Skip to main content
  • Research article
  • Open access
  • Published:

Chloroplast phylogenomic insights into the evolution of Distylium (Hamamelidaceae)

Abstract

Background

Most Distylium species are endangered. Distylium species mostly display homoplasy in their flowers and fruits, and are classified primarily based on leaf morphology. However, leaf size, shape, and serration vary tremendously making it difficult to use those characters to identify most species and a significant challenge to address the taxonomy of Distylium. To infer robust relationships and develop variable markers to identify Distylium species, we sequenced most of the Distylium species chloroplast genomes.

Results

The Distylium chloroplast genome size was 159,041–159,127 bp and encoded 80 protein-coding, 30 transfer RNAs, and 4 ribosomal RNA genes. There was a conserved gene order and a typical quadripartite structure. Phylogenomic analysis based on whole chloroplast genome sequences yielded a highly resolved phylogenetic tree and formed a monophyletic group containing four Distylium clades. A dating analysis suggested that Distylium originated in the Oligocene (34.39 Ma) and diversified within approximately 1 Ma. The evidence shows that Distylium is a rapidly radiating group. Four highly variable markers, matK-trnK, ndhC-trnV, ycf1, and trnT-trnL, and 74 polymorphic simple sequence repeats were discovered in the Distylium plastomes.

Conclusions

The plastome sequences had sufficient polymorphic information to resolve phylogenetic relationships and identify Distylium species accurately.

Background

Distylium Sieb. et Zucc is a genus of flowering plants in the tribe Fothergilleae of the family Hamamelidaceae, which is endemic to Asia. Fifteen species have been reported in Distylium worldwide, with 12 species occurring in China (D. chinense, 2n = 24). Additionally, two species are found in Japan, one of which is found also in China, and one species each in Malaysia and India. They are evergreen shrubs or small trees that grow mostly in subtropical evergreen forests.

This genus has been introduced as a cultivar and thrives in warm temperate and subtropical climates in Europe and the United States. Distylium, with dense branches and deep evergreen leaves, a neat tree shape, small red flowers in spring, good soundproof effects, and strong resistance to smoke and dust and various toxic gases (e.g., sulfur dioxide and chlorine), are suitable as greening and ornamental plants in cities, and industrial and mining areas. They are commonly cultivated in urban gardens in the Yangtze River basin of China. Some species, such as D. chinense, are used to stabilize solid earth embankments because of their robust root system, flooding tolerance, and resistance to sand burial soaks [1, 2].

Most Distylium species are endangered. According to the threatened species list of China’s higher plants [3], two species are Critically Endangered species (D. macrophyllum and D. tsiangii), two are Endangered species (D. chinense and D. gracile), and two species are Vulnerable (D. chungii and D. elaeagnoides). Some Distylium species are narrowly distributed, such as D. lepidotum, which is endemic to the Ogasawara (Bonin) Islands, located in the northwestern Pacific approximately 1000 km south of Tokyo [4]. D. tsiangii is only located in Dushan and Bazai counties of Guizhou Province.

Distylium species lack significant differences in the morphology of their flowers and fruits, and are classified primarily based on leaf morphology. However, leaf size, shape, and serration vary tremendously and are difficult characters to use in most cases. For example, the range of leaf variation in D. buxifolium is very striking [5]. This variability has led to a proposed number of new species, which have been reduced to synonymy, as more material has been found to link extreme forms [5]. Due to the insufficient number of morphological diagnostic characters and highly polymorphic traits, taxonomic classification of Distylium species has been unclear. Chloroplast genome markers, such as atpB, atpB-rbcL, matK, rbcL, trnH-psbA, and trnL-F, and the internal transcribed spacer (ITS) has enabled molecular phylogenetic analyses of several Distylium species [6,7,8,9]. However, those markers have lower divergence among Distylium species; no study has inferred the phylogeny of this genus.

Whole chloroplast genome sequences have been widely used to infer phylogenetic relationships at different taxonomic levels, and provide an effective genetic resource for resolving complex evolutionary relationships and identifying ambiguous species. With the development of sequencing methods, complete chloroplast genome sequences are now available at low cost, extending gene-based phylogenetics to genome-based phylogenomics [10,11,12], extending gene-based species identification to genome-based super DNA barcoding [13, 14], and making it easier to study evolutionary events in plant species [15].

In this study, we specifically aimed to (1) develop and screen appropriate intrageneric markers in the chloroplast genome to establish DNA barcodes for Distylium; (2) estimate the effectiveness of a whole chloroplast genome data set in resolving the relationships within this radiating lineage; (3) estimate the divergence time of Distylium.

Results

Basic characteristics of the Distylium plastomes

The complete chloroplast genomes of the 12 newly sequenced Distylium species ranged in length from 159,041 bp (D. lepidoium) to 159,127 bp (D. gracile) (Table 1). The Distylium chloroplast genomes had a quadripartite structure typical of most angiosperm species, including large single copy (LSC) and small single copy (SSC) regions separated by two inverted repeat (IRa and IRb) regions (Fig. 1). The LSC regions ranged from 87,825 bp (D. pingpienense) to 87,863 bp (D. racemosum), the SSC regions varied between 18,770 bp (D. dunnianum) and 18,796 bp (D. lepidoium), and the IR regions ranged from 26,225 bp (D. elaeagnoides) to 26,241 bp (D. dunnianum). The GC content of the chloroplast genome sequences was 38.0%. A total of 114 unique genes was detected in the chloroplast genomes of the 11 Distylium species (Table S1), including 80 protein coding genes, 30 tRNA genes, and 4 rRNA genes, and the gene order was highly conserved (Fig. 1 and Table 1). A total of 18 genes (including 11 coding genes and seven tRNA genes) had introns, with 16 genes having one intron and two genes (ycf3 and clpP) having two introns in the Distylium chloroplast genomes.

Table 1 The basic plastomes information of 12 Distylium samples
Fig. 1
figure 1

Gene map of the Distylium plastomes. Genes shown inside the inner circle are transcribed counterclockwise and those outside the circle are transcribed clockwise. The GC content of the genome is indicated by the dashed area in the inner circle

Repetitive sequences

A total of 801 SSRs were identified across the chloroplast genomes of the 11 Distylium species (Fig. 2 and Table S2). The number of SSRs per species ranged from 70 (D. dunnianum) to 78 (D. gracile). The majority of the SSRs were mononucleotide repeats (78.65%), followed by dinucleotide (8.61%) and tetranucleotide (5.87%) repeats. There were no hexanucleotide repeats in the Distylium plastomes. The SSR A and T motifs were the most frequent. SSRs were particularly rich in AT in the Distylium plastomes. Among those SSRs, most were located in the LSC/SSC regions (94.01%).

Fig. 2
figure 2

Frequency of the simple sequence repeat (SSR) sequences in the Distylium plastomes. a. The number of SSRs detected in the 11 Distylium species; b. Frequency of SSRs with di- to penta-nucleotide motifs

A total of 96 unique SSRs and 74 SSRs were polymorphic across the 11 Distylium species. All polymorphic SSRs were located in the single copy regions, except two SSRs (Table 2). The mononucleotide repeat units A and T were also the most frequent polymorphic SSRs.

Table 2 Polymorphism of SSRs in Distylium plastomes

Indel variations

A total of 76 indels were discovered in the Distylium plastomes, including 59 normal indels and 17 repeat indels. Most of the indels (72.37%, 55 times) were located in the spacer regions, 15.79% (12 times) of indels occurred in the exons, and 11.84% (nine times) were found in the introns (Fig. 3). The trnT-trnL spacer had five indels, followed by ndhC-trnV (3 indels). The size of the normal indels ranged from 1 to 13 bp, with 8 bp and 9 bp length indels being the most common. The largest indel (13 bp) was located in the trnC-petN spacer and was a deletion in D. macrophyllum. The second largest indel was in the ycf1 exon of 12 bp length and was an insert in the two D. lepidoium samples. The length of the repeat indels ranged from 2 to 16 bp. The largest repeat indel occurred in the rpl20-rps12 spacer and the second largest repeat indel was located in the rps7-trnV spacer.

Fig. 3
figure 3

Analyses of indels in the Distylium plastomes. a. Number and size of the indels among the Distylium plastomes. b. Frequency of indel types and locations

Variation in the plastomes and molecular markers for Distylium species

The mVISTA results showed that the 11 Distylium chloroplast genomes were collinear and highly conserved (Figure S1). The entire chloroplast genome of the 11 Distylium species was 159,360 bp in length, including 298 polymorphic sites and 115 parsimony informative sites (Table 3). The overall nucleotide diversity (π) was 0.00045; however, each region of the chloroplast genome revealed different nucleotide diversity; the SSC exhibited the highest π value (0.00089) and the IR had the lowest π value (0.00006). All species had a unique chloroplast haplotype. The number of nucleotide substitutions among the 11 species varied from 7 to 109, and the p-distance varied from 0.0004 to 0.0069. The lowest divergence was observed between D. buxifolium and D. chinese, and the largest sequence divergence was observed between D. chinese and D. lepidoium.

Table 3 Sequences divergence of Distylium plastomes

The π value ranged from 0 to 0.0027 in an 800-bp sliding window size. In total, four peaks with π values > 0.002 were identified in the chloroplast genome (Fig. 4). Those regions included matK-trnK, ndhC-trnV, ycf1, and trnT-trnL. Three intergenic regions (matK-trnK, ndhC-trnV, and trnT-trnL) were located in the LSC region, and the ycf1 coding region was in the SSC region. The primers were designed for the four variable markers (Table S3) and tested the effective for amplification (Figure S2).

Fig. 4
figure 4

Nucleotide diversity (π) in the Distylium plastomes using sliding window method. The four mutation hotspot regions (π > 0.002) were annotated. π values were calculated in 800 bp sliding windows with 100 bp steps

We tested the variability in the hypervariable markers by comparing with the three universal DNA barcodes (matK, rbcL, and trnH-psbA). The variable information is shown in Table 4. The intergenic spacer marker trnH-psbA was 367 bp, including two variable sites and no parsimony informative sites. The rbcL and matK genes were 1428 bp with three variable and three informative sites, and 1515 bp with only one variable and no informative sites, respectively. Combining the three universal markers, the aligned length was 3310 bp, with six variable sites and three informative sites. The mean distance was 0.00045. The species identification analyses showed that the universal DNA barcodes had less discriminatory power; there were only four haplotypes when combining the three markers, and the ML tree had lower resolution and most of the samples were not distinguished (Table 4 and Fig. 5).

Table 4 Variability of the four highly mutation hotspot regions and the universal chloroplast DNA barcodes in Distylium
Fig. 5
figure 5

ML tree for Distylium using combine three universal plant DNA barcodes and four highly variable regions combinations

The four hypervariable markers ranged from 827 bp (matK-trnK) to 2306 bp (ycf1) in length. The ycf1 gene had the greatest number of variable sites (20 sites) followed by trnT-trnL (9 sites), matK-trnK (8 sites), and ndhC-trrnV had the fewest (6 sites). Combining the four hypervariable markers, there were 43 variable sites and 16 parsimony informative sites that produced the most current identification (Table 4). The identified hypervariable markers had higher resolution compared with the tree universal markers, based on the ML tree (Fig. 5). We also amplified and sequenced these four regions of two samples and used the tree-based methods to test their discrimination power. The results showed the two samples had successful identification (Figure S3).

Phylogenetic inference

Using the complete chloroplast genome sequences, we inferred the phylogenetic relationships among the 24 Hamamelidaceae samples. The best-fit model GTR + G from ModelFinder was used for ML and BI analyses. The topology of the ML and BI trees was nearly identical (Fig. 6). All Distylium species formed a monophyletic clade that was sister to Parrotia within Fothergilleae. Distylium had a short branch on the phylogenetic tree, indicating low divergence among Distylium species. Four clades were reconstructed in Distylium with a 100% bootstrap value. Clade I included the basal species D. lepidoium. Clade II included only D. myricoides. Clade III included only D. macrophyllum. Clade IV included the most advanced eight species, i.e., D. buxifolium, D. chinense, D. pingienense, D. cuspidatum, D. dunnianum, D. gracile, D. elaeagoides, and D. racemosum (Fig. 6).

Fig. 6
figure 6

Phylogenetic reconstruction of Hamamelidaceae from Maximum likelihood (ML) and Bayesian inference (BI) methods based on the plastome dataset. The ML tree is shown. Number of the branches represent ML bootstrap support value (BP) /Bayesian posterior probability (PP). The photos were taken by ZhiXiang Zhang

Estimate of divergence time

Divergence time estimates suggested that Hamamelioideae diverged from Hamamelidaceae about 99.38 Ma (95% HPD: 90.71–105.44 Ma) during the Cenomanian of the Upper Cretaceous (Fig. 7). The stem note of Fothergilleae was dated to 88.87 Ma (95% HPD: 97–91.18 Ma). The stem date for Distylium was estimated to be 34.39 Ma (95% HPD: 29.99–39.03 Ma) in the Oligocene and the Distylium crown date was 5.39 Ma (95%HPD: 0.82–12.3 Ma) in the Pliocene. Diversification within this genus occurred over a short time period of approximately 1 Ma.

Fig. 7
figure 7

Divergence times of Hamamelidaceae obtained from BEAST analysis based on the complete plastome sequences. Mean divergence time of the nodes were shown next to the nodes while the blue bars correspond to the 95% highest posterior density (HPD). Black circles indicate the three calibration points

Discussion

The genera Distyliopsis, Distylium, Fothergilla, Parrotia, Parrotiopsis, Shaniodendron, and Sycopsis occur in the tribe Fothergilleae of the subfamily Hamamedoideae [9]. According to the phylogenetic relationships based on the several chloroplast and nuclear ITS genes [6, 8], Distylium is sister to Distyliopsis [9]. This is the first use of molecular data to infer the Distylium phylogeny. The Distylium genus formed a well-defined monophyletic group according to the chloroplast genome data (Fig. 6). Moreover, the phylogenetic tree possessed a series of short internodes within Distylium and most species diversified < 1 Ma (Fig. 7), suggesting that this genus has undergone rapid radiation. D. lepidoium was at the base of the genus. This species was first described in 1918 and is endemic to the Ogasawara Islands [4]. D. myricoides formed a monotypic clade and is distributed in eastern and southeastern China. According to the morphological characteristics, D. myricoides resembles D. buxijolium most closely, from which it may be distinguished by its larger leaves [5]. However, this relationship was not supported by the present study. D. buxijolium and D. chinense were sister species and formed a group supported by morphological characteristics [5]. In this study, the chloroplast genome data provided information to infer the phylogeny of Distylium. However, due to rapid radiation, sampling of additional individuals from each species and extending more nuclear genes would provide additional evidence of the evolutionary history of Distylium.

Most Distylium species are rare and endangered; thus, the development of rapid and easily accessible species identification methods is essential. The variations in the morphological characteristics between species were continuous and uninterrupted. Therefore, it was difficult to distinguish species using morphological characteristics. DNA barcoding offers an opportunity to identify Distylium species. RbcL and matK are the two core DNA barcodes in plants. However, many studies have shown that these two markers have lower species identification power [16, 17]. Our study also showed that rbcL and matK or a combination of the two markers failed to discriminate Distylium species (Fig. 5), explaining the low resolution in previous studies and highlighting the importance of developing highly divergent markers.

Some studies have indicated that mutations are not random and are clustered as “mutation hotspots” or “highly variable regions” [10, 16, 18]. In this study, we compared the whole chloroplast genomes and identified the mutation hotspots in Distylium (Fig. 4). Four variable loci (matK-trnK, ndhC-trnV, ycf1, and trnT-trnL) were discovered. TrnT-trnL has been frequently used in plant phylogeny [19]. MatK-trnK and ycf1 are considered divergence hotspots in angiosperms based on our previous research [16]. NdhC-trnV has been less used in plant phylogeny and species identification and is prone to have large indels [20]. The coding region of the ycf1 locus is the most divergent marker in most groups, and has been suggested as the main plant DNA barcode [17]. MatK-trnK is located in the LSC region, and this locus is used less frequently in evolutionary biology. Some lineages have the ploy T structure [21]. Therefore, the lineage-specific, highly variable markers developed in this study will facilitate further phylogenetic reconstruction and DNA barcoding of rare and endangered Distylium species.

Conclusions

In this study, we report 10 newly sequenced chloroplast genomes of Distylium species. The overall genomic structure, including the gene number and gene order, was well-conserved. The phylogeny and divergence time analyses based on the plastome sequences showed that Distylium was a rapidly radiating group and most speciation events occurred < 1 Ma. A comparison of sequence divergence across the Distylium plastomes revealed that matK-trnK, ndhC-trnV, ycf1, and trnT-trnL were mutation hotspot regions. Overall, our study demonstrated that plastome sequences can be used to improve phylogenetic resolution and species discrimination. Extended sampling and additional nuclear markers are absolutely necessary in further studies.

Methods

Plant material and DNA extraction

A total of 12 individual samples representing 11 Distylium species were sampled from the Plant DNA Bank of China at the Institute of Botany, Chinese Academy of Sciences. All samples were identified based on morphological characters. The details of the plant samples are presented in Table 5. Total genomic DNA was extracted from the leaf tissues of herbarium specimens of this genus following the modified CTAB DNA extraction protocol [22].

Table 5 Sampling information for the 12 Distylium samples

Sequence, chloroplast genome assembly, and annotation

The total DNA was fragmented ultrasonically to construct350-bp insert libraries according to the manufacturer’s instructions, which was then used for sequencing. Paired-end sequencing was performed on an Illumina HiSeq X-ten at Novogene (Tianjin, China), yielding approximately 4 Gb of high-quality 150-bp paired-end reads per sample.

The raw reads obtained from Novogene were filtered using Trimmomatic 0.39 [23] with the following parameters: LEADING = 20, TRAILING = 20, SLIDING WINDOW = 4:15, MIN LEN = 36, and AVG QUAL = 20. High-quality reads were assembled de novo using the SPAdes 3.6.1 program [24]. The chloroplast genome sequence contigs were selected from the initial assembled reads in SPAdes by performing a BLAST search using several related Hamamelidaceae chloroplast genome sequences as references. The chloroplast genome sequence contigs were further assembled using Sequencher 5.4.5. All plastid assemblies were annotated in Plann [25] using D. macrophyllum (GenBank Accession number: MN729500) as the reference, and missing or incorrect genes were checked in Sequin. A circular diagram for the chloroplast genome was generated using OGDRAW [26]. All chloroplast genomes assembled in this study have been deposited in GenBank under accession numbers of MW248109 - MW248120.

Microstructural mutation events

The Perl script microsatellite identification tool (MISA, http://pgrc.ipk-gatersleben.de/misa/misa.html) was used to identify the microsatellite regions of the chloroplast genome with the parameters set to 10 (repeat units ≥10) for mononucleotide simple sequence repeats (SSRs), 6 (repeat units ≥6) for dinucleotides, 5 (repeat units ≥5) for trinucleotides, 4 (repeat units ≥4) for tetranucleotides, and 3 (repeat units ≥3) for pentanucleotides and hexanucleotides.

The chloroplast genomes sequences were aligned using MAFFT [27] followed by manually examination and adjustment. Based on the aligned sequence matrix, the indels were manually checked and divided into categories of repeat indels and normal indels, according to Dong et al. [15]. D. dunnianum was used as the reference to determine the size and position of the indel events.

Sequence divergence analysis

The mVISTA program was used to compare the variability of Distylium chloroplast genome using the Shuffle-LAGAN mode [28]. Single nucleotide substitutions and the genetic p-distances were calculated using MEGA 7.0 [29] based on the aligned chloroplast genome sequences. To assess sequence divergence andexplore highly variable chloroplast markers, nucleotide diversity (π) was calculated by sliding window analysis using DnaSP v6 [30] with a widow size of 800 bp and a step size of 100 bp. The primers for amplifying the highly variable regions were designed using FastPCR [31]. The PCR amplifications were performed following Dong et al. [32].

Nucleotide diversity and the number of haplotypes were used to assess marker variability for all barcodes (hype-variable markers and the universal plant DNA barcodes, rbcL, matK, and trnH-psbA). The tree-based method was utilized to evaluate discrimination power. A maximum-likelihood (ML) tree was prepared in IQ-TREE2 using the GTR model [33].

Phylogenetic analyses

To elucidate the phylogenetic positions of Distylium within Hamamelidaceae and the interspecific phylogenetic relationships within Distylium, multiple alignments were performed using the whole chloroplast genome of 24 Hamamelidaceae samples representing 11 genera, including Cercidiphyllum japonicum, Daphniphyllum oldhamii, and Liquidambar formosana as outgroups. The Hamamelidaceae chloroplast genomes were aligned using MAFFT, and ambiguous alignment regions were trimmed with Gblocks 0.91b [34]. The maximum-likelihood (ML) analysis was run with RAxML-NG [35] with the best-fit model from ModelFinder [36]. Branch support was assessed by fast bootstrap methodology using non-parametric bootstrapping and 500 ML pseudo-replicates.

Mrbayes v3.2 [37] was used to infer the Bayesian inference (BI) tree. The BI analysis was run for 20 million generations, in which a tree was sampled every 1000 generations. Two independent Markov Chain Monte Carlo (MCMC) analyses were performed and each chain started with a random tree. The first 25% of the sampled trees was discarded as burn-in, while the remaining trees were constructed in a majority-rule consensus tree to estimate posterior probabilities.

Molecular clock dating

We used BEAST v2.5.1 [38] to estimate the divergence times of Hamamelidaceae using three priors based on the complete plastome sequences. Based on the average value obtained by Xiang et al. [9] in a calibrated analysis, three priors were used: (i) the average age of the most recent common ancestor (TMRCA) of Hamamelidaceae (the root of the tree) was 108 Ma; (ii) the crown age of Hamamelideae/Fothergilleae was 89 Ma; and (iii) the crown age of Mytilarioideae was 58.3 Ma. Each secondary prior was placed under a normal distribution with a standard deviation of 1.

The GTR nucleotide substitution model and the prior tree Yule model were selected with the uncorrelated lognormal distribution relaxed molecular clock model. The MCMC run had a chain length of 400,000,000 generations with sampling every 10,000 generations. The stationary phase was examined through Tracer 1.6 [39] to evaluate convergence and to ensure sufficient and effective sample size for all parameters surpassing 200. A burn-in of 10% generations was discarded, and TreeAnnotator v2.4.7 was used to produce a maximum clade credibility tree.

Availability of data and materials

The chloroplast genome of Distylium assembled in this study have been deposited in the National Center for Biotechnology and Information (NCBI) under the following accession as summarized in Table 1. The other sequences used in this study were downloaded from the NCBI.

Abbreviations

BI:

Bayesian Inference

bp:

Base pairs

Gb:

Gigabases

LSC:

Long single copy

Ma:

Million years ago

MCMC:

Markov chain Monte Carlo

ML:

Maximum likelihood

NCBI:

National Center for Biotechnology Information

NGS:

Next generation sequencing

π:

Nucleotide diversity

rRNA:

Ribosomal RNA

SSC:

Short single copy

SSR:

Simple sequence repeat

tRNA:

Transfer RNA

References

  1. Liu Z, Cheng R, Xiao W, Guo Q, Wang N. Effect of off-season flooding on growth, photosynthesis, carbohydrate partitioning, and nutrient uptake in Distylium chinense. PLoS One. 2014;9(9):e107636. https://doi.org/10.1371/journal.pone.0107636.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Xiang L, Li X-L, Wang X-S, Yang J, Lv K, Xiong Z-Q, et al. Genetic diversity and population structure of Distylium chinense revealed by ISSR and SRAP analysis in the three gorges reservoir region of the Yangtze River, China. Glob Ecol Conserv. 2020;21:e00805. https://doi.org/10.1016/j.gecco.2019.e00805.

    Article  Google Scholar 

  3. Qin H, Yang Y, Dong S, He Q, Jia Y, Zhao L, et al. Threatened species list of China’s higher plants. Biodivers Sci. 2017;25(7):696–744. https://doi.org/10.17520/biods.2017144.

    Article  Google Scholar 

  4. Yagi H, Xu J, Moriguchi N, Miyagi R, Moritsuka E, Sato E, et al. Population genetic analysis of two species of Distylium: D. racemosum growing in East Asian evergreen broad-leaved forests and D. lepidotum endemic to the Ogasawara (Bonin) Islands. Tree Genet Genomes. 2019;15(6):77.

    Article  Google Scholar 

  5. Walker EH. A revision of Distylium and Sycopsis (Hamamelidaceae). J Arnold Arboretum. 1944;25(3):319–41. https://doi.org/10.5962/bhl.part.15299.

    Article  Google Scholar 

  6. Shi S, Chang HT, Chen Y, Qu L, Wen J. Phylogeny of the Hamamelidaceae based on the ITS sequences of nuclear ribosomal DNA. Biochem Syst Ecol. 1998;26(1):55–69. https://doi.org/10.1016/S0305-1978(97)00075-6.

    Article  CAS  Google Scholar 

  7. Li J, Bogle AL, Klein AS. Phylogenetic relationships in the Hamamelidaceae: evidence from the nucleotide sequences of the plastid gene matK. Plant Syst Evol. 1999;218(3):205–19. https://doi.org/10.1007/BF01089228.

    Article  CAS  Google Scholar 

  8. Li J, Bogle AL, Klein AS. Phylogenetic relationships of the Hamamelidaceae inferred from sequences of internal transcribed spacers (ITS) of nuclear ribosomal DNA. Am J Bot. 1999;86(7):1027–37. https://doi.org/10.2307/2656620.

    Article  CAS  PubMed  Google Scholar 

  9. Xiang X, Xiang K, Ortiz RDC, Jabbour F, Wang W. Integrating palaeontological and molecular data uncovers multiple ancient and recent dispersals in the pantropical Hamamelidaceae. J Biogeogr. 2019;46(11):2622–31. https://doi.org/10.1111/jbi.13690.

    Article  Google Scholar 

  10. Dong W, Xu C, Li W, Xie X, Lu Y, Liu Y, et al. Phylogenetic resolution in Juglans based on complete chloroplast genomes and nuclear DNA sequences. Front Plant Sci. 2017;8:1148. https://doi.org/10.3389/fpls.2017.01148.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Dong W, Xu C, Wu P, Cheng T, Yu J, Zhou S, et al. Resolving the systematic positions of enigmatic taxa: manipulating the chloroplast genome data of Saxifragales. Mol Phylogenet Evol. 2018;126:321–30. https://doi.org/10.1016/j.ympev.2018.04.033.

    Article  CAS  PubMed  Google Scholar 

  12. Guo L, Guo S, Xu J, He L, Carlson JE, Hou X. Phylogenetic analysis based on chloroplast genome uncover evolutionary relationship of all the nine species and six cultivars of tree peony. Ind Crop Prod. 2020;153:112567. https://doi.org/10.1016/j.indcrop.2020.112567.

    Article  CAS  Google Scholar 

  13. Chen X, Zhou J, Cui Y, Wang Y, Duan B, Yao H. Identification of Ligularia Herbs Using the Complete Chloroplast Genome as a Super-Barcode. Front Pharmacol. 2018;9:695.

    Article  PubMed  PubMed Central  Google Scholar 

  14. Krawczyk K, Nobis M, Myszczynski K, Klichowska E, Sawicki J. Plastid super-barcodes as a tool for species discrimination in feather grasses (Poaceae: Stipa). Sci Rep. 2018;8(1):1924. https://doi.org/10.1038/s41598-018-20399-w.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Dong W, Xu C, Wen J, Zhou S. Evolutionary directions of single nucleotide substitutions and structural mutations in the chloroplast genomes of the family Calycanthaceae. BMC Evol Biol. 2020;20(1):96. https://doi.org/10.1186/s12862-020-01661-0.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Dong W, Liu J, Yu J, Wang L, Zhou S. Highly variable chloroplast markers for evaluating plant phylogeny at low taxonomic levels and for DNA barcoding. PLoS One. 2012;7(4):e35071. https://doi.org/10.1371/journal.pone.0035071.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Dong W, Xu C, Li C, Sun J, Zuo Y, Shi S, et al. ycf1, the most promising plastid DNA barcode of land plants. Sci Rep. 2015;5(1):8348. https://doi.org/10.1038/srep08348.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Li W, Liu Y, Yang Y, Xie X, Lu Y, Yang Z, et al. Interspecific chloroplast genome sequence diversity and genomic resources in Diospyros. BMC Plant Biol. 2018;18(1):210. https://doi.org/10.1186/s12870-018-1421-3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Hamzeh M, Dayanandan S. Phylogeny of Populus (Salicaceae) based on nucleotide sequences of chloroplast TRNT-TRNF region and nuclear rDNA. Am J Bot. 2004;91(9):1398–408. https://doi.org/10.3732/ajb.91.9.1398.

    Article  CAS  PubMed  Google Scholar 

  20. Shaw J, Lickey EB, Schilling EE, Small RL. Comparison of whole chloroplast genome sequences to choose noncoding regions for phylogenetic studies in angiosperms: the tortoise and the hare III. Am J Bot. 2007;94(3):275–88. https://doi.org/10.3732/ajb.94.3.275.

    Article  CAS  PubMed  Google Scholar 

  21. Wheeler GL, McGlaughlin ME, Wallace LE. Variable length chloroplast markers for population genetic studies in Acmispon (Fabaceae). Am J Bot. 2012;99(10):e408–10. https://doi.org/10.3732/ajb.1200129.

    Article  PubMed  Google Scholar 

  22. Li J, Wang S, Jing Y, Wang L, Zhou S. A modified CTAB protocol for plant DNA extraction. Chin Bull Bot. 2013;48(1):72–8.

    Article  Google Scholar 

  23. Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014;30(15):2114–20. https://doi.org/10.1093/bioinformatics/btu170.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS, et al. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J Comput Biol. 2012;19(5):455–77. https://doi.org/10.1089/cmb.2012.0021.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Huang DI, Cronk QCB. Plann: a command-line application for annotating plastome sequences. Appl Plant Sci. 2015;3(8):1500026. https://doi.org/10.3732/apps.1500026.

    Article  Google Scholar 

  26. Greiner S, Lehwark P, Bock R. OrganellarGenomeDRAW (OGDRAW) version 1.3.1: expanded toolkit for the graphical visualization of organellar genomes. Nucleic Acids Res. 2019;47(W1):W59–64. https://doi.org/10.1093/nar/gkz238.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 2013;30(4):772–80. https://doi.org/10.1093/molbev/mst010.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Frazer KA, Pachter L, Poliakov A, Rubin EM, Dubchak I. VISTA: computational tools for comparative genomics. Nucleic Acids Res. 2004;32(Web Server):W273–9. https://doi.org/10.1093/nar/gkh458.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Kumar S, Stecher G, Tamura K. MEGA7: molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol Biol Evol. 2016;33(7):1870–4. https://doi.org/10.1093/molbev/msw054.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Rozas J, Ferrer-Mata A, Sanchez-DelBarrio JC, Guirao-Rico S, Librado P, Ramos-Onsins SE, et al. DnaSP 6: DNA sequence polymorphism analysis of large data sets. Mol Biol Evol. 2017;34(12):3299–302. https://doi.org/10.1093/molbev/msx248.

    Article  CAS  PubMed  Google Scholar 

  31. Kalendar R, Khassenov B, Ramankulov Y, Samuilova O, Ivanov KI. FastPCR: an in silico tool for fast primer and probe design and advanced sequence analysis. Genomics. 2017;109(3):312–9. https://doi.org/10.1016/j.ygeno.2017.05.005.

    Article  CAS  PubMed  Google Scholar 

  32. Dong W, Liu H, Xu C, Zuo Y, Chen Z, Zhou S. A chloroplast genomic strategy for designing taxon specific DNA mini-barcodes: a case study on ginsengs. BMC Genet. 2014;15(1):138. https://doi.org/10.1186/s12863-014-0138-z.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Minh BQ, Schmidt HA, Chernomor O, Schrempf D, Woodhams MD, von Haeseler A, et al. IQ-TREE 2: new models and efficient methods for phylogenetic inference in the genomic era. Mol Biol Evol. 2020;37(5):1530–4. https://doi.org/10.1093/molbev/msaa015.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Castresana J. GBLOCKS: selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. EMBL. 2002;17:540.

    Google Scholar 

  35. Kozlov AM, Darriba D, Flouri T, Morel B, Stamatakis A. RAxML-NG: a fast, scalable and user-friendly tool for maximum likelihood phylogenetic inference. Bioinformatics. 2019;35(21):4453–5. https://doi.org/10.1093/bioinformatics/btz305.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Kalyaanamoorthy S, Minh BQ, Wong TKF, von Haeseler A, Jermiin LS. ModelFinder: fast model selection for accurate phylogenetic estimates. Nat Methods. 2017;14(6):587–9. https://doi.org/10.1038/nmeth.4285.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Ronquist F, Teslenko M, van der Mark P, Ayres DL, Darling A, Hohna S, et al. MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Syst Biol. 2012;61(3):539–42. https://doi.org/10.1093/sysbio/sys029.

    Article  PubMed  PubMed Central  Google Scholar 

  38. Bouckaert R, Heled J, Kuhnert D, Vaughan T, Wu CH, Xie D, et al. BEAST 2: a software platform for Bayesian evolutionary analysis. PLoS Comp Biol. 2014;10(4):e1003537. https://doi.org/10.1371/journal.pcbi.1003537.

    Article  CAS  Google Scholar 

  39. Rambaut A, Suchard M, Xie D, Drummond A: Tracer v1. 6. In.; 2014: Available from http://beast.bio.ed.ac.uk/Tracer.

Download references

Acknowledgments

We would like to thank the Plant DNA Bank of China in the Institute of Botany, Chinese Academy of Sciences for providing materials.

Funding

This study was funded by the National Key Research and Development Program of China (2017YFC1703700 & 2017YFC1703704), the National Natural Science Foundation of China (NSFC: 81891010 & 81891014).

Author information

Authors and Affiliations

Authors

Contributions

WD: Conceptualization, Methodology, Formal analysis, Investigation, Writing - original draft, Writing - review & editing. YL: Methodology, Data curation. CX: Resources, Writing - original draft. YG: Methodology, Software. QY: Resources, Methodology, Data curation. ZS: Resources, Writing - original draft, ZZ: Supervision, Writing - review & editing. JS: Supervision, Funding acquisition, Writing - original draft. The authors read and approved the final manuscript.

Corresponding authors

Correspondence to Zhixiang Zhang or Jiahui Sun.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1: Table S1.

List of genes found in the Distylium chloroplast genome.

Additional file 2: Table S2.

More detail of SSRs in Distylium species.

Additional file 3: Table S3.

The primers used for amplification the variable markers.

Additional file 4: Figure S1.

Visualization of the alignment of chloroplast genome sequences of Distylium. VISTA-based similarity graphical information illustrating the sequence identity of Distylium with reference D. chinese chloroplast genomes. The Y-scale axis represents the percent identity within 50–100%.

Additional file 5: Figure S2.

Gel profiles of fragments amplified from two species using four pairs of primers.

Additional file 6: Figure S3.

ML tree for Distylium using four highly variable regions combinations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dong, W., Liu, Y., Xu, C. et al. Chloroplast phylogenomic insights into the evolution of Distylium (Hamamelidaceae). BMC Genomics 22, 293 (2021). https://doi.org/10.1186/s12864-021-07590-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12864-021-07590-6

Keywords