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

Next-Generation Sequencing of the Chrysanthemum nankingense (Asteraceae) Transcriptome Permits Large-Scale Unigene Assembly and SSR Marker Discovery

  • Haibin Wang,

    Affiliation College of Horticulture, Nanjing Agricultural University, Nanjing, China

  • Jiafu Jiang,

    Affiliation College of Horticulture, Nanjing Agricultural University, Nanjing, China

  • Sumei Chen,

    Affiliation College of Horticulture, Nanjing Agricultural University, Nanjing, China

  • Xiangyu Qi,

    Affiliation College of Horticulture, Nanjing Agricultural University, Nanjing, China

  • Hui Peng,

    Affiliation College of Horticulture, Nanjing Agricultural University, Nanjing, China

  • Pirui Li,

    Affiliation College of Horticulture, Nanjing Agricultural University, Nanjing, China

  • Aiping Song,

    Affiliation College of Horticulture, Nanjing Agricultural University, Nanjing, China

  • Zhiyong Guan,

    Affiliation College of Horticulture, Nanjing Agricultural University, Nanjing, China

  • Weimin Fang,

    Affiliation College of Horticulture, Nanjing Agricultural University, Nanjing, China

  • Yuan Liao,

    Affiliation College of Horticulture, Nanjing Agricultural University, Nanjing, China

  • Fadi Chen

    chenfd@njau.edu.cn

    Affiliation College of Horticulture, Nanjing Agricultural University, Nanjing, China

Abstract

Background

Simple sequence repeats (SSRs) are ubiquitous in eukaryotic genomes. Chrysanthemum is one of the largest genera in the Asteraceae family. Only few Chrysanthemum expressed sequence tag (EST) sequences have been acquired to date, so the number of available EST-SSR markers is very low.

Methodology/Principal Findings

Illumina paired-end sequencing technology produced over 53 million sequencing reads from C. nankingense mRNA. The subsequent de novo assembly yielded 70,895 unigenes, of which 45,789 (64.59%) unigenes showed similarity to the sequences in NCBI database. Out of 45,789 sequences, 107 have hits to the Chrysanthemum Nr protein database; 679 and 277 sequences have hits to the database of Helianthus and Lactuca species, respectively. MISA software identified a large number of putative EST-SSRs, allowing 1,788 primer pairs to be designed from the de novo transcriptome sequence and a further 363 from archival EST sequence. Among 100 primer pairs randomly chosen, 81 markers have amplicons and 20 are polymorphic for genotypes analysis in Chrysanthemum. The results showed that most (but not all) of the assays were transferable across species and that they exposed a significant amount of allelic diversity.

Conclusions/Significance

SSR markers acquired by transcriptome sequencing are potentially useful for marker-assisted breeding and genetic analysis in the genus Chrysanthemum and its related genera.

Introduction

SSRs (simple sequence repeats) represent an informative class of genetic markers. They are distributed throughout the coding and non-coding regions of all eukaryotic genomes [1] and have been widely use to characterize genetic diversity, identify germplasm, construct linkage maps and tag genes for the purpose of marker-assisted breeding [2][4]. The development of an SSR assay requires DNA sequence to design the necessary PCR primers, and sufficient such sequence is as yet lacking in most species. The development of SSRs can be a costly and time-consuming endeavor [5], [6]. One approach taken to enhance the efficiency of the process has been to target EST (expressed sequence tag) sequence, which is acquired by sequencing a reverse-transcribed preparation of mRNA. SSR markers derived from EST sequence (“EST-SSRs”) are particularly attractive because their location within coding sequence enhances the probability of successful cross-species transferability [7][9]. Thus, for example, when a set of EST-SSRs was generated from peach cDNA, each was shown to be functional in six other Prunus species [10]. Some concern has been expressed that EST-SSRs may be less informative than those targeting non-coding DNA, but in sesame at least, EST-SSRs proved to be sufficiently informative [11]. EST-SSRs therefore can simultaneously circumvent cross-species constraints and be effective in exposing polymorphism.

Next-generation sequencing (NGS) has facilitated the study of gene expression, gene regulation and gene networks in both model and non-model organisms [12][14]. A large volume of sequence acquired in this way has already been deposited in GenBank. In particular, NGS has enhanced the value of EST libraries by expanding sequence read lengths. As a result, EST databases have become an increasingly valuable resource for SSR marker development [6].

The Asteraceae tribe Anthemideae Cass. includes the genus Chrysanthemum sensu lato. Chrysanthemum species form a polyploid series (2× to 10×) based on x = 9. The genus is one of the largest within the Asteraceae, comprising some 1,741 species [15]. Asteraceae genomes are typically very large and highly heterozygous, so it is unlikely that any of their full genomic sequences will be acquired in the near future. To date only a small number of Chrysanthemum ESTs (7,180 sequences deposited in GenBank as of September 2012) has been generated, and barely any SSR markers have been developed.

The diploid species C. nankingense (Nakai) Tzvel is a native of China and has a relatively small genome [16], [17]. The species is particularly well adapted to environments which are exposed to either extreme temperatures (both high and low), low soil fertility and/or drought. Young leaves are consumed as a vegetable, and the plant contains cancer antagonistic flavonoids and various aromatic oils [18]. Here, we describe our effort to characterize the C. nankingense transcriptome, based on the use of the Illumina paired-end sequencing platform. In conjunction with the EST sequences already deposited in GenBank, the newly acquired sequences were then used to develop EST-SSR markers which should find applications from linkage mapping to marker assisted breeding in the genus Chrysanthemum and other related genera.

Materials and Methods

Plant Material and RNA Extraction

The C. nankingense and other germplasm used are maintained by the Chrysanthemum Germplasm Resource Preserving Centre, Nanjing Agricultural University, China. For the RNA required for the transcriptome sequencing, stems and leaves were harvested from three 30 day old C. nankingense cuttings rooted on MS media and grown at a constant temperature of 25°C and a 16 h photoperiod (provided by cool white fluorescent lamps producing 36 µmol m−2 s−1).

A Total RNA Isolation System (Takara, Japan) was employed to extract RNA from the plant tissue, following the manufacturer’s instructions. The quality of the RNA (RNA Integrity Number (RIN)>8.5 and 28S:18S>1.5) was verified using a 2100 Bioanalyzer RNA Nanochip (Agilent, Santa Clara, CA) and its concentration ascertained using an ND-1000 Spectrophotometer (NanoDrop, Wilmington, DE). The standards applied were 1.8≤ OD260/280≤2.2 and OD260/230≥1.8. At least 20 µg of RNA was pooled in an equimolar fashion from each of the three sample plants.

cDNA Library Construction and Sequencing

Illumina (San Diego, CA) sequencing based on a GAII platform was performed at the Beijing Genomics Institute (Shenzhen, China; http://www.genomics.cn/index.php), following the manufacturer’s protocols. Briefly, beads coated with oligo (dT) were used to isolate poly (A) mRNA from the total RNA. A fragmentation buffer was added to interrupt the mRNA and thereby generate fragments in the size range 100–400 bp. The resulting fragments served as a template for the synthesis of the first strand cDNA, employing as primer random hexamers (N6). Second strand cDNA was synthesized using a SuperScript Double-Stranded cDNA Synthesis kit (Invitrogen, Camarillo, CA), after which it was purified using a QiaQuick PCR extraction kit (Qiagen, Hilden, Germany) and resolved with EB buffer for end reparation and poly (A) addition. The products were ligated with one another using sequencing adapters, and after agarose gel electrophoresis, a suitable size range of fragments were selected for PCR amplification. The resulting library was sequenced using an Illumina HiSeq™ 2000 device.

Data Filtering and de novo Assembly

Image data output from the sequencing device were transformed into raw reads and stored in FASTQ format. These data were filtered to remove raw reads that include adapter sequence or were of low quality. The assembly of the transcriptome was achieved using the short-read assembly program Trinity [19]. The unigenes are divided into either clusters or singletons. BLASTX [20] alignment (applying an E-value of less than 10−5) between each unigene sequence and those lodged in Nr (non-redundant protein database, NCBI), Nt (non-redundant nucleotide database, NCBI), Swiss-Prot, GO (gene ontology, http://www.geneontology.org/) and COG (clusters of orthologous groups) databases were performed, and the best alignments used to infer the unigene’s directionality. Where the outcome from the various databases conflicted with one another, the priority order applied was Nr, Swiss-Prot, COG. Where no alignment was possible, the software tool ESTScan [21] was used to assign directionality.

Gene Annotation and Analysis

Functional annotation was assigned using the protein (Nr and Swiss-Prot), COG and GO databases. BLASTX was employed to identify related sequences in the protein databases based on an E-value of less than 10−5. The COG database is an attempt to classify proteins from completely sequenced genomes on the basis of the orthology concept [22]. GO’s aim is to standardize the representation of genes and their products, by insisting on a controlled vocabulary and a strictly defined concept [23], [24]. The annotations acquired from Nr were processed through the Blast2GO program [25] to obtain the relevant GO terms, and these were then analyzed by WEGO software [26] to assign a GO functional classification and to illustrate the distribution of gene functions.

Development of cDNA-derived SSR Markers and Primer Design

Unigenes containing putative SSRs were identified by MISA software (http://pgrc.ipk-gatersleben.de/misa/), applying the following parameters: a minimum of six repeats for dinucleotide motifs, of five for trinucleotides, of four for tetranucleotides and of three for penta- and hexa-nucleotides. Primer 3.0 software (http://sourceforge.net/projects/primer3) was used to design appropriate primers from the flanking sequence, based on the following criteria: primer length 18–22 bp (optimally 20 bp), Tm of 50–60°C (no more than a 4°C difference between the Tms of the forward and reverse primers) and an amplicon length in the range 100–400 bp. Primers were synthesized by Invitrogen (Shanghai, China).

EST-SSR Screening and Polymorphism Survey

Genomic DNA was isolated from leaves following published protocols [27]. The DNA was dissolved in 50 µl water, diluted to a final concentration of 10 ng/µl and stored at −20°C until required. Genetic diversity was analyzed by tailing the forward SSR primer with a fluorescently labeled M13 sequence (CACGACGTTGTAAAACGAC) [28]. Agarose gel electrophoresis separation of the amplicons produced at 100 randomly chosen SSR loci from C. nankingense DNA template was used to check the functionality of the primers. A set of 20 of the 81 functional primer pairs was then used to amplify the DNA of 42 commercial chrysanthemum varieties and 30 related wild species (Table 1). The PCR conditions comprised an initial denaturing step (95°C/3 min), followed by ten cycles of 94°C/30 s, 50–60°C/30 s, 72°C/30 s, then 25 cycles of 94°C/30 s, 55°C/30 s, 72°C/30 s, and finally by an elongation step (72°C/7 min). The amplicons were separated using an ABI3730xl (Applied Biosystems) device, following the manufacturer’s protocols. Each sample was supplemented with ABI GeneScan LIZ500 size standard in order to determine amplicon lengths with the aid of GeneMapper® v3.7 software (Applied Biosystems). For the statistical analysis, the patterns at all SSR loci were scored for each polymorphic band as 1 for band presence and 0 for band absence. This allowed an estimate at each locus of the number of alleles present (NA) and the PIC value. Similarity coefficients based on SSR profiles were calculated according to Nei and Li [29]. Genetic relationships between materials were examined using cluster analysis implemented in the NTSYS-pc2.1 software package and a dendrogram was constructed based on the unweighted pair group method of arithmetic average algorithm. Some representative PCR products were also separated through 6% denaturing polyacrylamide gels (19∶1 acrylamide: bisacrylamide, 7.5 M urea, 1 × Tris-borate-EDTA pH 7.8) and visualized by silver staining.

thumbnail
Table 1. SSR genotypes of species within the genera Chrysanthemum, Leucanthemella, Ajania, Artemisia, Opisthopappus, Pyrethrum, Tanacetum and Crossostephium.

https://doi.org/10.1371/journal.pone.0062293.t001

Results

Sequencing and de novo Assembly

In all, 53,720,166 sequence reads were generated, of which 51,622,828 were of acceptable quality (Table 2). The de novo assembly yielded 120,260 contigs of mean length 321 bp (Fig. 1A), and these were resolved into 70,895 unigenes, of which 26,650 were clusters and 44,245 singletons (of length at least 150 bp). The range in unigene length was from 150 bp to 9,032 bp (mean 585 bp) (Fig. 1B).

thumbnail
Figure 1. The distribution of contig and unigene sequence lengths.

https://doi.org/10.1371/journal.pone.0062293.g001

Structural and Functional Annotation

A total of 45,789 of the unigene sequences shared some similarity to known genes. The three most frequently encountered transcripts encoded a light-regulated protein, a catalase and an arabinogalactan peptide. Just 107 of the unigenes showed sequence homology with the Chrysanthemum spp. sequences lodged in the Nr database, and the number of hits with sunflower and lettuce (also Asteraceae species) sequences was, respectively, 679 and 277. The ranges in E-value and sequence similarity of the top hits in the Nr database were comparable, with 33.4% (E-value of 0 to −60) and 19.3% (100%–80%) of the sequences possessing homology (Fig. 2A, B). On a species basis, the highest proportion of matching sequences in the Nr database were derived from grapevine (47.7%), followed by soybean (16.5%) (Fig. 2C).

thumbnail
Figure 2. The distribution of E-value, sequence similarity, and species of data base hits with the C. nankingense unigenes.

https://doi.org/10.1371/journal.pone.0062293.g002

The COG analysis allowed the functional classification of 21,952 of the unigenes (Fig. 3). The most frequently identified classes were “general function” (4,503, 16.3%), followed by “transcription” (2,622, 9.5%), “replication, recombination and repair” (2,370, 8.6%), “post-translational modification, protein turnover and chaperones” (2,189, 7.9%), “signal transduction” (1,898, 6.9%) and “translation, ribosomal structure and biogenesis” (1,714, 6.2%). Nr annotation assigned 36,985 of the unigenes to the category of “biological process”; within this ontology, the two most common functions were “cellular process” (8,639, 22.6%) and “metabolic process” (8,502, 23.0%). At the level of localization, “cell” applied to 14,402 unigenes (34.8%), “cell part” to 13,003 (31.4%) and “organelle” to 9,765 (23.6%). Although the function of the unigenes covered a comprehensive range of GO categories, “catalytic activity” (10,724, 47.5%) and “binding” (9,738, 43.1%) proteins made up the majority, while “transporter activity”, “molecular transducer activity”, “enzyme regulator activity”, “receptor activity”, “antioxidant activity” and “protein binding transcription factor activity” proteins together made up only 9.4% (Fig. 4).

thumbnail
Figure 3. Functional classification of the C. nankingense unigenes according to COG criteria.

https://doi.org/10.1371/journal.pone.0062293.g003

thumbnail
Figure 4. The distribution of C. nankingense unigenes among the GO functional classes.

From left to right, “Biological process” (shown in blue): biological adhesion, biological regulation, cell killing, cellular component organization or biogenesis, cellular process, death, developmental process, establishment of localization, growth, immune system process, localization, locomotion, metabolic process, multi-organism process, multicellular organismal process, negative regulation of biological process, pigmentation, positive regulation of biological process, regulation of biological process, reproduction, reproductive process, response to stimulus, rhythmic process, signaling, viral reproduction; “Cellular components” (shown in red): cell, cell junction, cell part, extracellular region, extracellular region part, macromolecular complex, membrane-enclosed lumen, organelle, organelle part, virion; “Molecular function” (shown in green): antioxidant activity, binding, catalytic activity, enzyme regulator activity, molecular transducer activity, protein binding transcription factor activity, receptor activity, transporter activity.

https://doi.org/10.1371/journal.pone.0062293.g004

Development and Characterization of cDNA-derived SSR Markers

MISA analysis identified 2,813 putative microsatellites, equivalent to one locus per 14.7 kb of the C. nankingense transcriptome. Based on the length of the repeat motif, the sequences were into two groups: Class I were hypervariable markers, consisted of SSRs ≥20 bp; Class II, or potentially variable markers were consisted of SSRs 12–20 bp [30], [31]. In all, 1,788 primer pairs were designed (Table S1), of which 342 (19.1%) targeted Class I loci and the remaining ones Class II loci. Almost all the sequences (95.6%) shared high homology to known genes. A further 501 putative SSRs were located among the EST sequences lodged in GenBank; from these, 363 PCR primer pairs were designed (Table S1). The most abundant repeat motifs were trinucleotides (1,914, 57.8%), followed by dinucleotides (747, 22.5%) and tetranucleotides (243, 7.3%). Over 160 motifs were identified, of which the most frequent were AC/TG (8.3%), CA/GT (181, 5.4%), CCA/GGT (181, 5.4%), ACC/TGG (162, 4.9%), CAA/GTT (131, 4.1%) and ATC/TAG (3.9%) (Fig. 5).

thumbnail
Figure 5. Frequencies of the various repeat motifs present in the C. nankingense EST-SSRs.

https://doi.org/10.1371/journal.pone.0062293.g005

Polymorphism Survey with EST-SSR

All 20 primer pairs (targeting 16 Class I and four Class II loci) have amplicons in 42 chrysanthemum cultivated varieties and could be used for marker-assisted breeding in chrysanthemum (Fig. 6). However, not all of them amplified the DNA of all 30 wild relatives. For example, even though most wild materials displayed the polymorphism, we could not detect any amplicon bands using primer pair 170 in Artemisia absinthium. Nearly all of the templates - but not those of several of the Artemisia spp. - amplified a product using primer pair #254 (Table 1). The phylogeny of the wild species based on their EST-SSR genotype using 20 of the primer pairs (Fig. 7) illustrates that they fall into six clades. The set of Chrysanthemum and Ajania spp., along with Leucanthemella linearis, formed one clade, the Artemisia spp. a second clade, leaving the Opisthopappus, Pyrethrum, Tanacetum and Crossostephium spp. separated both from one another and from each of the two major clades.

thumbnail
Figure 6. SSR allelic variation among 42 commercial C. morifolium cultivars.

(A) Single product, non-polymorphic amplicon(primer #214), (B) Two product, polymorphic amplicon (primer #86), (C) Multiple product, highly polymorphic amplicon (primer #245). Lanes 1–42: cv. ‘Zhongshanzigui’, cv. ‘Zhongshanjingui’, cv. ‘Zhongshanhongying’, cv. ‘Zhongshanaihuang’, cv. ‘Zhongshanzihe’, cv. ‘Zhongshanchengguang’, cv. ‘Zhongshanhongfeng’, cv. ‘Zhongshanhuangtuogui’, cv. ‘Zhongshanguohuang’, cv. ‘Zhongshanbailu’, cv. ‘Zhongshanfenzhuang’, cv. ‘Zhongshanzilian’, cv. Zhongshanfendai’, cv. ‘Zhongshanqueyu’, cv. ‘Zhongshanhongxia’, cv. ‘Zhongshanhuangying’, cv. ‘Zhongshanhuangyu’, cv. ‘Zhongshanjinyu’, cv. ‘Zhongshanyinxing’, cv. ‘Zhongshanzaobai’, cv. ‘Zhongshanzixing’, cv. ‘Zhongshanziyu’, cv. ‘Zhongshanfengui’, cv. ‘Zhongshanjinzhen’, cv. ‘Zhongshanxuegui’, cv. ‘Zhongshanzuirong’, cv. ‘Jinlingaihuang’, cv. ‘Jinlingbaifeng’, cv. ‘Jinlingbaixue’, cv. ‘Jinlingbaiyu’, cv. ‘Jinlingbaohui’, cv. ‘Jinlingbaoxia’, cv. ‘Jinlingchixin’, cv. ‘Jinlingchunse’, cv. ‘Jinlingfanxing’, cv. ‘Jinlingfendai’, cv. ‘Jinlingguofen’, cv. ‘Jinlingguohong’, cv. ‘Jinlingguohuang’, cv. ‘Jinlingguozi’, cv. ‘Jinlinghonghe’, cv. ‘Jinlinghongpao’.

https://doi.org/10.1371/journal.pone.0062293.g006

thumbnail
Figure 7. UPGMA-based phylogeny of Chrysanthemum spp. and species belonging to closely related genera.

The tree was derived from genotype data from 20 EST-SSR loci. (A–F) Six clades were recognized.

https://doi.org/10.1371/journal.pone.0062293.g007

Discussion

Sequencing, Assembly and Annotation

NGS has been widely used for de novo transcriptome sequencing, especially in non-model organisms [32][37]. With respect to Asteraceae species, the representation of ESTs comprises sequences from lettuce, sunflower and gerbera. The current C. nankingense data consisting of 70,895 unigene sequences is comparable in size to the 81,330 lettuce EST collection, and larger than that of either Centaurea solstitialis (40,407) or Zinnia violacea (20,767) (http://www.ncbi.nlm.nih.gov/dbEST/dbEST_summary.html). It is also noteworthy that 44,245 of the C. nankingense unigenes matched genes with unique annotations, a proportion which is comparable with what has been experienced in diploid plant species, e. g., rice, Arabidopsis thaliana and Brachypodium distachyon [38], [39]. Unigene sequences have a number of applications, such as the identification of single nucleotide polymorphism markers [40], [41], alternative gene splicing products [42,42], homologous genes [43][45] and gene families [46], [47]. Over 64% of the products of the unigenes shared homology with known proteins, a proportion which is rather higher than in gerbera, where only 42% of unigene sequences satisfied this criterion [48]. About 80.8% of the annotated unigenes could be assigned to the category “biological process”, which indicates that a wide diversity of transcripts is represented in the data set. Thus this sequence resource should have utility in future efforts to either identify novel genes or to undertake genome-wide transcriptomic analyses in chrysanthemum.

Identification and Characterization of SSR Marker Sequences

The development of informative SSRs remains a labor intensive process [49], [50]. Mining the gerbera sequence for functional SSR markers only realised 99 assays [51], although 730 potentially informative assays were generated for the Asteraceae species Cichorium intybus [52]. Here we report the development of 1,788 and 363 EST-SSRs based on present transcriptome and previously published ESTs in GenBank, respectively. To our knowledge, this represents the first reported mass isolation of EST-SSRs in Chrysanthemum. NGS clearly offers a rapid means of acquiring the sequence needed to discover SSRs and to design the necessary primers to form an assay. EST-SSR markers are of particular interest, as allelic variants can entail alterations in the coding sequence which could underlie specific phenotypic variation [53], [54]. EST-SSR markers also tend to have a higher amplification efficiency and are more likely to be transferable across species than is the case for SSR markers derived from non-coding regions of the genome [55]. The frequency of EST-SSR loci and their ease of PCR amplification were consistent with the experience of other plant species [4], [24], and it was relatively easy to generate a substantial number of informative markers of use both within cultivated chrysanthemum germplasm and among its wild relatives. The overall results showed that nearly all plant materials have genetic polymorphism, suggesting that the regions flanking the repeats are highly conserved.

Several lines of evidence suggest that the genera Leucanthemella, Ajania and Chrysanthemum are closely related to one another, but that Artemisia is rather distant [56], [57]. The phylogenetic relationships between these genera inferred from ETS-SSR genotype were largely consistent with this picture. Opisthopappus is a recently described Compositae genus from China; it differs from Tanacetum in a number of morphological ways [58]. Meanwhile, Crossostephium has been defined as a small separate genus, distinct from Pyrethrum [59]. The EST-SSR phylogeny confirmed that Opisthopappus, Pyrethrum, Tanacetum and Crossostephium are independent genera, and was congruent with phylogenies based on variation in the nuclear ribosomal ITS and chloroplast trnL-F IGS sequences [60]. All those data validated the universality of the 20 randomly selected EST-SSRs.

Conclusions

The Asteraceae species form a large and poorly understood group of plants, which nevertheless contains a number of horticulturally significant species. Nearly all the SSR assays developed were able to distinguish the cultivated C. morifolium germplasm from the wild relative species. The markers described here should find application in a range of research activity, including the assessment of genetic diversity, the identification of germplasm, the construction of linkage maps and gene tagging for marker-assisted breeding. Any genomic variable that affects genetic function may have an evolutionary role more or less. Some experimental evidence has suggested that the rapid evolution of SSR sequences - specifically, the gain or loss of repeats at a locus - may provide a molecular basis for adaptation to unfamiliar environments [61], [62]. Whether or not this holds for the Asteraceae will require further research.

Supporting Information

Table S1.

Microsatellite markers developed from the de novo transcriptome sequence of C. nankingense.

https://doi.org/10.1371/journal.pone.0062293.s001

(XLS)

Table S2.

Microsatellite markers developed from the NCBI dbEST database.

https://doi.org/10.1371/journal.pone.0062293.s002

(XLS)

Author Contributions

Conceived and designed the experiments: HW FC JJ SC. Performed the experiments: HW FC JJ XQ. Analyzed the data: HW HP PL AS. Contributed reagents/materials/analysis tools: YL ZG WF. Wrote the paper: HW JJ SC.

References

  1. 1. Levinson G, Gutman GA (1987) Slipped-strand mispairing: a major mechanism for DNA sequence evolution. Mol Biol Evol 4: 203–221.
  2. 2. Tang S, Kishore VK, Knapp SJ (2003) PCR-multiplexes for a genome-wide framework of simple sequence repeat marker loci in cultivated sunflower. Theor Appl Genet 107: 6–19.
  3. 3. Neeraja CN, Maghirang-Rodriguez R, Pamplona A, Heuer S, Collard BC, et al. (2007) A marker-assisted backcross approach for developing submergence-tolerant rice cultivars. Theor Appl Genet 115: 767–776.
  4. 4. Kaur S, Cogan NO, Pembleton LW, Shinozuka M, Savin KW, et al. (2011) Transcriptome sequencing of lentil based on second-generation technology permits large-scale unigene assembly and SSR marker discovery. BMC Genomics 12: 265.
  5. 5. Squirrell J, Hollingsworth PM, Woodhead M, Russell J, Lowe AJ, et al. (2003) How much effort is required to isolate nuclear microsatellites from plants? Mol Ecol 12: 1339–1348.
  6. 6. Chapman MA, Hvala J, Strever J, Matvienko M, Kozik A, et al. (2009) Development, polymorphism, and cross-taxon utility of EST-SSR markers from safflower (Carthamus tinctorius L.). Theor Appl Genet 120: 85–91.
  7. 7. Zwenger SR, Alsaggaf R, Basu C (2010) Does an expressed sequence tag (EST) library of Salsola iberica (tumbleweed) help to understand plant responses to environmental stresses? Plant Signal Behav 5: 1330–1335.
  8. 8. Sadd BM, Kube M, Klages S, Reinhardt R, Schmid-Hempel P (2010) Analysis of a normalised expressed sequence tag (EST) library from a key pollinator, the bumblebee Bombus terrestris. BMC Genomics 11: 110.
  9. 9. Lindqvist C, Scheen AC, Yoo MJ, Grey P, Oppenheimer DG, et al. (2006) An expressed sequence tag (EST) library from developing fruits of an Hawaiian endemic mint (Stenogyne rugosa, Lamiaceae): characterization and microsatellite markers. BMC Plant Biol 6: 16.
  10. 10. Vendramin E, Dettori MT, Giovinazzi J, Micali S, Quarta R, et al. (2007) A set of EST-SSRs isolated from peach fruit transcriptome and their transportability across Prunus species. Mol Ecol Notes 7: 307–310.
  11. 11. Wei WL, Qi XQ, Wang LH, Zhang YX, Hua W, et al. (2011) Characterization of the sesame (Sesamum indicum L.) global transcriptome using Illumina paired-end sequencing and development of EST-SSR markers. BMC Genomics 12: 451.
  12. 12. Parchman TL, Geist KS, Grahnen JA, Benkman CW, Buerkle CA (2010) Transcriptome sequencing in an ecologically important tree species: assembly, annotation, and marker discovery. BMC Genomics 11: 180.
  13. 13. Meyer E, Aglyamova GV, Wang S, Buchanan-Carter J, Abrego D, et al. (2009) Sequencing and de novo analysis of a coral larval transcriptome using 454 GSFlx. BMC Genomics 10: 219.
  14. 14. Sun C, Li Y, Wu Q, Luo H, Sun Y, et al. (2010) De novo sequencing and analysis of the American ginseng root transcriptome using a GS FLX Titanium platform to discover putative genes involved in ginsenoside biosynthesis. BMC Genomics 11: 262.
  15. 15. Bremer K, Humphries CJ (1993) Generic monograph of the Asteraceae-Anthemideae. Bulletin of the Natural History Museum Botany series 23: 71–177.
  16. 16. Yang W, Glover BJ, Rao GY, Yang J (2006) Molecular evidence for multiple polyploidization and lineage recombination in the Chrysanthemum indicum polyploid complex (Asteraceae). New Phytol 171: 875–886.
  17. 17. Zhao H, Liu Z, Hu X, Yin J, Li W, et al. (2009) Chrysanthemum genetic resources and related genera of Chrysanthemum collected in China. Genetic resources and crop evolution 56: 937–946.
  18. 18. Yang N, Ren A, Hu W, Qian S, Duan J, et al. (2005) Chemical constituents of Dendranthema nankingense. JOURNAL-CHINA PHARMACEUTICAL UNIVERSITY 36: 402.
  19. 19. Grabherr MG, Haas BJ, Yassour M, Levin JZ, Thompson DA, et al. (2011) Full-length transcriptome assembly from RNA-Seq data without a reference genome. Nat Biotechnol 29: 644–652.
  20. 20. Cameron M, Williams HE, Cannane A (2004) Improved gapped alignment in BLAST. IEEE/ACM Trans Comput Biol Bioinform 1: 116–129.
  21. 21. Iseli C, Jongeneel CV, Bucher P (1999) ESTScan: a program for detecting, evaluating, and reconstructing potential coding regions in EST sequences. Proc Int Conf Intell Syst Mol Biol: 138–148.
  22. 22. Tatusov RL, Galperin MY, Natale DA, Koonin EV (2000) The COG database: a tool for genome-scale analysis of protein functions and evolution. Nucleic Acids Res 28: 33–36.
  23. 23. Xia Z, Xu H, Zhai J, Li D, Luo H, et al. (2011) RNA-Seq analysis and de novo transcriptome assembly of Hevea brasiliensis. Plant Mol Biol 77: 299–308.
  24. 24. Wang Z, Fang B, Chen J, Zhang X, Luo Z, et al. (2010) De novo assembly and characterization of root transcriptome using Illumina paired-end sequencing and development of cSSR markers in sweet potato (Ipomoea batatas). BMC Genomics 11: 726.
  25. 25. Conesa A, Gotz S, Garcia-Gomez JM, Terol J, Talon M, et al. (2005) Blast2GO: a universal tool for annotation, visualization and analysis in functional genomics research. Bioinformatics 21: 3674–3676.
  26. 26. Ye J, Fang L, Zheng H, Zhang Y, Chen J, et al. (2006) WEGO: a web tool for plotting GO annotations. Nucleic Acids Res 34: W293–297.
  27. 27. Attitalla IH (2011) Modified CTAB method for high quality genomic DNA extraction from medicinal plants. Pak J Biol Sci 14: 998–999.
  28. 28. Schuelke M (2000) An economic method for the fluorescent labeling of PCR fragments. Nat Biotechnol 18: 233–234.
  29. 29. Nei M, Li WH (1979) Mathematical model for studying genetic variation in terms of restriction endonucleases. Proc Natl Acad Sci 76: 5269–5273.
  30. 30. Temnykh S, DeClerck G, Lukashova A, Lipovich L, Cartinhour S, et al. (2001) Computational and experimental analysis of microsatellites in rice (Oryza sativa L.): frequency, length variation, transposon associations, and genetic marker potential. Genome Res 11: 1441–1452.
  31. 31. Cho YG, Ishii T, Temnykh S, Chen X, Lipovich L, et al. (2000) Diversity of microsatellites derived from genomic libraries and GenBank sequences in rice (Oryza sativa L.). Theor Appl Genet 100: 713–722.
  32. 32. Metzker ML (2010) Sequencing technologies - the next generation. Nat Rev Genet 11: 31–46.
  33. 33. Glenn TC (2011) Field guide to next-generation DNA sequencers. Mol Ecol Resour 11: 759–769.
  34. 34. Rendon-Anaya M, Delaye L, Possani LD, Herrera-Estrella A (2012) Global Transcriptome Analysis of the Scorpion Centruroides noxius: New Toxin Families and Evolutionary Insights from an Ancestral Scorpion Species. PLoS One 7: e43331.
  35. 35. Zhang XM, Zhao L, Larson-Rabin Z, Li DZ, Guo ZH (2012) De Novo Sequencing and Characterization of the Floral Transcriptome of Dendrocalamus latiflorus (Poaceae: Bambusoideae). PLoS One 7: e42082.
  36. 36. Xiao L, Zhang L, Yang G, Zhu H, He Y (2012) Transcriptome of protoplasts reprogrammed into stem cells in Physcomitrella patens. PLoS One 7: e35961.
  37. 37. Loman NJ, Misra RV, Dallman TJ, Constantinidou C, Gharbia SE, et al. (2012) Performance comparison of benchtop high-throughput sequencing platforms. Nat Biotechnol 30: 434–439.
  38. 38. Vogel JP, Garvin DF, Mockler TC, Schmutz J, Rokhsar D, et al. (2010) Genome sequencing and analysis of the model grass Brachypodium distachyon. Nature 463: 763–768.
  39. 39. Bevan M, Walsh S (2005) The Arabidopsis genome: A foundation for plant research. Genome Res 15: 1632–1642.
  40. 40. Trick M, Long Y, Meng J, Bancroft I (2009) Single nucleotide polymorphism (SNP) discovery in the polyploid Brassica napus using Solexa transcriptome sequencing. Plant Biotechnol J 7: 334–346.
  41. 41. Bundock PC, Eliott FG, Ablett G, Benson AD, Casu RE, et al. (2009) Targeted single nucleotide polymorphism (SNP) discovery in a highly polyploid plant species using 454 sequencing. Plant Biotechnol J 7: 347–354.
  42. 42. Lalonde E, Ha KC, Wang Z, Bemmo A, Kleinman CL, et al. (2011) RNA sequencing reveals the role of splicing polymorphisms in regulating human gene expression. Genome Res 21: 545–554.
  43. 43. Sultan M, Schulz MH, Richard H, Magen A, Klingenhoff A, et al. (2008) A global view of gene activity and alternative splicing by deep sequencing of the human transcriptome. Science 321: 956–960.
  44. 44. Patel RV, Nahal HK, Breit R, Provart NJ (2012) BAR expressolog identification: expression profile similarity ranking of homologous genes in plant species. Plant J: 1038–1050.
  45. 45. Pavesi G, Mereghetti P, Zambelli F, Stefani M, Mauri G, et al. (2006) MoD Tools: regulatory motif discovery in nucleotide sequences from co-regulated or homologous genes. Nucleic Acids Res 34: W566–570.
  46. 46. Strozycki PM, Szymanski M, Szczurek A, Barciszewski J, Figlerowicz M (2010) A new family of ferritin genes from Lupinus luteus–comparative analysis of plant ferritins, their gene structure, and evolution. Mol Biol Evol 27: 91–101.
  47. 47. Fourquin C, Ferrandiz C (2012) Functional analyses of AGAMOUS family members in Nicotiana benthamiana clarify the evolution of early and late roles of C-function genes in eudicots. Plant J: 990–1001.
  48. 48. Laitinen RA, Immanen J, Auvinen P, Rudd S, Alatalo E, et al. (2005) Analysis of the floral transcriptome uncovers new regulators of organ determination and gene families related to flower organ differentiation in Gerbera hybrida (Asteraceae). Genome Res 15: 475–486.
  49. 49. Sonah H, Deshmukh RK, Sharma A, Singh VP, Gupta DK, et al. (2011) Genome-wide distribution and organization of microsatellites in plants: an insight into marker development in Brachypodium. PLoS One 6: e21298.
  50. 50. Szulkin M, David P (2011) Negative heterozygosity-fitness correlations observed with microsatellites located in functional areas of the genome. Mol Ecol 20: 3949–3952.
  51. 51. Gong L, Deng ZA (2010) EST-SSR markers for gerbera (Gerbera hybrida). Mol Breeding 26: 125–132.
  52. 52. Cadalen T, Morchen M, Blassiau C, Clabaut A, Scheer I, et al. (2010) Development of SSR markers and construction of a consensus genetic map for chicory (Cichorium intybus L.). Mol Breeding 25: 699–722.
  53. 53. Coetzee G, Irvine R (2002) Size of the androgen receptor CAG repeat and prostate cancer: does it matter? J Clin Oncol 20: 3572–3573.
  54. 54. Field D, Wills C (1996) Long, polymorphic microsatellites in simple organisms. Proc Biol Sci 263: 209–215.
  55. 55. Varshney RK, Graner A, Sorrells ME (2005) Genic microsatellite markers in plants: features and applications. Trends Biotechnol 23: 48–55.
  56. 56. Bremer K, Humphries C (1993) The generic monograph of the Asteraceae–Anthemideae. Bull Br Museum Bot Ser 23: 71–177.
  57. 57. Kondo K, Abd El-Twab MH (2002) Analysis of inter- and intrsgeneric relationships sensu stricto among the members of Chrysanthemum sensu lato by using fluorescent in situ hybridization. Chromosome Sci 6: 87–100.
  58. 58. Shih C (1979a) Opisthopappus Shih-a new genus of Compositae from China. Acta Phytotaxon Sin 17: 110–112.
  59. 59. Oberprieler C, Vogt R, Watson LE (2006) Tribe Anthemideae Cass. In: The families and genera of vascular plants VIII. Flowering plants. Eudicots. Asterales. Springer, Heidelberg, 342–374.
  60. 60. Zhao HB, Chen FD, Chen SM, Wu GS, Guo WM (2010) Molecular phylogeny of Chrysanthemum, Ajania and its allies (Anthemideae, Asteraceae) as inferred from nuclear ribosomal ITS and chloroplast trnL-F IGS sequences. Plant Syst Evol 284: 153–169.
  61. 61. Trifonov EN (2004) The tuning function of tandemly repeating sequences: a molecular device for fast adaptation. Evolutionary Theory and Processes: Modern Horizons, Kluwer Academic Publishers: 115–138.
  62. 62. King DG, Soller M, Kashi Y (1997) Evolutionary tuning knobs. Endeavour 21: 36–40.