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

RNA-Seq Based De Novo Transcriptome Assembly and Gene Discovery of Cistanche deserticola Fleshy Stem

  • Yuli Li ,

    Contributed equally to this work with: Yuli Li, Xiliang Wang, Tingting Chen

    Affiliations CAS Key Laboratory of Genome Sciences and Information, Beijing Institute of Genomics, Chinese Academy of Sciences, Beijing, China, University of Chinese Academy of Sciences, Beijing, China

  • Xiliang Wang ,

    Contributed equally to this work with: Yuli Li, Xiliang Wang, Tingting Chen

    Affiliations CAS Key Laboratory of Genome Sciences and Information, Beijing Institute of Genomics, Chinese Academy of Sciences, Beijing, China, University of Chinese Academy of Sciences, Beijing, China

  • Tingting Chen ,

    Contributed equally to this work with: Yuli Li, Xiliang Wang, Tingting Chen

    Affiliation Core Genomic Facility, Beijing Institute of Genomics, Chinese Academy of Sciences, Beijing, China

  • Fuwen Yao,

    Affiliations Core Genomic Facility, Beijing Institute of Genomics, Chinese Academy of Sciences, Beijing, China, University of Chinese Academy of Sciences, Beijing, China

  • Cuiping Li,

    Affiliation Core Genomic Facility, Beijing Institute of Genomics, Chinese Academy of Sciences, Beijing, China

  • Qingli Tang,

    Affiliation HongKui CongRong Group, Alashan, Inner Mongolia, China

  • Min Sun,

    Affiliation Core Genomic Facility, Beijing Institute of Genomics, Chinese Academy of Sciences, Beijing, China

  • Gaoyuan Sun,

    Affiliation Core Genomic Facility, Beijing Institute of Genomics, Chinese Academy of Sciences, Beijing, China

  • Songnian Hu,

    Affiliation CAS Key Laboratory of Genome Sciences and Information, Beijing Institute of Genomics, Chinese Academy of Sciences, Beijing, China

  • Jun Yu,

    Affiliation CAS Key Laboratory of Genome Sciences and Information, Beijing Institute of Genomics, Chinese Academy of Sciences, Beijing, China

  • Shuhui Song

    songshh@big.ac.cn

    Affiliations CAS Key Laboratory of Genome Sciences and Information, Beijing Institute of Genomics, Chinese Academy of Sciences, Beijing, China, Core Genomic Facility, Beijing Institute of Genomics, Chinese Academy of Sciences, Beijing, China

Abstract

Backgrounds

Cistanche deserticola is a completely non-photosynthetic parasitic plant with great medicinal value and mainly distributed in desert of Northwest China. Its dried fleshy stem is a crucial tonic in traditional Chinese medicine with roles of mainly improving male sexual function and strengthening immunity, but few mechanistic studies have been conducted partly due to the lack of genomic and transcriptomic resources.

Results

In this study, we performed deep transcriptome sequencing in fleshy stem of C. deserticola, and about 80 million reads were generated using Illumina pair-end sequencing on HiSeq2000 platform. Using trinity assembler, we obtained 95,787 transcript sequences with transcript lengths ranging from 200bp to 15,698bp, having an average length of 950 bases and the N50 length of 1,519 bases. 63,957 transcripts were identified actively expressed with FPKM ≥ 0.5, in which 30,098 transcripts were annotated with gene descriptions or gene ontology terms by sequence similarity analyses against several public databases (Uniprot, NR and Nt at NCBI, and KEGG). Furthermore, we identified key enzyme genes involved in biosynthesis of lignin and phenylethanoid glycosides (PhGs) which are known to be the primary active ingredients. Four phenylalanine ammonia-lyase (PAL) genes, the first key enzyme in lignin and PhG biosynthesis, were identified based on sequences comparison and phylogenetic analysis. Two biosynthesis pathways of PhGs were also proposed for the first time.

Conclusions

In all, we completed a global analysis of the C. deserticola fleshy stem transcriptome using RNA-seq technology. A collection of enzyme genes related to biosynthesis of lignin and phenylethanoid glysides were identified from the assembled and annotated transcripts, and the gene family of PAL was also predicted. The sequence data from this study will provide a valuable resource for conducting future phenylethanoid glysides biosynthesis researches and functional genomic studies in this important medicinal plant.

Introduction

C. deserticola is a worldwide genus of perennial desert plants from the Orobanchaceae family, and is a completely non-photosynthetic species and usually grows underground holoparasitic plant [1]. It is parasitized on the roots of psammophyte Haloxylon ammodendron (Chenopodiaceae) [1, 2], which mainly inhabits deserts and semi-deserts due to its high tolerance to drought and salinity [1, 3]. C. deserticola shows strong resistance to harsh environmental conditions and is mainly distributed in Northwest China [46], especially in Inner Mongolia, Gansu and Xinjiang. It is considered to be an endangered wild species in recent years due to increased consumption by humans [5, 6]. C. deserticola which is often called desert ginseng is commonly known as desert-broomrape and the dried fleshy stem has been extensively used as a traditionally important tonic in China and Japan for many years [4, 710]. It was initially recorded in Shen Nong Ben Cao Jing (Dictionary of Chinese Materia Medica, 1977) [11] approximately 1800 years ago and was regarded as one of the main sources of the Chinese medicinal herba Cistanche.

The extracts of C. deserticola possess a wide range of medicinal functions, especially for use in improving sexual function, tonifying kidney, protecting liver, aperient activity, enhancing memory, immunomodulatory, antioxidative activity, anti-inflammatory, antiviral activity etc [710, 1215]. The major bioactive components of C. deserticola are Phenylethanoid glycosides (PheGs, PhGs) [2, 9, 10, 14, 15]. To date, more than 20 phenylethanoid glycosides have been isolated from the succulent stem of C.deserticola [9, 14, 16]. Among them, acteoside and echinacoside are two main components with significant pharmacological activities [2, 16], and documented as the quality standards of C. deserticola in the Chinese pharmacopeia (2005 and 2010 editions). Three chemical components of PhGs are organic acid, saccharide and phenylethanoid, however, the details concerning phenylethanoid biosynthetic pathways remain poorly understood in C.deserticola.

Despite the commercial and medicinal importance of C.deserticola, the genomic and transcriptomic data of this species are very limited. There is no ESTs available in the NCBI database and the complete genome information for this species remains unavailable except for the chloroplast genome sequence [1]. The limited transcriptomic data hinder the study of PhG biosynthetic mechanisms. RNA-seq technology can generate sequences of the expressed parts of targeted genome [17] and identify genes [18] using the NGS technology platforms (such as Applied Biosystems SOLiD, Illumina HiSeq and Roche 454). It is becoming increasingly popular in transcriptome de novo assembly [1922], since it is a cost-effective and powerful approach with high resolution and broad dynamic range [2325], especially that it has an advantage to explore low abundance transcripts [26]. Because of the various advantages, RNA-seq is specifically attractive for non-model organisms with limited genetic resources [2729]. But there is no any detailed research of C. deserticola transcriptome by RNA-seq.

In this study, we globally sequenced the stem transcriptome for C. deserticola using Illumina Hiseq2000 platform, and got 7.9G raw data. By assembly and annotation, we mined the genes involved in biosynthesis of PhG and the genes responsible for entire lignin biosynthesis. Our RNA-seq analysis generated the first C. deserticola consensus trancriptome and provided new insights into comprehensive understanding of the medicinal value of C. deserticola. Additionally, the method described here can be widely applied to profile transcriptomes to facilitate the discovery of genes involved in specific medicinal components biosynthesis pathway in other medicinal plant with very limited genomic resources.

Materials and Methods

Plant material collection

The fresh succulent stem for C. deserticola in excavation stage was collected from a plant base in BayanHot City of Alxa League in Inner Mongolia in northwestern China. The collecting permit was obtained from the owner (HongKui CongRong Group) of the plant base. The voucher specimen was deposited in the Core Genomic Facility at Beijing Institute of Genomics, Chinese Academy of Sciences. After cleaning, the succulent stem tissues were cut into small pieces and immediately frozen in liquid nitrogen, and then stored at -80°C until further processing.

RNA extraction, cDNA library construction and Illumina sequencing

Total RNA was extracted from the succulent stem using TRIzol Reagent (Invitrogen Inc., California, USA) according to the manufacturer’s instruction. The resulting samples were treated with DNase I to remove any genomic DNA. Extracted RNAs were quantified using an Agilent 2100 bioanalyzer (Agilent Technologies) and checked for integrity using denaturing agarose gel electrophoresis with ethidium bromide staining. RNA samples with A260/A280 ratios between 1.9 and 2.1, RNA 28S:18S ratios higher than 1.0, and RNA integrity numbers (RINs) ≥ 8.5 were used in subsequent analyses.

The RNA-seq libraries were generated using Illumina Truseq RNA Sample Preparation Kits. Poly(A)+ RNA was isolated from total RNA using Dynal ligo(dT)25 beads according to the manufacturer’s instructions. Following purification, fragmentation buffer was added to break the mRNA into short fragments. First-strand cDNA was synthesized using these short fragments as templates, along with SuperScript III reverse transcriptase and N6 random hexamer primer. Second-strand cDNA was then synthesized using buffer, dNTPs, RNaseH and DNA polymerase I. The resulting double-stranded cDNA was subjected to end-repair using T4 DNA polymerase, DNA polymerase I Klenow fragment, and T4 polynucleotide kinase, and ligated to adapters using T4 DNA ligase. Adaptor-ligated fragments were purified using a QiaQuick PCR extraction kit and eluted with EB buffer. After analysis using agarose gel electrophoresis, suitable fragments were selected as templates for PCR amplification. Sequencing of the resulting cDNA library was carried out with an Illumina HiSeq 2000 system.

Transcripts de novo assembly and gene expression quantification

Raw reads generated from sequencing was cleaned by removing the adaptor sequences (ATCTCGTATGCCGTC) using in-house method. We then carried out a stringent low-quality filtering process. Firstly, bases with phred quality score lower than 20 would be trimmed from the 3’end of the sequence, until running into one base with a higher quality (≥20). If reads length was short than 50bp, it would be discarded. Secondly, reads will be further filtered by the criterion that 70% bases in one read have high quality scores (≥20). Thirdly, only paired-end reads were used for further assembly. De novo transcript assembly was conducted using Trinity release_20130216 [30] which consisted of three successive software modules: Inchworm, Chrysalis, and Butterfly. The assembly parameters were set as below:—seqType fq—JM 300G —min_contig_length 200—CPU 20—inchworm_cpu 20—bflyCPU 20.

To quantify transcript abundance, the sequenced pair-end reads were re-aligned to the assembled transcripts using a script in Trinity [31]. Mapped reads were used for quantification by RSEM (RNA-Seq by Expectation Maximization) software [32]. Gene or isoform abundance was represented by the fragment per kilobase of transcript per million fragment mapped (FPKM) value, those transcripts with FPKM value equal or larger than 0.05 were defined as expressed.

Functional annotation of expressed transcripts

For there is no any gene annotation sets of C. deserticola except for chloroplast genome [1]. We annotated the expressed transcripts by comparing to Genbank Nt, Genbank Nr, and TAIR10_pep_20101214_updated datasets separately using BLAST program (E< = 1e-20). Meanwhile, all expressed transcripts were translated into potential proteins according to ORF prediction by TransDecoder [30] and predicated for the conserved domains based on Pfam database [33].

Gene Ontology and KEGG pathway annotation

By sequence similarity alignment to Uniprot database (http://www.uniprot.org/), the Gene Ontology (GO) annotation of all assembled transcripts was obtained by using association file downloaded from (ftp://ftp.ebi.ac.uk/pub/databases/GO/goa/UNIPROT/gene_association.goa_uniprot.gz). GO terms clustering of expressed genes were conducted by using custom scripts, and we annotated genes at the fourth level for the CC, BP and MF categories separately.

KEGG pathway information was assigned for all predicted proteins sequences using online tool KAAS (KEGG Automatic Annotation Sever) [34]. Sequences in fasta format were submitted to KAAS request, and the resulting files of all pathways information related to C. deseticola stem transcriptome were downloaded. 13 plant organisms’ gene data sets in KEGG were used for annotation using the BBH (bi-directional best hit) method.

RT-qPCR analysis

After digestion with DNase I, approximately 5μg of total RNA was converted into first-strand cDNA via the reverse-transcription reaction with oligo(dT)15 primers and GoScript Reverse Transcription System (Promega). The cDNA products were then diluted 10-fold with nuclease free deionized water before use as a template in real-time PCR. Specific cDNAs were amplified by GoTaq 2-Step RT-qPCR system (Promega) in a volume of 20 ul. PCR amplification was performed at the annealing temperature of 60°C with the 7500 Real-Time PCR Detection System (Applied Biosystems) according to the manufacturer’s instructions. Relative transcript abundances were calculated by the comparative cycle threshold method with gene “comp10579_c0” as an internal standard, using the 7500 Manager software.

Primer pairs for RT-PCR were designed based on online software (http://primer3.ut.ee/) and are listed in S1 Dataset.

Results

RNA sequencing and de novo transcriptome assembly of C. deserticola fleshy stem

Stem of C. deserticola has been extensively used as a traditionally important tonic in China and Japan for many years. To obtain a global overview of gene expression in the C. deserticola fleshy stem, we collected C. deserticola stem samples of the same plant base in 2013 and 2014, respectively. Total RNAs were extracted and polyA+ RNAs were purified for constructing paired-end RNA-seq libraries. 79,433,734 and 86,019,176 pair-end reads corresponding to nearly 8 billion and 8.6 billion bases of sequence were obtained using Illumina HiSeq 2000 sequencing platform in 2013-year and 2014-year samples (Table 1). After removing adaptor sequences and filtering low quality reads (see details in Methods), 64,831,040 high quality pair-end reads in 2013-year sample were used for de novo transcriptome assembly. Using Trinity sequence assembler [30], 51,719 genes and 95,787 transcript sequences were generated with transcript lengths ranging from 200 bp to 15,698 bp. The average length of assembled transcripts is 950 bases and the N50 length is 1,519 bases. The number of transcript in different lengths revealed that 57.32% of the assemble transcript were about 500-bp or longer (Fig 1A). High quality pair-end reads in 2014-year sample were mapped to the assembled transcriptome. Besides, we found that the transcript number for each assembled genes varied and 69% genes with one expressed isoform while 31% genes expressed two or more transcripts (Fig 1B).

thumbnail
Table 1. Description of sequencing data and de novo assembly.

https://doi.org/10.1371/journal.pone.0125722.t001

thumbnail
Fig 1. Distribution of assembled transcript length and statistic of isoform pattern.

(a) Distribution of assembled transcript length binned at 250bp intervals. (b) The percentage of transcript isoform pattern ranging from a single isoform to 6 or more isoforms in a single assembled gene.

https://doi.org/10.1371/journal.pone.0125722.g001

Expression quantification and functional annotation of assembled transcripts

Gene or transcript abundance was quantified using RSEM package, in which the sequenced reads were re-aligned to the assembled genes or transcripts sequences using Bowtie and those mapped reads were used for quantification. FPKM value for each gene or transcript was calculated, and finally we identified 63,957 and 52,857 actively expressed transcripts (FPKM value ≥ 0.5) in C. deserticola fleshy stem samples in 2013 and 2014, respectively. 44,776 transcripts (70.01% in 2013-year sample, 84.71% in 2014-year sample) were commonly expressed in the two replicates, and the correlation (Pearson correlation coefficient: 0.91979) of their expression data was showed in S1 Fig. The sequencing raw data had been uploaded to NCBI SRA database (accession numbers: SRX857402 and SRX858938). We used expressed genes identified in 2013-year sample for further analysis. Functional annotation information for all expressed transcripts was obtained using two methods. Firstly, all expressed transcripts were aligned to known nucleotide (GenBank nt) and peptide sequence databases (GenBank nr and Arabidopsis peptide) separately by BLAST algorithm. Out of 63,957 expressed transcripts, 29,220 (45.7%) were annotated and showed homology to sequences in any of the three subject databases with E-value cutoff 1e-20. Meanwhile, the candidate coding regions for all expressed transcript sequences were predicted using TransDecoder software and the longest ORFs for each transcript were used for Pfam domain search. As a result, 21,358 (33.4%) transcripts were annotated based on Pfam database. Overall, 30,098 (47.1%) transcripts were significantly matched to known genes in the public databases by combining the two methods above. The complete expressed transcripts list with function annotation were showed in supplemental data (S2 Dataset).

We surveyed the top 20 most highly expressed transcripts (Table 2) corresponding to 18.99% of all sequencing reads, and found that most of them are genes responding to abiotic stress stimulus. Dehydrin (DHNs), a class of hydrophilic and thermostable stress proteins with a high number of charged amino acids that belong to the Group II Late Embryogenesis Abundant (LEA) family, is the most highly expressed genes. Three different Dehyrin transcripts (comp28713_c0_seq1/2/4) were detected highly expressed in fleshy stem which may be involved in protecting cells from damage caused by drought stress. Other stress-related genes such as heat shock protein, pathogen-related protein and metallothionein were also found expressed highly, which may be related to its severe survival environment. Additionally, some constitutive genes including 26S ribosomal RNA gene (comp22329_c2_seq1), auxin-repressed/dormancy-associated protein (comp20999_c0_seq1), ADP-ribosylation factor (comp20499_c0_seq1) were also highly transcribed.

thumbnail
Table 2. The top 20 highly expressed genes in fleshy stem tissue of C. deserticola.

https://doi.org/10.1371/journal.pone.0125722.t002

Functional classification of all expressed transcripts based on Gene Ontology and KEGG databases

Gene Ontology (GO) annotation was obtained from Uniprot annotation and identities association file. In total, 20,907 transcripts, accounting for 32.69% of the total expressed sequences, were assigned to 1,745 functional terms. Of the total functional GO terms, assignments to the biological process made up the majority (1,116, 63.95%) followed by cellular component (329, 18.85%) and molecular function (300, 17.20%). The assigned functions of expressed transcripts covered a broad range of GO categories, and the top 10 GO terms with most annotated transcripts were listed in Table 3. And we provide all expressed transcripts distribution in three Gene Ontology categories (Molecular Function, Cellular component and biological process) in supplemental file (S3 Dataset). GO terms related to binding functions and transferase activity were predominantly represented in the molecular function category. About the binding functions, cation binding (4,394 transcripts) represented the most abundant, followed by nucleotide/nucleoside binding (3,404 transcripts in average) and protein binding (2,422 transcripts). While in transferase activity group, the most is those with transferring phosphorus-containing groups (2,256 transcripts, 65.77%). Among the cellular component category, transcripts were more located to intracellular (10,581 transcripts in average), while among the biological process category, transcripts were more involved in biopolymer metabolic process (6,683 transcripts in average), followed by regulation of cellular process (4,841 transcripts), gene expression (4,678 transcripts) and transport (3,512 transcripts).

thumbnail
Table 3. The top 10 GO terms with the highest transcript number in each of the three ontologies.

https://doi.org/10.1371/journal.pone.0125722.t003

In order to mine genes involved in biosynthesis of lignin and PhG, 21,358 non-redundant potential protein sequences were searched against gene sequences of 13 plant organisms in KEGG database, and they were assigned to 275 KEGG pathways with at least 5 hits. The top 10 pathways with most aligned sequences were listed in Table 4. Most pathways were involved in primary metabolic processes, such as amino acid or protein metabolism (ko01230, ko04141 and ko04120), carbohydrate metabolism (ko01200 and ko00500) and nucleotide or nucleoside metabolism (ko03018, ko00230, and ko00240). Besides, there are 27 secondary metabolism related pathways (Fig 2), such as terpenoid backbone biosynthesis, phenylpropanoid biosynthesis, carotenoid biosynthesis, isoquinoline alkaloid biosynthesis and tropane, piperidine and pyridine alkaloid biosynthesis. These results provide further indication that active metabolic processes were underway in the C. deserticola stem tissue. All expressed transcripts associated with KEGG pathways were listed in supplemental file (S4 Dataset). Although there are some significantly changed pathways between C. deserticola and other plants, such as rice [35] (S5 Dataset), our main goal in this study aims to reveal the whole transcriptome profile of C. deserticola stem, and to picture the related pathways of PhGs biosynthesis which could be useful for guiding the cultivation.

thumbnail
Table 4. The top 10 KEGG pathways with most protein sequences.

https://doi.org/10.1371/journal.pone.0125722.t004

thumbnail
Fig 2. The KEGG pathways involved in secondary metabolism in C. deserticola.

KEGG classification based on secondary metabolite categories sorted by counts of transcripts in a single pathway.

https://doi.org/10.1371/journal.pone.0125722.g002

Candidate genes encoding enzymes involved in the biosynthesis of lignin

Lignin is the second most abundant natural terrestrial polymer in plant kingdom, composing up to one-third of the material found in plant cell walls [36, 37]. As an important component of cell wall, lignins help water transport, provide mechanical support and structural integrity, and defend against pathogens and herbivores [3840]. Those roles of lignin are much valuable to support underground erective growth of C. deserticola in desert. In this study, we presented the complete picture of lignin biosynthesis pathways in C. deserticola (Fig 3), in which the lignin monomers are biosynthesized from phenylalanine through a series of enzymatic reactions, including hydroxylation, methylation, reduction, and oxidative polymerization process. Lignin biosynthesis-related enzymes were detected for three mainly synthesized forms in vascular tissue (p-hydroxyl-phenyl (H), guaiacyl (G) and syringyl (S) lignin) and 5-hydroxyl-guaiacyl lignin which was only identified in COMT (caffeic acid 3-O-methyltransferase, EC 2.1.1.68) deficient (such as knock-down) plants [41, 42].

thumbnail
Fig 3. Detailed pathways and genes involved in the synthesis of lignin and PhGs in C. deserticola stem.

The top and bottom panels showed the detailed pathways and genes involved in the biosynthesis of four types lignin and PhGs, respectively. Red diamonds represent the common pathway shared by lignin and PhGs. Green dotted lines were used to represent the synthesis pathway of 5-Hydroxyl-Guaiacyl lignin, which was identified in COMT deficient plants. Abbreviation: PAL(Phenylalanine ammonialyas), 4CL(4-coumarate-CoA ligase), CYP73A(trans-cinnamate 4-monooxygenase), CCR(cinnamoyl-CoA reductase), HCT(shikimate o-hydroxycinnamoyltransferase), F5H(ferulate-5-hydroxylase), COMT(3-O-methyltransferase), CCoAOMT(caffeoyl-CoA O-methyltransferase), CAD(cinnamyl-alcohol dehydrogenase), PER(peroxidase), CYP98A3(coumaroylquinate/coumaroylshikimate 3'-monooxygenase), CSE(caffeoylshikimate esterase), GOT1(aspartate aminotransferase), TAT(tyrosine aminotransferase), hisC(histidinol-phosphate aminotransferase) and AOC3(primary-amine oxidase).

https://doi.org/10.1371/journal.pone.0125722.g003

Phenylalanine ammonialyas (PAL, EC 4.3.1.24) is the first key enzyme in lignin biosynthesis pathway (Fig 3) which transforms phenylalanine into cinnamic acid by non-oxidative deamination [43, 44]. A total of 6,297 PALs reads were sequenced and 7 PAL transcripts were assembled in C. deserticola (Table 5). By sequence similarities comparison, we found that 4 of them (comp28550_c1_seq1/2/3/5) had more than 95% similarity with the known mRNA sequence of C. deserticola (gi|289595227|gb|ADD12041.1|), while comp28550_c1_seq4 and comp25940_c0_seq1 had 77% and 82% similarities, respectively. ORF prediction revealed 5 transcripts had potentials of encoding proteins and carried with aromatic amino acid lyase domain (PF00221.14). Among them, only the comp28550_c1_seq4 transcript could encode a complete protein sequence of 718 amino acid residues. It has been reported that PAL was encoded by a small multigene family in most plant species, such as 4 in Arabidopsis thaliana [39, 45], 5 in Populus trichocarpa [46, 47], 3 in Scutellaria baicalensis [48], and 7 Cucumis sativus [43, 49] etc. Our phylogenetic analysis suggested that there were 4 PAL encoding genes in C. deserticola and we named them CdPAL1, CdPAL2, CdPAL3 and CdPAL4, respectively (S2 Fig). 4-coumarate-CoA ligase (4CL, EC 6.2.1.12) and trans-cinnamate 4-monooxygenase (CYP73A, EC 1.14.13.11) are two enzymes responsible for transforming cinnamic acid to p-coumaroyl-CoA in two reverse orders. They are also in backbones, and their expression FPKM values are 39.57 and 51.93, respectively.

thumbnail
Table 5. Phenylalanine ammonia lyase (PAL) mRNA in de novo assembly transcriptome of C. deserticola.

https://doi.org/10.1371/journal.pone.0125722.t005

The four type lignins were biosynthesized by different pathways which were controlled by three key enzymes, cinnamoyl-CoA reductase (CCR, EC 1.2.1.44), shikimate o-hydroxycinnamoyltransferase (HCT, EC 2.3.1.133), and ferulate-5-hydroxylase (F5H, EC 1.14.-.-). CCR was reported as a control point of lignins pathway [50, 51] which catalyzed X-CoA (X including p-coumaroyl, caffeoyl, feruloyl, 5-hydroxyl-feruloyl and sinapoyl) into Y-aldehyde (Y including p-coumar, caffeyl, coniferyl, 5-hydroxyl-coniferyl and sinap), while HCT catalyzed p-coumaroyl-CoA to p-coumaroyl shikimic acid/p-coumaroyl quinic acid. The two enzymes, just like a switch, regulated biosynthesis of P-hydroxyl-phenyl lignins or the other three type lignins. F5H was another branch switch that regulated syringyl lignin and 5-hydroxyl-guaiacyl lignin. Other important enzymes including caffeic acid 3-O-methyltransferase (COMT, EC 2.1.1.68), caffeoyl-CoA O-methyltransferase (CCoAOMT, EC 2.1.1.104) and cinnamyl-alcohol dehydrogenase (CAD, EC 1.1.1.195) were also detected expressed. Detailed expression information was listed in Table 6. These enzyme genes identified in this study will provide a valuable resource for functional genomic studies in this important medicinal plant. 10 genes related to lignins biosynthesis pathway in Table 6 were selected for RT-qPCR verification to confirm our RNA-seq results (Fig 4), and their high correlations (Pearson correlation coefficient: 0.90343) indicated high accuracy and reproducibility of our transcriptome analysis. S1 Dataset lists the primer sequences used in this analysis.

thumbnail
Table 6. Expression profile of related genes in biosynthesis of lignin and PhGs.

https://doi.org/10.1371/journal.pone.0125722.t006

thumbnail
Fig 4. RT-qPCR analysis. Real-time PCR results for 10 selected genes related to lignin biosynthesis pathway.

Plot in the top panel of the figure presents RNA-seq (line graph) and qPCR (bar graph) quantification results, and the bottom panel shows their high correlations (r>0.9). Corresponding gene names and unique gene identifier numbers in the assembly are bglX (comp10579_c0), 4CL (comp26895_c0), CAD_1 (comp30193_c0), CAD_2 (comp30398_c0), PER_1 (comp21305_c0), PER_2 (comp25087_c0), PER_3 (comp30272_c0), CYP98A3 (comp26902_c8), CSE (comp19323_c0), and cCOMT (comp28020_c1). (Notes: as comp30193_c0 and comp30398_c0 were both annotated as CAD gene, we added subscripts to differentiate them. The same thing was done to comp21305_c0, comp25087_c0 and comp30272_c0, which were all annotated as PER gene).

https://doi.org/10.1371/journal.pone.0125722.g004

Candidate genes encoding enzymes involved in the biosynthesis of PhGs

Phenylethanoid glycosides (PhGs) are known to be the primary active ingredients in C. deserticola with activities of improving sexual potency, scavenging free radical and anti-aging [710]. Three chemical components of PhGs are organic acid, saccharide and phenylethanol aglycon (Fig 3). The organic acid including caffeic acid, ferulic acid and coumalic acid are products of phenylpropanoid biosynthesis pathway. The components of saccharide including glucose and rhamnose are products of carbohydrate metabolism pathways, such as starch and sucrose metabolism, amino sugar and nucleotide sugar metabolism, fructose and mannose metabolism etc. However, the biosynthesis pathway of phenylethanol part is not clear yet. Here, we proposed two possible phenylethanol biosynthesis pathways based on our sequence data. One is the reported caffeic acid or ferulic acid pathway, also known as cinnamiv acid pathway which is similar to the lignin biosynthesis backbone pathway. Another is based on phenylalanine metabolism pathway (Fig 3), in which the phenylalanine to phenylethanol was achieved by a known ‘Enrlich pathway’ that was first found in yeast one century ago and validated in petunia flowers [52, 53], tomato [54], and rose [55, 56]. Four enzyme genes encoding aspartate/tyrosine aminotransferase, histidinol-phosphate aminotransferase, and primary-amine oxidase which are responsible for the conversion of phenylalanine to phenylethanol were detected expressed in stem of C. deserticola. The product of phenylethanol may be further oxidized by monooxygenase or methylated by methytransferase into its derivates (phenylethanol aglycon) which took part in PhG biosynthesis. In summary, two putative biosynthesis pathways of phenylethanol aglycon were proposed for C. deserticola but still needs more study in further.

Discussions

Recent years, plant genomics have developed rapidly with the application of next-generation sequencing technology, while few researches have been focused on the genomics of desert medicinal plants [57]. It is urgently necessary to perform genomic or transcriptomic research to understand its adaption to drought and salinity environment and the biosynthesis pathway of the major bioactive components. The de novo transcriptome discovery for some medical plants, such as Panax ginseng [58], Ginkgo biloba [59] and Glycyrrhiza uralensis [60] have been first exploited using Roche 454 platform for its long read length. Because of the effectively assembly ability with short reads, especially the advantaged paired-end reads [23, 61], Illumina-based transcriptome sequencing and assembly has also been extensively used for model [20, 22, 6264] and non-model organisms [17, 19, 21, 65, 66]. In the present study, we generated about 8G of 101 bp paired-end reads and produced longer unigene sequences with 725 bp average length. Large-scale stem-specific transcriptome data could provide useful reference data and be used to mine secondary metabolism of bioactive components of C. deserticola. There is 81.62% of the total raw reads passed stringent quality filters (including adaptor trimming and low-quality reads discarding) before assembly, suggesting the high-quality of our sequencing data, and in which 82.08% of the high-quality reads were useful for assemble. Other reads failed to use for assemble may come from sequencing error [67], assembly parameters [23] et al. Those unused high-quality reads remained helpful to improve de novo assembly combined with longer reads from other platform (such as Roche 454) in the future.

A great number of assembled transcripts (30,098) showed high sequences similarities to know genes in public databases, suggesting that our Illumina-based paired-end data covered a substantial fraction of transcripts of C. deserticola. Transcripts with no BLAST hits may due to 3’ or 5’ untranslated regions, non-coding RNA [23] or new gene sequences of C. deserticola. The expressed transcripts were annotated to a wide range of GO categories and KEGG pathways (Tables 3 and 4), in which many transcripts were assigned to secondary metabolism related pathways. As we known, phenylpropanoid can function as inducible antimicrobial compounds with great salutary for an underground lifestyle [1], and also act as signal molecule in plant-microbe interactions besides for its medicinal utility [68, 69]. Terpenoid is used for biosynthesis of bioactive components (such as 6- deoxycatalpol) [70]. We found genes involved in the phenylpropanoid and terpenoid backbone biosynthesis pathway were highly abundant in C. deserticola. More importantly, the discovery of well represented pathways of lignin biosynthesis (Fig 3) indicated the active metabolic process of lignin in C. deserticola stem. All known enzyme genes involved in biosynthesis of lignin (Fig 3) were detected expressed, and four key enzymes including PAL, CCR, HCT and F5H had lower expression abundance (FPKM 26.47, 3.89, 3.4 and 3.83, respectively) compared with other enzyme genes (Table 6). Whether or not the expression change of those three genes could influence lignin production in C. deserticola is worthy of further study. PAL is a key enzyme in lignin biosynthesis and also involving in the biosynthesis of phenylpropanoid, resveratrol, flavonoid and coumarin [7174]. We detected four distinct PAL genes in C. deserticola genome (S2 Fig) which was coincident with that PAL was encoded by a small multigene family [39, 43, 4549] and further proved it may play important roles in metabolic carbon flux.

PhG is the primary active ingredients in C. deserticola. Genes involved in the biosynthesis of phenylethanol are important for the quality of C. deserticola. We deduced two different biosynthesis pathways of phenylethanol and 17 enzyme genes involved in PhG biosynthesis in C. deserticola stem. The possible post-caffeic/ferulic acid processes (Fig 3) were also deduced for the first time based on structural formula of intermediates and catalytic properties of corresponding enzymes, in which the caffeic/ferulic acid would be first oxidized into phenyl-pyruvate derivate; then, the carboxyl group was deprived by decarboxylases; finally, aldehyde group were converted back into alcohol group by dehydrogenase. This is the first application of Illumina paired-end sequencing technology to investigate the whole transcriptome of C. deserticola and to assemble RNA-seq reads without a reference genome. This study will provide useful resources and gene sequences for functional genomics and proteomics research on C. deserticola in future.

Conclusions

In this study, we profiled the transcriptome of C. deserticola stem based on high-throughput sequencing data, and identified genes involved in biosynthesis pathways of lignin and also inferred potential biosynthesis pathway of PhGs for the first time, which will certainly accelerate the understanding of the ambiguous physiological processes and the great medicinal value in molecular level. Up to now, this is the first attempt to de novo assemble the whole transcriptome of C. deserticola stem and to detect biosynthesis pathway of medicinal components using Illumina-based sequencing datasets. Our study may promote the development of natural medicines and the selection of cultivars with medicinal traits.

Supporting Information

S1 Fig. Correlation between RNA-seq expression data of two replicated C. deserticola samples.

Normalized read count of two replicates (2013-year sample and 2014-year sample) were plotted, and high correlation showed reproducibility of our RNA-seq data.

https://doi.org/10.1371/journal.pone.0125722.s001

(TIF)

S2 Fig. Phylogenetic analyses of PAL genes in C. deserticola.

Phylogenetic distances between known PAL gene and assembled potential PAL sequences in C. deserticola.

https://doi.org/10.1371/journal.pone.0125722.s002

(TIF)

S1 Dataset. All primer sequences for 10 selected genes related to lignins biosynthesis pathway.

https://doi.org/10.1371/journal.pone.0125722.s003

(XLSX)

S2 Dataset. All expressed transcripts with annotation in assembled C. deserticola stem transcriptome.

https://doi.org/10.1371/journal.pone.0125722.s004

(XLSX)

S3 Dataset. GO annotation of all expressed transcripts in C. deserticola stem transcriptome.

https://doi.org/10.1371/journal.pone.0125722.s005

(XLSX)

S4 Dataset. KEGG pathway annotation of all expressed transcripts with predicted ORF in C. deserticola stem transcriptome.

https://doi.org/10.1371/journal.pone.0125722.s006

(XLSX)

S5 Dataset. Significantly changed KEGG pathways between C. deserticola and rice (tissues of callus and leaf).

https://doi.org/10.1371/journal.pone.0125722.s007

(XLSX)

Acknowledgments

The authors would like to thank Dongmei Tian for assisting in sequencing data analysis.

Author Contributions

Conceived and designed the experiments: SS YL XW. Performed the experiments: YL XW TC MS GS. Analyzed the data: YL XW TC. Contributed reagents/materials/analysis tools: FY CL QT. Wrote the paper: YL XW SS. Revised the manuscript: SH JY.

References

  1. 1. Li X, Zhang TC, Qiao Q, Ren ZM, Zhao JY, Yonezawa T, et al. (2013) Complete Chloroplast Genome Sequence of Holoparasite Cistanche deserticola (Orobanchaceae) Reveals Gene Loss and Horizontal Gene Transfer from Its Host Haloxylon ammodendron (Chenopodiaceae). PLoS ONE 8: e58747. pmid:23554920
  2. 2. Cheng XY, Zhou HY, Cui X, Ni W and Liu CZ (2006) Improvement of phenylethanoid glycosides biosynthesis in Cistanche deserticola cell suspension cultures by chitosan elicitor. Journal of Biotechnology 121: 253–260. pmid:16143416
  3. 3. Song J, Feng GU, Tian CY and Zhang FS (2005) Strategies for Adaptation of Suaeda physophora, Haloxylon ammodendron and Haloxylon persicum to a Saline Environment During Seed-Germination Stage. Annals of Botany 96: 399–405. pmid:16002418
  4. 4. Tu PF, He YP and Lou ZC (1994) Herbalogical Studies on Rou Congrong (Herba Cistanchis). China Academic Journal 19: 1–5.
  5. 5. Tu PF, He YP and Lou ZC (1994) Survey and Protection of Medicinal Resources of Desertliving Cistanche (Cistanche deserticola). China Academic Journal 25: 205–208.
  6. 6. Cao R, Ma H and Wang YC (2004) Morphotype Diversity of Cistanche deserticola. China Journal of Chinese Materia Medica 29: 35–37. pmid:15709376
  7. 7. Wang T, Zhang XY and Xie WY (2012) Cistanche deserticola Y. C. Ma, "Desert Ginseng": A Review. The American Journal of Chinese Medicine 40: 1123–1141. pmid:23227786
  8. 8. Gu L, Xiong WT, Wang C, Sun HX, Li GF, Liu X. (2013) Cistanche deserticola decoction alleviates the testicular toxicity induced by hydroxyurea in male mice. Asian Journal of Andrology 15: 838–840. pmid:23817499
  9. 9. Sarsenbayev KN, Sarsenbayeva AB and Balmukanov KU (2013) Chemical Composition of Low-Molecular Weigth Organic Compounds (LMWOC) of Water Extracts from Cistanche Deserticola Stolones Depending on Treatment. World Applied Sciences Journal 25: 28–35.
  10. 10. Stefanova NA, Fursova AZ, Sarsenbaev KN and Kolosova NG (2011) Effects of Cistanche deserticola on behavior and signs of cataract and retinopathy in senescence-accelerated OXYS rats. Journal of Ethnopharmacology 138: 624–632. pmid:22020276
  11. 11. College JNM and Press SSaT (1977) Dictionary of Chinese Materia Medica [M]. 895–897.
  12. 12. Lu MC (1998) Studies on the sedative effect of Cistanche deserticola. Journal of Ethnopharmacology 59: 161–165. pmid:9507899
  13. 13. Wang XW, Wang XY, Wu LY and Wang XF (2001) Scavenging effect of glycosides of Cistanche deserticola on free radicals and its protection against OH-induced DNA damage in vitro. Chin Pharm J 36: 29–31.
  14. 14. Jiang Y and Tu PF (2009) Analysis of chemical constituents in Cistanchespecies. J Chromatogr A 1216: 1970–1979. pmid:18691718
  15. 15. Dong Q, Yao J, Fang JN and Ding K (2007) Structural characterization and immunological activity of two cold-water extractable polysaccharides from Cistanche deserticola Y. C. Ma. Carbohydrate Research 342: 1343–1349. pmid:17442280
  16. 16. Jimenez C and Riguera R (1994) Phenylethanoid glycosides in plants: structure and biological activity. Natural Product Reports 11: 591–606. pmid:15209134
  17. 17. Zhu L, Zhang Y, Guo W, Xu XJ and Wang Q (2014) De Novo Assembly and Characterization of Sophora japonica Transcriptome Using RNA-seq. BioMed Research International 2014: 9.
  18. 18. Clark M, Thorne M, Vieira F, Cardoso J, Power D, Peck L. (2010) Insights into shell deposition in the Antarctic bivalve Laternula elliptica: gene discovery in the mantle transcriptome using 454 pyroseqeuencing. BMC Genomics 11: 1–14. pmid:20044946
  19. 19. Zhai LL, Xu L, Wang Y, Cheng H, Chen YL, Gong YQ, et al. (2014) Novel and useful genic-SSR markers from de novo transcriptome sequencing of radish (Raphanus sativus L.). Molecular Breeding 33: 611–624.
  20. 20. Li PH, Ponnala L, Gandotra N, Wang L, Si YQ, Tausta S, et al. (2010) The developmental dynamics of the maize leaf transcriptome. Nat Genet 42: 1060–1067. pmid:21037569
  21. 21. Duan JL, Xia C, Zhao GY, Jia JZ and Kong XY (2012) Optimizing de novo common wheat transcriptome assembly using short-read RNA-Seq data. BMC Genomics 13: 1–12. pmid:22214261
  22. 22. Lu TT, Lu GJ, Fan DL, Zhu CR, Li W, Zhao Q, et al. (2010) Function annotation of the rice transcriptome at single-nucleotide resolution by RNA-seq. Genome Research 20: 1238–1249. pmid:20627892
  23. 23. Wang ZY, Fang BP, Chen JY, Zhang XJ, Luo ZX, Huang LF, et al. (2010) De novo assembly and characterization of root transcriptome using Illumina paired-end sequencing and development of cSSR markers in sweetpotato (Ipomoea batatas). BMC Genomics 11: 726. pmid:21182800
  24. 24. Varshney R, Nayak S, May G and Jackson S (2009) Next-generation sequencing technologies and their implications for crop genetics and breeding. Trends Biotechnol 27: 522–530. pmid:19679362
  25. 25. Haas BJ, Papanicolaou A, Yassour M, Grabherr M, Blood PD, Bowden J, et al. (2013) De novo transcript sequence reconstruction from RNA-seq using the Trinity platform for reference generation and analysis. Nat Protocols 8: 1494–1512. pmid:23845962
  26. 26. Mortazavi A, Williams B, McCue K, Schaeffer L and Wold B (2008) Mapping and quantifying mammalian transcriptomes by RNA-Seq. Nat Methods 5: 621–628. pmid:18516045
  27. 27. Dassanayake M, Haas JS, Bohnert HJ and Cheeseman JM (2009) Shedding light on an extremophile lifestyle through transcriptomics. New Phytologist 183: 764–775. pmid:19549131
  28. 28. Gayral P, Melo-Ferreira J, Glemin S, Bierne N, Carneiro M, Nabholz B, et al. (2013) Reference-Free Population Genomics from Next-Generation Transcriptome Data and the Vertebrate-Invertebrate Gap. PLoS Genet 9: e1003457. pmid:23593039
  29. 29. Alagna F, D'Agostino N, Torchia L, Servili M, Rao R, Pietrella M, et al. (2009) Comparative 454 pyrosequencing of transcripts from two olive genotypes during fruit development. BMC Genomics 10: 399. pmid:19709400
  30. 30. Grabherr MG, Haas BJ, Yassour M, Levin JZ, Thompson DA, Amit I, et al. (2011) Full-length transcriptome assembly from RNA-Seq data without a reference genome. Nat Biotechnol 29: 644–U130. pmid:21572440
  31. 31. Langmead B, Trapnell C, Pop M and Salzberg SL (2009) Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol 10: R25. pmid:19261174
  32. 32. Li B and Dewey CN (2011) RSEM: accurate transcript quantification from RNA-Seq data with or without a reference genome. Bmc Bioinformatics 12: 323. pmid:21816040
  33. 33. Finn RD, Bateman A, Clements J, Coggill P, Eberhardt RY, Eddy SR, et al. (2014) Pfam: the protein families database. Nucleic Acids Research 42: D222–D230. pmid:24288371
  34. 34. Moriya Y, Itoh M, Okuda S, Yoshizawa AC and Kanehisa M (2007) KAAS: an automatic genome annotation and pathway reconstruction server. Nucleic Acids Research 35: W182–W185. pmid:17526522
  35. 35. Li YL, Wang XL, Li CP, Hu SN, Yu J, Song SH. (2014) Transcriptome-wide N(6)-methyladenosine profiling of rice callus and leaf reveals the presence of tissue-specific competitors involved in selective mRNA modification. RNA biology 11: 1180–1188. pmid:25483034
  36. 36. Pouteau C, Dole P, Cathala B, Averous L and Boquillon N (2003) Antioxidant properties of lignin in polypropylene. Polymer Degradation and Stability 81: 9–18.
  37. 37. Vinardell MP, Ugartondo V and Mitjans M (2008) Potential applications of antioxidant lignins from different sources. Industrial Crops and Products 27: 220–223.
  38. 38. Guo DJ, Chen F, Inoue K, Blount JW and Dixon RA (2001) Downregulation of caffeic acid 3-O-methyltransferase and caffeoyl CoA 3-O-methyltransferase in transgenic alfalfa: impacts on lignin structure and implications for the biosynthesis of G and S lignin. The Plant Cell Online 13: 73–88. pmid:11158530
  39. 39. Raes J, Rohde A, Christensen JH, Van de Peer Y and Boerjan W (2003) Genome-wide characterization of the lignification toolbox in Arabidopsis. Plant Physiology 133: 1051–1071. pmid:14612585
  40. 40. Rogers LA, Dubos C, Cullis IF, Surman C, Poole M, Willment J, et al. (2005) Light, the circadian clock, and sugar perception in the control of lignin biosynthesis. Journal of Experimental Botany 56: 1651–1663. pmid:15878986
  41. 41. Marita JM, Ralph J, Lapierre C, Jouanin L and Boerjan W (2001) NMR characterization of lignins from transgenic poplars with suppressed caffeic acid O-methyltransferase activity. Journal of the Chemical Society, Perkin Transactions 1: 2939–2945.
  42. 42. Ralph J, Lapierre C, Marita JM, Kim H, Lu FC, Hatfield RD, et al. (2001) Elucidation of new structures in lignins of CAD- and COMT-deficient plants by NMR. Phytochemistry 57: 993–1003. pmid:11423146
  43. 43. Hou XM, Shao FJ, Ma YM and Lu SF (2013) The phenylalanine ammonia-lyase gene family in Salvia miltiorrhiza: genome-wide characterization, molecular cloning and expression analysis. Mol Biol Rep 40: 4301–4310. pmid:23644983
  44. 44. Bagal U, Leebens-Mack J, Lorenz W and Dean J (2012) The phenylalanine ammonia lyase (PAL) gene family shows a gymnosperm-specific lineage. BMC Genomics 13: 1–9. pmid:22214261
  45. 45. Wanner LA, Li GQ, Ware D, Somssich IE and Davis KR (1995) The phenylalanine ammonia-lyase gene family in arabidopsis thaliana. plant molecular biology 27: 327–338. pmid:7888622
  46. 46. Shi R, Yang CM, Lu SF, Sederoff R and Chiang V (2010) Specific down-regulation of PAL genes by artificial microRNAs in Populus trichocarpa. Planta 232: 1281–1288. pmid:20725738
  47. 47. Shi R, Sun YH, Li QZ, Heber S, Sederoff R, Chiang VL. (2010) Towards a Systems Approach for Lignin Biosynthesis in Populus trichocarpa: Transcript Abundance and Specificity of the Monolignol Biosynthetic Genes. Plant and Cell Physiology 51: 144–163. pmid:19996151
  48. 48. Xu H, Park NI, Li XH, Kim YK, Lee SY, Park SU. (2010) Molecular cloning and characterization of phenylalanine ammonia-lyase, cinnamate 4-hydroxylase and genes involved in flavone biosynthesis in Scutellaria baicalensis. Bioresource Technology 101: 9715–9722. pmid:20708926
  49. 49. Shang QM, Li L and Dong CJ (2012) Multiple tandem duplication of the phenylalanine ammonia-lyase genes in Cucumis sativus L. Planta 236: 1093–1105. pmid:22572777
  50. 50. Lacombe E, Hawkins S, Van Doorsselaere J, Piquemal J, Goffner D, Poeydomenge O, et al. (1997) Cinnamoyl CoA reductase, the first committed enzyme of the lignin branch biosynthetic pathway: cloning, expression and phylogenetic relationships. The Plant Journal 11: 429–441. pmid:9107033
  51. 51. Li LG, Cheng XF, Leshkevich J, Umezawa T, Harding SA, Chiang V. (2001) The Last Step of Syringyl Monolignol Biosynthesis in Angiosperms Is Regulated by a Novel Gene Encoding Sinapyl Alcohol Dehydrogenase. The Plant Cell Online 13: 1567–1586. pmid:11449052
  52. 52. Negre F, Kish CM, Boatright J, Underwood B, Shibuya K, Wagner C, et al. (2003) Regulation of methylbenzoate emission after pollination in snapdragon and petunia flowers. Plant Cell 15: 2992–3006. pmid:14630969
  53. 53. Verdonk JC, de Vos CHR, Verhoeven HA, Haring MA, van Tunen AJ, Schuurink RC. (2003) Regulation of floral scent production in petunia revealed by targeted metabolomics. Phytochemistry 62: 997–1008. pmid:12590126
  54. 54. Tieman D, Taylor M, Schauer N, Fernie AR, Hanson AD, Klee HJ. (2006) Tomato aromatic amino acid decarboxylases participate in synthesis of the flavor volatiles 2-phenylethanol and 2-phenylacetaldehyde. Proceedings of the National Academy of Sciences of the United States of America 103: 8287–8292. pmid:16698923
  55. 55. Watanabe S, Hayashi K, Yagi K, Asai T, MacTavish H, Picone J, et al. (2002) Biogenesis of 2-phenylethanol in rose flowers: incorporation of [2H8]L-phenylalanine into 2-phenylethanol and its beta-D-glucopyranoside during the flower opening of Rosa 'Hoh-Jun' and Rosa damascena Mill. Bioscience, biotechnology, and biochemistry 66: 943–947. pmid:12092844
  56. 56. Sakai M, Hirata H, Sayama H, Sekiguchi K, Itano H, Asai T, et al. (2007) Production of 2-phenylethanol in roses as the dominant floral scent compound from L-phenylalanine by two key enzymes, a PLP-Dependent decarboxylase and a phenylacetaldehyde reductase. Biosci Biotech Bioch 71: 2408–2419. pmid:17928708
  57. 57. Chen SL, Xiang L, Guo X and Li QS (2011) An introduction to the medicinal plant genome project. Front Med 5: 178–184. pmid:21695623
  58. 58. Chen S, Luo H, Li Y, Sun Y, Wu Q, Niu Y, et al. (2011) 454 EST analysis detects genes putatively involved in ginsenoside biosynthesis in Panax ginseng. Plant Cell Rep 30: 1593–1601. pmid:21484331
  59. 59. Lin XH, Zhang J, Li Y, Luo HM, Wu Q, Sun C, et al. (2011) Functional genomics of a living fossil tree, Ginkgo, based on next-generation sequencing technology. Physiologia Plantarum 143: 207–218. pmid:21834857
  60. 60. Li Y, Luo HM, Sun C, Song JY, Sun YZ, Wu Q, et al. (2010) EST analysis reveals putative genes involved in glycyrrhizin biosynthesis. BMC Genomics 11: 1–11. pmid:20044946
  61. 61. Maher C, Palanisamy N, Brenner J, Cao X, Kalyana-Sundaram S, Luo S, et al. (2009) Chimeric transcript discovery by paired-end transcriptome sequencing. Proc Natl Acad Sci USA 106: 12353–12358. pmid:19592507
  62. 62. Trick M, Long Y, Meng J and Bancroft I (2009) Single nucleotide polymorphism (SNP) discovery in the polyploid Brassica napus using Solexa transcriptome sequencing. Plant Biotechnol J 7: 334–346. pmid:19207216
  63. 63. Li RQ, Zhu HM, Ruan J, Qian WB, Fang XD, Shi ZB, et al. (2010) De novo assembly of human genomes with massively parallel short read sequencing. Genome Res 20: 265–272. pmid:20019144
  64. 64. Hegedus Z, Zakrzewska A, Agoston VC, Ordas A, Racz P, Mink M, et al. (2009) Deep sequencing of the zebrafish transcriptome response to mycobacterium infection. Molecular Immunology 46: 2918–2930. pmid:19631987
  65. 65. Li RQ, Fan W, Tian G, Zhu HM, He L, Cai J, et al. (2010) The sequence and de novo assembly of the giant panda genome. Nature 463: 311–317. pmid:20010809
  66. 66. Collins LJ, Biggs PJ, Voelckel C and Joly S (2008) An approach to transcriptome analysis of non-model organisms using short-read sequences. Genome Informatics 10: 3–14.
  67. 67. Meyer E, Aglyamova G, Wang S, Buchanan-Carter J, Abrego D, Colbourne JK, et al. (2009) Sequencing and de novo analysis of a coral larval transcriptome using 454 GSFlx. BMC Genomics 10: 219. pmid:19435504
  68. 68. Hao DC, Ma P, Mu J, Chen SL, Xiao PG, Peng Y, et al. (2012) De novo characterization of the root transcriptome of a traditional Chinese medicinal plant Polygonum cuspidatum. Sci China Life Sci 55: 452–466. pmid:22645089
  69. 69. Naoumkina MA, Zhao Q, Gallego-Giraldo L, Dai XB, Zhao PX, Dixon RA, et al. (2010) Genome-wide analysis of phenylpropanoid defence pathways. Molecular Plant Pathology 11: 829–846. pmid:21029326
  70. 70. Tu PF, He YP and Lou ZC (1997) Studies on the chemical constituents of the cultivated desertliving Cistanche (Cistanche deserticola). Nat Prod Res Dev 9: 452–502.
  71. 71. Halls C and Yu O (2008) Potential for metabolic engineering of resveratrol biosynthesis. Trends in Biotechnology 26: 77–81. pmid:18191264
  72. 72. Shi CY, Yang H, Wei CL, Yu O, Zhang ZZ, Jiang CJ, et al. (2011) Deep sequencing of the Camellia sinensis transcriptome revealed candidate genes for major metabolic pathways of tea-specific compounds. BMC Genomics 12: 131. pmid:21356090
  73. 73. Stalman M, Koskamp A-M, Luderer R, Vernooy JHJ, Wind JC, Wullems GJ, et al. (2003) Regulation of anthraquinone biosynthesis in cell cultures of Morinda citrifolia. Journal of Plant Physiology 160: 607–614. pmid:12872482
  74. 74. Luo HM, Li Y, Sun C, Wu Q, Song JY, Steinmetz A, et al. (2010) Comparison of 454-ESTs from Huperzia serrata and Phlegmariurus carinatus reveals putative genes involved in lycopodium alkaloid biosynthesis and developmental regulation. BMC Plant Biology 10: 209. pmid:20854695