENES
GPPS
Methylbutenol
Isoprene
DMAPP
OPP
DMAPP
IPP
GPP
OH
OH
OH
Geraniol
Linalool β-Ocimene Myrcene Limonene
α-Terpinene Sabinene Menthol α-Pinene
SESQUITERPENES
OH
FPPS
DMAPP
2 IPP
OPP
FPP
(E )-4,8-dimethyl
-1,3,7-nonatriene
(DMNT)
Nerolidol
OH
OH
O
OH
O
O
OH
O
FPP
SQS
α-Humulene β-Caryophyllene β-Cadinene α-Farnesene Germacrene B
5-epi-Aristolochene Capsidiol
Juvocimene I
Rishitin
Juvenile
hormone III
TRITERPENES
OAc
OH
O
OH
O
HO
HO
O
α−Amyrin
SQUALENE
Cucurbitacin-E
OH
GGPPS
DITERPNES
HO
HO O
OPP
DMAPP
3 IPP
GGPP
Abietadiene
Taxadiene
Carnosic acid
IRREGULAR
DMAPP
GPP
OPP
+
GPP
MONOTERPENES
OPP
DITERPENES
OPP
Geranylgeraniol
OH
O
OH
H
11E,13E-labda11,13-diene-8a,
15-diol (WAF-1)
DMAPP
Casbene
OH
O
OH
+
Kaurene
OH
H
OPP
OH
Scareolide
H
Sclareol
PPO
PPO
Lavandulyl
diphosphate
Chrysanthemyl
diphosphate
Geranyl linalool
4,8,12-trimethyltrideca
-1,3,7,11-tetraene
(TMTT)
CH2OH
OR
Anisotomene
Fig. 4. Structure of some isoprenoid derivatives branching at different steps of the pathway. Enzyme abbreviations. GPPS, geranyl
diphosphate synthase; FPPS, farnesyl diphosphate synthase; SQS, squalene synthetase; GGPPS, geranylgeranyl diphosphate synthase.
Substrate abbreviations. DMAPP, dimethylallyl diphosphate; FPP, farnesyl diphosphate; GGPP, geranylgeranyl diphosphate; GPP,
geranyl diphosphate; IPP, isopentenyl diphosphate.
reminiscent of heteromeric bacterial prenyltransferase subgroups that lead to hexaprenyldiphosphate (C30)
and heptaprenyldiphosphate (C35). These enzymes consist of two non-identical subunits that can be dissociated. None of the isolated subunit has the catalytic activity [75]. Thus further characterization of peppermint
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
Plant
GGPP synthases
GPP synthase
GPP synthases large subunit
SPP synthases
Cyanobacteria
Plant
Plant
GPP synthases
369
ScHPPS
GsHPPS SeFPPS
EcFPPS
CaGGPPS
SnSPPS
AtSPPS1
AtSPPS2
MpGPPSlsu
AmGPPSlsu
AtGGPPS
TcGGPPS
CsGPPS
AtGPPS
AgGPPS
EcOPPS
ScGGPPS
SsGGPPS
Insect GPPS
MtFPPS
IpGPPS
AfGGPPS
ScFPPS
CbGPPSssu
OsFPPS
AmGPPSssu
CaFPPS
Aa2FPPS
Aa1FPPS
Atr1FPPS
Atr2FPPS
AtFPPS
MpGPPSssu
Plant
GPP synthases
small subunit
HsFPPS CcCPPS
Atr5FPPS
HbcisPT
0.1
AtcisPT
Plant
FPP synthases
Plant irregular
terpene synthases
Plant
cis-prenyltransferases
Fig. 5. Unrooted phylogenetic tree of prenyl diphosphate synthases and related homologues. Alignments were performed with ClustalX
and visualized with TreeView. Distances between sequences are expressed as 0.1 changes per amino acid residue. Accession Numbers and
abbreviations. CaGGPPS (CAA56554.1); MpGPPSlsu (AF182828_1); AmGPPSlsu (AAS82860.1); AtGGPPS (NP_195399); TcGGPPS
(AAD16018.1); AgGPPS (AAN01133.1); ScGGPPS (AAA83262.1); IpGPPS (AAX55632.1); ScFPPS (P08524); OsFPPS (T03687);
CaFPPS (CAA59170.1); Aa2FPPS (AAD32648.1); Aa1FPPS (JC4846); Atr1FPPS (AAP74720.1); Atr2FPPS (AAP74719.1); AtFPPS
(AAB07248.1); Atr5FPPS (AAP74721.1); CcCPPS (I13995); HsFPPS (P14324); AtcisPT (AAF22257); HbcisPT (BAB83522);
MpGPPSssu (AF182827_1); AmGPPSssu (AAS82859.1); CbGPPSssu (AAS82870.1); AfGGPPS (F69535); MtFPPS (O26156); SsGGPPS
(P95999); EcOPPS (AAC76219.1); AtGPPS, (CAC16849.1); CsGPPS (CAC16851.1); AtSPPS2 (BAD88534.1); AtSPPS1 (BAD88533.1);
SnSPPS (BAA16579); ScHPPS (AAA34686.1); GsHPPS (BAA08725.1); EcFPPS (P22939); SeFPPS (BAA82614). Aa, Artemisia annua;
Af, Archaeglobus fulgidus; Ag, Abies grandis; Am, Antirrhinum majus; At, Arabidopsis thaliana; Ca, Capsicum annuum; Cb, Clarkia breweri;
Cc, Chrysanthemum cinerariaefolium; Cs, Citrus sinensis; Ec, Escherichia coli; Gs, Geobacillus stearothermophilus; Hb, Hevea brasiliensis;
Hs, Homo sapiens; Ip, Ips pini; Mp, Mentha piperita; Mt, Methanobacterium thermoautotrophicum; Os, Oryza sativa; Sc, Saccharomyces
cerevisiae; Se, Synechococcus elongatus; Sn, Synechocystis sp.; Ss, Sulfolobus solfataricus; Tc, Taxus canadensis.
GGPPS will be necessary to substantiate the model. In contrast, the GPPS of Abies grandis (AgGPPS,
AF513112) was characterized as a homodimer [93] and clusters with GGPPS (Fig. 5). In Clarkia and Anthirrhinum flowers, a functional GGPPS has been characterized in parallel with two proteins sharing 53% identity
with peppermint GPPS-SSU [94]. When Clarkia and Anthirrhinum GGPPS and GPPS-SSU were incubated
together, GPP was synthesized. Further informations concerning SPP synthases (SPPS), cis-polyprenyl synthases, irregular terpene synthases and prenyltransferases alkylating non-isoprenoid acceptors are reported
later in appropriate sections.
370
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
3.2. Seminal informations gathered from FPP synthase and GGPP synthase
3.2.1. Enzyme structure and classification
Crystallographic studies using avian FPP synthase as a prototype reveal that trans-prenyl transferases are entirely composed of a-helices that form a large central cavity. The cavity consists of 10 antiparallel a-helices containing conserved residues [95]. Unexpectedly, squalene-hopane cyclase, pentalenene
synthase and protein farnesyl transferase have the same fold, thus leading to the concept of ‘‘terpenoid
synthase fold’’ [96]. Indeed these 3D data are reminiscent of previous studies demonstrating that under
constrained conditions, avian liver FPP synthase can in addition to FPP produce cyclic sesquiterpenes
[97].
trans-Prenyltransferases display several conserved regions and in particular two aspartate-rich domains
characterized as the first aspartate-rich motif (FARM) and the second aspartate-rich motif (SARM) which
is strongly conserved (Fig. 6). The FARM and the SARM are placed on distinct a-helices and are positioned
on opposite walls of the central cavity of FPPS. As might be inferred from previous studies using avian FPPS,
in trans-prenyltransferases, the pyrophosphate group of the allylic substrate is bound with Mg2+ chelated by
the FARM motif and the pyrophosphate of IPP, the homoallylic substrate is bound with Mg2+ coordinated
by the SARM [95,98]. Based on the FARM and SARM motifs, trans-prenyltransferases could be grouped into
several types.
Type I FPPSs also termed (eukaryotic) have a FARM motif (DDXXD) and a two aromatic amino acid
residues at the fourth and fifth positions upstream from the FARM. On the other hand, type II FPPSs also
termed (eubacterial FPPS) have a four amino acid insertion in the FARM DDXXXXD and a single aromatic
amino acid in the fifth position upstream from the FARM. A third FPPS group has been functionally identified recently in Mycobacterium tuberculosis [99]. In the latter case, the FARM motif has a DDXXD sequence
but deviates from type I because only one aromatic amino acid is at the fifth position. The mycobacterial FPP
is more closely related to archaeal type I GGPPS.
In the case of type I GGPPS also termed archaeal GGPPS, only two amino acids are inserted in the
FARM and one aromatic amino acid residue at the fifth position upstream from the FARM. Plant GGPPS
like eubacterial GGPPS belong to type II GGPPS. In these enzymes, four amino acids are inserted in the
Fig. 6. Structure of prenyl diphosphate synthase motif regulating the chain-length of final products. The first aspartate-rich motif
(FARM) and the second aspartate-rich motif (SARM) specifying the chain-length is indicated for the different prenyltransferases.
Abbreviations. GPPS (LSU), geranyl diphosphate synthase large subunit; FPPS, farnesyl diphosphate synthase; GGPPS, geranylgeranyl
phosphate synthase and SPPS, solanesyl diphosphate synthase.
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
371
FARM, like type II FPPS. Two of these inserted amino acids have been shown to be important for the
catalytic activity [100]. Type II GGPPS is devoid of two aromatic amino acid residues at the fourth and
fifth positions upstream from the FARM. Type III GGPPS (eukaryotic except plants) contains only two
inserted amino acids like type I GGPPS. But unlike type I GGPPS, one of the inserted residues in type
III GGPPS is glutamic acid. Type III GGPPS is also devoid of the characteristic aromatic upstream to
the FARM. In addition, it has been shown that this class of prenyltransferases did not accept DMAPP
as a substrate [101].
3.2.2. Mechanism of chain-length control
Data based on crystallographic studies [95,98,102–104] and site-directed mutagenesis of FPPS established that the product chain-length was mainly governed by amino acid residues located in the fourth
and fifth positions before the FARM. Thus, this region and the FARM have been considered as the
chain-length determination domain (CLD) of trans-prenyltransferases. Basically, when a mutation is introduced at these positions, the chain-length of the prenyl diphosphate resulting from the catalytic activity is
inversely proportional to the steric hindrance of the introduced amino acid residue. For instance, the
mutants F112A and F113S that introduce small residues in the FARM of avian FPPS led to FPPS
mutants that produced, respectively, GGPP and or geranylfarnesyl diphosphate [98]. Further shortening
of the active cavity of avian FPPS by introducing the mutation A116W reduced the binding of GPP
to the allylic site [105]. Also, a mutation S113F of Bacillus stearothermophilus, a position equivalent to
F113 in avian FPPS, resulted in the following distribution of products: GPP/FPP: 27:1 [106]. Interestingly, the first animal GPPS involved in pheromones biosynthesis has been isolated from the bark beetle
Ips pini [107]. This enzyme produced almost exclusively GPP. Although this enzyme is structurally related
to avian FPPS [95], its FARM structure possesses F and Y residues at the fourth and fifth positions,
consistent with the fact that relatively bulky residues play an important role in shortening the prenyl
chain-length. In the same line of reasoning, when bulky residues, i.e.; M135Y/S136F were introduced
to the FARM of GGPPS of the diterpene producing plant Scoparia dulcis, the mutant GGPPS lost
the ability to use FPP as an allylic substrate. Under these conditions, DMAPP was accepted and led
to 3% GGPP while FPP and GPP represented 55.3% and 9.3% of the final products [108]. This trend
was also verified in the case of Arabidopsis thaliana SPPSs (AtSPPSs). The fourth and the fifth regions
upstream from the FARM of AtSPPS1 and AtSPPS2 do not possess aromatic residues [109]. However
additional amino acid residues play a key role in determining the chain-length. When the P143 and
the C144 residues were deleted in the FARM domain of S. dulcis, the percentage of the products were:
GGPP, 1%; FPP, 20.9%; and GPP, 5.5% [108]. The difficulty in predicting the chain-length was also illustrated by the fact that unlike the plants GGPPS, archebacterial GGPPS have an aromatic residue at the
fifth position upstream from FARM and have only two amino acid residues inserted in the FARM,
instead of four like Arabidopsis GGPPS. Also, it has been noted that the chain-length was changed when
mutations were introduced in GGPPS, outside of the CLD domain and before the conserved G(Q/E)
motif [110,111].
Alternatively, the ratio between the concentration of IPP and DMAPP could alter the product specificity of
FPPS and GGPPS. At higher DMAPP concentration, FPPS produces increased amount of GPP [100]. In a
similar vein, at high DMAPP concentrations, pepper chromoplast GGPPS produced significant amount of
GPP [88]. Interestingly, it has been shown that the concentration of plastid DMAPP could reach millimolar
value in isoprene- and methylbutenol-emitting and non-emitting plants [112]. Also, the DMAPP concentration
was light-dependent [112] and could reach concentrations similar to that of other plastid metabolites such as 3phosphoglycerate, dihydroxyacetone phosphate and fructose1,6-bisphosphates [113,114]. Under these conditions, the allylic substrate could compete with nascent polyprenyl chain and thus induces the abortion of the
condensation reaction and the subsequent release of GPP. Alternatively, it has been shown that when the rate
of the allylic carbocation prone to condensation was accelerated by Ca2+ and Mn2+, the formation of long
prenyl diphosphate was favored [115]. Finally, it is worth noting that irregular terpene synthases that cluster
close to FPPSs (Fig. 6) could catalyze the synthesis of prenyl diphosphate having different chain-lengths
[116,117].
372
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
4. Polyprenyl polymer biogenesis
4.1. cis-Polyprenyltransferases
In eucaryotes, cis-prenyltransferases are involved in the synthesis of C55–C100 dolichols, which serve as
sugar carriers in protein glycosylation [118,119]. A dehydrodolichyl diphosphate synthase has been characterized from Arabidopsis [120]. The Arabidopsis gene was mainly expressed in roots [120]. In general, these prenyl
transferases are activated by lipids and detergents. The peptide sequence of these enzymes do not have the
most conservative FARM and the SARM that are characteristic of trans-type prenyltransferases, although
they require Mg2+ for activity. Data gathered from Escherichia coli undecaprenyl diphosphate synthase
revealed that the active-site is organized differently from trans-prenyl transferase. In cis-prenyltransferases,
the active-site is organized as a hydrophobic tunnel composed of a-helices and b-strands [121]. Recent crystallographic data using undecaprenyl diphosphate synthase revealed that no metal was detected in the FPP
active-site in complex with FPP. Only the binding of IPP and the condensation reaction require Mg2+.
FPP is bound to positively charged arginine residues and the hydrocarbon moiety interacts with hydrophobic
amino acids [122,123].
4.2. Rubber transferases
Despite the fact that more than 2000 plant species produce natural rubber, a cis-1,4-polyisoprenoid,
Hevea brasiliensis is the only commercial source of rubber [124]. In this plant, rubber prenyl-transferase synthesizes a polymer containing thousands of IPP units. A small rubber particle protein (SRPP) gene has been
cloned from H. brasiliensis [125]. SRPP is homologous to the rubber elongation factor previously characterized from H. brasiliensis rubber particles [126] and to a protein recently characterized from Guayule (Parthenium argentatum) [127]. Like the SRPP, the guayule homologue increased the incorporation of IPP
into high molecular weight rubbers [127]. Search of the protein databases revealed that the deduced peptide
sequences of the above depicted proteins have up to 47% identity to stress-related proteins from Arabidopsis
and Phaseolus vulgaris, which apparently do not produce rubber [127]. Thus the specific role of such protein
in rubber biosynthesis requires further investigations. Subsequent studies indicated that Arabidopsis possesses
a cis-prenyltransferase that catalyzed the formation of polyprenyl chains ranging number from C100 to C130
[128]. Because rubber prenyl-transferase synthesizes a polymer containing thousands of IPP units, the Arabidopsis cis-prenyltransferase alone is not competent for rubber synthesis. Recently, a cDNA coding for a
novel cis-prenyl chain elongating enzyme has been identified from H. brasiliensis. The corresponding gene
is expressed predominantly in the latex compared with other Hevea tissues examined. The recombinant protein catalyzed the cis-1,4-polymerization of isoprene units and yielded a polyprenyl chain of 2 · 103–1 · 104
Da. The highest polymerization activity required the addition of soluble factors present in the washed particles [129].
5. Lower terpenoid synthases
5.1. Isoprenoids branching from DMAPP and GPP
5.1.1. Isoprene and methylbutenol
Several plants belonging to mosses, ferns, gymnosperms and angiosperms emit isoprene and methylbutenol
(for a review, see [130]; Fig. 4). The reactions are catalyzed by isoprene synthase (IS) and methylbutenol synthase (MBS), two plastid enzymes that convert DMAPP to isoprene and methylbutenol [130]. In contrast to
several plastid biosynthetic enzymes, IS and MBS possess a relatively high KmDMAPP. The gene encoding IS
has been characterized from poplar species [131–133]. IS display significant identity and strong similarity to
monoterpene synthases and particularly limonene synthases. This trend is reinforced by the fact that IS has
a marginal limonene synthase activity [131]. The expression of IS is induced by heat stress and light, but
not by jasmonic and salicylic acids [133]. Although Arabidopsis does not emit isoprene, IS has extended
homology to Arabidopsis myrcene/ocimene synthase. One key difference between IS and monoterpene syn-
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
373
thase is the exclusive presence of two phenylalanine residues in IS active-site that could narrow and limit the
interaction with longer allylic diphosphates [132].
5.1.2. Cytokinins
The two physiologically active plant cytokins, trans-zeatin and isopentenyladenine and the less active ciszeatin are prenylated by DMAPP (Fig. 7) [134–136]. The reaction is catalyzed by specific prenyltransferases. In
Arabidopsis, four plastid adenosine phosphate-isopentenyl transferases divert the DXP-derived DMAPP to
the synthesis of trans-zeatin, whereas cis-zeatin is prenylated using MVA-derived DMAPP by a cytosolic prenyltransferase [137] (Fig. 7).
Overexpression of Arabidopsis FPPS1 isoform in Arabidopsis did not modify the sterol content but induced
premature cell death and senescence-like phenotype in transgenic plants compared to the control. This phenotype has been attributed to reduced levels of endogenous zeatin-type cytokinins in leaves, due to reduced
availability of IPP and DMAPP for cytokinin biosynthesis [138]. A quantitative trait loci (QTL) approach
has been used to characterize the gene encoding cytokinin oxidase/dehydrogenase (OsCKX2) from rice
[139]. OsCKX2 degraded cytokinins by cleavage of the unsaturated N6-isoprenoid side chains [136]. In rice,
when the expression of OsCKX2 was downregulated, the number of reproductive organs and the production
of grains were enhanced [139].
5.1.3. Phenylpropanoid derivatives
Several flavonoids are prenylated, for a review see [140]. Feeding experiments revealed that prenylated
flavanones from Glycyrrhiza glabra [141] and Sophora flavescens [142], humulone in Humulus lupulus [143],
CYTOSOL
PLASTID
MVA
PATHWAY
OPP
OPP
IPP
IPP
DXP
PATHWAY
OPP
+ ATP
OPP
DMAPP
+ tRNA
DMAPP
IPT
IPT
iPRMP
tZRMP
cZRMP
iPRMP
iPR
tZR
cZR
iPR
OH
OH
HN
HN
N
N
N
H
N
iP
N
N
N
H
N
tZ
HN
HN
N
N
N
cZ
N
H
N
N
N
H
N
iP
Fig. 7. Cytokinin biosynthetic pathways. Enzyme abbreviations. IPT, isopentenyltransferase. Substrate abbreviations. ATP, adenosine 5 0 triphosphate; cZ, cis-zeatin; cZR, cis-zeatin riboside; cZRMP, cis-zeatin riboside monophosphate; DMAPP, dimethylallyl diphosphate;
DXP, 1-deoxy-D-xylulose-5-phosphate; IPP, isopentenyl diphosphate; iP, isopentenyladenine; iPR, isopentenyladenine riboside; iPRMP,
isopentenyladenine riboside monophosphate; MVA, mevalonate; tRNA transfer RNA; tZ, trans-zeatin; tZR, trans-zeatin riboside;
tZRMP, trans-zeatin riboside monophosphate.
374
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
furanocoumarins in Apium graveolens [144] and the prenyl group of several isoprenyl phenyl ethers [145] derive
from the plastidial DXP pathway. These data contrast with the fact that in shikonin, the geranyl group derive
from the MVA pathway [89]. Because the flavanone ring is formed in the extraplastidial compartment, this
suggests that a tight cooperation could exist between plastid and ER–cytosol. A very similar mechanism could
operate for glyceollin and phaseollin, two prenylated isoflavonoid phytoalexins produced by elicited soybean
and bean plants [146]. To facilitate the connexion between these spatially distinct compartments, it has been
reported that an elicitor inducible flavonoid–DMAPP transferase is associated with the chloroplast membrane
envelope [146].
5.2. Semio-terpene synthases and traumato-terpene synthases
5.2.1. General characteristics
The function lower terpene synthases is prominently dedicaded to the synthesis of volatile and non-volatile
chemicals used for communication and defense responses (Fig. 4). Thus they are referred as semio-terpene synthases and traumato-terpene synthases from the Greek root ‘‘semion’’ and ‘‘traumatos’’ meaning, respectively,
mark or sign and wound or injury. Operationally these enzymes function as monoterpene, sesquiterpene and
diterpene synthases. In earliest studies, monoterpene synthases were found in a soluble fraction and the role of
plastids was challenged. On the other hand, sesquiterpene synthase and diterpene synthase were assigned,
respectively, to the cytosolic and the plastid compartments (for a review, see [147]). Our appreciation of this
compartmentation has changed through the use of genetic, molecular biology and more refined biochemical
techniques. The plastid is now recognized as a main site of the initial steps monoterpene synthesis in lower and
higher plants, whether or not they differentiate resin ducts, secretory pockets or trichome structures
[88,148,149].
Mechanistically these enzymes proceed via several transient steps including: ionization of the diphosphate
moiety with the aid of divalent cations, rearrangement of the resulting carbocation, deprotonation or nucleophile capture. The process is basically similar to that of prenyltransferases. Although endogenous phosphatase could yield terpenols [150], the synthesis of geraniol is catalyzed by specific monoterpene synthases
designated geraniol synthase [151,152]. Thus the biogenesis of geraniol like that of linalool and methybutenol
is based on hydration of transient allylic carbocation. Detailed aspects concerning the biosynthesis and the
molecular biology of semio-terpene synthases and traumato-terpene synthases have been largely reviewed
[153–155].
The number of semio-terpene synthases and traumato-terpene synthases in nature is presently unknown.
However, studies from known examples indicate that multiple products could be generated by the same
enzyme. For instance, (trans)-b-farnesene synthase, a sequiterpenoid cyclase, produced 5% of d-cadinene
and 8% of (cis)-b-farnesene in addition to (trans)-b-farnesene [156]. At the other extreme, the d-selinene
and c-humulene synthases from grand fir generated at least 34 and 52 products, respectively [157]. In the genome of Arabidopsis 32 putative members could be identified [158]. Available data indicate that most of these
genes were almost expressed in the flower organs [159]. Recent data indicate that only two sesquiterpene synthases catalyze the 20 sesquiterpene volatiles produced in Arabidopsis flowers [160].
Despite the fact that diverse sesquiterpene cyclases, FPPS and squalene synthetase (SQS) lack significant
amino acid sequence identity, their X-ray structures reveal that they adopt the terpenoid synthase fold. The
presently available crystal structure reveal that they have the same three-dimensional structures and differ
by the specific arrangements of the amino acid side chains in the active-site [96,161–163]. It has been suggested
that starting conformation of FPP in the active-site coupled to limited changes in the active site residues is a
critical determinant of product diversity [164,165]. Thus mutant aristolochene synthase Y92A, a sesquiterpene
synthase from Penicillium roqueforti, produced mostly the linear product (trans)-b-farnesene (72.6%), due to
the increased freedom of FPP and its derived transient intermediates [166].
5.2.2. Pathway induction by biotic and abiotic factors and biological roles
In addition to being implicated in the discrimination between host and non-host plants for pollinisation,
volatile isoprenoids are implicated in indirect defense. Most of the volatile emitted isoprenoids belong to
the monoterpene and sesquiterpene series but homoterpenes such as (3E)-4,8-dimethyl-1,3,7-nonatriene
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
375
(DMNT) and 4,8,12-trimethyltrideca-1,3,7,11-tetraene (TMTT) are also synthesized, respectively, from FPP
and GGPP (Fig. 4) [167]. The latter were synthesized via a cleavage pathway including the action of cytochrome P450 oxygenases [168,169]. Interestingly, feeding experiments revealed that under these stressing conditions, DMNT and TMTT were flexibly synthesized according to the mevalonate and the non-mevalonate
pathways [167]. Volatile isoprenoids emitted by herbivore infested plants serve as attractants for herbivore
predators or parasites of plants and provide indirect defense against herbivores (for a review, see [170–
174]). Their biosynthesis and emission are highly regulated. In brussels sprouts, the same volatile isoprenoid
blend was emitted following wounding and herbivore feeding [175]. In maize, soybean, lima bean, cotton and
cucumber, isoprenoid volatiles emitted during wounding were different from those due to herbivore feeding
[173]. Among these monoterpenes, b-ocimene plays a key role because of its specific emission pattern and wide
spread roles following herbivore feeding (indirect defense), resistance to pathogens [176] and tolerance to high
temperature [177]. These molecules could potentiate defense response as shown by the fact that in uninfested
lima bean leaves exposed to DMNT, TMTT and b-ocimene, the expression of several defence genes and FPPS
were induced [178].
In the Arabidopsis genome, 40 semio-terpene synthases and traumato-synthase-like genes have been identified by genome in silico analysis [158,179]. Among these A. thaliana terpene synthases, (AtTS3) and
(AtTS10) have been characterized at the molecular level. AtTS10 encodes a monoterpene synthase that catalyzes mainly the synthesis of myrcene (56%) and diverse monoterpenes including (trans)-b-ocimene (20%)
[180], while AtTS3 is almost dedicated to the synthesis of (trans)-b-ocimene (94%) and minor terpene including (cis)-b-ocimene (4%) and myrcene (2%) [181]. Arabidopsis infested by Pieris rapae emitted several volatiles
that serve to attract its parasitic predator Cotesia rubecula [182]. Under these conditions, the expression of
AtTS3 and AtTS10 was induced but only myrcene was emitted in the atmosphere [182]. On the other hand,
in response to jasmonic acid and mechanical wounding, the expression AtTS3 was induced in parallel to the
emission of (trans)-b-ocimene [181]. This indicates that the expression of AtTS was finely tuned according to
a complex and unknown mechanism. Furthermore, this phenomenon was regulated by a diurnal rhythm
[183].
The accumulation of semio-isoprenoids and traumato-isoprenoids is often induced in stressed plants (for a
review, see [184]). This phenomenon is often paralleled by pathway redirections as observed in diverse plants
including tobacco [185,186], potato [187], pepper [188] and cell-suspension cultures of Tabernaemontana
divaricata [189] and Ammi majus [190].
Finally, the secretion of oleoresin [a mixture of monoterpene olefins (turpentine) and diterpene resin acids
(rosin)] represents the primary response of conifers to wounding. Subsequent volatilization of turpentine
results in crystallization and polymerization of the nonvolatile resin acid fraction and subsequent sealing of
the wound site [191].
The molecular control of semio- and traumato-terpene synthases is poorly understood. The pathway is regulated by light through the action of phytochrome, at least in thyme seedlings [192] and Satureja douglasii
[193]. Calcium influxes, in combination with phosphorylation dephosphorylation cascades have also been
implicated [178]. The first transcription factor implicated in monoterpenoid metabolism was termed ORCA.
ORCA is a member of the AP2/ERF-domain family of plant transcription factors. ORCA regulates the biogenesis of terpenoid indole alkaloids [194]. Detailed work has also been performed using cotton. Sesquiterpene
aldehydes including gossypol are common phytoalexins of cotton. Their formation through the (+)-d-cadinene skeleton is developmentally and spatially regulated [195,196]. (+)-d-Cadinene synthase (CAD) is encoded
by a gene family divided into two subgroups that include CAD1-A represented by a single member and CAD1C which comprises several members. Members of the CAD1-C are highly expressed in flower organs and in
developing seeds concomitantly to the pigmentation of the glands and the accumulation of sesquiterpenoid
aldehydes [196]. On the other hand, CAD1-A is weakly expressed in these tissues, but is highly expressed in
elicited suspension-cultured cells of cotton that produce sesquiterpene aldehydes [196]. Recent data demonstrate that under these conditions, a specific cotton WRKY transcription upregulates the expression of
CAD1-A [197].
Like other aspects of isoprenoid synthesis in plants, further work is required to master the control of the
pathway. This will help explain why the overexpression of S-linalool synthase gene in Petunia hybrida not only
resulted in trace amount of volatile S-linalool but induced the accumulation of S-linalyl-b-D-glucopyranoside,
376
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
its conjugated derivative [198]. Also, Arabidopsis overexpressing strawberry linalool/nerolidol synthase produced linalool, in addition to its glycosylated and hydroxylated derivatives [199]. Similarly, ripening tomato
fruit expressing Clarkia breweri S-linalool synthase accumulated trace amounts of S-linalool and hydroxylinalool [200]. At the other extreme, transgenic tobacco expressing three monoterpene synthases c-terpinene
synthase, limonene synthase and b-pinene synthase produced 10–25-fold monoterpenoids [201]. Collectively,
these findings are reminiscent of the fact that high level production of novel fatty acid in plants induced a futile
cycle of fatty acid synthesis and degradation [202]. Thus one may anticipate that obtaining increased and
unmodified final products will implicate genes which induce the differentiation of sequestery structures such
as trichomes, resin ducts and secretory pockets.
6. Post-squalene sterol biogenesis
6.1. Background
Sterols are essential components of fungal, animal and plant membranes. They regulate membrane fluidity
and permeability and interact with lipids and proteins within the membrane. These structural and membrane
roles of sterols in different organisms or models, including the formation of ‘‘rafts’’, have been reviewed elsewhere in a number of well documented papers [203–209], including some dedicated more specifically to plant
sterols [210]. The sterols and the different biosynthetic routes operating in various organisms are now well
established (Fig. 8) and have been the object of a number of reviews [211,212] including plants [213–217]. Sterols are precursors of a vast array of compounds involved in important cellular and developmental processes in
animals (reviewed in [218]) and plant sterols are particularly linked to brassinosteroids synthesis for review:
[219–221]. More recently, functional analysis of sterol function in plants has involved a range of approaches
including genetic studies. Indeed, mutational and transgenesis studies have given new insights into the role of
sterols in plant development and have been reviewed elsewhere [222–225]. In yeast, all relevant genes of the
ergosterol pathway have been cloned allowing isolation of mutants by gene disruption 24, [226] for review.
More recently in plants, functions of sterol biosynthetic enzymes have been confirmed by cloning and characterization of eleven of the putative seventeen genes in the portion of the pathway committed to the transformation of oxidosqualene to sitosterol (4) (Fig. 9) and reviewed elsewhere [179,217,227].
6.2. Pathway of phytosterol synthesis
The 21 enzymatic steps in the synthesis of phytosterols from oxidosqualene are all catalyzed by membrane-bound enzymes of the endoplasmic reticulum [214,228]. The transformation of cycloartenol (6) to phytosterols has been mapped by a wealth of isotopic labeling experiments mainly based on (+)-R-mevalonic
acid incorporation [211,229] and, or, determination of sterol profile following treatment with inhibitors of
sterol biosynthesis [230,231]. Another approach consisted in isolation, biochemical and enzymological characterization of microsome-bound enzymes that catalyze sterol transformation [214,228]. However, until very
recently little had been known about the active-site of any of these enzymes. More recently, the genes coding
for a number of enzymes in the oxidosqualene to phytosterol segment of the pathway has been expressed in
the corresponding yeast defective mutants [216,227] allowing molecular and structure–function studies of the
corresponding enzymes. Fig. 8 represents a comparative pathway of sterol biosynthesis in mammals, fungi
and higher plants. As most pathways of terpenoids biosynthesis, enzymes involved are grouped into two
major classes of reactions. A first major group comprises reactions involving a common carbocationic mechanism and catalyse electrophilic cyclizations, alkylations, reductions, and isomerizations, and ultimately
quenches the initially produced positive charge by proton elimination, alkyl group or hydride addition
(Fig. 10). Enzymes from the second group catalyze key oxidation reactions involving the insertion of molecular oxygen (Fig. 11). They are, in part, mediated by members of the extremely broad family of cytochrome
P450 monooxygenases, or as shown later in this report, by several members of a small but increasing family
of membrane-bound non-heme iron oxygenases involved in lipid biosynthesis. In these lines it should be
emphasized the key role of the insertion of oxygen en route to the remodeling of bioactive natural terpenoids
as shown in Fig. 12.
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
ANIMALS
377
HIGHER PLANTS
ALGAE
FUNGI
O
24
24
14
14
HO
4
HO
HO
1. 14 DM
2. 14 SR
HO
1. 4 DM
2. 4 DM
4 DM1
HO
HO
ANIMALS
FUNGI
HO
HO
8,7 SI
8,7 SI
HO
HO
4 DM2
HO
5 DES
5 DES
HO
4
5 DES
HO
HO
R
HO
HO
cholesterol (1)
HO
ergosterol (2)
R = CH3
campesterol (3)
sitosterol (4)
R = C 2H 5
R = C2H5 , ∆ 22 stigmasterol (5)
Fig. 8. Compared sterol biosynthetic pathway in different organisms.
378
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
HO
HO
HO
(2)
(1)
HO
HO
(3)
(5)
(4)
H
HO
HO
HO
(6)
HO
(7)
(8)
H
H
(D)H
HO
H(D)
(9)
(10)
H
H
HO
HO
H
HO
(11)
HO
HO
(12)
(14)
13
(15)
H
H
H
H
HO
HO
HO
HO
(18)
(17)
(16)
HO
(19)
(20)
N
HO
HO
(21)
HO
(26)
(23)
(22)
HO
(24)
(25)
HO
HO
COOH
HO
HO
(27)
(28)
Fig. 9. Chemical structure of the sterols and triterpenoids considered in the present work. Numbers are those used in the text.
6.3. Electrophilic processes in phytosterol synthesis
6.3.1. Oxidosqualene-cyclases
(3S)-2,3-Oxidosqualene cyclases (OSC) are versatile catalysts in plants which together are responsible for
the formation of a vast array of tetra- and pentacyclic triterpenoids. Since the first hypothesis concerning the
course of cyclization of squalene followed by rearrangements to lanosterol [232], and the proposed stereochemical rules for the process [233,234], the tools of synthetic chemistry, protein chemistry, molecular biology, and structural biology have been used to unravel the subtile interactions between oxidosqualene and its
cyclases. Most of the work has been done with the mammalian and prokaryotic enzymes and has been
reviewed comprehensively in several papers [235,236]. The last 10 years have witnessed the cloning, primary
amino acid sequences and putative function elucidation of more than 15 plant cyclases. In addition to the
formation of cycloartenol (6), higher plant OSCs catalyze the cyclization of oxidosqualene into a variety
of pentacyclic triterpenoids. The first plant OSC was cloned from A. thaliana, and encoded a cycloartenol
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
379
20
8
HO
2
H
HO
HO
HO
cycloartenol synthase
cycloartenol synthase
24
R
C24 methyl transferase
cycloartenol synthase
R
R
R
H
9
9 8
8
HO
HO
HO
∆88−∆77-sterol isomerase
Cycloeucalenol isomerase
14
HO
∆8,14-sterol-∆14-reductase
5
6
7
∆5,7-sterol-∆7-reductase
Fig. 10. Selected carbocationic high energy intermediates involved in reactions of plant sterol biosynthesis.
HO
4
HO
CH2OH
CO2H
C4 methyl oxidase
C4 methyl oxidase
HO
HO
CH2OH
CO2H
C4 methyl oxidase
C4 methyl oxidase
14
CHO
CH2OH
HO
HO
C14 demethylase
C14 demethylase
Fig. 11. Selected oxygenated intermediates produced by oxygenases of sterol biosynthesis.
(6) synthase [237]. More recently, a number of plant triterpene synthases genes have been cloned and functionally characterized in plants. This includes for example: a,b-amyrin (10) synthase from Panax ginseng
[238], lupeol (11) synthase, a- (12) and b-amyrins, W-taraxasterol (13), taraxasterol (14), bauerenol (15), multiflorenol (16), butyrospermol (17), tirrucalla-7,21-dienol (18) synthases from A. thaliana [239–241], Medicago truncatula and Lotus japonicus [242,243], and cucurbitadienol synthase from Cucurbita pepo [243].
The polypeptide encoded by the A. thaliana lupeol synthase was shown to be a novel multifunctional triterpene synthase. Schuffling domain mutagenesis indicated that the second quarter from the N-terminus of the
synthase was shown to have a crucial importance in determining the cyclization toward lupeol or amyrin
380
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
HO
O
OR
CO2H
CO2Me
OH
O
O
H
O
H
H
OH
H
CO2H CO2H
AcO
MeO2C
H
COOH
GA66
OH
O
abietic acid
azadirachtin
OH
OH
OH
OH
HO
HO
HO
OH
HO
HO
O
O
O
brassinolide BR1
20-hydroxyecdysone
Fig. 12. Examples of highly oxidized bioactive terpenoids.
[244]. Similarly, a new a-amyrin producing enzyme from Pisum sativum was shown to be a multifunctional
triterpene synthase [245].
(3S)-2,3-Oxidosqualene is also channeled to the synthesis of triterpene saponins. Triterpene saponins are
restricted to a large class of dicotyledonous plants, that exclude Arabidopsis. Although cereals are deficient
in saponins, oats synthesize specific saponins termed avenacins [246]. Their synthesis from 2,3-oxidosqualene
involves cyclization of (3S)-2,3-oxidosqualene to give primarily oleanane (b-amyrin) or dammarane triterpenoid. The resulting terpenoid skeleton is subject to various modifications including oxidation, substitution and
glycosylation, catalyzed by cytochrome P450-dependent monooxygenases, glycosyltransferases and other
enzymes [247]. In oats these genes are clustered in the genome, thus giving a rare example of gene clustering
in the the biosynthesis of plant secondary metabolites [248].
In 1999 the first three-dimensional structure of a triterpene cyclase (squalene hopene cyclase from Alicyclobacillus acidocaldarius) was obtained [249]. However X-ray structural information on OSC is still lacking.
Based on the structure of A. acidocaldarius squalene hopene cyclase and by exploiting homologies between
the different OSC, homology modeling studies on human OSC lanosterol cyclase have been presented [250].
The model is in accordance with previously performed studies with mechanism-based suicide inhibitors
and the aforementioned mutagenesis experiments. It offers another way to apply structure-based drug-design
methods aiming to develop novel antifungal, hypocholesterolemic and phytotoxic compounds, an interest
which is still unbroken. This combination of structural and biological studies has produced important mechanistic insights into both squalene cyclase (SC) and OSC and the data have led to a unified mechanism that is
consistent with the chemical and structural findings.
The relaxed substrate and product specificities of a number of plant oxidosqualene cyclases, and also bacterial squalene-cyclases, point to the possibility of generating novel triterpenes by the directed evolution of
these enzymes. Directed evolution approaches have been described to find mutations that allow A. thaliana
or Dictyostelium discoideum cycloartenol synthase to biosynthesize lanosterol (19) [251,252], parkeol (20)
[253] or other altered triterpene product profiles [254–256]. These precise analyses of mutations responsible
for altered product profile may provide important insights into the role of residues close or distant from
the active-site of the cyclases.
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
381
6.3.2. Other enzymes
Other carbocationic steps in the conversion of cycloartenol to stigmasterol include: methylation of cycloartenol (6) to 24(28)-methylenecycloartanol (21); cleavage of the 9b,19-cyclopropyl ring of cycloeucalenol
(22) to produce a D8(9)-sterol, obtusifoliol (23); reduction of D14(15)-bond of D8,14-sterols to produce a
D8(9)-sterol; isomerization of the 8(9)-bond to the 7(8)-position; methylation of the side-chain D24(28)-bond
to produce a 24(28)-ethylidene derivative; reduction of the D7-double bond; isomerization of the D24(28)bond to the 24(25)-position and reduction of the D24(25)-double bond. In addition to early work concerning
the enzymology of the two methylation steps in plants [257,258], more recent data concerning this step present only in plants and fungi, have been extensively reviewed [216,259,260]. Enzymology of other anaerobic
steps in plant sterol biosynthesis has been reviewed previously [216,230,231], and a lot of gaps still remain in
the knowledge of molecular events involved in the course of these enzymatic reactions in plants. Molecular
biology and mutagenesis analysis should yield new insights in defining the course of these mechanistically
linked enzymes.
6.3.3. Carbocationic transition state analogs
The first described nitrogen containing sterol biosynthesis inhibitor is triparanol that was shown to lower
serum cholesterol in hypercholesterolemic rats and to lead to an accumulation of 24-dehydrocholesterol (desmosterol) (24) in the liver and serum of these rats, suggesting blockade of the sterol C24-reductase [261]. The
first sterol biosynthesis inhibitor which was proposed to function as a carbocationic high energy intermediate
analog (HEIA) (transition state analog; TSA), was 25-azacycloartanol (25) which was found to be a potent
inhibitor of SAM-cycloartenol-C24-methyltransferase and SAM-24-methylenelophenol-C28-methyl-transferase [262]. Indeed, further studies [257,258] confirmed that (25) and analogs did not behave as substrate for
N-methylation catalyzed by the C24 methyltransferase, as it was proposed for the inhibition of C24-methyltransferase by triparanol [263]. Subsequently a vast array of rationally designed carbocationic transition state
analogs possessing a heteroatom with a steady positive charge in the place of the carbenium-C were synthesized and could simulate HEIs and potently inhibit specific enzymes of terpenoids biosynthesis including:
oxidosqualene cyclases, C24- and C28-methyl transferases, cycloeucalenol isomerase, D8,14-sterol D14 reductase, D8-D7-sterol isomerase and D5,7-sterol D7 reductase, for example, for review: [228,230,231,263], supporting the carbocationic mechanism of the reactions catalyzed by these enzymes. Interest to use such an approach
to design novel antifungal compounds is still unbroken [264,265].
6.4. Membrane-bound non-heme iron oxygenases involved in phytosterol synthesis
A key step in terpenoid biosynthesis is the insertion of oxygen en route to the formation of endpathway
products. Such a process is involved in the synthesis of many bioactive terpenoids possessing additional hydroxyl or carboxylic-acid groups in their structure (Fig. 12). In addition, it is also involved in the remodeling of
sterol precursors to furnish the final physiologically active sterols, i.e., cholesterol, ergosterol or sitosterol.
Indeed, removal of the three extra methyl groups at C4 and C14 and insertion of the D5 and D22 double-bonds
in sterol precursors result from such an oxidative process. This process is often mediated by members of the
extremely broad and thoroughly characterized protein family of cytochrome P450 monooxygenases, including
obtusifoliol-14a-demethylase and sterol-D22-desaturase of plant sterol synthesis. However, the past decade
has shed light on another smaller but increasing family of membrane-bound non-heme iron oxygenases catalysing highly stereoselective and regioselective O2 dependent hydroxylations, desaturations, epoxygenations
and acetylenations of lipids. These oxygenases are mostly metabolizing fatty acids; however, recent studies
indicate that oxygenases involved in the C4-demethylation and C5(6)-desaturation of sterol precursors are
members of this family. While considerable experimental data are available on the biochemistry and the role
of residues affecting the activity of P450 monoxygenases, less is known about the enzymology and the molecular determinants affecting the activity of these non-heme iron oxygenases.
6.4.1. Membrane-bound D7-sterol C5(6)-desaturase
Very few enzymatic systems able to carry out in vitro lipid desaturation reactions have been described so
far. We have recently shown in higher plants [266] that D7-sterol-C5(6)-desaturase (EC 1.3.3.2) (5-DES)
382
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
processes the desaturation of D7-sterol precursors to produce D5,7-sterols, as in mammals [267] and yeast [268]
(Fig. 8). We succeeded in isolating microsomal preparations from a higher plant (Zea mays), and from an erg3
yeast strain defective in endogenous 5-DES and expressing the A. thaliana 5-DES, which functioned in the
C5(6)-desaturation reaction. Enzymatic activity requiring molecular oxygen, was strongly promoted by addition of NADH or NADPH, was inhibited by cyanide, hydrophobic metal chelators and cytochrome c, and
was insensitive to carbon monoxide. In addition, the maize microsomal 5-DES was strongly and completely
inhibited by IgG raised against plant cytochrome b5 [269] indicating that membrane-bound cytochrome b5 is
an obligatory electron carrier from NAD(P)H through NAD(P)H cytochrome b5 reductase to the plant 5-DES
as in the mammalian 5-DES process [270]. These properties showed strong similarities to those of soluble and
membrane-bound fatty acid desaturases such as the membrane-bound D12-desaturase from safflower for
example. They were not consistent with the involvement of cytochrome P450 in the reaction. Taken together,
these biochemical data suggested that the 5-DES is catalyzed by a non-heme iron oxygenase [266]. The substrate selectivity study for the maize 5-DES indicated that the presence of a C7–C8 double bond was necessary
for desaturation [266]. The C7 double bond may be required for the activation of the allylic C6 hydrogen
atom. Alternatively, the enzyme may be highly specific, in which case the absence of a double bond at C7–
C8 would lead to alterations of the orientation of the substrate binding during catalysis and particularly of
the hydrogen atoms at C5 and C6, thus affecting the course of the reaction. Taken together, the results demonstrated directly that during plant sterol synthesis the D5-bond is introduced via the sequence D7-sterol ! D5,7-sterol ! D5-sterol (Fig. 8).
A cDNA encoding 5-DES has been cloned in A. thaliana by functional complementation of a defective
yeast mutant (erg3) [271]. The A. thaliana 5-DES protein shares 29%, 35% and 29% identity with the yeast
[272], human [273] and Candida 5-DES, respectively, which differ slightly in size. The heterologous expression
of A. thaliana 5-DES in the yeast erg3 mutant led to a functional microsomal preparation of recombinant 5DES, and constituted, to our knowledge, the first in vitro functional recombinant plant membrane-bound
lipid desaturation yeast system described so far. It allowed isolation of sufficient amounts of 5-DES for enzymological and mechanistic studies, and site-directed mutational analysis.
Investigation of the role of a variety of evolutionary conserved residues in the 5-DES reaction was carried out using site-directed mutagenesis and expression of the mutated enzymes in the erg3 yeast strain
defective in 5-DES. The first group of mutants was affected in the eight conserved histidines from three
histidines-rich motifs. Site-directed mutagenesis experiments in which each of the eight conserved histidines
residues (His 147, 151, 161, 164, 165, 238, 241 and 242) was individually converted to leucine (which is
hydrophobic and electrically neutral), totally eliminated the 5-DES activity both in vivo and in vitro,
and established that all the histidines are essential for catalysis [274]. The role of the above mentioned histidine-rich motifs was probed by mutating them to glutamic acid residues, which could provide ligands for
a possible Fe center and change the reactivity of the 5-DES, as has been proposed for soluble desaturases,
the R2 subunit of ribonucleotide reductase [275,276] and the methane monooxygenase hydroxylase [277].
However, glutamic acid substitutions led to completely inactive mutants. In contrast, mutation of conserved histidines or other residues not part of the consensus motifs indicated that these residues were
not essential for the catalysis but contributed to the activity through conformational or other effects
[274]. Hydropathy plot of the plant 5-desaturase from different sources indicated that the three histidinerich motifs were located in hydrophilic domains from the enzyme. One possible function for the tripartite
consensus sequence, HX3H, HX2HH and HX2HH, of the 5-DES, consistent with the mutagenesis and biochemical data, would be to provide the ligands for a presumed catalytic Fe center, as previously proposed
for a number of integral membrane enzymes catalysing desaturation and hydroxylations [278,279] and is
discussed later.
A nuclear and recessive mutant of A. thaliana, ste1, defective in the sterol-C5-desaturase has been isolated and identified [280]. This mutant accumulates D7-sterols, (24R)-24-ethyl-5a-cholest-7-en-3b-ol (7),
and (24n)-24-methyl-5a-cholest-7-en-3b-ol (8) at the expense of the normally occuring campesterol (3)
and sitosterol (4). The 5-DES ORF from the A. thaliana leaky mutant ste1, contains a single mutation,
T114I, which was proposed to be responsible for the 5-DES deficiency of this mutant. The T114I mutant
was expressed in the erg3 yeast mutant and further enzymatically analyzed in the corresponding microsomal
preparation. Threonine 114 was found not to be essential for the 5-DES activity. However, replacement of
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
383
Threonine 114 by the functionally non-conservative isoleucine led to a 10-fold reduced catalytic efficiency of
the 5-DES and the corresponding transformed yeast accumulated substantial amounts of D7-sterols at the
expense of D5,7-sterols in accordance with data obtained with the ste1 mutant [271,274]. An hydroxyl function is conserved at position 114 in the ERG3 family. Thr114 was replaced by the functionally conservative
serine residue and the catalytical properties of mutants T114S were compared to that of the wild-type 5DES. Mutant enzyme T114S displayed a 28-fold higher Vmax value than the wild-type enzyme. Moreover,
V/K values were 4-fold higher than that of the native 5-DES, indicating that a hydroxyl containing sidechain at position 114 substantially increased the catalytic rate of the 5-DES. This data supported the role
of threonine 114 in the stabilization of the transition state of a rate-controlling step in the 5-DES catalytic
pathway [274].
It has been previously proposed that enzymes maximize rates by binding transition states strongly and substrate weakly [281], and that they should have evolved in this way [282]. From this point of view, a serine residue at position 114, brings an apparent additional stabilization of the transition state and decreases the
apparent binding energy of the substrate in the ground-state. In contrast, for mutant T114I, the free energy
of the substrate complex both in the ground-state and in the rate determining transition state was approximatively uniformly increased. The result was a marked decrease in the activation energy to reach the transition
state and a substantial improvement of the reaction rate of this mutant.
The ultra-selective oxidation chemistry and the detailed mechanism of the sterol-C5(6)-desaturase was
probed using an approach based on intermolecular primary deuterium kinetic isotope effects. Deuteriumlabeled 5a-cholest-7-en-3b-ol (9) bearing one or two deuteriums at the C-5a and (or) C-6a positions were synthesized in high isotopic and chiral purity. These compounds were used as substrates with the microsomal
wild-type Z. mays and recombinant A. thaliana D7-sterol-C5(6)-desaturases to probe directly the stereochemistry and the mechanism of the enzymatic reaction. Clearly, in the conversion of (9) by both 5-DESs, the 6ahydrogen was removed [283]. This data indicated a stereospecific syn desaturation during catalysis by both
plant 5-DES. It is in accord with former labeling pattern data obtained in vivo in plants [284], fungi [285]
and rat liver [286]. 6a-2H-5a-cholest-7-en-3b-ol showed a substantial intermolecular deuterium kinetic isotope
effect (DKIE) on V and V/K, D6V = 2.6 ± 0.3, D6V/K = 2.4 ± 0.1 and D6V = 2.3 ± 0.3, D6V/K = 2.3 ± 0.2 for
the Z. mays and A. thaliana wild-type 5-DES, respectively. In contrast, negligible or minor isotope effects,
D5
V = 0.99 ± 0.04, D5V/K = 0.91 ± 0.08; and D5V = 0.93 ± 0.06, D5V/K = 0.96 ± 0.04, respectively, were
observed with 5a-2H-cholest-7-en-3b-ol. The observed pattern of isotope effects indicated that: (i) the desaturation involves asynchronous scission of the two C–H bonds at C5 and C6. The data clearly indicated a stepwise mechanism, and ruled out a concerted mechanism, (ii) the 5-DES initiates oxidation by cleavage of the
chemically activated C6a–H bond, (iii) this chemical step involving considerable C6a–H bond stretching, is
partially rate-limiting the overall desaturation process. A possible mechanism for the 5-DES consistent with
the data would involve an initial, energetically difficult C6a–H activation step, executed by a compound Qtype oxidant [287] to produce a C6 carbon-centered radical/ FeOH pair (Fig. 13). This step would be similar
to that proposed for a number of reactions catalyzed by cytochrome P450-dependent hydroxylations and
desaturations involving an hypervalent iron-oxo species [288,289]. This radical could collapse to olefin via different pathways. Obtained experimental evidences favored a disproportionation reaction (pathway a) to give
an olefinic product and iron-bound water [283] (Fig. 13). In the case of the plant 5-DES, obtained data did not
favor pathway c including a C6a- or C6b-hydroxyl intermediates previously suggested [290], since such intermediates could not be detected during the reaction, and second, synthetic samples of these intermediates were
not converted into 7-dehydrocholesterol by the enzymatic preparation [283]. The mechanism found for the
present steroid desaturase is similar with that recently demonstrated for a number of membrane-bound
fatty-acid desaturases and hydroxylases which also involve a large deuterium KIE for one of the C–H cleavage
step while a negligible effect is found at the proximal carbon [291,292]. The switch controlling the ratio of
desaturation versus hydroxylation remains an intriguing question. What governs the orientation of the reaction pathway to desaturation at C5(6) versus possible hydroxylation at C6 is not currently understood and
presumably reflects the precise positioning of the substrate relative to the oxidant species within the Michaelis
complex, to produce a preferential loss of the C5 hydrogen to form an olefin, instead of a C–O bond formation
to form an alcohol at C6. That is, as suggested before, if access to the proximal carbon hydrogen is hampered,
then hydroxylation of substrate would be favored (negative catalysis) [293].
384
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
HO
H
*
H
OH
III
III Fe
Fe
O
*
H
5
HO
H
HO
6
H
O
IV
Fe
O
*
*
H
IV Fe
HO
6
H
H
H2
O
OH
III
III Fe
III
IV Fe
Fe
Fe
O
O
*
s
H
HO
H
OH
III
III Fe
Fe
O
Fig. 13. Pathway a: proposed mechanism for D7-sterol-C5(6)-desaturase; experimental data are not consistent with pathways b or c.
The preceding methodology was combined with variation of enzyme structure to find information on the
role of threonine 114 in catalysis by the 5-DES from A. thaliana. The data indicated the same stereochemical
integrity for the T114 mutants and a primary DKIE pattern of desaturases that is qualitatively similar to that
observed for the wild-type desaturase, suggesting a similar reaction pathway for these isoforms. In addition,
the data indicated that a hydroxyl-containing side-chain at position 114 substantially increases the catalytical
rate of the 5-DES and makes the deuterium sensitive step less rate limiting [283]. There is no evidence that the
intrinsic isotope effect is changed in the mutants. Rather it is likely that the chemistry is slower in the T114I
mutant (STE1), decreasing the commitment and resulting in a more fully expressed KIE, closer to the intrinsic
value [294]. Conversely, in the case of mutant T114S, the chemistry is faster but subsequent steps, as product
release, change less, so that the observed KIEs are smaller.
6.4.2. Membrane-bound sterol-4a-methyl-oxidases from the C4-demethylation complex
In all organisms, the sterol molecule becomes functional only after removal of the three methyl groups at
C4 and C14 [203]. The inefficiency of sterol precursors possessing C4-methyl groups has been shown particularly in the yeast Saccaromyces cerevisiae, where ERG25 the gene coding the sterol-4a-methyl-oxidase
(SMO) is essential [295]. In animal and yeast, following removal of the C14 methyl group of lanosterol, both
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
385
C4-methyl groups are sequentially eliminated by the same C4-demethylation enzymes complex involving a single C4-oxidase gene, ERG25 [295,296]. The C4 demethylation process is central in sterol biosynthesis and
involves 10 of the 20 enzymatic steps from oxidosqualene to phytosterols. It is energetically costly since it
requires six upon the eleven oxygen molecules and eight upon the eighteen NAD(P)H molecules consumed
in the post squalene segment of sterol biosynthesis. Moreover, since at least three crucial enzymes are involved,
there are more opportunities for regulation and inhibition.
During the conversion of cycloartenol to phytosterols in photosynthetic eukaryotes, early in vivo labelling
experiments [297,298] and more recent biochemical data [299] strongly suggested that two biosynthetically
distant and distinct oxidative systems are involved in the C4-demethylation of sterol precursors and thus,
is profoundly different with the animal and fungal systems. It has been enzymatically demonstrated in animals and in plants that the C4-demethylation complex involves a SMO which produces a 4a-carboxy-sterol
derivative [299,300]. It is subsequently oxidatively decarboxylated by the 4a-carboxysterol-C3-dehydrogenase/C4-decarboxylase (4a-CD) [301–303] to produce a 3-oxosteroid which is stereospecifically reduced by
a NADPH-dependent sterone reductase [304–306] (Fig. 14). In yeast, the genes encoding these components
have been cloned and identified, that is, ERG25 [295], ERG26 [302] and ERG27 104 [306]. Recently the enzymatic system corresponding to the SMO component of this complex in yeast has been described [307]. It
allowed to specifically investigate the enzymological properties of the yeast SMO and, particularly, to detect
and identify the immediate oxidized metabolite by the yeast SMO, that is a C4-hydroxymethyl-sterol derivative which was identified by MS analysis of its diacetate derivative. Thus, in yeast, as well as in animal [300]
and plants [299], the SMO is a multifunctional oxidase catalyzing three successive monooxygenations of the
C4-methyl group to produce the C4-carboxysterol-derivative. Finally, similarly as for the 5-DES, biochemical
properties of the plant and yeast SMOs were consistent with the involvement of a non-heme iron oxygenase
[299,307].
Fig. 14. Putative multienzymatic complex involved in sterol C4-demethylation.
386
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
By combining homology and sequence motif searches with knowledge relating to sterol biosynthetic genes
across species, five SMO cDNAs from A. thaliana (AtSMO) belonging to two families have been recently
cloned [308,309].
The two families of plant SMOs share only 37% sequence identity to each other and 26–34% identity to the
yeast and mammalians SMOs. They cluster well together and with ERG25 from animals and yeast, as shown
in the phylogenetic tree of Fig. 15. They also cluster with other identified membrane-bound non-heme iron
hydroxylases such as cholesterol 25 hydroxylase (CH25H) [310] or C4 sphingolipid hydroxylase (SUR2)
[311]. Heterologous expression of these cDNAs in a yeast erg25 ergosterol auxotroph, lacking sterol-4amethyl-oxidase activity, led to two groups of complementation: The SMO2 group of complementation
Fig. 15. Phylogenetic tree for selected members of the membrane-bound non-heme iron oxygenases family. SMO, sterol methyl oxidase;
AtSMO21, At1g07420; AtSMO22, At2g29390; LeSMO21, TC125471; TaSMO2, TC66074; OsSMO1, OSJNBa0001014.3; HsSMO,
NM006745; ScSMO, NP013157; C25H, cholesterol 25 hydroxylase; HsC25H, AF059214; MmC25H, AAC87480; SUR, C4-shingolipid
hydroxylase; ScSUR2, AAA16608; AtSUR21, At1g14290; AtSUR22, At1g69640; 5DES, D7-sterol-C5(6) desaturase; Nt5DES2,
AAD20458; Nt5DES, AAD04034; At5DES1, At3g02580; At5DES2, At3g02590; Hs5DES, NP008849; Mm5DES, BAA33730; Cg5DES,
AAB02330; CYMAA, p-cymene monooxygenase, hydroxylase; PsCYMAA, AAB62299; ALK, alkane-1-monooxygenase; PsALKB,
CAC40951; D12FA, D12-fatty acid epoxygenase; CpD12A, Y16283; 12AC, D12-fatty acid acetylenase; Ca12AC, CAA76158; FAD2, x-3
fatty acid desaturase (endoplasmic reticulum); AtFAD2, P48623; FAD7, x-3 fatty acid desaturase (chloroplast); AtFAD7, NM111953,
At3g11170; FAD6, x-6 fatty acid desaturase (chloroplast); AtFAD6, NP-194824; FAT, x-3 fatty acid desaturase; CeFAT1, AAA67369;
b5DES, D8-shingolipid desaturase; Atb5DES, At2g46210; D8SDES, D8 desaturase; AtD8SDES, At3g15870; D8DES, D8 sphingolipid
desaturase; Ha D8DES, S68358; D6DES, fatty acid D6-desaturase; BoD6DES, AF007561.1; FAH, fatty acid hydroxylase; AtFAH1,
At4g20870; AtFAH2, At2g34770; PiFA, fatty acid desaturase, CAD77651; SynD9DES, D9-desaturase; ScOLE1, fatty acyl-CoA D9deasturase, NP011460.1; PaCRTW, b-carotene ketolase, BAA09591.1; CAH, carotenoid hydroxylase; CaCAH1, Y0225; CaCAH2,
CAA70888; CER, aldehyde decarbonylase; AtCER1, At1g02205.2; AtCER1 like, At2g37700; AtCER1/2 like, At1g02190.
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
387
restored growth, but only low levels of ergosterol biosynthesis. The SMO1 group did not complement erg2525c [308,309]. This data suggest that only one family of SMOs is functional for the oxidation of 4,4-dimethylzymosterol, the physiological substrate of the yeast SMO accumulating in erg25. These results represent
the first functional identification of sterol-4a-methyl-oxidase genes from plants. All AtSMOs amino-acid
sequences were characterized by the presence of three histidine-rich motifs, HX3H X8 or 9 HX2HH X73 or 75
HX2HH, which exhibit a topology and spacing of amino acids within the histidine motifs typical of the
ERG3–ERG25 family, PFAM01598. Moreover, the spacing and topology of the histidine motifs are clearly
distinct from those found in the extended family of membrane-bound fatty acid desaturases/hydroxylases.
The identification of conserved motifs in different enzymes implies that they are evolutionarily related provided that the structure and relative spacing of the motifs is comparable. Indeed, elements of the motif are
equally spaced in the ERG3–ERG25 family and are also placed in equivalent positions with respect to potential membrane-spanning domains [274,296,308]. In the case of fatty acid desaturase–cytochrome b5 fusion proteins, the motif is predicted to occur on the same face of the membrane as the cytochrome b5 [312,313].
Because the histidine-rich motifs are essential to 5-DES as well as to other enzymes of the family, it has been
proposed that they are involved in coordinating an active-site metal center [278,279]. Moreover, spectroscopic
studies of alkane x-hydroxylase provided evidence that a di-iron center is present in at least one member of
this class of integral-membrane enzymes [314]. In the absence of crystallographic data for this membranebound desaturase, no information about specific details of the active-site including the proposed histidine
coordinating ligands and the location of the hydrophobic binding pocket is available. The only data in this
context have been obtained from the crystal structure of the related soluble castor seed D9 stearoyl-ACPdesaturase. In that case, the iron atom is buried in the middle of the structure, resulting in an isolated local
environment suitable for reactive oxygen chemistry [315].
To elucidate the precise functions of SMO1 and SMO2 genes families, their expression was reduced by
using a virus-induced gene silencing (VIGS) approach in N. benthamiana. Two cDNA fragments of 387
and 498 bp corresponding, respectively, to NtSMO1 and NtSMO2 were cloned from N.tabacum and were
inserted into the viral TTO1 vector, and N. benthamiana inoculated with the viral transcripts. The sterol profile of infected plants was analyzed: following silencing with SMO1, a substantial accumulation of 4,4dimethyl-9b,19-cyclopropyl-sterols (i.e., 24-methylenecycloartanol, (21)) was obtained, while qualitative
and quantitative levels of 4a-methylsterols were not affected. In the case of silencing with SMO2, an important accumulation of 4a-methyl-D7-sterols (i.e., 24-ethylidenelophenol (27) and 24-ethyllophenol) was found,
with no change in the levels of 4,4-dimethylsterols.
These clear and distinct biochemical phenotypes demonstrate that, in contrast to animals and fungi, in photosynthetic eukaryotes, these two novel families of cDNAs are coding two distinct types of C4 methylsteroloxidases controlling, respectively, the level of 4,4-dimethyl- and 4a-methylsterols precursors [309].
To our knowledge, plant mutants affected in SMO activity have not been reported yet. In this respect, the
present study could give important clues for the elucidation of the physiological roles of 4,4-dimethyl- and 4amethyl-sterols, and of the biological significance of the existence of two distinct C4-demethylation complexes
in photosynthetic eukaryotes.
Finally, the phylogenetic tree of the membrane-bound non-heme iron oxygenases family, shown in Fig. 15
clearly reveals the diversity of their functions.
6.4.3. Other components of the C4-demethylation complex
The enzymes corresponding to the remaining two steps in the plant C4-demethylation process have been
identified and their properties studied in maize. Evidence for the existence of a microsome-bound NAD+dependent 4a-carboxysterol-C3-dehydrogenase/C4-decarboxylase (4a-CD) in plant was directly obtained
from a maize microsomal preparation able to oxidatively decarboxylate several substrates, including 4a-carboxy-cholest-7-en-3b-ol (26). The BriJ W-1 solubilized enzyme was partially purified 290-fold and the apparent molecular mass of 4a-CD in sodium cholate was estimated to be 45 kDa [303]. The results supported the
contention that cleavage of the C4–C32 bond of plant C4 methylsterols resulted of two separate processes:
An oxygen and NAD(P)H-dependent oxidation of the 4a-methyl group to produce the corresponding 4acarboxylic acid, followed by oxygen-independent cleavage that is consistent with the conclusion that this
388
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
oxosteroid reductase was associated with the C4-demethylation process. In plants, study of genes for decarboxylation and ketone reduction remain to be carried out. Together with the SMOs genes, these genes are
interesting because no corresponding plant mutants have been detected in A. thaliana. This might mean that
the ability to demethylate is essential for viability of the plant cell in accordance with the essentiality of
ERG25, ERG26 and ERG27 in yeast [295,301,306]. In addition, a recent S. cerevisiae microarray expression
study indicated that an ORF, YER044C, designed ERG28, was strongly coregulated with ergosterol synthesis [316]. Disruption of the ERG28 gene in yeast results in slow growth and accumulation of sterol intermediates of the C4 demethylation process. Two-hybrid analysis indicated that Erg25p, Erg26p, and Erg27p
components required for the C4-demethylation, form a complex within the yeast cell [317]. Coimmunoprecipitation studies suggested that Erg28p works as a transmembrane scaffold to tether Erg27p and possibly
other C4 demethylation proteins (Erg25p, Erg26p), forming a demethylation complex in the yeast endoplasmic reticulum [318] (Fig. 14).
Although, there are at least three enzymes involved in the C4 demethylation complex, no agriculturally or medically useful inhibitors such as the widely used triazoles or morpholines series [230,231] have
emerged. However, the antifungal agent, 6-amino-2-n-pentylbenzothiazole was shown to directly inhibit,
in vitro, the yeast SMO, but was inactive with the plant SMOs [307]. In addition, fenhexamid, a recently
developed botryticide was shown to inhibit sterol biosynthesis in the fungus Botryotinia fuckeliana and
led to the accumulation of 3-oxo compounds suggesting an inhibition of the 3-oxosteroid reductase
[319].
6.5. Kinetically favored pathway in the conversion of cycloartenol to D5-sterols
Most of the reports on the enzymology of sterol transformation in plants have been obtained with a microsomal preparation isolated from Z. mays seedlings. The described data include apparent kinetical parameters,
Km and Vmax for substrate, cofactors, and a variety of substrate analogs, allowing the substrate specificities of
these distinct enzymes to be determined. Since the enzymes were all studied in the same microsomal preparation
isolated from the same plant tissue, the different Km and Vmax values may reveal, together with the control by
the substrate specificity, firstly, the kinetically favored pathway which leads directly to the end-product during
active cell proliferation, and secondly, the rate limiting steps in the pathway. Examination of the enzymological
literature includes studies of S-adenosylmethionine-sterol-C24- and C28-methyl-transferases [257–260,320],
4,4-dimethylsterol- and 4a-methyl-sterol-C4-methyl-oxidases [299], 4a-carboxysterol-C3-dehydrogenase/
C4-decarboxylase [303], 3-oxosteroid-reductase [305], cycloeucalenol-isomerase [321], obtusifoliol-14a-methyl
demethylase [322], D8,14-sterol-D14 reductase [323], D8-D7-sterol isomerase [324,325], D7-sterol-C5(6)-desaturase [266], and D5,7-sterol-D7-reductase [326] (Table 1, Fig. 16). Firstly, it turns (Table 1) that maize possesses a
number of enzymes that exhibit a high degree of substrate specificity, including particularly monooxygenases
involved in the removal of the C4 and C14 methyl groups. The specificity of these enzymes is particularly strict
for structural motifs of the tetracyclic nucleus while variation of the side-chain structure are more often
accepted. For example, SMO1 requires a 9b,19-cyclopropyl-ring [299], SMO2 a D7-unsaturation [299] and
14DM absence of a 4b-methyl group [322]. If minor pathways are poorly active in the case of these enzymes
so that only certain intermediates may finally lead to D5-sterols, affecting their rate should be associated with
a change of the amount of end pathway D5-sterols. Secondly, another control of the tetracyclic compounds
pathway could be provided by enzymes with low turnover such as the C24-methyl transferase, the 4a-methyl
oxidases and the 14a-demethylase.
This kinetic control of the pathway is exemplified by recent studies of plant mutants affected in postsqualene enzymes. A first example is given by the biochemical characterization of a tobacco mutant (LAB 1–4) that
overproduces 10-fold more sterols [327] due to an increase of the 3-hydroxy-3-methyglutaryl-coenzyme A
reductase (HMGR) activity [328]. In this mutant, high accumulation of cycloartenol, 24-methylenecycloartanol (21) and 24-ethylidenelophenol (27), and substantial accumulation of obtusifoliol (23) were observed in
calli.
More recently, co-expression of HMGR and C24-sterol methytransferase type 1 in transgenic tobacco
enhanced carbon flux towards end-product sterols and resulted in elevated accumulation of substrates of
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
389
Table 1
Apparent kinetically favored sterol synthesis pathway in Zea mays
Step
Enzyme
Favored substrate
Km
(lM)
Vmax
(nmol mg1 h1)
V/K
(h1)
Substrate
specificity
References
a
b
c, d, e
f
g
h
i
j
SqO-cycloartenol synthase
SAM-sterol-C24-methyltransferase
Sterol-4a-methyl oxidase
4a-Carboxysterol decarboxylase
3-Oxosteroid reductase
Cycloeucalenol isomerase
Sterol-14a-methyl demethylase
D8,14-Sterol-D14 reductase
150
30
500
nd
220
100
160
100
1.5
1.2
8.1
126
9.1
36
3.9
13.8
0.08
0.16
0.06
nd
0.16
1.44
0.10
0.55
++
++
++++
+
+
++
++++
++
[235,258]
[257–260]
[299]
[303]
[305]
[321]
[322]
[323]
k
D8-D7-Sterol isomerase
75
12
0.64
++
[324,325]
l
m, n, o
p
q
r
s
SAM-sterol-C28-methyltransferase
Sterol-4a-methyl oxidase
4a-Carboxysterol decarboxylase
3-Oxosteroid reductase
D7-Sterol-C5(6) desaturase
D5,7-Sterol-D7 reductase
Oxidosqualene
Cycloartenol
24-Methylenecycloartanol
4a-Carboxy,4b-methylsterol
4a-Methyl-3-oxosteroid
Cycloeucalenol
Obtusifoliol
4a-Methyl-8,14,24(28)
ergostatrienol
4a-Methyl-8,24(28)
ergostadienol
24-Methylenelophenol
24-Ethylidenelophenol
4a-Carboxy-sterol
3-Oxosteroid
D7-Sterol
D5,7-Sterol
1.8
9.6
97.2
9.6
7.1
36.6
0.24
0.08
0.75
0.16
0.20
0.32
++
++++
+
+
+++
++
[257–260]
[299]
[303]
[305]
[266]
[326]
30
480
520
245
140
460
SqO
28
a 1.5
b
14
HO
28
28
c
14
4
4
HO
1.2
8.1
d
14
4
HO
g
9
4
h
9
4
HO
36
9.1
8
l
HO
7
1.8
3.9
4
q
O
7
OH
7.1
5
4
12
p
7
HO
O
s
HO
5
k
4
o
HO
r
HO
9.6
4
HO
29
n
HO
9.6
126
OH
8
j
13.8
29
m
HO
O
14
4
HO
29
4
HO
i
f
14
O
4
HO
e
14
4
HO
OH
O
28
24
HO
O
97.2
t
HO
HO
36.6
Fig. 16. Kinetically favored pathway to 24-ethyl sterols in Zea mays. Numbers with the reaction refer to the catalytic constant Vmax
(nmol mg1 h1) of the corresponding enzymatic reaction in a Zea mays microsomal preparation, as mentioned in Table 1.
the C4-demethylation reactions [329]. These data are consistent with the conclusion that the methylation of
cycloartenol, the two C4-demethylation and the C14 demethylation reactions limit carbon flow to end-product
sterols, at least in a physiological situation when the carbon flow is upregulated, in accordance with kinetic
data obtained in vitro in the above mentioned maize enzymatic preparations.
Another example is given by the sterol profile of a number of A. thaliana mutants in the sterol synthetic pathway. This is the SMT1, cephalopod mutant which encodes a defective SAM-cycloartenol
390
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
C24-methyltransferase [330–332], the fackel (fk) mutant which encodes a defective sterol C14 reductase
[333,334] and the hydra1 mutant which encodes a defective D8-D7-sterol isomerase [331,335,336]. The fackel
callus accumulated up to 76% of D8,14-sterols including mainly D8,14-sitosterol (28) (59%) consistent with
the pathway of Fig. 16 where the subsequent D8-D7-sterol isomerase is specific for D8-sterols. Accordingly,
unmetabolized D8,14-sterol substrates, and no D7,14-sterols, were recovered in microsomal preparation or
plant cell cultures containing both sterol C14 reductase and D8-D7-sterol isomerase in the presence of a specific inhibitor (15-azasterol) of the sterol C14 reductase [323,337]. The sterol profile of the fackel mutant
totally confirmed the high level of D8,14-sterols, and reduction of end-products of sterol synthesis, obtained
after treatment of plant cell cultures with an inhibitor of C14-reductase, 15-azasterol [337]. Although the
detailed sterol profile of the hydra1 mutant was not determined, it showed a dramatic decrease in campesterol and sitosterol concentrations similar to that obtained in the fackel mutant [333,334]. Genetic interactions between cephalopod (CPH), fackel and hydra1, revealed that fackel and hydra1 act sequentially, thus
confirming the previous biosynthetic data [323,337], while CPH acts independently of these genes to produce
essential sterols [331]. The SMT1 null mutant was shown to still produce alkylated sterols [330], that is
consistent with the previously observed overlap in the substrate specificity of sterol methyl transferases in
phytosterol synthesis [258,338,339].
Taken together, the data obtained recently by powerful genetic analysis are globally consistent with the previous enzymological data of Table 1 and Fig. 16. They suggest a more linear pathway for transformation of
the tetracyclic nucleus, and a more independent and parallel construction of the sterol side-chain in the case of
a non-perturbated sterol biosynthesis. The modeling of the sterol structure is not only controlled by the substrate selectivity of the enzymes involved, but also by their high product selectivity as shown by mechanistic
data on reaction course carried out in corn. Indeed, cleavage of the 9b,19-cyclopropane ring proceeds regioselectively to the D8(9)-double bond [340,341], demethylation at C14 produces a D14(15) double bond [322],
and the D8 double bond is regioselectively isomerized to the D7 position without formation of a D8(14) double
bond [325].
Beside this possible kinetic regulation of the plant sterol biosynthetic pathway, and in contrast to the situation in animals [342,343] and yeast [344,345], much less is known regarding how plant cells monitor their
sterol content and regulate sterol synthesis.
A series of previous works have addressed the rate-limiting nature of HMGR for sterol biosynthesis in
plants [328,346,347], and recently a feedback upregulatory effect occurring at the HMGR protein level in
response to a depletion of end-products was reported in plants [348]. Squalene synthetase (SQS), a putative
branch point diverting carbon from the central isoprenoid pathway to sterol has been investigated as a possible means by which sterol biosynthesis may be controlled. Data obtained suggested that the suppression of
SQS in elicitor-treated tobacco cells was regulated at several different levels: probable decreased transcription
of the SQS gene and no change in the absolute level of the SQS polypeptide, implicating a post-translational
control for this enzyme activity [349]. It has been proposed that the C-24 methylation of cycloartenol (SMT1)
is a major site of regulation in the sterol biosynthetic pathway [259,346,347,350]. While an Arabidopsis sterol
methyltransferase 1 (SMT1) deletion mutant has been shown to still accumulate alkylated sterols, but at
altered levels [330], overexpression of SMT1 in tobacco seed increased the amount of total sterols in seed tissue
by up to 44% and decreased cycloartenol and cholesterol levels by up to 53% and 34%, respectively. A concomitant increase in endogenous HMGR activity, presumably upregulated by reduced level of cycloartenol,
was also observed. Thereby SMT1 controls the flux of carbon into sterol biosynthesis in tobacco seed
[329,351].
Several investigations have also sought to determine the contribution of specific biosynthetic steps by selection of genetic mutants or by the use of genetic engineering approaches previously mentioned in this report
[330,331,333,334,336]. These studies have demonstrated the important role of sterols in plant growth and
development but have not directly addressed how plant sterol biosynthesis is regulated. However, they suggested that sterols other than brassinolides may serve as novel signals controlling cell fate during plant
embryogenesis [222,331,334].
It has been suggested that genes coding for enzymes making up a specific metabolon must have similar transcriptional networks to coordinate expression of the metabolic unit. Many enzymes involved in plant sterol
biosynthesis are coded by multiple genes [216] and genes situated downstream of cycloartenol are involved
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
391
only in sterol biosynthesis. The biological significance of these paralogs remains to be elucidated. In these
lines, study of the substrate specificities of the different SMO isoenzymes might reveal one possible role for
this multiplicity.
Finally, further works on the role of the protein in the mechanism of the enzymes committed to the different
reactions of the post-squalene sterol biosynthesis in plants, based on structural elucidation, mutants characterization and directed evolution, are required to enhance our understanding of these enzymes. These studies
will be facilitated by the possibility to express cDNAs encoding these plant enzymes in heterologous systems
allowing particularly the rigorous analysis of mutations responsible for altered enzymological properties. Precious clues could also be obtained from crystal structure of enzymes from bacterial sources such as the A. acidocaldarius hopene synthase [249] and the M. tuberculosis lanosterol 14-demethylase [352]. They could open
new ways for the design of novel or more selective biocides affecting key enzymes of sterol biosynthesis, a field
presenting a constant need for new compounds.
7. Biogenesis of mitochondrial isoprenoids
7.1. Ubiquinones
It can be inferred from previous studies that IPP is imported from the cytosol and used for ubiquinone
biosynthesis [353] according to a pathway not yet definitively established in plants, but probably very similar to that operating in Gram-negative bacteria and yeasts [354] (Fig. 17). Feeding experiments show that
the side-chain of ubiquinone is synthesized from mevalonate-derived IPP [355]. Although mitochondrial
FPPS [356] and GGPPS [85] isoforms have been characterized, specific prenyltransferases catalyzing the synthesis of the side-chain of ubiquinone have not been characterized. It has been proposed that ubiquinone-9
(UQ-9) is synthesized in the ER–Golgi membranes of spinach leaves [357] and transported in the mitochondria [358]. Intracellular transport of UQs is supported by the fact that the inability of yeast coq mutants to
grow on a non-fermentable carbon source was restored following the addition of UQ-6 in the medium [359].
Also, the uptake and transport of dietary UQ-8 to mitochondria prevented the growth arrest and sterility
phenotypes of Caenorhabditis elegans clk-1 mutants [360]. In this context, it is interesting to note that A.
thaliana possesses a plastidial and cytosolic solanesyl diphosphate synthases (SPPS) [109,361]. One could
suggest that the cytosolic SPPS could be involved in the biogenesis of the prenyl side-chain of UQ-9. This
hypothesis is strengthened by experiments in which p-hydroxybenzoate polyprenyl diphosphate transferase
(PHBPT), encoded by COQ2 was deliberately targeted to the ER of yeast and tobacco. Under these conditions, the UQ-6 content of transgenic yeasts and the UQ-10 content of transgenic yeasts and tobacco
plants were 3- and 6-fold increased. The same trend was observed in transgenic yeasts and tobacco expressing COQ2 properly targeted to the mitochondria [362]. The increased synthesis of UQ-10 in tobacco was
paralleled by an increased tolerance to high-salt concentration and to oxidative stress induced by methylviologen [362].
Following the prenylation of p-hydroxybenzoate (PHB) three C-hydroxylations, C- and O-methylations
and a decarboxylation steps are required (Fig. 17). How these individual steps are sequentially ordered in
plants is not known. Genetic and biochemical information obtained from bacteria and eucaryotes [354,363]
could now help clarify the sequence of the individual steps operating in plants. The gene encoding Arabidopsis PHBPT has been characterized [364]. PHBPT accepted geranyl diphosphate (GPP) as the allylic substrate and could direct in vivo the synthesis of UQ-6 and UQ-10, the main UQs that accumulate in S.
cerevisiae and S. pombe, respectively [364]. This indicated that PHBPT has broad substrate specificity
and suggests that the side-chain of ubiquinone is primarily determined by the specificity of the mitochondrial prenyltransferase generating the side-chain. The PHBT mRNA was predominantly expressed in the
flower and disruption of the gene in Arabidopsis resulted in the arrest of embryo development at an early
stage of zygotic embryogenesis. These results demonstrate that the A. thaliana PHBT gene involved in the
biosynthesis of mitochondrial UQ plays an essential role in embryo development in Arabidopsis [364], in
good agreement with previous data reported from animal studies [365,366]. The nature of the chain-length
of plant ubiquinone varies according to species. For instance ferns accumulate mostly UQ-10, while most
392
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
Tyrosine
COOH
OH
PHB
PP
H
n
PHBPT
PolyprenylPP
COOH
H
n
OH
PPHB
DECARBOXYLATION
HYDROXYLATION
METHYLATION
O
CH3
CH3O
CH3O
O
H
n
Ubiquinone
Fig. 17. Ubiquinone biosynthetic pathway. Enzyme abbreviation. PHBPT, p-hydroxybenzoate polyprenyl-diphosphate transferase.
Substrate abbreviations. PHB, p-hydroxybenzoate; PolyprenylPP, polyprenyl diphosphate.
other plants accumulate UQ-9 and UQ-10. In some plants, UQ-9 dominates while in others Q10 dominate.
UQ-8 is also detected as a minor homologue in some plants. In Arabidopsis the main homologue is UQ-9,
while in Capsicum organs, the main homologue is UQ-10, followed by minor UQ-11, UQ-9 and UQ-8
homologues [367]. Since the specificity polyprenyl diphosphate defines the UQ chain-length [368], one must
consider the existence of diverse mitochondrial prenyltransferases dedicated to the synthesis of mitochondrial side-chain containing C8, C9, C10 and C11 isoprene units. Concerning the steps post-PHB, one could
note that Arabidopsis AtCOQ3 homologous to yeast COQ3 has been cloned [369]. AtCOQ3 encodes 3,4dihydroxy-5-hexaprenylbenzoate methyltransferase.
7.2. Isoprenoid oxidation
It has been reported that plant mitochondria is involved in the oxidative catabolism of isoprenoids
in a pathway that requires in addition the interaction with plastids and microbodies [370–372]. This
pathway is reminiscent of the modified b-oxidation pathway that operates in Pseudomonas citronellolis
[373].
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
393
8. Biogenesis of plastid isoprenoids branching from GGPP
8.1. Diterpene biogenesis
The plastid is the main if not the unique site for the synthesis of GGPP that constitutes the backbone for the
synthesis of diverse diterpenes. However, it must be mentioned that in the case of irregular diterpenes such as
anisotomenes, two GPPs are used instead of one GGPP [79] (Fig. 4). The initial steps of GGPP transformation
are catalyzed by diterpene cyclases and is usually followed by further modification in the ER–cytosol. The gibberellin pathway serves to illustrate this point (Fig. 18). Diterpene cyclases have been classified into cyclases A
Fig. 18. Gibberellin biosynthetic pathway and compartmentation. Enzyme abbreviations. CPS, ent-copalyl diphosphate synthase; KS, entkaurene synthase; KO, ent-kaurene oxidase; KAO, ent-kaurenoic acid oxidase; GA12S, GA12-aldehyde synthase; GA7OX, GA 7-oxidase;
GA13H, GA 13-hydroxylase. Substrate abbreviation. GGPP, geranylgeranyl diphosphate.
394
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
and B [374]. In type B, a N-terminal conserved aspartate motif E/DxD(D,N) is involved in the protonationinitiated cyclization of GGPP. In A type cyclase, a C-terminal conserved aspartate DDXXD motif is involved
in the diphosphate ionization-initiated cyclization of the substrate [374,375]. Thus in the plant gibberellin
pathway, the consecutive action copalyl diphosphate synthase (CPS) (B-type cyclase) and ent-kaurene synthase (KS) (A-type cyclase) are required to convert GGPP into kaurene (Fig. 18). On the other hand, in fungi
that produce gibberellin, both activities reside on a single bifunctional protein (CPS/KS) that has CPS and KS
[376,377]. Domains A and B are present in abietane synthase involved in the synthesis of resin acids. Interestingly, abietane synthase catalyzes two consecutive and mechanistically different cyclizations [375,378]. Taxadiene synthase [379] and levopimaradiene [380] that catalyze the first committed step of taxol and ginkgolide A
have both domains.
With respect to the gibberellin pathway, ent-kaurene is further oxidized by a plastid envelope ent-kaurene
oxidase (KO), a cytochrome P450 mono-oxygenase [381], that establishes a link between the plastid and the
endoplasmic reticulum–cytosol for further modifications by membrane-bound cytochrome P450 mono-oxygenases and soluble or cytosolic 2-oxoglutarate-dependent dioxygenases (Fig. 18). The Arabidopsis genome
contains only one gene encoding each of the diterpene biosynthetic enzyme CPS, KS and KO, whereas in
the rice genome 3 CPS, 8 KS and 5 KO have been identified [382]. Based on expression and mutant analyses,
it has been established that as in Arabidopsis, rice CPS, KS and KO specifically dedicated to gibberellin synthesis are encoded by single genes. This suggests that the remaining rice genes encode enzymes involved in the
biosynthesis of rice phytoalexins [382]. Indeed, fifteen phytoalexins have been identified in rice and, with the
exception of sakurane, a flavonoid, all phytoalexins belong to the diterpene family [383]. Indeed, in suspension-cultured rice cells, diterpenoid phytoalexins were produced in response to UV and exogenously applied
elicitors [384] according to a mechanism involving transcriptional expression of diterpene cyclases [385]. Thus,
it has been shown that in rice, CPSent1 and CPSent2 were dedicated, respectively, to gibberellin and defensive
phytoalexin synthesis [386,387]. In a similar vein, one could note that two kaurene synthase genes KS1 and
KS2 have been characterized in Stevia rebaudia leaves [388]. This phenomenon is reminiscent of the active biosynthesis of diterpene steviol glycosides, that accumulate in the leaves [388]. Similarly, it could be anticipated
that the expression of specific diterpene cyclases accounts for the increased synthesis of diterpene antioxidants
in plants belonging to the Labiate such as sage (Salvia officinalis) and rosemary (Rosmarinus officinalis) [389].
It has been shown that their synthesis is induced by light and drought stress [390,391]. The same mechanism
could apply for 11E,13E-k-11,13-diene-8a,15-diol (WAF-1), an activator of MAPK cascade, involved in plant
defenses [392].
Finally, either in the case of gibberellins or non-gibberellin-related diterpenoids, the later steps of diterpene
synthesis often involve modifications of plastid products by cytosolic enzymes. How this export is mediated is
not known. Potential plant diterpene ABC transporters have been identified in Nicotiana plumbaginifolia
[393,394]. However presently identified proteins are located in the plasmalemma.
8.1.1. Reduction of geranylgeranyl chain to phytyl chain
In plastids, GGPP is in channeled toward the synthesis of the phytyl side-chain of several prenyllipids
including chlorophylls, tocopherols and phylloquinone (Fig. 19). The reduction of the geranylgeranyl chain
to phytyl chain is catalyzed by geranylgeranyl reductase (GGR). The bacterial and cyanobacterial and plant
genes encoding GGR were identified previously [395–397]. Consistent with its role in the reduction of the geranylgeranyl chain [398], deletion of GGR induced the accumulation of GG-prenylated chlorophyll (GG-Chl)
[396,397]. We have shown that GGR sequentially catalyzes the reduction of geranylgeranyl-prenylated chlorophyll (GG-Chl) a into phytyl-Chla as well as the reduction of free GGPP into phytyl-PP [399]. In plastids,
the same geranylgeranyl reductase is recruited into the chlorophyll, the tocopherol and the phylloquinone
pathways (Fig. 19). The GGR gene is up-regulated during etioplast to chloroplast and chloroplast to chromoplast development [399]. In contrast to presently known GGRs, Rhosdospirillum rubrum GGR only catalyzes
the reduction of geranylgeranyl-prenylated pheophytin into phytyl-prenylated pheophytin [400].
The expression of an antisense GGR construct in tobacco resulted in a decreased content of tocopherol and
chlorophyll (Chl). It was observed that up to 58% of the Chls were mainly in the form of GG-Chl [401]. The
deficiency of tocopherol was compensated by a selective accumulation of xanthophyll cycle carotenoids [402].
The photosynthetic capacity of these transgenic plants was not altered under optimal growing conditions, thus
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
395
Homogentisate
PLASTID
CH2OPP
TOCOTRIENOL
GGPP
R = Chlorophyllide
O
GG-Chl
GGR
R
CH2OPP
GGR
GGR
O
DihydroGG-Chl
R
CH2OPP
GGR
GGR
O
TetrahydroGG-Chl
R
PhytylPP
O
CHLOROPHYLLS
TetrahydroGGPP
CH2OPP
GGR
Phytyl-Chl
DihydroGGPP
R
Homogentisate
TOCOPHEROL
1,4-Dihydroxy2-naphthoate
PHYLLOQUINONE
Fig. 19. Roles of geranylgeranyl reductase. The steps catalyzed by geranylgeranyl reductase yield the prenyl side-chain of chlorophylls,
tocopherols, tocotrienols and phylloquinone as indicated. Enzyme abbreviation. GGR, geranylgeranyl reductase. Substrate abbreviations.
GGPP, geranylgeranyl diphosphate; GG-Chl, geranylgeranyl-chlorophyll; DihydroGG-Chl, dihydrogeranylgeranyl-chlorophyll; TetrahydroGG-Chl, tetrahydrogeranylgeranyl-chlorophyll; Phytyl-Chl, phytyl-chlorophyll; DihydroGGPP, dihydrogeranylgeranyl diphosphate; TetrahydroGGPP, tetrahydrogeranylgeranyl diphosphate; PhytylPP, phytyl diphosphate.
suggesting that the geranylgeranyl side-chain had no influence on the transfer of light energy [403]. A very similar situation has been described using the rice mutants M249 and M134 [404,405]. The rice mutants were
devoid of tocopherol, and accumulated GG-Chls and phylloquinones prenylated with GG and partially
reduced GG [404]. In a similar vein, the inactivation of GGR in Synechocystis sp. PCC 6803 resulted in the
accumulation of GG-Chl, and low amount of tocotrienol instead of tocopherol [406]. Also, the mutant could
not grow photoautotrophically due to instability and rapid degradation of the photosystems in the absence of
added glucose [406]. The influence of the side-chain of chlorophylls seems also marked for the organization of
antenna complexes LHC1 and LHC2 in purple photosynthetic bacteria. For instance in the GGR mutant of R.
sphaeroides, that produces GG-bacteriochlorophyll (GG-Bchl) in place of phytyl-Bchl, the assembly of the
antenna complex LHC2 was altered [396]. These changes are usually explained by the fact that phytyl sidechain is more flexible, or less rigid than the unsaturated geranylgeranyl chain. In addition to being involved
in the reduction of GG-Chl, it has been demonstrated that R. sphaeroides GGR could also reduce the side GGprenylated bacteriopheophytin (GG-Ph) into phytyl-Ph [400]. When transposed to higher plants, and more
generally to organisms possessing two photosystems, these data strongly suggest the phytyl side-chain of phytyl-pheophytin is formed directly via GGR [400]. In other words, phytyl-pheophytin is not only formed via
magnesium de-chelation of preformed chlorophylls.
8.2. Carotenoid biogenesis
8.2.1. Phytoene backbone and mechanistically related irregular isoprenoids
The conversion of geranylgeranyl pyrophosphate into phytoene is catalyzed in two distinct steps by phytoene synthase (PS). Previous studies have shown that phytoene synthase from tomato [407] and pepper [408] are
integrated into a high molecular mass complex of 190–200 kDa. PS has been characterized as, a bifunctional
enzyme [409]. The first catalytic step involves the abstraction of a diphosphate of a GGPP donor and is
396
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
followed by a 1–1 or head to head condensation with a GGPP acceptor, thus giving a stable prephytoene PP
(PPPP). During the second step, PPPP undergoes a complex concerted rearrangement involving removal of
the diphosphate group and a carbocation neutralization (Fig. 20). Plant [409,410] and bacterial PS [411]
are strictly dependent upon Mn2+and are stimulated by neutral detergents [411]. The reaction catalyzed by
PS and squalene synthetase (SQS) are essentially identical, except that SQS requires Mg2+ for activity and catalyzes in addition the NADPH-dependent reduction of the central bond of dehydrosqualene to give squalene.
This mechanistic similarity is reflected by the fact that SQS produces dehydrosqualene, a C30 phytoene analogue, when incubated in the presence of Mn2+ and in the absence of NADPH [412] or in the presence of dihydroNADPH, an unreactive analogue [413]. Also in the absence of NADPH and in the presence of Mg2+, SQS
produces 12-hydroxysqualene and 10-hydroxybotryococcene in addition to dehydrosqualene [414]. No obvious sequence similarity to any known NADPH binding site could be found in the peptide sequence of SQS.
This led to the proposal that the NADPH-binding site is fulfilled by the most flexible part of SQS, i.e., the J–K
loop (VKIRK) and part of the K helix which are not present in phytoene synthase [415]. These data point to
convergent evolution between SQS and PS. Interestingly, Staphylococcus aureus and phototrophic bacteria
(Heliobacillus mobilis, Heliophilum fasciatum, Heliobacterium chlorum, Heliobacterium modesticaldum, and
Heliobacterium gestii) synthesize C30 carotenoids [416,417]. Available data from S. aureus indicate that these
carotenoids derive from diapophytoene, a C30 homologue of phytoene. Diapophytoene results from the headto-head condensation of two FPPs through a reaction catalyzed by diapophytoene synthase [418]. Indeed,
data gained from bacterial diapophytoene synthase using mutagenic libraries and domain-swapping revealed
OPP
GGPP
+
OPP
CH2OPP
PS
H
H+ PPi
+
GGPP
H PPi
CH3
PPPP
PS
Phytoene
CPPS
OPP
+
DMAPP
OPP
DMAPP
OPP
FPP
+
OPP
FPP
OPP
H+ PPi
CPP
OPP
SQS
PSPP
PPi
NADPH
NADP+
NADPH
SQS
PPi
aPS
BS
PPi
Squalene
NADP+
PPi
Botryococcene
Dehydrosqualene = Diapophytoene
Fig. 20. Biosynthesis of phytoene and mechanistically related isoprenoids. Enzyme abbreviations. PS, phytoene synthase; CPPS,
chrysanthemyl diphosphate synthase; SQS, squalene synthetase; aPS, apophytoene synthase; BS, botryococcene synthase. Substrate
abbreviations. CPP, chrysanthemyl diphosphate; DMAPP, dimethylallyl diphosphate; FPP, farnesyl diphosphate; GGPP, geranylgeranyl
diphosphate; PPPP, prephytoene diphosphate; PSPP, presqualene diphosphate.
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
397
that there is much more constraints to convert phytoene synthase into diapophytoene synthase than the
reverse. For instance a single change of Phe26 by non-aromatic amino acid residue in the peptide sequence
of diapophytoene allowed diapophytoene synthase to produce the C40 phytoene [419]. It has been also shown
from E. coli complementation that diapophytoene synthase has low phytoene synthase activity, while phytoene synthase has no diapophytoene synthase activity [420]. On the other hand, the desaturases seem more flexible since phytoene desaturase and diapophytoene desaturase could use phytoene as well as diapophytoene
[420]. Because the active-site of SQS is organized into a-helices [415], like other trans-prenyltransferases
and terpene cyclases, one may suggest that PS also adopt the same folding [96].
Mechanistically the 1 0 –1 fusion between GGPP or FPP to give phytoene or squalene could be envisioned
for the biosynthesis of irregular or non-head-to-tail derived isoprenoids [77]. This class of compounds include
the monoterpene chrysanthemic ester which is a typical component of natural pyrethrin insecticide extracted
from Chrysanthemum cinerarnaefolium flower heads [421] (Fig. 20). Also, botryococcene, a component of the
hydrocracking oil produced by the green alga Botryococcus brauni [422] belongs to this class.
The initial step of chrysanthemic ester is catalyzed by chrysanthemyl diphosphate synthase (CPPS) which
condenses two DMAPPs into CPP. Surprisingly, CPPS has no homology to SQS and PSY, but displayed
extended homology to FPPS (up to 80%) [423]. However, there are two main differences between CPPS and
the orthodox FPPS family of prenyltransferases. CPPS possesses a plastid targeting sequence, thus suggesting a new role of C. cinerarnaefolium chromoplasts in the synthesis of pyrethrins. In this connection, it is
worth noting that a maize sesquiterpene synthase displays a transit peptide-like motif [424] and that the
immunochemical localization of FPPS in rice, wheat and tobacco chloroplasts has been reported [425].
In addition, the second aspartate-rich domain DDXXD of FPPS is changed into NDXXD in CPPS
[423]. These above characteristics are shared with Artemisia tridentata CPPS. When incubated with
DMAPP, A. tridentata CPPS produced CPP and lavandulyl diphosphate. However, A. tridentata CPPS catalyzed the synthesis of lavandulyl diphosphate and geranyl diphosphate in the ratio: 4:1:1 when incubated
with IPP and DMAPP [117]. Further studies using chimeric FPPS and CPPS substantiated the mechanistic
relationship between the cyclopropanation and the chain elongation activities of prenyltransferase [116] and
this led to suggest that phytoene and squalene synthetase emerged from an ancestry chain elongation
enzyme [116].
Apparently the lavandulyl chain could be formed according to another mechanism as shown in synthesis of
sophoraflavanone G, a branched monoterpenoid-conjugated flavanone characteristic to S. flavescens [426].
Here, two plastidial prenyltransferases were involved. The first condenses DMAPP at position 8 of leachianone G and the second prenyltransferase transfers a second DMAPP to the 2 0 position of the dimethylallyl
group attached at position 8 of leachianone G to form sophoraflavanone G [426].
Several lines of evidence based on feeding [427] and in vitro studies [428] suggest that botryococcene is synthesized from FPP by an enzyme similar to SQS or phytoene synthase. The reaction proceeds through the condensation of two FPPs and is dependent upon Mg2+ and NADPH or NADH. Botryococcene synthesis is
inhibited by squalestatin like SQS. In addition, botryococcene synthase activity is stimulated by Tween 80,
as shown in previous studies with SQS [429] and PS [430].
Phylogenetic tree analysis reveals that higher plant phytoene synthases are evolutionary related to algae and
cyanobacteria phytoene synthase (Fig. 21). On the other hand, plant SQSs cluster with diapophytoene
synthase.
Arabidopsis possesses only one PS gene, while tomato and tobacco have two genes. The duplication of PS
gene seems more widespread in plants belonging to the poaceae and both genes are functional [431]. Finally,
available evidence implicates PS as a primary determinant in the control of carotenoid accumulation in sink
tissues [432–437].
8.2.2. Phytoene desaturation
In oxygenic photosynthetic organisms, the desaturation of cis-phytoene to trans-lycopene involves four
steps that are catalyzed by phytoene desaturase (PD) and f-carotene desaturase (Fig. 22). This situation constrasts with that occurring in non-photosynthetic bacteria and fungi that use only one desaturase CrtI. However more recently a CrtI-type desaturase has been characterized from the cyanobacterium Gloeobacter
violaceus PCC 7421 [438]. Thus G. violaceus is the first oxygenic photosynthetic organism possessing a
398
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
Plant
phytoene synthases
Mt
Os
Pa
Dc
Np
Te
Ha
Ca
Le
At
Hb
Zm
Db
No
Bl
Sn1
Sn2
Algae
Cyanobacteria
phytoene synthases
Rc
Cha
SadPS
Tt
CaSQS
NtSQS
0.05
AtSQS
Plant
squalene synthetases
Fig. 21. Unrooted phylogenetic tree of phytoene synthases and related homologues. Alignments were performed with ClustalX and
visualized with TreeView. Distances between sequences are expressed as 0.05 changes per amino acid residue. Accession Numbers and
abbreviations. Dc (BAA84763); Os (AAK07735); Np (S54135); Te (AAM45379); Ha (CAC19567); Ca (CAA48155); Le (CAA42969); At
(At5g17230); Zm (S68307); Db (T10702); No (NP_485873); Sn1 (NP_441168); Sn2 (CAA45350); Cha (ZP_00359166); SadPS
(CAA52097.1); AtSQS (P53799); NtSQS (U60057); CaSQS (AF124842); Tt (YP_006040); Rc (P17056); Bl (AAF65581); Hb
(NP_280449); Pa (BAA14128); Mt (NP_276914.1). At, Arabidopsis thaliana; Bl, Brevibacterium linens; Ca, Capsicum annuum; Cha,
Chloroflexus aurantiacus; Db, Dunaliella bardawil; Dc, Daucus carota; Ha, Helianthus annuus; Hb, Halobacterium sp.; Le, Lycopersicum
esculentum; Mt, Methanothermobacter thermoaututrophicus; No, Nostoc sp.; Np, Narcissus pseudonarcissus; Nt, Nicotiana tabacum; Os,
Oryza sativa; Pa, Pantoea ananatis; Rc, Rhodobacter capsulatus; Sa, Staphylococcus aureus; Sn1, Synechocystis sp.; Sn2, Synechococcus sp.;
Te, Tagetes erecta; Tt, Thermus thermophilus; Zm, Zea mays.
CrtI-type desaturase and that does not need a carotenoid isomerase which is apparently present in the G. violaceus genome [438]. This may be related to the fact that G. violaceus belongs to an early group of the cyanobacterial phylum. This phylum is characterized by the fact that the photochemical reaction takes place in the
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
399
PLASTID
Geranylgeranyl diphosphate
PS
PD
GGPP
Phytoene
PD
Phytofluene
ZD
ζ-Carotene
ZD
CRTISO
Neurosporene
Pro-Lycopene
Lycopene
Fig. 22. Pathway of phytoene desaturation. Enzyme abbreviations. PS, phytoene synthase; PD, phytoene desaturase; ZD, f-carotene
desaturase; CRTISO, carotenoid isomerase. Substrate abbreviation. GGPP, geranylgeranyl diphosphate.
periplasmic membrane due to the lack of thylakoid membranes [439]. One may envision that a very similar
situation exists in the case of lycopene cyclase, because none of the known cyanobacterial lycopene cyclase
genes [440,441] are observed in the entire genome of G. violaceus PCC 7421 [439].
The situation is different in non-oxygenic photosynthetic organisms such as Rhodobacter capsulatus and R.
sphaeroides and in non-photosynthetic organisms such as fungi. In general these organisms possess a PD that
could lead to lycopene. However in R. capsulatus and R. sphaeroides, the reaction yields mainly neurosporene en route to the synthesis of the acyclic xanthophyll spheroidenone. However exceptions could be
observed in the case of the green sulfur bacterium Chlorobium tepidum, a strict anaerobe and an obligate
photoautotroph. Here, the PD pathway is similar to that of cyanobacteria and plants [442]. On the other
400
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
hand, in Neurospora five desaturations lead to the formation of 3,4-dihydrolycopene, an intermediate of
torulene synthesis [443].
The reaction sequence 15-cis-phytoene to trans-lycopene obviously proceeds via an isomerization step. The
point at which in the biosynthetic sequence the cis to trans isomerization occurs and how this transition is
accomplished is not entirely resolved. Earlier studies have shown that several intermediates accumulate during
in vitro studies. In particular, cis- and trans-phytofluenes, cis- and trans-f-carotenes were formed under these
conditions [444]. Also, it is well known that carotenoids possessing the trans-configuration are selected for
light harvesting complexes while photoactive carotenoids having cis-configuration are usually incorporated
into the reaction centers of photosynthetic organisms [445]. The stage at which this isomerization occur is also
not known.
Etioplasts of the Arabidopsis crtiso mutant accumulates poly-cis-carotenes in the dark and do not differentiate prolamellar bodies [446]. Similarly, tomato crtiso previously known as tangerine mutant accumulates prolycopene [447]. CRTISO encodes a carotenoid isomerase catalyzing the isomerization of prolycopene to
lycopene (Fig. 22) [446,448]. CRTISO catalyzed in vitro the conversion of 7,9,9 0 -tri-cis-neurosporene to 9 0 cis-neurosporene and 7 0 ,9 0 -di-cis-lycopene into all-trans-lycopene [448]. When grown under light-activated
heterotrophic conditions, the crtiso mutant of Synechocystis sp. PCC 6803 accumulated mainly cis-lycopene
and traces of all trans-carotenoids including b-carotene and xanthophylls. Concomitantly, the D1 protein
of the PSII was not detected and the PSII activity could not be measured [449]. Under continuous illumination
the capacity to synthesize b-carotene and xanthophylls was restored as well as the PSII activity. This led to the
conclusion that carotene isomerization was required for the assembly of PSII [449]. When R. sphaeroides, an
organism devoid of CRTISO, was used for reconstitution experiments with trans-spheroidenone, the xanthophyll was converted to the 15,15 0 -cis configuration. This change occured without any exogenous isomerase
and in particular in the absence of a CRTISO [450]. Detailed analysis suggested that spontaneous thermal
isomerization of trans-spheroidenone into 15,15 0 -cis-spheroidenone is required before the binding process
[450].
In plants, the desaturation of phytoene is a target for photobleaching herbicides including pyridazinones,
pyridinecarboxamides and phenoxybutanamides [451]. On the other hand, the phytoene desaturase from Pantoea that carries the four desaturation steps between phytoene and lycopene is not sensitive to the bleaching
herbicide Norflurazon [452]. A mutated PD from Synechococcus PCC 7942 (mutant NFZ4) that conferred
resistance to Norflurazon was introduced in tobacco. This conferred 58-fold and 3-fold resistance to the
bleaching herbicide Norflurazon and fluridone [452]. This was paralleled by an increased stability of the PSII
protein D1. On the other hand, the plants were not resistant to f-carotene desaturase inhibitors [453]. In an
attempt to control the proliferation of invasive aquatic weed, a fluridone-resistant Hydrilla verticillata, has
been naturally selected. The resistance has been ascribed to independent mutations of a conserved Arg residue:
Arg304Ser, Arg304Cys, Arg304His [454,455]. The mechanism of phytoene desaturation and inhibition is largely unknown. Available data implicate plastoquinone-9 as a component of the carotene desaturase system
[456,457]. Based on kinetic studies, it has been shown that the bleaching ketomorpholines do not act competitively with respect to phytoene [458]. On the other hand, it has been suggested that the bleaching herbicide
Norflurazon acts competitively with respect to the plastoquinone cofactor on its PD binding site [459].
8.2.3. Lycopene cyclization
Lycopene cyclase (LC), catalyzes the cyclization of trans-lycopene at both ends [460,461] (Fig. 23). There is
practically no similarity between the peptide sequences of LC belonging to different taxa. LC can be grouped
into three main types. Type-1 LCs (or plant-type LCs) possess a pyridine nucleotide binding motif whose function is unknown [462]. It has been previously reported that NAD(P)H is required for the cyclization reaction
in vitro [463]. However, these data showed that the hydrogen atom introduced at the position C(2) during the
cyclization is derived from water and not from NADPH [463]. Thus no hydrogen was transferred from
NADPH to the substrate. Type-1 LC can be further subdivided into lycopene b-bicyclases (LCB), lycopene
e-bicylase (LCE), lycopene b-monocyclase and lycopene e-monocyclase and bifunctional lycopene b,e-bicyclase (Fig. 24). Type-2 LCs include the fungal enzymes. Here, lycopene cyclase is a bifunctional protein that
has phytoene synthase activity at the carboxy terminus and lycopene cyclase activity at the amino terminus
[464]. Type-3 LCs have been characterized from Gram-positive bacteria-like Brevibacterium linens. The lyco-
HO
HO
HO
Lutein 5',6'-epoxide
Loroxanthin
Lactucaxanthin
ε-Carotene
O
CH2OH
OH
OH
OH
HO
HO
HO
LCE
CHBNHDIM
Capsorubin
OH
O
Capsanthin 5,6-epoxide
O
Capsanthin
Lutein
Zeinoxanthin
α -Carotene
δ -Carotene
CCS
CHEHBP450
(CHBHBP450)
LCB
LCE
O
OH
O
O
OH
OH
HO
HO
HO
CCS
HO
OH
HO
OH
OH
O
LCB
CCS
Lycopene
NSY
ZEP
ZEP
CHBNHDIM
CHBNHDIM
LCB
O
O
Neoxanthin
O
Violaxanthin
VDE
Antheraxanthin
VDE
Zeaxanthin
(CHBHBP450)
β-Cryptoxanthin
(CHBHBP450)
β-Carotene
γ -Carotene
OH
OH
OH
KETCrtW
KETCrtO
OH
KETCrtW
KETCrtO
HO
HO
O
O
O
O
KETCrtW
CHBNHDIM
KETCrtW
O
Echinenone
KETCrtO
Astaxanthin
O
Adonixanthin
O
Adonirubin
O
Canthaxanthin
KETCrtO
OH
CHBNHDIM
OH
OH
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
401
Fig. 23. Xanthophyll biosynthetic pathway. Enzyme abbreviations. LCB, lycopene b-cyclase; LCE, lycopene e-cyclase; CHBNHDIM
carotenoid b-hydoxylase non-heme-di-iron monooxygenase; CHBHBP450 carotenoid b-hydoxylase heme-binding cytochrome P450;
CHEHBP450 carotenoid e-hydoxylase heme-binding cytochrome P450; KETCrtW/CrtO b-carotene ketolase; ZEP, zeaxanthin epoxidase.
VDE, violaxanthin de-epoxidase; NSY neoxanthin synthase; CCS, capsanthin-capsorubin synthase.
402
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
Bacterial β-cyclases
AaCrtY
Bacterial β-monocyclase
FCrtYm
PaCrtY
Plant ε-monocyclases
Bacterial β-monocyclase
ApLCE
ReCrtLm
TeLCE
LsLCE
Plant β-cyclases
AtLCE
NpLCB
NtLCB
LeLCE
LeLCB
CaLCB
Pm1LCB
AtLCB
Bifunctional
β-ε-cyclases
CtCCS
Pm2LCB
StNSY
LeNSY
Pm2LCE
CaCCS
Plant k-cyclases
SnLCB
DeCrtLm
Bacterial β-monocyclase
Pm1LCE
Cyanobacterial
cyclases
β-bicyclases
0.1
Fig. 24. Unrooted phylogenetic tree of lycopene cyclases. Alignments were performed with ClustalX and visualized with TreeView.
Distances between sequences are expressed as 0.1 changes per amino acid residue. Accession Numbers and abbreviations. ApLCE
(AF321535_1); TeLCE (AAG10428); LsLCE (AF321538_1); AtLCE (Q38932); LeLCE (O65837); Pm1LCB (NP_893181.1); Pm2LCB
(CAE21298); Pm2LCE (NP_895600.1); Pm1LCE (CAE19092.1); SnLCB (Q55276); DeCrtLm (AAF10377.1); CaCCS (Q42435); LeNSY
(CAB93342.1); StNSY (CAB92977.1); CtCCS (Q9SEA0); AtLCB (AAA81880); CaLCB (Q43415); LeLCB (Q43503); NtLCB (Q43578);
NpLCB, (Q40424); ReCrtLm (AAR98749); PaCrtY (BAA14126.1); AaCrtY (P54974); FCrtYm (BAC77673.1). Aa, Agrobacterium
aurantiacum; Ap, Adonis palaestina; At, Arabidopsis thaliana; Ca, Capsicum annuum; Ct, Citrus sinensis; De, Deinococcus radiodurans; F,
Flavobacterium sp.; Le, Lycopersicon esculentum; Ls, Lactuca sativa; Np, Narcissus pseudonarcissus; Nt, Nicotiana tabacum; Pa, Pantoea
ananatis; Pm, Prochlorococcus marinus; Re, Rhodococcus erythropolis; Sn, Synechococcus sp.; St, Solanum tuberosum; Te, Tagetes erecta.
pene cylase activity is the result of two distinct genes crtYc and crtYd encoding proteins of 125 and 107 amino
acids [465]. These two separable subunits are probably organized as a heterodimer. In archaea, the N-terminal
and C-terminal halves of LC correspond to the subunits crtYc and crtYd of type-3 LCs, as shown in Sulfolobus solfataricus [466] and Halobacterium salinum [467]. The apparent flexibility of lycopene cyclase probably
deserves further variation, since, sequence homologous to known plant or cyanobacterial LC could not be
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
403
identified in the sequenced genome of several cyanobacteria including Synechocystis sp. strain PCC 6803, T.
elongatus, Nostoc punctiforme and Anabaena sp. Strain PCC7120, G. violaceus and Trichodesmium erythraeum
[441,442]. Also, the genome of the green sulfur phototrophic bacterium C. tepidum did not encode any homolog of known types of LCs [442]. This led to the suggestion that the CrtU homolog that encodes a chlorobactene synthase could fulfill this role [442]. This suggestion is reinforced by the fact, that cyanobacteria that
possess an identifiable lycopene cyclase such as Prochlorococcus marinus and Synechococcus sp. (strain
WH8102) lack homologs of CrtU [442]. All lycopene cyclases, irrespective of class, proceed via a carbocationic
mechanism [468]. The mechanism is shared by capsanthin-capsorubin synthase [469]. Indeed, the primary
structures of LC and CCS display 55% identity.
In higher plants, the formation of a-carotene requires the interaction of the lycopene b-bicyclase and the
lycopene e-monocyclase [470]. In lactuca sativa a single lycopene bicyclase catalyzes the synthesis of e-carotene, the precursor of lactucaxanthin [471]. Unlike higher plants, in Prochlorococcus marinus MED4 only a
single lycopene bicyclase catalyzes the formation of the b-carotene, a-carotene and e-carotene [441,472]. Functional studies revealed that the capacities to synthesize one or two e-ring are specified by active amino acid
residues located in the C-terminal domain of lycopene cyclase [471].
In Synechocystis, the lycopene b-monocyclase could be involved in the biosynthesis of the monocyclic
myxol dimethyl-fucoside. This is reinforced by the fact that a marine bacterium, strain P99-3 produces myxol
and possesses a lycopene b-monocyclase [473]. Genes encoding lycopene b-monocyclases that convert lycopene into c-carotene have also been characterized from a marine bacterium, strain P99-3 [473], Rhodococcus
erythropolis and Deinococcus radiodurans [474]. The most important diversity of LCs was encountered in
cyanobacteria, belonging to the Prochlorococcus genus. Indeed, these cyanobacteria display a guenine isoprenoid metabolism due to the presence of chlorophyll b in addition to chlorophyll a and of a-carotene (63%),
b-carotene (11%), e-carotene epoxide (1%), d-carotene (1%) b-cryptoxanthin (3%), 3-hydroxy-a-carotene
(4%), zeaxanthin (16%) [441,472].
By virtue of its carbocationic mechannsim, LC activity is strongly inhibited by amine derivatives positively
charged at physiological pH [468,475–478].
8.2.4. Carotene to xanthophyll conversion
Two types of genes encoding oxygenases converting carotene into xanthophylls were first cloned in plants.
These include b-carotene hydroxylases (CHB) and b-carotene ketolases (KET). Two genes encoding b-carotene
hydroxylases (CHB1, CHB2) that convert b-carotene into C3-hydroxy-b-xanthophylls were first cloned from
Arabidopsis [479] and pepper [395] (Fig. 23). Genes encoding two structurally divergent ketolases that introduce
a keto group in the C4 of carotenoid rings have also been cloned. These include the non-phylogenetically related
CrtO and the CrtW genes (Fig. 25). The genome sequence of Nostoc sp. PCC7120 contains both CrtO and
CrtW homologues. The CrtO type displays six conserved domains, including a putative FAD-binding domain
[480]. These enzymes could function as monoketolase as shown for Synechocystis [481] or as mono and diketolase [480]. The CrtW ketolase could function as monoketolase or diketolase. These enzymes are found in
Agrobacterium aurantiacum [482] and also in H. pluvialis [483] and N. punctiforme PCC 73102 [484]. Blast search
revealed the presence of a gene homologous to H. pluvialis b-carotene ketolase in Chlamydomonas reinhardtii.
Although this organism did not produce astaxanthin under trophic changes [485], the putative ketolase corresponded to an expressed gene, encoding a protein with a C terminal extension of 115 amino acid not observed in
currently characterized ketolase [485]. However, C. nivalis under low temperatures and high level of UV irradiation produced astaxanthin as a natural screen and thus remain photosynthetically active [486,487]. Based on
this evidence the putative C. reinhardtii could encode a true ketolase. Alternatively, due to the mechanistic similarity between CHB and KET emphasized below, the putative KET could catalyze an hydroxylation steps,
possibly in the pathway leading to loroxanthin. Blast search also revealed that a putative non-heme-di-iron
CHB is present in the genome of the archaea S. solfataricus (Accession No. NP_344225).
It has been recognized that pepper CHB and algal KET of the CrtW types display similar histidine motifs:
HX3H, HX2HH, HX4H, HX2HH for CHB and HX3H, HX2HHXH, HX2HH for KET [395]. These histidine
motifs are in relative positioning with respect to hydrophobic domains of the proteins in plant and bacterial
CrtZ-type CHB [395]. Site-directed mutagenesis of residues within the conserved histidine motifs completely
abolished the hydroxylase activity of pepper CHB [395]. This provided the experimental support that the
404
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
Bacterial phytoene
desaturases
Cyanobacterial non-hemedi-iron hydroxylases
PsCrtI
ReCrtI
Ketolases type-2
SnCrtO
Np2CrtO
MaCrtI
Cm
Np1CrtO
Sn1
Pm1
GvCrtO
Pm2
DrCrtO
Tse
Sn2
ReCrtO
ACrtZ
HpCrtW
FCrtZ
BaCrtW
Ps
PanCrtZ
ACrtW
PaCrtZ
PmCrtW
Pp
Le2
NpCrtW
NCrtW
Non-heme-di-iron
ketolases type-1
Np
Cs
Os
Te
At2
At1 Vv
Ca1
Le1
Ca2
Plants
0.1
Bacteria
Non-heme-di-iron
hydroxylases
Fig. 25. Unrooted phylogenetic tree of hydroxylases and related homologues. Alignments were performed with ClustalX and visualized
with TreeView. Distances between sequences are expressed as 0.1 changes per amino acid residue. Accession Numbers and abbreviations.
Le2 (CAB55626); Ca1, (CAA70427.1); Le1 (CAB55625); Ca2 (CAA70888); Vv (AAM77007); At1 (NP_194300.1); At2 (BAA98075.1); Te
(AAG10430); Os (NP_922503); Cs (CAC95130); Np (CAC06712); NCrtW (NP_487229); NpCrtW (ZP_00111258); PmCrtW (CAB56059);
ACrtW (Q44261); BaCrtW (AAN86030.2); HpCrtW (Q39982); Sn2 (NP_440788); Tse (NP_682690); Pm2 (CAE21991); Pm1 (AAP99312);
Sn1 (CAE06806); Cm (NP_848964); MaCrtI (CAB94794.1); ReCrtI (AAW23161.1); PsCrtI (AAN85599.1); SnCrtO (S76617); Np2CrtO
(ZP_00111616); Np1CrtO (ZP_00112486.1); GvCrtO (BAC88335); DrCrtO (E75561); ReCrtO (AAW23159.1); ACrtZ (Q44262); FCrtZ
(AAC44852); Ps (AAN85601); PanCrtZ (P21688); PaCrtZ (Q01332); Pp (NP_745389). A, Alcaligenes sp.; At, Arabidopsis thaliana; Ba,
Brevundimonas aurantiaca; Ca, Capsicum annuum; Cm, Cyanidioschyzon merolae; Cs, Crocus sativus; Dr, Deinococcus radiodurans; Gv
Gloeobacter violaceus; F, Flavobacterium; HpCrtW Haematococcus pluvialis; Le, Lycopersicon esculentum; Ma Mycobacterium aurum; Np,
Narcissus pseudonarcissus; N, Nostoc sp.; Np, Nostoc punctiforme; Os, Oryza sativa; Pa, Pantoea agglomerans; Pan, Pantoea ananatis; Pm,
Paracoccus marcusii; Pm, Prochlorococcus marinus; Pp, Pseudomonas putida; Ps, Pantoea stewartii; PsCrtI Pantoea stewartii phytoene
desaturase (AAN85599.1); Re, Rhodococcus erythropolis; Sn Synechocystis sp.; Sn1-2, Synechococcus sp.; Tse, Thermosynechococcus
elongatus; Te, Tagetes erecta; Vv, Vitis vinifera.
histidine motif of b-carotene hydroxylase was equivalent to those previously identified in fatty acid desaturases
and alkane hydroxylases [278]. This suggested that CHB belong to the non-heme-di-iron monooxygenase
(NHDIM) family (Fig. 23) where these histidine residues play a vital role such as coordinating the Fe ions
in the active-site [279]. This assumption was strengthened by the fact that CHB was inhibited by iron chelating
reagents such as o-phenanthroline and 8-hydroxyquinoline [395]. Based on this evidence, a mechanism was
proposed [395] to explain the presence of iso-cryptoxanthin in place of echinenone in Dictyococcus cinnabari-
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
405
nus cell cultures treated with the carotenogenic inhibitor diphenylamine [488]. Mechanistically CHB and typeCrtW KET use activated molecular oxygen to break the C–H bond with the concomitant formation of unsaturated bonds, retention of hydroxyl or keto groups [489] as shown for fatty acids [490–492]. Also, it has been
reported that the D5 desaturase [493] and the D22 desaturase [494] involved in yeast ergosterol biosynthesis
could operate via hydroxy sterol intermediates. These considerations are supported by the recent identification
of a Brevundimonas sp. hydroxylase that introduces a C2 hydroxy group into astaxanthin [495]. Brevundimonas
sp. hydroxylase is phylogenetically more related closely related to sterol C5 desaturase than to other bacterial
or plant carotenoid hydroxylases (Fig. 26). In a similar vein, it has been shown that a carotenoid ketolase isolated from Adonis palaestina operates via a desaturase-like mechanism [496]. Thus, through the subtile variation of one protein motif, nature has an invariant means to perform the large spectrum of reactions
catalyzed by desaturases, epoxidases, acetylenases, conjugases, ketolases, decarbonylase and methyl oxidases
(Fig. 27).
The above hydroxylation mechanism discussed above could not account for the hydroxylation of the e-ring
of a-carotene en route to lutein. Lutein biosynthesis involves hydroxylation of the two cyclohexene rings,
Sterol C5-desaturases
MmDES
Beta-carotene C3-hydroxylases
AtDES
PanCrtZ
RnDES
Bacteria
PaCrtZ
HsDES
ACrtZ
Carotenoid
C2-hydroxylase
BCrtG
Ca1
Ca2
Beta-carotene
C4-ketolase
At1
HpKET
At2
ApKET2
Plants
ApKET1
AtFAD
Beta-carotene
desaturases-hydroxylases
AtFAD2
Fatty acid
desaturases
PpALKB
AaCrtZ
Alkane hydroxylase
Beta-carotene C3-hydroxylase
0.1
Fig. 26. Unrooted phylogenetic tree of non-heme-di-iron hydroxylases, ketolases and related homologues. Alignments were performed
with ClustalX and visualized with TreeView. Distances between sequences are expressed as 0.1 changes per amino acid residue. Accession
Numbers and abbreviations. PanCrtZ (P21688); PaCrtZ (Q01332); ACrtZ (Q44262); Ca1 (CAA70427.1); Ca2 (CAA70888); At1
(NP_194300.1); At2 (BAA98075.1); ApKET2 (AAV85453.1); ApKET1 (AAV85452.1); PpALKB (CAB54049.1); AaCrtZ (P54973);
AtFAD2 (AAA32782.1); AtFAD (AAA61773.1); HpKET, (Q39982); BCrtG (BAD99415.1); HsDES (BAA33729); MmDES
(AAH24132.1); RnDES (BAB19798.1); AtDES (CAA62079.1). Aa, Agrobacterium aurantiacum; A, Alcaligenes sp.; Ap, Adonis palaestina;
At, Arabidopsis thaliana; B Brevundimonas sp.; Ca, Capsicum annuum; Hp, Haematococcus pluvialis; HsDES, Homo sapiens; Mm, Mus
musculus; Pa, Pantoea agglomerans; Pan, Pantoea ananatis; Pp, Pseudomonas putida; Rn, Rattus norvegicus.
406
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
1
2
6
5
3
HO
4
β-ring
C2-Hydroxy
HO
C3-Hydroxy
HYDROXYLASE
DESATURASE
KETOLASE
OH
C4-Hydroxy
C3-4-Didehydro
O
C4-Keto
Fig. 27. Flexible transformation exerted by non-heme-di-iron hydroxylases, ketolases and related homologues.
b- and e- of a-carotene. Although the b-ring of lutein is hydroxylated by a non-heme-di-iron monooxygenase,
the e-ring is not. This point was clarified following the isolation of the Arabidopsis mutant lut1 [497]. Lut1
mutant has decreased lutein content (up to 95%) and accumulated zeinoxanthin, a mono-hydroxy xanthophyll
[497]. To identify the molecular basis of this biochemical modification, the LUT1 gene was cloned [498].
Molecular analysis revealed that the LUT1 gene encodes a cytochrome P450-type oxygenase, CYP97C1.
The gene complemented the lut1 mutant, thus establishing that mixed function oxygenases such as P450s
are involved in the hydroxylation of the e-ring of lutein [498]. Further construction of an Arabidopsis mutant
with all three known hydroxylase genes deficient, resulted in plant mutants still producing b-hydroxy xanthophylls [498]. This led to the hypothesis that Arabidopsis CYP97A3 which is 49% identical to LUT1, could
encode an additionnal b-hydroxylase [498]. In this context, it is worth noting that CYP175A1 from Thermus
thermophilus has been characterized as a CHB [499]. Thus the introduction of hydroxy group into the ring of
carotenoid is catalyzed by structurally unrelated enzymes such as non-heme-di-iron oxygenases and cytochrome P450-type oxygenases (Fig. 23).
8.2.5. Xanthophyll interconversion
The gene encoding zeaxanhin epoxidase (ZEP) that catalyzes the conversion of zeaxanthin into antheraxanthin and violaxanthin has been first cloned from N. plumbaginifolia [500] (Fig. 23). ZEP catalyzes the epoxidation of 3-hydroxy-b-cyclohexenyl ring and requires the presence of FAD and ferredoxin for activity
[501,502]. The resulting antheraxanthin and violaxanthin are subject to de-epoxidation by violaxanthin deepoxidase (VDE). VDE was purified and its gene cloned from romaine lettuce [503]. VDE and ZEP are members of the lipocalin family [504]. Proteins belonging to this family are characterized by their ability to bind
small lipophilic molecules [505]. Interestingly, VDE binds to monogalactosyldiacylglycerol (MGDG) at low
pH [506]. VDE activity depends upon the availability of violaxanthin, the acidic pH of the lumen and protonated ascorbate [507] (Fig. 28). Thus pretreatment with L-galactono-1,4-lactone, an ascorbic acid precursor,
induces the accumulation of zeaxanthin in leaves [508]. Also VDE is inhibited by dithiothreitol, mercaptoethanol, cysteine and o-phenanthroline [509]. Available evidence shows that during the light period, VDE
is bound to the thylakoid membranes [507,510] concomitantly to the acidification of the lumen. This agrees
with the fact that in vitro VDE binds to the thylakoid membranes at pH 5.2 and is released at pH 7 [506].
VDE is specific for 3-hydroxy-5,6-epoxy-carotenoids which are in a 3S,5R,6S-configuration [506]. Thus
epoxy-lutein could constitute an additional substrate for VDE in plants such as Cuscuta [511], oak [512],
the tropical legume, Inga sp. [512] and plants belonging to the loranthaceae and viscaceae [513]. Recent data
indicate that in woody plants, lutein-5,6-epoxide is compartmentalized mainly in the buds and leaf primordia
[514] and is light-induced [514]. Except in vascular plants, sequences similar to VDE could not be identified in
currently annoted cyanobacterial or algal sequences [485]. This could be due to the fact that algal VDE is
structurally unrelated to vascular plant VDE. Indeed, in vitro studies demonstrated that VDE from the prim-
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
407
Fig. 28. The xanthophyll cycle and its relationship with abscisic acid biosynthesis. Enzyme abbreviations. GR, glutathione reductase;
DHAR, dehydroascorbate reductase; VDE, violaxanthin de-epoxidase; ZEP, zeaxanthin epoxidase; NSY, neoxanthin synthase; NCED,
nine-cis-epoxycarotenoid dioxygenase; SDR, short-chain dehydrogenase/reductase; AO, aldehyde oxidase. Substrate abbreviations:
ASC, ascorbate; ASCH, ascorbic acid; DHA, dehydroascorbate, GSH, reduced glutathione, GSSG, oxidized glutathione; ABAaldehyde, abscisic acid aldehyde.
itive green alga Mantoniella squamata has reduced affinity for antheraxanthin and for mono epoxy-xanthophyll [515,516].
5,6-Epoxy cyclohexenyl xanthophylls are rearranged into j-xanthopylls by the first characterized plant xanthophyll biosynthetic gene capsanthin-capsorubin synthase (CCS) [469]. The reaction yields capsanthin and
capsorubin, the red pepper keto-carotenoids [469]. Mechanistically CCS operate like LC [468]. A CCS homolog displaying neoxanthin synthase (NSY) has been isolated from tomato [517] and potato tuber [518].
Although an Arabidopsis homologue has not been characterized, Blast search revealed that NSY is not chromoplast specific as anticipated [519], but is expressed in diverse tissues and organs. Mechanistically NSY operate like CCS, except we could not unambiguously demonstrate its cryptic lycopene cyclase.
8.2.6. The xanthophyll cycle
In photosynthetic organisms, the function of ZEP and VDE is integrated in the photoprotective xanthophyll cycle [520]. The xanthophyll cycle is present in vascular plants, green algae and some brown and red
algae but not in cyanobacteria [520]. In marine diatoms, the xanthophyll cycle is termed the diadinoxanthin
cycle and includes reversible reaction between diadinoxanthin and diatoxanthin [521] (Fig. 28).
Protection against excess light involves dissipation of light energy as heat and is known as the qE component of non-photochemical quenching (npq) of chlorophyll fluorescence. The process involves a buildup of a
proton gradient across the thylakoid membrane [522], the presence of the PSII protein PsbS [523] and the
deepoxidation of violaxanthin to zeaxanthin via antheraxanthin [520] (Fig. 28). Arabidopsis mutant devoid
of VDE activity has been isolated using a screen for plants displaying lower npq [524]. Also, Tobacco plants
expressing antisense VDE have reduced npq [525].
The isolation of Arabidopsis mutant deficient in vitamin C (vtc1) [526] suggests that a subtile regulation
seems to operate between the xanthophyll cycle, ascorbate and ABA synthesis, at least in photosynthetic
408
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
tissues (Fig. 28). Indeed, the ABA content of vtc1 leaves is 60% greater than those measured in control plants
[527]. These conditions favor the downregulation of VDE [528]. Thus reduced VDE activity could increase the
flux of violaxanthin and neoxanthin used for ABA synthesis. This contention is supported by the fact that the
expression of the 9-cis-epoxycarotenoid dioxygenase (NCED) catalyzing the synthesis of xanthoxin is
increased in vtc1 [527].
8.2.7. Organization of carotenogenic enzymes
Earlier studies showed that in plastid the initial steps of carotenoid synthesis up to the level of phytoene
could take place in the stroma while later steps require tight cooperation between the stromal phase and
the membrane components [430,529]. However beyond this fact, we do not know how the carotenogenic
enzymes are organized. This could have impact on the fate of the final products as shown by the accumulation
of c-carotene [473,530] or to b-cryptoxanthin [479].
Available data suggest that the xanthophyll cycle operate via binding and release from pigment protein
complexes [531]. Also the lipid environment influences the activity of carotenogenic enzymes. For instance,
VDE activity, is optimal when the concentration of MGDG relative to violaxanthin is 28:1 [532,533]. Also
overexpression of PsbS, an intrinsic chlorophyll-binding protein of PSII [523] increased the VDE activity
[534]. PsbS apparently binds free zeaxanthin [523] which otherwise could inhibit VDE [509,531].
8.2.8. Molecular regulation
Photosynthetic organisms control expression of photosynthesis genes in response to light intensity and
redox conditions. In anoxygenic photosynthetic bacteria the biosynthesis of the photosynthetic apparatus
and its isoprenoid components is upregulated by moderate light intensity and downregulated under aerobic
conditions [535,536]. To this end, Bradyrhizobium and Erythrobacter, possess two spatially and genetically distinct carotenogenic pathways that operate concomitantly in the same cell during the photosynthetic growth.
The photosynthetic pathway leads to spirilloxanthin which is incorporated in the photosynthetic membranes,
while the other, a non-photosynthetic pathway leads to canthaxanthin or other ketocarotenoids [537] which
are incorporated in non-photosynthetic membranes [538,539] to protect them by virtue of their efficient scavenging roles [540], 34 [541]. The regulation of the carotenoid biosynthesis under these conditions involves the
homologous ppsR/CrtJ, and tspO/CrtK [542,543]. TspO and CrtK encode a sensory transducer that negatively
regulate carotenogenic genes in response to light and oxygen [542]. PpsR and CrtJ possess a DNA and PAS
binding domains. The DNA binding domain blocks the transcription of carotenoid genes by binding to the
palindromic motif (TGTN12ACA) of the promoters [544,545], while the PAS domain is involved in sensing
light, redox potential and interacting with small ligands [546]. In the photosynthetic bacteria Bradyrhizobium
strain ORS278 and Rhodopseudomonas palustris, the repressive effect of PpsR and CrtJ is antagonized by the
red light absorbing form (Pr) of the bacteriophytochrome [538,547,548]. Interestingly, bacteriophytochrome
also control the synthesis of the carotenoid deinoxanthin in the non-photosynthetic bacteria Deinococcus radidurans [549].
The influence of light on carotenoid synthesis and carotenogenic gene expression has been reported in green
tissues [550–555]. Overexpression of the CHB gene in Arabidopsis increased the xanthophyll cycle pool 2-fold.
The resulting plants were more tolerant to high light and high temperatures as shown by reduced necrosis,
reduced lipid peroxide accumulation and anthocyanin production [556]. Similarly, tobacco plants expressing
a bacterial CHB produced more zeaxanthin and was more resistant when challenged by UV light [557]. On the
other hand, expression of an antisense construct of CHB in Arabidopsis reduced the content of violaxanthin
and neoxanthin, without changing the lutein content [558]. The level of lutein could be increased following the
overexpression of lycopene e-cyclase [558].
Through the use of 3-(3,4-dichlorophenyl)-1,1-dimethyl-urea (DCMU) and 2,5-dibromo-3-methyl-6-isopropyl-p-benzoquinone (DBMIB), two photosynthetic electron transport inhibitors that respectively inhibit
the reduction of plastoquinone (PQ) and the oxidation of plastoquinol, it has been shown that the expression
of ZEP and CHB are redox controlled. When the reduction of plastoquinone is inhibited by DCMU, the
expression of ZEP and CHB is upregulated. Conversely, when the oxidation of plastoquinol is inhibited by
DBMIB, the expression of ZEP and CHB is downregulated [555]. One could notice that in green tissue,
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
409
the reduction of PQ following DCMU treatment is paralleled by an increased accumulation of carotenoid
[555]. These data based on studies with photosynthetic tissues and those previously reported for chromoplast
containing tissue [489] enlarge the role of redox regulation of carotenogenic genes in plants. Demonstration of
this type of regulation under natural condition is provided by the protective role of eschscholtzxanthin, monoanhydroeschscholtzxanthin, anhydroeschscholtzxanthin which accumulate in red leaves of Buxus sempervirens
as a response to photoinhibitory conditions during winter [559].
In the case of chromoplast containing tissues, a transcriptional regulation of carotenogenesis seems to prevail [188,489,551,560]. Light also regulates the accumulation of carotenoids in non-photosynthetic tissues
through the action of phytochrome [561]. Further experiments show that specific tomato fruit phytochrome
controls the accumulation of lycopene during tomato ripening, independently of ethylene. This control is
essentially due to phytochrome A since its mRNA accumulates 11.4-fold during the ripening period [562]. Furthermore, the availability of deposit structures characteristic of chromoplasts [563] influences the level of
carotenoid that can be accumulated [564–566]. A mutation of the Or gene induces carotenoid accumulation
in orange-curd-high carotene cauliflower inbreds [567]. The Or gene has been cloned [568] and shown to
encode a plastid protein that display transmembrane domains. In addition, Or protein belongs to a protein
complex in vivo [569].
QTL analysis reveals that several largely uncharacterized genes are involved in the regulation of carotenoid
biosynthesis [570]. This is supported by recent data indicating a connection between carotenoid accumulation
and photomorphogenesis in tomato. Analysis of tomato mutants high pigment (hp) 1 and 2 and dark green
(dg) that induce increased photoresponsiveness and increased fruit pigmentation [571,572] shed new light.
It has been shown that these mutations are not due to a lesion in structural carotenogenic genes, but to the
tomato homolog of A. thaliana DE-ETIOlATED1 (DET1), a negative regulator of photomorphogenesis in
the case of hp2 [573] and dg [574] and to the UV-DAMAGED DNA-BINDING PROTEIN1 (DDB1) [575]
in the case of hp1 [576,577]. These data demonstrate that manipulation of the light signal cascade regulates
the accumulation of carotenoid in tomato fruits. This is further strengthened by RNA interference repression
of tomato DET1, DDB1 genes, as well as tomato COP1-like and HY5 [577–579]. Arabidopsis long hypocotyl
(hy) HY5 is a basic leucine zipper transcription factor that promotes photomorphogenesis whereas COP1, is a
negative regulator of photomorphogenesis that directly interacts with nuclear HY5 [580]. However, this regulation is not pathway specific, but participates in the pleiotropic regulation of antioxidant synthesis, because
in hp2 mutant the concentration of carotenoid, tocopherol, ascorbic acid as well as phenolic derivatives
possessing antioxidant effects were elevated [581,582]. Finally, the link between carotenogenesis and photomorphogenesis is also demonstrated by the fact that tomatoes overexpressing cryptochrome, a blue light
receptor, overproduce chlorophylls in leaves and lycopene in fruits [583].
8.2.9. Integration of carotenoids in lipoprotein structures
In chloroplast approximatively, 80% of the xanthophyll are bound to pigment–protein complexes and thus
are not involved in the lipid bilayer of the membranes [584]. Thus less than 20% of the xanthophyll help structure the membranes according to a rivet-like mechanism [585]. The nature and the amount of carotenoids that
could be incorporated into plastid membranes is highly regulated in green tissues compared to non-green tissues
that sequester excess carotenoids in specific structures [563]. Data demonstrating flexible regulation of this process could be inferred from previous studies. For instance when R. sphaeroides are grown semiaerobically, the
reaction center contains spheroidenone [586], while spheroidene is incorporated the reaction center under
anaerobic conditions [587]. Also non-native carotenoids can be functionally incorporated into photosynthetic
membranes and complexes. In R. sphaeroides when the native phytoene desaturase leading to neurosporene is
replaced by the 4-step phytoene desaturase from Pantoea phytoene desaturase, the resulting non-native lycopene is functionally integrated into the photosynthetic membranes [539,588]. Similarly, the chromoplast specific
carotenoids capsanthin and capsorubin could be incorporated functionally in N. benthamiana [589]. Finally,
evidence based on in vitro reconstitution experiments [590–593], algal mutants [594–596] and higher plant
mutants [497,597–599] indicate that zeaxanthin could replace lutein in light harvesting complexes (LHC). In
addition the relative proportions of the different chlorophyll antenna proteins are modified in these mutants
[596,600]. These data collectively suggest that lutein, loroxanthin, violaxanthin and neoxanthin are not essential
for the organization and the function of LHCII and LHCI in photosynthetic organisms.
410
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
The reassembly monomeric recombinant LHCIIb in the presence of diverse non-native xanthophylls,
revealed a specific requirement for the presence of a hydroxy group at C-3 on a single b-end group [601]. However, the requirement for a 3-hydroxy-b-xanthophyll could be bypassed in cyanobacteria. For instance, in Synechocystis defective in CHB, the resulting absence of zeaxanthin and myxol dimethyl-fucoside, did not impair
the growth and the photosynthetic capacities [602]. The isomeric configuration plays a decisive role as shown
by the fact that cis-isomers of zeaxanthin, violaxanthin, and lutein could not be incorporated, in contrast to
trans-neoxanthin [601]. However, one could note that some parasitic plants [603], such as Cuscuta reflexa, are
devoid of neoxanthin [511]. In C. reflexa, neoxanthin is replaced in the LHCs by 9-cis-violaxanthin [604], in
spite of the fact that neoxanthin binding site is highly selective for this xanthophyll [590]. It has also been
reported that complexes containing 30–50% trans-neoxanthin are not stable and display low energy transfer
at high temperatures [591]. On the other hand, lactucaxanthin and eschscholtzxanthin could be incorporated
in the absence of lutein [601]. With respect to the binding mechanism, it has been demonstrated that intermolecular p–p stacking between carotenoids and the aromatic residues in LHC play key role in the binding of
carotenoids to these proteins [605].
8.2.10. Carotenoid cleavage oxygenases
The rigid chromophore of carotenoids is attacked by several oxygenases to give apocarotenoids (Fig. 29)
for a review, see [606,607]. In plants, these reactions lead to the formation of abscisic acid, an important plant
hormone [608,609] and to several apocarotenoids derivatives that have diverse biological roles [607,610].
Carotenoid oxygenases have been characterized from Arabidopsis [611], tomato and peppers [612]. Similarly,
carotenoid dioxygenases catalyzing the initial steps of crocin synthesis in Crocus sativus stigma and bixin synthesis in Bixa orellana seeds have been characterized. In C. sativus, a chromoplast zeaxanthin 7,8(7 0 ,8 0 )-cleavage dioxygenase initiates the reaction to yield crocetin dialdehyde [612] whereas a lycopene 5,6(5 0 ,6 0 )-cleavage
dioxygenase is involved in the initial step of bixin synthesis [613]. All carotenoid oxygenases proteins display
four conserved histidines that could serve as ligands for non-heme iron. This is reinforced by the fact that the
human b-carotene 15,15 0 -monooxygenase [614] which catalyzes the first step in the synthesis of vitamin A, is
drastically inhibited by micromolar concentrations of the metal chelating agents a,a-bipyridyl and o-phenanthroline [615] and desferrioxamine [616]. X-ray analysis of the crystal structure of the apocarotenoid-15,15 0 oxygenase (ACO) from Synechocystis sp. PCC 6803 reveals that ACO contains an Fe2+-4-His arrangement at
the axis of a seven-bladed b-propeller chain fold [617]. Recent data based on site-directed mutagenesis using
mouse b-carotene 15,15 0 -monooxygenase-1 support the above evidence [618].
With respect to the biological roles of apocarotenoids, recent studies have further substantiated their potential involvement in the control of Arabidopsis [619] pea [620] and Petunia [621] ramification. In particular, the
putative apocarotenoid signal is modified via a cytochrome P450 enzyme [619].
Finally, like several membrane lipids with established functions, carotenoids are targeted to fulfill additional roles in other context (Fig. 30). In a more wider context, the process is triggered by lipid cleaving
and/or oxidizing enzymes. Thus unsaturated plant fatty acids (linolenic acid), sterols (campesterol) and diverse
carotenoids are, respectively, converted into jasmonic acid, brassinosteroids and apocarotenoids which regulate plant growth, development and response to biotic and abiotic factors (Fig. 30).
8.3. Chlorophyll biogenesis and degradation
The different steps of the pathway of chlorophyll biosynthesis and degradation have been elucidated
recently through biochemical analysis and molecular cloning (Fig. 31), for reviews, see [622,623]. The last steps
involve the prenylation of chlorophyllide by phytyl diphosphate or geranylgeranyl diphosphate catalyzed by
chlorophyll synthase (Fig. 31). In vitro studies revealed that chlorophyll a synthase accepts chlorophyllide a,
but not bacteriochlorophyllide a, as a substrate, whereas bacteriochlorophyll a synthase accepts bacteriochlorophyllide a, but not chlorophyllide [624]. In addition, bacteriochlorophyll a synthase in Rhodobacter, Synechocystis and Arabidopsis have a preference for phytyl diphosphate over geranylgeranyl diphosphate [624].
However in R. rubrum GGPP is used instead of phytyl diphosphate [625]. In some lower photosynthetic
organisms, farnesyl diphosphate is used as the prenyl donor. For instance, farnesyl diphosphate is used for
the prenylation of bacteriochlorophyll g is Heliobacteria. This phenomenon is due to the fact that Heliobac-
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
411
4'
7
1
13
11
15
14'
12'
10'
8'
3'
6'
6
2
9
5'
8
10
14
12
15'
13'
11'
9'
7'
2'
1'
5
3
4
CAROTENOIDS
CLEAVAGE-OXIDATION
APOCAROTENOIDS
CYCLIC
LINEAR
O
CHO
O
CHO
HOH2C
O
β -Cyclocitral Safranal
CH2OH
Rosafluin
Oxoisophorone β -Ionone
O
O
O
O
C14 dialdehyde
O
O
Dihydroactinidiolide β -Damascone
β -Damascenone
CHO
OHC
Crocetin dialdehyde
O
O
O
CHO
HO
O
COOH
R-O
O-R
O
Xanthoxin
Theaspirone
OH
O
Abscisic acid
CHO
Crocetin glycosides
CHO
O
Acycloretinal
HO
Retinal
3-Hydroxy-retinal
OHC
CHO
CHO
Bixin dialdehyde
β -Apo-10'-Carotenal
CH2OH
O
O
O
OH
O
Blumenin
HOOC
OH
OH
COOH
O
COOCH3
Bixin
OH
OH
O
O
Marmelo Lactones
Fig. 29. Structure of some apocarotenoids.
teria are apparently unable to produce geranylgeraniol (C20) [416]. Beside farnesyl, several non-isoprenoid
side chains could be used in photosynthetic bacteria. These include, dodecanyl-, hexadecenyl-, pentadecenyl-, tetradecenyl-side chains [626]. Mechanistic studies of the prenylation reactions in R. rubrum indicate that
the process involves the nucleophilic attack by the carboxylate group of the chlorophyllide on the activated
prenyl residue [627,628].
The first step of chlorophyll degradation involves the detachment of phytol by chlorophyllase (Fig. 31).
This reaction is followed by the removal of the magnesium atom and partial oxygenolytic breakdown of
412
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
Fig. 30. General role of membrane lipids in the formation of plant regulators. In addition to their structural roles, membrane lipids are
targeted to generate bioactive derivatives. The process is initiated by lipid cleaving and/or oxidizing enzymes and culminates for instance
with the production of jasmonic acid, brassinosteroids and apocarotenoids that serve diverse roles.
the tetrapyrrole macrocycle to give fluorescent chlorophyll catabolites (FCCs). These reactions are sequentially catalyzed by magnesium dechelatase (MDCH), pheophorbide oxygenase (PAO) and red chlorophyll
catabolite (RCC) reductase (RCCR) [623,629]. The FCCs are finally exported to the vacuole to give the final
products termed as non-fluorescent chlorophyll catabolites (NCCs) [630,631] (Fig. 31). The green alga Chlorella protothecoides is devoid of RCCR and RCC-like derivatives are excreted into the medium [632]. This catabolic pathway is enhanced during senescence and non-green plastid differentiation. The induction of PAO
and the activity of PAO are restricted to senescing tissues [633]. In the case of senescing tissues, the degradation of chlorophyll is not associated to the general remobilization of nutrient occurring during this period as
previously thought [634]. Chlorophyll catabolites are not exported from senescent leaves [635]. This catabolism most probably avoid the accumulation of photodynamic chlorophyll derivatives. The fate of the phytyl
chain is unknown. It has been proposed that the phytyl chain is subject to a photooxidative degradation process [636,637]. Finally, one could note that chlorophyll-free chromoplasts, maintain the potential for chlorophyll biosynthesis as shown by in vitro incubation with exogenous chlorophyll precursors [638–640].
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
413
Fig. 31. Pathways of chlorophyll biosynthesis and catabolism. (A) Biosynthesis. Enzyme abbreviations. GTS, glutamyl-tRNA synthetase;
GTR, glutamyl-tRNA reductase; GSAT, glutamate-1-semialdehyde aminotransferase; ALAD, 5-aminolevulinate dehydratase; PBD,
porphobilinogen deaminase; UROS, uroporphyrinogen III synthase; UROD, uroporphyrinogen III decarboxylase; CPO, coproporphyrinogen III oxidase; PPX, protoporphyrinogen IX oxidase; MCH, magnesium chelatase; MTF, magnesium protoporphyrin IX
methyltransferase; MTC, magnesium protoporphyrin IX monomethylester cyclase; DCR, divinyl protochlorophyllide reductase; POR,
NADPH-protochlorophyllide oxidoreductase; GGR, geranylgeranyl reductase; CS, chlorophyll synthase; CAO, chlorophyll a oxygenase;
CBR, chlorophyll b reductase; HCR, hydroxychlorophyll a reductase. Substrate abbreviations. GGPP, geranylgeranyl diphosphate;
PhytylPP, phytyl diphosphate. (B) Catabolism. Enzyme abbreviations. CASE, chlorophyllase; MDCH, magnesium dechelatase; PAO,
pheophorbide a oxygenase; RCCR, red chlorophyll catabolite reductase. Substrate abbreviations. FCC, fluorescent chlorophyll catabolite;
NCCs, non-fluorescent chlorophyll catabolites; RCC, red chlorophyll catabolite.
414
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
8.4. Tocopherol and tocotrienol biogenesis
The pathway of tocopherol and tocotrienol biosynthesis is shown in Fig. 32. It is apparent from structural
considerations that the biosynthetic steps could be divided into the origin of the quinol ring, the methyl groups
and the formation of the isoprenoid side-chain. Although the enzymatic steps and the compartmentation of
component enzymes [641–643] of the tocopherol pathways and the purification of S-adenosyl methionine-c-
O
H
3
O
HO
Phylloquinone
H
O
8
SAM
DMPQMT
Plastochromanol-8
O
TC
H
3
O
Tyrosine
DMPQ
H
HO
9
OH
DHNAPT
HPP
Chorismate
H
9
MSBQMT
SAM
SPP
OH
COOH
Plastoquinol-9
PP
HPPD
O
H
HO
O
OH
HST
HOOCH2C
DHNA
9
OH
MSBQ
HGA
PP
PP
H
3
PhytylPP
HPT
H
4
GGPP
HGGT
SAM
SAM
H
HO
3 MPBQMT
OH
TC
DMPBQ
H
HO
3
OH
TC
MPBQ
O
γ -Tocopherol
SAM
H
3
δ -Tocopherol
SAM
H
3
HO
HO
O
α -Tocopherol
H
3
δ -Tocotrienol
SAM
H
3
β -Tocopherol
H
3
O
γ -Tocotrienol
SAM
SAM γ -MT3
HO
O
TC
HO
O
SAM γ -MT
3
OH
DMGGBQ
HO
O
SAM γ -MT
TC
H
HO
3 MGGBQMT
OH
MGGBQ
HO
HO
H
HO
H
3
SAM γ -MT3
HO
O
β -Tocotrienol
H
3
O
α-Tocotrienol
H
3
Fig. 32. Phylloquinone, vitamin E (tocopherols and tocotrienols) and plastoquinone biosynthetic pathways. Enzyme abbreviations. HPPD,
p-hydroxyphenylpyruvate dioxygenase; HPT, 2, homogentisate phytyl transferase; MPBQMT, 2-methyl-6-phytyl-1,4-benzoquinol
methyltransferase; TC, tocopherol cyclase; SAM c-MT, S-adenosyl-L-methionine c-tocopherol methyltransferase; HGGT, homogentisate
geranylgeranyl transferase; MGGBQMT, 2-methyl-6-geranylgeranyl benzoquinol methyltransferase; SAM c-MT3, S-adenosyl-L-methionine c-tocotrienol methyltransferase; DHNAPT, DHNA phytyl transferase; DMPQMT, demethylphylloquinone methyltransferase; HST,
3, homogentisate solanesyl transferase; MSBQMT, 2-methyl-6-solanesyl-1,4-benzoquinol methyltransferase. Substrate abbreviations.
DHNA, 1,4-dihydroxy-2-naphtoate; DMGGBQ, 2,3-dimethyl-5-geranylgeranyl benzoquinol; DMPBQ, 2,3-dimethyl-5-phytyl benzoquinol; DMPQ, demethylphylloquinone; GGPP, geranylgeranyl diphosphate; HGA, homogentisic acid; HPP, p-hydroxyphenylpyruvate;
MGGBQ, 2-methyl-6-geranylgeranyl benzoquinol; MPBQ, 2-methyl-6-phytyl-1,4-benzoquinol; MSBQ, 2-methyl-6-solanesyl-1,4-benzoquinol; PhytylPP, phytyl diphosphate; SAM, S-adenosyl-L-methionine; SPP, solanesyl diphosphate.
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
415
tocopherol (SAM-c-T) [644] have been carried out previously, detailed progress in the elucidation of the pathways has been made through the use of Arabidopsis mutants (for a review, see [645]).
p-Hydroxyphenylpyruvate dioxygenase (HPPD) catalyzes the synthesis of homogentisic acid (HGA), a precursor of the aromatic ring of tocopherols, tocotrienols and plastoquinone (Fig. 32). However, it appears that
HPPD is not required for plastoquinone biosynthesis in cyanobacteria and notably in Synechocystis [646].
Available data point to the fact that in contrast to downstream enzymes involved in the biosynthesis of plastid
prenyllipids, HPPD is localized in the cytosol [647,648]. Thus we do not know how homogentesic acid (HGA)
is recruited for tocopherol and tocotrienol biosynthesis.
HPPD is inhibited by several secondary metabolites produced by lichens and plants. These inhibitors
include triketones, benzoquinones, naphthoquinones and anthraquinones. In vitro studies reveal that triketone natural products behave as competitive tight-binding inhibitors of HPPD [649]. HPPD is the target of
several bleaching herbicides such as sulcotrione and isoxaflutole [650,651]. The herbicidal activities of these
derivatives lead to the inhibition of tocopherol and plastoquinone biosynthesis and to concomitant buildup
of phytoene [652]. These herbicides per se did not inhibit phytoene desaturase [653] but inhibit the synthesis
of plastoquinone, a cofactor involved in the desaturation of phytoene [457].
The next step in the pathway is the prenylation of HGA acid by two distinct prenyltransferases. The prenylation of homogentesic acid by homogentisate-phytyl transferase (HPT) initiates the tocopherol branch. On
the other hand, for the tocotrienol branch, the prenylation is carried out by homogentisate-geranylgeranyl
transferase (HGGT) (Fig. 32). The tocotrienol pathway is particularly active in some plants belonging to
monocot [654]. Either in the tocopherol or tocotrienol pathways, the resulting prenyl derivatives are subjected
to methylation and cyclization as depicted in Fig. 32. Recent biochemical data indicate that Capsicum and
Arabidopsis SAM-c-T convert d- and c- into b- and a-tocopherol, respectively. b-tocopherol is not accepted
as a substrate. Kinetic analysis with the Arabidopsis suggests an iso-ordered bi–bi type reaction mechanism.
In Arabidopsis and in cyanobacteria the same methyltransferase catalyzes the methylation of 2-methyl-6phytyl benzoquinol (MPBQ), the tocopherol precursor and 2-methyl-6-solanesyl-1,4-benzoquinol (MSBQ),
the plastoquinone precursors [655,656]. Unexpectedly, the Arabidopsis and the cyanobacterial enzymes share
only 18% identity [656]. Second, MPBQ and 2-methyl-6-geranylgeranyl benzoquinol (MGGBQ), which are,
respectively, the tocopherol and the tocotrienol precursors are cyclized by the same enzyme [657].
Normal leaves are practically devoid of HGGT activity [641,658,659]. Recent data revealed that Arabidopsis HPT expressed in insect cells had approximately a 50- to 80-fold preference for Phytyl PP over GGPP [660].
These data may explain in part why leaves do not synthesize tocotrienols. However, it has been shown that
tobacco leaves synthesized massive amount of tocotrienols under constrained conditions, i.e., when overexpressing HPPD and yeast prephenate dehydrogenase [661]. On the other hand, overexpression of tocopherol
cyclase in Arabidopsis did not lead to the synthesis of tocotrienols [662]. The gene encoding HGGT has been
cloned from cereals and is significantly different from HPT [663]. When HGGT was expressed in tobacco callus and in Arabidospsis, tocotrienol accumulated [663].
In the maize mutant Sucrose export defective Sxd1 [664] which encodes tocoperol cyclase (TC) and in
potato plants where the tocopherol cyclase is silenced using a RNAi-mediated silencing approach [665], the
absence of tocopherol leads to pleiotropic effects. In these plants the plasmodesmata size exclusion limit or
closure is altered by callose deposits. Under these conditions, the regulation of symplastic transport exerted
by plasmodesmata is blocked in these plants. As a consequence, mutant and silenced plants accumulated
starch and soluble sugars at high levels compared to wild-type plants [664]. In silenced potatoes, this phenomenon was paralleled by repression of photosynthesis genes (rbcS, Cab genes) and by an induced accumulation
of defense-related genes such as the proteinase inhibitor 1 and 2 [664]. These two features are characteristic of
the molecular events induced by excess sucrose in plants [666]. However in the vitamin E deficient1 Arabidopsis
mutant (vte1) that lacks tocopherol due to the deficiency TC activity, no such pleitropic changes have been
observed [657,667]. vte1 is devoid of tocopherol but accumulates DMPBQ, the substrate of tocopherol cyclases [657,667]. It has been shown that DMPBQ could compensate for tocopherol deficiency in germinating
seedlings [668]. It has been suggested that the absence of carbohydrate-deficient phenotype in Arabidopsis
was due to the fact, that in this species, the phloem loading is predominantly apoplastic while in maize a symplastic pathway is preferentially used [669]. Callose deposition is usually associated with strong membrane
lipid peroxidation [670]. Thus tocopherol may be required to counteract this deleterious effect [391,671–673].
416
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
Genes encoding enzymes of the initial steps of tocopherol and plastoquinone synthesis are subject to regulation under stress conditions. In Arabidopsis, the activity of tyrosine aminotransferase, which catalyzes the
conversion of tyrosine to p-hydroxyphenylpyruvate, the precursor of HGA, is increased following jasmonic
acid treatment. This trend is followed by an increased accumulation of tocopherol [674]. In a similar vein,
the transcript level of barley 4-hydroxyphenylpyruvate dioxygenase HPPD is increased during senescence
and this trend is induced by application of methyl jasmonate, ethylene as well as by treatment with the herbicides paraquat and DCMU or hydrogen peroxide [675]. It has been shown that in Arabidopsis overexpressing TC, the tocopherol content is increased 7-fold and is paralleled by a decrease in ascorbate and glutathione
[662]. At the other extreme, the deficiency of TC is paralleled by an increase in ascorbate and glutathione [662].
Thus, these data suggest that a sensing mechanism operates that links the redox state of the cell to the ascorbate-glutathione cycle implicated in the regeneration of oxidized tocopherol [671].
Using hpt and tc mutants of Arabidopsis, it has been shown that tocopherols control during the mobilization of lipids during germination and early seedling development [668]. The seed longevity is reduced for both
mutants compared to the wild-type. The hpt mutant totally devoid of tocopherols, accumulates high level of
fatty acid hydroperoxides and their corresponding hydroxy derivatives. These changes were not observed in
the tocopherol cyclase mutants that accumulate DMPBQ, probably due to the potential antioxidant effect
of this compound [668]. These data are diagnostic features that the tocopherol content in germinating seeds
is required to protect the unsaturated lipid reserves from oxidative degradation. This phenomenon could have
a crucial role during the heterotrophic phase of oil seed germination. This contention is reinforced by the fact
that the transcription of GGR is stimulated by light [399,676]. Interestingly, the protective role of tocopherol
toward lipid oxidation is conserved in cyanobacteria [677] and algae. For instance, in Chlamydomonas mutant
devoid of phytoene synthase, the lack of carotenoid is paralleled by partial deficiency in chlorophylls while the
tocopherol content is increased 2-fold [678].
Knowledge gained from the molecular analysis of the tocopherol pathway has been exploited to increase
the tocopherol content in model plants and agricultural crops (for a review, see [679]). In the absence of specific transcription factor, the upregulation of the expression of the SAM-c-tocopherol gene has been achieved
using a synthetic hybrid transcription factor [680]. The approach involved first a Dnase I hypersensitive mapping of the region of SAM-c-tocopherol gene that could interact with the DNA binding domain of zinc finger
proteins. Subsequently, a tripartite transcription factor construct was made that comprises a zinc finger protein designed to bind to the putative target sequence of SAM-c-tocopherol methyltransferase promoter followed by a maize C1 activation domain and a nuclear targeting signal. When introduced in Arabidopsis,
the hybrid transcription factor increased up to 20-fold the a-tocopherol content of the seeds [680].
8.5. Phylloquinone and plastoquinone biogenesis
In photosynthetic membrane, phylloquinone also known as vitamin K1 functions as a bound one-electron
cofactor in the A1 site of photosystem I (PSI). On the other hand, plastoquinone-9 (PQ-9) is implicated in the
function of photosystem II (PSII), as a bound one-electron cofactor in the QA site and as an exchangeable
two-electron two proton cofactor in the QB site. In its reduced form, PQH2-9 diffuses in the membrane and
is oxidized by the cytochrome b6f complex. Their biosynthesis has not been fully investigated. Because bacterial menaquinones (vitamin K2) are structurally related to phylloquinone and because of the pathway leading
to menaquinone is known (see http://metacyc.org), it has been hypothesized that a very similar pathway operates in cyanobacteria and plants, according to 8 steps [681]. These informations led first to the cloning of 1,
4-dihydroxy-2-naphtoate phytyltransferase DHNA-phytyltrasferase (DHNAPT) [681], 2-succinyl-6-hydroxyl2,4-cyclohexadiene-1-carboxylate synthase and O-succinylbenzoic acid-CoA synthase from Synechocystis
[682] (Fig. 32). An Arabidopsis mutant showing pale-green young leaves and white old leaves with a lesion
in the gene designed abc4 has been isolated [683]. Cloning and subsequent functional characterization of
the ABC4 gene indicated that it encoded the DHNAPT [683]. The mutation led to the absence of phylloquinone, and to a reduced PQ-9 content (about 3% of that of the wild-type). Similarly, in cyanobacteria when the
phylloquinone biosynthetic gene encoding 2-succinyl-6-hydroxyl-2 4-cyclohexadiene-1-carboxylate synthase
and O-succinylbenzoic acid-CoA synthase, DHNA synthase and DHNAPT were inactivated, no phylloquinone was produced. However, under these conditions, the PSII was not affected while the PSI activity was
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
417
50–60% preserved due to the recruitment of plastoquinone into the A1 site of PSI [681,684]. Obviously this
compensation mechanism between phylloquinone and PQ-9 did not operate in higher plants. The binding flexibility of the A1 site of cyanobacteria is further strengthened by the fact that it could tolerate the biosynthetic
intermediate demethylphylloquinone (DMPQ) [685] nonnative quinones [686].
The apg1 mutant was isolated as a Ds-tagged Arabidopsis pale-green mutant. The mutant displays pale
green leaves that turn albino in old leaves [687]. The apg1 mutant could not grow autotrophically but remain
green when cultured in the presence of sucrose and under moderate light conditions (75 lmol quanta m2 s1).
Under these conditions the mutant has qualitatively the same pigment composition than the wild-type. Under
strong light conditions (400 lmol quanta m2 s1) the mutant was bleached [687]. In the apg1 mutant, the
amount of D1, LHC and OE23 was drastically reduced. The APG1 encodes a plastid SAM-dependent methyltransferase highly identical to the 37 kDa chloroplast inner envelope membrane protein [688]. The apg1
mutant is deficient in PQ-9 and did not apparently accumulate 2-methyl-6-solanesyl-1,4-benzoquinol (MSBQ).
Because the apg1 mutant was devoid of PQ-9, it has been inferred that APG1 protein catalyzed the last steps
of plastoquinone synthesis, i.e., the methylation of MSBQ into PQ-9 [687]. Interestingly the PQ-9 deficiency
was correlated with phytoene accumulation. This represents a diagnostic feature of the inhibition of phytoene
desaturation as shown in the Arabidopsis mutant devoid of plastoquinone and tocopherol due to mutation in
the HPPD gene [648]. Recent data based on transgenic expression of maize and Arabidopsis TC suggest that
TC has a broader substrate specificity and catalyzed the formation of the chromanol ring of plastochromanol8 from PQ-9 [689] (Fig. 32). These data could explain the presence of plastochromanol-8 as a minor component in seed oils [690].
PQ-9 are prenylated by solanesyl diphosphate (SPP), a C45 isoprenoid derivative. Previous studies have
established that solanesyl diphosphate synthase (SPPS) and prenyltransferases catalyzing the synthesis of
(all-trans)-long prenyldiphosphates are usually activated by exogenous activator that are thought to facilitate
the desorption of the products from the active-site of the enzyme [691]. Two genes encoding SPPS have been
characterized from A. thaliana and designed AtSPP1 and AtSPPS2 [109,361]. Recombinant AtSPPS accepted
FPP and GGPP as allylic substrates but did not use DMAPP or GPP and the activity was stimulated by Triton X-100 [692]. When GGPP was used as the allylic substrate, only solanesyl diphosphate was synthesized.
On the other hand, when FPP was used several C20–C40 intermediates were released [692]. The kcat/Km value
for GGPP was 5.1-fold higher than that of FPP [361]. AtSPPS2 is localized in the plastids, while AtSPS1 is
localized in the endoplasmic reticulum [109,361]. Both genes are highly expressed in leaves compared in roots
and the expression of AtSPPS1 is higher than that of AtSPPS2 [361]. The plastidial localization AtSPPS2 suggests that it is most likely involved in the synthesis of the side-chain of PQ-9. The existence of an extraplastidial SPPS is reminiscent of the fact that the endoplasmic reticulum–Golgi membranes of spinach leaves could
synthesize plastoquinone [357] which is further transported to the plastids and mitochondria [358].
9. Conclusion
The regulation of the biosynthesis of isoprenoids in plant cells is poorly understood. In general, the amount
of induced accumulation of isoprenoids in plant tissues is low. In fact we do not master several factors. We do
not know how the metabolite fluxes between the primary and the secondary metabolisms are orchestrated.
Practically, for all studied pathways, the concentration of the substrates and products of each step are presently unknown to evaluate which reactions are in equilibrium and which are displaced from the equilibrium.
In addition, due to the complexity of the pathways, feedback control mechanisms should be envisioned. Also
in several cases, the accumulation of secondary metabolites requires the differentiation of sequestery structures
or conjugations and transport systems. All these specificities are difficult to delineate by insertion of limited
number of genes.
Specific transcriptional regulators that could control one or multiple steps of the pathways have not been in
depth characterized, as shown for the flavonoid pathway [693] and the jasmonate-responsive transcriptional
regulators which induces the genes involved in alkaloid biosynthesis [694].
Finally, given the fact that agriculturally important traits are regulated by a number of genes known as
quantitative trait loci (QTLs) derived from natural allelic variations [695], exploitation of QTLs and syntenic
genomes would be a much powerful approach to control the pathway.
418
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
Acknowledgements
We are indebted to the Europe Union Research Program QLK3-2000-00809 for research grants.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]
[53]
Pichersky E, Gang DR. TIPS 2000;5:439–45.
Baker ME, Blasco R. FEBS Lett 1992;301:89–93.
Fraser PD, Bramley PM. Prog Lipid Res 2004;43:228–65.
Yalovsky S, Rodr Guez-Concepcion M, Gruissem W. Trends Plant Sci 1999;4:439–45.
Galichet A, Gruissem W. Curr Opin Plant Biol 2003;6:530–5.
Eisenreich W, Rohdich F, Bacher A. Trends Plant Sci 2001;6:78–84.
Rohdich F, Kis K, Bacher A, Eisenreich W. Curr Opin Chem Biol 2001;5:535–40.
Lichtenthaler HK. Annu Rev Plant Physiol Plant Mol Biol 1999;50:47–65.
Rodriguez-Concepcion M, Boronat A. Plant Physiol 2002;130:1079–89.
Rohmer M. Comprehensive natural product chemistry, vol. 2. Oxford: Pergamon; 1999. p. 45–67.
Schwender J, Gemunden C, Lichtenthaler HK. Planta 2001;212:416–23.
Jomaa H, Wiesner J, Sanderbrand S, Altincicek B, Weidemeyer C, Hintz M, et al. Science 1999;285:1573–6.
Lange BM, Rujan T, Martin W, Croteau R. Proc Natl Acad Sci USA 2000;97:13172–7.
Lange BM, Croteau R. Proc Natl Acad Sci USA 1999;96:13714–9.
Mandel MA, Feldmann KA, Herrera-Estrella L, Rocha-Sosa M, Leon P. Plant J 1996;9:649–58.
Carretero-Paulet L, Ahumada I, Cunillera N, Rodriguez-Concepcion M, Ferrer A, Boronat A, et al. Plant Physiol
2002;129:1581–91.
Hsieh MH, Goodman HM. Plant Physiol 2005;138:641–53.
Bouvier F, dÕHarlingue A, Suire C, Backhaus RA, Camara B. Plant Physiol 1998;117:1423–31.
Lois LM, Rodriguez-Concepcion M, Gallego F, Campos N, Boronat A. Plant J 2000;22:503–13.
Gutierrez-Nava Mde L, Gillmor CS, Jimenez LF, Guevara-Garcia A, Leon P. Plant Physiol 2004;135:471–82.
Estevez JM, Cantero A, Romero C, Kawaide H, Jimenez LF, Kuzuyama T, et al. Plant Physiol 2000;124:95–104.
Estevez JM, Cantero A, Reindl A, Reichler S, Leon P. J Biol Chem 2001;276:22901–9.
Mahmoud SS, Croteau RB. Proc Natl Acad Sci USA 2001;98:8915–20.
Botella-Pavia P, Besumbes O, Phillips MA, Carretero-Paulet L, Boronat A, Rodriguez-Concepcion M. Plant J 2004;40:188–99.
Balmer Y, Koller A, del Val G, Manieri W, Schurmann P, Buchanan BB. Proc Natl Acad Sci USA 2003;100:370–5.
Lemaire SD, Guillon B, Le Marechal P, Keryer E, Miginiac-Maslow M, Decottignies P. Proc Natl Acad Sci USA 2004;101:7475–80.
Okada K, Hase T. J Biol Chem 2005;280:20672–9.
Wolfertz M, Sharkey TD, Boland W, Kuhnemann F. Plant Physiol 2004;135:1939–45.
Guevara-Garcia A, San Roman C, Arroyo A, Cortes ME, de la Luz Gutierrez-Nava M, Leon P. Plant Cell 2005;17:628–43.
Kemp LE, Alphey MS, Bond CS, Ferguson MA, Hecht S, Bacher A, et al. Acta Crystallogr D Biol Crystallogr 2005;61:45–52.
Fellermeier M, Sagner S, Spiteller P, Spiteller M, Zenk MH. Phytochemistry 2003;64:199–207.
Learned RM, Connolly EL. Plant J 1997;11:499–511.
Rodriguez-Concepcion M, Fores O, Martinez-Garcia JF, Gonzalez V, Phillips MA, Ferrer A, et al. Plant Cell 2004;16:144–56.
Leivar P, Gonzalez VM, Castel S, Trelease RN, Lopez-Iglesias C, Arro M, et al. Plant Physiol 2005;137:57–69.
Suzuki M, Kamide Y, Nagata N, Seki H, Ohyama K, Kato H, et al. Plant J 2004;37:750–61.
Schwarz MK. Eidgenössische Technische Hochschule, Zürich; 1994..
Arigoni D, Sagner S, Latzel C, Eisenreich W, Bacher A, Zenk MH. Proc Natl Acad Sci USA 1997;94:10600–5.
De-Eknamkul W, Potduang B. Phytochemistry 2003;62:389–98.
Yang JW, Orihara Y. Tetrahedron 2002;58:1265–70.
Kasahara H, Hanada A, Kuzuyama T, Takagi M, Kamiya Y, Yamaguchi S. J Biol Chem 2002;277:45188–94.
Kim D, Filtz MR, Proteau PJ. J Nat Prod 2004;67:1067–9.
Nabeta K, Saitoh T, Adachi K, Komuro K. Chem Commun 1998;6:671–2.
Itoh D, Karunagoda RP, Fushie T, Katoh K, Nabeta K. J Nat Prod 2000;63:1090–3.
Itoh D, Kawano K, Nabeta K. J Nat Prod 2003;66:332–6.
Hemmerlin A, Hoeffler JF, Meyer O, Tritsch D, Kagan IA, Grosdemange-Billiard C, et al. J Biol Chem 2003;278:26666–76.
Laule O, Furholz A, Chang HS, Zhu T, Wang X, Heifetz PB, et al. Proc Natl Acad Sci USA 2003;100:6866–71.
Hampel D, Mosandl A, Wust M. Phytochemistry 2005;66:305–11.
Luan F, Wust M. Phytochemistry 2002;60:451–9.
Dudareva N, Andersson S, Orlova I, Gatto N, Reichelt M, Rhodes D, et al. Proc Natl Acad Sci USA 2005;102:933–8.
Budziszewski GJ, Lewis SP, Glover LW, Reineke J, Jones G, Ziemnik LS, et al. Genetics 2001;159:1765–78.
Nagata N, Suzuki M, Yoshida S, Muranaka T. Planta 2002;216:345–50.
Page JE, Hause G, Raschke M, Gao W, Schmidt J, Zenk MH, et al. Plant Physiol 2004;134:1401–13.
Sharkey TD, Loreto F, Delwiche CF. In: Sharkey TD, Holland EA, Mooney HA, editors. Trace gas emission from plants. San
Diego: Academic Press; 1991. p. 153–84.
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
[54]
[55]
[56]
[57]
[58]
[59]
[60]
[61]
[62]
[63]
[64]
[65]
[66]
[67]
[68]
[69]
[70]
[71]
[72]
[73]
[74]
[75]
[76]
[77]
[78]
[79]
[80]
[81]
[82]
[83]
[84]
[85]
[86]
[87]
[88]
[89]
[90]
[91]
[92]
[93]
[94]
[95]
[96]
[97]
[98]
[99]
[100]
[101]
[102]
[103]
[104]
[105]
[106]
[107]
[108]
[109]
[110]
[111]
[112]
[113]
[114]
419
Sharkey TD, Chen X, Yeh S. Plant Physiol 2001;125:2001–6.
Loreto F, Pinelli P, Brancaleoni E, Ciccioli P. Plant Physiol 2004;135:1903–7.
Loreto F, Velikova V. Plant Physiol 2001;127:1781–7.
Schuhr CA, Radykewicz T, Sagner S, Latzel C, Zenk MH, Bacher A, Rohdich F, Eisenreich W. Phytochem. Rev 2003;2:3–16.
Soler E, Clastre M, Bantignies B, Marigo G, Ambid C. Planta 1993;205:324–9.
Bick JA, Lange BM. Arch Biochem Biophys 2003;415:146–54.
Flugge UI, Gao W. Plant Biol (Stuttg) 2005;7:91–7.
Burlat V, Oudin A, Courtois M, Rideau M, St-Pierre B. Plant J 2004;38:131–41.
Wouters J, Oudjama Y, Barkley SJ, Tricot C, Stalon V, Droogmans L, et al. J Biol Chem 2003;278:11903–8.
Carrigan CN, Poulter CD. J Am Chem Soc 2003;125:9008–9.
Kaneda K, Kuzuyama T, Takagi M, Hayakawa Y, Seto H. Proc Natl Acad Sci USA 2001;98:932–7.
Hemmi H, Ikeda Y, Yamashita S, Nakayama T, Nishino T. Biochem Biophys Res Commun 2004;322:905–10.
Ershov Y, Gantt RR, Cunningham FX, Gantt E. FEBS Lett 2000;473:337–40.
Poliquin K, Ershov YV, Cunningham Jr FX, Woreta TT, Gantt RR, Gantt E. J Bacteriol 2004;186:4685–93.
Barkley SJ, Desai SB, Poulter CD. J Bacteriol 2004;186:8156–8.
Nakamura A, Shimada H, Masuda T, Ohta H, Takamiya K. FEBS Lett 2001;506:61–4.
Ramos-Valdivia AC, van der Heijden R, Verpoorte R, Camara B. Eur J Biochem 1997;249:161–70.
Ramos-Valdivia AC, van der Heijden R, Verpoorte R. Planta 1997;203:155–61.
Ramos-Valdivia AC, van der Heijden R, Verpoorte R. Nat Prod Rep 1997;14:591–603.
Rilling HC. Pure Appl Chem 1979;51:597–608.
Sacchettini JC, Poulter CD. Science 1997;277:1788–9.
Koyama T. Biosci Biotechnol Biochem 1999;63:1671–6.
Liang PH, Ko TP, Wang AH. Eur J Biochem 2002;269:3339–54.
Poulter CD. Acc Chem Res 1990;23:70–7.
van Klink J, Becker H, Andersson S, Boland W. Org Biomol Chem 2003;1:1503–8.
van Klink JW, Becker H, Perry NB. Org Biomol Chem 2005;3:542–5.
Delourme D, Lacroute F, Karst F. Plant Mol Biol 1994;26:1867–73.
Dogbo O, Camara B. Biochim Biophys Acta 1987;920:140–8.
Kuntz M, Römer S, Suire C, Hugueney P, Weil JH, Schantz R, et al. Plant J 1992;2:25–34.
Sagami H, Korenaga T, Ogura K, Steiger A, Pyun HJ, Coates RM. Arch Biochem Biophys 1992;297:314–20.
Zenitani S, Shindo K, Tashiro S, Sekiguchi M, Nishimori M, Suzuki K, et al. J Antibiot (Tokyo) 2003;56:658–60.
Okada K, Saito T, Nakagawa T, Kawamukai M, Kamiya Y. Plant Physiol 2000;122:1045–56.
Zhu XF, Suzuki K, Saito T, Okada K, Tanaka K, Nakagawa T, et al. Plant Mol Biol 1997;35:331–41.
Zhu XF, Suzuki K, Okada K, Tanaka K, Nakagawa T, Kawamukai M, et al. Plant Cell Physiol 1997;38:357–61.
Bouvier F, Suire C, dÕHarlingue A, Backhaus RA, Camara B. Plant J 2000;24:241–52.
Li SM, Hennig S, Heide L. Tetrahedron Lett 1998;39:2721–4.
Ohara K, Ujihara T, Endo T, Sato F, Yazaki K. J Exp Bot 2003;54:2635–42.
Burke CC, Wildung MR, Croteau R. Proc Natl Acad Sci USA 1999;96:13062–7.
Burke C, Croteau R. J Biol Chem 2002;277:3141–9.
Burke C, Croteau R. Arch Biochem Biophys 2002;405:130–6.
Tholl D, Kish CM, Orlova I, Sherman D, Gershenzon J, Pichersky E, et al. Plant Cell 2004;16:977–92.
Tarshis LC, Yan M, Poulter CD, Sacchettini JC. Biochemistry 1994;33:10871–7.
Lesburg CA, Zhai G, Cane DE, Christianson DW. Science 1997;277:1820–4.
Saito A, Rilling HC. Arch Biochem Biophys 1981;208:508–11.
Tarshis LC, Proteau PJ, Kellogg BA, Sacchettini JC, Poulter CD. Proc Natl Acad Sci USA 1996;93:15018–23.
Dhiman RK, Schulbach MC, Mahapatra S, Baulard AR, Vissa V, Brennan PJ, et al. J Lipid Res 2004;45:1140–7.
Ohnuma S, Hirooka K, Ohto C, Nishino T. J Biol Chem 1997;272:5192–8.
Sagami H, Morita Y, Ogura K. J Biol Chem 1994;269:20561–6.
Ohnuma S, Hirooka K, Hemmi H, Ishida C, Ohto C, Nishino T. J Biol Chem 1996;271:18831–7.
Ohnuma S, Nakazawa T, Hemmi H, Hallberg AM, Koyama T, Ogura K, et al. J Biol Chem 1996;271:10087–95.
Ohnuma S, Hirooka K, Tsuruoka N, Yano M, Ohto C, Nakane H, et al. J Biol Chem 1998;273:26705–13.
Stanley Fernandez SM, Kellogg BA, Poulter CD. Biochemistry 2000;39:15316–21.
Narita K, Ohnuma S, Nishino T. J Biochem (Tokyo) 1999;126:566–71.
Gilg AB, Bearfield JC, Tittiger C, Welch WH, Blomquist GJ. Proc Natl Acad Sci USA 2005;102:9760–5.
Sitthithaworn W, Kojima N, Viroonchatapan E, Suh DY, Iwanami N, Hayashi T, et al. Chem Pharm Bull (Tokyo)
2001;49:197–202.
Jun L, Saiki R, Tatsumi K, Nakagawa T, Kawamukai M. Plant Cell Physiol 2004;45:1882–8.
Hemmi H, Noike M, Nakayama T, Nishino T. Eur J Biochem 2003;270:2186–94.
Kawasaki T, Hamano Y, Kuzuyama T, Itoh N, Seto H, Dairi T. J Biochem (Tokyo) 2003;133:83–91.
Rosenstiel TN, Fisher AJ, Fall R, Monson RK. Plant Physiol 2002;129:1276–84.
Winter H, Robinson DG, Heldt HW. Planta 1994;193:530–5.
Leidreiter K, Kruse A, Heineke D, Robinson DG, Heldt HW. Bot Acta 1995;108:439–44.
420
[115]
[116]
[117]
[118]
[119]
[120]
[121]
[122]
[123]
[124]
[125]
[126]
[127]
[128]
[129]
[130]
[131]
[132]
[133]
[134]
[135]
[136]
[137]
[138]
[139]
[140]
[141]
[142]
[143]
[144]
[145]
[146]
[147]
[148]
[149]
[150]
[151]
[152]
[153]
[154]
[155]
[156]
[157]
[158]
[159]
[160]
[161]
[162]
[163]
[164]
[165]
[166]
[167]
[168]
[169]
[170]
[171]
[172]
[173]
[174]
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
Ohnuma S, Koyama T, Ogura K. Biochem Biophys Res Commun 1993;192:407–12.
Erickson HK, Poulter CD. J Am Chem Soc 2003;125:6886–8.
Hemmerlin A, Rivera SB, Erickson HK, Poulter CD. J Biol Chem 2003;278:32132–40.
Sato M, Fujisaki S, Sato K, Nishimura Y, Nakano A. Genes Cells 2001;6:495–506.
Swiezewska E, Danikiewicz W. Prog Lipid Res 2005;44:235–58.
Cunillera N, Arro M, Fores O, Manzano D, Ferrer A. FEBS Lett 2000;477:170–4.
Fujihashi M, Zhang YW, Higuchi Y, Li XY, Koyama T, Miki K. Proc Natl Acad Sci USA 2001;98:4337–42.
Chang SY, Ko TP, Chen AP, Wang AH, Liang PH. Protein Sci 2004;13:971–8.
Chen AP, Chang SY, Lin YC, Sun YS, Chen CT, Wang AH, et al. Biochem J 2005;386:169–76.
Backhaus RA. Isr J Bot 1985;34:283–93.
Oh SK, Kang H, Shin DH, Yang J, Chow KS, Yeang HY, et al. J Biol Chem 1999;274:17132–8.
Dennis MS, Light DR. J Biol Chem 1989;264:18608–17.
Kim IJ, Ryu SB, Kwak YS, Kang H. J Exp Bot 2004;55:377–85.
Oh SK, Han KH, Ryu SB, Kang H. J Biol Chem 2000;275:18482–8.
Asawatreratanakul K, Zhang YW, Wititsuwannakul D, Wititsuwannakul R, Takahashi S, Rattanapittayaporn A, et al. Eur J
Biochem 2003;270:4671–80.
Sharkey TD, Yeh S. Annu Rev Plant Physiol Plant Mol Biol 2001;52:407–36.
Miller B, Oschinski C, Zimmer W. Planta 2001;213:483–7.
Sharkey TD, Yeh S, Wiberley AE, Falbel TG, Gong D, Fernandez DE. Plant Physiol 2005;137:700–12.
Sasaki K, Ohara K, Yazaki K. FEBS Lett 2005;579:2514–8.
Inoue T, Higuchi M, Hashimoto Y, Seki M, Kobayashi M, Kato T, et al. Nature 2001;409:1060–3.
Yamada H, Suzuki T, Terada K, Takei K, Ishikawa K, Miwa K, et al. Plant Cell Physiol 2001;42:1017–23.
Mok DSM, Mok MC. Annu Rev Plant Physiol Plant Mol Biol 2001;52:89–118.
Kasahara H, Takei K, Ueda N, Hishiyama S, Yamaya T, Kamiya Y, et al. J Biol Chem 2004;279:14049–54.
Masferrer A, Arro M, Manzano D, Schaller H, Fernandez-Busquets X, Moncalean P, et al. Plant J 2002;30:123–32.
Ashikari M, Sakakibara H, Lin S, Yamamoto T, Takashi T, Nishimura A, et al. Science 2005;309:741–5.
Barron D, Ibrahim RK. Phytochemistry 1996;43:921–82.
Asada Y, Li W, Yoshikawa T. Phytochemistry 2000;55:323–6.
Yamamoto H, Zhao P, Inoue K. Phytochemistry 2002;60:263–7.
Goese M, Kammhuber K, Bacher A, Zenk MH, Eisenreich W. Eur J Biochem 1999;263:447–54.
Stanjek V, Piel J, Boland W. Phytochemistry 1999;50:1141–5.
Barlow AJ, Becker H, Adam KP. Phytochemistry 2001;57:7–14.
Biggs DR, Welle JL, Grisebach H. Planta 1990;181:244–8.
Gershenzon J, Croteau R. Recent Adv Phytochem 1990;24:99–160.
Turner G, Gershenzon J, Nielson EE, Froehlich JE, Croteau R. Plant Physiol 1999;120:879–88.
Suire C, Bouvier F, Backhaus RA, Bégu D, Bonneu M, Camara B. Plant Physiol 2000;124:1–8.
Nah J, Song SJ, Back K. Plant Cell Physiol 2001;42:864–7.
Yang T, Li J, Wang HX, Zeng Y. Phytochemistry 2005;66:285–93.
Iijima Y, Gang DR, Fridman E, Lewinsohn E, Pichersky E. Plant Physiol 2004;134:370–9.
Dewick PM. Nat Prod Rep 2002;19:181–222.
Dewick PM. Nat Prod Rep 1997;14:111–44.
Bohlmann J, Meyer-Gauen G, Croteau R. Proc Natl Acad Sci USA 1998;95:4126–33.
Crock J, Wildung M, Croteau R. Proc Natl Acad Sci USA 1997;94:12833–8.
Steele CL, Crock J, Bohlmann J, Croteau R. J Biol Chem 1998;273:2078–89.
Aubourg S, Lecharny A, Bohlmann J. Mol Genet Genom 2002;267:730–45.
Chen F, Tholl D, DÕAuria JC, Farooq A, Pichersky E, Gershenzon J. Plant Cell 2003;15:481–94.
Tholl D, Chen F, Petri J, Gershenzon J, Pichersky E. Plant J 2005; Online publication date: 26-Apr-2005. doi: doi:10.1111/j.1365313X.2005.02419.x.
Starks CM, Back K, Chappell J, Noel JP. Science 1997;277:1815–20.
Caruthers JM, Kang I, Rynkiewicz MJ, Cane DE, Christianson DW. J Biol Chem 2000;275:25533–9.
Rynkiewicz MJ, Cane DE, Christianson DW. Proc Natl Acad Sci USA 2001;98:13543–8.
Cane DE. Chem Rev 1990;90:1089–103.
Calvert MJ, Ashton PR, Allemann RK. J Am Chem Soc 2002;124:11636–41.
Deligeorgopoulou A, Allemann RK. Biochemistry 2003;42:7741–7.
Piel J, Donath J, Bandemer K, Boland W. Angew Chem Int Ed 1998;37:2478–81.
Donath J, Boland W. J Plant Physiol 1994;143:473–8.
Degenhardt J, Gershenzon J. Planta 2000;210:815–22.
Arimura G, Kost C, Boland W. Biochim Biophys Acta 2005;1734:91–111.
Walling LL. J Plant Growth Regul 2000;19:195–216.
Baldwin IT, Preston CA. Plant 1999;208:137–45.
Pare PW, Tumlinson JH. Plant Physiol 1999;121:325–32.
Pichersky E, Gershenzon J. Curr Opin Plant Biol 2002;5:237–43.
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
[175]
[176]
[177]
[178]
[179]
[180]
[181]
[182]
[183]
[184]
[185]
[186]
[187]
[188]
[189]
[190]
[191]
[192]
[193]
[194]
[195]
[196]
[197]
[198]
[199]
[200]
[201]
[202]
[203]
[204]
[205]
[206]
[207]
[208]
[209]
[210]
[211]
[212]
[213]
[214]
[215]
[216]
[217]
[218]
[219]
[220]
[221]
[222]
[223]
[224]
[225]
[226]
[227]
[228]
[229]
[230]
[231]
[232]
[233]
[234]
[235]
421
Mattiacci L, Dicke M, Posthumus MA. J Chem Ecol 1994;20:2229–47.
Loreto F, Nascetti P, Graverini A, Mannozzi M. Funct Ecol 2000;14:589–95.
Singsaas E. New Phytol 2000;146:1–4.
Arimura G, Ozawa R, Shimoda T, Nishioka T, Boland W, Takabayashi J. Nature 2000;406:512–5.
Lange BM, Ghassemian M. Plant Mol Biol 2003;51:925–48.
Bohlmann J, Martin D, Oldham NJ, Gershenzon J. Arch Biochem Biophys 2000;375:261–9.
Faldt J, Arimura G, Gershenzon J, Takabayashi J, Bohlmann J. Planta 2003;216:745–51.
Van Poecke RMP, Posthumus MA, Dicke M. J Chem Ecol 2001;27:1911–28.
Arimura G, Huber DP, Bohlmann J. Plant J 2004;37:603–16.
Zhao J, Davis LC, Verpoorte R. Biotechnol Adv 2005;23:283–333.
Threlfall DR, Whitehead IM. Phytochemistry 1988;27:2565–80.
Vögeli U, Chappell J. Plant Physiol 1988;88:1291–6.
Brindle PA, Kuhn PJ, Threlfall DR. Phytochemistry 1988;27:133–50.
Hugueney P, Bouvier F, Badillo A, Quennemet J, dÕHarlingue A, Camara B. Plant Physiol 1996;111:619–26.
van der Heijden R, Threlfall DR, Verpoorte R, Whitehead IM. Phytochemistry 1989;28:2981–8.
Fulton DC, Kroon PA, Matern U, Threlfall DR, Whitehead IM. Phytochemistry 1993;34:139–45.
Phillips MA, Croteau RB. Trends Plant Sci 1999;4:184–90.
Tanaka S, Yamaura T, Shigemoto R, Tabata M. Phytochemistry 1989;28:2955–7.
Peer WA, Langenheim JH. Biochem Syst Ecol 1998;26:25–34.
Memelink J, Verpoorte R, Kijne JW. Trends Plant Sci 2001;6:212–9.
Chen XY, Chen Y, Henstein P, Davission VJ. Arch Biochem Biophys 1995;324:255–66.
Tan XP, Liang WQ, Liu CJ, Luo P, Heinstein P, Chen XY. Planta 2000;210:644–51.
Xu YH, Wang JW, Wang S, Wang JY, Chen XY. Plant Physiol 2004;135:507–15.
Lucker J, Bouwmeester HJ, Schwab W, Blaas J, van der Plas LH, Verhoeven HA. Plant J 2001;27:315–24.
Aharoni A, Giri AP, Deuerlein S, Griepink F, de Kogel WJ, Verstappen FW, et al. Plant Cell 2003;15:2866–84.
Lewinsohn E, Schalechet F, Wilkinson J, Matsui K, Tadmor Y, Nam KH, et al. Plant Physiol 2001;127:1256–65.
Lucker J, Schwab W, van Hautum B, Blaas J, van der Plas LH, Bouwmeester HJ, et al. Plant Physiol 2004;134:510–9.
Eccleston VS, Ohlrogge JB. Plant Cell 1998;10:613–22.
Bloch KE. Crit Rev Biochem 1983;14:47–92.
Ohvo-Rekilä H, Ramstedt B, Leppimäki P, Slotte Jp. Prog Lipid Res 2002;41:66–97.
Miao L, Nielsen M, Thewalt J, Ipsen JH, Bloom M, Zuckermann MJ, et al. Biophys J 2002;82:1429–44.
Barenholtz Y. Prog Lipid Res 2002;41:1–5.
Fielding JF, Fielding PE. Biochim Biophys Acta 2000;1529:210–22.
Parks L, Smith SJ, Crowley JH. Lipids 1995;30:227–30.
Silvius JR. Biochim Biophys Acta 2003;1610:174–83.
Hartmann MA. Trends Plant Sci 1998;2:137–43.
Caspi E. Tetrahedron 1986;42:3–50.
Volkman JK. Appl Microbiol Biotechnol 2003;60:495–506.
Nes WR. Adv Lipid Res 1977;15:233–324.
Benveniste P. Ann Rev Plant Physiol 1986;37:275–308.
Goad LJ. In: Charlwood BV, Banthorpe DV, editors. Methods in plant biochemistry, vol. 7. Academic Press; 1991. p. 369–434 [13
volumes].
Benveniste P. In: The arabidopsis book; 2002. Available from: www.aspb.org/publication..
Benveniste P. Annu Rev Plant Physiol Plant Mol Biol 2004;55:429–57.
Moebius FF, Fitzky U, Glossmann H. Trends Endoc Metab 2000;11:106–14.
Sakurai A. Plant Physiol Biochem 1999;37:351–61.
Mandava N. Annu Rev Plant Physiol Plant Mol Biol 1988;39:23–52.
Bishop GJ, Yokota T. Plant Cell Physiol 2001;42:114–20.
Chappell J. Curr Opin Plant Biol 2001;5:151–7.
Clouse SD. Plant Cell 2002;14:1995–2000.
Schaller H. Prog Lipid Res 2003;42:163–75.
Lindsey K, Pullen ML, Topping JF. Trends Plant Sci 2003;8:521–5.
Lees ND, Skaggs B, Kirsch DR, Bard M. Lipids 1995;30:221–6.
Bach TJ, Benveniste P. Prog Lipid Res 1997;36:197–226.
Benveniste P, Rahier A. In: Köller W, editor. Targets sites of fungicide action. Boca Raton: CRC Press; 1992. p. 207–25.
Goad LJ, Goodwin TW. Biochem J 1966;99:735–46.
Mercer EI. Prog Lipid Res 1993;32:357–416.
Rahier A, Taton M. Pestic Biochem Physiol 1997;57:1–27.
Woodward RB, Bloch KJ. J Am Chem Soc 1953;75:2023–4.
Eschenmoser A, Ruzicka L, Jeger O, Arigoni D. Helv Chim Acta 1955;38:1890–904.
Stork G, Burgenstahler AW. J Am Chem Soc 1955;77:5068–77.
Abe I, Rohmer M, Prestwich GD. Chem Rev 1993;93:2189–206.
422
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
[236]
[237]
[238]
[239]
[240]
[241]
[242]
[243]
[244]
[245]
[246]
[247]
[248]
[249]
[250]
[251]
[252]
[253]
[254]
[255]
[256]
[257]
[258]
[259]
[260]
[261]
[262]
[263]
[264]
[265]
[266]
[267]
[268]
[269]
[270]
[271]
[272]
[273]
[274]
[275]
[276]
[277]
[278]
[279]
[280]
[281]
[282]
[283]
[284]
[285]
[286]
[287]
[288]
[289]
[290]
[291]
[292]
[293]
[294]
[295]
[296]
[297]
Wendt KU, Schulz GE, Corey EJ, Liu DR. Angew Chem Int Ed 2000;39:2812–33.
Corey EJ, Matsuda SPT, Bartel B. Proc Natl Acad Sci USA 1993;90:11628–32.
Kushiro M, Shibuya M, Ebizuka Y. Eur J Biochem 1998;256:238–44.
Herrera JBR, Bartel B, Wilson WK, Matsuda SPT. Phytochemistry 1998;49:1905–98.
Husselstein-Muller T, Schaller H, Benveniste P. Plant Mol Biol 2001;42:75–92.
Kushiro T, Shibuya M, Masuda K, Ebizuka Y. Tetrahedron Lett 2000;41:7705–10.
Iturbe-Ormaetxe I, Haralampidis K, Papadopoulo K, Osbourn AE. Plant Mol Biol 2003;51:731–43.
Shibuya M, Adachi S, Ebizuka Y. Tetrahedron 2004;60:6995–7003.
Kushiro T, Shibuya M, Ebizuka Y. J Am Chem Soc 1999;121:1208–16.
Morita M, Shibuya M, Kushiro T, Masuda K, Ebizuka Y. Eur J Biochem 2000;267:3453–60.
Osbourn AE. Phytochemistry 2003;62:1–4.
Haralampidis K, Trojanowska M, Osbourn AE. Adv Biochem Eng Biotechnol 2002;75:31–49.
Qi X, Bakht S, Leggett M, Maxwell C, Melton R, Osbourn A. Proc Natl Acad Sci USA 2004;101:8233–8.
Wendt KU, Lenhart A, Schulz GE. J Mol Biol 2000;286:175–87.
Schulz-Gasch T, Stahl M. J Comput Chem 2003;24:741–53.
Hart EA, Hua L, Darr LB, Wilson WK, Pang JH, Matsuda SPT. J Am Chem Soc 1999;121:9887–8.
Meyer MM, Xu R, Matsuda SPT. Org Lett 2000;4:1395–8.
Segura MJR, Lodeiro S, Meyer MM, Patel AJ, Matsuda SPT. Org Lett 2002:4459–62.
Meyer MM, Segura MJR, Wilson WK, Matsuda SPT. Angew Chem 2000;112:4256–8.
Meyer MM, Segura MJR, Wilson WK, Matsuda SPT. Angew Chem Int Ed 2000;39:4090–2.
Lodeiro S, Segura M, Tahl M, Schulz-Gasch T, Matsuda SPT. Chembiochem 2004;5:1581–5.
Rahier A, Génot JC, Schuber F, Benveniste P, Narula AS. J Biol Chem 1984;259:15215–23.
Rahier A, Taton M, Bouvier-Navé P, Schmitt P, Benveniste P, Schuber F, et al. Lipids 1986;21:52–62.
Nes WD. Biochim Biophys Acta 2000;1529:63–88.
Nes WD. Phytochemistry 2003;64:75–95.
Avignan J, Steinberg D, Vroman HE, Thompson MJ, Mosettig E. J Biol Chem 1960;235:3123–30.
Rahier A, Narula AS, Benveniste P, Schmitt P. Biochem Biophys Res Commun 1980;92:20–5.
Rahier A, Taton M, Benveniste P. Biochem Soc Trans 1990;18:48–52.
Gössnitzer E, Malli R, Schuster S, Favre B, Ryder NS. Arch Pharm Pharm Med Chem 2002;11:535–46.
Debieu D, Bach J, Arnold A, Brousset S, Gredt M, Taton M, et al. Pest Biochem Physiol 2000;67:85–94.
Taton M, Rahier A. Arch Biochem Biophys 1996;325:279–88.
Kawata S, Trzaskos JM, Gaylor JL. J Biol Chem 1985;260:6609–17.
Osumi T, Nishimo T, Katsuki H. J Biochem 1979;85:819–26.
Rahier A, Smith M, Taton M. Biochem Biophys Res Commun 1997;236:434–7.
Reddy VVR, Kupfer D, Caspi E. J Biol Chem 1977;252:2797–801.
Gachotte D, Husselstein T, Bard M, Lacroute F, Benveniste P. Plant J 1996;9:391–8.
Arthington BA, Bennett LG, Skatrud PL, Guynn CJ, Barbuch RJ, Ulbright CE, et al. Gene 1991;102:39–44.
Matsushima M, Inazawa J, Takahashi E, Suzumori K, Nakamura Y. Cytogenet Cell Genet 1996;74:252–4.
Taton M, Husselstein T, Benveniste P, Rahier A. Biochemistry 2000;39:701–11.
Moënne-Loccoz P, Baldwin J, Ley BA, Loehr TM, Bollinger JM. J Biochemistry 1998;37:14659–63.
Nordlund P, Sjöberg BM, Eklund H. Nature 1990;345:593–8.
Rosenzweig AC, Frederick CA, Lippard SJ, Nordhund P. Nature 1993;366:537–43.
Shanklin J, Whittle E, Fox BG. Biochemistry 1994;33:12787–94.
Shanklin J, Cahoon Eb. Annu Rev Plant Physiol Plant Mol Biol 1998;49:611–41.
Gachotte D, Meens R, Benveniste P. Plant J 1995;8:407–16.
Pauling L. Chem Eng News 1946;24:1375–9.
Fersht Ar. In: Freeman W.H., editor, Enzyme structure and mechanism. 2nd ed. San Francisco and Oxford; 1985. p. 324–9..
Rahier A. Biochemistry 2001;40:256–67.
Goad LJ, Gibbons GF, Bolger LM, Rees HH, Goodwin TW. Biochem J 1969;114:885–92.
Goad LJ, Goodwin TW. Prog Phytochem 1972;3:113–98.
Akhtar M, Marsh S. Biochem J 1967;102:462–7.
Merkx M, Kopp DA, Sazinsky MH, Blazyk JL, Muller J, Lippard SJ. Angew Chem Int Ed Engl 2001;40:2782–807.
Rettie AE, Boberg M, Rettenmeier AW, Baillie TA. J Biol Chem 1988;263:13733–8.
Lee-Robichaud P, Shyadehi AZ, Wright JN, Akhtar ME, Akhtar M. Biochemistry 1995;34:14104–13.
Slaytor M, Bloch K. J Biol Chem 1965;240:4598–602.
Buist PH, Beharouzian B. J Am Chem Soc 1996;118:6295–6.
Beharouzian B, Buist PH. Curr Opin Chem Biol 2002;6:577–82.
Retey J. Angew Chem Int Ed Engl 1990;29:355–61.
Northrop Pb. Methods Enzymol 1982;87:607–25.
Bard M, Bruner Pa, Pierson Ca, Lees Nd, Biermann B, Frye L, et al. Proc Natl Acad Sci USA 1996;93:186–90.
Li Z, Kaplan J. J Biol Chem 1996;271:16927–33.
Ghisalberti EI, De Souza NJ, Rees HH, Goad LJ, Goodwin TW. J Chem Soc Chem Commun 1969;16:1403–5.
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
[298]
[299]
[300]
[301]
[302]
[303]
[304]
[305]
[306]
[307]
[308]
[309]
[310]
[311]
[312]
[313]
[314]
[315]
[316]
[317]
[318]
[319]
[320]
[321]
[322]
[323]
[324]
[325]
[326]
[327]
[328]
[329]
[330]
[331]
[332]
[333]
[334]
[335]
[336]
[337]
[338]
[339]
[340]
[341]
[342]
[343]
[344]
[345]
[346]
[347]
[348]
[349]
[350]
[351]
[352]
[353]
[354]
[355]
[356]
[357]
[358]
[359]
Knapp FF, Nicholas HJ. J Chem Soc Chem Commun 1970;19:399–400.
Pascal S, Taton M, Rahier A. J Biol Chem 1993;268:11639–54.
Trzaskos JM, Gaylor JL. Enzymes Biol Membr 1985;2:177–204.
Rahimtula AD, Gaylor JL. J Biol Chem 1972;247:9–15.
Gachotte D, Barbuch R, Gaylor J, Nickel E, Bard M. Proc Natl Acad Sci USA 1998;95:13794–9.
Rondet S, Taton M, Rahier A. Arch Biochem Biophys 1999;366:249–60.
Billheimer JT, Alcorn M, Gaylor JL. Arch Biochem Biophys 1981;211:430–8.
Pascal S, Taton M, Rahier A. Arch Biochem Biophys 1994;312:260–71.
Gachotte D, Sen SE, Bard M. Proc Natl Acad Sci USA 1999;96:12655–60.
Darnet S, Rahier A. Biochem Biophys Acta 2003;1633:106–17.
Darnet S, Bard M, Rahier A. FEBS Lett 2001;508:39–43.
Darnet S, Rahier A. Biochem J 2004;378:889–98.
Lund EG, Kerr TA, Sakai J, Li W, Russell DW. J Biol Chem 1998;273:34316–27.
Sperling P, Ternes P, Moll H, Franke S, Zähringer U. FEBS Lett 2001;494:90–4.
Mitchell AG. J Biol Chem 1995;270:29766–72.
Sperling P, Schmidt H, Heinz E. Eur J Biochem 1995;232:798–805.
Fox BG, Shanklin J, Somerville C, Münck E. Proc Natl Acad Sci USA 1993;90:2486–90.
Lindqvist Y, Huang W, Schenider G, Shanklin J. EMBO J 1996;15:4081–92.
Gachotte D, Eckstein J, Barbuch R, Hughes T, Roberts C, Bard M. J Lipid Res 2001;42:150–4.
Baudry K, Swain E, Rahier A, Germann M, Batta A, Rondet S, et al. J Biol Chem 2001;276:12702–11.
Mo C, Valachovic M, Randall SK, Nickels JT, Bard M. Proc Natl Acad Sci USA 2002;99:9739–44.
Debieu D, Bach J, Hugon M, Malosse C, Leroux P. Pest Manage Sci 2001;57:1060–7.
Nes WD, Jansen GG, Bergensträhle A. J Biol Chem 1991;266:15202–12.
Rahier A, Taton M, Benveniste P. Eur J Biochem 1989;181:615–26.
Taton M, Rahier A. Biochem J 1991;277:483–92.
Taton M, Benveniste P, Rahier A. Eur J Biochem 1989;185:605–14.
Rahier A, Taton M, Schmitt P, Benveniste P, Place P, Anding C. Phytochemistry 1985;24:1223–32.
Taton M, Benveniste P, Rahier A. Pest Sci 1987;21:269–80.
Taton M, Rahier A. Biochem Biophys Res Commun 1991;181:465–73.
Maillot-Vernier P, Gondet L, Schaller H, Benveniste P, Belliard G. Mol Gen Genet 1991;231:33–40.
Gondet L, Weber T, Maillot-Vernier P, Benveniste P, Bach TJ. Biochem Biophys Res Commun 1992;186:888–93.
Homberg N, Harker M, Wallace Ad, Clayton JC, Gibbard CL, Safford R. Plant J 2003;36:12–20.
Diener C, Haoxia L, Zhou W, Whoriskey WJ, Nes WD, Fink GR. Plant Cell 2000;12:853–70.
Schrick K, Mayer U, Horrichs A, Kuhnt C, Bellini C, Dangl J, et al. Genes Dev 2000;14:1471–84.
Willemsen V, Friml J, Grebe M, Van Den Toorn A, Palme K, Scheres B. Plant Cell 2003;15:612–25.
Schrick K, Mayer U, Martin G, Bellini C, Kuhnt C, Schmidt J, et al. Plant J 2002;31:61–73.
Jang J, Fujioka S, Tasaka M, Seto H, Takatsuto S, Ishii A, et al. Genes Dev 2000;14:1485–97.
Topping J, May V, Muskett P, Lindsey K. Development 1997;124:4415–24.
Souter M, Topping J, Pullen M, Friml J, Palme K, Hackett R, et al. Plant Cell 2002;14:1017–31.
Schmitt P, Scheid F, Benveniste P. Phytochemistry 1980;19:525–30.
Bouvier-Nave P, Husselstein T, Desprez T, Benveniste P. Eur J Biochem 1997;246:518–29.
Bouvier-Nave P, Husselstein T, Benveniste P. Eur J Biochem 1998;256:88–96.
Heintz R, Benveniste P. J Biol Chem 1974;249:4267–74.
Rahier A, Cattel L, Benveniste P. Phytochemistry 1977;16:1187–92.
Goldstein Jl, Brown MS. Nature 1990;343:425–30.
Brown MS, Goldstein JL. Cell 1997;89:331–40.
Karst F, Lacroute F. Mol Gen Genet 1977;154:269–77.
Kennedy MA, Barbuch R, Bard M. Biochim Biophys Acta 1999;1445:110–22.
Schaller H, Grausem B, Benveniste P, Chye ML, Tan CT, Song YH, et al. Plant Physiol 1995;109:761–70.
Chappell J, Wolf F, Proulx J, Cuellar R, Saunders C. Plant Physiol 1995;109:1337–43.
Wentzinger LF, Bach T, Hartmann MA. Plant Physiol 2002;130:334–46.
Devarenne TP, Ghosh A, Chappell J. Plant Physiol 2002;129:1095–106.
Hartmann MA, Benveniste P. Phytochemistry 1974;13:2667–72.
Holmberg N, Harker M, Gibbard CL, Wallace AD, Clayton JC, Rawlins S, et al. Plant Physiol 2002;130:303–11.
Podust LM, Poulos TL, Waterman MR. Proc Natl Acad Sci USA 2001;98:3068–73.
Lutke-Brinkhaus F, Liedvogel B, Kleinig H. Eur J Biochem 1984;141:537–41.
Meganathan R. FEMS Microbiol Lett 2001;203:131–9.
Disch A, Hemmerlin A, Bach TJ, Rohmer M. Biochem J 1998;331:615–21.
Cunillera N, Boronat A, Ferrer A. J Biol Chem 1997;272:15381–8.
Swiezewska E, Dallner G, Andersson B, Ernster L. J Biol Chem 1993;268:1494–9.
Wanke M, Dallner G, Swiezewska E. Biochim Biophys Acta 2000;1463:188–94.
Do TQ, Hsu AY, Jonassen T, Lee PT, Clarke CF. J Biol Chem 2001;276:18161–8.
423
424
[360]
[361]
[362]
[363]
[364]
[365]
[366]
[367]
[368]
[369]
[370]
[371]
[372]
[373]
[374]
[375]
[376]
[377]
[378]
[379]
[380]
[381]
[382]
[383]
[384]
[385]
[386]
[387]
[388]
[389]
[390]
[391]
[392]
[393]
[394]
[395]
[396]
[397]
[398]
[399]
[400]
[401]
[402]
[403]
[404]
[405]
[406]
[407]
[408]
[409]
[410]
[411]
[412]
[413]
[414]
[415]
[416]
[417]
[418]
[419]
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
Jonassen T, Marbois BN, Faull KF, Clarke CF, Larsen PL. J Biol Chem 2002;277:45020–7.
Hirooka K, Izumi Y, An CI, Nakazawa Y, Fukusaki E, Kobayashi A. Biosci Biotechnol Biochem 2005;69:592–601.
Ohara K, Kokado Y, Yamamoto H, Sato F, Yazaki K. Plant J 2004;40:734–43.
Hsu AY, Do TQ, Lee PT, Clarke CF. Biochim Biophys Acta 2000;1484:287–97.
Okada K, Ohara K, Yazaki K, Nozaki K, Uchida N, Kawamukai M, et al. Plant Mol Biol 2004;55:567–77.
Jonassen T, Larsen PL, Clarke CF. Proc Natl Acad Sci USA 2001;98:421–6.
Levavasseur F, Miyadera H, Sirois J, Tremblay ML, Kita K, Shoubridge E, et al. J Biol Chem 2001;276:46160–4.
Schindler S, Lichtenthaler HK. In: Siegenthaler PA, Eichenberger W, editors. Structure, function and metabolism of plant lipids.
Amsterdam: Elsevier Science Publishers B.V.; 1984. p. 273–6.
Okada K, Suzuki K, Kamiya Y, Zhu X, Fujisaki S, Nishimura Y, et al. Biochim Biophys Acta 1996;1302:217–23.
Avelange-Macherel MH, Joyard J. Plant J 1998;14:203–13.
Guan X, Diez T, Prasad TK, Nikolau BJ, Wurtele ES. Arch Biochem Biophys 1999;362:12–21.
Anderson MD, Che P, Song J, Nikolau BJ, Wurtele ES. Plant Physiol 1998;118:1127–38.
Che P, Wurtele ES, Nikolau BJ. Plant Physiol 2002;129:625–37.
Fall RR, Hector ML. Biochemistry 1977;16:4000–5.
MacMillan J, Beale MH. In: Comprehensive natural products chemistry No. 2, Isoprenoid including carotenoids and steroids.
Oxford: Elsevier Science; 1999. p. 217–43.
Peters RJ, Croteau RB. Biochemistry 2002;41:1836–42.
Kawaide H, Imai R, Sassa T, Kamiya Y. J Biol Chem 1997;272:21706–12.
Kawaide H, Sassa T, Kamiya Y. J Biol Chem 2000;275:2276–80.
Peters RJ, Ravn MM, Coates RM, Croteau RB. J Am Chem Soc 2001;123:8974–8.
Wildung MR, Croteau R. J Biol Chem 1996;271:9201–4.
Schepmann HG, Pang J, Matsuda SP. Arch Biochem Biophys 2001;392:263–9.
Helliwell CA, Sullivan JA, Mould RM, Gray JC, Peacock WJ, Dennis ES. Plant J 2001;28:201–8.
Sakamoto T, Miura K, Itoh H, Tatsumi T, Ueguchi-Tanaka M, Ishiyama K, et al. Plant Physiol 2004;134:1642–53.
Grayer RJ, Kokubun T. Phytochemistry 2001;56:253–63.
Nemoto T, Cho EM, Okada A, Okada K, Otomo K, Kanno Y, et al. FEBS Lett 2004;571:182–6.
Cho EM, Okada A, Kenmoku H, Otomo K, Toyomasu T, Mitsuhashi W, et al. Plant J 2004;37:1–8.
Prisic S, Xu M, Wilderman PR, Peters RJ. Plant Physiol 2004;136:4228–36.
Otomo K, Kenmoku H, Oikawa H, Konig WA, Toshima H, Mitsuhashi W, et al. Plant J 2004;39:886–93.
Richman AS, Gijzen M, Starratt AN, Yang Z, Brandle JE. Plant J 1999;19:411–21.
Munne-Bosch S, Alegre L. Plant Physiol 2001;125:1094–102.
Munne-Bosch S, Alegre L. Planta 2000;210:925–31.
Munne-Bosch S, Alegre L. Plant Physiol 2003;131:1816–25.
Seo S, Seto H, Koshino H, Yoshida S, Ohashi Y. Plant Cell 2003;15:863–73.
Jasinski M, Stukkens Y, Degand H, Purnelle B, Marchand-Brynaert J, Boutry M. Plant Cell 2001;13:1095–107.
van den Brule S, Muller A, Fleming AJ, Smart CC. Plant J 2002;30:649–62.
Bouvier F, Keller Y, dÕHarlingue A, Camara B. Biochim Biophys Acta 1998;1391:320–8.
Addlesee HA, Hunter CN. J Bacteriol 1999;181:7248–55.
Addlesee HA, Gibson LCD, Jensen PE, Hunter CN. FEBS Lett 1996;389:126–30.
Rudiger W, Hedden P, Kost HP, Chapman DJ. Biochem Biophys Res Commun 1977;74:1268–72.
Keller Y, Bouvier F, dÕHarlingue A, Camara B. Eur J Biochem 1998;251:413–7.
Addlesee HA, Hunter CN. J Bacteriol 2002;184:1578–86.
Tanaka R, Oster U, Kruse E, Rudiger W, Grimm B. Plant Physiol 1999;120:695–707.
Havaux M, Lutz C, Grimm B. Plant Physiol 2003;132:300–10.
Grasses T, Grimm B, Koroleva O, Jahns P. Planta 2001;213:620–8.
Shibata M, Tsuyama M, Takami T, Shimizu H, Kobayashi Y. J Exp Bot 2004;55:1989–96.
Shibata M, Mikota T, Yoshimura A, Iwata N, Tsuyama M, Kobayashi Y. Plant Sci 2004;166:593–600.
Shpilyov AV, Zinchenko VV, Shestakov SV, Grimm B, Lokstein H. Biochim Biophys Acta 2005;1706:195–203.
Porter JW, Spurgeon SL. Pure Appl Chem 1979;51:609–22.
Camara B. Pure Appl Chem 1985;57:675–7.
Dogbo O, Laferrière A, dÕHarlingue A, Camara B. Proc Natl Acad Sci USA 1988;85:7054–8.
Dogbo O, Camara B. Plant Sci 1987;49:103–9.
Iwata-Reuyl D, Math SK, Desai SB, Poulter CD. Biochemistry 2003;42:3359–65.
Nakashima T, Inoue T, Oka A, Nishino T, Osumi T, Hata S. Proc Natl Acad Sci USA 1995;92:2328–32.
Blagg BS, Jarstfer MB, Rogers DH, Poulter CD. J Am Chem Soc 2002;124:8846–53.
Jarstfer MB, Zhang DL, Poulter CD. J Am Chem Soc 2002;124:8834–45.
Pandit J, Danley DE, Schulte GK, Mazzalupo S, Pauly TA, Hayward CM, et al. J Biol Chem 2000;275:30610–7.
Takaichi S, Inoue K, Akaike M, Kobayashi M, Oh-oka H, Madigan MT. Arch Microbiol 1997;168:277–81.
Taylor RF. Microbiol Rev 1984;48:181–98.
Wieland B, Feil C, Gloria-Maercker E, Thumm G, Lechner M, Bravo JM, et al. J Bacteriol 1994;176:7719–26.
Umeno D, Tobias AV, Arnold FH. J Bacteriol 2002;184:6690–9.
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
[420]
[421]
[422]
[423]
[424]
[425]
[426]
[427]
[428]
[429]
[430]
[431]
[432]
[433]
[434]
[435]
[436]
[437]
[438]
[439]
[440]
[441]
[442]
[443]
[444]
[445]
[446]
[447]
[448]
[449]
[450]
[451]
[452]
[453]
[454]
[455]
[456]
[457]
[458]
[459]
[460]
[461]
[462]
[463]
[464]
[465]
[466]
[467]
[468]
[469]
[470]
[471]
[472]
[473]
[474]
[475]
[476]
[477]
[478]
425
Raisig A, Sandmann G. Biochim Biophys Acta 2001;1533:164–70.
Casida JE. Environ Health Perspect 1980;34:189–202.
Metzger P, Largeau C. In: Cohen Z, editor. Chemicals from microalgae. London: Taylor & Francis; 1999. p. 205–60.
Rivera SB, Swedlund BD, King GJ, Bell RN, Hussey Jr CE, Shattuck-Eidens DM. Proc Natl Acad Sci USA 2001;98:4373–8.
Shen B, Zheng Z, Dooner HK. Proc Natl Acad Sci USA 2000;97:14807–12.
Sanmiya K, Ueno O, Matsuoka M, Yamamoto N. Plant Cell Physiol 1999;40:348–54.
Zhao P, Inoue K, Kouno I, Yamamoto H. Plant Physiol 2003;133:1306–13.
Huang Z, Poulter CD. J Am Chem Soc 1989;111:2713–5.
Okada S, Devarenne TP, Murakami M, Abe H, Chappell J. Arch Biochem Biophys 2004;422:110–8.
Zhang D, Jennings SM, Robinson GW, Poulter CD. Arch Biochem Biophys 1993;304:133–43.
Camara B, Bardat F, Moneger R. Eur J Biochem 1982;127:255–8.
Gallagher CE, Matthews PD, Li F, Wurtzel ET. Plant Physiol 2004;135:1776–83.
Burkhardt PK, Beyer P, Wunn J, Kloti A, Armstrong GA, Schledz M, et al. Plant J 1997;11:1071–8.
Fray RG, Wallace A, Fraser PD, Valero D, Hedden P, Bramley PM, et al. Plant J 1995;8:693–701.
Shewmaker CK, Sheehy JA, Daley M, Colburn S, Ke DY. Plant J 1999;20:401–12.
Lindgren LO, Stalberg KG, Hoglund AS. Plant Physiol 2003;132:779–85.
Ducreux LJ, Morris WL, Hedley PE, Shepherd T, Davies HV, Millam S, et al. J Exp Bot 2005;56:81–9.
Fraser PD, Romer S, Shipton CA, Mills PB, Kiano JW, Misawa N, et al. Proc Natl Acad Sci USA 2002;99:1092–7.
Tsuchiya T, Takaichi S, Misawa N, Maoka T, Miyashita H, Mimuro M. FEBS Lett 2005;579:2125–9.
Nakamura Y, Kaneko T, Sato S, Mimuro M, Miyashita H, Tsuchiya T, et al. DNA Res 2003;10:137–45.
Cunningham Jr FX, Sun Z, Chamovitz D, Hirschberg J, Gantt E. Plant Cell 1994;6:1107–21.
Stickforth P, Steiger S, Hess WR, Sandmann G. Arch Microbiol 2003;179:409–15.
Frigaard NU, Maresca JA, Yunker CE, Jones AD, Bryant DA. J Bacteriol 2004;186:5210–20.
Hausmann A, Sandmann G. Fungal Genet Biol 2000;30:147–53.
Camara B, Monéger R. Physiol Vég 1982;20:757–73.
Koyama Y, Fuji R. In: Frank HA, Young AJ, Britton G, Cogdell RJ, editors. The photochemistry of carotenoids, vol. 8. Dordrecht:
Kluwer Academic Publishers; 1999. p. 161–88.
Park H, Kreunen SS, Cuttriss AJ, DellaPenna D, Pogson BJ. Plant Cell 2002;14:321–32.
Isaacson T, Ronen G, Zamir D, Hirschberg J. Plant Cell 2002;14:333–42.
Isaacson T, Ohad I, Beyer P, Hirschberg J. Plant Physiol 2004;136:4246–55.
Masamoto K, Hisatomi S, Sakurai I, Gombos Z, Wada H. Plant Cell Physiol 2004;45:1325–9.
Roszak AW, McKendrick K, Gardiner AT, Mitchell IA, Isaacs NW, Cogdell RJ, et al. Structure (Camb) 2004;12:765–73.
Dayan FE, Duke SO. In: Plimmer JR, Gammon DW, Ragsdale NN, Roberts T, editors. Encyclopedia of agrochemicals, vol. 2. New
York: John Wiley & Sons; 2003. p. 744–9.
Misawa N, Yamano S, Linden H, de Felipe MR, Lucas M, Ikenaga H, et al. Plant J 1993;4:833–40.
Wagner T, Windhovel U, Romer S. Z Naturforsch [C] 2002;57:671–9.
Arias RS, Netherland MD, Scheffler BE, Puri A, Dayan FE. Pest Manage Sci 2005;61:258–68.
Michel A, Arias RS, Scheffler BE, Duke SO, Netherland M, Dayan FE. Mol Ecol 2004;13:3229–37.
Mayer MP, Beyer P, Kleinig H. Eur J Biochem 1990;191:141–50.
Norris SR, Barrette TR, DellaPenna D. Plant Cell 1995;7:2139–49.
Sandmann G, Mitchell G. J Agric Food Chem 2001;49:138–41.
Breitenbach J, Zhu C, Sandmann G. J Agric Food Chem 2001;49:5270–2.
Camara B, Dogbo O, dÕHarlingue A, Kleinig H, Monéger R. Biochim Biophys Acta 1985;836:262–6.
Camara B, Dogbo O. Plant Physiol 1986;80:172–4.
Bellamacina CR. FASEB J 1996;10:1257–69.
Hornero-Mendez D, Britton G. FEBS Lett 2002;515:133–6.
Verdoes JC, Krubasik KP, Sandmann G, van Ooyen AJ. Mol Gen Genet 1999;262:453–61.
Krubasik P, Sandmann G. Mol Gen Genet 2000;263:423–32.
Hemmi H, Ikejiri S, Nakayama T, Nishino T. Biochem Biophys Res Commun 2003;305:586–91.
Peck RF, Johnson EA, Krebs MP. J Bacteriol 2002;184:2889–97.
Bouvier F, dÕHarlingue A, Camara B. Arch Biochem Biophys 1997;346:53–64.
Bouvier F, Hugueney P, dÕHarlingue A, Kuntz A, Camara B. Plant J 1994;6:45–54.
Cunningham FXJ, Pogson B, Sun ZR, McDonald KA, Dellapenna D, Gantt E. Plant Cell 1996;8:1613–26.
Cunningham Jr FX, Gantt E. Proc Natl Acad Sci USA 2001;98:2905–10.
Hess RH, Rocap G, Ting CS, Larimer F, Stilwagen S, Lamerdin J, et al. Photosynth Res 2001;70:53–71.
Teramoto M, Takaichi S, Inomata Y, Ikenaga H, Misawa N. FEBS Lett 2003;545:120–6.
Tao L, Picataggio S, Rouviere PE, Cheng Q. Mol Genet Genom 2004;271:180–8.
Camara B, Dogbo O, dÕHarlingue A, Bardat F. Phytochemistry 1985;24:2751–2.
Camara B, dÕHarlingue A, Dogbo O, Agnès C, Hénaff J. C R Acad Sci Paris 1985;300(ser III):297–300.
Yokoyama H, Hsu WJ, Poling SM, Hayman E. In: Brotton G, Goodwin TW, editors. Carotenoid chemistry and biochemistry.
Oxford, New York: Pergamon Press; 1982. p. 371–85.
Fedtke C, Depka B, Schallner O, Tietjen K, Trebst A, Wollweber D, et al. Pest Manage Sci 2001;57:278–82.
426
[479]
[480]
[481]
[482]
[483]
[484]
[485]
[486]
[487]
[488]
[489]
[490]
[491]
[492]
[493]
[494]
[495]
[496]
[497]
[498]
[499]
[500]
[501]
[502]
[503]
[504]
[505]
[506]
[507]
[508]
[509]
[510]
[511]
[512]
[513]
[514]
[515]
[516]
[517]
[518]
[519]
[520]
[521]
[522]
[523]
[524]
[525]
[526]
[527]
[528]
[529]
[530]
[531]
[532]
[533]
[534]
[535]
[536]
[537]
[538]
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
Sun Z, Gantt E, Cunningham Jr FX. J Biol Chem 1996;271:24349–52.
Tao L, Cheng Q. Mol Genet Genom 2004;272:530–7.
Fernandez-Gonzalez B, Sandmann G, Vioque A. J Biol Chem 1997;272:9728–33.
Misawa N, Satomi Y, Kondo K, Yokoyama A, Kajiwara S, Saito T, et al. J Bacteriol 1995;177:6575–84.
Kajiwara S, Kakizono T, Saito T, Kondo K, Ohtani T, Nishio N, et al. Plant Mol Biol 1995;29:343–52.
Steiger S, Sandmann G. Biotechnol Lett 2004;26:813–7.
Lohr M, Im CS, Grossman AR. Plant Physiol 2005;138:490–515.
Gorton HL, Vogelmann TC. Photochem Photobiol 2003;77:608–15.
Gorton HL, Williams WE, Vogelmann TC. Photochem Photobiol 2001;73:611–20.
Gribanovski-Sassu O. Phytochemistry 1972;11:3195–8.
Bouvier F, Backhaus RA, Camara B. J Biol Chem 1998;273:30651–9.
Behrouzian B, Savile CK, Dawson B, Buist PH, Shanklin J. J Am Chem Soc 2002;124:3277–83.
Broadwater JA, Whittle E, Shanklin J. J Biol Chem 2002;277:15613–20.
Broun P, Shanklin J, Whittle E, Somerville C. Science 1998;282:1315–7.
Topham RW, Gaylor JL. Biochem Biophys Res Commun 1972;47:180–6.
Topham RW, Gaylor JL. Biochem Biophys Res Commun 1967;27:644–9.
Nishida Y, Adachi K, Kasai H, Shizuri Y, Shindo K, Sawabe A, et al. Appl Environ Microbiol 2005;71:4286–96.
Cunningham Jr FX, Gantt E. Plant J 2005;41:478–92.
Pogson B, McDonald KA, Truong M, Britton G, DellaPenna D. Plant Cell 1996;8:1627–39.
Tian L, Musetti V, Kim J, Magallanes-Lundback M, DellaPenna D. Proc Natl Acad Sci USA 2004;101:402–7.
Blasco F, Kauffmann I, Schmid RD. Appl Microbiol Biotechnol 2004;64:671–4.
Marin E, Nussaume L, Quesada A, Gonneau M, Sotta B, Hugueney P, et al. EMBO J 1996;15:2331–42.
Buch K, Stransky H, Hager A. FEBS Lett 1995;376:45–8.
Bouvier F, dÕHarlingue A, Hugueney P, Marin E, Marion-Poll A, Camara B. J Biol Chem 1996;271:28861–7.
Bugos RC, Yamamoto HY. Proc Natl Acad Sci USA 1996;93:6320–5.
Hieber AD, Bugos RC, Yamamoto HY. Biochim Biophys Acta 2000;1482:84–91.
Flower DR. Biochem J 1996;318:1–14.
Yamamoto HY, Higashi RM. Arch Biochem Biophys 1978;190:514–22.
Bratt CE, Arvidsson PO, Carlsson M, Akerlund HE. Photosynth Res 1995;45:169–75.
Leipner J, Stamp P, Fracheboud Y. Planta 2000;210:964–9.
Havir EA, Tauska SL, Peterson RB. Plant Sci 1997;123:57–66.
Hager A, Holocher K. Planta 1994;192:581–9.
Bungard RA, Ruban AV, Hibberd JM, Press MC, Horton P, Scholes JD. Proc Natl Acad Sci USA 1999;96:1135–9.
Garcı́a-Plazaola JI, Hernández A, Errasti E, Becerril JM. Funct Plant Biol 2002;29:1075–80.
Matsubara S, Morosinotto T, Bassi R, Christian AL, Fischer-Schliebs E, Luttge U, et al. Planta 2003;217:868–79.
Garcı́a-Plazaola JI, Hormaetxe K, Hernández A, Olano JM, Becerril JM. Funct Plant Biol 2004;31:815–23.
Goss R. Planta 2003;217:801–12.
Frommolt R, Goss R, Wilhelm C. Planta 2001;213:446–56.
Bouvier F, dÕHarlingue A, Backhaus RA, Kumagai MH, Camara B. Eur J Biochem 2000;267:6346–52.
Al-Babili S, Hugueney P, Schledz M, Welsch R, Frohnmeyer H, Laule O, et al. FEBS Lett 2000;485:168–72.
Ronen G, Carmel-Goren L, Zamir D, Hirschberg J. Proc Natl Acad Sci USA 2000;97:11102–7.
Yamamoto H, Bugos RC, Hieber AD. In: Frank HA, Young AJ, Britton G, Cogdell RJ, editors. The photochemistry of
carotenoids, vol. 8. Dordrecht: Kluwer Academic Publishers; 1999. p. 293–303.
Olaizola M, La Roche J, Kolber Z, Falkowski PG. Photosynth Res 1994;41:357–70.
Muller P, Li XP, Niyogi KK. Plant Physiol 2001;125:1558–66.
Li XP, Bjorkman O, Shih C, Grossman AR, Rosenquist M, Jansson S, et al. Nature 2000;403:391–5.
Niyogi KK, Grossman AR, Bjorkman O. Plant Cell 1998;10:1121–34.
Chang SH, Sun WH, Yamamoto HY. Photosynth Res 2000;64:95–103.
Conklin PL, Pallanca JE, Last RL, Smirnoff N. Plant Physiol 1997;115:1277–85.
Pastori GM, Kiddle G, Antoniw J, Bernard S, Veljovic-Jovanovic S, Verrier PJ, et al. Plant Cell 2003;15:939–51.
Muller-Moule P, Conklin PL, Niyogi KK. Plant Physiol 2002;128:970–7.
Kreuz K, Beyer P, Kleinig H. Planta 1982;154:66–9.
De la Guardia MD, Aragon CM, Murillo FJ, Cerda-Olmedo E. Proc Natl Acad Sci USA 1971;68:2012–5.
Morosinotto T, Baronio R, Bassi R. J Biol Chem 2002;277:36913–20.
Yamamoto HY, Chenchin EE, Yamada DK. In: Avron M, editor. Proceedings of the third international congress on
photosynthesis. Amsterdam: Elsevier Scientific Publishing Company; 1974. p. 1999–2006.
Aspinall-OÕDea M, Wentworth M, Pascal A, Robert B, Ruban A, Horton P. Proc Natl Acad Sci USA 2002;99:16331–5.
Hieber AD, Kawabata O, Yamamoto HY. Plant Cell Physiol 2004;45:92–102.
Bauer C, Elsen S, Swem LR, Swem DL, Masuda S. Philos Trans R Soc Lond B Biol Sci 2003;358:147–53 [discussion 153–144].
Zeilstra-Ryalls JH, Kaplan S. Cell Mol Life Sci 2004;61:417–36.
Yurkov VV, Beatty JT. Microbiol Mol Biol Rev 1998;62:695–724.
Jaubert M, Zappa S, Fardoux J, Adriano JM, Hannibal L, Elsen S, et al. J Biol Chem 2004;279:44407–16.
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
[539]
[540]
[541]
[542]
[543]
[544]
[545]
[546]
[547]
[548]
[549]
[550]
[551]
[552]
[553]
[554]
[555]
[556]
[557]
[558]
[559]
[560]
[561]
[562]
[563]
[564]
[565]
[566]
[567]
[568]
[569]
[570]
[571]
[572]
[573]
[574]
[575]
[576]
[577]
[578]
[579]
[580]
[581]
[582]
[583]
[584]
[585]
[586]
[587]
[588]
[589]
[590]
[591]
[592]
[593]
[594]
[595]
[596]
[597]
427
Giraud E, Hannibal L, Fardoux J, Jaubert M, Jourand P, Dreyfus B, et al. J Biol Chem 2004;279:15076–83.
Terao J. Lipids 1989;24:659–61.
Woodall AA, Lee SW, Weesie RJ, Jackson MJ, Britton G. Biochim Biophys Acta 1997;1336:33–42.
Yeliseev AA, Kaplan S. J Biol Chem 1995;270:21167–75.
Gomelsky M, Horne IM, Lee HJ, Pemberton JM, McEwan AG, Kaplan S. J Bacteriol 2000;182:2253–61.
Ponnampalam SN, Bauer CE. J Biol Chem 1997;272:18391–6.
Elsen S, Ponnampalam SN, Bauer CE. J Biol Chem 1998;273:30762–9.
Taylor BL, Zhulin IB. Microbiol Mol Biol Rev 1999;63:479–506.
Giraud E, Fardoux J, Fourrier N, Hannibal L, Genty B, Bouyer P, et al. Nature 2002;417:202–5.
Giraud E, Zappa S, Jaubert M, Hannibal L, Fardoux J, Adriano JM, et al. Photochem Photobiol Sci 2004;3:587–91.
Davis SJ, Vener AV, Vierstra RD. Science 1999;286:2517–20.
Bugos RC, Chang SH, Yamamoto HY. Plant Physiol 1999;121:207–14.
Corona V, Aracri B, Kosturkova G, Bartley GE, Pitto L, Giorgetti L, et al. Plant J 1996;9:505–12.
von Lintig J, Welsch R, Bonk M, Giuliano G, Batschauer A, Kleinig H. Plant J 1997;12:625–34.
Wetzel CM, Rodermel SR. Plant Mol Biol 1998;37:1045–53.
Romer S, Fraser PD. Planta 2005;221:305–8.
Woitsch S, Romer S. Plant Physiol 2003;132:1508–17.
Davison PA, Hunter CN, Horton P. Nature 2002;418:203–6.
Gotz T, Sandmann G, Romer S. Plant Mol Biol 2002;50:129–42.
Pogson BJ, Rissler HM. Philos Trans R Soc Lond B Biol Sci 2000;355:1395–403.
Hormaetxe K, Hernandez A, Becerril JM, Garcia-Plazaola JI. Plant Biol (Stuttg) 2004;6:325–32.
Ronen G, Cohen M, Zamir D, Hirschberg J. Plant J 1999;17:341–51.
Thomas RL, Jen JJ. Plant Physiol 1975;56:452–3.
Alba R, Cordonnier-Pratt MM, Pratt LH. Plant Physiol 2000;123:363–70.
Camara B, Hugueney P, Bouvier F, Kuntz M, Monéger R. Inter Rev Cytol 1995;163:175–247.
Deruère J, Römer S, dÕHarlingue A, Backhaus RA, Kuntz M, Camara B. Plant Cell 1994;6:119–33.
Rabbani S, Beyer P, Lintig J, Hugueney P, Kleinig H. Plant Physiol 1998;116:1239–48.
Vishnevetsky M, Ovadis M, Vainstein A. Trends Plant Sci 1999;4:232–5.
Li L, Paolillo DJ, Parthasarathy MV, Dimuzio EM, Garvin DF. Plant J 2001;26:59–67.
Li L, Garvin DF. Genome 2003;46:588–94.
Lu S, OÕHalloran DM, Cosman KM, Yang Y, Xiang B, Thannhauser TW, et al. In: American society of plant biologists annual
meeting; 2005. p. 606.
Liu YS, Gur A, Ronen G, Causse M, Damidaux R, Buret M, et al. Plant Biotechnol 2003;1:195–207.
Wann EV, Jourdain EL, Pressay R, Lyon BG. J Am Soc Hort Sci 1985;110:212–5.
Peters JL, Schreuder MEL, Verduin SJW, Kendrick RE. Photochem Photobiol 1992;56:75–82.
Mustilli AC, Fenzi F, Ciliento R, Alfano F, Bowler C. Plant Cell 1999;11:145–57.
Levin I, Frankel P, Gilboa N, Tanny S, Lalazar A. Theor Appl Genet 2003;106:454–60.
Schroeder DF, Gahrtz M, Maxwell BB, Cook RK, Kan JM, Alonso JM, et al. Curr Biol 2002;12:1462–72.
Lieberman M, Segev O, Gilboa N, Lalazar A, Levin I. Theor Appl Genet 2004;108:1574–81.
Liu Y, Roof S, Ye Z, Barry C, van Tuinen A, Vrebalov J, et al. Proc Natl Acad Sci USA 2004;101:9897–902.
Davuluri GR, van Tuinen A, Mustilli AC, Manfredonia A, Newman R, Burgess D, et al. Plant J 2004;40:344–54.
Davuluri GR, van Tuinen A, Fraser PD, Manfredonia A, Newman R, Burgess D, et al. Nat Biotechnol 2005;23:890–5.
Hardtke CS, Deng XW. Plant Physiol 2000;124:1548–57.
Bino RJ, Ric de Vos CH, Lieberman M, Hall RD, Bovy A, Jonker HH, et al. New Phytol 2005;166:427–38.
Mochizuki T, Kamimura S. In: 9th Meeting of the Eucarpia Tomato Workshop, Eucarpia Tomato Working Group, Wageningen,
The Netherlands; 1984. p. 8–13..
Giliberto L, Perrotta G, Pallara P, Weller JL, Fraser PD, Bramley PM, et al. Plant Physiol 2005;137:199–208.
Yamamoto HY, Bassi R. In: Oert DR, Yocum CF, editors. Oxygenic photosynthesis. Dordrecht, The Netherlands: Kluwer
Academic Publishers; 1996. p. 539–63.
Milon A, Wolff G, Ourisson G, Nakatani Y. Helv Chim Acta 1986;69:12–24.
McAuley KE, Fyfe PK, Ridge JP, Cogdell RJ, Isaacs NW, Jones MR. Biochemistry 2000;39:15032–43.
Yeates TO, Komiya H, Chirino A, Rees DC, Allen JP, Feher G. Proc Natl Acad Sci USA 1988;85:7993–7.
Garcia-Asua G, Cogdell RJ, Hunter CN. Mol Microbiol 2002;44:233–44.
Kumagai MH, Keller Y, Bouvier F, Clary D, Camara B. Plant J 1998;14:305–15.
Croce R, Weiss S, Bassi R. J Biol Chem 1999;274:29613–23.
Hobe S, Niemeier H, Bender A, Paulsen H. Eur J Biochem 2000;267:616–24.
Caffarri S, Croce R, Breton J, Bassi R. J Biol Chem 2001;276:35924–33.
Gastaldelli M, Canino G, Croce R, Bassi R. J Biol Chem 2003;278:19190–8.
Bishop NI, Bulga B, Senger H. Bot Acta 1998;111:231–5.
Heinze I, Pfündel E, Hühn M, Dau H. Biochim Biophys Acta 1997;1320:188–94.
Polle JE, Niyogi KK, Melis A. Plant Cell Physiol 2001;42:482–91.
Pogson BJ, Niyogi KK, Bjorkman O, DellaPenna D. Proc Natl Acad Sci USA 1998;95:13324–9.
428
[598]
[599]
[600]
[601]
[602]
[603]
[604]
[605]
[606]
[607]
[608]
[609]
[610]
[611]
[612]
[613]
[614]
[615]
[616]
[617]
[618]
[619]
[620]
[621]
[622]
[623]
[624]
[625]
[626]
[627]
[628]
[629]
[630]
[631]
[632]
[633]
[634]
[635]
[636]
[637]
[638]
[639]
[640]
[641]
[642]
[643]
[644]
[645]
[646]
[647]
[648]
[649]
[650]
[651]
[652]
[653]
[654]
[655]
[656]
[657]
[658]
[659]
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
Tardy F, Havaux M. J Photochem Photobiol B 1996;34:87–94.
Hurry V, Anderson JM, Chow WS, Osmond CB. Plant Physiol 1997;113:639–48.
Havaux M, DallÕOsto L, Cuine S, Giuliano G, Bassi R. J Biol Chem 2004;279:13878–88.
Phillip D, Hobe S, Paulsen H, Molnar P, Hashimoto H, Young AJ. J Biol Chem 2002;277:25160–9.
Lagarde D, Vermaas W. FEBS Lett 1999;454:247–51.
Matsubara S, Gilmore AM, Osmond B. Aust J Plant Physiol 2001;28:793–800.
Snyder AM, Clark BM, Robert B, Ruban AV, Bungard RA. J Biol Chem 2004;279:5162–8.
Wang Y, Hu X. J Am Chem Soc 2002;124:8445–51.
Camara B, Bouvier F. Arch Biochem Biophys 2004;430:16–21.
Bouvier F, Isner JC, Dogbo O, Camara B. Trends Plant Sci 2005;10:187–94.
Rock CD, Zeevaart JAD. Proc Natl Acad Sci USA 1991;88:7496–9.
Parry AD, Horgan R. Phytochemistry 1991;30:815–21.
Winterhalter P, Rouseff RL. Carotenoid-derived aroma compounds, ACS symposium series, Washington DC; 2002..
Schwartz SH, Qin X, Zeevaart JAD. J Biol Chem 2001;276:25208–11.
Bouvier F, Suire C, Mutterer J, Camara B. Plant Cell 2003;15:47–62.
Bouvier F, Dogbo O, Camara B. Science 2003;300:2089–91.
Leuenberger MG, Engeloch-Jarret C, Woggon WD. Angew Chem Int Ed 2001;40:2614–7.
Lindqvist A, Andersson S. J Biol Chem 2002;277:23942–8.
During A, Smith MK, Piper JB, Smith JC. J Nutr Biochem 2001;12:640–7.
Kloer DP, Ruch S, Al-Babili S, Beyer P, Schulz GE. Science 2005;308:267–9.
Poliakov E, Gentleman S, Cunningham Jr FX, Miller-Ihli NJ, Redmond TM. J Biol Chem 2005;280:29217–23.
Booker J, Sieberer T, Wright W, Williamson L, Willett B, Stirnberg P, et al. Dev Cell 2005;8:443–9.
Foo E, Bullier E, Goussot M, Foucher F, Rameau C, Beveridge CA. Plant Cell 2005;17:464–74.
Snowden KC, Simkin AJ, Janssen BJ, Templeton KR, Loucas HM, Simons JL, et al. Plant Cell 2005;17:746–59.
Matile P, Hortensteiner S, Thomas H. Annu Rev Plant Physiol Plant Mol Biol 1999;50:67–95.
Eckhardt U, Grimm B, Hortensteiner S. Plant Mol Biol 2004;56:1–14.
Oster U, Bauer CE, Rüdiger W. J Biol Chem 1997;272:9671–6.
Katz JJ, Strain HH, Harkness AL, Studier MH, Svec WA, Janson TR, et al. J Am Chem Soc 1972;94:7938–9.
Glaeser J, Baneras L, Rutters H, Overmann J. Arch Microbiol 2002;177:475–85.
Akhtar M, Ajaz AA, Corina DL. Biochem J 1984;224:187–94.
Emery VC, Akhtar M. Biochemistry 1987;26:1200–8.
Hortensteiner S, Wuthrich KL, Matile P, Ongania KH, Krautler B. J Biol Chem 1998;273:15335–9.
Hinder B, Schellenberg M, Rodoni S, Ginsburg S, Vogt E, Martinoia E, et al. J Biol Chem 1996;271:27233–6.
Tommasini R, Vogt E, Fromenteau M, Hortensteiner S, Matile P, Amrhein N, et al. Plant J 1998;13:773–80.
Hortensteiner S, Chinner J, Matile P, Thomas H, Donnison IS. Plant Mol Biol 2000;42:439–50.
Pruzinska A, Tanner G, Aubry S, Anders I, Moser S, Muller T, et al. Plant Physiol 2005. doi:10.1104/pp.105.065870.
Himelblau E, Amasino RM. J Plant Physiol 2001;158:1317–23.
Hortensteiner S, Feller U. J Exp Bot 2002;53:927–37.
Rontani JF, Aubert C. Rapid Commun Mass Spectrom 2005;19:637–46.
Rontani JF, Cuny P, Grossi V. Phytochemistry 1996;42:347–51.
Kreuz K, Kleinig H. Plant Cell Rep 1984;1:40–2.
Lutzow M, Kleinig H. Arch Biochem Biophys 1990;277:94–100.
Dogbo O, Bardat F, Camara B. Physiol Vég 1984;22:75–82.
Soll J, Kemmerling M, Schultz G. Arch Biochem Biophys 1980;204:544–50.
Schultz G, Soll J, Fiedler E, Schulze-Siebert D. Physiol Plant 1985;64:123–9.
Camara B, Bardat F, Sèye A, dÕHarlingue A, Monéger R. Plant Physiol 1982;70:1562–3.
dÕHarlingue A, Camara B. J Biol Chem 1985;260:15200–3.
DellaPenna D. J Plant Physiol 2005;162:729–37.
Dahnhardt D, Falk J, Appel J, van der Kooij TA, Schulz-Friedrich R, Krupinska K. FEBS Lett 2002;523:177–81.
Garcia I, Rodgers M, Lenne C, Rolland A, Sailland A, Matringe M. Biochem J 1997;325:761–9.
Norris SR, Shen X, DellaPenna D. Plant Physiol 1998;117:1317–23.
Meazza G, Scheffler BE, Tellez MR, Rimando AM, Romagni JG, Duke SO, et al. Phytochemistry 2002;60:281–8.
Matringe M, Sailland A, Pelissier B, Rolland A, Zink O. Pest Manage Sci 2005;61:269–76.
Moran GR. Arch Biochem Biophys 2005;433:117–28.
Soeda T, Uchida T. Pestic Biochem Physiol 1987;29:35–42.
Sandmann G, Böger P, Kumita I. Pestic Sci 1990;30:353–5.
Ong ASH. In: Packer L, Fuchs J, editors. Vitamin E in health and disease. New York: Marcel Dekker, Inc.; 1993. p. 3–8.
Shintani DK, Cheng Z, DellaPenna D. FEBS Lett 2002;511:1–5.
Cheng Z, Sattler S, Maeda H, Sakuragi Y, Bryant DA, DellaPenna D. Plant Cell 2003;15:2343–56.
Porfirova S, Bergmuller E, Tropf S, Lemke R, Dormann P. Proc Natl Acad Sci USA 2002;99:12495–500.
Hutson KG, Threlfall DR. Biochem Biophys Acta 1980;632:630–48.
Collakova E, DellaPenna D. Plant Physiol 2001;127:1113–24.
F. Bouvier et al. / Progress in Lipid Research 44 (2005) 357–429
[660]
[661]
[662]
[663]
[664]
[665]
[666]
[667]
[668]
[669]
[670]
[671]
[672]
[673]
[674]
[675]
[676]
[677]
[678]
[679]
[680]
[681]
[682]
[683]
[684]
[685]
[686]
[687]
[688]
[689]
[690]
[691]
[692]
[693]
[694]
[695]
429
Hunter S, Cahoon EB. In: Midwest American Society of plant biologists sectional meeting; 2005. p. 66.
Rippert P, Scimemi C, Dubald M, Matringe M. Plant Physiol 2004;134:92–100.
Kanwischer M, Porfirova S, Bergmuller E, Dormann P. Plant Physiol 2005;137:713–23.
Cahoon EB, Hall SE, Ripp KG, Ganzke TS, Hitz WD, Coughlan SJ. Nat Biotechnol 2003;21:1082–7.
Provencher LM, Miao L, Sinha N, Lucas WJ. Plant Cell 2001;13:1127–41.
Hofius D, Hajirezaei MR, Geiger M, Tschiersch H, Melzer M, Sonnewald U. Plant Physiol 2004;135:1256–68.
Rolland F, Moore B, Sheen J. Plant Cell 2002;14(Suppl.):S185–205.
Sattler SE, Cahoon EB, Coughlan SJ, DellaPenna D. Plant Physiol 2003;132:2184–95.
Sattler SE, Gilliland LU, Magallanes-Lundback M, Pollard M, DellaPenna D. Plant Cell 2004;16:1419–32.
Munne-Bosch S. New Phytol 2005;166:363–6.
Yamamoto Y, Kobayashi Y, Matsumoto H. Plant Physiol 2001;125:199–208.
Fryer MJ. Plant Cell Environ 1992;15:381–92.
Munné-Bosch S, Alegre L. Crit Rev Plant Sci 2002;21:31–57.
Trebst A, Depka B, Hollander-Czytko H. FEBS Lett 2002;516:156–60.
Sandorf I, Hollander-Czytko H. Planta 2002;216:173–9.
Falk J, Krauss N, Dahnhardt D, Krupinska K. J Plant Physiol 2002;159:1245–53.
Giannino D, Condello E, Bruno L, Testone G, Tartarini A, Cozza R, et al. J Exp Bot 2004;55:2063–73.
Maeda H, Sakuragi Y, Bryant DA, Dellapenna D. Plant Physiol 2005;138:1422–35.
McCarthy SS, Kobayashi MC, Niyogi KK. Genetics 2004;168:1249–57.
Sattler SE, Cheng Z, DellaPenna D. Trends Plant Sci 2004;9:365–7.
Van Eenennaam AL, Li G, Venkatramesh M, Levering C, Gong X, Jamieson AC, et al. Metab Eng 2004;6:101–8.
Johnson TW, Shen G, Zybailov B, Kolling D, Reategui R, Beauparlant S, et al. J Biol Chem 2000;275:8523–30.
Wade Johnson T, Naithani S, Stewart Jr C, Zybailov B, Daniel Jones A, Golbeck JH, et al. Biochim Biophys Acta 2003;1557:67–76.
Shimada H, Ohno R, Shibata M, Ikegami I, Onai K, Ohto MA, et al. Plant J 2005;41:627–37.
Johnson TW, Zybailov B, Jones AD, Bittl R, Zech S, Stehlik D, et al. J Biol Chem 2001;276:39512–21.
Sakuragi Y, Zybailov B, Shen G, Jones AD, Chitnis PR, van der Est A, et al. Biochemistry 2002;41:394–405.
Sakuragi Y, Zybailov B, Shen G, Bryant DA, Golbeck JH, Diner BA, et al. J Biol Chem 2005;280:12371–81.
Motohashi R, Ito T, Kobayashi M, Taji T, Nagata N, Asami T, et al. Plant J 2003;34:719–31.
Dreses-Werringloer U, Fischer K, Wachter E, Link TA, Flugge UI. Eur J Biochem 1991;195:361–8.
Kumar R, Raclaru M, Schusseler T, Gruber J, Sadre R, Luhs W, et al. FEBS Lett 2005;579:1357–64.
Olejnik D, Gogolewski M, Nogala-Kalucka M. Nahrung 1997;41:101–4.
Ohnuma SI, Koyama T, Ogura K. J Biol Chem 1991;266:23706–13.
Hirooka K, Bamba T, Fukusaki E, Kobayashi A. Biochem J 2003;370:679–86.
Bovy A, de Vos R, Kemper M, Schijlen E, Almenar Pertejo M, Muir S, et al. Plant Cell 2002;14:2509–26.
van der Fits L, Memelink J. Science 2000;289:295–7.
Yano M. Curr Opin Plant Biol 2001;4:130–5.