ArticlePDF Available

Cloning and Functional Analysis of a cDNA Encoding Ginkgo biloba Farnesyl Diphosphate Synthase

Authors:

Abstract and Figures

Farnesyl diphosphate synthase (FPS; EC2.5.1.1/EC2. 5.1.10) catalyzes the synthesis of farnesyl diphosphate, and provides precursor for biosynthesis of sesquiterpene and isoprenoids containing more than 15 isoprene units in Ginkgo biloba. Here we report the cloning, characterization and functional analysis of a new cDNA encoding FPS from G. biloba. The full-length cDNA (designated GbFPS) had 1731 bp with an open reading frame of 1170 bp encoding a polypeptide of 390 amino acids. The deduced GbFPS was similar to other known FPSs and contained all the conserved regions of trans-prenyl chain-elongating enzymes. Structural modeling showed that GbFPS had the typical structure of FPS, the most prominent feature of which is the arrangement of 13 core helices around a large central cavity. Southern blot analysis revealed a small FPS gene family in G. biloba. Expression analysis indicated that GbFPS expression was high in roots and leaves, and low in stems. Functional complementation of GbFPS in an FPS-deficient strain confirmed that GbFPS mediates farnesyl diphosphate biosynthesis.
Content may be subject to copyright.
Mol. Cells, Vol. 18, No. 2, pp. 150-156
Cloning and Functional Analysis of a cDNA Encoding
Ginkgo biloba Farnesyl Diphosphate Synthase
Peng Wang
1
, Zhihua Liao
1
, Liang Guo
1
, Wenchao Li
1
, Min Chen
2
, Yan Pi
1
, Yifu Gong
3
, Xiaofen Sun
1
, and
Kexuan Tang
1,3,
*
1
State Key Laboratory of Genetic Engineering, School of Life Sciences, Fudan-SJTU-Nottingham Plant Biotechnology R&D Center,
Morgan-Tan International Center for Life Sciences, Fudan University, Shanghai 200433, China;
2
School of Pharmacy, Fudan University, Shanghai 200032, China;
3
Plant Biotechnology Research Center, Fudan-SJTU-Nottingham Plant Biotechnology Research and Development Center, School of
Agriculture and Biology, Shanghai Jiaotong University, Shanghai 200030, China.
(Received March 30, 2004; Accepted June 1, 2004)
Farnesyl diphosphate synthase (FPS; EC2.5.1.1/EC2.
5.1.10)
catalyzes the synthesis of farnesyl diphosphate,
and provides precursor for biosynthesis of sesquiter-
pene and isoprenoids containing more than 15 iso-
prene units in Ginkgo biloba. Here we report the clon-
ing, characterization and functional analysis of a new
cDNA encoding FPS from G. biloba. The full-length
cDNA (designated GbFPS) had 1731 bp with an open
reading frame of 1170 bp encoding a polypeptide of
390 amino acids. The deduced GbFPS was similar to
other known FPSs and contained all the conserved
regions of trans-prenyl chain-elongating enzymes. Struc-
tural modeling showed that GbFPS had the typical
structure of FPS, the most prominent feature of which
is the arrangement of 13 core helices around a large
central cavity. Southern blot analysis revealed a small
FPS gene family in G. biloba. Expression analysis in-
dicated that GbFPS expression was high in roots and
leaves, and low in stems. Functional complementation
of GbFPS in an FPS-deficient strain confirmed that
GbFPS mediates farnesyl diphosphate biosynthesis.
Keywords: Bilobalide; Farnesyl Diphosphate Synthase;
Ginkgo biloba; RACE; Yeast Complementation.
Introduction
Isoprenoids are the most widespread and structurally di-
verse family of natural products. They are present in all
* To whom correspondence should be addressed.
Tel: 86-21-65642772; Fax: 86-21-65643552
E-mail: kxtang1@yahoo.com
plants. Many play important roles as growth hormones,
photosynthetic pigments and membrane components, and
are also involved in defense and communication. Some
are commercially important compounds such as flavors,
fragrances, medicines, and natural rubbers.
Isoprenoids are derived from a common precursor,
isopentenyl diphosphate (IPP). The major steps involved
in isoprenoid biosynthesis are sequential condensations of
IPP with a growing allylic polyisoprenoid diphosphate.
These chain elongation reactions are catalyzed by a fam-
ily of enzymes designated prenyltransferases. Farnesyl
diphosphate synthase (FPS) is a representative of this
family (Kellogg and Poulter, 1997).
FPS catalyzes the sequential head-to-tail coupling of
DMAPP with two molecules of IPP, producing farnesyl
diphosphate (FPP), which lies at the branch point of the
pathway of isoprenoid synthesis. FPP has three functions.
First, it is the precursor of a structurally diverse class of
sesquiterpenes that are widely distributed in plant king-
dom including a diversity of functionally important com-
pounds, such as phytoalexins, antibiotic compounds that
respond to microbial challenge, and anti-feedants that dis-
courage opportunistic herbivory. In addition, sesquiter-
penes are found in essential oils and act as anticancer and
antimalarial drugs. Second, FPP is the precursor of
squalene synthase which condenses two molecules of FPP
to produce squalene, the committed precursor for triter
Abbreviations: CTAB, cetyltrimethylammonium bromide; DMAPP,
dimethylallyl diphosphate; DXP, 1-deoxyxylulose-5-phosphate;
FPP, farnesyl diphosphate; FPS, farnesyl diphosphate synthase;
GbFPS, farnesyl diphosphate synthase from Ginkgo biloba; GPP,
geranyl diphosphate; IPP, isopentenyl diphosphate; ORF, open
reading frame; RACE, rapid amplification of cDNA ends.
Molecules
and
Cells
KSMCB 2004
Peng Wang et al. 151
pene synthesis. This large class of molecules includes the
brassinosteroids, certain phytoalexins, and various toxins
and components of surface waxes. Finally, FPP provides
precursors for further chain elongation in the biosynthesis
of diterpenes, tetraterpenes and polyterpenes (Croteau et
al., 2000). Because of its crucial status, FPS has been ex-
tensively and thoroughly studied. It has been purified
from a number of organisms and its three-dimensional
structure has been characterized by Tarshis et al. (1994).
cDNAs encoding FPS have been obtained from various
angiosperms, such as Arabidopsis thaliana (Cunillera et
al., 1997), Artemisia annua (Hemmerlin et al., 2003), and
Zea mays (Li and Larkins, 1996). However, as far as we
know, there has been no report of the cloning of FPS from
gymnosperms.
The gymnosperm, Ginkgo biloba, one of the oldest liv-
ing tree species in the world, is the sole extant representa-
tive of the order Ginkgoles. It thrived 125 million years
ago, and the genus has remained virtually unchanged
from that time (Biswas and Johri, 1997). G. biloba is a
great survivor in polluted environments (Kim et al., 1997),
and it is exceptionally resistant to fungal (Aoki, 1997) and
insect attack (Honda, 1997). It is now considered a me-
dicinal plant in the United States and Europe, as has been
recognized as such since ancient times in East Asia
(Hasebe, 1997). The tree produces a variety of secondary
metabolites, among which the terpene trilactones, gink-
golides and bilobalide, are the most characteristic. Gink-
golides are diterpenes that have been detected both in
roots and leaves, and the sesquiterpene bilobalide is the
major terpene trilactone in leaves and a minor component
in roots. These compounds all utilize FPP as precursor
(Carrier et al., 1998; Park et al., 2004). Ginkgolides are a
structurally unique family of diterpenoids that are highly
specific platelet-activating factor receptor antagonists
(Guinot and Braquet, 1994), and bilobalide has neuropro-
tective effects and is effective in neurodegenerative dis-
eases (Thompson, 1995). It is important to elucidate the
pathway of biosynthesis of these terpene trilactones in G.
biloba at both the genetic and enzymatic level because of
their pharmaceutical status and potential biological func-
tion. There have been a few reports of the cloning of
genes involved in the biosynthesis of isoprenes such as
ginkgolide in G. biloba, (Liao et al., 2004; Schepmann et
al., 2001) but they are far from exhaustive. Here, we re-
port the cloning and characterization of a full-length
cDNA encoding G. biloba FPS. We also studied its role
using a complementation test in a mutant yeast strain.
Materials and Methods
Plant material, yeast strains and culture methods Ginkgo
biloba was grown on the campus of Fudan University, Shanghai,
China, and fresh roots, stems and leaves were collected, frozen
immediately in liquid nitrogen and stored at 80
o
C prior to total
RNA extraction.
Saccharomyces cerevisiae strain CC25 (MATa/MATalpha,
deltaERG20/+) was obtained from the American Type Culture
Collection (ATCC number 4021258; for further details see Cu-
nillera et al., 2000; Winzeler et al., 1999 and the ATCC website:
www.atcc.org). The strain was maintained in YPD medium with
1% (w/v) yeast extract, 2% (w/v) bactopeptone and 2% (w/v)
glucose. Unless otherwise stated, the cells were grown at 28°C
in liquid culture or on 2% agar. When required, ergosterol (80
mg/ml in agar plates) was added to the growth medium.
Cloning of full-length GbFPS cDNA by rapid amplification of
cDNA ends (RACE) Total RNA was isolated from young leaves
of G. biloba with TRIzol reagent according to the manufac-
turer’s instructions (Invitrogen, USA). Single-stranded cDNAs
were reverse transcribed from 5 µg of total RNA with an oligo(dT)
primer according to the manufacturer’s protocol (PowerScript,
CLONTECH, USA). After RNaseH treatment, the single-stranded
cDNA mixtures were used as templates for polymerase chain
reaction (PCR). Two degenerate oligonucleotide primers, DFFPS
(5-TGGTG(C/T)AT(A/T/C)GAATGGCT(T/C)CA(A/G)GC-3)
and DRFPS (5-TC(A/G)TCCTG(A/T/C/G)ACTTG(A/G)AA(A/G)
TA(T/G)(A/G)T(A/T/C)CCCAT-3), encoding the highly con-
served amino acid sequences (GWCVEWLQ and MG(I/T)
YFQVQDD) of plant FPSs were synthesized and used in gradient
PCR-amplification of the core cDNA fragment. The gradient PCR
was carried out by denaturing the cDNA at 94°C for 3 min fol-
lowed by 29 cycles of amplification (94°C for 45 s, 5159°C for
45 s, 72°C for 1 min) and extension at 72°C for 6 min. The core
fragment was amplified at the annealing temperature of 55.5°C,
and subcloned into pGEM T-easy vector (Promega, USA). Se-
quencing and a blast-n search confirmed that it was highly ho-
mologous to other plant FPS genes. The core fragment was sub-
sequently used to design and synthesize gene-specific primers to
clone the full-length cDNA by RACE.
A SMART RACE cDNA Amplification Kit (Clontech,
USA) was used to isolate the 3- and 5-ends of GbFPS cDNA.
First-strand 3-RACE-ready and 5-RACE-ready cDNA samples
from G. biloba were prepared according to the manufacturer’s
protocol (SMART RACE cDNA Amplification Kit, User
Manual, Clontech, USA) and used as templates for 3-RACE
and 5-RACE, respectively. The 3-end of GbFPS cDNA was
amplified using 3-gene-specific primers and the universal prim-
ers provided by Clontech. For the first round PCR amplification
of 3-RACE, GBFPS3-1 (5-GGACAATTTTGTAGCTGTCAA-
GAAC-3) and UPM (Universal Primer Mix, provided by Clon-
tech) were used as the PCR primers (3-RACE), and 3-RACE-
ready cDNA was used as template. For nested PCR amplifica-
tion of 3-RACE, GBFPS3-2 (5-TGCAGATGGGAACCTA-
TTTTCAAG-3) and NUP (Nested Universal Primer, provided
by Clontech) were used as PCR primers (3-RACE), and the
first round PCR products were used as templates. The 5-end of
GbFPS cDNA was amplified using 5-gene-specific primers and
the universal primers provided. For the first round PCR amplifi-
152 FPS from Ginkgo biloba
cation of 5-RACE, GBFPS5-1 (5-TGTGAGACCCATCCATA-
ATGTCATC-3) and UPM were used as primers (5-RACE),
and 5-RACE-ready cDNA, as template. For the nested PCR
amplification of 5-RACE, GBFPS5-2 (5-AATACAAGGAAA-
TAGGCTTGAAGCC-3) and NUP were used as primers (5-
RACE) and the first round PCR products, as templates. For the
first and nested PCR amplification of 3 and 5-ends of GbFPS
cDNA, we used an Advantage 2 PCR Kit (CLONTECH, USA)
and carried out PCR with 25 cycles of amplification (30 s at
94°C, 30 s at 68°C, 3 min at 72°C). The 3 and 5-RACE prod-
ucts were subcloned into pGEM T-easy vector followed by se-
quencing. By assembling the sequences of the 3 and 5 RACE
products and the core fragment on Contig Express (Vector NTI
Suite 6.0), we deduced the full-length cDNA sequence of
GbFPS and subsequently amplified it by PCR using primers
FGBFPS (5-CCAATCTCTTACTAGTTCACAGATATTC-3)
and RGBFPS (5- TCCGCTTTCTTCAATCCAAGAAAG-3) in
the following conditions: 3 min at 94°C followed by 29 cycles
of amplification (50 s at 94°C, 50 s at 60°C, 3 min at 72°C) and
6 min at 72°C. The amplified PCR product was purified and
cloned into pGEM T-easy vector (Promega, USA) and se-
quenced. In total, we picked three independent positive clones
and sequenced them to confirm the sequence and avoid PCR
errors. The ORF of GbFPS was predicted by ORF Finder at
NCBI (http://www.ncbi.nlm.nih.gov/gorf/gorf.html).
Sequence analysis We analuzed the sequence of GbFPS online at
the websites http://www.ncbi.nlm.nih.gov and http://cn.expasy.org.
SubLoc 1.1 (Hua and Sun, 2001) was used to predict subcellular
localization and CLUSTRAL X (Thompson et al., 1997) for mul-
tiple alignment of GbFPS and other FPSs. Homology-based struc-
tural modeling was performed by Swiss-Model (Peitsch and Jon-
geneel, 1993; Schwede et al., 2003), and Web Lab View Lite 4.0
was used to display 3-D structures. We constructed the phyloge-
netic tree of FPSs using MEGA2 by the neighbor-joining method
with 1,000 repeats (Kumar et al., 2001).
Southern blot analyses Aliquots of DNA (20 µg/sample) were
digested overnight at 37°C with HindIII and NcoI respectively,
which does not cut within the probe region, fractionated by
0.85% agarose gel electrophoresis and transferred to a Hybond-
N
+
nylon membrane (Amersham Pharmacia, UK). The 461-bp
probe was generated by PCR using the 1173 bp coding sequence
of GbFPPS as template with primers FFPSPIN (5-TGGTGC-
ATTGAATGGCTTCAAGCCT-3) and RFPSPIN (5-TCATCC-
TGTACTTGAAAATAGGTTCCCA-3). Probe labeling (biotin),
hybridization and signal detection were performed using Gene
Images random prime labeling module and CDP-Star detection
module following the manufacturer’s instructions (Amersham
Pharmacia, UK). The film was washed under stringent condi-
tions (60°C) and signals were visualized by exposure to Fuji X-
ray film at room temperature for 1.5 h.
Semi-quantitative RT-PCR We extracted total RNA from
roots, stems and leaves by a modified CTAB method. The qual-
ity and concentration of the RNA samples were examined by
EB-stained agarose gel electrophoresis and spectrophotometric
analysis. After establishing agreement between the OD values of
the RNAs from the three tissues, we performed RT-PCR using a
One Step RT-PCR Kit (TaKaRa, Japan) with the primer pairs
RTF1 (5-AAGCTTATGCAATTCCCTTCTCTCAGAAAA-3)
and RTR1 (5-AATACAAGGAAATAGGCTTGAAGCC-3). The
procedure was: 50°C for 30 min, 94°C for 2 min followed by 20
cycles of amplification (94°C for 30 s, 60°C for 30 s, 72°C for 1
min). A fragment of 18S rRNA (108 bp) was also amplified in
the RT-PCR with primers 18SF1 (5-ATGATAACTCGACG-
GATCGC-3) and 18SR1 (5-CTTGGATGTGGTAGCCGTTT-
3), as control. The densities of the target bands were measured
with a Furi FR-200A ultraviolet analyzer (Furi Tech., China).
Functional complementation of GbFPS in yeast Saccharomy-
ces cerevisiae strain CC25 was used to test the activity of
GbFPS. A fragment containing the coding region of GbFPS was
amplified by PCR using the pair of primers (5-AAGCTTAT-
GCAATTCCCTTCTCTCAGAAAA-3 and 5-GGATCCCTA-
CTTCTGTCGCTTGTATATCTTC-3). After digesting the frag-
ment with HindIII and BamHI, the coding region was cloned
into the expression vector pYES2 (Invitrogen, USA), to yield
pYES2 + GbFPS, which was used to transform E. coli DH5α
cells. Transformants were selected on solid LB medium contain-
ing carbenicillin (125 mg/L), and the extracted plasmids were
used to transform S. cerevisiae CC25. Positive colonies were
selected by plating on YPD agar at 42°C for 16 h, followed by
37°C for 2 d.
Results and Discussion
Molecular cloning of full-length GbFPS cDNA Using
total RNA isolated from young leaves of G. biloba and
the degenerate sense primer, dffps, and anti-sense primer,
drfps, we amplified a 461-bp product by RT-PCR, ligated
it into pGEM T-easy vector and sequenced it. A BLAST-
n search of the sequence showed that the fragment was
highly homologous to FPS genes from other plant species
(data not shown). We synthesized gene-specific primers
from the sequence and used them to generate 5-end and
3-end DNA fragments which were subsequently ampli-
fied by RACE (see Materials and Methods). By aligning
the sequences obtained, we deduced the full-length cDNA
sequence of GbFPS (GenBank accession number AY
389818) and amplified it. The full-length cDNA of
GbFPS had 1731 bp with an ORF of 1170 bp, encoding a
polypeptide of 390 amino acids, flanked by a 172-bp 5-
untranslated region containing a TATA box, and a 389-bp
3-untranslated region including a poly(A) tail of 28 bp.
The predicted GbFPS protein had a calculated molecular
mass of 44.68 kDa and a theoretical pI of 5.75.
Comparison of FPS from G. biloba and those of other
Peng Wang et al. 153
Fig. 1. Alignment of the deduced amino acid sequences of GbFPS
and FPSs of other model organisms. Identical and conserved
amino acid residues are denoted by black and gray backgrounds,
respectively. The five conserved domains of prenyltransferases
are boxed and numbered. The highly conserved aspartate-rich
motifs (DDXX(XX)D) is present in domains II and V. GenBank
accession numbers: A. thaliana, AAL17614; O. sativa, T03687; H.
sapiens, NP_001995; D. melanogaster, CAA08919; S. cerevisiae,
P08524; E. coli, BAA00599.
species, and construction of a phylogenetic tree A
BLAST search of the protein database at NCBI showed
that the polypeptide sequence of G. biloba (GbFPS) had
4776% identity and 6488% similarity with FPS’s from
other species. The polypeptide sequence of G. biloba FPS
(GbFPS) was aligned with FPS sequences from Arabidop-
sis thaliana, Oryza sativa, Homo sapiens and Saccharo-
myces cerevisiae. The result showed that GbFPS had the
five conserved regions identified by Chen et al. (1994)
that are characteristic of prenyltransferases that synthesize
isoprenoid diphosphates with E-double bonds (Fig. 1).
The highly conserved aspartate-rich motif DDXX(XX)D
was present in domains II and V. All other amino acids
known to be important for FPS activity or substrate bind-
ing were present.
Phylogenetic tree analysis showed that GbFPS ap-
peared at the base of the clade of the plant kingdom, and
that FPSs evolved vertically from a common ancestor (Fig.
Fig. 2. Phylogenetic tree of the amino acid sequences of FPSs of
different organisms constructed by the neighbor-joining method
on MEGA 2. Bacteria-derived FPSs are indicated by ; animal-
derived FPSs by ▲; fungi-derived FPSs by ; plant-derived
FPSs by . GenBank accession numbers: Lupinus albus, P49351;
Gossypium arboreum, CAA72793; Hevea brasiliensis, AY135188;
Mentha piperita, AAK63847; Arabidopsis thaliana, AAL17614;
Capsicum annuum, CAA59170; Lycopersicon esculentum, T06272;
Malus domestica, AAM08927; Helianthus annuus, AAC78557;
Artemisia annua, P49350; Parthenium argentatum, CAA57892;
Humulus lupulus, BAB40665; Eucommia ulmoides, AB052681;
Zea mays, T03291; Oryza sativa, T03687; G. biloba, AY389818;
Mucor circinelloides, CAD42869; Saccharomyces cerevisiae,
p08524; Kluyveromyces lactis, CAA 53614; Schizosaccharomy-
ces pombe, NP593299; Gibberella fujikuroi, CAA65641; Neuro-
spora crassa, CAD21355; Drosophila melanogaster, CAA08919;
Gallus gallus, P08836; Mus musculus, AAL09445; Homo sapiens,
NP001995; Bos Taurus, AAL58886; Caenorhabditis elegans,
CAB03221; Micrococcus luteus, BAA 25265; Yersinia pestis,
NP406651; Salmonella enterica, NP806170; Escherichia coli,
BAA00599; Shigella flexneri, NP706309.
2). The rbcL and rRNA data have given similar results
(Nickrent and Soltis, 1995). As the results were based on
relatively little data and the bootstrap values were high,
there was no prior criterion to decide which datum was
the best, and additional data are needed to confirm our
results. Due to the fact that FPSs are present in a wide
variety of species and have the same evolutionary pattern
as other proteins, they may be ideal molecular markers for
evolutionary studies.
Homology modeling of GbFPS By combining the se-
quence information about GbFPS with knowledge of the
three-dimensional structure of avian FPS, we obtained a
homology model of GbFPS. This exhibited a fold com-
posed of α-helices joined by connecting loops and a spot
of sheets (Fig. 3). The fold consisted
of a large central
154 FPS from Ginkgo biloba
Fig. 3. The 3-D structure of GbFPS established by homology-
based modeling. The helix, sheet and random coil are indicated
by the column-shape, arrow plate-shape and rope-shape, respec-
tively. The two aspartate-rich substrate-binding motifs, DDIMD
and DDYLD, are indicated by balls.
cavity formed by a bundle of 13 α-helices, and was the
putative catalytic site. This finding is consistent with re-
sults obtained with other FPSs (Tarshis et al. 1994). The
two aspartate-rich DDXX(XX)D sequences that are
highly conserved
in isoprenyl diphosphate synthases were
located on opposite walls of this cavity, facing each other.
All five of the identified conserved regions are clustered
around this cavity.
Southern blot analysis To investigate the genomic or-
ganization of GbFPS, we carried out a Southern blot
analysis by digesting genomic DNA of G. biloba with
HindIII and NcoI, followed by hybridization with a 461
bp probe generated by PCR using the 1173 bp coding
sequence of GbFPS cDNA as template. As shown in Fig.
4A, more than two hybridized bands were detected in
each lane, indicating that GbFPS belongs to a small gene
family in G. biloba. This is consistent with previous re-
ports in other plant species (Cunillera et al., 1996; Hem-
merlin et al., 2003; Li and Larkins, 1996).
Prediction of the subcellular localization of GbFPS IPP
is synthesized via two pathways: the mevalonate and the
1-deoxyxylulose-5-phosphate (DXP) pathways (Kuzuyama,
2002). The mevalonate pathway is found primarily in eu-
karyotes and archaea, whereas the non-mevalonate path-
way is found primarily in prokaryotes and in the plastids
of plants. Prenyltransferases, which utilize IPP, are also
thought to be distributed in three subcellular compart-
ments: the cytosol, mitochondria and plastids (Gray, 1987;
A B
Fig. 4. Southern blot and expression analyses of GbFPS. A.
Southern blot analysis. Genomic DNA of G. biloba was di-
gested with HindIII and NcoI, separated on a 0.85% agarose gel,
blotted onto a positively charged nylon membrane and probed
with a biotin-labeled GbFPS fragment. B. Expression analysis
of GbFPS in various tissues by semi-quantitative RT-PCR. To-
tal RNA was extracted from roots, stems and leaves, and sub-
jected to semi-quantitative RT-PCR analysis. The 18S rRNA
was used for normalization of RNA loading.
Fig. 5. Functional complementation of the leaky mutant yeast
strain CC25 with plasmid pYES2 + GbFPS. Strain CC25 and
strain CC25 transformed with pYES2 + GbFPS were streaked
onto YPD plates with/without 80 mg/L ergosterol and incubated
at 42°C for 16 h followed by 37°C for 2 d.
Kleinig, 1989). However, the fact that there is crosstalk
between the various subcellular compartments involved in
isoprenoid biosynthesis (Laule et al., 2003) makes it dif-
ficult to establish the subcellular localization of prenyl-
transferases. For example, FPS in Arabidopsis is targeted
to both the cytosol and the mitochondria
(Cunillera et al.,
1997), and there are reports that FPSs are localized in the
plastid compartment (Hemmerlin et al., 2003; Sanmiya et
al., 1999). These observations may be due to the fact that
FPSs belong to a small gene family and that particular
forms are present in particular subcellular compartments
Peng Wang et al. 155
where they perform specific functions. In the present case,
prediction of the subcellular localization of GbFPS indi-
cated that it was most likely to be cytoplasmic. Taken
together with the results of the Southern blot analysis, we
may deduce that a number of genes encode FPSs in G.
biloba and that the different products occupy different
subcellular compartments, the product of the full-length
cDNA we cloned being cytoplasmic.
Semi-quantitative RT-PCR Semi-quantitative RT-PCR
was carried out to investigate the expression pattern of
GbFPS in different tissues. The result showed that there is
high GbFPS expression in roots and leaves, and low
expression in stems (Fig. 4B), reflecting the fact that the
biosynthesis of ginkgolides and bilobalide occurs in roots
and leaves (Carrier et al., 1998).
Confirmation of GbFPS activity by complementation
of the mutant yeast strain, CC25 We used the sterol-
auxotrophic yeast strain CC25 bearing a disrupted copy of
the FPS gene to confirm that the cDNA of GbFPS en-
coded the anticipated functional enzyme. CC25 (MATa/
MATalpha, deltaERG20/+) is thermo-sensitive: it bears a
leaky mutation, erg20, affecting the ability of FPS to con-
dense GPP with IPP to yield FPP and, as a result, gener-
ates an amount of geraniol that is toxic (Blanchard and
Karst, 1993), and affects viability above 36°C. This phe-
notype becomes more pronounced as the temperature is
increased and above 42°C the cells are almost unviable.
The addition of ergosterol complements the phenotype
and restores viability at non-permissive temperature. As
shown in Fig. 5, both the mutant strain and the pYES2 +
GbFPS transformants were viable when the plates were
supplemented with ergosterol (80 mg/ml) at non-permissive
temperature (42°C for 16 h and then 37°C for 2 d). In
contrast, in the absence of ergosterol, the cells were al-
most inviable whereas the pYES2 + GbFPPS transfor-
mants grew well. This result confirms that the cloned
cDNA encodes a functional GbFPS.
Acknowledgments We are grateful to Wei Cao for technical
assistance, to Juan Lin for critical reading of the manuscript,
and to Xin Li, Xiaojun Liu, and Zhonghai Chen for valuable
advice on the manuscript. This research was supported by the
China National “863” High-Tech Program, China Ministry of
Education, Shanghai Science and Technology Committee and
Shanghai Agriculture Committee.
References
Aoki, T. (1997) Fungal association with Ginkgo biiloba; in
Ginkgo Biloba-A Global Treasure: From Biology to Medi-
cine, Hori, T., Ridge, R. W., Tulecke, W., and Tredici, P. D.
(eds.), pp. 251258, Springer-Verlag, Tokyo.
Biswas, C. and Johri, B. M. (1997) The Gymnosperm, pp. 98-
126, Narosa Publishing House, New Delhi, India.
Blanchard, L. and Karst, F. (1993) Characterization of a lysine-
to-glutamic acid mutation in a conservative sequence of far-
nesyl diphosphate synthase from Saccharomyces cerevisiae.
Gene 125, 185189.
Carrier, D. J., van Beek, T. A., van der Heijden, R., and Ver-
poort, R. (1998) Distribution of ginkgolides and terpenoid
biosynthetic activity in G. biloba. Phytochemistry 48, 8992.
Chen, A., Kroon, P. A., and Poulter, C. D. (1994) Isoprenyl
diphosphate synthases: protein sequence comparisons, a phy-
logenetic tree, and predictions of secondary structure. Pro-
tein Sci. 3, 600607.
Croteau, R., Kutchan, T. M., and Lewis, N. G. (2000) Natural
products (secondary metabolites); in Biochemistry & Molecu-
lar Biology of Plants, Buchanan, B. B., Gruissen, W., and
Jones, R. L. (eds.), pp. 1251-1267, American Society of Plant
Physiologists, Rockeville.
Cunillera, N., Boronat, A., and Ferrer, A. (1997) The Arabidop-
sis thaliana FPS1 gene generates a novel mRNA that encodes
a mitochondrial farnesyl-diphosphate synthase isoform. J.
Biol. Chem. 272, 1538115388.
Cunillera, N., Arro, M., Fores, O., Manzano, D., and Ferrer, A.
(2000) Characterization of dehydrodolichyl diphosphate syn-
thase of Arabidopsis thaliana, a key enzyme in dolichol bio-
synthesis. FEBS Lett. 477, 170174.
Gray, J. C. (1987) Control of isoprenoid biosynthesis in higher
plants. Adv. Bot. Res. 14, 2591.
Guinot, P. and Braquet, P. (1994) Effects of the PAF antagonists,
ginkgolides (bn-52063, bn-52021), in various clinical indica-
tions. J. Lipid. Mediat. Cell. 10, 141146.
Hasebe, M. (1997) Molecular phylogeny of G. biloba: close
relationship between G. biloba and cycads; in G. biloba-A
Global Treasure: From Biology to Medicine. Hori, T., Ridge,
R. W., Tulecke, W., and Tredici, P. D. (eds.), pp. 73182,
Springer-Verlag, New York.
Hemmerlin, A., Rivera, S. B., Erickson, H. K., and Poulter, C. D.
(2003) Enzymes encoded by the farnesyl diphosphate syn-
thase gene family in the big sagebrush Artemisia tridentata
ssp spiciformis. J. Biol. Chem. 278, 3213232140.
Honda, H. (1997) Ginkgo and insects; in Ginkgo biloba-A
Global Treasure: From Biology to Medicine, Hori, T., Ridge,
R. W., Tulecke, W., and Tredici, P. D. (eds.), pp. 243250,
Springer-Verlag, Tokyo.
Hua, S. J. and Sun, Z. R. (2001) Support vector machine ap-
proach for protein subcellular localization prediction. Bioin-
formatics 17, 721728.
Kellogg, B. A. and Poulter, C. D. (1997) Chain elongation in the
isoprenoid biosynthetic pathway. Curr. Opin. Chem. Biol. 1,
570578.
Kim, Y. S., Lee, J. K., and Chung, G. C. (1997) Tolerance and
susceptibility of Ginkgo to air pollution; in Ginkgo biloba-A
Global Treasure: From Biology to Medicine, Hori, T., Ridge,
R. W., Tulecke, W., and Tredici, P. D. (eds.), pp. 233242,
Springer-Verlag, Tokyo.
Kleinig, H. (1989) The role of plastids in isoprenoid biosynthe-
sis. Annu. Rev. Plant Physiol. Plant Mol. Biol. 40, 3959.
Kumar, S., Tamura, K., Jakobsen, I. B., and Nei, M. (2001)
MEGA2: molecular evolutionary genetics analysis software.
156 FPS from Ginkgo biloba
Bioinformatics 17, 12441245.
Kuzuyama, T. (2002) Mevalonate and nonmevalonate pathways
for the biosynthesis of isoprene units. Biosci. Biotechnol.
Biochem. 66, 16191627.
Laule, O., Furholz, A., Chang, H. S., Zhu, T., Wang, X., Heifetz,
P. B., Gruissem, W., and Lange, B. M. (2003) Crosstalk be-
tween cytosolic and plastidial pathways of isoprenoid bio-
synthesis in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA
100, 68666871.
Li, C. P. and Larkins, B. A. (1996) Identification of a maize
endosperm-specific cDNA encoding farnesyl pyrophosphate
synthetase. Gene 171, 193196.
Liao, Z., Tan, Q., Chai, Y., Zuo, K., Chen, M., Gong, Y., Wang,
P., Pi, Y., Tan, F., Sun, X., and Tang, K. (2004) Cloning and
characterisation of the gene encoding HMG-CoA reductase
from Taxus Media and its functional identification in yeast.
Funct. Plant Biol. 31, 7381.
Nickrent, D. L. and Soltis, D. E. (1995) A comparison of angio-
sperm phylogenies from nuclear 18SrDNA and rbcL se-
quences. Ann. Mol. Bot. Gard. 82, 208234.
Park, Y. G., Kim, S. J., Jung, H. Y., Kang, Y. M., Kang, S. M.,
Prasad, D. T., Kim, S. W., and Choi, M. S. (2004) Variation
of ginkgolides and bilobalide contents in leaves and cell cul-
tures of Ginkgo biloba L. Biotechnol. Bioproc. E. 9, 3540.
Peitsch, M. C. and Jongeneel, V. (1993) A 3-dimensional model
for the CD40 ligand predicts that it is a compact trimer simi-
lar to the tumor necrosis factors. Int. Immunol. 5, 233238.
Sanmiya, K., Ueno, O., Matsuoka, M., and Yamamoto, N. (1999)
Localization of farnesyl diphosphate synthase in chloroplasts.
Plant Cell. Physiol. 40, 348354.
Schepmann, H. G., Pang, J. H., and Matsuda, S. P. T. (2001)
Cloning and characterization of G. biloba levopimaradiene
synthase, which catalyzes the first committed step in gink-
golide biosynthesis. Arch. Biochem. Biophys. 392, 263269.
Schwede, T., Kopp, J., Guex, N., and Peitsch, M. C. (2003)
SWISS-MODEL: an automated protein homology-modeling
server. Nucleic Acids Res. 31, 33813385.
Tarshis, L. C., Yan, M. J., Poulter, C. D., and Sacchettini, J. C.
(1994) Crystal-structure of recombinant farnesyl diphosphate
synthase at 2.6-angstrom resolution. Biochemistry 33, 10871
10877.
Thompson, C. B. (1995) Apoptosis in the pathogenesis and
treatment of disease. Science 267, 14561462.
Thompson, J. D., Gibson, T. J., Plewniak, F., Jeanmougin, F.,
and Higgins, D. G. (1997) The ClustalX windows interface:
flexible strategies for multiple sequence alignment aided by
quality analysis tools. Nucleic Acids Res. 24, 48764882.
Winzeler, E. A., Shoemaker, D. D., Astromoff, A., Liang, H.,
Anderson, K., Andre, B., Bangham, R., Benito, R., Boeke, J.
D., Bussey, H., Chu, A. M., Connelly, C., Davis, K., Dietrich,
F., Dow, S. W., EL Bakkoury, M., Foury, F., Friend, S. H.,
Gentalen, E., Giaever, G., Hegemann, J. H., Jones, T., Laub,
M., Liao, H., Liebundguth, N., Lockhart, D. J., Lucau-Danila,
A., Lussier, M., M’Rabet, N., Menard, P., Mittmann, M., Pai,
C., Rebischung, C., Revuelta, J. L., Riles, L., Roberts, C. J.,
Ross-MacDonald, P., Scherens, B., Snyder, M., Sookhai-
Mahadeo, S., Storms, R. K., Veronneau, S., Voet, M., Vol-
ckaert, G., Ward, T. R., Wysocki, R., Yen, G. S., Yu, K. X.,
Zimmermann, K., Philippsen, P., Johnston, M., and Davis, R.
W. (1999) Functional characterization of the S. cerevisiae
genome by gene deletion and parallel analysis. Science 285,
901906.
... FPSs have been identified in a few plants 11,22,23,25 . FPSs belong to a small multigenic family which encodes at least two different isoforms in plants. ...
... For example, in Arabidopsis FPS1 is predominantly expressed in roots and inflorescences, whereas FPS2 accumulates preferentially mRNA in inflorescences 22 . In Ginkgo biloba, the higher GbFPS expression level was detected in roots and leaves 23 , in which the ginkgolides and bilobalide are synthesized 24 . In Euphorbia pekinensis, EpFPS had a high transcription level in roots, in which terpenoids are synthesized 25 . ...
... Plant hormones have crucial important roles in regulating natural rubber biosynthesis 39,43,44 . WRKY transcription factors are involved in SA, AB, ET, and JA signaling pathways and plays a vital role in the signal crosstalk of the SA, AB, ET, and JA signaling pathways 23,45,46 . The promoter of HbWRKY27 had a few cis-acting elements related to hormone responses and HbWRKY27 is simultaneously up-regulated by SA, AB, ET, and JA, suggesting that HbWRKY27 might integrate plant hormones signals and regulates natural rubber biosynthesis. ...
Article
Full-text available
Farnesyl pyrophosphate synthase (FPS) is a key enzyme that catalyzes the formation of farnesyl pyrophosphate, the main initiator for rubber chain initiation in Hevea brasiliensis Muell. Arg. The transcriptional regulatory mechanisms of the FPS gene still not well understood. Here, a WRKY transcription factor designated HbWRKY27 was obtained by screening the latex cDNA library applied the HbFPS1 promoter as bait. HbWRKY27 interacted with the HbFPS1 promoter was further identified by individual Y1H and EMSA assays. HbWRKY27 belongs to group IIe WRKY subfamily which contains a typical WRKY domain and C-X5-CX23-HXH motif. HbWRKY27 was localized to the nucleus. HbWRKY27 predominantly accumulated in latex. HbWRKY27 was up-regulated in latex by ethrel, salicylic acid, abscisic acid, and methyl jasmonate treatment. Transient expression of HbWRKY27 led to increasing the activity of the HbFPS1 promoter in tobacco plant, suggesting that HbWRKY27 positively regulates the HbFPS1 expression. Taken together, an upstream transcription factor of the key natural rubber biosynthesis gene HbFPS1 was identified and this study will provide novel transcriptional regulatory mechanisms of the FPS gene in Hevea brasiliensis.
... Gene sequencing and sequence analysis revealed that PcFPPS contained an integrated open reading frame (ORF) with 1,050 bp, encoded 349 amino acid residues, and its estimated molecular weight was 40.11 kDa. Consistent with FPPSs from different plants (Wang et al., 2004), the deduced PcFPPS protein contains five highly conserved domains, two aspartate-rich motifs (DDxxD) FARM (First Aspartic Rich Motif) and SARM (Second Aspartic Rich Motif) located in conserved domains II and V, respectively. These motifs are considered the binding sites of IDP and DMAPP ( Figure 1A). ...
Article
Full-text available
Farnesyl pyrophosphate synthase (FPPS) plays an important role in the synthesis of plant secondary metabolites, but its function and molecular regulation mechanism remain unclear in Pogostemon cablin. In this study, the full-length cDNA of the FPP synthase gene from P. cablin (PcFPPS) was cloned and characterized. The expressions of PcFPPS are different among different tissues (highly in P. cablin flowers). Subcellular localization analysis in protoplasts indicated that PcFPPS was located in the cytoplasm. PcFPPS functionally complemented the lethal FPPS deletion mutation in yeast CC25. Transient overexpression of PcFPPS in P. cablin leaves accelerated terpene biosynthesis, with an ~47% increase in patchouli alcohol. Heterologous overexpression of PcFPPS in tobacco plants was achieved, and it was found that the FPP enzyme activity was significantly up-regulated in transgenic tobacco by ELISA analysis. In addition, more terpenoid metabolites, including stigmasterol, phytol, and neophytadiene were detected compared with control by GC-MS analysis. Furthermore, with dual-LUC assay and yeast one-hybrid screening, we found 220 bp promoter of PcFPPS can be bound by the nuclear-localized transcription factor PcWRKY44. Overexpression of PcWRKY44 in P. cablin upregulated the expression levels of PcFPPS and patchoulol synthase gene (PcPTS), and then promote the biosynthesis of patchouli alcohol. Taken together, these results strongly suggest the PcFPPS and its binding transcription factor PcWRKY44 play an essential role in regulating the biosynthesis of patchouli alcohol.
... fdps serves as a substrate for the first committed reactions of several branched pathways leading to the synthesis of compounds required for growth and development (e.g., phytosterols and sesquiterpenoid phytoalexins) (Cunillera et al., 1996). In this study, the down-regulation of fdps gene may block the synthesis pathway of sterol and cause sterol depletion (Wang et al., 2004). Involved in the biosynthesis of isoprene diphosphate and dimethylallyl diphosphate, ispE catalyzes the monophosphate dependent phase of 4-diphosphate-2-methyl-D-erythritol-2phosphate. ...
Article
The occurrence of hormesis in the algal growth inhibition test is a major challenge in the dose-response characterization, whereas the molecular mechanism remains unraveled. The aim of this study is therefore to investigate the changes in the molecular pathways in a model green alga Raphidocelis subcapitata treated with erythromycin (ERY; 4, 80, 120 μg L⁻¹) by transcriptomic analysis. After 7 day exposure, ERY at 4 μg L⁻¹ caused hormetic effects (21.9%) on cell density, whereas 52.0% and 65.4% were inhibited in two higher exposures. By using adj p < 0.05 and absolute log2 fold change > 1 as a cutoff, we identified 218, 950, and 2,896 differentially expressed genes in 4, 80, 120 μg L⁻¹ treatment groups, respectively. In two higher ERY treated groups, genes involved in phases I, II & III metabolism processes and porphyrin and chlorophyll metabolism pathway were consistently suppressed. Interestingly, genes (e.g., prim2, mcm2, and mcm6) enriched in DNA replication process were up-regulated in 4 μg L⁻¹ group, whereas these genes were all repressed in 120 μg L⁻¹ group. Alteration trend in gene expression was consistent with algal growth. Taken together, our results unveiled the molecular mechanism of action in ERY- stimulated/ inhibited growth in green alga.
... In different plants, the expression of the FPS genes showed organic specificity, and the expression level was correlated to their accumulation of isoprenoidderivants (Gupta et al., 2011;Lan et al., 2013;Wang et al., 2004;Xiang et al., 2010;Yin et al., 2011). In the present study, this kind of organic specificity was not very obvious ( Fig. 9), but consistent with the general biosynthesis of terpenoids in all organs of A. roxburghii and A. formosanus (Chappell, 1995;Falcone Ferreyra et al., 2012;Kim et al., 2010;Szkopinska & Płochocka, 2005). ...
... The biosynthesis of bilobalide was highly correlated with the mevalonic acid pathway, and the biosynthesis of ginkgolides was highly correlated with the methylerythritol 4-phosphate pathway (Chen et al., 2015). Most structural genes of the terpene trilactone biosynthetic pathway in G. biloba, including DXS (Gong et al., 2006;Kim et al., 2006a), DXR (Gong et al., 2005), MECT (Kim et al., 2006b), CMK (Kim et al., 2008a), MECS (Kim et al., 2006c), HDS (Kim and Kim, 2009), HDR (Kim et al., 2008b;Lu et al., 2008), FPS (Wang et al., 2004), GGPPS , LPS (Schepmann et al., 2001), HMGR , MVD (Pang et al., 2006), AACT and MVK , have been isolated. WRKY binding sites are found in the 1101 ...
Article
Full-text available
Ginkgo biloba is widely planted, and the extracts of leaves contain flavonoids, terpene esters and other medicinal active ingredients. WRKY proteins are a large transcription factor family in plants, which play an important role in the regulation of plant secondary metabolism and development, as well as the response to biotic and abiotic stress. In our study, we identified 40 genes with conserved WRKY motifs in the G. biloba genome and classified into groups I (groups I-N and -C), II (groups IIa, b, c, d, and e), and III, which include 12, 26, and 2 GbWRKY genes, respectively. Meanwhile, the expression patterns of 10 GbWRKY (GbWRKY2, GbWRKY3, GbWRKY5, GbWRKY7, GbWRKY11, GbWRKY15, GbWRKY23, GbWRKY29, GbWRKY31, GbWRKY32) under different tissue and abiotic stress conditions were analyzed. Under stress treatment, the expression patterns of 10 WRKY genes were changed. 10 ginkgo WRKY transcription factors were induced by ETH and SA, but there are two different induced response modes. The expression of 10 WRKY genes was inhibited under low temperature, high temperature and MeJA hormone induction. Most WRKY genes were up-regulated under the induction of high salt and ABA. GbWRKYs were differentially expressed in various tissues after abiotic stress and plant hormone treatments, thereby indicating their possible roles in biological processes and abiotic stress tolerance and adaptation. Our results provided insight into the genome-wide identification of GbWRKYs, as well as their differential responses to stresses and hormones. These data can also be utilized to identify potential molecular targets to confer tolerance to various stresses in G. biloba. ********* In press - Online First. Article has been peer reviewed, accepted for publication and published online without pagination. It will receive pagination when the issue will be ready for publishing as a complete number (Volume 47, Issue 4, 2019). The article is searchable and citable by Digital Object Identifier (DOI). DOI link will become active after the article will be included in the complete issue. *********
... Several studies showed that the expression pattern of FPPS in plant tissues significantly varies across different plants. For example, FPPS is strongly expressed in leaf and root, moderately expressed in stem, and weakly expressed in the stem of G. biloba (Wang et al., 2004). FPPS genes are commonly found in tomato plants and are regulated during development (Gaffe et al., 2000). ...
Article
Full-text available
Farnesyl diphosphate synthase (FPPS), an isopentenyl transferase, catalyzes the condensation reaction of five carbon isopentenyl pyrophosphate (IPP) and dimethylallyl pyrophosphate (DMAPP) to form fifteen carbon farnesyl pyrophosphate (FPP), which is the key precursor for sesquiterpene biosynthesis. In this study, a FPPS gene (CnFPPS) was cloned from Chamaemelum nobile. The full-length cDNA of CnFPPS is 1239 bp and contains an open reading frame (ORF) of 1029 bp encoding 342 amino acids. The theoretical molecular weight and pI of the CnFPPS protein are 39.38 kDa and 5.59, respectively. Multiple alignment analysis showed the protein sequence of CnFPPS had a high homology with FPPS proteins from other plants. The deduced amino acid of CnFPPS contained five conservative domains such as substrate binding pocket, substrate-Mg2+ binding site, catalytic site, aspartate-rich region 1 and 2, suggesting CnFPPS is one member of FPPS family in C. nobile. Phylogenetic analysis based on the amino acid sequences of FPPSs showed that CnFPPS was closely related to the FPPS of Matricaria chamomilla. The result of qRT-PCR revealed that CnFPPS gene was constitutively expressed in different tissues of C. nobile, with the highest expression in the root. These findings improve the understanding of the synthesis and regulation of the terpenoid compounds at the molecular level and lay a foundation for studying the regulatory functions of CnFPPS in terpenoid biosynthetic pathway in C. nobile.
... The parameters added and modified in the nexus file for tree reconstruction were as follows: Statefreqpr = dirichlet (0.2722, 0.2343, 0.2413, and 0.2522), revmatpr = dirichlet (1.4781, 3.1597, 1.1667, 1.1255, 4.6277, and 1.0000), shapepr = fixed (2.2202), pinvarpr = fixed (0.0054), unlinkshape = (3), mcmcp ngen = 10,000,000, and samplefreq = 10,000; mcmc. There were 10 million generations with sampling every 10 thousand generations [56,57]. After completing the MrBayes analysis, the first 250,000 generations were discarded from every run. ...
Article
Full-text available
Background Farnesyl pyrophosphate synthase (FPS) belongs to the short-chain prenyltransferase family, and it performs a conserved and essential role in the terpenoid biosynthesis pathway. However, its classification, evolutionary history, and the forces driving the evolution of FPS genes in plants remain poorly understood. ResultsPhylogeny and positive selection analysis was used to identify the evolutionary forces that led to the functional divergence of FPS in plants, and recombinant detection was undertaken using the Genetic Algorithm for Recombination Detection (GARD) method. The dataset included 68 FPS variation pattern sequences (2 gymnosperms, 10 monocotyledons, 54 dicotyledons, and 2 outgroups). This study revealed that the FPS gene was under positive selection in plants. No recombinant within the FPS gene was found. Therefore, it was inferred that the positive selection of FPS had not been influenced by a recombinant episode. The positively selected sites were mainly located in the catalytic center and functional areas, which indicated that the 98S and 234D were important positively selected sites for plant FPS in the terpenoid biosynthesis pathway. They were located in the FPS conserved domain of the catalytic site. We inferred that the diversification of FPS genes was associated with functional divergence and could be driven by positive selection. Conclusions It was clear that protein sequence evolution via positive selection was able to drive adaptive diversification in plant FPS proteins. This study provides information on the classification and positive selection of plant FPS genes, and the results could be useful for further research on the regulation of triterpenoid biosynthesis.
Article
Rosa rugosa , a renowned ornamental plant, is cultivated for its essential oil containing valuable monoterpenes, sesquiterpenes, and other compounds widely used in the floriculture industry. Farnesyl diphosphate synthase (FPPS) is a key enzyme involved in the biosynthesis of sesquiterpenes and triterpenes for abiotic or biotic stress. In this study, we successfully cloned and characterized a full-length FPPS- encoding cDNA identified as RrFPPS1 using RT-PCR from R. rugosa . Phylogenetic analysis showed that RrFPPS1 belonged to the angiosperm-FPPS clade. Transcriptomic and RT-qPCR analyses revealed that the RrFPPS1 gene had tissue-specific expression patterns. Subcellular localization analysis using Nicotiana benthamiana leaves showed that RrFPPS1 was a cytoplasmic protein. In vitro enzymatic assays combined with GC-MS analysis showed that RrFPPS1 produced farnesyl diphosphate (FPP) using isopentenyl diphosphate (IPP) and dimethylallyl diphosphate (DMAPP) as substrates to provide a precursor for sesquiterpene and triterpene biosynthesis in the plant. Additionally, our research found that RrFPPS1 was upregulated under salt treatment. These substantial findings contribute to an improved understanding of terpene biosynthesis in R. rugosa and open new opportunities for advancements in horticultural practices and fragrance industries by overexpression of the RrFPPS1 gene in vivo increased FPP production and subsequently led to elevated sesquiterpene yields in the future. The knowledge gained from this study can potentially lead to the development of enhanced varieties of R. rugosa with improved aroma, medicinal properties, and resilience to environmental stressors.
Chapter
Terpenoids represent the largest class of secondary metabolites and usually do not contain nitrogen or sulfur in their structures. Many terpenoids serve as defence compounds against microbes and herbivores and/or are signal molecules to attract pollinating insects, fruit‐dispersing animals or predators which can destroy insect herbivores. As a consequence, many terpenoids have pronounced pharmacological activities and are therefore interesting for medicine and biotechnology. The first part of the biosynthesis is the generation of a C5 unit, such as isopentenyl diphosphate (IPP) or dimethylallyl diphosphate (DMAPP). Two independent pathways have been discovered that can produce the C5 unit: the mevalonate and the methylerythritol phosphate (MEP) pathway. Depending on the number of C5 units, we distinguish hemiterpenes C5, monoterpenes including iridoids (C10), sesquiterpenes (C15), diterpenes (C20), sesterterpenes (C25), triterpenes (including steroids) (C30), tetraterpenes (C40) and polyterpenes (>C40). The biosynthesis (including enzymes, genes and their regulation) of mevalonate and the methylerythritol phosphate pathway and the consecutive pathways leading to mono‐, sesqui‐ and diterpenes are discussed in this chapter in detail.
Article
Eucommia ulmoides Oliver is rich in trans-polyisoprene rubber (Eu-Rubber), a high-molecular mass polymer of isoprene units with a trans-configuration. Farnesyl diphosphate (FPP) synthase (FPS) is a key enzyme, which involved in the production of important precursors of different terpenoids. In this study, we cloned and characterized five novel FPS genes from E. ulmoides. The full-length synthases were named EuFPS1-5 and their deduced amino acid sequences exhibited high homology to those from other plant isoforms. EuFPS1 and EuFPS4 were observed to be highly expressed in leaves, EuFPS2 and EuFPS3 were present at low levels in leaves and fruit throughout the plant development, and EuFPS5 was highly expressed exclusively in young fruit. Expression of EuFPS5 correlated with the accumulation rate of Eu-Rubber and might be responsible for it. This study is expected to enhance our understanding of the role of EuFPSs in biosynthesis and regulation of useful secondary metabolites in E. ulmoides.
Chapter
Full-text available
The ginkgo, or maidenhair tree, Ginkgo biloba L., is the only living species of the family Ginkgoaceae, which were gymnosperms that thrived about 175 to 200 Million years ago. Native to southern China, Ginkgo is now distributed widely. Ginkgo was introduced to Japan about 900 years ago and has been used locally for firewood and handcrafts, and the seed roasted as a delicacy. It has been exploited mainly as a landscape tree rather than as a source of timber. For centuries, because of its high desirability as an ornamental tree with handsome foliage and autumn colors, its resistance to insect pests and plant diseases and its adaptability to urban conditions, planting of Ginkgo has increased along road-sides, in public parks, and in private gardens. However, I will note in this chapter some insect pests that occur Ginkgo trees and the relationships between insects and Ginkgo.
Chapter
Full-text available
Trees face a variety of environmental stresses and must endure conditions unfavorable for their growth. The stress induced by numerous anthropogenic stress factors can, however, be calculated to a critical threshold value by morphological and physiological adaptations of woody plants. If the sum of the various stresses exceeds this critical value, trees start to develop symptoms in the plant organs [1].
Article
Full-text available
To investigate the phylogenetic utility of entire, nuclear-encoded small-subunit (18S) ribosomal DNA sequences, we compared the rate of evolution and phylogenetic resolution of entire 18S sequences with those for the chloroplast gene rbcL using a suite of 59 angiosperms and 3 gymnosperms (Gnetum, Ephedra, and Zamia) as outgroups. For rbcL, 482 (33.6%) of the 1431 base positions were phylogenetically informative, whereas for 18S rDNA 341 (18.4%) of the 1853 positions were informative. Pairwise comparisons within the angiosperms show that rbcL is generally about three times more variable than 18S rDNA. However, because the 18S region is approximately 400 base pairs longer than rbcL, the ratio of the number of phylogenetically informative sites per molecule is only about 1.4 times greater for rbcL compared to 18S rDNA. Not only are sites more variable in rbcL than in 18S rDNA, but this variability is more evenly distributed over the length of rbcL. In contrast, 18S rDNA shows highly variable regions interspersed with regions of extreme conservation. Minimum-length Fitch trees were constructed for each matrix, and the results were compared to a tree derived from a previous global analysis of rbcL sequences based on 499 seed plants. Parsimony analyses showed that several clades are strongly supported by both data sets, such as Gnetales, monocots, paleoherbs, Santalales, and various clades within Rosidae s.l. and Asteridae s.l. Some clades (e.g., Santalales) have higher base substitution rates for 18S rDNA, permitting the assessment of inter- and intrafamilial relationships. This comparative study indicates that 18S rDNA sequences contain sufficient information to conduct phylogenetic studies at higher taxonomic levels (family and above) within angiosperms. rDNA sequences are best applied to such deep divergences, but the amount of variation differs significantly among taxonomic groups.
Chapter
Higher plants contain a bewildering array of isoprenoid compounds with a wide variety of structures and functions. In higher plants, many of these isoprenoid compounds play vital roles in the metabolism and development of the plant. The plant growth regulators, abscisic acid and gibberellins, are isoprenoid compounds and many cytokinins contain an isoprenoid side chain. Isoprenoid side chains are also found in many other biologically active molecules; including chlorophylls, plastoquinone, and other prenylquinones in chloroplasts, where together with carotenoids they are involved in photosynthesis. Isoprenoid side chains are also found in the mitochondria1 electron transfer chain components, ubiquinone, and haem α of cytochrome oxidase. Plants must be able to produce this wide range of isoprenoid compounds in different amounts in different parts of the plant at different stages of growth and development. As all these compounds are produced by a common biosynthetic pathway, the plant must have exquisite control mechanisms to ensure the synthesis of the necessary compounds in the right place at the right time. This chapter considers the knowledge of the control of isoprenoid biosynthesis in higher plants, particularly with respect to the properties and subcellular locations of the enzymes involved.
Chapter
In the Gymnospermae, coniferous substrata have been well examined for associating fungal species, e.g., in phytopathological studies, in floral investigation including edible or toxic mushrooms, and in the research of fungal succession phenomena following decomposition of leaf litter or wood [1,2]. Fungi in relation to Ginkgo biloba have not been well studied, however. Because of its less common distribution in the wild and because of the limited industrial use of Ginkgo wood, interests of many mycologists until now have been mainly confined to fungi pathogenic to Ginkgo street trees.
Article
The functions of many open reading frames (ORFs) identified in genome-sequencing projects are unknown. New, whole-genome approaches are required to systematically determine their function. A total of 6925 Saccharomyces cerevisiae strains were constructed, by a high-throughput strategy, each with a precise deletion of one of 2026 ORFs (more than one-third of the ORFs in the genome). Of the deleted ORFs, 17 percent were essential for viability in rich medium. The phenotypes of more than 500 deletion strains were assayed in parallel. Of the deletion strains, 40 percent showed quantitative growth defects in either rich or minimal medium.
Article
The terpene trilactone content (bilobalide and ginkgolides) of extracts prepared from leaves of terminal buds, rosettes and side branches, from stem and root bark, and from root and root meristems of three-year-old Ginkgo biloba plants was determined. The aerial parts were relatively rich in bilobalide while ginkgolides were the major constituents of the underground parts. The formation of farnesyl and geranylgeranyl pyrophosphate was monitored by incubating cell-free extracts prepared from the corresponding plant parts with [1-14C]isopentenyl pyrophosphate. Extracts prepared from leaves of the terminal buds displayed terpenoid biosynthetic activity, suggesting that terpene trilactone synthesis might occur in actively growing tissues.