Comparative analyses of Agrimonia complete chloroplast genomes with newly assembled chloroplast genomes of A. coreana and A. nipponica (Rosaceae)

Article information

Korean J. Pl. Taxon. 2024;54(1):47-71
Publication date (electronic) : 2024 March 31
doi : https://doi.org/10.11110/kjpt.2024.54.1.47
1Infoboss Inc. and Infoboss Research Center, Seoul 07766, Korea
2Department of Biology, Daejeon University, Daejeon 34520, Korea
Corresponding author Jongsun PARK, E-mail: starflr@infoboss.co.kr
Received 2023 February 28; Revised 2024 March 27; Accepted 2024 March 28.

Abstract

The genus Agrimonia L. (1753) is a small group consisting of 19 species and three varieties. We completed two Agrimonia chloroplast genomes, Agrimonia coreana Nakai and Agrimonia nipponica Koidz, collected in Korea and conducted comprehensive analyses of six Agrimonia chloroplast genomes. Two of the Agrimonia chloroplasts genomes have a typical quadripartite structure with lengths ranging from 155,161 bp to 151,362 bp, similar to the remaining Agrimonia chloroplast genomes. High nucleotide diversity was detected in the ycf1 gene, which can serve as a molecular marker. Intraspecific variations of four A. pilosa chloroplast genomes and interspecific variations of three Agrimonia species revealed that KY419942 is distinct to the other three A. pilosa chloroplast genomes. A phylogenetic analysis demonstrated that KY419942 was clustered with A. nipponica, requiring further analyses to understand Agrimonia species in East Asia. A comparison of simple sequence repeats identified from the six Agrimonia chloroplast genomes suggests potential molecular markers to distinguish species or populations of the same species. Our results define the phylogenetic relationship of three Agrimonia species and provide insight into the intraspecific features of Agrimonia chloroplast genomes with which to gain a better understanding of the genetic structure of Agrimonia species.

INTRODUCTION

The subtribe Agrimoniinae (Rosaceae, Rosoideae), one of two members of the tribe Sanguisorbeae, displays polyploidy and morphological variations, both of which are good sources of information when attempting to understand the corresponding biogeographical patterns and endemism (Chung, 2008). This subtribe consists of the five genera, Agrimonia (19 species in Eurasia and South America), Hagenia (one species in east Africa), Leucosidea (one species in southern Africa), Spenceria (one species in central Asia), and Aremonia (one species in southern Europe) (Potter et al., 2007; Chung 2008), with all genera except Agrimonia being monotypic. Given this biased distribution of species, partial or complete chloroplast genomes for four out of the five genera except Aremonia have been determined thus far: two complete and one partial determination for Agrimonia pilosa Ledeb. (Zhang et al., 2017; Heo et al., 2020; Liu et al., 2020b), one complete determination for Agrimonia pilosa var. nepalensis (D. Don) Nakai (Yang et al., 2021), one complete and one partial determination for Hagenia abyssinica Willd (Gichira et al., 2017; Zhang et al., 2017), one partial determination for Leucosidea sericea Eckl. & Zeyh. (Zhang et al., 2017), and one partial determination for Spenceria ramalana Trimen (Zhang et al., 2017).

Agrimonia L. (1753), consisting of 19 species and three varieties, is characterized by two major morphological characteristics: interrupted pinnately compound leaves and bristly indumentum, within Rosaceae (Chung, 2008). Agrimonia species are usually distributed in temperate regions, including North America, Central America, the West Indies, southern South America, Europe, Asia, and southern Africa (Hutchinson, 1964; Skalický, 1973; Robertson, 1974; Murata and Umemoto, 1983; Chung and Kim, 2000; Kalkman, 2004; Chung, 2008). The base chromosome number is x = 7 (Darlington and Wylie, 1955) and the polyploidy levels of Agrimonia species range from tetraploid (2n = 4x = 28) to octaploid (2n = 8x = 56) (Chung, 2008). Various natural hybrids observed in Agrimonia hinder a clear delimitation of its species boundary (Wallroth, 1842; Skalický, 1962; Chung, 2008), indicating that alternative methods, such as the use of molecular markers, are required to understand the species boundary.

The Agrimonia pilosa complex is an excellent case for investigating molecular variation at the genomic level. The A. pilosa complex, with three morphologically similar species A. pilosa, A. coreana Nakai, and A. nipponica Koidz, is characterized by having relatively small fruits with spreading, erect, or connivent prickles (Chung and Kim, 2000; Chung, 2008). All of these species are distributed in East Asia with extensions to Eastern Europe. An examination of the intraspecific variations of chloroplast genomes together with their morphological characteristics would provide fundamental information about the levels and patterns of variations within and among closely related species, especially considering that data pertaining to four chloroplast genomes of A. pilosa are available in China and Korea (Zhang et al., 2017; Heo et al., 2020; Liu et al., 2020b). In addition, plants of the A. pilosa complex have been used to treat boils, eczema, taeniasis, abdominal pain, sore throat, headaches, and heat stroke in traditional medicine in China, Nepal, and Korea (World Health Organization, 1998). Careful and thorough investigations have also found that A. pilosa extracts have various useful effects, among them antiviral (Shin et al., 2010), whitening (Kim et al., 2011a; Kim et al., 2011b), antinociceptive (Park et al., 2012), antioxidant (Seo et al., 2008; Kim et al., 2011b; Chen and Kang, 2014), anticancer (Seo et al., 2008), anti-aging (Yoon et al., 2012), and anti-inflammatory (Jung et al., 2010) effects, suggesting the potential value of investigating the genetic background of Agrimonia species in Korea.

Owing to next-generation sequencing and third-generation sequencing technologies (Zhou et al., 2010; Roberts et al., 2013; Van Dijk et al., 2014; Deamer et al., 2016; Goodwin et al., 2016), nearly ten thousand chloroplast genomes have been sequenced thus far (Park et al., 2021d). These chloroplast genomes have been used for phylogenetic or phylogenomic analyses (Hassemer et al., 2019; Alzahrani et al., 2020; Liang et al., 2020; Chang et al., 2021), to identify the phylogenetic positions of new species candidates (Kim et al., 2019f; Oh et al., 2019a; Park et al., 2021b), to investigate intraspecific variations on chloroplast genomes (Li et al., 2018; Bum et al., 2020; Baek et al., 2021), and to develop useful molecular markers (Li et al., 2020a; Wang et al., 2020). Complete chloroplast genomes can be utilized to understand the relationships between morphological features and their phylogenetic positions. Examples of this possibility include Chenopodium album L. (Park et al., 2021d), Suaeda japonica Makino (Kim et al., 2020), the Allium genus (Xie et al., 2019), and the Fagopyrum genus (Wang et al., 2017). Due to the morphological complexity and the presence of intermediate forms in the A. pilosa complex, additional chloroplast genomes of A. coreana, and A. nipponica beyond the three chloroplast genomes of A. pilosa already reported (Zhang et al., 2017; Heo et al., 2020; Liu et al., 2020b) can be helpful to understand the relationships, morphological diversity and phylogenetic positions of this complex.

Here, we complete the two chloroplast genomes of A. coreana and A. nipponica and conduct comparative analyses of the six Agrimonia chloroplast genomes available to date. Our comprehensive analyses of nucleotide diversity, intraspecific and interspecific variations, and phylogeneticity in these genomes and comparisons of simple sequence repeats (SSRs) demonstrate that two groups of A. pilosa, one a sister to A. nipponica and the other a sister to A. coreana, reflect that A. coreana, A. nipponica, and A. pilosa have complex evolutionary histories. Also presented are the chromosome configurations of the three species. Further analyses of additional chloroplast genomes and the continuing whole genome sequencing of other Agrimonia species will be required to understand the precise evolutionary history of the Agrimonia genus.

MATERIALS AND METHODS

DNA extraction of two Agrimonia samples isolated in Korea

Two Agrimonia species, A. coreana and A. nipponica, were collected on the Korean Peninsula (Table 1). Both are native Korean plant species (Park et al., 2020a) collectable without special permission in Korea. The sampling process for Agrimonia was conducted while also adhering to all local, national, and international guidelines and laws. All vouchers of the two Agrimonia samples were deposited into the InfoBoss Cyber Herbarium (IN) (Table 1). Their total DNA was extracted from fresh leaves of the two samples using a DNeasy Plant Mini Kit (QIAGEN, Hilden, Germany).

List of six Agrimonia chloroplast genomes analyzed in this study.

Genome sequencing and de novo assembly of the two Agrimonia chloroplast genomes

Genome sequencing was performed with the NovaSeq 6000 system at Macrogen Inc. in Korea using the DNA extracted from the two Agrimonia species. De novo assembly was then done, with confirmation accomplished with Velvet v1.2.10 (Zerbino and Birney, 2008) after filtering raw reads using Trimmomatic v0.33 (Bolger et al., 2014). After obtaining the first draft of the chloroplast genome sequences, gaps were filled with GapCloser v1.12 (Zhao et al., 2011) and all bases from the assembled sequences were confirmed by checking each base in the alignment (tview mode in SAMtools v1.9) (Li et al., 2009) against the assembled chloroplast genome generated with BWA v0.7.17 (Li, 2013). All of these processes were conducted under the environment of Genome Information System (GeIS; http://geis.infoboss.co.kr/), which has been utilized in a range of studies (Kim et al., 2021b, 2021c; Park et al., 2021c, 2021h), including work on plant organelle genomes (Suh et al., 2021; Park et al., 2022a, 2022c; Yoo et al., 2023).

Chloroplast genome annotation

Geneious Prime 2020.2.4 (Biomatters Ltd., Auckland, New Zealand) was used for genome annotation of the two Agrimonia species based on the A. pilosa chloroplast genome (GenBank accession number: MT415946) (Heo et al., 2020) by transferring annotations while correcting exceptional cases, including missing start or stop codons. tRNA was predicted and confirmed based on predictions by tRNAscan-SE v2.0.6 (Lowe and Chan, 2016). Circular maps of the two Agrimonia chloroplasts were drawn using the OGDRAW v1.3.1 (Greiner et al., 2019).

Identification of sequence variations from complete chloroplast genomes

Single nucleotide polymorphisms (SNPs) and insertions and deletions (INDELs) were identified by means of a pair-wise alignment of the two selected chloroplast genomes done using MAFFT v7.450 (Katoh and Standley, 2013). When the numbers of INDELs were calculated, continuous INDEL bases were considered as one INDEL. When partial chloroplast genomes without one of the inverted repeat (IR) regions were used to identify sequence variations, SNPs and INDELs located in the IR regions were counted once more with the assumption that the IR region missing in the partial chloroplast genome is identical to the IR region in the partial chloroplast genomes. These methods have been utilized in previous studies (Kim et al., 2021a, 2021e; Park et al., 2021g).

Identification of SSRs

SSRs were identified on the chloroplast genome sequence using the pipeline of the SSR database (SSRDB; http://ssrdb.infoboss.co.kr/; Park et al., in preparation), as utilized in previous studies (Hong et al., 2022). Based on the conventional definition of an SSR on a chloroplast genome, monoSSR (1 bp) to hexaSSR (6 bp), the total length of the SSRs on the chloroplast genome exceeds 10 bp. Owing to the different criteria pertaining to SSRs on chloroplast genomes, we adopted the criteria used with the organelle genomes of Dysphania species (Kim et al., 2019d), Arabidopsis thaliana L. (Park et al., 2020c), Chenopodium album (Park et al., 2021d), Diarthron linifolium Turcz. (Kim et al., 2021d), Campsis grandiflora (Park and Xi 2022), and Rosa rugosa Thunb. (Park et al., 2020d), as follows; the monoSSR (unit sequence length of 1 bp) to hexaSSR (6 bp) are used as normal SSRs, and the heptaSSR (7 bp) to decaSSR (10 bp) are defined as the extended SSRs. Among the normal SSRs, pentaSSRs, and hexaSSRs with a unit sequence repeat number of 2 are classified as potential SSRs. Regional classifications of the chloroplast genome were conducted using the method described in the section above.

Comparison of SSRs identified from the six Agrimonia chloroplast genomes

SSRs identified from ten Agrimonia chloroplast genomes were compared based on their flanking sequences under the SSRDB environment (http://ssrdb.infoboss.co.kr/) (Park et al., in preparation). The pipeline of the SSR comparison implemented in the SSRDB relies on various organelle genome studies (Park et al., 2021i, 2022b; Kim et al., 2023) and was used with the following conditions: a cut-off e-value of 1e-10 and a maximum flanking sequence for the comparison of 60 bp.

Nucleotide diversity analysis

Nucleotide diversity was calculated using the method proposed by Nei and Li (Nei and Li, 1979) based on the multiple sequence alignment of Agrimonia chloroplast genomes using a Perl script used in previous studies (Lee et al., 2020; Choi et al., 2021). The window size was set to 500 bp and the step size was 200 bp when using the sliding-window method. The genomic coordination of each window was compared to the gene annotation of the chloroplast genome under the GenomeArchive (http://www.genomearchive.net/) environment for further analyses.

Construction of phylogenetic trees

Thirty-three whole chloroplast genomes of tribe Sanguisorbeae and two outgroup species, Rosa and Potentilla (Table 2), were aligned by MAFFT v7.450 (Katoh and Standley, 2013) and the alignment quality was checked manually. The maximum likelihood (ML) tree was reconstructed in IQ-TREE v1.6.6 (Nguyen et al., 2015). In the ML analysis, a heuristic search was used with nearest-neighbor interchange branch swapping, the GTR + F + R4 model, and uniform rates among sites. All other options used default settings. Bootstrap analyses with 1,000 pseudoreplicates were conducted with the same options. The posterior probability of each node was estimated by means of Bayesian inference (BI) using the MrBayes v3.2.6 (Huelsenbeck and Ronquist, 2001) plug-in implemented in Geneious Prime 2020.2.4 (Biomatters Ltd.). The HKY85 model with gamma rates was used as a molecular model. A Markov-chain Monte Carlo (MCMC) algorithm was employed for 1,100,000 generations, sampling trees every 200 generations, with four chains running simultaneously. Trees from the first 100,000 generations were discarded as burn-in.

List of intraspecific variations identified from various plant species isolated from Korea and China using pair-wise comparison method.

RESULTS AND DISCUSSION

Two complete Agrimonia chloroplast genomes, A. coreana and A. nipponica

Two Agrimonia species, A. coreana and A. nipponica, were sampled on the Korean Peninsula (Table 1) (Chung 2017; Park et al., 2020e). Chloroplast genomes of the two Agrimonia species were successfully completed, with corresponding coverage rates of 133.88x and 150.31x. It was thus found that the A. coreana chloroplast genome was 151,362 bp long (Fig. 1A), the longest among the six Agrimonia chloroplast genomes, while that of A. nipponica was 155,161 bp long (Fig. 1B), similar to the other Agrimonia chloroplast genomes (Table 1). A small single-copy (SSC) region of A. nipponica was the longest (19,825 bp) among the six Agrimonia chloroplast genomes, while the IR region of A. nipponica was the shortest (25,411 bp). A large single-copy (LSC) region of A. coreana was longest (84,597 bp) (Table 1). The two Agrimonia chloroplast genome sequences sequenced in this study can be accessed via accession numbers MZ604439 and MZ604440 in the NCBI GenBank.

Fig. 1.

Circular gene map of Agrimonia coreana and Agrimonia nipponica chloroplast genomes. A. Agrimonia coreana, B. Agrimonia nipponica. Genes shown outside are transcribed clockwise, and those inside the circle are transcribed counter-clockwise. Genes are color-coded to distinguish different functional groups. The dark grey and the light grey plots in the inner circle correspond to the GC content and AT content, respectively.

The overall GC contents of A. coreana and A. nipponica amounted to 36.9%, identical to the remaining Agrimonia chloroplast genomes, except for the partial chloroplast genome (GenBank accession number: KY419942) (Table 1). The GC contents of the LSC in both A. coreana and A. nipponica amounted to 34.8%, whereas the corresponding SSC and IR rates of A. coreana were 30.3% and 42.6%, lower than those of A. nipponica (Table 1), caused by different range of the IR region.

Both of the Agrimonia chloroplast genomes contained 130 genes, consisting of 84 protein-coding genes (PCGs), eight ribosomal RNAs (rRNAs), 37 transfer RNAs (tRNAs), and one pseudogene (ycf1) (Online Supplementary Material S1).

The 17 genes duplicated in IR regions consisted of seven PCGs (rpl2, rpl23, ycf2, ndhB, rps7, rps12, and ycf1), four rRNAs (rrn16, rrn23, rrn4.5, and rrn5), and seven tRNAs (trnI-CAU, trnL-CAA, trnV-GAC, trnI-GAU, trnA-UGC, trnR-ACG, and trnN-GUU). The ycf1 gene in the IRb region was considered as a pseudogene because the putative start methionine was inside ndhF. However, ycf1 in A. pilosa var. nepalensis was annotated as a normal gene by the start methionine located at 89-bp inside ndhF, addressing the need for additional research on the correct position of ycf1 in the IR region. This caused one additional PCG in the A. pilosa var. nepalensis chloroplast genome (Table 1). In addition, the annotation of the three genes, rpl16, petD, and petB, in the A. pilosa var. nepalensis chloroplast indicated that they had no intron; however, those in the remaining Agrimonia chloroplast genomes consisted of two exons. The annotation of these genes should also be corrected for a more accurate analysis.

In the six Agrimonia chloroplast genomes, including the two Agrimonia chloroplast genomes sequenced in this study after correction of the corrected three genes in A. pilosa var. nepalensis, eight PCGs (rps16, rpoC1, petB, petD, rpl16, ndhB, ndhA, and rpl2) contain one intron and clpP, ycf3, and rps12 have two introns, which are conserved across Agrimonia chloroplast genomes. The gene structure of the six Agrimonia chloroplast genomes is well conserved.

Nucleotide diversity of six Agrimonia chloroplast genomes

To investigate the landscape of nucleotide diversity in six Agrimonia chloroplast genomes, nucleotide diversity from the alignment of the six Agrimonia chloroplast genomes was assessed (see Materials and Methods). The LSC and SSC regions displayed a high level of nucleotide diversity, whereas two IR regions exhibited a low level, congruent to those of other chloroplast genomes (Du et al., 2017; Jiang et al., 2017; Li et al., 2018, 2019; Liu et al., 2020a; Loeuille et al., 2021). Most highly diverse regions were intergenic regions, and some regions contained tRNA genes, specifically trnQ, trnS, trnT, trnS, and trnT (Fig. 2), which is a reasonable observation in that most genes in chloroplast genomes are essential for photosynthesis and self-replication (Online Supplementary Material S1). One exceptional case is ycf1, displaying high nucleotide diversity throughout the coding genes in comparison to other PCGs (Online Supplementary Material S2). In detail, 29 non-synonymous SNPs, 15 synonymous SNPs, and 9-bp insertions were found in ycf1 (Online Supplementary Material S2). The ratio of non-synonymous to synonymous SNPs (1.93) was high enough in comparison to those of all intraspecific variations of the PCGs of the C. album chloroplast genome (15:11 = 1.36), which is also considered as a high ratio in comparison to other plant species (Park et al., 2021d). In addition, ycf1 displays high nucleotide diversity within the genus (Hong et al., 2017; Liu et al., 2018; de Souza et al., 2019; Celiński et al., 2020; Loeuille et al., 2021) and even among certain species (Kim et al., 2020; Park et al., 2021d), which is also congruent with previous studies showing that ycf1 is the most promising plastid DNA barcode in land plants (Neubig et al., 2009; Dong et al., 2015). Interestingly, ycf1 is reportedly a target gene of PBR1 in Arabidopsis (Yang et al., 2016) as well as an essential part of the chloroplast protein import machinery (Kikuchi et al., 2013; de Vries et al., 2015). Moreover, ycf1 and ycf2 in the tobacco chloroplast genome reportedly have an essential function according to a transformation experiment (Drescher et al., 2000). Taken together, the selection pressure imposed on ycf1 may not be high enough in comparison to other PCGs in the chloroplast genome and appeared to be distributed evenly on the gene region (Online Supplementary Material S2), also addressing the need for further investigations of ycf1 through land plant species.

Fig. 2.

Nucleotide diversity of the six Agrimonia chloroplast genomes. The X-axis represents the coordination of the chloroplast genome and the Y-axis represents the nucleotide diversity values from 500-bp window sliding in 200-bp steps. Highest peaks of the nucleotide diversity graph display the names of the genetic regions. The three blue arrow bars below the X-axis indicate the large single-copy (LSC), inverted repeat (IR), and small single-copy (SSC) regions, respectively.

Intraspecific variations identified from the four complete chloroplast genomes of Agrimonia pilosa

The intraspecific variations of the four A. pilosa complete chloroplast genomes were investigated based on pair-wise comparisons, as was done in previous studies (Kwon et al., 2019a; Park et al., 2019f, 2019h, 2021f; Oh et al., 2021), as this method offers advantage of estimating the degree of intraspecific variation. SNPs and INDELs identified from the four A. pilosa chloroplast genomes ranged from 22 to 280 and 16 to 105 (86 bp to 633 bp in length) (Fig. 3A). It should be noted that the partial chloroplast genome KY419942 (Zhang et al., 2017) exhibited a high degree of intraspecific variation against the remaining three chloroplast genomes, including different varieties of A. pilosa and A. pilosa var. nepalensis (Fig. 3B). This phenomenon was also found in two chloroplast genomes of Duchesnea indica (Andrews) Th. Wolf (Kim et al., 2020), among which one is also from the same study (Zhang et al., 2017). For a comparison of the degrees of intraspecific variation in other plant species calculated from various studies which investigated intraspecific variation in plant samples isolated in Korea and China (Table 2), the proportions of SNPs and INDELs against the length of the alignment sequences were calculated. This shows that the ratios of the SNPs range from 0.000226 to 0.001801 and those of the INDELs are from 0.000522 to 0.004073. The degrees of SNPs and INDELs for A. pilosa are separated by two groups (blue and orange arrows in Fig. 3B) due to the numerous variations found in KY419942 chloroplast genome (Fig. 3A).

Fig. 3.

Plot of intraspecific variations identified from the chloroplast genome of various plant species: (A) display of the plot on which the Xaxis denotes the proportion of single nucleotide polymorphisms (SNPs) and the Y-axis is the proportion of insertions and deletions (INDELs) identified from various plant species. Three color-filled circles, shown in blue, orange, and dark blue, represent the Sanguisorbeae species, between Chinese isolates, between Korean and Chinese isolates, and between Korean isolates, respectively. Circles filled in with white represent the four groups, other species between Chinese isolates, those between Chinese isolates, those between Korean and Chinese isolates, those between Korean isolates, those between Korean and unknown isolates. B. Enlarged graph from the dotted lines in the graph in (A). Blue arrows and yellow arrows indicate the proportions of the SNPs and INDELs identified from A. pilosa chloroplast genomes.

To evaluate the degrees of SNPs and INDELs identified among available Agrimonia chloroplast genomes, the degrees of intraspecific variation from five species in the tribe Sanguisorbeae which contains more than one chloroplast genome were also identified (Table 3). Based on the geographical locations of the species in Sanguisorbeae, including Agrimonia species, three major groups were defined: (1) one within Chinese isolates, (2) one between Korean and Chines isolates, and (3) one within Korean isolates. The proportions of SNPs and INDELs identified from all three groups were distributed from low to high degrees (Fig. 3B), suggesting that these proportions are not strongly affected by the geographical distribution. This finding is also congruent with those associated with other plant species investigated in this study (Table 2); however, certain Orchidaceae species, including Goodyera schlechtendaliana Rchb. f. and Gastrodia electa Blume, displayed very high SNP and INDEL proportions (above 0.01) (Fig. 3A); these can be considered family-specific characteristics.

List of chloroplast genomes used for comparative analyses in this study.

Two hundred and ninety-two non-redundant SNPs identified based on a multiple-sequence alignment of four A. pilosa chloroplast genomes indicated that most of the SNPs are common among the SNPs identified based on six pair-wise comparisons of A. pilosa, as the largest number of SNPs among A. pilosa is 280, found between MT040192 and KY419942 (Fig. 3A). In addition, 78 out of 292 SNPs and two INDELs were found in 36 of 84 PCGs (42.85%), which is a large number in comparison with other intraspecific variations of Arabidopsis thaliana L. (11 PCGs) (Park et al., 2020f), C. album (nine PCGs) (Park et al., 2021d), and Campanula takesimana Nakai (eight PCGs) (Park et al., 2021a). It is remarkable that only three synonymous SNPs identified in three PCGs were from MT040192; two synonymous SNPs were from MT415946 and no variations were found in A. pilosa var. nepalensis (MW387437), indicating that the only five PCGs, matK, psbC, psaB, psbJ, and psbB, have only synonymous SNPs apart from KY419942, lower than those identified in the three species used in this comparison.

In the same sense, ycf1 exhibited 10 synonymous SNPs, 11 non-synonymous SNPs, and one 9-bp insertion, which did not cause a frameshift only from KY419942. It is also interesting that three A. pilosa chloroplast genomes but not KY419942 displayed no intraspecific variation in ycf1, which usually presents numerous intraspecific variations (Kim et al., 2020; Park et al., 2020f, 2021d). This provides indirect evidence of the distance between KY419942 and the remaining three A. pilosa chloroplast genomes.

An additional interesting factor is that certain proportions of SNPs and INDELs of Agrimonia species are highest among those of Sanguisorbeae species. Determination of these proportions was from pair-wise comparisons with the KY419942 chloroplast genome, which displays a high level of intraspecific variation (Fig. 3A). This difference supports the hypothesis that the KY419942 chloroplast genome evolved independently through, for instance, long-term geographical isolation, akin to island plant species (Zhang et al., 2019a; Celiński et al., 2020).

Interspecific variations identified from the chloroplast genomes of three Agrimonia species

Given the fewest intraspecific variations between A. pilosa and A. pilosa var. nepalensis among the six pair-wise comparisons of A. pilosa chloroplasts conducted here (Fig. 4A), the chloroplast genomes of A. coreana, A. nipponica, and A. pilosa were used to investigate the interspecific variations. Due to the differences from the KY419942 chloroplast genome (Figs. 4A), four Agrimonia chloroplast genomes were used for this investigation (Fig. 4B). The numbers of SNPs identified from the four chloroplast genomes range from 244 to 311 (Fig. 4B), higher than those of SNPs within A. pilosa (Fig. 4A) except for that between A. coreana and A. pilosa (GenBank accession number: MT415496). The numbers of INDEL regions and the total lengths of the INDEL regions identified from A. pilosa (GenBank accession number: MT415496) are slight lower than those identified from another A. pilosa chloroplast genome (GenBank accession number: KY419942) (Fig.4B). Moreover, ycf1, which has a high number of intraspecific variations in KY419942, presented more variations than the intraspecific variations of A. pilosa: 22 non-synonymous SNPs, 14 synonymous SNPs, and one 9-bp insertion. These variations show the presence of four non-synonymous SNPs and four synonymous SNPs from A. coreana; six non-synonymous SNPs and one synonymous SNP from A. nipponica; seven non-synonymous SNPs, five synonymous SNPs, and a 9-bp insertion from KY419942 (A. pilosa); and five non-synonymous SNPs and one synonymous SNP from MT415496 (A. pilosa). In addition, three synonymous SNPs showed that half were from A. coreana and MT415496 and half were from A. nipponica and KY419942, suggesting that KY419942 is a different species from A. pilosa based on the investigation of sequence variations, a case similar to certain cryptic plant species (Gurushidze et al., 2008; Okuyama and Kato 2009; Myszczyński et al., 2017; Li et al., 2020b; Liu et al., 2020b) or hybrid-origin species such as Arabidopsis suecica (r.) Norrl. (O'Kane et al., 1996; Novikova et al., 2017) and Arabis species (Kawabe et al., 2018).

Fig. 4.

Numbers of single nucleotide polymorphisms (SNPs) and insertions and deletions (INDELs) identified among the four Agrimonia chloroplast genomes. Yellow, grey, and blue boxes indicate the A. coreana, A. nipponica, and A. pilosa chloroplast genomes, respectively. Numbers of SNPs and INDEL regions with the total lengths of INDEL regions are displayed on the lines connecting two chloroplast genomes. The comparisons with the KY419942 chloroplast genome exhibit additional information about the numbers of SNPs and INDELs in the IR region due to a lack of one of the IR regions in the KY419942 chloroplast genome.

The chromosome numbers for A. pilosa isolated in China and on Mt. Baekdu were reported to be 2n = 4x = 28 (Kwon et al., 2005) and 2n = 2x = 16 (Mitrenina et al., 2020), respectively. These different chromosomal configurations of A. pilosa suggest at least two possible origins of A. pilosa. First, a tetraploid of A. pilosa may have been formed via allopolyploidization, as in the case of Coffea arabica L. which originates from C. canephora Pierre ex A. Froehner and C. eugenioides S. Moore (Clarindo and Carvalho, 2008). In this case, two types of chloroplast genomes from two parental species can be found in a hybrid species if allopolyploidization occurred in both directions, in agreement with the two distinct types of chloroplast genomes found in this study. The second scenario involves two different specification events, resulting in two different basic chromosome numbers, n = 7 (Kwon et al., 2005) and n = 8 (Mitrenina et al., 2020). From the two speciation routes, two different types of chloroplast genomes could be passed down to A. pilosa, which can explain the current distinct between the two chloroplast genomes.

Phylogenetic analysis of the Agrimonia chloroplast genome sequence

ML and BI phylogenetic trees were constructed based on the 33 Sanguisorbeae species, including the two Agrimonia chloroplast genomes assembled in this study. Agrimonia exhibited monophyletic characteristics in that it formed one clade clustered with the Hagenia and Leucosidea genera (Fig. 5), congruent with the taxonomy of the subtribe Agrimoniinae (Chung, 2008) and previous phylogenetic analyses (Helfgott et al., 2000; Zhang et al., 2017). In addition, A. pilosa var. nepalensis was clustered with two A. pilosa complete chloroplast genomes isolated in Korea and China (Fig. 5), also in agreement with the low number of intraspecific variations among the four A. pilosa chloroplast genomes (Fig. 4A).

Fig. 5.

Maximum likelihood tree of 33 Sanguisorbeae chloroplast genomes with two outgroup species. Numbers on branches in the phylogenetic tree indicate the bootstrap values of ML and the prior possibility of Bayesian inference. Chloroplast genomes sequenced in this study are presented in bold text.

In contrast, the A. pilosa chloroplast (GenBank accession number: KY419942), which is different from the remaining three A. pilosa chloroplast genomes, is a sister to A. nipponica (Fig. 5), in contrast to the expected results based on the numbers of SNPs and INDELs (Fig. 3B). This indicates a need for an additional investigation of A. pilosa to test the suggested hypotheses that A. pilosa was formed from hybridization or independent speciation events. The chromosome numbers of the samples of A. pilosa from which chloroplast genomes were determined are unknown. The chromosome number heterogeneity of A. pilosa, as discussed above, associated with hybridization and polyploidization, may be responsible for the close relationship between the A. pilosa chloroplast (GenBank accession number: KY419942) and A. nipponica. The chromosome numbers of A. nipponica and A. coreana are 2n = 4x = 28 (Iwatsubo et al., 1993), reflecting the complex speciation history of Agrimonia species. Based on leaf morphological characteristics, A. coreana and A. nipponica are closely related by having 1–2 pairs of lateral leaflets in the mid-caule region and an abaxial surface of the leaflets with densely glandular and tomentose hairs (Chung, 2017), in contrast to the phylogenetic tree (Fig. 5). This indicates that additional research is needed to investigate the corresponding morphological features possibly to support the current phylogenetic tree or to consider the possibility of hybridized individuals of A. pilosa. Furthermore, A. pilosa may contain cryptic species, or there may even have been a misidentification of A. pilosa chloroplast genome (KY419942), as in the C. album (Park et al., 2021d) and Magnolia insignis (Wall.) Blume and Magnolia alba Figlar & Noot cases. Park (2020) can be referenced for further analyses.

Comparative analysis of SSR polymorphisms on six Agrimonia chloroplast genomes

SSRs were identified from the six Agrimonia chloroplast genomes, denoting the total numbers of SSRs, including normal SSRs, potential SSRs, and the extended SSRs ranging from 546 to 558 apart from KY419942 due to the lack of one IR region (Table 4). PentaSSR and hexaSSR identified in the chloroplast genome are usually classified into the potential SSR of which the length is 10 or 12 bp with two repeats. Only the A. coreana chloroplast genome contained two normal pentaSSRs (P0000056 and P0000057) with the three repeats (Fig. 6A). The numbers of normal SSRs for the six chloroplast genomes ranged from 73 to 83 (Fig. 6A). The four A. pilosa chloroplast genomes exhibited a range of 73 to 74, in good agreement with other plant species, such as C. album (53 to 55) (Park et al., 2021d). In addition, KY419942 without one IR region displayed the same level of normal SSRs (Fig. 6A), reflecting that the IR region of A. pilosa does not contain normal SSRs. The A. coreana and A. nipponica chloroplast genomes contained more normal SSRs than those of A. pilosa: A. coreana covers more monoSSRs and pentaSSRs, while A. nipponica had more diSSRs and tetraSSRs (Fig. 6A).

List of normal SSRs, potential SSRs, and extended SSRs identified from the six Agrimonia chloroplast genomes.

Fig. 6.

Number of normal simple sequence repeats (SSRs) and distribution of SSR groups identified on six Agrimonia chloroplast genomes. A. Number of SSR types in each of the six Agrimonia chloroplast genomes. The X-axis represents the six Agrimonia chloroplasts and the Yaxis indicates the number of normal SSRs. The five different colors indicate the five types of SSRs, i.e., monoSSRs, diSSRs, triSSRs, tetraSSRs, and pentaSSRs. The table below the X-axis represents the numbers of normal SSRs along with the six Agrimonia samples. B. Number of SSR groups along with the number of samples in each SSR group. The X-axis is the number of species in the SSR groups and the Y-axis is the number of SSR groups. The orange-colored bar indicates singleton SSRs. C. Pie graph presenting SSR groups, showing the length differences in the SSR group.

Based on a comparison of the flanking sequences of SSRs identified from the six Agrimonia chloroplast genomes, 601 SSR groups and 272 singleton SSRs were identified (Fig. 6B). In addition, 281 of 601 SSR groups (46.76%), covering six SSRs from six different chloroplast genomes, were identified (Fig. 6B). This indicates that fewer than half of the Agrimonia SSRs are shared among the six chloroplast genomes, referred to as common SSR groups. One hundred and fifty-one SSR groups (25.12%) contained five SSRs that originated from five chloroplast genomes (Fig. 6B). Seventeen out of 281 common SSR groups (6.05%) exhibited length variations in the SSR regions (Fig. 6C), representing a good molecular marker candidate with which to distinguish species or populations. The length differences in the SSR region ranged from 1 bp to 6 bp (Fig. 6C). Twelve out of 17 common SSR groups (70.59%) were monoSSR and the remaining five common SSR groups consisted of two diSSRs, two pentaSSRs, and one pentaSSR/hexaSSR (Table 5). The greatest difference (6 bp found in SSR Group 281) was caused by three additional time repeats of diSSRs, indicating that the SSRs in KY419942 and the A. coreana chloroplast genomes were eight-fold TA diSSR while the remaining SSRs are five-fold TA diSSR (Table 5). Interestingly, SSR Group 189 displayed one of six SSRs that was the pentaSSR type (A. nipponica), while the remaining SSRs were hexaSSRs (Fig. 6B, C). These variations in the 17 common SSR groups can serve as a good example for developing molecular markers based on SSRs in chloroplast genomes. In detail, the 151 SSR groups covering five Agrimonia chloroplast genomes can be considered as another type of molecular marker candidate because one of the six chloroplast genomes lacks the SSR, which is clear evidence allowing the detection of a specific Agrimonia sample. Moreover, because the number of SSRs can be affected by the evolution of chloroplast genomes (Sawicki et al., 2020), SSRs identified from the six Agrimonia chloroplast genomes can be used to understand the corresponding evolutionary features with the additional chloroplast genome of Agrimonia and a neighbor genus in the near future. Taken together, the variations identified in the SSRs from six Agrimonia chloroplast genomes exhibit not only the dynamic characteristics of Agrimonia SSRs but also suggest potential uses of Agrimonia chloroplast genomes.

List of SSRs from common SSR groups displaying length difference of SSR regions.

CONCLUSION

We completed the chloroplast genomes of A. coreana and A. nipponica isolated in Korea and conducted comparative analyses of six Agrimonia chloroplast genomes originating from three species. The Agrimonia coreana and A. nipponica chloroplast genomes are 151,362 bp and 155,161 bp long, respectively. High nucleotide diversity was detected in the ycf1 gene, congruent with earlier work (Hong et al., 2017; Jiang et al., 2017; Liu et al., 2018; de Souza et al., 2019; Kim et al., 2019d; Li et al., 2019; Celiński et al., 2020; Park et al., 2020e; Loeuille et al., 2021). The intraspecific variations of four A. pilosa chloroplast genomes revealed that KY419942 is distinct from the remaining three A. pilosa chloroplast genomes. In addition, the interspecific variations among A. coreana, A. nipponica, and A. pilosa indicated that the distance between two A. pilosa chloroplast genomes is similar to that between A. coreana and A. nipponica. A phylogenetic analysis found that KY419942 was clustered with A. nipponica, suggesting that KY419942 (A. pilosa) might have been misidentified or a cryptic species caused by complex evolutionary histories. A comparison of SSRs identified from the six Agrimonia chloroplast genomes suggests potential molecular markers with which to distinguish between species or among populations of the same species. Considering these aspects together with the results of this study, our A. coreana and A. nipponica chloroplast genomes provide insight into the phylogenetic relationships among Agrimonia species in Korea and about the morphological as well as interspecific and intraspecific features of chloroplast genomes, including variations of chloroplast genomes. Moreover, information about the two Agrimonia chloroplast genomes serves to unravel the complex evolutionary history of A. pilosa, A. coreana, and A. nipponica, which require further analyses that can include deciphering additional chloroplast genomes of A. pilosa and neighboring species.

ONLINE SUPPLEMENTARY MATERIALS

S1 and S2 are available at https://doi.org/10.11110/kjpt.2024.54.1.147.

Supplementary Fig.1.

Nucleotide diversity of ycf1 in the six Agrinonia chloroplast genomes

X-axis presents coordination of chloroplast genome and Y-axis presents values of nucleotide diversity value of 500-bp window sliding by 200-bp step displaying ycf1 genic regions. Below X-axis, positions of sequence variations identified in ycf1 were presented as black lines in the six grey bars, corresponding to the six Agrinonia chloroplast genomes. Blue-colored circles under the gene diagram means synonymous SNPs and orange-colored circles indicate non-synonymous SNPs. INDEL was depicted as blue-colored triangle.

kjpt-54-1-47-Supplementary-Material-1.pptx
Supplementary Fig.2.

Numbers and proportions of SNPs and INDELs identified among the four Agrimonia pilosa chloroplast genomes

(A) Blue boxes indicate A. pilosa chloroplast genomes and light-blue box means A. pilosa var. nepalensis. Number of SNPs and INDEL regions with total length of INDEL regions were displayed on the lines connecting two chloroplast genomes. The comparisons with KY419942 chloroplast genome exhibit additional information of # of SNPs and INDELs in the IR region due to lack of one of IR regions in KY419942 chloropalst genome. (B) Blue boxes indicate A. pilosa chloroplast genomes and light-blue box means A. pilosa var. nepalensis. Proportion of SNPs and INDEL regions against the alignment length and alignment length were displayed on the lines connecting two chloroplast genomes.

kjpt-54-1-47-Supplementary-Material-2.pptx

Acknowledgements

We would like to express our appreciation of Dr. Sang-Hun Oh for the useful advice about this manuscript. This work was supported by an InfoBoss Research Grant (IBG-0035).

Notes

CONFLICTS OF INTEREST

The authors declare that there are no conflicts of interest.

References

Alzahrani D. A., Yaradua S. S., Albokhari E. J., Abba A.. 2020;Complete chloroplast genome sequence of Barleria prionitis, comparative chloroplast genomics and phylogenetic relationships among. Acanthoideae BMC Genomics 21:393.
Baek J., Park S., Lee J., Min J., Park J., Lee G. W.. 2021;The complete chloroplast genome of Chrysanthemum zawadskii Herbich (Asteraceae) isolated in Korea. Mitochondrial DNA Part B Resources 6:1956–1958.
Bolger A. M., Lohse M., Usadel B.. 2014;Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 30:2114–2120.
Bum H. M., Kim H. K., Kim N. S., Kwon W., Park J.. 2020;The complete chloroplast genome of Douinia plicata (Lindb.) Konstant. & Vilnet (Scapaniaceae, Jungermanniales). Mitochondrial DNA Part B Resources 5:3680–3682.
Cai Z., Penaflor C., Kuehl J. V., Leebens-Mack J., Carlson J. E., dePampjilis C. W., Boore J. L., Jansen R. K.. 2006;Complete plastid genome sequences of Drimys, Liriodendron, and Piper: Implications for the phylogenetic relationships of magnoliids. BMC Evolutionary Biology 6:77.
Celiński K., Kijak H., Wiland-Szymańska J.. 2020;Complete chloroplast genome sequence and phylogenetic inference of the Canary Islands Dragon Tree (Dracaena draco L.). Forests 11:309.
Chang H., Zhang L., Xie H., Liu J., Xi Z., Xu X.. 2021;The conservation of chloroplast genome structure and improved resolution of infrafamilial relationships of Crassulaceae. Frontiers in Plant Science 12:631884.
Chen L., Kang Y.-H.. 2014;Antioxidant activities of Agrimonia pilosa ledeb: In vitro comparative activities of its different fractions. Korean Journal of Plant Resources 27:642–649.
Cheon K.-S., Kim K.-A., Jang S.-K., Yoo K.-O.. 2016;Complete chloroplast genome sequence of Campanula takesimana (Campanulaceae), an endemic to Korea. Mitochondrial DNA Part A, DNA Mapping, Sequencing, and Analysis 27:2169–2171.
Cho M.-S., Kim Y., Kim S.-C., Park J.. 2019;The complete chloroplast genome of Korean Pyrus ussuriensis Maxim. (Rosaceae): Providing genetic background of two types of P. ussuriensis. Mitochondrial DNA Part B Resources 4:2424–2425.
Choi N. J., Xi H., Park J.. 2021;A comparative analyses of the complete mitochondrial genomes of fungal endosymbionts in Sogatella furcifera, white-backed planthoppers. International Journal of Genomics 2021:6652508.
Choi Y. G., Yun N., Park J., Xi H., Min J., Kim Y., Oh S.-H.. 2020;The second complete chloroplast genome sequence of the Viburnum erosum (Adoxaceae) showed a low level of intra-species variations. Mitochondrial DNA Part B Resources 5:271–272.
Chung K.-S.. 2008. A systematic study of genus Agrimonia (Rosaceae). PhD dissertation,. The University of Oklahoma, Norman; OK, USA: 183.
Chung K.-S.. 2017;The first record of Agrimonia gorovoii Rumjantsev in Korea (Rosaceae). Korean Journal of Plant Taxonomy 47:132–136.
Chung K. S., Kim Y.-S.. 2000;A taxonomic study on the genus Agrimonia (Rosaceae) in Korea. Korean Journal of Plant Taxonomy 30:315–337.
Clarindo W. R., Carvalho C. R.. 2008;First Coffea arabica karyogram showing that this species is a true allotetraploid. Plant Systematics and Evolution 274:237–241.
Darlington C. D., Wylie A. P.. 1955. Chromosome Atlas of Flower Plants George Allen and Union Ltd.. London: p. 519.
de Souza U. J., Nunes B. R., Targueta C. P., Diniz-Filho F., de Campos Telles M. P.. 2019;The complete chloroplast genome of Stryphnodendron adstringens (Leguminosae -Caesalpinioideae): Comparative analysis with related Mimosoid species. Scientific Reports 9:14206.
de Vries J., Sousa F. L., Bölter B., Soll J., Gould S. B.. 2015;YCF1: A green TIC? The Plant Cell 27:1827–1833.
Deamer D., Akeson M., Branton D.. 2016;Three decades of nanopore sequencing. Nature Biotechnology 34:518–524.
Dong W., Xu C., Li C., Sun J., Zuo Y., Shi S., Cheng T., Guo J., Zhou S.. 2015;ycf1, the most promising plastid DNA barcode of land plants. Scientific Reports 5:8348.
Drescher A., Ruf S., Calsa T. Jr, Carrer H., Bock R.. 2000;The two largest chloroplast genome?encoded open reading frames of higher plants are essential genes. The Plant Journal 22:97–104.
Du Y.-P., Bi Y., Yang F.-P., Zhang M.-F., Chen X.-Q., Xue J., Zhang X.-H.. 2017;Complete chloroplast genome sequences of Lilium: Insights into evolutionary dynamics and phylogenetic analyses. Scientific Reports 7:5751.
Gichira A. W., Li Z., Saina J. K., Long Z., Hu G., Gituru R. W., Wang Q., Chen J.. 2017;The complete chloroplast genome sequence of an endemic monotypic genus Hagenia (Rosaceae): Structural comparative analysis, gene content and microsatellite detection. PeerJ 5:e2846.
Gil H.-Y., Kim Y., Kim S.-H., Jeon J.-H., Kwon Y., Kim S.-C., Park J.. 2019;The complete chloroplast genome of Pyrus ussuriensis Maxim. (Rosaceae). Mitochondrial DNA Part B Resources 4:1000–1001.
Goodwin S., McPherson J. D., McCombie W. R.. 2016;Coming of age: Ten years of next-generation sequencing technologies. Nature Reviews Genetics 17:333–351.
Greiner S., Lehwark P., Bock R.. 2019;OrganellarGenome-DRAW (OGDRAW) version 1.3. 1: Expanded toolkit for the graphical visualization of organellar genomes. Nucleic Acids Research 47:W59–W64.
Gurushidze M., Fritsch R. M., Blattner F. R.. 2008;Phylogenetic analysis of Allium subg. Melanocrommyum infers cryptic species and demands a new sectional classification. Molecular Phylogenetics and Evolution 49:997–1007.
Hassemer G., Bruun-Lund S., Shipunov A. B., Briggs B. G., Meudt H. M., Rønsted N.. 2019;The application of highthroughput sequencing for taxonomy: The case of Plantago subg. Plantago (Plantaginaceae). Molecular Phylogenetics and Evolution 138:156–173.
Helfgott D. M., Francisco-Ortega J., Santos-Guerra A., Jansen R. K., Simpson B. B.. 2000;Biogeography and breeding system evolution of the woody Bencomia alliance (Rosaceae) in Macaronesia based on ITS sequence data. Systematic Botany 25:82–97.
Heo K.-I., Park J., Xi H., Min J.. 2020;The complete chloroplast genome of Agrimonia pilosa Ledeb. isolated in Korea (Rosaceae): Investigation of intraspecific variations on its chloroplast genomes. Mitochondrial DNA Part B Resources 5:2264–2266.
Hong K.-J., Ki W., Lee I.-J., Lee H., Park J., Lee W.. 2022;The complete mitochondrial genome of Aclees taiwanensis Kôno, 1933 (Coleoptera: Curculionidae). Mitochondrial DNA Part B Resources 7:1460–1462.
Hong S.-Y., Cheon K.-S., Yoo K.-O., Lee H.-O., Cho K.-S., Suh J.-T., Kim S.-J., Nam J.-H., Sohn H.-B., Kim Y.-H.. 2017;Complete chloroplast genome sequences and comparative analysis of Chenopodium quinoa and C. album. Frontiers in Plant Science 8:1696.
Huelsenbeck J. P., Ronquist F.. 2001;MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics 17:754–755.
Hutchinson J.. 1964. Agrimonia. In The Genera of Flowering Plants. Vol 1. Angiospermae Dicotyledones Clarendon Press. Oxford: p. 206.
Iwatsubo Y., Mishima M., Naruhashi N.. 1993;Chromosome studies of Japanese Agrimonia (Rosaceae). Cytologia 58:453–461.
Jiang D., Zhao Z., Zhang T., Zhong W., Liu C., Yuan Q., Huang L.. 2017;The chloroplast genome sequence of Scutel laria baicalensis provides insight into intraspecific and interspecific chloroplast genome diversity in Scutellaria. Genes 8:227.
Jung C. H., Kim J.-H., Park S., Kweon D.-H., Kim S.-H., Ko S.- G.. 2010;Inhibitory effect of Agrimonia pilosa Ledeb. on inflammation by suppression of iNOS and ROS production. Immunological Investigations 39:159–170.
Kalkman C.. 2004. Rosaceae. Flowering Plants · Dicotyledons In : Kubitzki K., ed. Springer. Berlin: p. 343–386.
Kang M.-J., Kim S.-C., Lee H.-R., Lee S.-A., Lee J.-W., Kim T.-D., Park E.-J.. 2020;The complete chloroplast genome of Korean Gastrodia elata Blume. Mitochondrial DNA Part B Resources 5:1015–1016.
Katoh K., Standley D. M.. 2013;MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Molecular Biology and Evolution 30:772–780.
Kawabe A., Nukii H., Furihata H. Y.. 2018;Exploring the history of chloroplast capture in Arabis using whole chloroplast genome sequencing. International Journal of Molecular Sciences 19:602.
Kikuchi S., Bédard J., Hirano M., Hirabayashi Y., Oishi M., Imai M., Takase M., Ide T., Nakai M.. 2013;Uncovering the protein translocon at the chloroplast inner envelope membrane. Science 339:571–574.
Kim D.-H., An B.-J., Lee J.-Y.. 2011a;Whitening activities of the Agrimonia pilosa L. extracts. Journal of Applied Biological Chemistry 54:284–289.
Kim D. E., Kim P., Lee H., Kim N.-H., Kim D., Lee M.-J., Ban Y.-G., Jang B.-J., Park J.. 2023;Comprehensive analysis of the complete mitochondrial genome of Melanoplus differentialis (Acrididae: Melanoplinae) captured in Korea. Entomological Research 53:66–81.
Kim J., Kim Y., Park J.. 2019a;Complete chloroplast genome sequence of the Salix koriyanagi Kimura ex Goerz (Salicaceae). Mitochondrial DNA Part B Resources 4:549–550.
Kim M.-H., Park S., Lee J., Baek J., Park J., Lee G. W.. 2021a;The complete chloroplast genome of Glycyrrhiza uralensis Fisch. isolated in Korea (Fabaceae). Korean Journal of Plant Taxonomy 51:353–362.
Kim M., Xi H., Park J.. 2021b;Genome-wide comparative analyses of GATA transcription factors among 19 Arabidopsis ecotype genomes: Intraspecific characteristics of GATA transcription factors. PLoS ONE 16:e0252181.
Kim M., Xi H., Park S., Yun Y., Park J.. 2021c;Genome-wide comparative analyses of GATA transcription factors among seven Populus genomes. Scientific Reports 11:16578.
Kim S.-H., Cho C. H., Yang M., Kim S.-C.. 2017;The complete chloroplast genome sequence of the Japanese Camellia (Camellia japonica L.). Mitochondrial DNA Part B Resources 2:583–584.
Kim S.-T., Oh S.-H., Park J.. 2021d;The complete chloroplast genome of Diarthron linifolium (Thymelaeaceae), a species found on a limestone outcrop in eastern Asia. Korean Journal of Plant Taxonomy 51:345–352.
Kim T. H., Kim J. M., Baek J. M., Kim T. W., Kim D. J., Park J. H., Choe M.. 2011b;Antioxidant and whitening effects of Agrimonia pilosa Ledeb water extract. Korean Journal of Medicinal Crop Science 19:177–184.
Kim Y., Chung Y., Park J.. 2019b;The complete chloroplast genome sequence of Dysphania pumilio (R. Br.) Mosyakin & Clemants (Amaranthaceae). Mitochondrial DNA Part B Resources 4:403–404.
Kim Y., Heo K.-I., Lee S., Park J.. 2019c;The complete chloroplast genome sequence of Pseudostellaria palibiniana (Takeda) Ohwi (Caryophyllaceae). Mitochondrial DNA Part B Resources 4:973–974.
Kim Y., Park J., Chung Y.. 2019d;Comparative analysis of chloroplast genome of Dysphania ambrosioides (L.) Mosyakin & Clemants understanding phylogenetic relationship in genus Dysphania R. Br. Korean Journal of Plant Resources 32:644–668.
Kim Y., Park J., Chung Y.. 2019e;The complete chloroplast genome of Suaeda japonica Makino (Amaranthaceae). Mitochondrial DNA Part B Resources 4:1505–1507.
Kim Y., Park J., Chung Y.. 2020;The comparison of the complete chloroplast genome of Suaeda japonica Makino presenting different external morphology (Amaranthaceae). Mitochondrial DNA Part B Resources 5:1616–1618.
Kim Y., Xi H., Park J.. 2021e;The complete chloroplast genome of Limonium tetragonum (Plumbaginaceae) isolated in Korea. Korean Journal of Plant Taxonomy 51:337–344.
Kim Y., Yi J.-S., Min J., Xi H., Kim D. Y., Son J., Park J., Jeon J.-I.. 2019f;The complete chloroplast genome of Aconitum coreanum (H. Lév.) Rapaics (Ranunculaceae). Mitochondrial DNA Part B Resources 4:3404–3406.
Kwon W., Kim Y., Park J.. 2019a;The complete chloroplast genome of Korean Marchantia polymorpha subsp. ruderalis Bischl. & Boisselier: Low genetic diversity between Korea and Japan. Mitochondrial DNA Part B Resources 4:959–960.
Kwon W., Kim Y., Park J.. 2019b;The complete mitochondrial genome of Dumortiera hirsuta (Sw.) Nees (Dumortieraceae, Marchantiophyta). Mitochondrial DNA Part B Resources 4:1586–1587.
Kwon W., Kim Y., Park J.. 2019c;The complete mitochondrial genome of Korean Marchantia polymorpha subsp. ruderalis Bischl. & Boisselier: Inverted repeats on mitochondrial genome between Korean and Japanese isolates. Mitochondrial DNA Part B Resources 4:769–770.
Kwon W., Min J., Xi H., Park J.. 2019d;The complete chloroplast genome of Fissidens nobilis Griff. (Fissidentaceae, Bryophyta). Mitochondrial DNA Part B Resources 4:2225–2226.
Kwon Y. J., Sul M. R., An J. K., Kim C. H., Sun B.-Y.. 2005;Chromosome numbers of vascular plants of Mt. Baekdu and adjacent area in China. Korean Journal of Plant Taxonomy 35:47–55.
Lee B., Park J.. 2021a;The complete chloroplast genome of Zoysia japonica Steud. isolated in Korea (Poaceae): Investigation of potential molecular markers on Z. japonica chloroplast genomes. Plant Biotechnology Reports 15:707–715.
Lee B., Park J.. 2021b;The complete chloroplast genome of Zoysia matrella (L.) Merr. isolated in Korea (Poaceae): Investigation of intraspecific variations on chloroplast genomes. Mitochondrial DNA Part B Resources 6:572–574.
Lee J., Park J., Xi H., Park J.. 2020;Comprehensive analyses of the complete mitochondrial genome of Figulus binodulus (Coleoptera: Lucanidae). Journal of Insect Science 20:10.
Lee Y. S., Park J. Y., Kim J.-K., Lee H. O., Park H.-S., Lee S.-C., Kang J. H., Lee T. J., Sung S. H., Yang T.-J.. 2016;Complete chloroplast genome sequence of Artemisia fukudo Makino (Asteraceae). Mitochondrial DNA Part B Resources 1:376–377.
Li C., Zheng Y., Huang P.. 2020a;Molecular markers from the chloroplast genome of rose provide a complementary tool for variety discrimination and profiling. Scientific Reports 10:12188.
Li H.. 2013. Aligning sequence reads, clone sequences and assembly contigs with BWA-MEM. arXiv preprint at: https://arxiv.org/abs/1303.3997.
Li H., Handsaker B., Wysoker A., Fennell T., Ruan J., Homer N., Marth G., Abecasis G., Durbin R.. 1000;Genome Project Data Processing Subgroup. 2009. The sequence alignment/map format and SAMtools. Bioinformatics 25:2078–2079.
Li W., Liu Y., Yang Y., Xie X., Lu Y., Yang Z., Jin X., Dong W., Suo Z.. 2018;Interspecific chloroplast genome sequence diversity and genomic resources in Diospyros. BMC Plant Biology 18:210.
Li W., Zhang C., Guo X., Liu Q., Wang K.. 2019;Complete chloroplast genome of Camellia japonica genome structures, comparative and phylogenetic analysis. PLoS ONE 14:e0216645.
Li Z.-Z., Ngarega B. K., Lehtonen S., Gichira A. W., Karichu M. J., Wang Q.-F., Chen J.-M.. 2020b;Cryptic diversity within the African aquatic plant Ottelia ulvifolia (Hydrocharitaceae) revealed by population genetic and phylogenetic analyses. Journal of Plant Research 133:373–381.
Liang H., Zhang Y., Deng J., Gao G., Ding C., Zhang L., Yang R.. 2020;The complete chloroplast genome sequences of 14 Curcuma species: Insights into genome evolution and phylogenetic relationships within zingiberales. Frontiers in Genetics 11:802.
Liu Q., Li X., Li M., Xu T., Schwarzacher W., Heslop-Harrison J. S.. 2020a;Comparative chloroplast genome analyses of Avena: insights into evolutionary dynamics and phylogeny. BMC Plant Biology 20:406.
Liu T., Qian J., Du Z., Wang J., Duan B.. 2020b;Complete genome and phylogenetic analysis of Agrimonia pilosa Ldb. Mitochondrial DNA Part B Resources 5:1435–1436.
Liu W., Kong H., Zhou J., Fritsch P. W., Hao G., Gong W.. 2018;Complete chloroplast genome of Cercis chuniana (Fabaceae) with structural and genetic comparison to six species in Caesalpinioideae. International Journal of Molecular Sciences 19:1286.
Loeuille B., Thode V., Siniscalchi C., Andrade S., Rossi M., Pirani J. R.. 2021;Extremely low nucleotide diversity among thirty-six new chloroplast genome sequences from Aldama (Heliantheae, Asteraceae) and comparative chloroplast genomics analyses with closely related genera. PeerJ 9:e10886.
Lowe T. M., Chan P. P.. 2016;tRNAscan-SE On-line: Integrating search and context for analysis of transfer RNA genes. Nucleic Acids Research 44:W54–W57.
Meng X.-X., Xian Y.-F., Xiang L., Zhang D., Shi Y.-H., Wu M.-L., Dong G.-Q., Ip S.-P., Lin Z.-X., Wu L., Sun W.. 2018;Complete chloroplast genomes from Sanguisorba: Identity and variation among four species. Molecules 23:2137.
Min J., Kim Y., Xi H., Heo K.-I., Park J.. 2019a;The complete chloroplast genome of coffee tree, Coffea arabica L.‘Typica’ (Rubiaceae). Mitochondrial DNA Part B Resources 4:2240–2241.
Min J., Kim Y., Xi H., Jang T., Kim G., Park J., Park J.-H.. 2019b;The complete chloroplast genome of a new candidate cultivar, Sang Jae, of Abeliophyllum distichum Nakai (Oleaceae): Initial step of A. distichum intraspecies variations atlas. Mitochondrial DNA Part B Resources 4:3716–3718.
Min J., Kwon W., Xi H., Park J.. 2019c;The complete chloroplast genome of Leucobryum juniperoideum (brid.) C. Müll. (Leucobryaceae, Bryophyta). Mitochondrial DNA Part B Resources 4:2962–2963.
Min J., Park J., Kim Y., Kwon W.. 2019d;The complete chloroplast genome of Artemisia fukudo Makino (Asteraceae): Providing insight of intraspecies variations. Mitochondrial DNA Part B Resources 4:1510–1512.
Mitrenina E., Skaptsov M., Kutsev M., Kuznetsov A., Ikeda H., Erst A.. 2020;A new diploid cytotype of Agrimonia pilosa (Rosaceae). Caryologia 73:67–73.
Murata G., Umemoto K.. 1983;Critical notes on the Asiatic Agrimonia with special reference to the distribution patterns of calcium oxalate crystals. Acta Phytotaxonomica et Geobotanica 34:166–177.
Myszczyński K., Bączkiewicz A., Buczkowska K., Ślipiko M., Szczecińska M., Sawicki J.. 2017;The extraordinary variation of the organellar genomes of the Aneura pinguis revealed advanced cryptic speciation of the early land plants. Scientific Reports 7:9804.
Nei M., Li W.-H.. 1979;Mathematical model for studying genetic variation in terms of restriction endonucleases. Proceedings of the National Academy of Sciences of the United States of America 76:5269–5273.
Neubig K. M., Whitten W. M., Carlsward B. S., Blanco M. A., Endara L., Williams N. H., Moore M.. 2009;Phylogenetic utility of ycf 1 in orchids: A plastid gene more variable than mat K. Plant Systematics and Evolution 277:75–84.
Nguyen L.-T., Schmidt H. A., von Haeseler A., Minh B. Q.. 2015;IQ-TREE: A fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Molecular Biology and Evolution 32:268–274.
Niu Z., Xue Q., Zhu S., Sun J., Liu W., Ding X.. 2017;The complete plastome sequences of four orchid species: Insights into the evolution of the Orchidaceae and the utility of plastomic mutational hotspots. Frontiers in Plant Science 8:715.
Novikova P. Y., Tsuchimatsu T., Simon S., Nizhynska V., Voronin V., Burns R., Fedorenko O. M., Holm S., Säll T., Prat E., Marande W., Castric V., Nordborg M.. 2017;Genome sequencing reveals the origin of the allotetraploid Arabidopsis suecica. Molecular Biology and Evolution 34:957–968.
Oh S.-D., Lee S.-K., Yun D.-W., Sun H.-J., Kang H.-G., Lee H.-Y., Xi H., Park J., Lee B.. 2021;The complete chloroplast genome of Zoysia macrostachya (Poaceae): Insights into intraspecific variations and species delimitation of the Zoysia species. Korean Journal of Plant Taxonomy 51:326–331.
Oh S.-H., Park J.. 2020;The complete chloroplast genome of Euscaphis japonica (Thunb.) Kanitz (Staphyleaceae) isolated in Korea. Mitochondrial DNA Part B Resources 5:3751–3753.
Oh S.-H., Suh H. J., Park J., Kim Y., Kim S.. 2019a;The complete chloroplast genome sequence of a morphotype of Goodyera schlechtendaliana (Orchidaceae) with the column appendages. Mitochondrial DNA Part B Resources 4:626–627.
Oh S.-H., Suh H. J., Park J., Kim Y., Kim S.. 2019b;The complete chloroplast genome sequence of Goodyera schlechtendaliana in Korea (Orchidaceae). Mitochondrial DNA Part B Resources 4:2692–2693.
O'Kane S. L. Jr, Schaal B. A., Al-Shehbaz I. A.. 1996;The origins of Arabidopsis suecica (Brassicaceae) as indicated by nuclear rDNA sequences. Systematic Botany 21:559–566.
Okuyama Y., Kato M.. 2009;Unveiling cryptic species diversity of flowering plants: Successful biological species identification of Asian Mitella using nuclear ribosomal DNA sequences. BMC Evolutionary Biology 9:105.
Park I., Kim W.-J., Yang S., Yeo S.-M., Li H., Moon B. C.. 2017;The complete chloroplast genome sequence of Aconitum coreanum and Aconitum carmichaelii and comparative analysis with other Aconitum species. PLoS ONE 12:e0184257.
Park I., Yang S., Kim W. J., Noh P., Lee H. O., Moon B. C.. 2018;Complete chloroplast genome of Sanguisorba × tenuifolia Fisch. ex Link. Mitochondrial DNA Part B Resources 3:909–910.
Park J.-S., Jin D.-P., Park J.-W., Choi B.-H.. 2019a;Complete chloroplast genome of Fagus multinervis, a beech species endemic to Ulleung Island in South Korea. Mitochondrial DNA Part B Resources 4:1698–1699.
Park J., An J.-H., Kim Y., Kim D., Yang B.-G., Kim T.. 2020a;Database of National Species List of Korea: The taxonomical systematics platform for managing scientific names of Korean native species. Journal of Species Research 9:233–246.
Park J., Bae Y., Kim B.-Y., Nam G.-H., Park J.-M., Lee B. Y., Suh H.- J., Oh S.-H.. 2021a;The complete chloroplast genome of Campanula takesimana Nakai from Dokdo Island in Korea (Campanulaceae). Mitochondrial DNA Part B Resources 6:135–137.
Park J., Choi Y. G., Yun N., Xi H., Min J., Kim Y., Oh S.-H.. 2019b;The complete chloroplast genome sequence of Viburnum erosum (Adoxaceae). Mitochondrial DNA Part B Resources 4:3278–3279.
Park J., Heo K.-I., Kim Y., Kwon W.. 2019c;The complete chloroplast genome of Potentilla fragarioides var. major Maxim. Mitochondrial DNA Part B Resources 4:1265–1266.
Park J., Kim Y., Kwon W., Xi H., Kwon M.. 2019d;The complete chloroplast genome of tulip tree, Liriodendron tulifipera L. (Magnoliaceae): Investigation of intra-species chloroplast variations. Mitochondrial DNA Part B Resources 4:2523–2524.
Park J., Kim Y., Kwon W., Xi H., Park C.-H.. 2021b;The complete chloroplast genome sequence of new species candidate of Plantago depressa Willd. in Korea (Plantaginaceae). Mitochondrial DNA Part B Resources 6:1961–1963.
Park J., Kim Y., Min J., Kim S.-T.. 2022a;The complete chloroplast genome of Atriplex gmelinii C. A. Mey. ex Bong. (Amaranthaceae). Mitochondrial DNA Part B Resources 7:541–543.
Park J., Kim Y., Xi H.. 2019e;The complete chloroplast genome sequence of male individual of Korean endemic willow, Salix koriyanagi Kimura ex Goerz (Salicaceae). Mitochondrial DNA Part B Resources 4:1619–1621.
Park J., Kim Y., Xi H., Nho M., Woo J., Seo Y.. 2019f;The complete chloroplast genome of high production individual tree of Coffea arabica L. (Rubiaceae). Mitochondrial DNA Part B Resources 4:1541–1542.
Park J., Lee J., Park J.. 2022b;The investigation of intraspecific characteristics and comparative analyses of the complete mitochondrial genome of Stegobium paniceum (Linnaeus, 1758) (Coleoptera: Ptinidae) assembled from public NGS raw reads of the black truffle, Tuber melanosporum. Science Progress 105:368504211072355.
Park J., Lee S. H., Kim J. H.. 2021c;Complete genome sequence of the endosymbiotic bacterium “Candidatus Riesia pediculicola”. Microbiology Resource Announcements 10:e01181–20.
Park J., Min J., Kim Y., Chung Y.. 2021d;The comparative analyses of six complete chloroplast genomes of morphologically diverse Chenopodium album L. (Amaranthaceae) collected in Korea. International Journal of Genomics 2021:6643444.
Park J., Min J., Kim Y., Xi H., Kwon W., Jang T., Kim G., Park J.- H.. 2019g;The complete chloroplast genome of a new candidate cultivar, Dae Ryun, of Abeliophyllum distichum Nakai (Oleaceae). Mitochondrial DNA Part B Resources 4:3713–3715.
Park J., Oh S.-H.. 2020;A second complete chloroplast genome sequence of Fagus multinervis Nakai (Fagaceae): Intraspecific variations on chloroplast genome. Mitochondrial DNA Part B Resources 5:1868–1869.
Park J., Park S., Jang T., Kim G., Park J.-H.. 2021e;The complete chloroplast genome of Abeliophyllum distichum f. lilacinum Nakai (Oleaceae) from the Chungbuk Province, Korea. Mitochondrial DNA Part B Resources 6:1754–1756.
Park J., Suh Y., Kim S.. 2020b;A complete chloroplast genome sequence of Gastrodia elata (Orchidaceae) represents high sequence variation in the species. Mitochondrial DNA Part B Resources 5:517–519.
Park J., Xi H.. 2021;Investigation of nucleotide diversity based on 17 sea cucumber mitochondrial genomes and assessment of sea cucumber mitochondrial gene markers. Advances in Oceanography & Marine Biology 2:2021.
Park J., Xi H.. 2022;The complete chloroplast genome of Campsis grandiflora (Bignoniaceae). Korean Journal of Plant Taxonomy 52:156–172.
Park J., Xi H., Kim Y.. 2020c;The complete chloroplast genome of Arabidopsis thaliana isolated in Korea (Brassicaceae): An investigation of intraspecific variations of the chloroplast genome of Korean A. thaliana. International Journal of Genomics 2020:3236461.
Park J., Xi H., Kim Y.. 2021f;The complete mitochondrial genome of Arabidopsis thaliana (Brassicaceae) isolated in Korea. Korean Journal of Plant Taxonomy 51:176–180.
Park J., Xi H., Kim Y., Heo K.-I., Nho M., Woo J., Seo Y., Yang J. H.. 2019h;The complete chloroplast genome of cold hardiness individual of Coffea arabica L. (Rubiaceae). Mitochondrial DNA Part B Resources 4:1083–1084.
Park J., Xi H., Kim Y., Kim M.. 2021g;Complete genome sequence of Lentilactobacillus parabuchneri strain KEM. Microbiology Resource Announcements 10:e01208–20.
Park J., Xi H., Kim Y., Kim S.-T.. 2022c;The complete chloroplast genome of Utricularia tenuicaulis Miki (Lentibulariaceae) isolated in Korea. Mitochondrial DNA Part B Resources 7:1143–1145.
Park J., Xi H., Kim Y., Nam S., Heo K.-I.. 2020d;The complete mitochondrial genome of new species candidate of Rosa rugosa (Rosaceae). Mitochondrial DNA Part B Resources 5:3435–3437.
Park J., Xi H., Oh S.-H.. 2020e;Comparative chloroplast genomics and phylogenetic analysis of the Viburnum dilatatum complex (Adoxaceae) in Korea. Korean Journal of Plant Taxonomy 50:8–16.
Park J., Xi H., Park J.. 2021h;Complete genome sequence of a Blochmannia endosymbiont of Colobopsis nipponica. Microbiology Resource Announcements 10:e01195–20.
Park J., Xi H., Park J., Nam S. J., Lee Y.-D.. 2020f;Complete genome sequence of the Blochmannia endosymbiont of Camponotus nipponensis. Microbiology Resource Announcements 9:e00703–20.
Park J., Xi H., Park J., Seo B. Y.. 2021i;A new mitochondrial genome of Sogatella furcifera (Horváth) (Hemiptera: Delphacidae) and mitogenome-wide investigation on polymorphisms. Insects 12:1066.
Park J., Yun N., Oh S.-H.. 2019i;The complete chloroplast genome of an endangered species in Korea, Halenia corniculata (L.) Cornaz (Gentianaceae). Mitochondrial DNA Part B Resources 4:1539–1540.
Park S.-H., Sim Y.-B., Kang Y.-J., Lee J.-K., Lim S.-S., Suh H.-W.. 2012;Effect of Agrimonia pilosa Ledeb extract on the antinociception and mechanisms in mouse. The Korean Journal of Physiology & Pharmacology 16:119–123.
Park S.. 2020. A phylogenomic study of Magnoliaceae and its evolutionary implications. PhD dissertation,. Sungshin Womens University; Seoul, Korea: 283.
Potter D., Eriksson T., Evans R. C., Oh S., Smedmark J. E. E., Morgan D. R., Kerr M., Robertson K. R., Arsenault M., Dickinson T. A., Cambell C. S.. 2007;Phylogeny and classifi cation of Rosaceae. Plant Systematics and Evolution 266:5–43.
Roberts R. J., Carneiro M. O., Schatz M. C.. 2013;The advantages of SMRT sequencing. Genome Biology 14:405.
Robertson K. R.. 1974;The genera of Rosaceae in the southeastern United States. Journal of the Arnold Arboretum 55:344–401.
Sawicki J., Bączkiewicz A., Buczkowska K., Górski P., Krawczyk K., Mizia P., Myszczyński K., Ślipiko M., Szczecińska M.. 2020;The increase of simple sequence repeats during diversification of Marchantiidae, an early land plant lineage, leads to the first known expansion of inverted repeats in the evolutionarily-stable structure of liverwort plastomes. Genes 11:299.
Seo H. S., Chung B. H., Cho Y. G.. 2008;Antioxidant and anticancer effects of agrimony (Agrimonia pilosa L.) and Chinese Lizardtail (Saururus chinensis Baill). Korean Journal of Medicinal Crop Science 16:139–143.
Shin W.-J., Lee K.-H., Park M.-H., Seong B.-L.. 2010;Broad?spectrum antiviral effect of Agrimonia pilosa extract on influenza viruses. Microbiology and Immunology 54:11–19.
Skalický V.. 1962;Ein Beitrag zur Erkenntnis der europäischen Arten der Gattung Agrimonia. Acta Horti Botanici Pragensis 1962:87–108.
Skalický V.. 1973;Amerikanische Arten der Gattung Agrimonia L. ser. Tuberosae ser. nova. Folia Geobotanica and Phytotaxonomica 8:95–104.
Suh H.-J., Min J., Park J., Oh S.-H.. 2021;The complete chloroplast genome of Aruncus dioicus var. kamtschaticus (Rosaceae). Mitochondrial DNA Part B Resources 6:1256–1258.
Van Dijk E. L., Auger H., Jaszczyszyn Y., Thermes C.. 2014;Ten years of next-generation sequencing technology. Trends in Genetics 30:418–426.
Wallroth C. F. W.. 1842;Monographisher Versuch über die Gewächs-Gattung Agrimonia Celsi. Beitrage zur Botanik (Wallroth) 1:1–66.
Wang C.-L., Ding M.-Q., Zou C.-Y., Zhu X.-M., Tang Y., Zhou M.-L., Shao J.-R.. 2017;Comparative analysis of four buckwheat species based on morphology and complete chloroplast genome sequences. Scientific Reports 7:6514.
Wang H., Park S.-Y., Song S.-H., San M.-L., Kim Y.-C., Ham T.-H., Kim D.-Y., Kim T.-S., Lee J., Kwon S.-W.. 2020;Analysis of complete chloroplast genome sequence of Korean landrace Cymbidium goeringii. 3 Biotech 10:29.
World Health Organization. 1998. Medicinal Plants in the Republic of Korea: Information on 150 Commonly Used Medicinal Plants WHO Regional Office for the Western Pacific. Manila: p. 316.
Xiang S., Liu X.-D., Sun W.-H., Lan S.-R., Liu Z.-J., Zou S.-Q.. 2019;The complete chloroplast genome sequence of Euscaphis japonica (Staphyleaceae). Mitochondrial DNA Part B Resources 4:3484–3485.
Xie D.-F., Yu H.-X., Price M., Xie C., Deng Y.-Q., Chen J.-P., Yu Y., Zhou S.-D., He X.-J.. 2019;Phylogeny of Chinese Allium species in section Daghestanica and adaptive evolution of Allium (Amaryllidaceae, Allioideae) species revealed by the chloroplast complete genome. Frontiers in Plant Science 10:460.
Xu Z., Xin T., Bartels D., Li Y., Gu W., Yao H., Liu S., Yu H., Pu X., Zhou J., Xu J., Xi C., Lei H., Song J., Chen S.. 2018;Genome analysis of the ancient tracheophyte Selaginella tamariscina reveals evolutionary features relevant to the acquisition of desiccation tolerance. Molecular Plant 11:983–994.
Yang J., Takayama K., Youn J.-S., Pak J.-H., Kim S.-C.. 2020;Plastome characterization and phylogenomics of East Asian beeches with a special emphasis on Fagus multinervis on Ulleung Island, Korea. Genes 11:1338.
Yang L., Teng Y., Gong S.-F., Feng Z.-M., Zhang Y.. 2021;The complete chloroplast genome of Agrimonia pilosa var. nepalensis (D. Don) Nakai. Mitochondrial DNA Part B Resources 6:938–940.
Yang X.-F., Wang Y.-T., Chen S.-T., Li J.-K., Shen H.-T., Guo F.-Q.. 2016;PBR1 selectively controls biogenesis of photosynthetic complexes by modulating translation of the large chloroplast gene Ycf1 in Arabidopsis. Cell Discovery 2:16003.
Yoo S.-C., Oh S.-H., Park J.. 2021;Phylogenetic position of Daphne genkwa (Thymelaeaceae) inferred from complete chloroplast data. Korean Journal of Plant Taxonomy 51:171–175.
Yoo S.-C., Park J., Oh S.-H.. 2023;Complete chloroplast genome of Daphne pseudomezereum var. koreana (Thymelaeaceae). Mitochondrial DNA Part B Resources 8:305–309.
Yoon J.-Y., Lee S.-Y., Jun H.-J., Lee J.-Y.. 2012;Anti-aging effects of solvent fraction from Agrimonia pilosa L. extracts. Journal of Applied Biological Chemistry 55:35–39. (in Korean).
Yuan Y., Jin X., Liu J., Zhao X., Zhou J., Wang X., Wang D., Lai C., Xu W., Huang J., Zha L., Liu D., Ma X., Wang L., Zhou M., Jiang Z., Meng H., Peng H., Liang Y., Li R., Jiang C., Zhao Y., Nan T., Jin Y., Zhan Z., Yang J., Jiang W., Huang L.. 2018;The Gastrodia elata genome provides insights into plant adaptation to heterotrophy. Nature Communications 9:1615.
Zerbino D. R., Birney E.. 2008;Velvet: Algorithms for de novo short read assembly using de Bruijn graphs. Genome Research 18:821–829.
Zhang S.-D., Jin J.-J., Chen S.-Y., Chase M. W., Soltis D. E., Li H.-T., Yang J.-B., Li D.-Z., Yi T.-S.. 2017;Diversification of Rosaceae since the Late Cretaceous based on plastid phylog enomics. New Phytologist 214:1355–1367.
Zhang W., Zhao Y., Yang G., Peng J., Chen S., Xu Z.. 2019a;Determination of the evolutionary pressure on Camellia oleifera on Hainan Island using the complete chloroplast genome sequence. PeerJ 7:e7210.
Zhang Z., Jiang X., Chen Y., Zhu P., Li L., Zeng Y., Tang T.. 2019b;Characterization of the complete chloroplast genome sequence of Conyza canadensis and its phylogenetic implications. Mitochondrial DNA Part B Resources 4:2028–2030.
Zhao Q.-Y., Wang Y., Kong Y.-M., Luo D., Li X., Hao P.. 2011;Optimizing de novo transcriptome assembly from shortread RNA-Seq data: A comparative study. BMC Bioinformatics 12:S2.
Zhou X., Ren L., Meng Q., Li Y., Yu Y., Yu J.. 2010;The nextgeneration sequencing technology and application. Protein & Cell 1:520–536.

Article information Continued

Fig. 1.

Circular gene map of Agrimonia coreana and Agrimonia nipponica chloroplast genomes. A. Agrimonia coreana, B. Agrimonia nipponica. Genes shown outside are transcribed clockwise, and those inside the circle are transcribed counter-clockwise. Genes are color-coded to distinguish different functional groups. The dark grey and the light grey plots in the inner circle correspond to the GC content and AT content, respectively.

Fig. 2.

Nucleotide diversity of the six Agrimonia chloroplast genomes. The X-axis represents the coordination of the chloroplast genome and the Y-axis represents the nucleotide diversity values from 500-bp window sliding in 200-bp steps. Highest peaks of the nucleotide diversity graph display the names of the genetic regions. The three blue arrow bars below the X-axis indicate the large single-copy (LSC), inverted repeat (IR), and small single-copy (SSC) regions, respectively.

Fig. 3.

Plot of intraspecific variations identified from the chloroplast genome of various plant species: (A) display of the plot on which the Xaxis denotes the proportion of single nucleotide polymorphisms (SNPs) and the Y-axis is the proportion of insertions and deletions (INDELs) identified from various plant species. Three color-filled circles, shown in blue, orange, and dark blue, represent the Sanguisorbeae species, between Chinese isolates, between Korean and Chinese isolates, and between Korean isolates, respectively. Circles filled in with white represent the four groups, other species between Chinese isolates, those between Chinese isolates, those between Korean and Chinese isolates, those between Korean isolates, those between Korean and unknown isolates. B. Enlarged graph from the dotted lines in the graph in (A). Blue arrows and yellow arrows indicate the proportions of the SNPs and INDELs identified from A. pilosa chloroplast genomes.

Fig. 4.

Numbers of single nucleotide polymorphisms (SNPs) and insertions and deletions (INDELs) identified among the four Agrimonia chloroplast genomes. Yellow, grey, and blue boxes indicate the A. coreana, A. nipponica, and A. pilosa chloroplast genomes, respectively. Numbers of SNPs and INDEL regions with the total lengths of INDEL regions are displayed on the lines connecting two chloroplast genomes. The comparisons with the KY419942 chloroplast genome exhibit additional information about the numbers of SNPs and INDELs in the IR region due to a lack of one of the IR regions in the KY419942 chloroplast genome.

Fig. 5.

Maximum likelihood tree of 33 Sanguisorbeae chloroplast genomes with two outgroup species. Numbers on branches in the phylogenetic tree indicate the bootstrap values of ML and the prior possibility of Bayesian inference. Chloroplast genomes sequenced in this study are presented in bold text.

Fig. 6.

Number of normal simple sequence repeats (SSRs) and distribution of SSR groups identified on six Agrimonia chloroplast genomes. A. Number of SSR types in each of the six Agrimonia chloroplast genomes. The X-axis represents the six Agrimonia chloroplasts and the Yaxis indicates the number of normal SSRs. The five different colors indicate the five types of SSRs, i.e., monoSSRs, diSSRs, triSSRs, tetraSSRs, and pentaSSRs. The table below the X-axis represents the numbers of normal SSRs along with the six Agrimonia samples. B. Number of SSR groups along with the number of samples in each SSR group. The X-axis is the number of species in the SSR groups and the Y-axis is the number of SSR groups. The orange-colored bar indicates singleton SSRs. C. Pie graph presenting SSR groups, showing the length differences in the SSR group.

Table 1.

List of six Agrimonia chloroplast genomes analyzed in this study.

Species Voucher GenBank accession Length (bp) Gap GC contents No. of genes Ref.



Whole LSC SSC IR Whole (%) LSC (%) SSC (%) IR (%) No. of PCGs No. of tRNAs No. of rRNAs
A. coreana IB-01066 MZ604439 155,362 84,597 18,807 25,979 0 36.9 34.8 30.3 42.6 84 37 8 This study
A. nipponica IB-01069 MZ604440 155,161 84,514 19,825 25,411 0 36.9 34.8 30.7 42.8 84 37 8 This study
A. pilosa IB-01030 MT415946 155,125 84,458 18,737 25.965 0 36.9 34.9 30.4 42.6 84 37 8 Heo et al. (2020)
MT415946
A. pilosa N/A MT040192 155,188 84,521 18,737 25,965 5 36.9 34.9 30.4 42.6 84 37 8 Liu et al. (2020b)
A. pilosaa N/A KY419942 129.106 84,399 18,743 25,964 0 35.8 34.9 30.4 42.6 78 30 4 Zhang et al. (2017)
A. pilosa var. nepalensis 2020Jul SH08 MW387437 155,147 84.480 18,737 25,965 0 36.9 34.8 30.4 42.6 85 37 8 Yang et al. (2021)

LSC, large single-copy; SSC, small single-copy; IR, inverted repeat; PCG, protein-coding gene; N/A, not available.

a

Partial chloroplast genome with only one IR region.

Table 2.

List of intraspecific variations identified from various plant species isolated from Korea and China using pair-wise comparison method.

Family Species name Source Target No. of SNPs Proportion of SNPs No. of INDEL regions INDEL length (bp) Proportion of INDELs Alignment length (bp) Reference
Selaginellaceae Selaginella tamariscina Korean isolate (MN894555) Chinese isolate (NC_041646) 1,223 0.009603 376 1,635 0.012839 127,350 Xu et al. (2018), Park et al. (2020b)
Rosaceae Hagenia abyssinica Chinese isolate (KX008604) Chinese isolate (KY420026) 82 0.000529 28 262 0.001689 155,127 Gichira et al. (2017), Zhang et al. (2017)
Sanguisorba officinalis Chinese isolate (NC_044694) Korean isolate (MZ145058) 85 0.000546 61 301 0.001934 155,596 Meng et al. (2018), Unpublished
Chinese isolate (NC_044694) Korean isolate (MK696193) 75 0.000482 60 290 0.001864 155,581
Korean isolate (MZ145058) Korean isolate (MK696193) 10 0.000064 17 105 0.000675 155,455
Sanguisorba tenuifolia Chinese isolate (NC_044692) Chinese isolate (NC_042223) 27 0.000174 32 138 0.000887 155,499 Meng et al., (2018), Park et al. (2018), Unpublished
Chinese isolate (NC_044692) Korean isolate (MK696194) 91 0.000585 64 338 0.002172 155,605
Chinese isolate (NC_042223) Korean isolate (MK696194) 89 0.000572 66 344 0.002211 155,581
Sanguisorba stipulata Korean isolate (MZ145059) Korean isolate (MK696195) 78 0.000502 77 367 0.002360 155,540 Unpublished
Sanguisorba filiformis Chinese isolate (NC_044693) Chinese isolate (KY419920) 45 0.000291 12 478 0.003090 154,707 Zhang et al. (2017), Meng et al. (2018)
Amaranthaceae Chenopodium album CAGAP004 (MW446246) CAGOH01 (MW446245) 28 0.000184 12 19 0.000125 152,207 Hong et al. (2017), Park et al. (2021d)
CAGAP004 (MW446246) CAJEJG05 (MW446243) 2 0.000013 2 2 0.000013 152,199
CAGAP004 (MW446246) CCANG01 (MW446241) 27 0.000177 17 315 0.002070 152,210
CAGAP004 (MW446246) CSJUK01 (MW446242) 21 0.000138 12 18 0.000118 152,207
CAGAP004 (MW446246) CVHUP01 (MW446244) 14 0.000092 13 33 0.000217 152,291
CAGAP004 (MW446246) NC_034950 17 0.000112 8 36 0.000237 152,201
CAGOH01 (MW446245) CAJEJG05 (MW446243) 30 0.000197 9 35 0.000230 152,207
CAGOH01 (MW446245) CCANG01 (MW446241) 0 0.000000 7 298 0.001958 152,200
CAGOH01 (MW446245) CSJUK01 (MW446242) 21 0.000138 7 15 0.000099 152,204
CAGOH01 (MW446245) CVHUP01 (MW446244) 30 0.000197 9 40 0.000263 152,293
CAGOH01 (MW446245) NC_034950 33 0.000217 6 37 0.000243 152,200
CAJEJG05 (MW446243) CCANG01 (MW446241) 29 0.000191 36 331 0.002175 152,210
CAJEJG05 (MW446243) CSJUK01 (MW446242) 23 0.000151 14 34 0.000224 151,961
CAJEJG05 (MW446243) CVHUP01 (MW446244) 16 0.000105 15 49 0.000322 152,045
CAJEJG05 (MW446243) NC_034950 19 0.000125 10 52 0.000342 151,955
CCANG01 (MW446241) CSJUK01 (MW446242) 21 0.000138 10 309 0.002030 152,206
CCANG01 (MW446241) CVHUP01 (MW446244) 30 0.000197 17 336 0.002206 152,296
CCANG01 (MW446241) NC_034950 33 0.000217 11 333 0.002188 152,203
CSJUK01 (MW446242) CVHUP01 (MW446244) 23 0.000151 13 39 0.000256 152,293
CSJUK01 (MW446242) NC_034950 26 0.000171 7 36 0.000237 152,200
CVHUP01 (MW446244) NC_034950 18 0.000118 8 57 0.000374 152,287
Dysphania pumilio Anyang city, Gyeonggi-do (MH936550) Gangseo-gu, Seoul (MK541016) 25 0.000165 1 2 0.000013 151,962 Kim et al. (2019b), Park et al. (2019i)
Suaeda japonica Ganghwa type (MK764271) Julpo type (MK558824) 3 0.000020 3 3 0.000020 152,112 Kim et al., (2019e), Kang et al. (2020)
Magnoliaceae Magnolia kobus Jejudo (NC_023237) Jejudo (MN894553) 50 0.000314 28 63 0.000395 159,457 Unpublished
Liriodendron tulipifera Korean isolate (MK477550) Chinese isolate (NC_008326) 12 0.000075 0 0 0.000000 159,886 Cai et al. (2006), Kwon et al. (2019d)
Theaceae Camellia japonica Soyeonpyeong-do (MK353210) Seogwang-ri in Jejudo (MK353211) 25 0.000159 2 2 0.000013 156,976 Kim et al., (2017), Kwon et al. (2019c)
Soyeonpyeong-do (MK353210) Wimi-ri in Jejudo island (KU951523) 25 0.000159 2 2 0.000013 156,972 Li et al. (2019), Min et al. (2019c)
Seogwang-ri in Jejudo (MK353211) Wimi-ri in Jejudo (KU951523) 25 0.000159 0 0 0.000000 156,971
Cv. Huaheling (MW602996) Soyeonpyeong-do (MK353210) 8 0.000051 7 36 0.000229 157,004
Cv. Huaheling (MW602996) Seogwang-ri in Jejudo (MK353211) 8 0.000050 9 38 0.000242 157,005
Cv. Huaheling (MW602996) Wimi-ri in Jejudo (KU951523) 33 0.000210 9 38 0.000242 157,005
S288C (NC_036830) Soyeonpyeong-do (MK353210) 78 0.000496 60 644 0.004099 157,110
S288C (NC_036830) Seogwang-ri in Jejudo (MK353211) 78 0.000496 61 645 0.004105 157,111
S288C (NC_036830) Wimi-ri in Jejudo (KU951523) 103 0.000656 61 645 0.004105 157,111
S288C (NC_036830) Cv. Huaheling (MW602996) 80 0.000509 61 671 0.004270 157,139
Caryophyllaceae Pseudostellaria palibiniana Mt. Taebaeksan (MK120981) Mt. Gwangdeoksan (MK309611) 84 0.000561 43 175 0.001169 149,741 Kim et al. (2019c), Park et al. (2019d)
Salicaceae Salix koriyanagi Male, Seoul (MK541017) Female, Seoul (MK120982) 0 0.000000 0 0 0.000000 155,548 Kim et al. (2019a), Park et al. (2019e)
Rosaceae Pyrus ussuriensis Bonghwa-gun Mt. Hambeaksan 121 0.000753 61 781 0.004859 160,727 Cho et al. (2019), Gil et al. (2019)
Fagaceae Fagus multinervis Ulleungdo (MN894556) Ulleungdo (MK518070) 2 0.000013 2 2 0.000013 158,350 Park et al. (2019a), Park and Oh (2020), Yang et al. (2020)
Ulleungdo (MN894556) Ulleungdo (MT762296) 2 0.000013 3 3 0.000019 158,350
Ulleungdo (MK518070) Ulleungdo (MT762296) 0 0.000000 0 0 0.000000 158,348
Asteraceae Artemisia fukudo Aphaedo, Shinan-gun (NC_044156) Jeungdo, Shinan-gun (MG951488) 7 0.000046 5 12 0.000079 151,838 Lee et al. (2016), Min et al. (2019d)
Orchidaceae Goodyera schlechtendaliana Korean isolate (MK144665) Korean isolate (MK134679) 200 0.001294 59 511 0.003305 154,607 Niu et al. (2017), Oh et al. (2019a, 2019b)
Chinese isolate (NC_029364) Korean isolate (MK144665) 842 0.005434 130 1,779 0.011480 154,964
Chinese isolate (NC_029364) Korean isolate (MK134679) 740 0.004780 95 1,470 0.009495 154,820
Chinese isolate (LC085346) Chinese isolate (NC_029364) 514 0.003310 110 2,133 0.013733 155,322
Chinese isolate (LC085346) Korean isolate (MK144665) 700 0.004526 112 1,366 0.008832 154,665
Chinese isolate (LC085346) Korean isolate (MK134679) 597 0.003863 72 1,065 0.006892 154,525
Chinese isolate (AB893949) Chinese isolate (LC085346) 445 0.002887 29 415 0.002693 154,119
Chinese isolate (AB893949) Chinese isolate (NC_029364) 864 0.005573 123 2,045 0.131912 155,027
Chinese isolate (AB893949) Korean isolate (MK144665) 282 0.001828 86 1,060 0.006871 154,261
Chinese isolate (AB893949) Korean isolate (MK134679) 163 0.001058 43 652 0.004232 154,067
Gastrodia elata Korean isolate (MN026874) Korean isolate (MN296709) 324 0.009119 76 630 0.017731 35,530 Yuan et al. (2018), Kang et al. (2020), Park et al. (2020b)
Chinese isolate (NC_037409) Korean isolate (MN026874) 493 0.013851 79 651 0.018291 35,592
Chinese isolate (NC_037409) Korean isolate (MN296709) 457 0.012842 75 671 0.018855 35,587
Oleaceae Abeliophyllum distichum NC_031445 MN127986 93 0.000596 15 57 0.000365 156,023 Min et al. (2019a), Min et al. (2019b), Park et al. (2019g), Park et al. (2021e)
NC_031445 MK616470 93 0.000596 19 64 0.000410 156,023
NC_031445 MF407183 93 0.000594 15 57 0.000364 156,567
NC_031445 MN116559 102 0.000654 19 64 0.000410 156,032
NC_031445 MW426545 99 0.000634 18 72 0.000461 156,034
MN127986 MK616470 0 0.000000 0 0 0.000000 156,008
MN127986 MF407183 0 0.000000 1 1 0.000006 156,553
MN127986 MN116559 9 0.000058 10 12 0.000077 156,019
MN127986 MW426545 6 0.000038 7 20 0.000128 156,021
MK616470 MF407183 0 0.000000 1 1 0.000006 156,553
MK616470 MN116559 9 0.000058 10 11 0.000071 156,019
MK616470 MW426545 6 0.000038 7 19 0.000122 156,021
MF407183 MN116559 9 0.000057 10 11 0.000070 156,564
MF407183 MW426545 6 0.000038 8 21 0.000134 156,566
MN116559 MW426545 7 0.000045 9 23 0.000147 156,028
Adoxaceae Viburnum erosum MN641480 MN218778 16 0.000101 8 50 0.000315 158,630 Park et al. (2019b), Choi et al. (2020)
Ranunculaceae Aconitum coreanum NC_031421 KU318669 29 0.000185 26 61 0.000388 157,075 Park et al. (2017), Kim et al. (2019f)
NC_031421 Namhansan Fotress (MN400660) 5 0.000032 9 52 0.000331 157,052
Namhansan Fotress (MN400660) (KU318669) 19 0.000121 25 92 0.000586 157,088
Thymelaeaceae Daphne genkwa Jindo island (MT754180) Buan 59 0.000444 90 404 0.003039 132,923 Yoo et al. (2021), Unpublished
Jindo island (MT754180) Chinese isolate (NC_045891) 69 0.000518 225 772 0.005798 133,156
Buan Chinese isolate (NC_045891) 85 0.000639 100 543 0.004080 133,083
Campanulaceae Campanula takesimana Dok-do island (MW013763) Ulleung-do island (NC_026203) 33 0.000194 39 662 0.003900 169,737 Cheon et al. (2016), Park and Xi (2021)
Poaceae Zoysia matrella Korean isolate (MT983887) Chiba Fair Green (AP014937) 28 0.000206 56 91 0.000670 135,894 Lee and Park (2021b)
Zoysia japonica Jeonju city (MW690657) Unknown (NC_036827) 68 0.000500 24 50 0.000405 135,894 Lee and Park (2021a)
Asteraceae Erigeron canadensis Korean isolate (MT806101) Chinese isolate (NC_046789) 103 0.000674 47 208 0.001360 152,926 Zhang et al. (2019b) Park et al. (in preparation)
Staphyleaceae Euscaphis japonica Chinese isolate (MN159078) Korean isolate (NC_052922) 424 0.002635 152 809 0.005067 160,941 Xiang et al. (2019), Oh and Park (2020)

SNP, single nucleotide polymorphism; INDEL, insertion and deletion.

Table 3.

List of chloroplast genomes used for comparative analyses in this study.

Tribe Subtribe Species name NCBI accession Total length (bp) No. of PCGs No. of tRNAs No. of rRNAs GC ratio (%) Reference
Sanguisorbeae Agrimoniinae Agrimonia pilosa MT415496 155,125 84 37 8 34.9 Heo et al. (2020)
Agrimonia pilosa MT040192 155,188 84 37 8 34.9 Liu et al. (2020b)
Agrimonia pilosaa KY419942 129.106 78 30 4 34.9 Zhang et al. (2017)
Agrimonia pilosa var. nepalensis MW387437 155,147 84b 37 8 34.8 Yang et al. (2021)
Agrimonia coreana MZ604439 155,362 84 37 8 36.9 This study
Agrimonia nipponica MZ604440 155,161 84 37 8 36.9 This study
Hagenia abyssinicaa KY420026 129,090 78 30 4 36.0 Zhang et al. (2017)
Hagenia abyssinica KX008604 154,961 85 37 8 37.1 Gichira et al. (2017)
Leucosidea sericeaa KY419929 128,924 78 30 4 35.9 Zhang et al. (2017)
Spenceria ramalanaa KY419995 128,291 77 30 4 36.0 Zhang et al. (2017)
Sanguisorbinae Sanguisorba officinalis NC_044694 155,749 84 37 8 37.2 Meng et al. (2018)
Sanguisorba officinalis MZ145058 155,412 85 37 8 37.2 Unpublished
Sanguisorba officinalis MK696193 155,393 N/A N/A N/A 37.2 Unpublished
Sanguisorba tenuifolia var. alba NC_044692 155,457 84 37 8 37.2 Meng et al. (2018)
Sanguisorba tenuifolia NC_042223 155,403 86 37 8 37.2 Park et al. (2018)
Sanguisorba tenuifolia MK696194 155,415 N/A N/A N/A 37.2 Unpublished
Sanguisorba stipulata MK696195 155,385 N/A N/A N/A 37.2 Unpublished
Sanguisorba stipulata MZ145059 155,328 85 37 8 37.2 Unpublished
Sanguisorba sitchensis NC_044691 155,127 84 37 8 37.2 Meng et al. (2018)
Sanguisorba filiformis NC_044693 154,282 84 37 8 37.3 Meng et al. (2018)
Sanguisorba filiformisa KY419920 128,491 78 30 4 36.3 Zhang et al. (2017)
Sanguisorba hakusanensis MZ145057 155,645 85 37 8 37.2 Unpublished
Bencomia exstipulata NC_039924 152,781 84 33 8 37.2 Unpublished
Bencomia moquinianaa KY420023 128,612 78 30 4 36.1 Zhang et al. (2017)
Bencomia sphaerocarpaa KY419986 128,439 78 30 4 36.1 Zhang et al. (2017)
Polylepis australisa KY419989 129,569 78 30 4 36.2 Zhang et al. (2017)
Polylepis reticulataa KY419921 129,512 78 30 4 36.2 Zhang et al. (2017)
Polylepis sp. SCZ-2017a KY419992 129,472 78 30 4 36.2 Zhang et al. (2017)
Acaena pinnatifidaa KY419984 129,416 78 30 4 36.2 Zhang et al. (2017)
Cliffortia repensa KY419983 129,326 78 30 4 36.1 Zhang et al. (2017)
Margyricarpus pinnatusa KY419972 129,276 78 30 4 36.2 Zhang et al. (2017)
Dendriopoterium menendeziia KY419966 128,700 78 30 4 36.2 Zhang et al. (2017)
Poterium spinosuma KY419948 128,592 78 30 4 36.1 Zhang et al. (2017)
Potentilleae Potentilla stolonifera NC_044418 156,440 84 37 8 36.9 Park et al. (2019c)
Roseae Rosa rugosa NC_044126 156,989 84 37 8 37.2 Kwon et al. (2019b)

IR, inverted repeat; PCG, protein-coding gene; N/A, not available.

a

These chloroplast genomes are partial chloroplast genome without one IR region.

b

Number of PCGs in this chloroplast genome was updated in this study.

Table 4.

List of normal SSRs, potential SSRs, and extended SSRs identified from the six Agrimonia chloroplast genomes.

SSR type A. coreana A. nipponica A. pilosa MT415946 A. pilosa MT040192 A. pilosa KY419942 A. pilosa var. nepalensis
Normal SSRs MonoSSR 62 57 55 55 56 54
DiSSR 10 12 9 9 9 9
TriSSR 3 3 3 3 2 3
TetraSSR 6 7 7 7 6 7
PentaSSR 2 0 0 0 0 0
Potential SSRs PentaSSR 307 310 312 312 248 311
HexaSSR 106 103 104 105 86 104
Extended SSRs HeptaSSR 32 31 33 32 28 32
OctaSSR 16 14 13 13 10 13
NonaSSR 9 9 9 9 10 9
DecaSR 5 4 4 4 5 4
Total 558 550 549 549 460 546

SSR, simple sequence repeat.

Table 5.

List of SSRs from common SSR groups displaying length difference of SSR regions.

Name Type A. coreana A. nipponica A. pilosa MT415946 A. pilosa MT040192 A. pilosa KY419942 A. pilosa var. nepalensis
SSRGroup 281 diSSR 16 10 10 10 8 10
SSRGroup 280 pentaSSR 15 10 10 10 10 10
SSRGroup 223 monoSSR 10 13 11 11 15 11
SSRGroup 58 pentaSSR 15 10 10 10 10 10
SSRGroup 567 monoSSR 10 11 10 10 13 10
SSRGroup 371 monoSSR 11 12 11 11 11 14
SSRGroup 120 monoSSR 11 11 10 11 13 11
SSRGroup 455 monoSSR 11 12 11 11 10 11
SSRGroup 373 monoSSR 11 10 11 10 12 11
SSRGroup 359 diSSR 14 12 12 12 12 12
SSRGroup 189 HexaSSR PentaSSR 12 10 12 12 12 12
SSRGroup 502 monoSSR 11 12 11 12 12 11
SSRGroup 338 monoSSR 10 10 11 11 10 11
SSRGroup 195 monoSSR 12 11 12 12 12 12
SSRGroup 92 monoSSR 13 13 12 12 13 12
SSRGroup 57 monoSSR 10 11 11 11 11 11
SSRGroup 36 monoSSR 10 11 11 11 11 11

SSR, simple sequence repeat.