Next Article in Journal
Synthesis and Bioactivity Evaluation of N-Arylsulfonylindole Analogs Bearing a Rhodanine Moiety as Antibacterial Agents
Previous Article in Journal
Modifications of Porphyrins and Hydroporphyrins for Their Solubilization in Aqueous Media
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Complete Chloroplast Genome Sequences of Fritillaria ussuriensis Maxim. and Fritillaria cirrhosa D. Don, and Comparative Analysis with Other Fritillaria Species

1
K-Herb Research Center, Korea Institute of Oriental Medicine, Daejeon 305-811, Korea
2
Department of Agronomy, Yanbian University Agriculture College, Yanji 133002, China
*
Author to whom correspondence should be addressed.
Molecules 2017, 22(6), 982; https://doi.org/10.3390/molecules22060982
Submission received: 6 March 2017 / Revised: 2 June 2017 / Accepted: 10 June 2017 / Published: 13 June 2017
(This article belongs to the Section Molecular Diversity)

Abstract

:
The genus Fritillaria belongs to the widely distributed Liliaceae. The bulbs of Fritillaria, F. ussuriensis and F. cirrhosa are valuable herbaceous medicinal ingredients. However, they are still used indiscriminately in herbal medicine. Identification and molecular phylogenic analysis of Fritillaria species are therefore required. Here, we report the complete chloroplast (CP) genome sequences of F. ussuriensis and F. cirrhosa. The two Fritillaria CP genomes were 151,524 and 151,083 bp in length, respectively, and each included a pair of inverted repeated regions (52,678 and 52,156 bp) that was separated by a large single copy region (81,732 and 81,390 bp), and a small single copy region (17,114 and 17,537 bp). A total of 111 genes in F. ussuriensis and 112 in F. cirrhosa comprised 77 protein-coding regions in F. ussuriensis and 78 in F. cirrhosa, 30 transfer RNA (tRNA) genes, and four ribosomal RNA (rRNA) genes. The gene order, content, and orientation of the two Fritillaria CP genomes exhibited the general structure of flowering plants, and were similar to those of other Fritillaria species. Comparison of the six Fritillaria species’ CP genomes indicated seven highly divergent regions in intergenic spacers and in the matK, rpoC1, rpoC2, ycf1, ycf2, ndhD, and ndhF coding regions. We established the position of the six species through phylogenic analysis. The complete chloroplast genome sequences of the two Fritillaria species and a comparison study are useful genomic information for identifying and for studying the phylogenetic relationship among Fritillaria species within the Liliaceae.

1. Introduction

The genus Fritillaria belongs to the Liliaceae, which consists of 140 known species. The bulbs of several Fritillaria species (called “Pae-mo” in Korean and “Bei-mu” in Chinese) are important ingredients in herbal drugs that are used in oriental medicine, and have great economic value in Asian countries. Fritillaria species are distributed in temperate regions of the Northern hemisphere [1]. Fritillaria (F.) cirrhosa is mainly distributed in the alpine regions of Northwestern China (Gansu, Qinghai, Sichuan, Xizang, and Yunnan provinces) at altitudes of 3200–4600 m, whereas F. ussuriensis is distributed in the lowland regions (0–500 m altitude) of the northeastern and the far-eastern areas of China, Russia, and Korea [2]. Fritillaria ussuriensis and F. cirrhosa display morphological differences in their floral bud, bract, stem, and capsule [2]. The stem length of F. ussuriensis is 50–60 cm (maximum: 100 cm); the stem length of F. cirrhosa is less than 60 cm. Fritillaria cirrhosa has three bracts per flower, whereas F. ussuriensis has two. The tapals are yellow or yellowish green in F. cirrhosa; the petals are purple in F. ussuriensis. The capsules of F. cirrhosa are narrowly winged; the capsules of F. ussuriensis are wingless. While it is difficult to classify the plants according to the morphological characteristics of their bulbs, the flowering plants can be easily distinguished. The bulbs are divided into two and five types in the Korean Pharmacopoeia and the Chinese Pharmacopoeia, respectively [3,4]. The dried bulbs of Fritillaria ussuriensis Maxim. and Fritillaria cirrhosa D. Don are used in different herbal medicines—namely, Fritillariae Ussuriensis Bulbus (Ping-bei-mu in Chinese) and Fritillariae Cirrhosae Bulbus (Chuan-bei-mu in Chinese), respectively. Fritillariae Cirrhosae Bulbus and Fritillariae Ussuriensis Bulbus have separately been used for clinical purposes in traditional Korean medicine. In detail, the former has been used to treat cough due to deficiency of the lung, asthenia of the viscera, and tidal fever, and the latter has mainly been used to treat cough due to exogenous dryness and deficiency of Um (Yin; Chinese traditional medicine technical term). Furthermore, Fritillariae Cirrhosae Bulbus and Fritillariae Ussuriensis Bulbus have been prescribed to treat pulmonary carbuncle and bloody sputum, respectively [5]. Although the bulbs of Fritillaria have value in herbal medicine, different Fritillaria species are still used indiscriminately because of their morphological similarity and similar names [6]. Therefore, accurate identification of Fritillaria species (e.g., using molecular markers) is required to identify medicinal plants and the drugs derived from them [6,7].
Chloroplasts play an important role in photosynthesis and carbon fixation as well as in the biosynthesis of starch, fatty acids, and amino acids [8,9]. The chloroplast (CP) genome ranges from 120 to 180 kb in higher plants and has a quadripartite structure consisting of a large single copy (LSC) region, a small single copy (SSC) region, and two copies of a larger inverted repeat (IR) [10,11]. The CP genome encodes 110 to 130 genes with up to 80 unique protein-coding genes, four ribosomal RNAs (rRNAs), and approximately 30 transfer RNAs (tRNAs). However, few parasitic plants have small chloroplast genomes due to their unique life cycles [12]. Since the CP genome of Marchantia polymorpha [13] was reported in 1986, more than 500 complete chloroplast genome sequences have been deposited in the GenBank database [14]. With the advancement of next-generation sequencing (NGS) technology, chloroplast genome assembly has become cheaper and easier compared to the Sanger method [15]. Through the comparison of chloroplast genomes, the development of molecular markers has also become more cost effective. The CP genome has been widely used for understanding phylogenetic relationships and discovering useful molecular markers, which are used in DNA barcoding to identify plant species and authenticating and in identifying herbal medicines. In particular, matK and rbcL are used as universal plant DNA barcodes [16].
Here, we report the de novo assembly of F. ussuriensis and F. cirrhosa CP genomes using the Illumina platform. This is the first comparative analysis of Fritillaria CP genomes in conjunction with previously reported CP genomes. This study aims to investigate global structural patterns of six Fritillaria CP genomes and also to discover highly divergent regions among the species. We also analyzed phylogenic relationships among the six Fritillaria species. The results provide basic knowledge on characteristics of Fritillaria species and enhance our understanding of Fritillaria species evolution within the Liliaceae.

2. Materials and Methods

2.1. Genome Sequencing and Assembly

Fresh leaves of F. ussuriensis (KY646166) and F. cirrhosa (KY646167) were collected from medicinal plant plantations, and the samples were used for CP genome sequencing. Fritillaria ussuriensis and F. cirrhosa were given identification numbers, and specimens were registered in the Korean Herbarium of Standard Herbal Resources (Index Herbariorum code KIOM) at the Korea Institute of Oriental Medicine (KIOM). The extraction of DNA was conducted with the DNeasy Plant Maxi kit (Qiagen, Valencia, CA, USA), according to the manufacturer’s instructions. Illumina short-insert paired-end sequencing libraries were constructed and generated using the NextSeq platform (Illumina, San Diego, Valencia, CA, USA) by LabGenomics, Korea. The CP genomes were obtained by the de novo assembly method from low-coverage whole-genome sequence derived from the Phyzen pipeline [17]. Trimmed paired-end reads (Phred scores ≥20) were assembled using CLC genome assembler (ver. 4.06 beta, CLC Inc, Rarhus, Denmark) with default parameters. SOAPdenovo gap closer was used to fill gaps based on alignment paired-end reads [18]. The principal contigs representing the CP genome were retrieved from the total contigs using Nucmer [19], and aligned contigs were ordered with the CP genome sequence of Fritillaria hupehensis (NC024736) [20].

2.2. Genome Annotation and Comparative Analysis

Gene annotation of F. ussuriensis and F. cirrhosa CP genomes was performed using DOGMA annotation [21], and manually corrected for codons and gene boundaries using BLAST searches. The tRNAs were confirmed with tRNAscan-SE 1.21 [22]. The circular maps of the two Fritillaria CP genomes were obtained using OGDRAW [23]. GC content and codons were analyzed using MEGA6 software [24]. The mVISTA program was used to compare the seven Fritillaria CP genomes using the F. ussuriensis CP genome as reference [25]. Five Fritillaria CP genomes were downloaded from GenBank (F. hupenesis: NC024736, F. tapaiensis: NC023247, F. unibracteata: KF769142, and F. thunbergii: KY646165).

2.3. Repeat Analysis

SSRs in F. ussuriensis and F. cirrhosa CP genomes were detected using MISA [26] with the parameters set to minimum number of repeats: 10, 5, 4, 3, 3, and 3 for mono-, di-, tri-, tetra-, penta-, and hexa-nucleotides, respectively. The tandem repeats were 20 bp or more with minimum alignment score and maximum period size set at 50 and 500, respectively, and the identity of repeats was set to ≥90% [27]. IRs were detected using Inverted Repeat Finder with default parameters. The IRs were 20 bp or more with 90% similarity [28].

2.4. Phylogenic and Divergence Analysis

A molecular phylogenetic tree was constructed using 74 protein-coding genes from 11 species. Among these 11 taxa, nine completed CP genomes were downloaded from NCBI: Fritillaria unibracteata (KF769142), Fritillaria taipaiensis (NC023247), Fritillaria hupehensis (NC024736), Fritillaria thunbergii (KY646165), Lilium superbum (NC026787), Lilium longiflorum (KC968977), Smilax china (HM536959), Acorus gramineus (NC026299), and Cocos nucifera (KX028884). A total of 71 protein-coding genes were aligned with MAFFT [29]. Maximum likelihood (ML) and maximum parsimony (MP) analyses were performed using MEGA6 with 1000 bootstrap replicates [24]. Six Fritillaria species CP genomes were aligned using MAFFT, and the sequences were manually adjusted with Bioedit [30]. To calculate nucleotide variability (Pi) between CP genomes, sliding window analysis was performed using DnaSP version 5.1 software [31]. Window length was set to 600 bp, and the step size was 200 bp.

3. Results

3.1. Chloroplast Genome Organization of Two Fritillaria Species

Illumina sequencing generated 5.0 and 4.5 Gb of trimmed paired-end reads from Fritillaria ussuriensis and Fritillaria cirrhosa, respectively. From the de novo assembly sequence that uses low-coverage whole-genome sequencing (WGS), we obtained ten and eight contigs covering the whole chloroplast genome sequences of F. ussuriensis and F. cirrhosa, respectively (Table S1). Single circular sequences were completed after gap filling and manual editing. The complete circular chloroplast genomes of F. ussuriensis and F. cirrhosa were 151,524 and 151,083 bp, with approximately 256× and 452× coverages, respectively (Table S2). Paired-end read mapping was conducted to validate the draft genome, which was compared to our draft genomes and the previously reported F. hupehensis genome using BLASTZ program (Figure S1). Both F. ussuriensis and F. cirrhosa chloroplast genomes had a quadripartite structure similar to most land plants consisting of a pair of IRs (52,678 and 52,156 bp, respectively). In addition, both F. ussuriensis and F. cirrhosa chloroplast genomes were also separated by large single copy (LSC; 81,732 and 81,390 bp) and small single copy (SSC; 17,114 and 17,537 bp) regions (Figure 1, Table 1). The Fritillaria chloroplast genomes were AT-rich (63% in both species), but the LSC (65.3% and 64.2% in F. ussuriensis and F. cirrhosa, respectively) and SSC (69.4% and 69.6% in F. ussuriensis and F. cirrhosa, respectively) regions were more AT-rich than the IR regions (57.6% and 57.4% in F. ussuriensis and F. cirrhosa, respectively), making the LSC and SSC regions more similar to other chloroplast genomes [10,32,33,34,35].
The gene content, order, and orientation were similar in Fritillaria CP genomes. There were 111 and 112 predicted genes in F. ussuriensis and F. cirrhosa, respectively. Of these, 94 in F. ussuriensis and 95 in F. cirrhosa were unique to the LSC and SSC regions, and 18 were duplicated in the IR regions (Table 1 and Table 2). The 111 and 112 unique genes consisted of 77 and 78 protein-coding regions in F. ussuriensis and F. cirrhosa, respectively. In the upstream region of the F. ussuriensis CP genome, one gene, cemA, had an internal stop codon (TGA). The gene cemA encodes a heme-binding protein in the chloroplast envelope membrane, which was lost in land plants due to the introduction of an internal stop codon. [8,10,36]. The two Fritillaria CP genomes had 30 tRNAs, with 17 duplicated genes including seven tRNAs (trnA-UGC, trnI-CAU, trnI-GAU, trnL-CAA, trnN-GUU, trnR-ACG, trnV-GAC), four rRNAs (rrn16, rrn23, rrn4.5, rrn5), and six protein-coding genes (ndhB, rpl2, rpl23, rps12, ycf1, ycf2). They also had 18 intron-containing genes, among which 15 (nine protein-coding genes and six tRNA genes) had a single intron and two genes (ycf3 and clpP) had two introns each (Table S4). Thirteen genes (nine protein-coding and four tRNA genes) were located in the LSC region, a protein-coding gene in the SSC region, and four genes (two protein-coding and two tRNA genes) in the IR regions. The protein-coding genes included five genes (ndhB, rpl2, rpl23, rps12, ycf2) that were duplicated in the IR regions. The rps12 gene was trans-spliced because the 5′ end was located in the LSC region and the 3′ end in the IR region. The trnK-UUU gene had the largest intron region (2613 bp in F. ussuriensis and 2562 bp in F. cirrhosa), including matK. The genes psbT, rpl2, and ndhD had the alternative start codon ACG, and rps19 started with GTG. Use of ACG and GTG as start codons are common features of a variety of genes in the chloroplast genomes of land plants [37,38,39,40].
Approximately 52% of Fritillaria chloroplast genomes consisted of protein-coding genes (78,951 bp in F. ussuriensis and 79,835 bp in F. cirrhosa), 1.9% of tRNAs (2876 bp in both species), and 6.0% of rRNAs (9048 bp in both species). The remaining 40.1% consisted of intergenic regions, non-coding introns, and pseudogenes. The ycf1 gene located between the inverted repeat b (IRb) and the small single copy (SSC) region had premature stop codons in the coding sequence, and has been annotated as a pseudogene in other angiosperm chloroplast genomes [32,41,42,43]. The codon usage and anticodon recognition patterns of the CP genomes are summarized in Table S5. Protein-coding genes comprised 26,317 codons in F. ussuriensis and 26,611 in F. cirrhosa. Among these codons, those for leucine and isoleucine were the most common in both Fritillaria genomes, as observed previously in other land plant CP genomes (Figure 2) [10,44]. The 30 tRNA genes included codons for all 20 amino acids required for biosynthesis. Within protein-coding regions, the AT content for the first, second, and third codons were 55.1%, 62.1%, and 70.4% in F. ussuriensis, respectively, and 59.3%, 63.1%, and 65.6% in F. cirrhosa, respectively (Table S3). The bias towards a higher AT content at the third position has been observed in other land plant CP genomes [45,46,47].

3.2. Repeat Analysis in Two Fritillaria Chloroplast Genomes

Simple sequence repeats (SSRs) or microsatellites are tandem repeat sequences consisting of 1–6 nt sequence motifs in prokaryotic and eukaryotic genomes [48,49]. SSRs were detected using a microsatellite identification tool—MISA—in both Fritillaria CP genomes. We detected 183 and 178 SSRs in F. ussuriensis and F. cirrhosa CP genomes, respectively. Mononucleotide motifs were the most abundant type of repeat, and di-nucleotides were the second most abundant in both of the Fritillaria CP genomes (Figure 3). Almost all SSR loci were composed of A or T, which contributed to the bias in base composition (A/T; 63%) in both Fritillaria CP genomes. Simple sequence repeats (SSRs) were more abundant in non-coding regions than in coding regions, as evidenced by the presence of 63% of all SSRs in the non-coding regions of both genomes. Furthermore, most SSRs were located in the LSC region. In addition, 48 polymorphic SSRs were detected between the two Fritillaria CP genomes (Table S8). Most SSRs were located in the intergenic region with the A or the T motif. The longest polymorphic SSR was found in the psbZ-trnG region, and had a 12 bp length difference.
We also identified 15 tandem repeats in F. ussuriensis and 13 tandem repeats in F. cirrhosa of more than 20 bp (Table S6). Of these, most were located in IGS, LSC, and IR regions. The longest tandem repeats were 108 bp in F. ussuriensis (located in the trnT-UGU/trnL-UAA IGS) and 94 bp in F. cirrhosa (located in the trnG-UCC/trnR-UCU IGS). Four (two in IGS, two in CDS) tandem repeats represented the same region in both CP genomes. Four and six palindromic repeats were also detected in F. ussuriensis and F. cirrhosa, respectively (Table S7). In F. ussuriensis, two of these were located in the LSC region and the other two were located in the IR regions. Both species had palindromic repeats at four locations—namely, the IGS of accD/psaI, petD/rpoA, ccsA/ndhD, and rps15/ycf1 regions.

3.3. Comparison of the Chloroplast Genomes with Those of Other Fritillaria Species

The two Fritillaria chloroplast genomes had approximately 98% sequence identity, and their gene content and order and genome structure were similar. The CP genome of F. ussuriensis was approximately 441 bp longer than F. cirrhosa (Table 1). The LSC and IR regions of F. ussuriensis were 342 bp and 522 bp longer, respectively, than F. cirrhosa. The SSC region of F. ussuriensis was 423 bp shorter than F. cirrhosa. IR contraction and expansion are common evolutionary events and contribute to genome size variation [10,32,50]. We analyzed the border structure of F. ussuriensis and F. cirrhosa genomes. Detailed comparison of the LSC, SSC, and IR regions are shown in Figure 4. The rps19 gene located in the LSC region extended into the IRb region by 27–46 bp. The border between IRb/SSC and SSC/IRa extended into the ycf1 genes if all Fritillaria species except for F. taipaiensis. Overlaps of 17–33 bp were observed between the ycf1 pseudogene and ndhF gene, except in F. taipaiensis. The trnH genes were all located in the IR region, 149–174 bp away from the IRa/LSC boundary. The locations of most of the other genes (e.g., ndhF and trnH) were similar in both CP genomes (Figure 4).
We performed multiple sequence alignments between six Fritillaria chloroplast genomes using mVISTA (Figure 5). The non-coding regions were more divergent than the coding regions. The most divergent regions were found in IGSs such as matK/trnK-UUU, trnK-UUU/rps16, rps16/trnQ-UUG, psbK/psbI, atpH/atpI, psbM/trnD-GUC, and ycf4/petD. For the coding regions, the most divergent regions included matK, rpoC1, rpoC2, ycf1, ycf2, ndhD, and ndhF. Previous studies reported similar divergent regions [32]. These regions are conserved regions with clusters of sequence variability called hotspots, containing single-nucleotide polymorphisms (SNPs) and indels [51,52]. The nucleotide variability—Pi—was calculated to show divergence at the sequence level of Fritillaria CP genomes. As expected, the IR regions were more conserved than the LSC and SSC regions. Between F. ussuriensis and F. cirrhosa CP genomes, Pi values (%) ranged from 0 to 15.8% with a mean of 0.9%. Ten highly divergent loci included matK, atpI, trnY-GUA, trnE-UUC, trnT, ycf3, rps4, ycf4, petA, rpl16, rps3, rps19, ccsA, ndhD, and ycf1 (Figure 6). The most divergent region in the ycf4-petA (IGS) region showed 15.8% sequence variability. In the coding region, rpl16 showed the highest degree of nucleotide variability (4.3%). The ten loci had a much higher sequence divergence value than the other regions (Pi > 2.7%). Among the six Fritillaria CP genomes, Pi values varied from 0 to 2.1% with a mean of 0.4% (Figure 6). The loci matK, atpI, trnY, trnE, trnT, rps19, ccsA, ndhD, and ycf1 were highly divergent among Fritillaria species. The highest degree of variability was found in ndhD (2%).

3.4. Phylogenic Analysis

Chloroplast genome sequences have been successfully used in numerous phylogenetic studies of angiosperms [53,54]. To identify the phylogenetic position of the six Fritillaria species within the Liliaceae, 74 protein-coding sequences shared by 11 CP genomes were aligned (Figure 7). Two species, Acorus gramineus and Cocos nucifera, were set as outgroups. The alignment covered 80,532 bp. Maximum likelihood (ML) and maximum parsimony (MP) analyses revealed that six out of eight nodes had 100% bootstrap values. Both the ML and MP phylogenetic results strongly indicated that the genera—Fritillaria and Lilium—were clearly separated, and the six Fritillaria species were closely related within Liliaceae. Fritillaria cirrhosa and F. unibracteata formed a cluster. Subsequently, a monophyletic clade formed a cluster with F. taipaiensis, F. hupehensis, and F. thunbergii, which were related to F. ussuriensis as a monophyletic branch.

4. Discussion

Advances in NGS technologies make it possible to complete the entire chloroplast genome with the discovery of molecular markers [55,56]. We used low-coverage WGS data to obtain chloroplast genomes—an approach that has been successfully used in several other studies [17,57,58]. This approach requires less time and has a lower cost than the previously used method. Here, we obtained two Fritillaria chloroplast genomes and applied comparative analysis to six Fritillaria species CP genomes. The two Fritillaria CP genomes contained a pair of IRs, and LSC and SSC regions. The two Fritillaria genomes have similar genome structures, gene orders, and gene contents, including introns and base compositions. They show characteristics that are typical of land plant CP genomes [8,32].
SSRs are suitable molecular markers because they are distributed throughout the whole genome and display high polymorphism between species, locus-specific co-dominance, and high transferability [59]. SSRs play an important role in CP genome rearrangement during evolution [41,60]. The repeat units—A and T—appear most frequently in SSRs, which contribute to the AT-richness of the CP genome [61,62,63]. Recently, plastid genomic information obtained by NGS analysis has provided a resource for high-throughput screening of SSR loci and permitted efficient SSR marker development in various plant taxa [64,65]. In this study, we identified 183 and 178 SSR loci in F. ussuriensis and F. cirrhosa CP genomes, respectively (Figure 3). In addition, we also identified 48 potential polymorphic SSRs between the two Fritillaria species, F. cirrhosa and F. ussuriensis (Table S8). These abundant SSR loci and potential inter-species specific polymorphic SSRs could provide useful genetic information and sequence resources for further molecular genetic studies of Fritillaria species, including species identification, assessment of genetic diversity, ecological genetic studies, and evolution studies [66].
IR regions are the most conserved regions in the chloroplast genome [9]. The contraction and expansion at the borders of IR regions, however, are common evolutionary events, and they are the major causes of rearrangements and size variations [10,34,50]. In this study, we compared the SC/IR boundaries among six Fritillaria genomes. The SC/IR boundaries showed only slight differences, for example, at ycf1 and ѱycf1 (Figure 4). This phenomenon is relatively common in other CP genomes [35,41,67]. Several genes, including ndhF, rps19, and trnH, had almost identical locations and sizes among Fritillaria. Multiple sequences of alignment between the six Fritillaria genomes indicated that the IR regions were more conserved than the LSC and SSC regions due to copy corrections by gene conversion in the IR regions [68]. The most divergent regions were found in the IGS, which have been used in phylogenetic studies [11,69]. For the coding regions, the most divergent regions included matK, rpoC1, rpoC2, ycf1, ycf2, ndhD, and ndhF. Previous studies reported similar divergent regions [9]. These regions are conserved regions of clustered variation called hotspots, containing SNPs and indels [51,52]. The rpoC2, rpoC1, and ycf1 genes are known to contain hotspots of sequence variation [11,51,52,70]. Therefore, as in other land plants, Fritillaria CP genomes contain general hotspot regions for genetic variation.
Chloroplast genomes provide an abundant genomic resource for phylogenetic analysis, and many studies have used protein-coding sequences or whole chloroplast genome sequences in these analyses [32,69,71]. In the present study, we conducted phylogenic analysis on the six Fritillaria species and found that the six species were well clustered according to the type of herbal medicine (Figure 7). Furthermore, F. ussuriensis could be clearly separated from other Fritillaria species, as previously reported [20,72]. In this study, phylogenetic analysis revealed clear positons of Fritillaria at the species level with 100% bootstrap values. Fritillaria hupehensis and F. thunbergii formed monophyly with 100% bootstrap values. Another monophyletic group clustered F. cirrhosa and F. unibracteata with F. taipaiensis. Fritillaria ussuriensis showed a distant genetic relationship to the other five Fritillaria species. In a previous study, a phylogenetic tree was constructed to verify the relationship between ninety-two Fritillaria species using a combined chloroplast region consisting of matK, rbcL, and rpl16. Ninety-two Fritillaria species were classified into eight Fritillaria subgenera while considering their geographic distributions [1]. Fritillaria thunbergii, F. cirrhosa, and F. ussuriensis were well classified into the Fritillaria subgenus—consistent with our results. Although the phylogenetic tree comprised only a few Fritillaria species, our results are well supported by the phylogenetic relationships of Fritillaria species reported previously [1]. Thus, phylogenetic analysis of CP genome sequences could provide useful information for uncovering relationships among Fritillaria species—in particular, details of their positions at the species level with 100% bootstrap values. However, the results suggested that additional large-scale genomic analyses using numerous accurately identified Fritillaria species are required to clarify the taxonomy and phylogenetic relationships of Fritillaria species at low taxonomic levels.
Molecular genetic tools using complete CP genome sequences provide an efficient and accurate way to authenticate herbal medicines. Universal DNA barcoding is widely used as a reliable genetic tool for identifying plant species; however, these methods are still limited in their ability to identify and discriminate medicinal plants at the species level for several taxa. The complete CP genome sequences could be useful for identifying plant species solely as a super DNA barcode, which provide high resolution at lower taxonomic levels [73]. In addition, the two complete Fritillaria CP genome sequences determined in this study could provide useful basic genomic information along with the previously reported Fritillaria species for studying phylogeny. The sequence variability from comparative analyses could be used as potential resources for further development of DNA molecular markers such as SSR or microsatellite marker. These additional developments could be used to study evolutionary properties and/or population genetics of Fritillaria species as well as DNA barcodes or SCAR markers.

Supplementary Materials

Supplementary Materials are available online.

Acknowledgments

We thank the “Classification and Identification Committee of the KIOM” for the identification of plant materials and the Herbarium of Korea Standard Herbal Resources (Index Herbarium code KIOM) for the provision of plant materials. This work was supported by a grant of the Development of Foundational Techniques for the Domestic Production of Authentic Herbal Medicines based on the Establishment of Molecular Authentication System (K17403) from the Korea Institute of Oriental Medicine (KIOM). The grant was funded by the Ministry of Science, ICT, and Future Planning (MSIP) of the Republic of Korea.

Author Contributions

I.P. performed experiments, carried out sequence analysis, and drafted the manuscript. W.J.K. and S.-M.Y. performed experiments. G.C. and Y.-M.K. and R.P. collected and identified plant materials. B.C.M. designed experiments, collected plant materials, and assembled sequencing results. All authors contributed to the experiments and approved the final manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

CPChloroplast
LSCLarge Single Copy
SSCSmall Single Copy
IRInverted Repeat
tRNATransfer RNA
rRNARibosomal RNA
NGSNext-Generation Sequencing
KIOMKorea Institute of Oriental Medicine
MLMaximum Likelihood
MPMaximum Parsimony
WGSWhole-Genome Sequencing
IRbInverted Repeat b
SSRsSimple Sequence Repeats
SNPsSingle-Nucleotide Polymorphisms

References

  1. Day, P.D.; Berger, M.; Hill, L.; Fay, M.F.; Leitch, A.R.; Leitch, I.J.; Kelly, L.J. Evolutionary relationships in the medicinally important genus Fritillaria L. (liliaceae). Mol. Phylogenet. Evol. 2014, 80, 11–19. [Google Scholar] [CrossRef] [PubMed]
  2. Xingi, C.; Mordak, H. Tulipa L. In Flora China Flagellariaceae Marantaceae; Science Press: Beijing, China, 2000; pp. 123–136. [Google Scholar]
  3. The State Pharmacopoeia Commission of the People’s Republic of China. Pharmacopoeia of the People’s Republic of China, 2010 ed.; Chemical Industry Press: Beijing, China, 2010; Volume 1, p. 191. [Google Scholar]
  4. Korea Institute of Oriental Medicine (KIOM). Defining Dictionary for Medicinal Herbs [Korean, ‘hanyak giwon sajeon’]. 2016. Available online: http://boncho.kiom.re.kr/codex/ (accessed on 15 March 2017).
  5. Lee, S.; Ju, Y. Identification of 11 species of paemo through each original plant and medicines. Korea J. Herbol. 2014, 29, 133–140. [Google Scholar] [CrossRef]
  6. Xiang, L.; Su, Y.; Li, X.; Xue, G.; Wang, Q.; Shi, J.; Wang, L.; Chen, S. Identification of Fritillariae bulbus from adulterants using ITS2 regions. Plant Gene 2016, 7, 42–49. [Google Scholar] [CrossRef]
  7. Xin, G.Z.; Lam, Y.C.; Maiwulanjiang, M.; Chan, G.K.; Zhu, K.Y.; Tang, W.L.; Dong, T.T.; Shi, Z.Q.; Li, P.; Tsim, K.W. Authentication of Bulbus Fritillariae Cirrhosae by RAPD-derived DNA markers. Molecules 2014, 19, 3450–3459. [Google Scholar] [CrossRef] [PubMed]
  8. Wicke, S.; Schneeweiss, G.M.; de Pamphilis, C.W.; Muller, K.F.; Quandt, D. The evolution of the plastid chromosome in land plants: Gene content, gene order, gene function. Plant Mol. Biol. 2011, 76, 273–297. [Google Scholar] [CrossRef] [PubMed]
  9. Daniell, H.; Lin, C.S.; Yu, M.; Chang, W.J. Chloroplast genomes: Diversity, evolution, and applications in genetic engineering. Genome Biol. 2016, 17, 134. [Google Scholar] [CrossRef] [PubMed]
  10. Yang, M.; Zhang, X.; Liu, G.; Yin, Y.; Chen, K.; Yun, Q.; Zhao, D.; Al-Mssallem, I.S.; Yu, J. The complete chloroplast genome sequence of date palm (Phoenix dactylifera L.). PLoS ONE 2010, 5, e12762. [Google Scholar] [CrossRef] [PubMed]
  11. Shaw, J.; Lickey, E.B.; Schilling, E.E.; Small, R.L. Comparison of whole chloroplast genome sequences to choose noncoding regions for phylogenetic studies in angiosperms: The tortoise and the hare iii. Am. J. Bot. 2007, 94, 275–288. [Google Scholar] [CrossRef] [PubMed]
  12. Wicke, S.; Muller, K.F.; de Pamphilis, C.W.; Quandt, D.; Wickett, N.J.; Zhang, Y.; Renner, S.S.; Schneeweiss, G.M. Mechanisms of functional and physical genome reduction in photosynthetic and nonphotosynthetic parasitic plants of the Broomrape family. Plant Cell 2013, 25, 3711–3725. [Google Scholar] [CrossRef] [PubMed]
  13. Umesono, K.; Inokuchi, H.; Ohyama, K.; Ozeki, H. Nucleotide sequence of Marchantia polymorpha chloroplast DNA: A region possibly encoding three trnas and three proteins including a homologue of E. coli ribosomal protein S14. Nucleic Acids Res. 1984, 12, 9551–9565. [Google Scholar] [CrossRef] [PubMed]
  14. Benson, D.A.; Karsch-Mizrachi, I.; Lipman, D.J.; Ostell, J.; Rapp, B.A.; Wheeler, D.L. Genbank. Nucleic Acids Res. 2000, 28, 15–18. [Google Scholar] [CrossRef] [PubMed]
  15. Varshney, R.K.; Nayak, S.N.; May, G.D.; Jackson, S.A. Next-generation sequencing technologies and their implications for crop genetics and breeding. Trends Biotechnol. 2009, 27, 522–530. [Google Scholar] [CrossRef] [PubMed]
  16. Dong, W.; Liu, J.; Yu, J.; Wang, L.; Zhou, S. Highly variable chloroplast markers for evaluating plant phylogeny at low taxonomic levels and for DNA barcoding. PLoS ONE 2012, 7, e35071. [Google Scholar] [CrossRef] [PubMed]
  17. Cho, K.S.; Yun, B.K.; Yoon, Y.H.; Hong, S.Y.; Mekapogu, M.; Kim, K.H.; Yang, T.J. Complete chloroplast genome sequence of tartary buckwheat (Fagopyrum tataricum) and comparative analysis with common buckwheat (F. esculentum). PLoS ONE 2015, 10, e0125332. [Google Scholar] [CrossRef] [PubMed]
  18. Luo, R.; Liu, B.; Xie, Y.; Li, Z.; Huang, W.; Yuan, J.; He, G.; Chen, Y.; Pan, Q.; Liu, Y.; et al. Soapdenovo2: An empirically improved memory-efficient short-read de novo assembler. Gigascience 2012, 1, 18. [Google Scholar] [CrossRef] [PubMed]
  19. Delcher, A.L.; Salzberg, S.L.; Phillippy, A.M. Using mummer to identify similar regions in large sequence sets. Curr. Protoc. Bioinform. 2003. [Google Scholar] [CrossRef]
  20. Li, Q.; Li, Y.; Song, J.; Xu, H.; Xu, J.; Zhu, Y.; Li, X.; Gao, H.; Dong, L.; Qian, J.; et al. High-accuracy de novo assembly and SNP detection of chloroplast genomes using a SMRT circular consensus sequencing strategy. New Phytol. 2014, 204, 1041–1049. [Google Scholar] [CrossRef] [PubMed]
  21. Wyman, S.K.; Jansen, R.K.; Boore, J.L. Automatic annotation of organellar genomes with DOGMA. Bioinformatics 2004, 20, 3252–3255. [Google Scholar] [CrossRef] [PubMed]
  22. Lowe, T.M.; Eddy, S.R. Trnascan-se: A program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Res. 1997, 25, 955–964. [Google Scholar] [CrossRef] [PubMed]
  23. Lohse, M.; Drechsel, O.; Bock, R. OrganellarGenomeDRAW (OGDRAW): A tool for the easy generation of high-quality custom graphical maps of plastid and mitochondrial genomes. Curr. Genet. 2007, 52, 267–274. [Google Scholar] [CrossRef] [PubMed]
  24. Tamura, K.; Stecher, G.; Peterson, D.; Filipski, A.; Kumar, S. Mega6: Molecular evolutionary genetics analysis version 6.0. Mol. Biol. Evol. 2013, 30, 2725–2729. [Google Scholar] [CrossRef] [PubMed]
  25. Frazer, K.A.; Pachter, L.; Poliakov, A.; Rubin, E.M.; Dubchak, I. Vista: Computational tools for comparative genomics. Nucleic Acids Res. 2004, 32, W273–W279. [Google Scholar] [CrossRef] [PubMed]
  26. Thiel, T. Misa—Microsatellite Identification Tool. 2003. Available online: http://pgrc.ipk-gatersleben. de/misa/ (accessed on 15 March 2017).
  27. Benson, G. Tandem repeats finder: A program to analyze DNA sequences. Nucleic Acids Res. 1999, 27, 573–580. [Google Scholar] [CrossRef] [PubMed]
  28. Warburton, P.E.; Giordano, J.; Cheung, F.; Gelfand, Y.; Benson, G. Inverted repeat structure of the human genome: The X-chromosome contains a preponderance of large, highly homologous inverted repeats that contain testes genes. Genome Res. 2004, 14, 1861–1869. [Google Scholar] [CrossRef] [PubMed]
  29. Katoh, K.; Misawa, K.; Kuma, K.I.; Miyata, T. Mafft: A novel method for rapid multiple sequence alignment based on fast fourier transform. Nucleic Acids Res. 2002, 30, 3059–3066. [Google Scholar] [CrossRef] [PubMed]
  30. Hall, T.A. Bioedit: A user-friendly biological sequence alignment editor and analysis program for windows 95/98/NT. Nucleic Acid Symp. Ser. 1999, 41, 95–98. [Google Scholar]
  31. Librado, P.; Rozas, J. DnaSP v5: A software for comprehensive analysis of DNA polymorphism data. Bioinformatics 2009, 25, 1451–1452. [Google Scholar] [CrossRef] [PubMed]
  32. Qian, J.; Song, J.; Gao, H.; Zhu, Y.; Xu, J.; Pang, X.; Yao, H.; Sun, C.; Li, X.; Li, C.; et al. The complete chloroplast genome sequence of the medicinal plant salvia miltiorrhiza. PLoS ONE 2013, 8, e57607. [Google Scholar] [CrossRef] [PubMed]
  33. Huang, Y.Y.; Matzke, A.J.; Matzke, M. Complete sequence and comparative analysis of the chloroplast genome of Coconut palm (Cocos nucifera). PLoS ONE 2013, 8, e74736. [Google Scholar] [CrossRef] [PubMed]
  34. Raubeson, L.A.; Peery, R.; Chumley, T.W.; Dziubek, C.; Fourcade, H.M.; Boore, J.L.; Jansen, R.K. Comparative chloroplast genomics: Analyses including new sequences from the angiosperms Nuphar advena and Ranunculus macranthus. BMC Genom. 2007, 8, 174. [Google Scholar] [CrossRef] [PubMed]
  35. Ye, C.Y.; Lin, Z.; Li, G.; Wang, Y.Y.; Qiu, J.; Fu, F.; Zhang, H.; Chen, L.; Ye, S.; Song, W.; et al. Echinochloa chloroplast genomes: Insights into the evolution and taxonomic identification of two weedy species. PLoS ONE 2014, 9, e113657. [Google Scholar] [CrossRef] [PubMed]
  36. Sanchez-Puerta, M.V.; Abbona, C.C. The chloroplast genome of Hyoscyamus niger and a phylogenetic study of the tribe Hyoscyameae (Solanaceae). PLoS ONE 2014, 9, e98353. [Google Scholar] [CrossRef] [PubMed]
  37. Kahlau, S.; Aspinall, S.; Gray, J.C.; Bock, R. Sequence of the Tomato chloroplast DNA and evolutionary comparison of Solanaceous plastid genomes. J. Mol. Evol. 2006, 63, 194–207. [Google Scholar] [CrossRef] [PubMed]
  38. Sasaki, T.; Yukawa, Y.; Miyamoto, T.; Obokata, J.; Sugiura, M. Identification of RNA editing sites in chloroplast transcripts from the maternal and paternal progenitors of Tobacco (Nicotiana tabacum): Comparative analysis shows the involvement of distinct trans-factors for ndhB editing. Mol. Biol. Evol. 2003, 20, 1028–1035. [Google Scholar] [CrossRef] [PubMed]
  39. Gao, L.; Yi, X.; Yang, Y.X.; Su, Y.J.; Wang, T. Complete chloroplast genome sequence of a tree fern Alsophila spinulosa: Insights into evolutionary changes in fern chloroplast genomes. BMC Evol. Biol. 2009, 9, 130. [Google Scholar] [CrossRef] [PubMed]
  40. Do, H.D.; Kim, J.S.; Kim, J.H. A trnI_CAU triplication event in the complete chloroplast genome of paris verticillata m.Bieb. (melanthiaceae, liliales). BMC Evol. Biol. 2014, 6, 1699–1706. [Google Scholar] [CrossRef] [PubMed]
  41. Curci, P.L.; De Paola, D.; Danzi, D.; Vendramin, G.G.; Sonnante, G. Complete chloroplast genome of the multifunctional crop globe artichoke and comparison with other Asteraceae. PLoS ONE 2015, 10, e0120589. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, H.; Li, C.; Miao, H.; Xiong, S. Insights from the complete chloroplast genome into the evolution of Sesamum indicum L. PLoS ONE 2013, 8, e80508. [Google Scholar] [CrossRef] [PubMed]
  43. Dong, W.; Xu, C.; Li, C.; Sun, J.; Zuo, Y.; Shi, S.; Cheng, T.; Guo, J.; Zhou, S. Ycf1, the most promising plastid DNA barcode of land plants. Sci. Rep. 2015, 5, 8348. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, J.; Hao, Z.; Xu, H.; Yang, L.; Liu, G.; Sheng, Y.; Zheng, C.; Zheng, W.; Cheng, T.; Shi, J. The complete chloroplast genome sequence of the relict woody plant Metasequoia glyptostroboides Hu et Cheng. Front. Plant Sci. 2015, 6, 447. [Google Scholar] [CrossRef] [PubMed]
  45. Liu, Q.; Xue, Q. Comparative studies on codon usage pattern of chloroplasts and their host nuclear genes in four plant species. J. Genet. 2005, 84, 55–62. [Google Scholar] [CrossRef] [PubMed]
  46. Wolfe, K.H.; Li, W.-H.; Sharp, P.M. Rates of nucleotide substitution vary greatly among plant mitochondrial, chloroplast, and nuclear DNAs. Proc. Natl. Acad. Sci. USA 1987, 84, 9054–9058. [Google Scholar] [CrossRef] [PubMed]
  47. Kim, K.-J.; Lee, H.-L. Complete chloroplast genome sequences from Korean ginseng (panax schinseng Nees) and comparative analysis of sequence evolution among 17 vascular plants. DNA Res. 2004, 11, 247–261. [Google Scholar] [CrossRef] [PubMed]
  48. Park, I.; Kim, J.; Lee, J.; Kim, S.; Cho, O.; Yang, K.; Ahn, J.; Nahm, S.; Kim, H. Development of SSR markers by next-generation sequencing of Korean landraces of chamoe (Cucumis melo var. makuwa). Mol. Bio. Rep. 2013, 40, 6855–6862. [Google Scholar] [CrossRef] [PubMed]
  49. Zalapa, J.E.; Cuevas, H.; Zhu, H.; Steffan, S.; Senalik, D.; Zeldin, E.; McCown, B.; Harbut, R.; Simon, P. Using next-generation sequencing approaches to isolate simple sequence repeat (SSR) loci in the plant sciences. Am. J. Bot. 2012, 99, 193–208. [Google Scholar] [CrossRef] [PubMed]
  50. Wang, R.J.; Cheng, C.L.; Chang, C.C.; Wu, C.L.; Su, T.M.; Chaw, S.M. Dynamics and evolution of the inverted repeat-large single copy junctions in the chloroplast genomes of monocots. BMC Evol. Biol. 2008, 8, 36. [Google Scholar] [CrossRef] [PubMed]
  51. Redwan, R.M.; Saidin, A.; Kumar, S.V. Complete chloroplast genome sequence of MD-2 pineapple and its comparative analysis among nine other plants from the subclass Commelinidae. BMC Plant Biol. 2015, 15, 196. [Google Scholar] [CrossRef] [PubMed]
  52. Song, Y.; Dong, W.; Liu, B.; Xu, C.; Yao, X.; Gao, J.; Corlett, R.T. Comparative analysis of complete chloroplast genome sequences of two tropical trees Machilus yunnanensis and Machilus balansae in the family Lauraceae. Front. Plant Sci. 2015, 6, 662. [Google Scholar] [CrossRef] [PubMed]
  53. Jansen, R.K.; Cai, Z.; Raubeson, L.A.; Daniell, H.; Depamphilis, C.W.; Leebens-Mack, J.; Muller, K.F.; Guisinger-Bellian, M.; Haberle, R.C.; Hansen, A.K.; et al. Analysis of 81 genes from 64 plastid genomes resolves relationships in angiosperms and identifies genome-scale evolutionary patterns. Proc. Natl. Acad. Sci. USA 2007, 104, 19369–19374. [Google Scholar] [CrossRef] [PubMed]
  54. Moore, M.J.; Bell, C.D.; Soltis, P.S.; Soltis, D.E. Using plastid genome-scale data to resolve enigmatic relationships among basal angiosperms. Proc. Natl. Acad. Sci. USA 2007, 104, 19363–19368. [Google Scholar] [CrossRef] [PubMed]
  55. Lin, C.S.; Chen, J.J.; Huang, Y.T.; Chan, M.T.; Daniell, H.; Chang, W.J.; Hsu, C.T.; Liao, D.C.; Wu, F.H.; Lin, S.Y.; et al. The location and translocation of ndh genes of chloroplast origin in the Orchidaceae family. Sci. Rep. 2015, 5, 9040. [Google Scholar] [CrossRef] [PubMed]
  56. Moore, M.J.; Dhingra, A.; Soltis, P.S.; Shaw, R.; Farmerie, W.G.; Folta, K.M.; Soltis, D.E. Rapid and accurate pyrosequencing of angiosperm plastid genomes. BMC Plant Biol. 2006, 6, 17. [Google Scholar] [CrossRef] [PubMed]
  57. Kim, K.; Lee, S.C.; Lee, J.; Lee, H.O.; Joh, H.J.; Kim, N.H.; Park, H.S.; Yang, T.J. Comprehensive survey of genetic diversity in chloroplast genomes and 45s nrDNAs within panax ginseng species. PLoS ONE 2015, 10, e0117159. [Google Scholar] [CrossRef] [PubMed]
  58. Garaycochea, S.; Speranza, P.; Alvarez-Valin, F. A strategy to recover a high-quality, complete plastid sequence from low-coverage whole-genome sequencing. Appl. Plant Sci. 2015, 3. [Google Scholar] [CrossRef] [PubMed]
  59. Saha, M.C.; Cooper, J.D.; Mian, M.A.; Chekhovskiy, K.; May, G.D. Tall fescue genomic SSR markers: Development and transferability across multiple grass species. Theor. Appl. Genet. 2006, 113, 1449–1458. [Google Scholar] [CrossRef] [PubMed]
  60. Yap, J.Y.; Rohner, T.; Greenfield, A.; Van Der Merwe, M.; McPherson, H.; Glenn, W.; Kornfeld, G.; Marendy, E.; Pan, A.Y.; Wilton, A.; et al. Complete chloroplast genome of the Wollemi pine (Wollemia nobilis): Structure and evolution. PLoS ONE 2015, 10, e0128126. [Google Scholar] [CrossRef] [PubMed]
  61. Ni, L.; Zhao, Z.; Dorje, G.; Ma, M. The complete chloroplast genome of Ye-Xing-Ba (Scrophularia dentata; Scrophulariaceae), an alpine Tibetan herb. PLoS ONE 2016, 11, e0158488. [Google Scholar] [CrossRef] [PubMed]
  62. Kong, W.; Yang, J. The complete chloroplast genome sequence of Morus mongolica and a comparative analysis within the Fabidae clade. Curr. Genet. 2016, 62, 165–172. [Google Scholar] [CrossRef] [PubMed]
  63. Zhang, Y.; Du, L.; Liu, A.; Chen, J.; Wu, L.; Hu, W.; Zhang, W.; Kim, K.; Lee, S.C.; Yang, T.J.; et al. The complete chloroplast genome sequences of five Epimedium species: Lights into phylogenetic and taxonomic analyses. Front. Plant Sci. 2016, 7, 306. [Google Scholar] [CrossRef] [PubMed]
  64. Arif, I.A.; Bakir, M.A.; Khan, H.A.; Al Farhan, A.H.; Al Homaidan, A.A.; Bahkali, A.H.; Sadoon, M.A.; Shobrak, M. A brief review of molecular techniques to assess plant diversity. Int. J. Mol. Sci. 2010, 11, 2079–2096. [Google Scholar] [CrossRef] [PubMed]
  65. Vieira, M.L.C.; Santini, L.; Diniz, A.L.; Munhoz, C.D.F. Microsatellite markers: What they mean and why they are so useful. Genet. Mol. Biol. 2016, 39, 312–328. [Google Scholar] [CrossRef] [PubMed]
  66. Moose, S.P.; Mumm, R.H. Molecular plant breeding as the foundation for 21st century crop improvement. Plant Physiol. 2008, 147, 969–977. [Google Scholar] [CrossRef] [PubMed]
  67. Yang, J.B.; Tang, M.; Li, H.T.; Zhang, Z.R.; Li, D.Z. Complete chloroplast genome of the genus Cymbidium: Lights into the species identification, phylogenetic implications and population genetic analyses. BMC Evol. Biol. 2013, 13, 84. [Google Scholar] [CrossRef] [PubMed]
  68. Khakhlova, O.; Bock, R. Elimination of deleterious mutations in plastid genomes by gene conversion. Plant J. 2006, 46, 85–94. [Google Scholar] [CrossRef] [PubMed]
  69. Nie, X.; Lv, S.; Zhang, Y.; Du, X.; Wang, L.; Biradar, S.S.; Tan, X.; Wan, F.; Weining, S. Complete chloroplast genome sequence of a major invasive species, Crofton weed (Ageratina adenophora). PLoS ONE 2012, 7, e36869. [Google Scholar] [CrossRef] [PubMed]
  70. Jheng, C.F.; Chen, T.C.; Lin, J.Y.; Chen, T.C.; Wu, W.L.; Chang, C.C. The comparative chloroplast genomic analysis of photosynthetic orchids and developing DNA markers to distinguish Phalaenopsis orchids. Plant Sci. 2012, 190, 62–73. [Google Scholar] [CrossRef] [PubMed]
  71. Yang, Y.; Zhou, T.; Duan, D.; Yang, J.; Feng, L.; Zhao, G. Comparative analysis of the complete chloroplast genomes of five Quercus species. Front. Plant Sci. 2016, 7, 959. [Google Scholar] [CrossRef] [PubMed]
  72. Li, Y.; Li, Q.; Li, X.; Song, J.; Sun, C. Complete chloroplast genome sequence of Fritillaria unibracteata var. wabuensis based on SMRT sequencing technology. Mitochondrial DNA 2016, 27, 3757–3758. [Google Scholar] [PubMed]
  73. Li, X.; Yang, Y.; Henry, R.J.; Rossetto, M.; Wang, Y.; Chen, S. Plant DNA barcoding: From gene to genome. Biol. Rev. Camb. Philos. Soc. 2015, 90, 157–166. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Not available.
Figure 1. Circular gene map of chloroplast genomes of two Fritillaria species. Genes drawn inside the outer layer circle are transcribed clockwise, and those outside the circle are transcribed counterclockwise. The darker gray in the inner circle corresponds to GC content. * cemA is pseudogene in F. ussuriensis.
Figure 1. Circular gene map of chloroplast genomes of two Fritillaria species. Genes drawn inside the outer layer circle are transcribed clockwise, and those outside the circle are transcribed counterclockwise. The darker gray in the inner circle corresponds to GC content. * cemA is pseudogene in F. ussuriensis.
Molecules 22 00982 g001
Figure 2. Amino acid frequencies in Fritillaria (F.) ussuriensis and F. cirrhosa protein-coding sequences.
Figure 2. Amino acid frequencies in Fritillaria (F.) ussuriensis and F. cirrhosa protein-coding sequences.
Molecules 22 00982 g002
Figure 3. Distribution of SSRs (simple sequence repeat) in the Fritillaria (F.) ussuriensis and F. cirrhosa chloroplast (CP) genomes. (A) SSR type distribution in the two Fritillaria CP genomes; (B) The proportion of SSRs in different genomic regions of Fritillaria CP genomes; (C) SSR distribution between coding and non-coding regions.
Figure 3. Distribution of SSRs (simple sequence repeat) in the Fritillaria (F.) ussuriensis and F. cirrhosa chloroplast (CP) genomes. (A) SSR type distribution in the two Fritillaria CP genomes; (B) The proportion of SSRs in different genomic regions of Fritillaria CP genomes; (C) SSR distribution between coding and non-coding regions.
Molecules 22 00982 g003
Figure 4. Comparison of large single copy (LSC), small single copy (SSC), and inverted repeat (IR) border regions among the six Fritillaria species’ chloroplast genomes. Colored boxes for genes represent the gene position. ψ: pseudogenes.
Figure 4. Comparison of large single copy (LSC), small single copy (SSC), and inverted repeat (IR) border regions among the six Fritillaria species’ chloroplast genomes. Colored boxes for genes represent the gene position. ψ: pseudogenes.
Molecules 22 00982 g004
Figure 5. Comparison of six Fritillaria chloroplast genomes using mVISTA alignment program. The complete chloroplast (CP) genomes of six Fritillaria species were used for comparisons with published CP genomes. Blue block: conserved gene; sky-blue block: transfer RNA (tRNA) and ribosomal RNA (rRNA); red block: conserved non-coding sequences (CNS). White peaks indicate regions with sequence variation among Fritillaria species.
Figure 5. Comparison of six Fritillaria chloroplast genomes using mVISTA alignment program. The complete chloroplast (CP) genomes of six Fritillaria species were used for comparisons with published CP genomes. Blue block: conserved gene; sky-blue block: transfer RNA (tRNA) and ribosomal RNA (rRNA); red block: conserved non-coding sequences (CNS). White peaks indicate regions with sequence variation among Fritillaria species.
Molecules 22 00982 g005
Figure 6. Sliding window analysis of the chloroplast (CP) genomes. (A) Comparison of the nucleotide variability (Pi) between Fritillaria (F.) ussuriensis and F. cirrhosa. (B) Comparison of the nucleotide variability (Pi) among six Fritillaria species’ cp genomes.
Figure 6. Sliding window analysis of the chloroplast (CP) genomes. (A) Comparison of the nucleotide variability (Pi) between Fritillaria (F.) ussuriensis and F. cirrhosa. (B) Comparison of the nucleotide variability (Pi) among six Fritillaria species’ cp genomes.
Molecules 22 00982 g006
Figure 7. Phylogenetic trees constructed with 74 protein-coding genes of 11 species using maximum likelihood (ML) and maximum parsimony (MP) methods. Numbers above are bootstrap support values (>50%).
Figure 7. Phylogenetic trees constructed with 74 protein-coding genes of 11 species using maximum likelihood (ML) and maximum parsimony (MP) methods. Numbers above are bootstrap support values (>50%).
Molecules 22 00982 g007
Table 1. Size comparison of two Fritillaria species’ chloroplast genomic regions.
Table 1. Size comparison of two Fritillaria species’ chloroplast genomic regions.
SpeciesFritillaria ussuriensisFritillaria cirrhosa
Total CP genome size (bp)151,524151,083
LSC (bp)81,73281,390
IR (bp)52,67852,156
SSC (bp)17,11417,537
GC content (%)36.9536.96
LSC (%)34.7134.79
IR (%)42.4042.60
SSC (%)30.6330.50
Total number of genes111112
Protein-coding gene7778
rRNA44
tRNA3030
CP: Chloroplast; LSC: Large single copy; IR: Inverted repeat; SSC: Small single copy.
Table 2. Genes present in the two Fritillaria chloroplast genomes.
Table 2. Genes present in the two Fritillaria chloroplast genomes.
Gene Products of Two Fritillaria Species
Photosystem IpsaA, B, C, I, J
Photosystem IIpsbA, B, C, D, E, F, H, I, J, K, L, M, N, T, Z
Cytochrome b6/fpetA, B (1), D (1), G, L, N
ATP synthaseatpA, B, E, F (1), H, I
RubiscorbcL
NADH oxidoreductasendhA (1), B (1,3), C, D, E, F, G, H, I, J, K
Large subunit ribosomal proteinsrpl2 (1,3), 14, 16 (1), 20, 22, 23 (3), 32, 33, 36
Small subunit ribosomal proteinsrps2, 3, 4, 73 (3), 8, 11, 12 (2–4), 14, 15, 16, 18, 19
RNA polymeraserpoA, B, C1 (1), C2
Unknown function protein-coding geneycf1 (3), 2 (3), 3 (2), 4
Other genesaccD, ccsA, cemA (5), clpP (2), matK
Ribosomal RNAsrrn16 (3), 23 (3), 4.5 (3), 5 (3)
Transfer RNAstrnA-UGC (1,3), trnC-GCA, trnD-GUC, trnE-UUC, trnF-GAA, trnG-UCC (1), trnG-GCC, trnH-GUG, trnI-CAU (3), trnI-GAU (1,3), trnK-UUU1 (3), trnL-UAA1 (3), trnL-UAG, trnL-CAA (3), trnM-CAU, trnfM-CAU, trnN-GUU (3), trnP-UGG, trnQ-UUG, trnR-ACG (3), trnR-UCU, trnS-GCU, trnS-GGA, trnS-UGA, trnT-GGU, trnT-UGU, trnV-UAC (1), trnV-GAC (3), trnW-CCA, trnY-GUA
(1) Gene containing a single intron; (2) gene containing two introns; (3) two gene copies in inverted repeats (IRs); (4) trans-splicing gene; (5) pseudogene in the chloroplast (CP) genome of Fritillaria ussuriensis.

Share and Cite

MDPI and ACS Style

Park, I.; Kim, W.J.; Yeo, S.-M.; Choi, G.; Kang, Y.-M.; Piao, R.; Moon, B.C. The Complete Chloroplast Genome Sequences of Fritillaria ussuriensis Maxim. and Fritillaria cirrhosa D. Don, and Comparative Analysis with Other Fritillaria Species. Molecules 2017, 22, 982. https://doi.org/10.3390/molecules22060982

AMA Style

Park I, Kim WJ, Yeo S-M, Choi G, Kang Y-M, Piao R, Moon BC. The Complete Chloroplast Genome Sequences of Fritillaria ussuriensis Maxim. and Fritillaria cirrhosa D. Don, and Comparative Analysis with Other Fritillaria Species. Molecules. 2017; 22(6):982. https://doi.org/10.3390/molecules22060982

Chicago/Turabian Style

Park, Inkyu, Wook Jin Kim, Sang-Min Yeo, Goya Choi, Young-Min Kang, Renzhe Piao, and Byeong Cheol Moon. 2017. "The Complete Chloroplast Genome Sequences of Fritillaria ussuriensis Maxim. and Fritillaria cirrhosa D. Don, and Comparative Analysis with Other Fritillaria Species" Molecules 22, no. 6: 982. https://doi.org/10.3390/molecules22060982

Article Metrics

Back to TopTop