Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Molecular Cloning and Characterisation of Farnesyl Pyrophosphate Synthase from Tripterygium wilfordii

  • Yu-Jun Zhao,

    Affiliation Capital Medical University School of Traditional Chinese Medicine, Capital Medical University, Beijing, People’s Republic of China

  • Xin Chen ,

    chenxin4283@126.com (XC); huangluqi01@126.com (LH); weigao@ccmu.edu.cn (WG)

    Affiliation Capital Medical University School of Traditional Chinese Medicine, Capital Medical University, Beijing, People’s Republic of China

  • Meng Zhang,

    Affiliation Capital Medical University School of Traditional Chinese Medicine, Capital Medical University, Beijing, People’s Republic of China

  • Ping Su,

    Affiliation Capital Medical University School of Traditional Chinese Medicine, Capital Medical University, Beijing, People’s Republic of China

  • Yu-Jia Liu,

    Affiliation Capital Medical University School of Traditional Chinese Medicine, Capital Medical University, Beijing, People’s Republic of China

  • Yu-Ru Tong,

    Affiliation Capital Medical University School of Traditional Chinese Medicine, Capital Medical University, Beijing, People’s Republic of China

  • Xiu-Juan Wang,

    Affiliation Capital Medical University School of Traditional Chinese Medicine, Capital Medical University, Beijing, People’s Republic of China

  • Lu-Qi Huang ,

    chenxin4283@126.com (XC); huangluqi01@126.com (LH); weigao@ccmu.edu.cn (WG)

    Affiliation National Resource Center for Chinese Materia Medica, China Academy of Chinese Medical Sciences, Beijing, People’s Republic of China

  • Wei Gao

    chenxin4283@126.com (XC); huangluqi01@126.com (LH); weigao@ccmu.edu.cn (WG)

    Affiliation Capital Medical University School of Traditional Chinese Medicine, Capital Medical University, Beijing, People’s Republic of China

Abstract

Farnesylpyrophosphate synthase (FPS) catalyzes the biosynthesis of farnesyl pyrophosphate (FPP), which is an important precursor of sesquiterpenoids such as artemisinin and wilfordine. In the present study, we report the molecular cloning and characterization of two full-length cDNAs encoding FPSs from Tripterygium wilfordii (TwFPSs). TwFPSs maintained their capability to synthesise FPP in vitro when purified as recombinant proteins from E. coli. Consistent with the endogenous role of FPS in FPP biosynthesis, TwFPSs were highly expressed in T. wilfordii roots, and were up-regulated upon methyl jasmonate (MeJA) treatment. The global gene expression profiles suggested that the TwFPSs might play an important regulatory role interpenoid biosynthesis in T. wilfordii, laying the groundwork for the future study of the synthetic biology of natural terpene products.

Introduction

Tripterygium wilfordii is a traditional Chinese medicinal plant used to treat inflammatory diseases because of its analgesic and anti-microbial properties [1]. This plant has been widely used in the treatment of immune and tumour diseases [24]. Terpenoids are the primary active substances (Fig 1) of T.wilfordii and include sesquiterpene, diterpenoids andtriterpenoids. Triptolide [5], which has been recognised as one of the primary active constituents of T. wilfordii, is a class of 3 epoxygroups and an α,β-unsaturated five-membered lactone ringand is also a unique abietane diterpene. T. wilfordii has attracted much attention due to its architecture and significant activities. Tripterine [6], which was the first monomer isolated from T. wilfordii has important biological activity.

thumbnail
Fig 1. The main medicinal active substances of Tripterygium wilfordii.

https://doi.org/10.1371/journal.pone.0125415.g001

FPS belongs to the family of short-chain prenyltransferases [7]. FPS catalyzes the head-to-tail condensation reaction of dimethylallyl pyrophosphate (DMAPP) with two molecules of isopentenyl pyrophosphate (IPP) to form FPP [89], which is the precursor of all sesquiterpenes [10], such as artemisinin and wilfordine.FPS provides substrate FPP to squalene synthase and sesquiterpene synthase. Squalene synthase brings the synthesis into steroids and saponins, which involved in cell membrane systems building; sesquiterpene synthase put the synthesis on cyclic sesquiterpene compounds [11]. Upon treatment with elicitors, FPP, which is normally employed in sterol biosynthesis, is diverted from the sterol pathway as large amounts of antibiotic sesquiterpenesare produced [1214]. This diversion is accomplished by the induction of sesquiterpene cyclases and coeval suppression of squalene synthesis, the initial committed step of phytosterol production [1314].

Thus far, FPS has been isolated from sorghum [15], corn [16], sandalwood [17] and 41 other species of plants. Recently, FPS has been isolated from Panax ginseng [18], Salvia miltiorrhiza, Acanthopanax senticosus, Artemisia annua [19], Ginkgo biloba [20] and other medicinal plants. International and domestic academics have studied the gene expression patterns of farnesyl pyrophosphate synthase in plants and found that this gene has tissue-specific expression with increased isoprenoid derivative contents. Gene expression studies have demonstrated that gene expression is related to the isoprene content in plants.

The effective components in T. wilfordii are difficult to obtain by traditional chemical methods. In addition, the products of bioactive compounds cannot be synthesized through microbial methods. The current information regarding terpenoid biosynthesis genes is limited, and studies regarding key enzymes in T. wilfordii are few. Based on the above issues, we present the cloning and characterisation of full-length FPS cDNAs of T. wilfordii (TwFPSs) for the first time.

Materials and Methods

Plant materials

Fresh leaves and stems of T. wilfordii were sheared, cleaned and disinfected. Then, the leaves were cut into 1.0 cm×1.0 cm pieces, and the stems were cut into 1.0 cm lengths after rinsing with sterile water. These tissues were cultivated in MS agar medium containing 2,4-D hormone at 25°C in dark. After two weeks, calluses began to grow at the explant slits. The calluses which had a white lustre, were soft and grew well, were cultured in MS agar medium containing 0.5 mg/L of 2,4-D, 0.1 mg/L of KT, and 0.5 mg/L of IBA at 25°C in the dark.

The calluses were cultured in MS agar medium containing 0.2 mg/L of IAA, 0.5 mg/L of KT, and 1.5 mg/L of 6-BA at 25°C under 16 h light/8 h dark conditions. The aseptic seedlings were harvested after subculturing once a month.

After 3 subcultures, we chose the calluses that grew well and that had a loose texture and clipped these calluses into small pieces with tweezers. These calluses were cultured in MS medium containing 0.5 mg/L of 2,4-D, 0.1 mg/L of KT, 0.5 mg/L of IBA, and cell suspensions of 2.0 g/40 mL in the dark at 25°C with rotary shaking at 120 rpm.

Elicitor treatment

T. wilfordii cell suspensions displaying good growth were chosen as the experimental material. After 12 days, 1 mM [21] methyl jasmonate (MeJA) [2223] was added to the treatment groups, and DMSO was added to the control groups. After treatment, the cell suspensions were harvested in liquid nitrogen at 0, 4, 12, 24 and 48 hand stored at -80°C.

Cloning of TwFPS1 and TwFPS2 from T. wilfordii

Total RNA was extracted from T. wilfordii cell suspensions using the cetyltrimethylammonium bromide (CTAB) [24] method. Genome or protein contamination was eliminated using DNase I and an RNA purification kit. A PrimeScript 1st Strand cDNA Synthesis Kit was used to obtain the cDNA. A SMART RACE cDNA Amplification Kit was used to transcribe the first-strand cDNA for 3′- and 5′-RACE. The 3′- and 5′-RACE products were subcloned into the pMD19-T plasmid. The cloning vector was amplified in E. coli DH5α cells. After sequencing and alignment, the full-length predicted cDNAs for TwFPS1 and TwFPS2 could be obtained. 3′-RACE Primer: 5′-TGCCTTGCTCGGATGGGCTTCG-3′ (TwFPS1), 5′-GGATGATTACCTGGACTGTTTTGGGG-3′ (TwFPS2). 5′-RACE primer: 5′-TGGTTGACCCCGCCGAGTAACAGAT-3′ (TwFPS1), 5′-TGAACAATGCGGCGGTGGAGTGAC-3′ (TwFPS2). New primers were designed to confirm the open reading frame (ORF) of T. wilfordii. TwFPS1: 5′-ACATGGGGATCGGCAGCCATAC-3′ (forward) and 5′-TCAGAAGCTACGGCAGAATCTAATGGAG-3′ (reverse). TwFPS2: 5′-TCTCTGTGTCTCCGCAAA-3′ (forward) and 5′-GAGTAACCATAAGCAGCAGAC-3′ (reverse).

Sequence and phylogenetic analyses

The nucleotide and protein sequences were compared using NCBI (http://www.ncbi.NLM.NIH.gov). The sequences were translated into amino acidsusing DNAMAN software. The ORF was searched using ORF Finder (www.ncbi.NLM.NIH.gov/Gorf/Gorf.html). The theoretical isoelectric point (pI) and molecular weight (Mw) were determined using the Compute pI/Mw tool (http://Web.ExPASy.org/compute_pi/). Multiple sequence alignments were performed using DNAMAN and ClustalW software. Phylogenetic analysis was performed using MEGA5.1 software to build evolutionary trees.

Expression of TwFPS1 andTwFPS2 in E. coli and purification of recombinant protein

Based on the cloning vector, the primers for recombination were designed; an NcoI site and a HindIII site were introduced at the start and stop codons, respectively. TwFPS1: 5′-GAGGAGCCATGGCTATGAGCGACACCAAGTCCAAG-3′ (forward) and 5′-GAGGAGAAGCTTCTACTTCTCTCGCTTGTATAT-3′ (reverse). The primers for recombination were designed; a KpnI site and an EcoRI site were introduced at the start and stop codons, respectively. TwFPS2: 5′-GAGGAGGGTACCATGGCGGATCTCAAGTCAACG-3′ (forward) and 5′-GAGGAGGAATTCCTACTTCTGTCTCTTGTATATC-3′ (reverse). The fragments were cloned into the expression vector pET-32a(+). E. coli BL21 (DE3) cells were used for recombinant plasmid expression and the induction was performed at 16°C for 20 h with the addition of 1 mM isopropyl thiogalactoside (IPTG). After collection, the cells were resuspended in 2 mL lysis buffer A (20 mM Tris—HCl(pH 8.0), 500 mM NaCl, 1 mM PMSF and bacteria protease inhibitor cocktail), lysed by sonication (lysed for 5s, paused for 5s) and centrifuged for 30 min at 13000 g. Thus, the supernatant protein was harvested.

This supernatant was loaded onto a Ni-NTA column and pre-equilibrated by lysis buffer A. The column was washed with 10 column volumes of lysis buffer A containing 20 mM imidazole. The target proteins were eluted by buffer B (20 mM Tris—HCl(pH 8.0) containing 500 mM NaCl and 50 mM imidazole) and buffer C (20 mM Tris—HCl(pH 8.0) containing 150 mM NaCl and 250 mM imidazole). The purified protein was examined by SDS-PAGE gel electrophoresis. The concentration of FPPS protein was determined by the Bradford method [25].

Assay for enzymatic activity

Scale assays (200μL) for the identification of enzymatic reaction were performed with purified FPPS proteins (200μg) using 100μM IPP and 100μM DMAPP in assay buffer (50 mM Tris—HCl, pH 7.6) containing 5 mM MgCl2, 25 mM DTT, and 10% [v/v] glycerol. The assays were incubated at 30°C for 2 hours. To stop the assay and hydrolyse all diphosphate esters, 200μL of solution containing 2 units of potato apyrase and 2 units of calfintestine alkaline phosphatasein 0.2 M Tris—HCl(pH 9.5)was added to all assays, followed by incubation for 8 h at 30°C. After enzymatic hydrolysis, the resulting isopropyl alcohols were extracted into 500μL hexane 3 times. The hexane phase was concentrated by passing N2 at the opening of the tube, and then the products were dissolved in 100μL hexane and used for GC—MS measurements.

GC—MS analysis was detected on an Agilent 6890N gas chromatograph (splitless; injector temperature, 250°C) with a 5975imass spectrometer (GC—MS). One microliter of dissolved organic phase was injected. The separation was performed on an HP-5MS column (50 m×0.25 mm×0.5μm) with helium as the carrier gas (flow rate of 1 mL/min) on a temperature gradient from 60°C, at 10°C per min to 90°C (hold 1 min),and then 3°C per min to 220°C (hold 1 min). Mass spectra, 70eV (in EI mode), ion trap heating, 230°C; scan range, 30–500 amu. The products were identified viamass spectrometry profiles, and farnesol could be identified using standard chemicals and retention times. The assay was performed with empty vector as a control.

Real-time quantitative PCR analysis of TwFPS1 andTwFPS2 expression

Total RNA from different tissues (roots, stems, and leaves) and different inductive stages was extracted separately as described above. The primers for real-time quantitative PCR analysis were designed using Primer Premier 5.0 software. TwFPS1: 5′-GGGTGTATTTGCGGAGT-3′ (forward) and 5′-CGGCAGAATCTAATGGAG-3′ (reverse); TwFPS2: 5′-CAGACCCTCACCTTCCATT-3′ (forward) and 5′-AAGAGTAACCATAAGCAGCAGAC-3′ (reverse). The βtactin gene was used as an endogenous control to normalize expression. The PCR reaction conditions were as follows: an initial incubation at 95°C for 3 min and then cycling at 95°C for 3 s and 60°C for 30 s for 40 cycles. There were three samples in each group and each sample was repeated for three times to insure the credibility of the data. The relative quantification of the TwFPS transcript levels was achieved by the 2-ΔΔCt [26] method using ABI 7500 Software v2.0.1 (PE Applied Biosystems).

Results

Molecular cloning of the full-length TwFPS1 andTwFPS2 cDNAs

RT-PCR was performed with total RNA from T. wilfordii. TwFPS gene fragments were obtained by 3′ rapid amplification of cDNA ends (3′-RACE-PCR) and 5′-RACE-PCR. The full-length cDNA encoding the FPS protein was isolated from T. wilfordii. The full-length cDNA of TwFPS1was 1345bp, with a1029bp ORF, which encodes a 342 amino acid polypeptide, flanked by an 81bp 5′-untranslated region and a 235bp 3′-untranslated region including a 28bp poly(A) tail. The predicted TwFPS1 protein had a calculated molecular mass of 39.54 kDa and a theoretical pI of 5.59 (GenBank accession number KM058711). The full-length cDNA of TwFPS2is1312bp, with a1029bp ORF, which encodes a 342 amino acid polypeptide, flanked by a 68bp 5′-untranslated region and a 215bp 3′-untranslated region including a of 26bp poly (A) tail. The predicted TwFPS2protein had a calculated molecular mass of 39.54 kDa and a theoretical pI of 5.28 (GenBank accession number KM058712).

A BLAST search of the NCBI protein database showed that the deduced amino acid sequence of TwFPS1 had 77–85% identity to the FPSs from Mangifera indica, Panax quinquefolius, Panax ginseng, Panax notoginseng, Eucommia ulmoides, Aralia elata, Centella asiatica, Astragalus membranaceus and Malus domestica. The deduced amino acid sequence of TwFPS2 had 78–88% identity to those FPSs from Glycyrrhiza uralensis, Astragalus membranaceus, Hevea brasiliensis, Euphorbia pekinensis, Medicago sativa, Pyrus communis, Aquilaria sinensis, Gossypium hirsutum, and Gentiana lutea. The two proteins are highly conserved; the sequence identity was 77.26%, and the amino acid identity was 80.12%. Both TwFPS1 and TwFPS2had five conserved regions, which were numbered I to V [27]. The highly conserved aspartate-rich motif in region II with the sequence DDXX(XX)D is called FARM (first Asp-rich motif), which is highly conserved in all known prenyltransferases and which has been designated as the chain length determination region [28]. Region V with the sequence DDXXD is called SARM (second Asp-rich motif). These regions marked with lines are characteristic of prenyltransferases that can be used to synthesise isoprenoid diphosphates [29] (Fig 2).

thumbnail
Fig 2. Comparison of the deduced amino acid sequences ofTwFPS1, TwFPS2 and related proteins.

The five conserved domains of prenyltransferases are boxed and numbered. The highly conserved aspartate-rich motif (DDXX(XX)D) was present in domains II and V.

https://doi.org/10.1371/journal.pone.0125415.g002

A phylogenetic tree of isoprenyl diphosphate synthases including GPPS, FPPS and GGPPS from different organisms was constructed to investigate the evolutionary relations. All of the plant isoprenyl diphosphate synthases sequences were separated into three main groups. T. wilfordii1 and T. wilfordii2 clustered with 28 FPPS sequences. A. grandis, G. rigescens GPPS and M. luteus, E. coli FPPS clustered with 11 GGPPS sequences. This tree showed that FPSs evolved from a common ancestor and that the two TwFPSs belong to the clade of the plants (Fig 3). T. wilfordii1 and Eucommia ulmoides were classified into one cluster, and T. wilfordii 2 and Malus domestica were classified into one cluster. These clusters mean that these plants had the closest evolutionary relations. Moreover, these plants all belong to Angiospermae Dicotyledoneae. However, the two genes are not close, which suggests to us that a certain difference exists between TwFPS1 and TwFPS2.

thumbnail
Fig 3. Phylogenetic tree of the amino acid sequences of isoprenyl diphosphate synthase of different organisms constructed by the neighbor-joining method on MEGA 5.

GenBank accession numbers: Hevea brasiliensis (AY135188); Euphorbia pekinensis (ACN63187); Lupinus albus (P49351); Malus domestica (AAM08927); Gossypium arboretum (CAA72793); Helianthus annuus (AAC78557); Parthenium argentatum (CAA57892); Matricaria chamomilla var. recutita (ABS11699); Artemisia annua (AAD17204); Centella asiatica (AAV58896); Panax ginseng (AAY87903); Panax notoginseng (AAY53905); Humulus lupulus (BAB40665); Eucommia ulmoides (AB052681); Capsicum annuum (CAA59170); Chimonanthus praecox (ACJ38671); Michelia chapensis (GQ214406); Musa acuminate (AAL82595); Taxus media (AAS19931); Ginkgo biloba (AY389818); Picea abies (ACA21460); Oryza sativa (O04882); Zea mays (P49353.1); Sorghum bicolor (XP_002441458); Saccharomyces cerevisiae (p08524); Fusarium fujikuroi (CAA65641); Mus musculus (AAL09445); Caenorhabditis elegans (CAB03221); Quercus robur (CAC20852); Citrus sinensis (CAC16851); Catharanthus roseus (AHA82035); Vitis vinifera (AAR08151); Salvia miltiorrhiza (AEZ55677); Arabidopsis thaliana (NP_001031483); Micrococcus luteus (BAA25265); Escherichia coli (BAA00599); Jatropha curcas (ADD82422); Pinus massoniana (AGU43761); Abies grandis (AAL17614.2); Abies grandis (AAN01133); Ginkgo biloba (AAQ72786); Taxus x media (AAS67008); Salvia miltiorrhiza (ACJ66778); Nicotiana attenuate (ABQ53935); Gentiana rigescens (AHK06853); Jasminum sambac (AIY24421); Corylus avellana (ABW06960); Elaeagnus umbellate (ACO59905); Medicago sativa (ADG01841).

https://doi.org/10.1371/journal.pone.0125415.g003

Characterisation of TwFPS1 andTwFPS2 recombinant protein

The entire reading frame of TwFPSs was cloned into the pET-32a(+) vector and expressed in E. coli BL21 (DE3) cells to obtain the TwFPS protein for characterising the farnesylpyrophosphate synthase activity. After induction by IPTG, the recombinant protein was expressed. The molecular mass of TwFPS1 (Fig 4A) fused with a Trx-tag and a His-tag on N-terminal is approximately 58 kDa, and the TwFPS2 (Fig 4B) fusion protein is approximately 58 kDa, as determined by SDS-PAGE. The Trx-tag is a fusion tag and advantageous to the soluble protein expression. His-tag is a purification tag. Highly purified preparation of the protein can be obtained through purification with Ni-column affinity chromatography.

thumbnail
Fig 4. SDS-PAGE analysis of recombinant TwFPS1and TwFPS2 protein expressed in E. coli.

A Lane M, protein molecular weight marker(low); Lane 1, the supernatant of the empty vector without the induction; Lane2, the sediment of the empty vector without the induction; Lane 3, the supernatant of the empty vector with the induction; Lane 4, the sediment of the empty vector with the induction; Lane 5, the supernatant of theTwFPS1 protein without the induction; Lane 6, the sediment of the TwFPS1 protein without the induction; Lane7, the supernatant of theTwFPS1 protein with the induction; Lane 8, the sediment of TwFPS1 protein with the induction; B Lane M, protein molecular weight marker (low); Lane 1, the supernatant of theTwFPS2protein with the induction; Lane 2,the supernatant of the empty vector bacteria with the induction; Lane 3,the supernatant of theTwFPS2 bacteria with the induction; Lane 4, the sediment of the empty vector with the induction; Lane 5, the supernatant of the empty vector with the induction; Lane 6, the sediment of TwFPS2 protein with the induction.

https://doi.org/10.1371/journal.pone.0125415.g004

The purified proteins were assayed for farnesylpyrophosphate synthase catalytic activity. When the purified enzyme was incubated with DMAPP and IPP, the products had the same retention time as the farnesol standards (Fig 5A5D). The GC retention time (RT) of farnesol was 29.588 min; TwFPS1 samples of the product, 29.590 min; and TwFPS2 samples of the product, 29.584 min. The blank control sample was not detected in the corresponding characteristic peak. Under GC-MS analysis (Fig 5E5G), the FPS sample product qualified as farnesol, which had the characteristic peaks, including m/z = 222.0 (Molecular ion: M+) and m/z = 69.10 (CH3(CH3) = CHCH2-). Thus, these results indicated that the coding regions of TwFPSs encode functional FPP synthase.

thumbnail
Fig 5. GC—MS analysis of reaction products catalyzed by purified recombinant TwFPS incubated with IPP and DMAPP.

A Control (the empty vector). B The reaction products catalyzed by purified recombinant TwFPS1 (IPP and DMAPP were added to the reaction mixture). C The reaction products catalyzed by purified recombinant TwFPS2 (IPP and DMAPP were added to the reaction mixture). D GC—MS analysis of dephosphorylated FPP (farnesol) as standards. E The mass spectrogram of the reaction products catalyzed by purified recombinant TwFPS1.F The mass spectrogram of the reaction products catalyzed by purified recombinant TwFPS2. G The mass spectrogram of the dephosphorylated FPP(farnesol).

https://doi.org/10.1371/journal.pone.0125415.g005

Tissue-specific and inducible expression of TwFPS1 andTwFPS2

When total RNAs were isolated from the roots, stems and leaves of T. wilfordii, the lowest level of FPPS mRNA expression was found in leaves. Its value was set up as 1. The other FPPS expression tissues were evaluated relative to the leaves level. We found that TwFPSs are preferentially expressed in the roots. Among different tissues of T. wilfordii, the highest transcript levels of TwFPS1 (Fig 6A) and TwFPS2 (Fig 6B) were observed in the roots, and the lowest levels of TwFPS1and TwFPS2expression were found in the leaves. The highest levels of TwFPS1 expression were observed in roots, about 9.5 fold higher than in leaves. The highest levels of TwFPS2 expression were also observed in roots, about 21.4 fold higher than in leaves. It is indicated that roots are the main source of active contents.

thumbnail
Fig 6. Expression patterns of TwFPS1and TwFPS2 in different T. wilfordii tissues.

Total RNA isolated from roots, stems and leaves. A TwFPS1 expression in leaves was set as1; B TwFPS2 expression in leaves was set as 1. Data are presented as mean±SE from three experimental replicates.

https://doi.org/10.1371/journal.pone.0125415.g006

Moreover, real-time PCR analysis with cell suspensions at different developmental stages was also performed to examine the changes in the expression of TwFPS genes upon MeJA treatment. Specifically, MeJA caused a significant increase in TwFPS levels in T. wilfordii cell suspensions. Interestingly, the dynamic range of the induction varied. The levels of TwFPS1(Fig 7A) and TwFPS2 (Fig 7B) expression both increased at first, peaked at 12 h, then decreased gradually, and the expression levels reached almost the same levels that of the control group after 48 h. The value of 0h was set up as 1 and the TwFPSs mRNA expression in other stage was evaluated relative to the 0h. The TwFPS1 expression of MeJA group at 12h was 2.4-hold higher than control group. The TwFPS2 expression of MeJA group at 12h was 4-hold higher than control group. These results indicated that the expression of TwFPS2relative to 0 h was higher than that of TwFPS1.

thumbnail
Fig 7. Expression profile of TwFPS1 and TwFPS2 when treated with 1mM methyl jasmonate (MeJA) over 48h.

RT-PCR analysis was performed using total RNA isolated from suspension cells of T. wilfordii. A TwFPS1 expression at 0h was set as 1; B TwFPS2 expression at 0h was set as 1. Data are presented as mean±SE from three experimental replicates.

https://doi.org/10.1371/journal.pone.0125415.g007

Discussion

FPS plays a key role [3032] in the catalytic reaction in isoprenoid biosynthesis; this step is considered rate-limiting. In the present study, we reported the molecular characterisation of two FPS genes from T. wilfordii for the first time. Knowledge regarding the regulation of sesquiterpene biosynthesis in medicinal plants is important. The results indicated that T. wilfordii contains a small FPS gene family that consists of at least two genes (TwFPS1 and TwFPS2). These two genes shared a high level of sequence similarity in the coding region but not in noncoding regions. The root of T. wilfordii is the medicinal portion used in Chinese medicine. Pharmacological experiments confirmed that the primary active terpenoids of T. wilfordii showed strong pharmacological activities. The tissue expression analysis showed high TwFPS expression in the roots of T. wilfordii. This indicated that the biosynthesis of sesquiterpene compounds, such as wilfordine, occurs at roots. Moreover, the relative expression level ofTwFPS2 was higher than that ofTwFPS1, suggesting thatTwFPS2plays a leading role in terpene synthase biosynthesis in T. wilfordii. The analysis of FPS gene expression patterns in various plants demonstrated not only that the gene expression is tissue-specific [3334] but also that the isoprenoid derivative contents increase [20, 30, 3538]. The gene expression analysis showed that FPS genes could increase the isoprenoid substance contents in the plants [3942]. In Poria cocos, there is a significant difference in total triterpenoids production between the control and experimental groups demonstrating that MeJA can potently stimulate triterpenoids biosynthesis [43]. The specific synthetic mechanism of terpene synthase biosynthesis requires further research.

In Arabidopsis thaliana, for example, FPS exists in the form of a small gene family. The FPS gene family encodes three isomers of FPS (FPS1S, FPS1L and FPS2). The primary difference between FPS1S and FPS1L is the N-terminus. FPS1S and FPS2 are found in plastids, and FPS1L is found in mitochondria [32]. The different metabolic channels determine their different functions and expression patterns. Thus, the secondary metabolism that FPS is involved in is complex, with great diversity and specificity. A study of rice shows that FPS exists in chloroplasts and in plastids [44]. FPS1 is expressed in different parts of the plants and participated in all life processes. FPS2 is primarily expressed in the flowers, as well as in the root tips of the lateral root and in the juncture with the primary root or secondary root [45]. The subcellular localisation prediction showed that the major localisation of TwFPS1 is the cytoplasm; however, the localisation of TwFPS2 is the cytoplasm or mitochondria. Therefore, a tentative inference of this result is that TwFPS1 and TwFPS2 participate in different metabolic channels.

These findings suggest that the expression of the native FPS genes in T. wilfordii result in the accumulation of the active components. The cloning and identification of key genes involved in the biosynthesis of active compounds from medicinal plants is important for the analysis of synthesis pathways. A recently established strategy in synthetic biology [4649] is to use microorganisms such as E. coli to synthesize active ingredients found in medicinal plants, which will provide a new effective strategy and research approach for the sustainable utilisation of medicinal plant resources.

Accession Number

The encoded nucleotide sequence can be found in the GenBank/NCBI data libraries. The number of TwFPS1 is KM058711 and the number of TwFPS2 is KM058712.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (81422053 and 81373906 to WG, and 81325023 to LH) and the Author of National Excellent Doctoral Dissertation of China (201188) and the Importation and Development of High-Caliber Talents Project of Beijing Municipal Institutions (CIT&TCD201304174) to WG.

Author Contributions

Conceived and designed the experiments: WG XC LH XW YZ. Performed the experiments: YZ PS MZ YL YT XC WG. Analyzed the data: YZ WG PS MZ YL YT. Contributed reagents/materials/analysis tools: WG XC XW LH. Wrote the paper: YZ PS MZ YL YT XC XW LH WG.

References

  1. 1. Brinker AM, Ma J, Lipsky PE, Raskin I. Medicinal chemistry and pharmacology of genus Tripterygium (Celastraceae). Phytochemistry. 2007; 68:732–766. pmid:17250858
  2. 2. Titov DV, Gilman B, He QL, Bhat S, Low WK, Dang Y, et al. XPB, a subunit of TFIIH, is a target of the natural product triptolide. Nat Chem Biol.2011; 7:182–188. pmid:21278739
  3. 3. Manzo SG, Zhou ZL, Wang YQ, Marinello J, He JX, Li YC, et al. Natural product triptolide mediates cancer cell death by triggering CDK7-dependent degradation of RNA polymerase II. Cancer Res. 2012; 72:5363–5373. pmid:22926559
  4. 4. Chugh R, Sangwan V, Patil SP. Dudeja V, Dawra RK, Banerjee S, et al. A preclinical evaluation of Minnelide as a therapeutic agent against pancreatic cancer. Science Transl Med.2012; 4:156ra139.
  5. 5. Kupchan SM, Court WA, Dailey RG Jr, Gilmore CJ, Bryan RF. Tumor inhibitors. LXXIV. Triptolide and tripdiolide, novel antileukemic diterpenoid triepoxides from Tripterygium wilfordii. J Am Chem Soc.1972; 94:7194–7195. pmid:5072337
  6. 6. Yang H, Chen D, Cui QC, Yuan X, Dou QP. Celastrol, a triterpene extracted from the Chinese “Thunder of God Vine,” is a potent proteasome inhibitor and suppresses human prostate cancer growth in nude mice. Cancer Res. 2006; 66:4758–4765. pmid:16651429
  7. 7. Chappell J. Biochemistry and molecular biology of the isoprenoid biosynthetic pathway in plants. Annu Rev Plant Biol.1995; 46:521–547.
  8. 8. Cornish K. The separate roles of plant cis and trans prenyl transferases in cis-1, 4-polyisoprene biosynthesis. Eur J Biochem.1993; 218:267–271. pmid:8243473
  9. 9. Lange BM, Rujan T, Martin W, Croteau R. Isoprenoid biosynthesis: the evolution of two ancient and distinct pathways across genomes. Proc Natl Acad Sci U S A.2000; 97:13172–13177. pmid:11078528
  10. 10. Szkopińska A, Płochocka D. Farnesyl diphosphate synthase; regulation of product specificity. Acta Biochim Pol.2005; 52: 45–55. pmid:15827605
  11. 11. McGarvey DJ, Croteau R. Terpenoid metabolism. Plant Cell. 1995; 7: 1015–1026. pmid:7640522
  12. 12. Chappell J, Nable R. Induction of sesquiterpenoid biosynthesis in tobacco cell suspension cultures by fungal elicitor. Plant Physiol. 1987; 85: 469–473. pmid:16665722
  13. 13. Threlfall DR, Whitehead IM. Co-ordinated inhibition of squalene synthetase and induction of enzymes of sesquiterpenoid phytoalexin biosynthesis in cultures of Nicotiana tabacum. Phytochemistry. 1988; 27: 2567–2580.
  14. 14. Vögeli U, Chappell J. Induction of sesquiterpene cyclase and suppression of squalene synthetase activities in plant cell cultures treated with fungal elicitor. Plant Physiol. 1988; 88: 1291–1296. pmid:16666457
  15. 15. Paterson AH, Bowers JE, Bruggmann R, Dubchak I, Grimwood J, Gundlach H, et al. The Sorghum bicolor genome and the diversification of grasses. Nature.2009; 457:551–556. pmid:19189423
  16. 16. Alexandrov NN, Brover VV, Freidin S, Troukhan ME, Tatarinova TV, Zhang H, et al. Insights into corn genes derived from large-scale cDNA sequencing. Plant Mol Biol.2009; 69: 179–194. pmid:18937034
  17. 17. Jones CG, Moniodis J, Zulak KG, Scaffidi A, Plummer JA, Ghisalberti EL, et al. Sandalwood fragrance biosynthesis involves sesquiterpene synthases of both the terpene synthase (TPS)-a and TPS-b subfamilies, including santalene synthases. J Biol Chem.2011; 286:17445–17454. pmid:21454632
  18. 18. Kim OT, Kim SH, Ohyama K, Muranaka T, Choi YE, Lee HY, et al. Upregulation of phytosterol and triterpene biosynthesis in Centellaasiatica hairy roots overexpressed ginseng farnesyl diphosphate synthase. Plant cell Rep.2010; 29:403–411. pmid:20195611
  19. 19. Matsushita Y, Kang W, Charlwood BV. Cloning and analysis of a cDNA encoding farnesyl diphosphate synthase from Artemisia annua. Gene.1996; 172:207–209. pmid:8682304
  20. 20. Wang P, Liao Z, Guo L, Li W, Chen M, Pi Y, et al. Cloning and functional analysis of a cDNA encoding Ginkgo biloba farnesyl diphosphate synthase. Mol Cells.2004; 18:150–156. pmid:15528989
  21. 21. Wang Y, Weathers PJ. Springer New York. Sugars proportionately affect artemisinin production. Plant Cell Rep. 2007; 26: 1073–1081. pmid:17221224
  22. 22. Nkembo KM, Lee JB, Hayashi T. Selective enhancement of scopadulcic acid B production in the cultured tissues of Scopariadulcis by methyl jasmonate. Chem Pharm Bull.2005; 53:780–782. pmid:15997134
  23. 23. Lambert E, Faizal A, Geelen DJ. Modulation of triterpene saponin production: in vitro cultures, elicitation, and metabolic engineering. Appl Biochem Biotechnol.2011; 164: 220–237. pmid:21267794
  24. 24. Del Sal G, Manfioletti G, Schneider C. The CTAB-DNA precipitation method: a common mini-scale preparation of template DNA from phagemids, phages or plasmids suitable for sequencing. Biotechniques.1989; 7:514–520. pmid:2699240
  25. 25. Bradford MM. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem.1976; 72:248–254. pmid:942051
  26. 26. Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Methods.2001; 25:402–408. pmid:11846609
  27. 27. Chen A, Kroon PA, Poulter CD.Isoprenyl diphosphate synthases: protein sequence comparisons, a phylogenetic tree, and predictions of secondary structure. Protein Sci.1994; 3: 600–607. pmid:8003978
  28. 28. Gupta P, Akhtar N, Tewari SK, Sangwan RS, Trivedi PK. Differential expression of farnesyl diphosphate synthase gene from Withania somnifera in different chemotypes and in response to elicitors. Plant Growth Regul.2011; 65:93–100.
  29. 29. Ohnuma Si, Narita K, Nakazawa T, Ishida C, Takeuchi Y, Ohto C, et al. A role of the amino acid residue located on the fifth position before the first aspartate-rich motif of farnesyl diphosphate synthase on determination of the final product. J Biol Chem.1996; 271:30748–30754. pmid:8940054
  30. 30. Liao ZH, Chen M, Gong YF, Zhu GL, Kai JZ, Peng W, et al. A new Farnesyl diphosphate synthase gene from Taxus media Rehder: cloning, characterization and functional complementation. J Integer Plant Biol.2006; 48: 692–699.
  31. 31. Thabet I, Guirimand G, Courdavault V, Papon N, Godet S, Dutilleul C, et al. The subcellular localisation of periwinkle farnesyl diphosphate synthase provides insight into the role of peroxisome in isoprenoid biosynthesis. J Plant Physiol.2011; 168: 2110–2116. pmid:21872968
  32. 32. Closa M, Vranová E, Bortolotti C, Bigler L, Arró M, Ferrer A, et al. The Arabidopsis thaliana FPP synthase isozymes have overlapping and specific functions in isoprenoid biosynthesis, and complete loss of FPP synthase activity causes early developmental arrest. Plant J.2010; 63: 512–525. pmid:20497375
  33. 33. Yin T, Cao X, Miao Q, Li C, Chen X, Zhou M,et al. Molecular cloning and functional analysis of an organ-specific expressing gene coding for farnesyl diphosphate synthase from Michelia chapensis Dandy.Acta Physiol Plant.2011; 33: 137–144.
  34. 34. Moco S, Capanoglu E, Tikunov Y, Bino RJ, Boyacioglu D, Hall RD, et al. Tissue specialization at the metabolite level is perceived during the development of tomato fruit. J Exp Bot.2007; 58: 4131–4146. pmid:18065765
  35. 35. Cunillera N, Arró M, Delourme D, Karst F, Boronat A, Ferrer A. Arabidopsis thaliana contains two differentially expressed farnesyl-diphosphate synthase genes. J Biol Chem.1996; 271:7774–7780. pmid:8631820
  36. 36. Cervantes-Cervantes M, Gallagher CE, Zhu C, Wurtzel ET. Maize cDNAs expressed in endosperm encode functional farnesyl diphosphate synthase with geranylgeranyl diphosphate synthase activity. Plant Physiol.2006; 141:220–231. pmid:16581875
  37. 37. Gaffe J, Bru JP, Causse M, Vidal A, Stamitti-Bert L, Carde JP, et al. LEFPS1, a tomato farnesyl pyrophosphate gene highly expressed during early fruit development. Plant Physiol.2000; 123:1351–1362. pmid:10938353
  38. 38. Li CP, Larkins BA. Identification of a maize endosperm-specific cDNA encoding farnesyl pyrophosphate synthetase. Gene.1996; 171:193–196. pmid:8666271
  39. 39. Devarenne TP, Shin DH, Back K, Yin S, Chappell J. Molecular characterization of tobacco squalene synthase and regulation in response to fungal elicitor. Arch Biochem Biophys.1998; 349: 205–215. pmid:9448707
  40. 40. Chen DH, Ye H, Li GF.Expression of a chimeric farnesyl diphosphate synthase gene in Artemisia annua L. transgenic plants via Agrobacterium tumefaciens-mediated transformation. Plant Sci.2000; 155:179–185. pmid:10814821
  41. 41. Han JL, Liu BY, Ye HC, Wang H, Li ZQ, Li GF. Effects of overexpression of the endogenous farnesyl diphosphate synthase on the artemisinin content in Artemisia annua L. J Intergr Plant Biol.2006; 48:482–487.
  42. 42. Kim OT, Bang KH, Jung SJ, Kim YC, Hyun DY, Kim SH, et al. Molecular characterization of ginseng farnesyl diphosphate synthase gene and its up-regulation by methyl jasmonate. Biol Plantarum.2010; 54:47–53.
  43. 43. Wang J, Li Y, Liu D. Cloning and Characterization of Farnesyl Diphosphate Synthase Gene Involved in Triterpenoids Biosynthesis from Poria cocos. Int J Mol Sci.2014; 15: 22188–22202. pmid:25474088
  44. 44. Sanmiya K, Ueno O, Matsuoka M, Yamamoto N. Localization of farnesyl diphosphate synthase in chloroplasts. Plant Cell Physiol.1999; 40:348–354. pmid:10353221
  45. 45. Cunillera N, Boronat A, Ferrer A. Spatial and temporal patterns of GUS expression directed by 5’regions of the Arabidopsis thaliana farnesyl diphosphate synthase genesFPS1 and FPS2. Plant Mol Biol.2000; 44:747–758. pmid:11202437
  46. 46. Moses T, Pollier J, Thevelein JM, Goossens A. Bioengineering of plant (tri)terpenoids: from metabolic engineering of plants to synthetic biology in vivo and in vitro. New Phytol.2013; 200: 27–43. pmid:23668256
  47. 47. Vervacke JS, Wang YC, Distefano MD. Photoactive analogs of farnesyl diphosphate and related isoprenoids: design and applications in studies of medicinally important isoprenoid-utilizing enzymes. Curr Med Chem.2013; 20:1585–1594. pmid:23409719
  48. 48. Banyai W, Kirdmanee C, Mii M, Supaibulwatana K. Overexpression of farnesyl pyrophosphate synthase (FPS) gene affected artemisinin content and growth of Artemisia annua L. Plant Cell Tiss Org (PCTOC).2010; 103: 255–265.
  49. 49. Alcaíno J, Romero I, Niklitschek M, Sepúlveda D, Rojas MC, Baeza M, et al. Functional characterization of the Xanthophyllomyces dendrorhous farnesyl pyrophosphate synthase and geranylgeranyl pyrophosphate synthase encoding genes that are involved in the synthesis of isoprenoid precursors. PLoS One.2014; 9: e96626. pmid:24796858