Skip to main content

Comparative chloroplast genomes of Dactylicapnos species: insights into phylogenetic relationships

Abstract

Background

Dactylicapnos is a climbing herbaceous vine, distributed from the Himalayas to southwestern China, and some of the species have important medicinal values. However, the chloroplast genomes of Dactylicapnos have never been investigated. In this study, chloroplast genomes of seven Dactylicapnos species covering all three sections and one informal group of Dactylicapnos were sequenced and assembled, and the detailed comparative analyses of the chloroplast genome structure were provided for the first time.

Results

The results showed that the chloroplast genomes of Dactylicapnos have a typical quadripartite structure with lengths from 172,344 bp to 176,370 bp, encoding a total of 133–140 genes, containing 88–94 protein-coding genes, 8 rRNAs and 37–39 tRNAs. 31 codons were identified as relative synonymous codon usage values greater than one in the chloroplast genome of Dactylicapnos genus based on 80 protein-coding genes. The results of the phylogenetic analysis showed that seven Dactylicapnos species can be divided into three main categories. Phylogenetic analysis revealed that seven species form three major clades which should be treated as three sections.

Conclusions

This study provides the initial report of the chloroplast genomes of Dactylicapnos, their structural variation, comparative genomic and phylogenetic analysis for the first time. The results provide important genetic information for development of medical resources, species identification, infrageneric classification and diversification of Dactylicapnos.

Peer Review reports

Introduction

Dactylicapnos Wall. belongs to Fumarioideae (DC.) Endlicher. in the family Papaveraceae Juss., established by Wallich in 1826 [1]. There are about 15 species in the genus, distributed from the Himalayas to southwestern China [2,3,4,5]. Dactylicapnos is a climbing herbaceous vine, distinguished by its branched tendrils at the end of leaves, pendent raceme inflorescences, yellow bisymmetric flowers and subquadrangular stigma with a papilla on each corner [2]. Taxa of Dactylicapnos are rich in active ingredients such as isoquinoline alkaloids [6], with the highest content of isocorydine and protopine [7], and is used in a Bai Nationality folk medicine due to its analgestic, anti-inflammatory, hemostatic and anti-hypertensive effects [8].

Despite some species of Dactylicapnos are very important medicinal plants, the relationship between these species is not clear. Recent classification of Dactylicapnos by Lidén and Pathak [5] based on morphology divided the genus into three sections (sect. Dactylicapnos, sect. Minicalcara and sect. Pogonosperma) and two informal groups. Validity of these morphologically defined sections and informal groups could not be confirmed due to lack of a systematic molecular study of the genus. A few molecular studies performed to date included only a few species of the genus, Lidén [9] used rps16 fragments to explore the systematic relationship between Dicentra and Dactylicapnos, Pérez-Gutiérrez [10, 11] conducted a molecular phylogenetic study of the Fumarioideae using five plastid markers, which included only four Dactylicapnos species, and only three species were included in the study by Chen [12]. There has never been a systematic molecular study to resolve the genus infrageneric phylogeny based on cp genome. Thus, the monophyly of these morphologically defined sections and informal groups could not be confirmed to verify the Lidén and Pathak [5] classification.

Chloroplasts are important organelles for photosynthesis in green plants which genome uniparently inherited. The chloroplast (cp) genome size of angiosperms is in the range of 120 to 160 kb [13], with a typical quadripartite structure, consisting of two-copy inverted repeat (IR) of 20–28 kb, a large single-copy regon (LSC) of about 80–90 kb and a small single-copy region of 16–27 kb [14], usually encoding for 120–150 genes. It is known that the cp genome encodes all the tRNA and rRNA molecules and partial proteins required for its own function [15,16,17]. Due to the highly conservative structure, rich in genetic information [18], slow nucleotide substitution rate [19], and uniparental inheritance [20], the cp genome has been widely used in phylogenetics analyses and identifications [21]. At present, chloroplast genome sequences of many species have been published, but the species of Dactylicapnos has not been published yet. The lack of systematic molecular studies hampers the development and application of Dactylicapnos. This motivated the current study of a comparative genomic analysis of seven of Dactylicapnos species covering all three sections and one informal group, in order to understanding the evolution of the genus structure and clarification of the phylogenetic relationship in Dactylicapnos species.

Results

Chloroplast genome structure and characteristics analyses of Dactylicapnos species

The lengths of the studied cp genomes varied from 172,344 bp (Dactylicapnos schneideri (Fedde) Lidén.) to 176,370 bp (Dactylicapnos grandifoliolata Merrill.), with a typical quadripartite structure, a pair of IR regions (28,530 bp–37,115 bp), LSC regions (89,195 bp–101,092 bp) and SSC (9303 bp–26,089 bp) (Fig. 1; Table 1). In the studies species there was an identical level of GC content with the total content 40.0%–40.6%, 41.6%–43.6% in IR, 39.1%–39.4% in LSC, and 35.3%–38.0% in SSC. The GC content of IR region was higher than LSC and SSC.

Fig. 1
figure 1

Gene map of the chloroplast genomes of D. scandens. Genes inside the circle are transcribed clockwise, and those on the outside are transcribed counter-clockwise. Genes belonging to different functional groups have been color-coded. The darker grey area in the inner circle corresponds to GC content, while the lighter grey corresponds to AT content

Table 1 The basic chloroplast genome information of seven Dactylicapnos species

The seven cp genomes have 133–140 genes, including 88–94 protein-coding genes, 37–39 tRNA genes, and 8 rRNA genes (Table 2). In the studied species, there are 18–26 genes with two copies, which were mostly comprised of seven protein-coding genes (ycf2, ycf15, ycf68, rps12, rps7, ndhB, ndhF), seven tRNA genes (trnI-CAU, trnL-CAA, trnR-AGC, trnA-UGC, trnI-GAU, trnV-GAC, trnN-GUU), and four rRNA (rrn5, rrn4.5, rrn23, rrn16), but the D. grandifoliolata also has six protein-coding genes (rpl32, ccsA, ndhD, psaC, ndhE, ndhG), two tRNA genes (trnH-GUG, trnL-UAG), D. schneideri also has one tRNA gene (trnH-GUG), and Dactylicapnos scandens Hutch. also have seven protein-coding genes (rpl32, ccsA, ndhD, psaC, ndhE, ndhG, ndhI) and one tRNA gene (trnL-UAG). Sixteen genes (trnG-UCC, atpF, rpoC1, trnL-UAA, trnV-UAC, petB, petD, rpl16, rpl2, trnA-UGC, trnI-GAU, rps12, ndhB, ndhA, trnK-UUU, rps16) contain a single intron, two genes (ycf3, clpP) have two introns, and the gene trnK-UUU has the largest intron, which contains the matK gene.

Table 2 The basic chloroplast genome information of seven Dactylicapnos species

Repeat sequence analysis

The studied cp genomes contained 546–878 dispersed repeats, including 360–467 forward repeats (F), 181–410 palindromic repeats (P), 3–23 complement repeats (C), and 1–20 reverse repeats (R), but some Dactylicapnos species do not have complement repeats and reverse repeats (Fig. 2A). Forward repeat was the most universal type, and most dispersed repeats were distributed in two-copy inverted repeat (IR) and large single-copy region (LSC) (Fig. 2B).

Fig. 2
figure 2

Dispersed repeated sequences analyses in cp genomes for seven Dactylicapnos species. A Statistics of four types of dispersed repeats sequences in seven cp genomes; Distribution of dispersed repeats sequences in seven cp genomes

The number of the tandem repeats ranged in the studied cp genomes from 54 to 72. Dactylicapnos torulosa (Hook.f. & Thomson) Hutch. had the most tandem repeats and Dactylicapnos macrocapnos Hutch. had the smallest (Fig. 3A). There were 4 cases of distribution of tandem repeated in the region, distributed in IRa/LSC region, IRb/LSC region, LSC region, SSC/LSC region, most of the tandem repeats are distributed in LSC region (Fig. 3B).

Fig. 3
figure 3

Tandem repeated sequence analyses for seven Dactylicapnos species. A Number of tandem repeats sequences in seven cp genomes; Distribution of tandem repeats sequences in seven cp genomes

Simple sequence repeats (SSRs) analyses

The SSRs were mainly distributed in the LSC region of Dactylicapnos species (Fig. 4A). A total of 327 SSRs were detected in the seven cp genomes, and the number of SSRs ranges from 37 (Dactylicapnos roylei Hutch.) to 60 (D. grandifoliolata), which had the largest number of mononucleotides (35–47), dinucleotides (2–9), trinucleotides (2), hexanucleotides (2), but some Dactylicapnos species did not have trinucleotides and hexanucleotides (Fig. 4B). These SSRs were dominated by mononucleotides (A/T) n. (Fig. 4C), suggesting that the base composition of SSRs is biased toward A/T base.

Fig. 4
figure 4

SSRs analyses for seven Dactylicapnos species. A The number of SSRs in LSC, SSC and IRs in seven cp genomes; Number of different SSRs types; Frequency of identified SSRs in different repeat class types

Codon usage analysis

In total, 64 types of codons encoding 20 amino acids were detected, including three termination codons, UAA(*), UAG(*) and UGA(*). The number of codons ranged from 22,187 to 23,325, with the highest number of codons found in D. schneideri, and the lowest number of codons found in Dactylicapnos lichiangensis (Fedde) Hand.-Mazz..

Relative synonymous codon usage (RSCU) values reflect a relationship between the number of actual codon emergence and the number of anticipated codon emergence [22], so that if the RSCU > 1, this mean that the condon has the strong preference. The RCUS calculated from 80 common CDS of the cp genomes of the studied species showed that all protein-coding sequence, 31 codons have RSCU > 1 (strong preference), 31 codons have RSCU < 1 (low preference), Methionine (Met) and threonine (Thr) have no bias (RSCU = 1). (Fig. 5).

Fig. 5
figure 5

The RSCU analysis of 80 common CDS in the chloroplast genomes of Dactylicapnos

IR contraction and expansion

There were differences in the boundary regions of the studied species. In D. macrocapnos, D. scandens, D. schneideri, D. torulosa, and D. lichiangensis, rpl23 was 160–240 bp to the left of the LSC/IRb boundary and trnI was 152–432 bp to the right of the LSC/IRb. The ndhA gene of D. macrocapnos and D. scandens covered the junction of SSC/IRa showed different sizes with 2183 bp and 2184 bp, extending into IRa by 1097 bp and SSC region by 1086 bp and 1087 bp. The ndhI gene of D. grandifoliolata also covered the junction of SSC/IRa, extending into IRa by 333 bp and SSC region by 162 bp. The gene trnH of D. schneideri and D. grandifoliolata was distributed on the left side of the border of IRa/LSC, and the gene trnH of the other species was distributed to the right of the IRa/LSC junction, with an interval of 38–127 bp from the border to the gene (Fig. 6).

Fig. 6
figure 6

Comparison of LSC, IR, and SSC junction positions among seven Dactylicapnos species in cp genomes. JLB denotes the LSC/IRb junction,JSB denotes the SSC/IRb junction, JSA denotes the SSC/IRa junction, and JLA denotes the LSC/IRa junction

Similarity analysis and synteny analysis

Analysis of the level of divergence among the studied species sequences, with D. roylei as a reference, done by mVISTA revealed that the bulk of among sequence variation is located in non-coding intergenic regions and that there were apparent deletions between the coding genes rps3rpl2 of D. lichiangensis, D. torulosa and D. grandifoliolata (Fig. 7).

Fig. 7
figure 7

Visualization of genome alignment of the chloroplast genomes of seven Dactylicapnos species using D. roylei as a reference by mVISTA. The x-axis represents the coordinate in the chloroplast genome. The Y-axis represents different species, and sequence similarity of aligned regions is displayed as horizontal bars, which expresses as a percentage within 50–100%

The synteny analysis revealed some genomic rearrangements and inversions in the seven cp genomes. Due to the expansion of IR region, the nucleotide sequences of in cp genomes of D. schneideri and D. grandifiliolata was rearranged, and some single copy regions of D. torulosa, D. macrocapnos, D. scandens and D. grandifoliolata were inverted (Fig. 8).

Fig. 8
figure 8

Synteny analysis for seven Dactylicapnos species in chloroplast genomes

Phylogenetic analysis

The maximum likelihood (ML) and bayesian inference (BI) phylogenetic trees (Fig. 9) were constructed using 78 common CDS of the cp genomes of 10 Fumarioideae species, including seven newly sequenced Dactylicapnos species and three outgroups including Lamprocapnos spectabilis (L.) Fukuhara, Corydalis adunca Maxim. and Corydalis edulis Maxim.. The ML and BI methods yielded identical tree topologies with full support for each node (MLBS = 100% and BIPP = 1). The genus was found to be monophyletic and the species of Dactylicapnos formed three distinct clades. The clade consistsing of D. schneideri, and D. grandifoliolata were sister to the rest of Dactylicapnos. D. schneideri formed an independent informal group, and was clustered together with D. grandifoliolata of sect. Pogonosperma. Section Dactylicapnos including D. scandens and D. macrocapnos were sister to sect. Minicalcara including D. lichiangensis, D. roylei and D. torulosa with full support (MLBS = 100% and BIPP = 1).

Fig. 9
figure 9

Maximum likelihood and Bayesian inference phylogeny of Dactylicapnos based on 78 common CDSs of 10 chloroplast genomes. From left to right, Numbers above branches indicate Bayesian posterior probability [BIPP], and Maximum Likelihood bootstrap support [MLBS], respectively. (A, D. macrocapnos; B, D. torulosa; C, D. grandifoliolata; D, D. schneideri)

Discussion

Structure and comparative analysis of Dactylicapnos species

Comparative analysis of cp genomes has been widely used in many plant taxa [23]. In this study, the cp genome of seven Dactylicapnos species were first sequenced, it is also the first time to explore Dactylicapnos species from the molecular analysis. As in the most angiosperms, the cp genome of Dactylicapnos has a typical quadripartite structure [14] but is very long 172,322–176,370 bp being one of the largest cp genomes sequenced to date [24], and the genomic size of the SSC region ranges from 9303 bp to 26,089 bp, with a number difference of about 16 kb, indicating the weakest conservatism and stability. The seven cp genomes are similar in structure, and had from 133 to 140 genes, which indicates that Dactylicapnos cp genomes are structurally conserved and rich in genetic information, which is a reliable molecular material for phylogenetic studies.

The highly conservative IR region is thought to play an important role in stabilizing the chloroplast genome structure [25]. Expansion and contraction of the IR region is a common phenomenon in plant evolutionary history responsible for cp genome length variation [26], which affects the cp genome’s rate of evolution [27, 28], examples are early-diverging eudicots [29, 30] and Apiales [31]. There have been many research about the expansion and contraction of the IR region, and the expansion mechanism of the IR region, the major viewpoint is that minor and apparently random IR expansion may be caused by gene conversion, and larger IR expansion may be achieved through double-strand DNA breaks and subsequent repair mechanism [32, 33], and the contraction mechanism of the IR region is also assumed to be the double-strand DNA breaks and subsequent repair mechanism [34]. In the present study, the IR region has significant expansion or contraction, forming a variety of boundary genes, and the seven cp genomes can be divided into three types according to their variability, which are consistent with the clustering results of the phylogenetic analysis. The gene location information in the boundary region can reveal the phylogenetic relationships between species to some extent [35]. In addition, as the expansion of the IR region at the LSC-IRb boundary, the trnH gene of D. schneideri and D. grandifoliolata entered the IR region leading to genomic rearrangement of these two sequences and the trnH gene becames gene with two copies. The chloroplast genome has multiple copies in the cell and has sufficient interspecific differentiation [35], chloroplast genome sequences for species identification is one of the best methods at present [36], while the cp genome of Dactylicapnos species have significant differences in expansion and contraction, and there are obvious differences in the size of the LSC, SSC, and IR regions of seven cp genomes, suggesting that Dactylicapnos species have a high degree of interspecies differentiation, which can be utilized to adequately demonstrate the phylogenetic relationships between Dactylicapnos species through the cp genomes.

Repeat sequences and SSRs

The plastid genome contains many oligonucleotide repeat sequences that are considered biomarkers of mutational hotspots [37, 38]. Repeat sequences have an important position in genome rearrangements and an important molecular marker in phylogenetic studies [39, 40]. In this present study, four different types of repeat sequences were detected, with the highest number of forward repeats (F) and the lowest number of complement repeats (C). The composition of different types of repeat sequences affects the inheritance and evolution of species [41]. There are small differences in the number and type of repeats among closely related species, both D. schneideri and D. grandifoliolata have four types of repetitive sequences with high similarity in type and number, inferring that the two species may have similarities in genetics and evolution [42]. SSRs are repeated DNA motifs with 1–6 nucleotides and have high polymorphism rates at the species level, have been extensively investigated in population genetics, phylogeography and variety identification [43, 44]. In this study, we found that the types of SSRs in seven cp genomes were found to be essentially the same, but the number of sequences contained in each type was different. Most SSRs loci were distributed in LSC region, with size ranging from 10–125 bp. The mononucleotide (A/T) was the highest proportion in the cp genomes of seven Dactylicapnos species, were found in all species. SSRs polymorphisms are repeat length polymorphisms caused by elongation or shortening of repeat units [45], it is a common molecular tool used to study the evolution of species. In the Camellia [46] and Triticum [47] plant, genetic diversity analysis was performed by amplifying SSRs primers, which led to the construction of genetic evolutionary relationships among species. The large number of SSRs detected in this research can be used as potential molecular markers for subsequent studies of Dactylicapnos species and also provide a theoretical basis for interspecific identification.

Codon usage analysis

The codons that encode the same amino acid are called synonymous codons [48]. In the process of species evolution, synonymous codons are not only associated with nature selection, mutation and genetic drift [49, 50], but also affected by factors such as genome size [51], tRNA abundance [52, 53] and gene expression levels [54], resulting in the genetic codes of different species tend to use one of several synonymous codons, called codon usage bias, which a common feature of eukaryotic genomes and is essential for the regulation of gene expression [55]. The results of the codon usage analysis showed that 31 codons had RSCU values > 1, indicating a codon bias in the amino acids, but unlike other dicotyledons plants [56], these 31 codons of Dactylicapnos do not prefer to end in A/U, It is possible that different levels of evolutionary pressures in Dactylicapnos species have biased the use of codons in this chloroplast genome, but the mechanisms involved need to be further explored [57, 58].

Phylogenetic analysis

The cp genome sequences have been successfully used to reveal phylogenetic relationships [59]. However, due to the different degree of gene rearrangement and inversion in Dactylicapnos, there are significant differences in gene order between sequences, and reliable phylogenetic relationships could not be established using the whole chloroplast genome. The analysis of 78 common CDS from the cp genomes of seven Dactylicapnos species and three outgroups showed that seven Dactylicapnos species were divided into three major clades with full support. Recent classification of Dactylicapnos based on morphology divided the genus into three sections and two informal groups [5]. Our study covered all three sections and one informal group, and our results basically clarified the infrageneric relationships between these three sections and one informal group. The first separated clade includes D. schneideri of an independent informal group sensu Lidén and Pathak [5] and D. grandifoliolata of sect. Pogonosperma sensu Lidén and Pathak [5]. The cp genomes of both D. schneideri and D. grandifoliolata had genomic rearrangements and contracted in the IR regions, and were clustered into the same clade, indicationg their close genetic relationship. The other two clades correspond to the two sections sensu Lidén and Pathak [5], sect. Dactylicapnos and sect. Pogonosperma, respectively. There were differences in the number of genes and GC content of these two clades, and there were also obvious differences in morphological characteristics. D. macrocapnos and D. scandens which were perennial plants with cylindrical stems and small flat globular elaiosomes [2], while D. torulosa, D. roylei and D. lichiangensis were all annual plants with winged-ridged stems and irregular mass elaiosomes [3]. D. scandens and D. macrocapnos of sect. Dactylicapnos were clustered together with full support, and D. lichiangensis, D. roylei and D. torulosa of sect. Minicalcara were also clustered together, so we confirmed sect. Dactylicapnos and sect. Minicalcara based on the plastome phylogenomics.

Conclusion

The cp genome of Dactylicapnos species had a typical tetrad structure and high sequence conservation. A total of 133–140 genes were annotated in the seven Dactylicapnos species, and a large number of repeat sequences and SSRs detected were important molecular markers in population genetics and phylogenetics. Expansion of IR regions and genomic rearrangements revealed by comparative genomic analysis played an important role in the evolution of Dactylicapnos species, and showed that the cp genomes of the D. macrocapnos and D. scandens were closer in structural variation, the D. schneideri was similar to and D. grandifoliolata, while the D. torulosa, D. lichiangensis and D. roylei were more consistent, which supported the results of phylogenetic analyses that categorized the seven species of Dactylicapnos into three clades. In addition, the most comprehensive and robust phylogeny covering all three sections and one informal group of Dactylicapnos based on cp genomes was reconstructed to basically clarify infrageneric relationships for the first time. Phylogenetic analysis showed that seven species separated into three major evolutionary clades, which suggested that this genus should be divided into three sections. The novel genomic resources provided here will aid future study in development of medicine resources, infrageneric classification, character evolution, diversification and biogeography. It also showed that the structural information and variation of chloroplast genomes were important for phylogenetic analysis, providing strong evidence for a deeper understanding of phylogenetic relationships and evolution among species.

Materials and methods

Plant material, DNA extraction and sequencing

Most of the material of Dactylicapnos species were fresh leaves collected in the field and dried with silica gel, and a few materials were obtained from the herbarium of KUN (Herbarium, Kunming Institute of Botany, CAS) and PE (Herbarium, Institute of Botany,CAS) (Table 3). The DNA extraction, library preparation and shallow sequencing were performed by Novogene, and the library was sequenced on the Illumina Hiseq 4000 platform with 150 bp paired-end reads. For the herbarium specimens, the method of Zeng et al. [60] was adopted for sequencing and library construction.

Table 3 Species Collection Information

Chloroplast genome assembly, annotation and codon usage

De novo assembly of the cp genome was carried out using GetOrganelle 1.7.6.1 [61]. We used the genome annotator PGA [62] to annotate the sequences that have been assembled into loops using the Lamprocapnos spectabilis (NC_039756) as the reference, and manually correct the position of the start and stop codons and the boundary between the exons and introns with Geneious Prime 2023.0.4 [63]. Finally, the physical maps of cp genome were created by using OrganellarGenomeDRAW (https://chlorobox.mpimp-golm.mpg.de/OGDraw.html) [64]. The RSCU was the ratio of the frequency of a specific codon to the expected frequency of that codon, which was obtained by Genepioneer platform, and plotted the heatmap of RSCU values with TBtools 1.116 [65].

Analysis of repeat sequences and SSRs

Repeat sequences in the cp genome were detected by REPuter [66], including forward, palindromic, reverse and complement repeats, the parameters were set with minimum repeat size 30 bp, and an hamming distance of 3. And exploring tandem repeats of cp genome by the Tandem Repeat Finder [67]. The simple sequence repeats (SSRs) were identified by using MISA online tool (https://webblast.ipk-gatersleben.de/misa/) [68], and the repeat thresholds for mononucleotide, dinucleotide, trinucleotide, tetrtanucleotide, pentanucleotide and hexanucleotide SSRs were 10, 6, 5, 5, 5, 5, respectively.

Comparative genomic analyses

The online program IRscope (https://irscope.shinyapps.io/irapp/) [69] was used to study the expansion and contraction of the IR region in the cp genome sequence of Dactylicapnos species. The genome comparison of the seven Dactylicapnos species in the cp genomes was analyzed by the mVISTA (https://genome.lbl.gov/vista/index.shtml) [70] program with the Shuffle-LAGAN mode, and the synteny analysis of cp genome was performed with Mauve [71].

Phylogenetic analysis

Phylogenetic analysis was performed based on 78 common CDS of the cp genomes of 10 Fumarioideae species, including seven Dactylicapnos cp genomes and three closely related species (Lamprocapnos spectabilis, Corydalis adunca and Corydalis edulis). These three species were selected as outgroups based on previous phylogenetic results[10,11,12], and these three plastomes were downloaded from GenBank. All sequences were aligned using MAFFT and maximum likelihood (ML) analysis was performed by RAxML-8.2.12 on CIPRES (https://www.phylo.org/portal2/) website with the GTRGAMM model, and 1000 bootstrap replicates. The best-fit model GTR + I + G was selected by AIC (Akaike Information Criterion) with jModelTest 2.1.10 [72], and the Bayesian inference (BI) analyses were conducted by MrBayes-3.2.7 on CIPRES website, with the settings: four MCMC simulations were run simultaneously and sampled every 1,000 generations for a total of two million generations, the first 25% of trees were discarded as burn-in.

Availability of data and materials

The datasets generated and analysed during the current study are available in the National Center for Biotechnology Information (NCBI) database, using the accession number OR568572, OR568573, OR589103, OR589104, OR589105, OR589106 and OR589107 (see Table 3 for details).

References

  1. Wallich N. Tentamen florae Napalensis illusrata, consisting of botanical descriptions and lithographic figures of select Nipal plants. Asiatic Lithographic Press. 1826;2:51–2.

  2. Zhang ML, Su ZY, Lidén M. Papaveraceae. In: Wu ZY, Raven PH, Hong DY, editors. Flora of China. Vol 7. Beijing: Science Press; 2008. p. 291–295.

  3. Wu ZY, Zhang X, Su ZY. Flora Reipublicae Popularis Sinicae. Beijing Science Press. 1999;32:88–93.

    Google Scholar 

  4. Lidén M. Three new species of Dactylicapnos (Fumariaceae) and a synopsis of the D. macrocapnos complex. Nord J Bot. 2010;28(6):656–60.

  5. Lidén M, Pathak MK. Studies in Dactylicapnos (Papaveraceae–Fumarioideae) part II Revision of Dactylicapnos sect. Pogonosperma sect. nov with D. arunachalensis sp nov. Nord J Bot. 2014;32(2):176–84.

  6. Wang FH, Hu X, Chen HL, Ma JP, Wang JX, Hou AJ. Alkaloids from Dactylicapnos scandens Hutch. China J Chin Materia Med. 2009;34(16):2057–9.

    CAS  Google Scholar 

  7. Guo CC. The metabolism and pharmacokinetics of isocorydine and protopine in Dactylicapnos scandens. Zhejiang University; 2013. p. 2–7.

  8. Wang B, Zhao YJ, Zhao YL, Liu YP, Li XN, Zhang HB, Luo XD. Exploring aporphine as anti-inflammatory and analgesic lead from Dactylicapnos scandens. Org Lett. 2019;22(1):257–60.

    Article  PubMed  Google Scholar 

  9. Lidén M, Fukuhara T, Rylander J, Oxelman B. Phylogeny and classification of Fumariaceae, with emphasis on Dicentra sl, based on the plastid gene rps16 intron. Plant Syst Evol. 1997;206:411–20.

    Article  Google Scholar 

  10. Perez-Gutierrez MA, Romero-Garcia AT, Salinas MJ, Blanca G, Fernandez MC, Suarez-Santiago VN. Phylogeny of the tribe Fumarieae (Papaveraceae s.l.) based on chloroplast and nuclear DNA sequences: evolutionary and biogeographic implications. American Journal Botany. 2012;99(3):517–28.

    Article  Google Scholar 

  11. Perez-Gutierrez MA, Romero-Garcia AT, Fernandez MC, Blanca G, Salinas-Bonillo MJ, Suarez-Santiago VN. Evolutionary history of fumitories (subfamily Fumarioideae, Papaveraceae): An old story shaped by the main geological and climatic events in the Northern Hemisphere. Molecular Phylogenetic and Evolution. 2015;88:75–92.

    Article  Google Scholar 

  12. Chen JT, Lidén M, Huang XH, Zhang L, Zhang XJ, Kuang TH, Landis JB, Wang D, Deng T, Sun H. An updated classification for the hyper-diverse genus Corydalis (Papaveraceae: Fumarioideae) based on phylogenomic and morphological evidence. J Integr Plant Biol. 2023;65(9):2138–56.

  13. Palmer JD. Comparative organization of chloroplast genomes. Annu Rev Genet. 1985;19(1):325–54.

    Article  CAS  PubMed  Google Scholar 

  14. Chumley TW, Palmer JD, Mower JP, Fourcade HM, Calie PJ, Boore JL, Jansen RK. The complete chloroplast genome sequence of Pelargonium× hortorum: organization and evolution of the largest and most highly rearranged chloroplast genome of land plants. Mol Biol Evol. 2006;23(11):2175–90.

    Article  CAS  PubMed  Google Scholar 

  15. Shinozaki K, Ohme M, Tanaka M, Wakasugi T, Hayashida N, Matsubayashi T, Zaita N, Chunwongse J, Obokata J, Yamaguchi-Shinozaki K, Ohto C, Torazawa K, Meng BY, Sugita M, Deno H, Kamogashira T, Yamada K, Kusuda J, Takaiwa F, Kato A, Tohdoh N, Shimada H, Sugiura M. The complete nucleotide sequence of the tobacco chloroplast genome: its gene organization and expression. EMBO J. 1986;5(9):2043–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Hiratsuka J, Shimada H, Whittier R, Ishibashi K, Sakamoto M, Mori M, Kondo C, Honji Y, Sun CR, Meng BY, Li YQ, Kanno A, Nishizawa Y, Hirai A, Shinozaki K, Sugiura M. The complete sequence of the rice (Oryza sativa) chloroplast genome: intermolecular recombination between distinct tRNA genes accounts for a major plastid DNA inversion during the evolution of the cereals. Mol Gen Genet MGG. 1989;217:185–94.

  17. Ohyama K, Fukuzawa H, Kohchi T, Shirai H, Sano T, Sano S, Umesono K, Shiki Y, Takeuch M, Chang Z, Aota SI, Inokuch H, Ozek H. Chloroplast gene organization deduced from complete sequence of liverwort Marchantia polymorpha chloroplast DNA. Nature. 1986;322(6079):572–4.

    Article  CAS  Google Scholar 

  18. Dong WP, Xu C, Cheng T, Lin K, Zhou S. Sequencing angiosperm plastid genomes made easy: a complete set of universal primers and a case study on the phylogeny of Saxifragales. Genome Biol Evol. 2013;5(5):989–97.

    Article  PubMed  PubMed Central  Google Scholar 

  19. Clegg MT, Gaut BS, Learn GH Jr, Morton BR. Rates and patterns of chloroplast DNA evolution. Proc Natl Acad Sci. 1994;91(15):6795–801.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Yang Z, Wang GX, Ma Q, Ma WX, Liang LS, Zhao TT. The complete chloroplast genomes of three Betulaceae species: implications for molecular phylogeny and historical biogeography. PeerJ. 2019;7:e6320.

    Article  PubMed  PubMed Central  Google Scholar 

  21. Wang YH, Wang S, Liu YL, Yuan QL, Sun JH, Guo LP. Chloroplast genome variation and phylogenetic relationships of Atractylodes species. BMC genomics. 2021;22(1):1–12.

    Google Scholar 

  22. Zhou JH, Zhang J, Chen HT, Ma LN, Liu YS. Analysis of synonymous codon usage in foot-and-mouth disease virus. Vet Res Commun. 2010;34:393–404.

    Article  PubMed  Google Scholar 

  23. Fan XG, Wang WC, Wagutu GK, Li W, Li XL, Chen YY. Fifteen complete chloroplast genomes of Trapa species (Trapaceae): insight into genome structure, comparative analysis and phylogenetic relationships. BMC Plant Biol. 2022;22(1):1–16.

    Article  Google Scholar 

  24. Hong Z, Wu ZQ, Zhao KK, Yang ZJ, Zhang NN, Guo JY, Tembrock LR, Xu DP. Comparative analyses of five complete chloroplast genomes from the genus Pterocarpus (Fabacaeae). Int J Mol Sci. 2020;21(11):3758.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Maréchal A, Brisson N. Recombination and the maintenance of plant organelle genome stability. New Phytologist Foundation. 2010;186(2):299–317.

    Article  Google Scholar 

  26. Wang RJ, Cheng CL, Chang CC, Wu TM, Chaw SM. Dynamics and evolution of the inverted repeat-large single copy junctions in the chloroplast genomes of monocots. BMC Evol Biol. 2008;8(1):1–14.

    Article  CAS  Google Scholar 

  27. Kim KJ, Lee HL. Complete chloroplast genome sequences from Korean ginseng (Panax schinseng Nees) and comparative analysis of sequence evolution among 17 vascular plants. DNA Res. 2004;11:247–61.

    Article  CAS  PubMed  Google Scholar 

  28. Zhang HY, Li C, Miao HM, Xiong SJ. Insights from the complete chloroplast genome into the evolution of Sesamum indicum L. PLoS ONE. 2013;8(11):e80508.

    Article  PubMed  PubMed Central  Google Scholar 

  29. Sun Y, Moore MJ, Zhang S, Soltis PS, Soltis DE, Zhao T, Meng A, Li X, Li J, Wang H. Phylogenomic and structural analyses of 18 complete plastomes across nearly all families of early-diverging eudicots, including an angiosperm-wide analysis of IR gene content evolution. Mol Phylogenet Evol. 2016;96:93–101.

    Article  PubMed  Google Scholar 

  30. Sun YX, Moore MJ, Meng AP, Soltis PS, Soltis DE, Li JQ, Wang HC. Complete plastid genome sequencing of Trochodendraceae reveals a significant expansion of the inverted repeat and suggests a Paleogene divergence between the two extant species. PLoS ONE. 2013;8(4):e60429.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Downie SR, Jansen RK. A comparative analysis of whole plastid genomes from the Apiales: expansion and contraction of the inverted repeat, mitochondrial to plastid transfer of DNA, and identification of highly divergent noncoding regions. Syst Bot. 2015;40(1):336–51.

    Article  Google Scholar 

  32. Goulding SE, Wolfe KH, Olmstead RG, Morden CW. Ebb and flow of the chloroplast inverted repeat. Mol Gen Genet MGG. 1996;252:195–206.

    Article  CAS  PubMed  Google Scholar 

  33. Wang RJ, Cheng CL, Chang CC, Wu CL, Su TM, Chaw SM. Dynamics and evolution of the inverted repeat-large single copy junctions in the chloroplast genomes of monocots. BMC Evol Biol. 2008;8(1):1–14.

    Article  CAS  Google Scholar 

  34. Peery RM. Understanding angiosperm genome interactions and evolution: insights from sacred lotus (Nelumbo nucifera) and the carrot family (Apiaceae). University of Illinois at Urbana-Champaign; 2015. p. 11–54.

  35. Chen MM, Zhang M, Liang ZS, He QL. Characterization and Comparative Analysis of Chloroplast Genomes in Five Uncaria Species Endemic to China. Int J Mol Sci. 2022;23(19):11617.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Hollingsworth PM, Graham SW, Little DP. Choosing and using a plant DNA barcode. PLoS ONE. 2011;6(5):13.

    Article  Google Scholar 

  37. Ibrar A, Biggs PJ, Matthews PJ, Collins LJ, Hendy MD, Lockhart PJ. Mutational Dynamics of Aroid Chloroplast Genomes. Genome Biol Evol. 2012;4(12):1316–23.

    Article  Google Scholar 

  38. Jungeun L, Kang Y, Chul SS, Hyun P, Hyoungseok L. Combined Analysis of the Chloroplast Genome and Transcriptome of the Antarctic Vascular Plant Deschampsia antarctica Desv. PLoS ONE. 2014;9(3):e92501.

    Article  Google Scholar 

  39. Cavalier-Smith T. Chloroplast evolution: secondary symbiogenesis and multiple losses. Curr Biol. 2002;12(2):R62–4.

    Article  CAS  PubMed  Google Scholar 

  40. Nazareno AG, Carlsen M, Lohmann LG. Complete chloroplast genome of Tanaecium tetragonolobum: the first Bignoniaceae plastome. PLoS ONE. 2015;10(6):e0129930.

    Article  PubMed  PubMed Central  Google Scholar 

  41. Keller J, Rousseau-Gueutin M, Martin GE, Morice J, Boutte J, Coissac E, Ourari M, Aïnouche M, Salmon A, Cabello-Hurtado F, Aïnouche A. The evolutionary fate of the chloroplast and nuclear rps16 genes as revealed through the sequencing and comparative analyses of four novel legume chloroplast genomes from Lupinus. DNA Res. 2017;24(4):343–58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. He Y, Xiao HT, Deng C, Xiong L, Yang J, Peng C. The complete chloroplast genome sequences of the medicinal plant Pogostemon cablin. Int J Mol Sci. 2016;17(6):820.

    Article  PubMed  PubMed Central  Google Scholar 

  43. Xue JH, Wang S, Zhou SL. Polymorphic chloroplast microsatellite loci in Nelumbo (Nelumbonaceae). Am J Bot. 2012;99(6):e240–4.

    Article  PubMed  Google Scholar 

  44. Zheng G, Wei LL, Ma L, Wu ZQ, Gu CH, Chen K. Comparative analyses of chloroplast genomes from 13 Lagerstroemia (Lythraceae) species: identification of highly divergent regions and inference of phylogenetic relationships. Plant Mol Biol. 2020;102:659–76.

    Article  CAS  PubMed  Google Scholar 

  45. Sonah H, Deshmukh RK, Sharma A, Singh VP, Gupta DK, Gacche RN, Rana JC, Singh NK, Sharma TR. Genome-wide distribution and organization of microsatellites in Plants: An insight into marker development in Brachypodium. PLoS ONE. 2011;6(6):e21298.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Wang LY, Liu BY, Jiang YH, Duan YS, Cheng H, Zhou J, Tang YC. Phylogenetic analysis of interspecies in section Thea through SSR markers. J Tea Sci. 2009;29(5):341–6.

    Google Scholar 

  47. Shi JX, Qiao YL, Ma YQ, Ji WQ, He PR, Weng YJ. Analysis on genetic evolution relation of A, B genomes between Triticum aestivum and T. dicoccoides by SSR. Acta Botanica Boreali Occidentalia Sinica. 2003;23(06):933–7.

  48. Lagerkvist ULF. “ Two out of three”: an alternative method for codon reading. Proc Natl Acad Sci. 1978;75(4):1759–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Bulmer M. The selection-mutation-drift theory of synonymous codon usage. Genetics. 1991;129(3):897–907.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Hershberg R, Petrov DA. Selection on codon bias. Annu Rev Genet. 2008;42(1):287–99.

    Article  CAS  PubMed  Google Scholar 

  51. Dong LN, Du XY, Zhou W. The complete plastid genome sequence of Begonia guangxiensis. Mitochondrial DNA Part B. 2019;4(2):3766–7.

    Article  PubMed  PubMed Central  Google Scholar 

  52. Duret L. tRNA gene number and codon usage in the C. elegans genome are co-adapted for optimal translation of highly expressed genes. Trends in Genetics. 2000;16(7):287–9.

  53. Olejniczak M, Uhlenbeck OC. tRNA residues that have coevolved with their anticodon to ensure uniform and accurate codon recognition. Biochimie. 2006;88(8):943–50.

    Article  CAS  PubMed  Google Scholar 

  54. Hiraoka Y, Kawamata K, Haraguchi T, Chikashige Y. Codon usage bias is correlated with gene expression levels in the fission yeast Schizosaccharomyces pombe. Genes Cells. 2010;14(4):499–509.

    Article  Google Scholar 

  55. Lyu XL, Liu Y. Nonoptimal codon usage is critical for protein structure and function of the master general amino acid control regulator CPC-1. Molecular Biology and Physiology. 2020;11(5):e02605–e2620.

    CAS  Google Scholar 

  56. Kawabe A, Miyashita NT. Patterns of codon usage bias in three dicot and four monocot plant species. Genes Genet Syst. 2003;78(5):343–52.

    Article  CAS  PubMed  Google Scholar 

  57. Kong WQ, Yang JH. The complete chloroplast genome sequence of Morus cathayana and Morus multicaulis, and comparative analysis within genus Morus L. PeerJ. 2017;5:e3037.

    Article  PubMed  PubMed Central  Google Scholar 

  58. Suzuki H, Morton BR. Codon adaptation of plastid genes. PLoS ONE. 2016;11(5):e0154306.

    Article  PubMed  PubMed Central  Google Scholar 

  59. Moore MJ, Bell CD, Soltis PS, Soltis DE. Using plastid genome-scale data to resolve enigmatic relationships among basal angiosperms. Proc Natl Acad Sci. 2007;104(49):19363–8.

    Article  PubMed  PubMed Central  Google Scholar 

  60. Zeng CX, Hollingsworth PM, Yang J, He ZS, Zhang ZR, Li DZ, Yang JB. Genome skimming herbarium specimens for DNA barcoding and phylogenomics. Plant Methods. 2018;14:1–14.

    Article  Google Scholar 

  61. Jin JJ, Yu WB, Yang JB, Song Y, de Pamphilis CW, Yi TS, Li DZ. GetOrganelle: a fast and versatile toolkit for accurate de novo assembly of organelle genomes. Genome Biology. 2020;21:1–31.

    Article  Google Scholar 

  62. Qu XJ, Moore MJ, Li DZ, Yi TS. PGA: a software package for rapid, accurate, and flexible batch annotation of plastomes. Plant Methods. 2019;15:1–12.

    Article  Google Scholar 

  63. Kearse M, Moir R, Wilson A, Stones-Havas S, Cheung M, Sturrock S, Buxton S, Cooper A, Markowitz S, Duran C, Thierer T, Ashton B, Meintjes P, Drummond A. Geneious Basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics. 2012;28(12):1647–9.

    Article  PubMed  PubMed Central  Google Scholar 

  64. Lohse M, Drechsel O, Bock R. OrganellarGenomeDRAW (OGDRAW): a tool for the easy generation of high-quality custom graphical maps of plastid and mitochondrial genomes. Curr Genet. 2007;52:267–74.

    Article  CAS  PubMed  Google Scholar 

  65. Chen C, Chen H, Zhang Y, Thomas HR, Frank MH, He YH, Xia R. TBtools: an integrative toolkit developed for interactive analyses of big biological data. Molecular Plant. 2020;13(8):1194–202.

    Article  CAS  PubMed  Google Scholar 

  66. Kurtz S, Schleiermacher C. REPuter: fast computation of maximal repeats in complete genomes. Bioinformatics (Oxford, England). 1999;15(5):426–7.

    CAS  PubMed  Google Scholar 

  67. Benson G. Tandem repeats finder: a program to analyze DNA sequences. Nucleic Acids Res. 1999;27(2):573–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Beier S, Thiel T, Münch T, Scholz U, Mascher M. MISA-web: a web server for microsatellite prediction. Bioinformatics. 2017;33(16):2583–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Amiryousefi A, Hyvönen J, Poczai P. IRscope: an online program to visualize the junction sites of chloroplast genomes. Bioinformatics. 2018;34(17):3030–1.

    Article  CAS  PubMed  Google Scholar 

  70. Frazer KA, Pachter L, Poliakov A, Rubin EM, Dubchak I. VISTA: computational tools for comparative genomics. Nucleic Acids Research. 2004;32(suppl_2):W273–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Darling ACE, Mau B, Blattner FR, Perna NT. Mauve: multiple alignment of conserved genomic sequence with rearrangements. Genome Res. 2004;14(7):1394–403.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Posada D. jModelTest: phylogenetic model averaging. Mol Biol Evol. 2008;25(7):1253–6.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank the staffs of Herbarium of Kunming Institute of Botany (KUN) for their help with this research. Thanks to Min-Shu Song (Kunming Institute of Botany, Chinese Academy of Sciences) for assistance on lab work. We are grateful to Peng-Rui Luo (Kunming Institute of Botany, Chinese Academy of Sciences), Bin Yang (Xishuangbanna Tropical Botanical Garden, Chinese Academy of Sciences), Xin-Tang Ma (Institute of Botany, Chinese Academy of Sciences) and Cheng Liu (Kunming Institute of Botany, Chinese Academy of Sciences) for providing materials and photos.

Data availability statement

We have uploaded the data to the NCBI website (https://www.ncbi.nlm.nih.gov/) and obtained the GenBank accession numbers, which is displayed in the Table 3.

Research involving plants

The sources of plant materials used in the study were collected, and the collection of plant material are licensed.

Funding

This research was funded by grants from the Second Tibetan Plateau Scientific Expedition and Research (STEP) program (2019QZKK0502), National Natural Science Foundation of China (32322006 and 32300173), Major Program for Basic Research Project of Yunnan Province (202101BC070002, 202005AB160005), the Key R&D Program of Yunnan (202103AF140005), Biological Resources Programme, Chinese Academy of Sciences (KFJ-BRP-017), the Key Projects of the Joint Fund of the National Natural Science Foundation of China (U23A20149).

Author information

Authors and Affiliations

Authors

Contributions

HS and TD conceived the study and acquired these fundings. SQY and JTC performed the data and drafted the earlier version of manuscript. ZML, KT, XHH, XZ and QL provided suggestions on structuring the article and revised the manuscript. All authors read and approved the final version of the manuscript.

Corresponding authors

Correspondence to Hang Sun or Tao Deng.

Ethics declarations

Ethics approval and consent to participate

The authors confirm that all methods comply with local and national regulations.

Consent for publication

Not applicable.

Competing of interests

The authors declare no competing interests.

Additional information

Publisher’s Note

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

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

Yang, S., Chen, J., Li, Z. et al. Comparative chloroplast genomes of Dactylicapnos species: insights into phylogenetic relationships. BMC Plant Biol 24, 350 (2024). https://doi.org/10.1186/s12870-024-04989-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12870-024-04989-7

Keywords