Next Article in Journal
4-O-Methylhonokiol Influences Normal Cardiovascular Development in Medaka Embryo
Previous Article in Journal
Ripe and Raw Pu-Erh Tea: LC-MS Profiling, Antioxidant Capacity and Enzyme Inhibition Activities of Aqueous and Hydro-Alcoholic Extracts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Complete Chloroplast Genome Sequences of Kaempferia Galanga and Kaempferia Elegans: Molecular Structures and Comparative Analysis

Guangdong Key Lab of Ornamental Plant Germplasm Innovation and Utilization, Environmental Horticulture Research Institute, Guangdong Academy of Agricultural Sciences, Guangzhou 510640, China
*
Author to whom correspondence should be addressed.
Molecules 2019, 24(3), 474; https://doi.org/10.3390/molecules24030474
Submission received: 27 December 2018 / Revised: 24 January 2019 / Accepted: 25 January 2019 / Published: 29 January 2019

Abstract

:
Kaempferia galanga and Kaempferia elegans, which belong to the genus Kaempferia family Zingiberaceae, are used as valuable herbal medicine and ornamental plants, respectively. The chloroplast genomes have been used for molecular markers, species identification and phylogenetic studies. In this study, the complete chloroplast genome sequences of K. galanga and K. elegans are reported. Results show that the complete chloroplast genome of K. galanga is 163,811 bp long, having a quadripartite structure with large single copy (LSC) of 88,405 bp and a small single copy (SSC) of 15,812 bp separated by inverted repeats (IRs) of 29,797 bp. Similarly, the complete chloroplast genome of K. elegans is 163,555 bp long, having a quadripartite structure in which IRs of 29,773 bp length separates 88,020 bp of LSC and 15,989 bp of SSC. A total of 111 genes in K. galanga and 113 genes in K. elegans comprised 79 protein-coding genes and 4 ribosomal RNA (rRNA) genes, as well as 28 and 30 transfer RNA (tRNA) genes in K. galanga and K. elegans, respectively. The gene order, GC content and orientation of the two Kaempferia chloroplast genomes exhibited high similarity. The location and distribution of simple sequence repeats (SSRs) and long repeat sequences were determined. Eight highly variable regions between the two Kaempferia species were identified and 643 mutation events, including 536 single-nucleotide polymorphisms (SNPs) and 107 insertion/deletions (indels), were accurately located. Sequence divergences of the whole chloroplast genomes were calculated among related Zingiberaceae species. The phylogenetic analysis based on SNPs among eleven species strongly supported that K. galanga and K. elegans formed a cluster within Zingiberaceae. This study identified the unique characteristics of the entire K. galanga and K. elegans chloroplast genomes that contribute to our understanding of the chloroplast DNA evolution within Zingiberaceae species. It provides valuable information for phylogenetic analysis and species identification within genus Kaempferia.

1. Introduction

The genus Kaempferia belongs to the family Zingiberaceae, which consists of approximately 50 species in the world [1,2,3]. Kaempferia species are distributed in tropical Asia regions [1,2]. Kaempferia species are grown primarily for their ornamental foliage rather than for their flowers [3]. In addition, several species have long been cultivated for their medicinal properties [3]. Kaempferia galanga and Kaempferia elegans are valuable herbal medicine and ornamental plants in this genus, respectively. K. galanga is mainly distributed in the regions of Southern and Northwestern China (Guangdong, Guangxi, Guizhou, Yunnan and Sichuan provinces) and is widely cultivated in Southeast Asia; whereas K. elegans is only produced in Sichuan province of China and is commonly cultivated in tropical Asia regions [1,2]. Morphology data had been used to determine the differences between K. galanga and K. elegans [1,2,3]. From these studies, the two Kaempferia species had been characterized differently by leaf shape, petioles, flower color and rhizomes [1,2,3]. The leaves of K. galanga spread flat on ground, subsessile, green, orbicular, 7–20 × 3–17 cm, glabrous on both surfaces or villous abaxially, margin usually white, apex mucronate or acute; the leaves of K. elegans have petioles, to 10 cm, leaf adaxially green, abaxially pale green, oblong or elliptic, 13–15 × 5–8 cm, margin usually red, base rounded, apex acute. The petals are white in K. galanga, whereas the petals are purple in K. elegans. The rhizomes of K. galanga are pale green or greenish white inside, tuberous and fragrant; the rhizomes of K. elegans are bearing globose tubers with fibrous roots. K. galanga can be used as aromatic medicinal plant and also as ornamental plant with important economic value, whereas K. elegans is mainly used as potted plant with ornamental value. In detail, the rhizomes of K. galanga have been used as aromatic stomachic, with effects of dissipating cold, dampness, warm the spleen and stomach, also used as flavoring spices, for example, a famous Guangdong delicacy of sand ginger salted chicken [2]. With increased demand for rhizomes of K. galanga, researches related to the high rhizomes yield for its cultivation has been lacking, leading to having a high price in the market. However, sometimes morphological identification of Kaempferia species was difficult owing to the morphological similarity of vegetative parts among species and other genera in Zingiberaceae, such as Boesenbergia, Cornukaempferia, Curcuma and Scaphochlamys [3,4]. In addition, intraspecific variation caused more complicated problems in the morphological taxonomy of the genus Kaempferia [3,4]. Based only on morphological characteristics, we could not conclusively distinguish and identify the Kaempferia species and hybrids and other genera species in Zingiberaceae. Therefore, the morphological classification and relationships within Kaempferia species need further investigation together with more molecular analyses. Combined evidences from morphology characteristics and chloroplast DNA have proven useful and powerful in species identification and phylogenetic relationships analysis [4,5,6,7,8].
In angiosperm plants, chloroplasts play an important role in photosynthesis and the metabolism of starch, fatty acids, nitrogen, amino acids and internal redox signals [9,10,11]. In general, the chloroplast genomes of angiosperms encode 110–130 genes with a size range of 120-180 kb, which have a typical quadripartite structure consisting of a large single copy (LSC) region, a small single copy (SSC) region and two copies of inverted repeats (IRs) [12,13,14,15]. As the chloroplast is the center of photosynthesis, the research of the chloroplast genome is important to discover the mechanisms of plant photosynthesis. The third-generation sequencing platform PacBio utilizes a single-molecule real-time sequencing technology, which has been successfully used to determine many chloroplast genome sequences [16,17,18]. The main advantage of this method is the long read length of over 10 kb on average, which provides many benefits in genome assembly, including longer contigs and fewer unresolved gaps [16,17]. However, PacBio sequencing has high rates of random error in its single-pass reads; therefore, in combination with Illumina sequencing can reduce these random errors [17,18].
In this study, we sequenced and analyzed the complete chloroplast genomes from K. galanga and K. elegans, respectively, using Illumina and PacBio sequencing. We also characterized the long repeats and SSRs detected in the genome, including repeat types, distribution patterns and so on. Comparative sequence analysis and phylogenetic relationships were also analyzed among other members in the family Zingiberaceae. These results will improve the genetic information of the genus Kaempferia we already have and will be beneficial for DNA molecular studies in Kaempferia.

2. Results

2.1. Chloroplast Genome Organization of Two Kaempferia Species

The complete chloroplast genome of K. galanga and K. elegans consisted of a single circular molecule with quadripartite structure (Figure 1). The sizes of K. galanga and K. elegans chloroplast genomes were 163,811 and 163,555 bp, respectively. They consisted of a pair of inverted repeats (IRs) of 29,797 bp in K. galanga and 29,773 bp in K. elegans, a large single copy (LSC) region of 88,405 bp in K. galanga and 88,020 bp in K. elegans and a small single copy (SSC) region of 15,812 bp in K. galanga and 15,989 bp in K. elegans (Figure 1 and Table 1). The GC content of the genomes was 36.1% both in K. galanga and K. elegans but the IR regions had higher GC contents (41.2% and 41.1% in K. galanga and K. elegans, respectively) than that of the LSC regions (33.9% both in K. galanga and K. elegans) and SSC regions (29.5% and 29.4% in K. galanga and K. elegans, respectively). Approximately 48.3–50.7% of the two Kaempferia species chloroplast genomes consisted of protein-coding genes (83,172 bp in K. galanga and 79,117 bp in K. elegans), 1.7% of tRNAs (2870 bp K. galanga and 2852 bp in K. elegans) and 5.5% of rRNAs (9046 bp in K. galanga and 9046 bp in K. elegans). For protein-coding regions, the AT content for the first, second and third codons were 55.4%, 62.6% and 71.1% in K. galanga, respectively and 66.9%, 56.7% and 64.7% in K. elegans, respectively (Table 1). The non-coding regions consisting of introns, pseudogenes and intergenic spacers accounted for 49.3% and 51.7% for the K. galanga and K. elegans chloroplast genomes, respectively (Table 1).
There were 111 predicted genes in the K. galanga chloroplast genome including 79 protein-coding genes, 28 tRNA genes and 4 rRNA genes, while 113 genes predicted in the K. elegans chloroplast genome consisted of 79 protein-coding genes, 30 tRNA genes and 4 rRNA genes (Table 2). Among the protein-coding genes in K. galanga chloroplast genome, 61 genes were located in the LSC region, 12 genes were in the SSC region and 8 genes were duplicated in the IR regions (Supplementary file 1). In total, there were 18 intron-containing genes in the K. galanga chloroplast genome, 16 of which contained one intron and two of which (ycf3 and clpP) contained two introns (Table 3). Among the protein-coding genes in the K. elegans chloroplast genome, 63 genes were located in the LSC region, 12 genes were in the SSC region and 6 genes were duplicated in the IR regions (Supplementary file 2). In total, there were 17 intron-containing genes in the K. elegans chloroplast genome, 15 of which contained one intron and two of which (ycf3 and clpP) contained two introns (Table 3).
Relative synonymous codon usage (RSCU) is the ratio between frequency of use and expected frequency of a particular codon. RSCU values <1.00 indicate use of a codon less frequent than expected, while codons used more frequently than expected have a score of >1.00 [19,20]. The codon usage of the K. galanga and K. elegans chloroplast genomes are summarized in Table S1. Protein-coding genes comprised 27,724 and 27,675 codons in both the K. galanga and K. elegans, respectively. Among these codons, those for leucine and isoleucine were the most common in both K. galanga and K. elegans chloroplast genomes (Figure 2 and Table S1). The use of the codons ATG and TGG, encoding Met and Trp respectively, exhibited no bias (RSCU = 1.00) in these two Kaempferia species chloroplast genomes (Table S1). Codons ending in A and/or U accounted for 71.1% and 64.7% of all protein-coding gene codons of the chloroplast genomes of K. galanga and K. elegans, respectively (Table 1 and Table S1). The findings also revealed that all of the types of preferred synonymous codons (RSCU>1.00) ended with A or U except for trL-CAA in these two Kaempferia species (Table S1).

2.2. Analysis of SSRs and Long Repeats

SSRs or microsatellites, are tandem repeat sequences consisting of 1-6 nucleotide repeat units and are widely distributed in chloroplast genomes [5,7,12]. SSRs were detected using MISA in both Kaempferia species chloroplast genomes. We detected 240 and 248 SSRs in K. galanga and K. elegans chloroplast genomes, respectively. Mononucleotide motifs were the most abundant type of repeat and dinucleotides were the second most abundant in both Kaempferia species chloroplast genomes (Figure 3 and Table S2). There were 177 momo-, 32 di-, 6 tri-, 21 tetra-, 3 penta- and one hexa-nucleotide SSRs in K. galanga chloroplast genome; by contrast, there were 188 momo-, 33 di-, 7 tri-, 19 tetra- and 1 penta-nucleotide SSRs in K. elegans chloroplast genome (Figure 3). The majority of SSRs were located in the LSC regions rather than in IR and SSC regions of both Kaempferia species chloroplast genomes (Figure 3 and Table S2). SSRs were more abundant in non-coding regions than in coding regions of both genomes (Figure 3). Furthermore, almost all SSR loci were composed of A or T, which contributed to the bias in base composition (A/T; both 63.9%) in the chloroplast genomes of the two Kaempferia species.
Long repeat sequences in the K. galanga and K. elegans chloroplast genomes were analyzed by REPuter and results shown in Figure 4 and Table S3. In the K. galanga chloroplast genome, 21 forward repeats, 20 palindrome repeats, 5 reverse repeats and 4 complement repeats were detected (identity>90%) (Figure 4 and Table S3). In comparison, in the K. elegans chloroplast genome, 26 forward repeats, 17 palindrome repeats, 4 reverse repeats and 2 complement repeats were detected (Figure 4 and Table S3). Out of the 50 repeats in K. galanga chloroplast genome, 38 repeats (76.0%) were 30–39 bp long, 9 repeats (18.0%) were 40–49 bp long and 3 repeats (6.0%) were ≥50 bp long (Figure 4 and Table S3). By contrast, of the 49 repeats in K. elegans chloroplast genome, 37 repeats (75.5%) were 30–39 bp long, 6 repeats (12.2%) were 40–49 bp long and 6 repeats (12.2%) were ≥50 bp long (Figure 4 and Table S3). The majority of these repeats were mainly forward and palindromic types with lengths mainly in the range of 30–50 bp in both Kaempferia species.

2.3. IR Contraction and Expansion

A detailed comparison was performed for four junctions, LSC/IRa, LSC/IRb, SSC/IRa and SSC/IRb, between the two IRs (IRa and IRb) and the two single-copy regions (LSC and SSC) among A. zerumbet, C. flaviflora and Z. spectabile in comparison to K. galanga and K. elegans (Figure 5). Although the IR region of the five Zingiberaceae species chloroplast genomes was highly conserved, structure variation was still found in the IR/SC boundary regions. As shown in Figure 5, the rpl22-rps19 genes were located in the junctions of the LSC/IRb regions in K. galanga, K. elegans, A. zerumbet and C. flaviflora, though the trnM-ycf2 sequence in Z. spectabile, one of which was missing the rpl22/-rps19 gene in the junctions of the LSC/IRb regions. The ycf1-ndhF genes were located at the junctions of the IRb/SSC regions in the five Zingiberaceae species chloroplast genomes. The ndhF gene was 23, 98, 251, 133 and 33 bp from the IRb/SSC border in K. galanga, K. elegans, A. zerumbet, C. flaviflora and Z. spectabile, respectively (Figure 5). The SSC/IRa junctions in the five Zingiberaceae species chloroplast genomes were crossed by the ycf1 gene, with 665–3888 bp in the IRa region. Like the IRb/SSC boundary regions, the IRa/LSC regions were also variable. The rps19-psbA genes were located in the junctions of the IRa/LSC regions in K. galanga, K. elegans, A. zerumbet and C. flaviflora, though the trnH-psbA genes in Z. spectabile, one of which was missing the rps19 gene in the junctions of the IRa/LSC region. The rps19-psbA genes of K. elegans were located at the junctions of IRa/LSC regions with 136 and 123 bp, respectively, separating the spacer from the end of the IRa region. However, in Z. spectabile, the trnH gene was the last gene at one end of the IRa region, 256 bp away from the IRa/LSC border. Overall, contraction and expansion of the IR regions was detected across the five Zingiberaceae species chloroplast genomes.

2.4. Comparative Chloroplast Genomic Analysis

To characterize genome divergence, we performed multiple sequence alignments between the five Zingiberaceae species chloroplast genomes using the program mVISTA, with K. galanga being used as a reference (Figure 6). The comparison demonstrated that the two IR regions were less divergent than the LSC and SSC. Moreover, the coding regions are more conserved than the non-coding regions. The most highly divergent regions among the five chloroplast genomes were found among the intergenic spacers, including trnH-psbA, rps16-psbK, atpH-atpI, petN-psbM, trnE-psbD psbC-rps14, rps4-ycf3, rps4-ndhJ, ndhC-atpE, ycf4-cemA and petA-psbJ in LSC as well as rpl32-ccsA, psaC-ndhG and ndhG-ndhI in SSC. Higher divergence in the coding regions was found in the matK, rpoA, rps16, rps19, ndhF, ccsA, psaC, ndhD, ndhE, ndhG, ndhI, ndhA and ycf1 sequences.
Furthermore, sliding window analysis using DnaSP detected highly variable regions in the chloroplast genomes between K. galanga and K. elegans (Figure 7A). The average value of nucleotide diversity (Pi) was 0.01075. The IR regions showed lower variability than the LSC and SSC regions. There were 7 mutational hotspots that exhibited remarkably higher Pi values (>0.03) and were located at the LSC and SSC regions, which included trnS-trnG, rps12-clpP, psbT-psbN, ycf1-ndhF, ndhF-rpl32, psaC-ndhE and ccsA-ndhD regions from the chloroplast genomes (Figure 7A). By contrast, there was only 1 mutational hotspot (rpl2-trnH) that exhibited remarkably higher Pi values (>0.03) located at the IR regions (Figure 7A).
Figure 7B showed that the average value of Pi was 0.01591 among two Kaempferia species, A. zerumbet, C. flaviflora and Z. spectabile. The Pi values of these five species were commonly higher than those of the two Kaempferia species. Particularly, seven highly divergent loci showed remarkably higher Pi values (>0.045), including trnS-trnG, psbT-psbN, trnH-rpl2, trnI-ycf2, ccsA-ndhD, psaC-ndhE and ycf2-trnI regions from the chloroplast genomes (Figure 7B). These regions may be undergoing rapid nucleotide substitution at the species level, indicating potential application of molecular markers for plant identification and phylogenetic analysis.
The chloroplast genomes of K. galanga and K. elegans were found to show a 256 bp difference in length (Table 1). In addition to the total length difference, we assessed SNP and Indel variations between the two Kaempferia species chloroplast genomes in their entirety. There were 536 SNPs identified in the two chloroplast genomes (Table S4). The most frequently occurring mutations were located in intergenic region, which included 357 SNPs. The coding regions contained 91 synonymous SNPs, 87 nonsynonymous SNPs and 1 stop mutation. There were 107 indels in the chloroplast genome identified between K. galanga and K. elegans (Table S5), including 47 deletions and 60 insertions. Of the 107 indel markers between K. galanga and K. elegans genomes, the longest indels (10 bp) were located within the two intergenic sequences (petA-psbJ and atpH-atpI) and two coding sequences (atpF and rps12).

2.5. Phylogenetic Analysis

In this study, phylogenetic trees were constructed with SNPs from eleven species using ML and MP methods, respectively, including nine Zingiberaceae plants and using C. pulverulentus and C. indica as outgroups (Figure 8). Both the ML and MP phylogenetic trees strongly indicated that K. galanga and K. elegans formed a cluster within Zingiberaceae and the C. pulverulentus and C. indica species were clearly separated from Zingiberaceae species (Figure 8). Among nine Zingiberaceae species, they were clustered into four clusters. The first cluster comprised the genus Kaempferia (K. galanga and K. elegans). The second cluster comprised the two genera—Zingiber and Curcuma (Z. spectabile, C. flaviflora and C. roscoeana). The third cluster comprised the genus Amomum (A. kravanh and A. compactum). The fourth cluster comprised the genus Alpinia (A. zerumbet and A. oxyphylla).

2.6. Potential RNA Editing Sites

In the present study, potential RNA editing sites were predicted for 34 genes; as a result, a total of 54 and 80 RNA editing sites were identified in the K. galanga and K. elegans chloroplast genomes, respectively (Table S6). No potential editing sites were identified in seven genes (petG, petL, psaB, psaI, psbL, rpl2, rpl23) in both chloroplast genomes. Of the 54 editing sites, which occurred in 21 genes, 15 (27.8%) and 39 (72.2%) were located at the first and the second codon position, respectively, in K. galanga. Of the 80 editing sites, which occurred in 26 genes, 21 (26.2%) and 59 (73.8%) were located at the first codon and the second codon position, respectively, in K. elegans. No editing sites were found at the third codon position in both Kaempferia species.
We also observed that RNA editing sites were all C to U conversion both in K. galanga and K. elegans. In K. galanga, the ndhB gene was predicted to have ten editing sites; accD and matK, five; rpoB and rpoC2, four; rpl20, rpoA and rps14, three; atpB, petB, psbE and rpoC1, two; and one each in atpA, atpF, atpI, ccsA, clpP, psbB, psbF, rps2 and rps8. Meanwhile in K. elegans, the gene of ndhB was predicted to have ten editing sites; ndhA and ndhD, seven; accD, matK, ndhF and rpoC2, five; rpoB and ycf3, four; rpl20, rpoA and rps14, three; atpB, ndhG, petB, psbB and rpoC1, two; and one each in atpA, atpF, atpI, ccsA, clpP, psbF, rps2, rps8 and rps16.

3. Materials and Methods

3.1. Plant Material and DNA Isolation

Fresh leaves were collected from potted K. galanga and K. elegans plants, respectively, from greenhouse in environmental horticulture research institute, Guangdong academy of agricultural sciences, Guangzhou, China. Total chloroplast DNA was extracted from about 100 g of leaves using the sucrose gradient centrifugation method as improved by Li et al. [21]. The chloroplast DNA concentration for each sample was estimated using an ND-2000 spectrometer (Nanodrop technologies, Wilmington, DE, USA), whereas visual examination was performed using gel electrophoresis.

3.2. Chloroplast Genome Sequencing and Genome Assembly

The chloroplast DNA was first fragmented into 300–500 bp using a Covaris M220 Focused-ultrasonicator (Covaris, Woburn, MA, USA) and used to construct short-insert libraries (insert size about 430 bp) according to the manufacturer’s instructions (Illumina, San Diego, CA, USA). The short fragments were sequenced using an Illumina Hiseq X Ten platform (Novogene, Beijing, China). The Illumina raw reads were cleaned by removing the adapter sequences and low quality sequences, which included the reads with ambiguous nucleotides and ones containing more than 10% nucleotides in read with Q-value ≤ 20 and short reads (length < 50 bp).
The chloroplast DNA was also fragmented into 8-10 kb fragments, which were subjected to DNA sequencing following the standard protocol provided by PacBio platform (Novogene, Beijing, China). The PacBio raw reads were pre-processed by trimming the adapter sequences, low quality (Q < 0.80) reads, short reads (length < 100 bp) and short subreads (length <500 bp).
Initially, the Illumina clean reads were assembled using SOAPdenovo (version 2.04, Hongkong, China) with default parameters into principal contigs [22] and all contigs were sorted and joined into a single draft sequence using the software Geneious version 11.0.4 (Auckland, New Zealand) [23]. Next, BLASR software (San Diego, CA, USA) was used to compare the PacBio clean data with the single draft sequence and to extract the correction and error correction [24]. Next, the corrected PacBio clean data were assembled using Celera Assembler (version 8.0, Rockville, MD, USA) with default parameters, thus generating scaffolds [25]. Next, the assembled scaffolds were mapped back to the Illumina clean reads using GapCloser (version 1.12, Hongkong, China) for gap closing [22]. Finally, the redundant fragments sequences were removed, thus generating the final assembled chloroplast genomic sequence.

3.3. Chloroplast Genome Annotation and Codon Usage

The initial gene annotation of the chloroplast genome was carried out with BLAST homology searches and DOGMA (Dual Organellar Genome Annotator) [26]. tRNA genes were identified using tRNAscanSE with default settings [27]. The gene homologies were confirmed by comparing them with National Center for Biotechnology Information (NCBI)’s non-redundant (Nr) protein database, Clusters of orthologous groups (COG) for eukaryotic complete genomes database (http://www.ncbi.nlm.nih.gov/COG), Kyoto Encyclopedia of Genes and Genomes (KEGG) (http://www.kegg.jp/), Gene Ontology (GO) (http://www.geneontology.org) and SWISS-PROT (http://web.expasy.org/docs/swissprotguideline.html) databases. The structural features of chloroplast genome maps were drawn using OGDRAWv1.2 (Potsdam-Golm, Germany) [28]. Codon usage was determined for all protein-coding genes. To examine the deviation in synonymous codon usage, the relative synonymous codon usage (RSCU) was calculated using MEGA6 software (Version 6.0, Jeddah, Saudi Arabia) [29]. Amino acid (AA) frequency was also calculated and expressed by the percentage of the codons encoding the same amino acid divided by the total number of codons. The final chloroplast genomic sequences have been submitted to GenBank under accession numbers MK209001 and MK209002 for K. galanga and K. elegans, respectively.

3.4. SSRs and Long Repeat Structure

SSRs were identified using MIcroSAtellite (MISA) [30]. The parameters for SSRs were adjusted for identification of perfect mono-, di-, tri-, tetra-, pena- and hexanucleotide motifs with a minimum of 8, 5, 4, 3, 3 and 3 repeats, respectively. The online REPuter software was used to identify and locate forward, palindrome, reverse and complement repeat sequences with repeat sizes ≥30 bp and sequences identity ≥90% [31].

3.5. Comparative and Divergence Analysis of Chloroplast Genomes of K. galanga and K. elegans

The complete chloroplast genome of K. galanga was employed as a reference and was compared with the chloroplast genomes of K. elegans, Alpinia zerumbet (JX088668), Curcuma flaviflora (KR967361) and Zingiber spectabile (JX088661), the last three of which were obtained from GenBank, using mVISTA program (http://genome.lbl.gov/vista/mvista/about.shtml) in the Shuffle-LAGAN mode [32]. To calculate nucleotide variability (Pi) between K. galanga and K. elegans chloroplast genomes, sliding window analysis was performed using DnaSP version 5.1 software [33] with window length of 600 bp and the step size of 200 bp.
The complete chloroplast sequences of K. galanga and K. elegans were also aligned using MUMmer software (Maryland, USA) [34] and adjusted manually where necessary using Se-Al 2.0 [35]. The single nucleotide polymorphisms (SNPs) and insertion/deletions (indels) were recorded separately as well as their locations in the chloroplast genome.

3.6. Phylogenetic Analysis

A molecular phylogenetic tree was constructed using SNP arrays from 11 species including K. galanga and K. elegans. Among these 11 species, nine complete chloroplast genome sequences were downloaded from NCBI: A. zerumbet (JX088668), C. flaviflora (KR967361), Z. spectabile (JX088661), C. roscoeana (NC_022928.1), Alpinia oxyphylla (NC_035895.1), Amomum kravanh (NC_036935.1), Amomum compactum (NC_036992.1), Costus pulverulentus (KF601573) and Canna indica (KF601570). K. galanga chloroplast genome was used as reference. Costus pulverulentus and Canna indica were set as outgroups of the family Zingiberaceae. Firstly, using MUMmer software [34], each chloroplast genome above was compared globally with the reference genome and the difference between each chloroplast genome and the reference genome found and preliminary filtering performed to detect the potential SNP sites. Secondly, the sequence of 100 bp on each side of the reference sequence SNP site was extracted and the extracted sequence and assembly results were compared using the BLAT software [36,37] to verify the SNP site. If the length of the alignment is less than 101 bp, it is considered to be a non-trusted SNP and is removed; if compared several times, the SNP that is considered to be a duplicate region and will also be removed; and finally a reliable SNP will be obtained. Thirdly, for each chloroplast genome, all SNPs are connected in the same order to obtain a sequence in FASTA format. Multiple FASTA format sequences alignments were carried out using ClustalX version 1.81 [38]. To examine the phylogenetic applications of rapidly evolving SNP markers, the maximum likelihood (ML) and maximum parsimony (MP) methods with 1000 bootstrap replicates were employed to construct phylogenetic trees using MEGA6 software, respectively [29].

3.7. RNA Editing Analysis

Thirty-four protein-coding genes of K. galanga and K. elegans chloroplast genomes were used to predict potential RNA editing sites using the online program Predictive RNA Editor for Plants (PREP) suite (http://prep.unl.edu/) with a cutoff value of 0.8 (Bielefeld, Germany) [39].

4. Discussion

In this study, we obtained the complete chloroplast genomes of K. galanga and K. elegans by using Illumina and PacBio sequencing, ranging from 163.5-163.8 kb in length. Both chloroplast genomes exhibit a typical quadripartite structure, as reported for other Zingiberaceae species, such as A. oxyphylla, A. zerumbet, C. flaviflora, Z. spectabile, C. roscoeana, A. compactum and A. kravanh [12]. Both genomes encode about 111-113 genes, including 79 protein-coding genes, 4 rRNA genes as well as 28 and 30 tRNA genes distributed throughout their genomes, respectively. This conformed with the protein-coding genes found in other Zingiberaceae members [12].
The molecular markers obtained from chloroplast genome sequences such as highly variable sequences, SSRs, SNPs and indels are useful tools in research. In Camellia species, 1.5% high divergent sequences were used for phylogenetics, taxonomy and species identification [40]. In this study, eight highly variable regions had been detected between K. galanga and K. elegans chloroplast genomes, including trnS-trnG, rps12-clpP, psbT-psbN, ycf1-ndhF, ndhF-rpl32, psaC-ndhE, ccsA-ndhD and rpl2-trnH (Figure 7A). As our results displayed, most of them occurred in the LSC and SSC regions but only one in the IR regions. Among these highly variable regions, ndhF-rpl32, ccsA-ndhD, trnS-trnG, psbT-psbN and psaC-ndhE had been reported as highly variable regions in many species, such as Papaver [5], Machilus [41], Citrullus [42], Fagopyrum [43], Citrus [44] and Oryza [45]. In addition, the ndhF-rpl32, ccsA-ndhD and trnS-trnG regions had been used as molecular markers for phylogenetic analysis, to resolve origin problems and phylogeographic studies [13,41,42,43,45]. Therefore, these eight highly variable regions could serve to enrich the molecular marker resources of genus Kaempferia in studies of phylogeny, evolution and species identification.
Besides the highly variable regions, we were able to retrieve SSRs and long repeats. Of the 240 SSRs identified in K. galanga, 64.58% (155 SSRs) were located in the LSC region, 16.66% (49 SSRs) in the SSC region and 18.75% (45 SSRs) in the IR regions. In contrast, out of the 248 SSRs identified in K. elegans, 64.11% (159 SSRs) were present in the LSC region, 16.93% (42 SSRs) in the SSC region and the remaining 19.75% (49 SSRs) located in the IR regions, as reported in other plants like A. kravanh [12], Talinum paniculatum [20] and Oryza minuta [45]. Among the SSRs types, the most abundant was found to be mononucleotides in both K. galanga and K. elegans (Figure 3). These findings were in agreement with results from previous studies in A. kravanh [12], T. paniculatum [20] and buckwheat species [43] but were different from O. minuta which possessed a majority of dinucleotide repeat motif SSRs [45]. AT/AT (12.5%) was the most frequent dinucleotide motifs followed by AAAT/ATTT in both K. galanga and K. elegans, respectively (Figure 3). These SSRs and long repeats identified in our study could be useful in molecular studies, such as genetic diversity and phylogenetic relationship analysis, species identification and evolution studies [7,12,21,40,43].
In the present study, we also identified 536 SNPs and 107 indels between the two Kaempferia species (Tables S4 and S5). From the SNPs results, 536 nucleotide substitutions were detected between K. galanga and K. elegans chloroplast genomes. It indicated that the nucleotide substitution events in the chloroplast genomes of Kaempferia species were more than that between species of Oryza, Machilus, cultivated Fagopyrum, Citrus and Panax but less than species of Solanum and wild Fagopyrum. Comparative analysis of chloroplast genomes found 159 SNPs between Oryza nivara and O. sativa [46], 231 SNPs between Machilus yunnanensis and M. balansae [41], 317 SNPs between cultivated species Fagopyrum dibotrys and F. tataricum [43], 330 SNPs between Citrus sinensis and C. aurantiifolia [44], 464 SNPs between Panax notoginseng and P. ginseng [47], 591 SNPs between Solanum tuberosum and S. bulbocastanum [48], 6260 SNPs between wild species F. luojishanense and F. esculentum [43]. Out of the 107 indels found between K. galanga and K. elegans chloroplast genomes, two longest intergenic sequences petA-psbJ and atpH-atpI were detected. The petA-psbJ and partial psbA-trnH spacer sequences can be used for species identification of most Kaempferia and outgroup species [4]. Similarly, a single large 241-bp deletion in S. tuberosum clearly discriminated a cultivated potato from the wild potato species S. bulbocastanum [48]. The indels and SNPs of 12 Triticeae species chloroplasts were used to estimate wheat, barley, rye and their relatives evolution [49]. These indels and SNPs found in our study could be useful in phylogenetic analysis, species identification and evolutionary studies as well as the 65 indels detected between M. yunnanensis and M. balansae [41], 156 indels between P. notoginseng and P. ginseng [47] and 53 indels between Aconitum pseudolaeve and A. longecassidatum [50].
The chloroplast genome sequences provide a useful genomic resource for phylogenetic studies and many studies have successfully used protein-coding sequences or whole chloroplast genome sequences in these analyses [7,12,20,43]. Specifically, the chloroplast psbA-trnH and partial petA-psbJ sequences and matK gene had been utilized in Zingiberaceae species phylogenetic studies before [4,51]. In this study, we constructed phylogenetic trees using ML and MP methods based on SNPs commonly present in the chloroplast genomes of eleven species, including two Kaempferia species from the current study. Our phylogenetic analysis clearly revealed that the two Kaempferia species clustered together, with bootstrap values of 100%, as well as Amomum and Alpinia species, which segregated in two sister clades (Figure 8). In a previous study, a phylogenetic tree constructed by using whole chloroplast genome sequences strongly supported the position in the Zingiberaceae of A. kravanh as a sister of the closely related species A. zerumbet [12]. Our phylogenetic trees using SNPs were in broad agreement with the previous study [12]. Therefore, phylogenetic analysis using SNPs among chloroplast genome sequences could provide useful information for revealing relationships among Zingiberaceae species.
In conclusion, we assembled and analyzed the complete chloroplast genomes of K. galanga and K. elegans and compared them with other Zingiberaceae species for the first time. The chloroplast genomes organization, gene order, GC content and codon usage of the two Kaempferia species showed high similarity. The location and distribution of SSRs and long repeat sequences were determined. Eight highly variable regions between the two Kaempferia species were identified and 643 mutation events, including 536 SNPs and 107 indels, were accurately located. Sequence divergences of chloroplast genomes were also calculated for the two Kaempferia species and related Zingiberaceae species. The phylogenetic analysis based on SNPs among eleven species strongly supported that K. galanga and K. elegans formed a cluster within Zingiberaceae. Our results provided insights into the characteristics of the entire K. galanga and K. elegans chloroplast genomes and the phylogenetic relationships within Zingiberaceae species.

Supplementary Materials

The Supplementary Materials are available online. Table S1: Codon Usage for K. galanga (Kg) and K. elegans (Ke), Table S2: SSRs, Table S3: Long repeats, Table S4: SNPs detected between K. galanga and K. elegans chloroplast genomes, Table S5: InDels detected between K. galanga and K. elegans chloroplast genomes, Table S6: RNA editing results, Supplementary file1: Genes distribution in the K. galanga chloroplast genome, Supplementary file2: Genes distribution in the K. elegans chloroplast genome.

Author Contributions

D.-M.L. designed the experiments, performed the experiments, collected plant materials, carried out sequence analysis and draft the manuscript. C.-Y.Z., X.-F.L. and D.-M.L. identified and collected plant materials. All authors contributed to the experiments and approved the final manuscript.

Funding

This research was funded by Guangzhou Municipal Science and Technology Project (No. 201607010101), Guangdong Science and Technology Project (No.2015A020209078, 2015B070701016) and the National Natural Science Foundation of China (31501788).

Conflicts of Interest

The authors declare that they have no conflict of interests.

Abbreviations

LSCLarge single copy
SSCSmall single copy
IRInverted repeat
tRNATransfer RNA
rRNARibosomal RNA
SSRsSimple Sequence Repeats
SNPsSingle-Nucleotide Polymorphisms
indelsinsertion/deletions

References

  1. Wu, D.; Larsen, K. Zingiberaceae. Flora China 2000, 24, 322–377. [Google Scholar]
  2. Wu, D.; Liu, N.; Ye, Y. The Zingiberaceae resources in China; Huazhong University of Science and Technology University Press: Wuhan, China, 2016; Volume 1, pp. 107–108. [Google Scholar]
  3. Branney, T.M.E. Hardy Gingers: Including Hedychium, Roscoea and Zingiber; Timber Press, Inc.: Portland, OR, USA, 2005; pp. 181–187. [Google Scholar]
  4. Techaprasan, J.; Klinbunga, S.; Ngamriabsakul, C.; Jenjittikul, T. Genetic variation of Kaempferia (Zingiberaceae) in Thailand based on chloroplast DNA (psbA-trnH and petA-psbJ) sequences. Genet. Mol. Res. 2010, 9, 1957–1973. [Google Scholar] [CrossRef] [PubMed]
  5. Zhou, J.; Cui, Y.; Chen, X.; Li, Y.; Xu, Z.; Duan, B.; Li, Y.; Song, J.; Yao, H. Complete chloroplast genomes of Papaver rhoeas and Papaver orientale: Molecular structures, comparative analysis and phylogenetic analysis. Molecules 2018, 23, 437. [Google Scholar] [CrossRef] [PubMed]
  6. Jiang, D.; Zhao, Z.; Zhang, T.; Zhong, W.; Liu, C.; Yuan, Q.; Huang, L. The chloroplast genome sequence of Scutellaria baicalensis provides insight into intraspecific and interspecific chloroplast genome diversity in Scutellaria. Genes 2017, 8, 227. [Google Scholar] [CrossRef]
  7. 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. [Google Scholar] [CrossRef] [PubMed]
  8. Wang, W.; Yu, H.; Wang, J.; Lei, W.; Gao, J.; Qiu, X.; Wang, J. The complete chloroplast genome sequences of the medicinal plant Forsythia suspense (Oleaceae). Int. J. Mol. Sci. 2017, 18, 2288. [Google Scholar] [CrossRef] [PubMed]
  9. Wicke, S.; Schneeweiss, G.M.; DePamphilis, 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]
  10. 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]
  11. Brunkard, J.O.; Runkel, A.M.; Zambryski, P.C. Chloroplast extend stromules independently and in response to internal redox signals. Proc. Natl. Acad. Sci. USA 2015, 112, 10044–10049. [Google Scholar] [CrossRef]
  12. Wu, M.; Li, Q.; Hu, Z.; Li, X.; Chen, S. The complete Amomum kravanh chloroplast genome sequence and phylogenetic analysis of the commelinids. Molecules 2017, 22, 1875. [Google Scholar] [CrossRef]
  13. 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]
  14. Li, Y.; Zhou, J.G.; Chen, X.L.; Cui, Y.X.; Xu, Z.C.; Li, Y.H.; Song, J.Y.; Duan, B.Z.; Yao, H. Gene losses and partial deletion of small single-copy regions of the chloroplast genomes of two hemiparasitic Taxillus species. Sci. Rep. 2017, 7, 12834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Lin, M.; Qi, X.; Chen, J.; Sun, L.; Zhong, Y.; Fang, J.; Hu, C. The complete chloroplast genome sequence of Actinidia arguta using the PacBio RSII platform. PLoS ONE 2018, 13, e0197393. [Google Scholar]
  16. Eid, J.; Fehr, A.; Gray, J.; Luong, K.; Lyle, J.; Otto, G.; Peluso, P.; Rank, D.; Baybayan, P.; Bettman, B.; et al. Real-time DNA sequencing from single polymerase molecules. Science 2009, 323, 133–138. [Google Scholar] [CrossRef]
  17. Ferrarini, M.; Moretto, M.; Ward, J.A.; Surbanovski, N.; Stevanovic, V.; Giongo, L.; Viola, R.; Cavalieri, D.; Velasco, R.; Cestaro, A.; et al. An evaluation of the PacBio RS platform for sequencing and de novo assembly of a chloroplast genome. BMC Genomics 2013, 14, 670. [Google Scholar] [CrossRef] [PubMed]
  18. Wu, Z.; Gui, S.; Guan, Z.; Pan, L.; Wang, S.; Ke, W.; Liang, D.; Ding, Y. A precise chloroplast genome of Nelumbo nucifera (Nelumbonaceae) evaluated with Sanger, Illumina MiSeq and PacBio RS II sequencing platforms: Insight into the plastid evolution of basal eudicots. BMC Plant Biol. 2014, 14, 289. [Google Scholar] [CrossRef] [PubMed]
  19. Shap, P.M.; Li, W.H. The codon adaptation index- a measure of directional synonymous codon usage bias and its potential applications. Nucleic Acids Res. 1987, 15, 1281–1295. [Google Scholar]
  20. Liu, X.; Li, Y.; Yang, H.; Zhou, B. Chloroplast genome of the folk medicine and vegetable plant Talinum paniculatum (Jacq.) Gaertn.: Gene organization, comparative and phylogenetic analysis. Molecules 2018, 23, 857. [Google Scholar]
  21. Li, X.; Hu, Z.; Lin, X.; Li, Q.; Gao, H.; Luo, G.; Chen, S. High-throughput pyrosequencing of the complete chloroplast genome of Magnolia officinalis and its application in species identification. Acta Pharm. Sin. 2012, 47, 124–130. [Google Scholar]
  22. 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-end de novo assembler. Gigascience 2012, 1, 18. [Google Scholar] [CrossRef]
  23. Kearse, M.; Moir, R.; Wilson, A.; Stoneshavas, S.; Cheung, M.; Sturrock, S.; Buxton, S.; Cooper, A.; Markowitz, S.; Duran, C.; et al. Geneious Basic: An integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 2012, 28, 1647–1649. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Chaisson, M.J.; Tesler, G. Mapping single molecule sequencing reads using basic local alignment with successive refinement (BLASR): Application and theory. BMC Bioinform. 2012, 13, 238. [Google Scholar] [CrossRef] [PubMed]
  25. Denisov, G.; Walenz, B.; Halpern, A.L.; Miller, J.; Axerlrod, N.; Levy, S.; Sutton, G. Consensus generation and variant detection by celera assembler. Bioinformatics 2008, 24, 1035–1040. [Google Scholar] [CrossRef]
  26. 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] [Green Version]
  27. Lowe, T.M.; Chan, P.P. tRNAscan-SE On-line: Search and Contextual Analysis of Transfer RNA Genes. Nucleic Acids Res. 2016, 44, W54–W57. [Google Scholar] [CrossRef] [PubMed]
  28. Lohse, M.; Drechsel, O.; Kahlau, S.; Bock, R. Organellar Genome DRAW—A suite of tools for generating physical maps of plastid and mitochondrial genomes and visualizing expression data sets. Nucleic Acids Res. 2013, 41, W575–W581. [Google Scholar] [CrossRef] [PubMed]
  29. 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]
  30. MISA-Microsatellite Identification Tool. Available online: http://pgrc.ipk-gatersleben.de/misa/ (accessed on 20 September 2017).
  31. Kurtz, S.; Choudhuri, J.V.; Ohlebusch, E.; Schleiermacher, C.; Stoye, J.; Giegerich, R. REPuter: The manifold applications of repeat analysis on a genomic scale. Nucleic Acids Res. 2001, 29, 4633–4642. [Google Scholar] [CrossRef]
  32. 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]
  33. Librado, P.; Rozas, J. DnaSP v5: A software for comprehensive analysis of DNA polymorphism data. Bioinformatics 2009, 25, 1451–1452. [Google Scholar] [CrossRef]
  34. Marcais, G.; Delcher, A.L.; Phillippy, A.M.; Coston, R.; Salzberg, S.L.; Zimin, A. MUMmer4: A fast and versatile genome alignment system. PLoS Comput. Biol. 2018, 14, e1005944. [Google Scholar] [CrossRef] [PubMed]
  35. Rambaut, A. Se-Al: Sequence Alignment Editor. Version 2.0. Available online: http://tree.bio.ed.ac.uk/software (accessed on 30 September 2017).
  36. Kent, W.J. BLAT—The BLAST-like alignment tool. Genome Res. 2002, 12, 656–664. [Google Scholar] [CrossRef]
  37. Bhagwat, M.; Young, L.; Robison, R.R. Using BLAT to find sequence similarity in closely related genomes. Curr. Protoc. Bioinform. 2012, 010, Unit10.8. [Google Scholar] [CrossRef]
  38. Thompson, J.D.; Gibson, T.J.; Plewniak, F.; Jeanmougin, F.; Higgins, D.G. The CLUSTAL_X windows interface: Flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Res. 1997, 25, 4876–4882. [Google Scholar] [CrossRef]
  39. Mower, J.P. The PREP Suite: Predictive RNA editors for plant mitochondrial genes, chloroplast genes and user-defined alignments. Nucleic Acids Res. 2009, 37, W253–W259. [Google Scholar] [CrossRef] [PubMed]
  40. Huang, H.; Shi, C.; Liu, Y.; Mao, S.Y.; Gao, L.Z. Thirteen Camellia chloroplast genome sequences determined by high-throughput sequencing: Genome structure and phylogenetic relationships. BMC Evol. Biol. 2014, 14, 151. [Google Scholar] [CrossRef]
  41. 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]
  42. Chomicki, G.; Renner, S.S. Watermelon origin solved with molecular phylogenetics including Linnaen material: Another example of museomics. New Phytol. 2015, 205, 526–532. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, C.L.; Ding, M.Q.; Zou, C.Y.; Zhu, X.M.; Tang, Y.; Zhou, M.L.; Shao, J.R. Comparative analysis of four buckwheat species based on morphology and complete chloroplast genome sequences. Sci. Rep. 2017, 7, 6514. [Google Scholar] [CrossRef]
  44. Su, H.J.; Hogenhout, S.A.; Al-Sadi, A.M.; Kuo, C.H. Complete chloroplast genome sequence of Omani lime (Citrus aurantiifolia) and comparative analysis within the rosids. PLoS ONE 2014, 9, e113049. [Google Scholar] [CrossRef] [PubMed]
  45. Asaf, S.; Waqas, M.; Khan, A.L.; Khan, M.A.; Kang, S.M.; Imran, Q.M.; Shahzad, R.; Bilal, S.; Yun, B.W.; Lee, I.J. The complete chloroplast genome of wild rice (Oryza minuta) and its comparison to related species. Front. Plant Sci. 2017, 8, 304. [Google Scholar] [CrossRef]
  46. Shahid Masood, M.; Nishikawa, T.; Fukuoka, S.; Njenga, P.K.; Tsudzuki, T.; Kadowaki, K. The complete nucleotide sequence of wild rice (Oryza nivara) chloroplast genome: First genome wide comparative sequence analysis of wild and cultivated rice. Gene 2004, 340, 133–139. [Google Scholar] [CrossRef] [PubMed]
  47. Dong, W.; Liu, H.; Xu, C.; Zuo, Y.; Chen, Z.; Zhou, S. A chloroplast genomic strategy for designing taxon specific DNA mini-barcodes: A case study on ginsengs. BMC Genet. 2014, 15, 138. [Google Scholar] [CrossRef]
  48. Chung, H.J.; Jung, J.D.; Park, H.W.; Kim, J.H.; Cha, H.W.; Min, S.R.; Jeong, W.J.; Liu, J.R. The complete chloroplast genome sequences of Solanum tuberosum and comparative analysis with Solanaceae species identified the presence of a 241-bp deletion in cultivated potato chloroplast DNA sequence. Plant Cell Rep. 2006, 25, 1369–1379. [Google Scholar] [CrossRef] [PubMed]
  49. Middleton, C.P.; Senerchia, N.; Stein, N.; Akhunov, E.D.; Keller, B.; Wicker, T.; Kilian, B. Sequencing of chloroplast genomes from wheat, barley, rye and their relatives provides a detailed insight into the evolution of the Triticeae tribe. PLoS ONE 2014, 9, e85761. [Google Scholar] [CrossRef] [PubMed]
  50. Park, I.; Yang, S.; Choi, G.; Kim, W.J.; Moon, B.C. The complete chloroplast genome sequences of Aconitum pseudolaeve and Aconitum longecassidatum and development of molecular markers for distinguishing species in the Aconitum subgenus Lycoctonum. Molecules 2017, 22, 2012. [Google Scholar] [CrossRef]
  51. Kress, W.J.; Prince, L.M.; Williams, K.J. The phylogeny and a new classification of the gingers (Zingiberaceae) evidence from molecular data. Am. J. Bot. 2002, 89, 1682–1696. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Samples of the compounds are available from the authors.
Figure 1. Circular gene map of chloroplast genomes of two Kaempferia species. The gray arrowheads indicate the direction of the genes. Genes shown inside the circle are transcribed clockwise and those outside are transcribed counterclockwise. Different genes are color coded. The innermost darker gray corresponds to GC content, whereas the lighter gray corresponds to AT content. IR, inverted repeat; LSC, large single copy region; SSC, small single copy region.
Figure 1. Circular gene map of chloroplast genomes of two Kaempferia species. The gray arrowheads indicate the direction of the genes. Genes shown inside the circle are transcribed clockwise and those outside are transcribed counterclockwise. Different genes are color coded. The innermost darker gray corresponds to GC content, whereas the lighter gray corresponds to AT content. IR, inverted repeat; LSC, large single copy region; SSC, small single copy region.
Molecules 24 00474 g001
Figure 2. Amino acid frequencies in K. galanga and K. elegans protein-coding sequences.
Figure 2. Amino acid frequencies in K. galanga and K. elegans protein-coding sequences.
Molecules 24 00474 g002
Figure 3. Distribution of SSRs in the chloroplast genomes of K. galanga and K. elegans. (A) Number of different SSR types detected in the two Kaempferia species chloroplast genomes; (B) Frequency of identified SSR motifs in different repeat class types; (C) SSR distribution in different genomic regions of two Kaempferia species chloroplast genomes; (D) SSR distribution between coding and non-coding regions of two Kaempferia species chloroplast genomes.
Figure 3. Distribution of SSRs in the chloroplast genomes of K. galanga and K. elegans. (A) Number of different SSR types detected in the two Kaempferia species chloroplast genomes; (B) Frequency of identified SSR motifs in different repeat class types; (C) SSR distribution in different genomic regions of two Kaempferia species chloroplast genomes; (D) SSR distribution between coding and non-coding regions of two Kaempferia species chloroplast genomes.
Molecules 24 00474 g003aMolecules 24 00474 g003b
Figure 4. Analysis of long repeat sequences in the chloroplast genomes of K. galanga and K. elegans. (A) Frequency of long repeats types; (B) Frequency of long repeats by length.
Figure 4. Analysis of long repeat sequences in the chloroplast genomes of K. galanga and K. elegans. (A) Frequency of long repeats types; (B) Frequency of long repeats by length.
Molecules 24 00474 g004
Figure 5. Comparison of the borders of the LSC, SSC and IR regions among five Zingiberaceae chloroplast genomes. Ψ, pseudogenes. Boxes above the main line indicate the adjacent border genes. The figure is not to scale with respect to sequence length and only shows relative changes at or near the IR/SC borders.
Figure 5. Comparison of the borders of the LSC, SSC and IR regions among five Zingiberaceae chloroplast genomes. Ψ, pseudogenes. Boxes above the main line indicate the adjacent border genes. The figure is not to scale with respect to sequence length and only shows relative changes at or near the IR/SC borders.
Molecules 24 00474 g005
Figure 6. Comparison of five chloroplast genomes, with K. galanga as a reference using mVISTA alignment program. Gray arrows and thick black lines above the alignment indicate gene orientation. Purple bars represent exons, sky-blue bars represent transfer RNA (tRNA) and ribosomal RNA (rRNA), red bars represent non-coding sequences (CNS) and white peaks represent differences of chloroplast genomes. The y-axis represents the identity percentage ranging from 50 to 100%.
Figure 6. Comparison of five chloroplast genomes, with K. galanga as a reference using mVISTA alignment program. Gray arrows and thick black lines above the alignment indicate gene orientation. Purple bars represent exons, sky-blue bars represent transfer RNA (tRNA) and ribosomal RNA (rRNA), red bars represent non-coding sequences (CNS) and white peaks represent differences of chloroplast genomes. The y-axis represents the identity percentage ranging from 50 to 100%.
Molecules 24 00474 g006
Figure 7. Sliding window analysis of the whole chloroplast genomes. Window length: 800 bp; step size: 200 bp. X-axis:position of the midpoint of a window. Y-axis: nucleotide diversity of each window. (A) Pi between K. galanga and K. elegans. (B) Pi among two Kaempferia species, Alpinia zerumbet, Curcuma flaviflora and Zingiber spectabile.
Figure 7. Sliding window analysis of the whole chloroplast genomes. Window length: 800 bp; step size: 200 bp. X-axis:position of the midpoint of a window. Y-axis: nucleotide diversity of each window. (A) Pi between K. galanga and K. elegans. (B) Pi among two Kaempferia species, Alpinia zerumbet, Curcuma flaviflora and Zingiber spectabile.
Molecules 24 00474 g007
Figure 8. Phylogenetic trees constructed with SNPs from 11 species using maximum likelihood (ML, left) and maximum parsimony (MP, right) methods. Numbers at nodes on the tree indicate bootstrap values (>50%).
Figure 8. Phylogenetic trees constructed with SNPs from 11 species using maximum likelihood (ML, left) and maximum parsimony (MP, right) methods. Numbers at nodes on the tree indicate bootstrap values (>50%).
Molecules 24 00474 g008
Table 1. Features of the chloroplast genomes of K. galanga and K. elegans.
Table 1. Features of the chloroplast genomes of K. galanga and K. elegans.
SpeciesRegionsPositionsLength (bp)T/U (%)C (%)A (%)G (%)AT/U (%)
K. galangaGenome 163,81132.218.331.717.763.9
LSC 88,40533.717.332.416.466.1
IRa 29,79728.819.830.021.258.8
SSC 15,81234.515.535.913.970.5
IRb 29,79728.819.830.021.258.8
Protein coding genes 83,17231.517.231.419.763.0
1st position27,72423.918.231.426.355.4
2nd position27,72432.420.030.117.362.6
3rd position27,72438.313.332.815.571.1
tRNA 2,87024.923.622.029.347.0
rRNA 9,04618.623.626.131.544.8
K. elegansGenome 163,55532.218.331.717.763.9
LSC 88,02033.717.432.416.566.1
IRa 29,77328.819.830.121.358.9
SSC 15,98934.615.536.113.870.6
IRb 29,77328.819.830.121.358.9
Protein coding genes 79,11731.617.331.219.962.8
1st position26,37235.014.331.918.866.9
2nd position26,37226.619.430.123.956.7
3rd position26,37233.318.231.417.164.7
tRNA 2,85224.923.722.029.446.9
rRNA 9,04618.723.626.131.544.8
Table 2. Genes present in the chloroplast genomes of K. galanga and K. elegans.
Table 2. Genes present in the chloroplast genomes of K. galanga and K. elegans.
CategoryGene Name
Photosystem IpsaA, psaB, psaC, psaI, psaJ
Photosystem IIpsbA, psbB, psbC, psbD, psbE, psbF, psbH, psbI, psbJ, psbK, psbL, psbM, psbN, psbT, lhbA
Cytochrome b/fpetA, petB *, petD *, petG, petL, petN
ATP synthaseatpA, atpB, atpE, atpF *, atpH, atpI
NADH dehydrogenasendhA *, ndhB(×2) *, ndhC, ndhD, ndhE, ndhF, ndhG, ndhH, ndhI, ndhJ, ndhK
RubiscorbcL
RNA polymeraserpoA, rpoB, rpoC1 *, rpoC2
Large subunit ribosomal proteinsrpl2(×2) *, rpl14, rpl16 *, rpl20, rpl22, rpl23(×2), rpl32, rpl33, rpl36
Small subunit ribosomal proteinsrps2, rps3, rps4, rps7(×2), rps8, rps11, rps12(×2) *, rps14, rps15, rps16 *, rps18, rps19(×2)
Other proteinsaccD, ccsA, cemA, clpP *, infA, matK
Proteins of unknown functionycf1(×2), ycf2(×2), ycf3*, ycf4
Ribosomal RNAsrrn4.5(×2), rrn5(×2), rrn16(×2), rrn23(×2)
Transfer RNAstrnA-UGC (×2) *, trnC-GCA, trnD-GUC, trnE-UUC, trnF-GAA, trnfM-CAU, trnG-GCC (Kg×2, Ke) **, trnG-UCC, trnH-GUG (×2), trnI-CAU (×2), trnI-GAU (×2) *, trnK-UUU (×2) *, trnL-CAA (×2), trnL-UAA (×2) *, trnL-UAG, trnM-CAU, trnN-GUU (×2), trnP-UGG, trnQ-UUG, trnR-ACG (×2), trnR-UCU, trnS-GCU (Kg×2, Ke), trnS-UGA, trnT-GGU (Kg×2, Ke), trnV-GAC (×2), trnV-UAC (×2) *, trnW-CCA, trnY-GUA, trnS-GGA (Ke), trnT-UGU (Ke)
Kg: K. galanga; Ke: K. elegans; ×2: Gene with two copies; *: Genes containing introns both in K. galanga and K. elegans; **: Genes containing introns only in K. galanga.
Table 3. Genes with introns in the chloroplast genomes of K. galanga and K. elegans, including the exon and intron lengths.
Table 3. Genes with introns in the chloroplast genomes of K. galanga and K. elegans, including the exon and intron lengths.
SpeciesGeneLocationExon I (bp)Intron I (bp)Exon II (bp)Intron II (bp)Exon III (bp)
K. galangatrnA-UGCIR3880135
trnG-GCCLSC1471148
trnI-GAUIR4293535
trnK-UUULSC35264637
trnL-UAALSC3553650
trnV-UACLSC3759838
rps12 *LSC/IR114-23154027
rps16LSC21274940
rpl2IR443650315
rpl16LSC40210589
petBLSC6783642
petDLSC8740481
atpFLSC425816145
ndhASSC5181083562
ndhBIR778673782
rpoC1LSC1632726432
clpPLSC25263630685660
ycf3LSC153794228714132
K. eleganstrnA-UGCIR3880135
trnI-GAUIR4293535
trnK-UUULSC35266337
trnL-UAALSC3553550
trnV-UACLSC3759838
rps12 *LSC/IR114-23154027
rps16LSC21272940
rpl2IR432659387
rpl16LSC40210569
petBLSC6784642
petDLSC8741481
atpFLSC411816144
ndhASSC5401079552
ndhBIR756700777
rpoC1LSC1632728432
clpPLSC25563629185469
ycf3LSC153794228723132
* The rps12 gene is divided into 5′-rps12 in the LSC region and 3′-rps12 in the IR region.

Share and Cite

MDPI and ACS Style

Li, D.-M.; Zhao, C.-Y.; Liu, X.-F. Complete Chloroplast Genome Sequences of Kaempferia Galanga and Kaempferia Elegans: Molecular Structures and Comparative Analysis. Molecules 2019, 24, 474. https://doi.org/10.3390/molecules24030474

AMA Style

Li D-M, Zhao C-Y, Liu X-F. Complete Chloroplast Genome Sequences of Kaempferia Galanga and Kaempferia Elegans: Molecular Structures and Comparative Analysis. Molecules. 2019; 24(3):474. https://doi.org/10.3390/molecules24030474

Chicago/Turabian Style

Li, Dong-Mei, Chao-Yi Zhao, and Xiao-Fei Liu. 2019. "Complete Chloroplast Genome Sequences of Kaempferia Galanga and Kaempferia Elegans: Molecular Structures and Comparative Analysis" Molecules 24, no. 3: 474. https://doi.org/10.3390/molecules24030474

Article Metrics

Back to TopTop