Next Article in Journal
Long-Chain and Medium-Chain Fatty Acids in Energy Metabolism of Murine Kidney Mitochondria
Previous Article in Journal
The Functional Characterization of Carboxylesterases Involved in the Degradation of Volatile Esters Produced in Strawberry Fruits
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Plastid Phylogenomics Provide Evidence to Accept Two New Members of Ligusticopsis (Apiaceae, Angiosperms)

Key Laboratory of Bio-Resources and Eco-Environment of Ministry of Education, College of Life Sciences, Sichuan University, Chengdu 610065, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(1), 382; https://doi.org/10.3390/ijms24010382
Submission received: 9 November 2022 / Revised: 20 December 2022 / Accepted: 22 December 2022 / Published: 26 December 2022
(This article belongs to the Section Molecular Plant Sciences)

Abstract

:
Peucedanum nanum and P. violaceum are recognized as members of the genus Peucedanum because of their dorsally compressed mericarps with slightly prominent dorsal ribs and narrowly winged lateral ribs. However, these species are not similar to other Peucedanum taxa but resemble Ligusticopsis in overall morphology. To check the taxonomic positions of P. nanum and P. violaceum, we sequenced their complete plastid genome (plastome) sequences and, together with eleven previously published Ligusticopsis plastomes, performed comprehensively comparative analyses. The thirteen plastomes were highly conserved and similar in structure, size, GC content, gene content and order, IR borders, and the patterns of codon bias, RNA editing, and simple sequence repeats (SSRs). Nevertheless, twelve mutation hotspots (matK, ndhC, rps15, rps8, ycf2, ccsA-ndhD, petN-psbM, psbA-trnK, rps2-rpoC2, rps4-trnT, trnH-psbA, and ycf2-trnL) were selected. Moreover, both the phylogenetic analyses based on plastomes and on nuclear ribosomal DNA internal transcribed spacer (ITS) sequences robustly supported that P. nanum and P. violaceum nested in Ligusticopsis, and this was further confirmed by the morphological evidence. Hence, transferring P. nanum and P. violaceum into Ligusticopsis genus is reasonable and convincing, and two new combinations are presented.

1. Introduction

Ligusticopsis Leute, a small flowering plant genus of Apiaceae, was established by Gerfried Horand Leute in 1969 with L. rechingeriana Leute as the type species, recognizable by conspicuous calyx teeth and strongly dorsally flattened mericarps with numerous vallecular vittae [1]. However, these characteristics can also be detected in some Ligusticum L. members [2]; therefore, the morphological delimitation between Ligusticopsis and Ligusticum is historically unclear. Furthermore, several Ligusticopsis species described by Leute do not have prominent calyx teeth [3], which further blurred the boundaries of this genus. Hence, the genus Ligusticopsis has been merged into the genus Ligusticum by some authors [2,4,5,6]. However, Li et al. [7] recently confirmed that the genus Ligusticopsis is a natural unit based on molecular and morphological evidence, and confirmed the presence of nine “true Ligusticopsis species”. Subsequently, the phylogenetic analyses based on plastome data performed by Ren et al. [8] also recovered Ligusticopsis as a monophyletic group and recognized two additional species. To date, the genus Ligusticopsis contains eleven validated species.
Peucedanum nanum R.H.Shan and M.L.Sheh and P. violaceum R.H.Shan and M.L.Sheh are species endemic to China, which grow on dry mountain slopes and in sparse forests or grassy places on riverbanks, respectively [9,10]. Both species are placed in Peucedanum L. owing to their dorsally compressed mericarps with slightly prominent dorsal ribs and narrowly winged lateral ribs [11]. However, P. nanum and P. violaceum are not similar to the type species of Peucedanum (P. officinale L.) [12] but resemble Ligusticopsis species in overall morphology (Figure 1). Additionally, the genus Peucedanum is not monophyletic [13,14,15,16,17,18,19], and its taxonomy has faced extreme challenges. Therefore, the taxonomic positions of P. nanum and P. violaceum need to be re-evaluated.
A robust molecular phylogenetic framework could provide valuable information for resolving the taxonomic positions of P. nanum and P. violaceum. Unfortunately, molecular data for both species are limited, and these species have not been included in previous phylogenetic studies. Hence, it is necessary to identify molecular markers to investigate the phylogenetic position of P. nanum and P. violaceum.
The plastid genome (plastome) sequence, possessing highly variable characters, gives us the potential to obtain a robust phylogenetic framework at low taxonomic levels [20,21,22,23,24,25,26,27,28,29]. With the development of next-generation sequencing, plastome sequences have been applied extensively and successfully to resolve the phylogenetic position of taxonomically difficult taxa [8,30,31,32,33,34,35]. In this study, we sequenced and assembled the plastomes of P. nanum and P. violaceum for the first time. Together with the previously published eleven Ligusticopsis plastomes, we carried out comprehensively comparative analyses to reveal the plastome features for P. nanum, P. violaceum, and Ligusticopsis species. Subsequently, we performed phylogenetic analyses based on the plastome data and nuclear ribosomal DNA internal transcribed spacer (ITS) sequences to investigate the phylogenetic positions of P. nanum and P. violaceum. Finally, by combining evidence from the comparative plastome analyses, molecular phylogeny, and morphology, taxonomic revisions for P. nanum and P. violaceum were conducted.

2. Results

2.1. Plastome Features

Illumina sequencing obtained 31,564,816 and 33,651,636 paired-end clean reads for P. nanum and P. violaceum, respectively (Table S1); among these reads, 1,025,204 and 543,725 reads were mapped to the assemblies, respectively. Based on these data, two high-quality plastomes for P. nanum and P. violaceum were generated with 1022.351× and 536.171× coverage, respectively.
The plastome features of eleven Ligusticopsis taxa and two Peucedanum species were comprehensively investigated. The overall size ranged from 146,900 bp (P. nanum) to 148,633 bp (L. brachyloba (Franch.) Leute) in the thirteen plastomes (Table 1). All of them exhibited typical quadripartite structures containing a pair of inverted repeat regions (IRs, 19,056–20,022 bp) divided by a large single-copy region (LSC, 91,480–92,305 bp) and a small single-copy region (SSC, 16,335–17,654 bp) (Table 1, Figure 2). The total GC content of the thirteen plastomes ranged from 37.3% to 37.5%, and 113 unique genes were identified, including 79 protein-coding genes, 30 tRNA genes, and four rRNA genes (Table 1 and Table S2).
The 79 protein-coding genes typically shared by the thirteen plastomes were extracted and connected for each species. These sequences were 67,566–67,896 bp in length and harbored 22,522–22,632 codons (Table S3). Among these codons, the least number of codons were used to encode the Cys, while the highest number of codons were used to encode the Leu. Additionally, the relative synonymous codon usage (RSCU) values of all codons ranged from 0.34 to 2.00 in the thirteen plastomes (Figure 3). Among them, the RSCU values of 30 codons were greater than 1.00 in all plastomes. All of these codons ended with A/U, except for UUG.
A total of 57–59 potential RNA editing sites were identified in the thirteen plastomes (Table S4). All detected RNA editing sites were Cytosine to Uracil (C-U) conversion, and most of them occurred in the second codon position (43–45) followed by the first codon position (14), but no site was located in the third codon position (Table S5). Moreover, the ndhB gene contained the highest number of RNA editing sites (10) in all plastomes (Table S6).
The total number of simple sequence repeats (SSRs) ranged from 67 to 84 among the thirteen plastomes (Figure 4). Among these, mononucleotide repeats were the most abundant (34–43) followed by dinucleotides (17–24). In addition, bases A and T were dominant for all the identified SSRs in the thirteen plastomes (Table S7).

2.2. Plastome Comparison

The borders between the IR and SC among the thirteen plastomes were compared (Figure 5). The junctions of IRa/LSC and IRb/LSC fell into the ycf2 gene and intergenic region of trnL-trnH, respectively. The borders of IRb/SSC fell into the ycf1 gene in all species, whereas the overlap between the ycf1 gene and the ndhF gene in the IRa/SSC junctions was only detected in L. capillacea (H.Wolff) Leute.
The gene arrangement among the thirteen plastomes was the same (Figure 6), and their sequences showed high similarity with 98.2% pairwise identity (Figure 7). Nevertheless, 12 mutation hotspot regions were identified, including five protein-coding genes (matK, ndhC, rps15, rps8, ycf2) that exhibited Pi > 0.00340 and 7 non-coding regions (ccsA-ndhD, petN-psbM, psbA-trnK, rps2-rpoC2, rps4-trnT, trnH-psbA, ycf2-trnL) that showed Pi > 0.01000 (Figure 8).

2.3. Phylogenetic Analyses

The analyses of maximum likelihood (ML) and Bayesian inference (BI) based on the plastome data generated identical tree topologies. As shown in Figure 9, the eleven Ligusticopsis taxa, P. nanum, and P. violaceum clustered as a clade (BI posterior probabilities, PP = 1.00, ML bootstrap values, BS = 100) within Selineae (PP = 1.00, BS = 100), which was clearly distant from other Ligusticum taxa. Within this clade, three lineages could be recognized: (1) L. daucoides (Franch.) Lavrova ex Pimenov and Kljuykov, L. hispida (Franch.) Lavrova and Kljuykov, L. involucrata (Franch.) Lavrova, L. oliveriana (H.Boissieu) Lavrova, and L. rechingeriana formed a clade (PP = 1.00, BS = 100) in which L. oliveriana early diverged from the reminders (PP = 1.00, BS = 100) followed by L. daucoides (PP = 1.00, BS = 100), and the sub-clade L. involucrata + L. rechingeriana (PP = 1.00, BS = 100) sister to L. hispida (PP = 1.00, BS = 100); (2) P. nanum and P. violaceum represented a clade (PP = 1.00, BS = 100); (3) the six remainders constituted another clade (PP = 1.00, BS = 82) in which L. capillacea + L. scapiformis (H.Wolff) Leute was sister to L. integrifolia (H.Wolff) Leute + L. modesta (Diels) Leute (PP = 1.00, BS = 100) and then clustered with L. brachyloba + L. wallichiana (DC.) Pimenov and Kljuykov (PP = 1.00, BS = 82).
Although phylogenetic analyses based on ITS sequences yielded topologies with low support and resolution, the results also indicated that the sister group of P. nanum and P. violaceum clustered with the Ligusticopsis species (PP = 1.00, BS = 97), and this clade was relatively distant from other Ligusticum taxa (Figure S1).

3. Discussion

3.1. Plastome Features

In this study, we conducted comprehensively comparative analyses for the plastomes of P. nanum, P. violaceum, and Ligusticopsis species. The thirteen plastomes showed typical quadripartite structures, including a pair of inverted repeat regions divided by a large single-copy region and a small single-copy region, which is the same as the other plastomes of Apiaceae [7,8,19,36,37,38,39,40,41]. Although gene loss and rearrangement have been reported in the plastomes of Apiaceae [19,38,39], the gene content and order in the thirteen studied plastomes were identical. All these plastomes also shared similar genomic size, total GC content, and IR borders. Furthermore, the patterns of codon bias, RNA editing sites, and SSR were extremely similar and have also been detected in the plastomes of Ligusticum and Peucedanum within Apiaceae [19,42]. These results indicated that the thirteen plastomes were highly conserved. Meanwhile, the conserved and similar plastome characters among P. nanum, P. violaceum, and Ligusticopsis species also implied that P. nanum and P. violaceum may be members of Ligusticopsis.
Although the thirteen plastomes showed high similarity, 12 mutation hotspot regions (matK, ndhC, rps15, rps8, ycf2, ccsA-ndhD, petN-psbM, psbA-trnK, rps2-rpoC2, rps4-trnT, trnH-psbA, ycf2-trnL) were still identified, which could be used as potential DNA barcodes for species identification and phylogenetic analysis of the Ligusticopsis species. Among them, the matK gene and the trnH-psbA fragment have been suggested as universal DNA barcodes [43,44,45,46], while the ycf2 gene and the petN-psbM region have been extensively used for phylogenetic analysis [47,48,49,50,51,52,53]. In future studies, the effects of these sequences on species identification and phylogenetic analysis of Ligusticopsis will be further validated.

3.2. Phylogenetic Inference

Since the establishment of the genus Ligusticopsis, its taxonomy has been controversial. Pu [2], Pu and Watson [5], Zhang [4], and Wang et al. [54] did not recognize Ligusticopsis as a distinct genus but merged it into Ligusticum based on morphological characteristics. However, based on carpoanatomical evidence, Pimenov et al. [55] accepted the establishment of Ligusticopsis. Subsequently, Pimenov [56] recognized 18 Ligusticopsis species in his checklist of Chinese Umbelliferae based on reviews of the type specimens and morphological evidence. Recently, plastome phylogenetic analyses performed by Li et al. [7] and Ren et al. [8] robustly confirmed the monophyly of Ligusticopsis, although limited samples of Ligusticopsis and Ligusticum were used in both studies. In the present study, twelve Ligusticum species and eleven Ligusticopsis taxa were included in the phylogenetic analyses. Both the phylogenies based on plastome data and on ITS sequences revealed that eleven Ligusticopsis species clustered as a clade and belonged to the Selineae tribe. Although the type species of Ligusticum (Ligusticum scoticum L.) was absent in our analyses, the phylogenetic position of this species, located in the Acronema Clade, was revealed by a previous study [18], which was obviously distant from the clade formed by the Ligusticopsis species. Our results with more extensive taxa sampling provided additional evidence to accept Ligusticopsis as a distinct genus.
Additionally, all our phylogenetic analyses based on plastome data and ITS sequences robustly supported that P. nanum and P. violaceum nested within Ligusticopsis. The type species of Peucedanum (P. officinale) was not included in our analyses; however, a previous study revealed its phylogenetic location was distant from Ligusticopsis [18]. These results implied that P. nanum and P. violaceum were distant from P. officinale but closely related to Ligusticopsis. Furthermore, the affinity between both species and Ligusticopsis was supported by the high similarity of their plastome sequences and also supported by the shared morphological features: stem base clothed in fibrous remnant sheaths, conspicuous calyx teeth, and strongly compressed dorsally mericarps with slightly prominent dorsal ribs, winged lateral ribs, and numerous vittae in the commissure and in each furrow [7,9,10]. However, hispid mericarps can easily distinguish P. nanum and P. violaceum from the glabrous mericarps of other Ligusticopsis species [7]. Moreover, P. nanum has densely hispid mericarps with slightly prominent dorsal ribs and six vittae in the commissure, whereas sparsely hispid mericarps with filiform dorsal ribs and eight vittae in the commissure are observed in P. violaceum [9,10]. Therefore, we could reasonably transfer P. nanum and P. violaceum into Ligusticopsis as two new members of this genus.
The sister relationship between P. nanum and P. violaceum was strongly supported by both the phylogenetic analyses based on plastome data and ITS sequences. The hispid mericarps shared by both species could further support this relationship [9,10]. Unfortunately, the relationship between this sister group and other Ligusticopsis species was not clearly resolved in our phylogenetic analyses. To confirm the phylogenetic position of the sister group of P. nanum and P. violaceum within Ligusticopsis, additional molecular sequences such as additional nuclear DNA fragments are required in future studies.

3.3. Taxonomic Treatment

Ligusticopsis nana (R.H.Shan and M.L.Sheh) C.K.Liu and X.J.He, comb. nov.
Peucedanum nanum R.H.Shan and M.L.Sheh in Act. Phytotax. Sin. 18 (3): 377. 1980
Type: China. Xizang: Lhasa, in clivis montibus, 3500–3700 m, 16 September 1970, Kuo 8109 (holotype HNWP; isotype NAS!).
Distribution and habitat: This species is endemic to China (Xizang) and grows on dry mountain slopes with elevations of 3500–3800 m.
Additional specimens examined: China. Xizang: Rikaze, 3800 m, 1960, G.X. Fu 1377 (PE); Lhasa, 3821 m, 17 October 2021, J.J. Deng and R.X. Zhou LCK20211017-01 (SZ).
Ligusticopsis violacea (R.H.Shan and M.L.Sheh) C.K.Liu and X.J.He, comb. nov.
Peucedanum violaceum R.H.Shan and M.L.Sheh in Act. Phytotax. Sin. 18 (3): 378. 1980
Type: China. Xizang: Mainling Xian, in locis arenosis montis Do Hsium, 2980 m, 11 August 1975, Qinghai-Xizang Exped. 751,309 (holotype PE; isotype KUN!).
Distribution and habitat: This species is endemic to China (Xizang) and occurs in sparse forests or grassy places on river banks with elevations of 2100–3500 m.
Additional specimens examined: China. Xizang; Lhoka City, Zhanang County, 3788 m, 25 August 2017, PE-Xiang Expedition PE5120 (PE); Nyingchi City, Mainling Country, 3100 m, 21 July 1972, Tibet Chinese Herbal Medicine Survey Team 3845 (PE); 3013 m, 16 September 2017, PE-Xiang Expedition PE6747 (PE); 2975 m, 19 October 2021, J.J. Deng and R.X. Zhou LCK2021101901 (SZ).

4. Materials and Methods

4.1. Plant Sample, DNA Extraction, Sequencing, and Assembly

Fresh leaves of P. nanum and P. violaceum were collected from their type localities and dried with silica gel. Voucher specimens were deposited in the herbarium of Sichuan University (Chengdu, China) (Table S1). Genomic DNA was extracted from the silica-gel-dried leaves using the modified CTAB method [57] and then fragmented into 400 bp to create a pair-end library according to the manufacturer’s protocol (Illumina, San Diego, CA, USA). Subsequently, the libraries were sequenced on the Illumina NovaSeq platform at Personalbio (Shanghai, China). The raw data yielded by Illumina sequencing were filtered with fastP v0.15.0 [58] to obtain high-quality reads with -n 10 and -q 15. The plastomes were then assembled based on the high-quality reads using NOVOPlasty v2.6.2 [59] with the default parameters and rbcL sequence extracted from the plastome of L. rechingeriana (MZ491175) as the seed. In addition, the ITS sequences were assembled using the GetOrganelle pipeline [60] with the ITS sequence of L. rechingeriana (MZ497220) as the reference.

4.2. Plastome Annotation and Feature Analyses

The assembled plastomes were initially annotated with the web server CPGAVAS2 (http://www.herbalgenomics.org/cpgavas2, accessed on 11 September 2022) [61]. Then, the start and stop codons and intron positions were manually corrected using Geneious v9.0.2 [62]. Finally, the online program OrganellarGenomeDRAW (OGDRAW) [63] was used to display the well-annotated plastomes.
Eleven plastomes of Ligusticopsis, which we have previously reported, were downloaded from the NCBI database (Table S8). In conjunction with two newly sequenced plastomes, codon usage of the thirteen plastomes was detected using CodonW v1.4.2 (Nottingham, UK). Subsequently, the potential RNA editing sites of the protein-coding genes for the thirteen plastomes were predicted using the online program Predictive RNA Editor for Plants suite with a cutoff value of 0.8 [64]. We also detected simple sequence repeats (SSRs) in the thirteen plastomes using MISA (http://pgrc.ipk-gatersleben.de/misa/, accessed on 11 September 2022). The minimum number of repeat units for mono-, di-, tri-, tetra-, penta-, and hexa-nucleotides was set to 10, 5, 4, 3, 3, and 3, respectively.

4.3. Comparative Plastome Analyses

The borders of the inverted repeat regions for the thirteen plastomes were compared in Geneious v9.0.2 [62]. Then, the gene order and sequence identity among the thirteen plastomes were investigated by using Mauve Alignment [65] implemented in Geneious v9.0.2 [62] and the mVISTA tool [66], respectively. To identify the mutation hotspot regions, the protein-coding genes, non-coding regions, and intron regions of the thirteen plastomes were extracted in Geneious v9.0.2 [62] and aligned with MAFFT v7.221 [67]. Alignments with more than 200 bp in length were used to calculate nucleotide diversity (Pi) using DnaSP v5.0 [68].

4.4. Phylogenetic Analyses

To resolve the phylogenetic positions of P. nanum and P. violaceum, 48 plastomes and 48 ITS sequences were used to reconstruct the phylogenetic tree (Tables S1 and S8). Among them, Chamaesium mallaeanum Farille and S.B.Malla and Chamaesium viridiflorum (Franch.) H.Wolff ex R.H.Shan were chosen as the outgroup based on a previous study [39]. The two datasets were aligned using MAFFT v7.221 [67]. Alignments were used for the maximum-likelihood analyses (ML) and Bayesian inference (BI). For the ML analyses, RAxML v8.2.8 [69] was used to reconstruct the phylogenetic tree with 1000 replicates and the GTRGAMMA model as suggested by the RAxML manual. The BI analyses were performed using MrBayes v3.2.7 [70]. The best-fit substitution models for plastome data (TVM+I+G) and ITS data (SYM+I+G) were tested using Modeltest v3.7 [71]. Two independent Markov chains were run for 1,000,000 generations with sampling every 100 generations and discarding the first 25% of the trees as burn-in.

5. Conclusions

The whole plastomes of P. nanum and P. violaceum were reported for the first time in the present study. The plastome comparisons among P. nanum, P. violaceum, and eleven Ligusticopsis species revealed that these plastomes were highly conserved and similar in terms of structure, size, GC content, gene content and order, IR borders, and the patterns of codon bias, RNA editing, and SSR. Nevertheless, 12 mutation hotspot regions (matK, ndhC, rps15, rps8, ycf2, ccsA-ndhD, petN-psbM, psbA-trnK, rps2-rpoC2, rps4-trnT, trnH-psbA, ycf2-trnL) were identified, which could serve as potential DNA markers for species identification and phylogenetic analysis of Ligusticopsis. Moreover, the phylogenetic analyses based on plastome data and ITS sequences robustly supported that P. nanum and P. violaceum nested in the genus Ligusticopsis. Considering also the morphological affinities, we transferred P. nanum and P. violaceum into Ligusticopsis and proposed two new combinations.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24010382/s1.

Author Contributions

Methodology, conceptualization, formal analysis, and original draft preparation, C.L.; software and validation, J.D.; data curation, investigation, and resources, R.Z. and B.S.; funding acquisition and project administration, S.Z.; funding acquisition, manuscript review and editing, visualization, and supervision, X.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (grant numbers 32170209 and 32070221) and the Survey on the Background Resources of Chengdu Area of Giant Panda National Park (grant number 510101202200376).

Data Availability Statement

Plastomes and ITS sequences of P. nanum and P. violaceum generated in the current study are available at the NCBI database (https://www.ncbi.nlm.nih.gov, accessed on 19 December 2022).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Leute, G.H. Untersuchungen über den Verwandtschaftskreis der Gattung Ligusticum L. (Umbelliferae)—I Teil. Ann. Naturhist. Mus. Wien 1969, 73, 55–98. [Google Scholar]
  2. Pu, F.D. A revision of the genus Ligusticum L. (Umbelliferae) in China. Acta Phytotax. Sin. 1991, 29, 385–393. [Google Scholar]
  3. Leute, G.H. Untersuchungen über den Verwandtschaftskreis der Gattung Ligusticum L. (Umbelliferae)—II Teil. Ann. Naturhist. Mus. Wien 1970, 74, 457–519. [Google Scholar]
  4. Zhang, H.C. Ligusticum L. In Flora Reipublicae Popularis Sinicae; Shan, R.H., She, M.L., Eds.; Science Press: Beijing, China, 1985; Volume 55, pp. 234–257. [Google Scholar]
  5. Pu, F.D.; Watson, M.F. Ligusticum L. In Flora of China; Wu, Z.Y., Raven, P.H., Hong, D.Y., Eds.; Science Press: Beijing, China, 2005; Volume 14, pp. 140–150. [Google Scholar]
  6. Sun, N.; He, X.J.; Zhou, S.D. Morphological cladistic analysis of Ligusticum (Umbelliferae) in China. Nord. J. Bot. 2008, 26, 118–128. [Google Scholar] [CrossRef]
  7. Li, Z.X.; Guo, X.L.; Price, M.; Zhou, S.D.; He, X.J. Phylogenetic position of Ligusticopsis (Apiaceae, Apioideae): Evidence from molecular data and carpological characters. AoB Plants 2022, 14, plac008. [Google Scholar] [CrossRef] [PubMed]
  8. Ren, T.; Xie, D.; Peng, C.; Gui, L.; Price, M.; Zhou, S.D.; He, X.J. Molecular evolution and phylogenetic relationships of Ligusticum (Apiaceae) inferred from the whole plastome sequences. BMC Eco. Evol. 2022, 22, 55. [Google Scholar] [CrossRef] [PubMed]
  9. Sheh, M.L. Peucedanum L. In Flora Reipublicae Popularis Sinica; Shan, R.H., Sheh, M.L., Eds.; Science Press: Beijing, China, 1992; Volume 55, pp. 123–175. [Google Scholar]
  10. Sheh, M.L.; Watson, M.F. Peucedanum L. In Flora of China; Wu, Z.Y., Raven, P.H., Hong, D.Y., Eds.; Science Press: Beijing, China, 2005; Volume 14, pp. 182–192. [Google Scholar]
  11. Shan, R.H.; Sheh, M.L.; Yuan, C.Q.; Wang, T.S. New taxa of the Umbelliferae from Xizang (Tibet). Acta Phytotax. Sin. 1980, 18, 374–379. [Google Scholar]
  12. Kadereit, J.W.; Bittrich, V. Flowering plants. Eudicots: Apiales, Gentianales (except Rubiaceae). In The Families and Genera of Vascular Plants; Kubitzki, K., Ed.; Springer: Berlin/Heidelberg, Germany, 2018; Volume 15, p. 168. [Google Scholar]
  13. Downie, S.R.; Watson, M.F.; Spalik, K.; Katz-Downie, D.S. Molecular systematics of Old World Apioideae (Apiaceae): Relationships among some members of tribe Peucedaneae sensu lato, the placement of several island-endemic species, and resolution within the apioid superclade. Can. J. Bot. 2000, 78, 506–528. [Google Scholar]
  14. Spalik, K.; Reduron, J.P.; Downie, S.R. The phylogenetic position of Peucedanum sensu lato and allied genera and their placement in tribe Selineae (Apiaceae, subfamily Apioideae). Plant Syst. Evol. 2004, 243, 189–210. [Google Scholar] [CrossRef]
  15. Valiejo-Roman, C.M.; Terentieva, E.I.; Samigullin, T.H.; Pimenov, M.G.; Ghahremani-Nejad, F.; Mozaffarian, V. Molecular data (nrITS-sequencing) reveal relationships among Iranian endemic taxa of the Umbelliferae. Feddes Repert. 2006, 117, 367–388. [Google Scholar] [CrossRef]
  16. Feng, T.; Downie, S.R.; Yu, Y.; Zhang, X.M.; Chen, W.W.; He, X.J.; Liu, S. Molecular systematics of Angelica and allied genera (Apiaceae) from the Hengduan Mountains of China based on nrDNA ITS sequences: Phylogenetic affinities and biogeographic implications. J. Plant Res. 2009, 122, 403–414. [Google Scholar] [CrossRef] [PubMed]
  17. Zhou, J.; Gong, X.; Downie, S.R.; Peng, H. Towards a more robust molecular phylogeny of Chinese Apiaceae subfamily Apioideae: Additional evidence from nrDNA ITS and cpDNA intron (rpl16 and rps16) sequences. Mol. Phylogenet. Evol. 2009, 53, 56–68. [Google Scholar] [CrossRef] [PubMed]
  18. Zhou, J.; Gao, Y.Z.; Wei, J.; Liu, Z.W.; Downie, S.R. Molecular phylogenetics of Ligusticum (Apiaceae) based on nrDNA ITS sequences: Rampant polyphyly, placement of the Chinese endemic species, and a much-reduced circumscription of the genus. Int. J. Plant Sci. 2020, 181, 306–323. [Google Scholar] [CrossRef]
  19. Liu, C.K.; Lei, J.Q.; Jiang, Q.P.; Zhou, S.D.; He, X.J. The complete plastomes of seven Peucedanum plants: Comparative and phylogenetic analyses for the Peucedanum genus. BMC Plant Biol. 2022, 22, 101. [Google Scholar] [CrossRef]
  20. Shaw, J.; Shafer, H.L.; Leonard, O.R.; Kovach, M.J.; Schorr, M.; Morris, A.B. Chloroplast DNA sequence utility for the lowest phylogenetic and phylogeographic inferences in angiosperms, the tortoise and the hare IV. Am. J. Bot. 2014, 101, 1987–2004. [Google Scholar] [CrossRef] [PubMed]
  21. Huang, Y.; Li, X.; Yang, Z.; Yang, C.; Yang, J.; Ji, Y. Analysis of complete chloroplast genome sequences improves phylogenetic resolution in Paris (Melanthiaceae). Front. Plant Sci. 2016, 7, 1797. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Tonti-Filippini, J.; Nevill, P.G.; Dixon, K.; Small, I. What can we do with 1000 plastid genomes? Plant J. 2017, 90, 808–818. [Google Scholar] [CrossRef] [Green Version]
  23. Ji, Y.; Yang, L.; Chase, M.W.; Liu, C.; Yang, Z.; Yang, J.; Yang, J.B.; Yi, T.S. Plastome phylogenomics, biogeography, and clade diversification of Paris (Melanthiaceae). BMC Plant Biol. 2019, 19, 543. [Google Scholar] [CrossRef] [Green Version]
  24. Dong, W.; Liu, Y.; Xu, C.; Gao, Y.; Yuan, Q.; Suo, Z.; Zhang, Z.; Sun, J. Chloroplast phylogenomic insights into the evolution of Distylium (Hamamelidaceae). BMC Genom. 2021, 22, 293. [Google Scholar] [CrossRef]
  25. Guo, X.; Liu, C.; Wang, H.; Zhang, G.; Yan, H.; Jin, L.; Su, W.; Ji, Y. The complete plastomes of two flowering epiparasites (Phacellaria glomerata and P. compressa): Gene content, organization, and plastome degradation. Genomics 2021, 113, 447–455. [Google Scholar] [CrossRef]
  26. Odago, W.O.; Waswa, E.N.; Nanjala, C.; Mutinda, E.S.; Wanga, V.O.; Mkala, E.M.; Oulo, M.A.; Wang, Y.; Zhang, C.F.; Hu, G.W.; et al. Analysis of the complete plastomes of 31 Species of Hoya group: Insights into their comparative genomics and phylogenetic relationships. Front. Plant Sci. 2021, 12, 814833. [Google Scholar] [CrossRef] [PubMed]
  27. Schneider, J.V.; Paule, J.; Jungcurt, T.; Cardoso, D.; Amorim, A.M.; Berberich, T.; Zizka, G. Resolving recalcitrant clades in the pantropical Ochnaceae: Insights from comparative phylogenomics of plastome and nuclear genomic data derived from targeted sequencing. Front. Plant Sci. 2021, 12, 638650. [Google Scholar] [CrossRef] [PubMed]
  28. Sielemann, K.; Pucker, B.; Schmidt, N.; Viehöver, P.; Weisshaar, B.; Heitkam, T.; Holtgräwe, D. Complete pan-plastome sequences enable high resolution phylogenetic classification of sugar beet and closely related crop wild relatives. BMC Genom. 2022, 23, 113. [Google Scholar] [CrossRef] [PubMed]
  29. Yu, J.; Fu, J.; Fang, Y.; Xiang, J.; Dong, H. Complete chloroplast genomes of Rubus species (Rosaceae) and comparative analysis within the genus. BMC Genom. 2022, 23, 32. [Google Scholar] [CrossRef]
  30. Liu, C.; Yang, Z.; Yang, L.; Yang, J.; Ji, Y. The complete plastome of Panax stipuleanatus: Comparative and phylogenetic analyses of the genus Panax (Araliaceae). Plant Divers. 2018, 40, 265–276. [Google Scholar] [CrossRef]
  31. Yang, L.; Yang, Z.; Liu, C.; He, Z.; Zhang, Z.; Yang, J.; Liu, H.; Yang, J.; Ji, Y. Chloroplast phylogenomic analysis provides insights into the evolution of the largest eukaryotic genome holder, Paris japonica (Melanthiaceae). BMC Plant Biol. 2019, 19, 293. [Google Scholar] [CrossRef] [Green Version]
  32. Liu, C.; Yang, J.; Jin, L.; Wang, S.; Yang, Z.; Ji, Y. Plastome phylogenomics of the East Asian endemic genus Dobinea. Plant Divers. 2021, 43, 35–42. [Google Scholar] [CrossRef]
  33. Asaf, S.; Ahmad, W.; Al-Harrasi, A.; Khan, A.L. Uncovering the first complete plastome genomics, comparative analyses, and phylogenetic dispositions of endemic medicinal plant Ziziphus hajarensis (Rhamnaceae). BMC Genom. 2022, 23, 83. [Google Scholar] [CrossRef]
  34. Cvetković, T.; Hinsinger, D.D.; Thomas, D.C.; Wieringa, J.J.; Velautham, E.; Strijk, J.S. Phylogenomics and a revised tribal classification of subfamily Dipterocarpoideae (Dipterocarpaceae). Taxon 2022, 71, 85–102. [Google Scholar] [CrossRef]
  35. Xia, M.Z.; Li, Y.; Zhang, F.Q.; Yu, J.Y.; Khan, G.; Chi, X.F.; Xu, H.; Chen, S.L. Reassessment of the phylogeny and systematics of Chinese Parnassia (Celastraceae): A thorough investigation using whole plastomes and nuclear ribosomal DNA. Front. Plant Sci. 2022, 13, 855944. [Google Scholar] [CrossRef]
  36. Guo, X.L.; Zheng, H.Y.; Price, M.; Zhou, S.D.; He, X.J. Phylogeny and comparative analysis of Chinese Chamaesium species revealed by the complete plastid genome. Plants 2020, 9, 965. [Google Scholar] [CrossRef] [PubMed]
  37. Gou, W.; Jia, S.B.; Price, M.; Guo, X.L.; Zhou, S.D.; He, X.J. Complete plastid genome sequencing of eight species from Hansenia, Haplosphaera and Sinodielsia (Apiaceae): Comparative analyses and phylogenetic implications. Plants 2020, 9, 1523. [Google Scholar] [CrossRef] [PubMed]
  38. Wang, M.; Wang, X.; Sun, J.; Wang, Y.; Ge, Y.; Dong, W.; Yuan, Q.; Huang, L. Phylogenomic and evolutionary dynamics of inverted repeats across Angelica plastomes. BMC Plant Biol. 2021, 21, 26. [Google Scholar] [CrossRef]
  39. Wen, J.; Xie, D.F.; Price, M.; Ren, T.; Deng, Y.Q.; Gui, L.J.; Guo, X.L.; He, X.J. Backbone phylogeny and evolution of Apioideae (Apiaceae): New insights from phylogenomic analyses of plastome data. Mol. Phylogenet. Evol. 2021, 161, 107183. [Google Scholar] [CrossRef] [PubMed]
  40. Jiang, Q.P.; Liu, C.K.; Xie, D.F.; Zhou, S.D.; He, X.J. Plastomes provide insights into differences between morphology and molecular phylogeny: Ostericum and Angelica (Apiaceae) as an example. Diversity 2022, 14, 776. [Google Scholar] [CrossRef]
  41. Yang, L.; Abduraimov, O.; Tojibaev, K.; Shomurodov, K.; Zhang, Y.M.; Li, W.J. Analysis of complete chloroplast genome sequences and insight into the phylogenetic relationships of Ferula L. BMC Genom. 2022, 23, 643. [Google Scholar] [CrossRef]
  42. Ren, T.; Li, Z.X.; Xie, D.F.; Gui, L.J.; Peng, C.; Wen, J.; He, X.J. Plastomes of eight Ligusticum species: Characterization, genome evolution, and phylogenetic relationships. BMC Plant Biol. 2020, 20, 519. [Google Scholar] [CrossRef] [PubMed]
  43. Hollingsworth, P.M. Refining the DNA barcode for land plants. Proc. Natl. Acad. Sci. USA 2011, 108, 19451–19452. [Google Scholar] [CrossRef] [Green Version]
  44. CBOL Plant Working Group. A DNA barcode for land plants. Proc. Natl. Acad. Sci. USA 2009, 106, 12794–12797. [Google Scholar] [CrossRef] [Green Version]
  45. Hollingsworth, P.M.; Graham, S.W.; Little, D.P. Choosing and using a plant DNA barcode. PLoS ONE 2011, 6, e19254. [Google Scholar] [CrossRef]
  46. Hollingsworth, P.M.; Li, D.Z.; van der Bank, M.; Twyford, A.D. Telling plant species apart with DNA: From barcodes to genomes. Philos. Trans. R. Soc. B 2016, 371, 20150338. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Gaskin, J.F.; Wilson, L.M. Phylogenetic relationships among native and naturalized Hieracium (Asteraceae) in Canada and the United States based on plastid DNA sequences. Syst. Bot. 2007, 32, 478–485. [Google Scholar] [CrossRef]
  48. Theis, N.; Donoghue, M.J.; Li, J. Phylogenetics of the Caprifolieae and Lonicera (Dipsacales) based on nuclear and chloroplast DNA sequences. Syst. Bot. 2008, 33, 776–783. [Google Scholar] [CrossRef]
  49. Yue, J.P.; Sun, H.; Baum, D.A.; Li, J.H.; Al-shehbaz, I.A.; Pee, R. Molecular phylogeny of Solms-Laubachia (Brassicaceae) s.l., based on multiple nuclear and plastid DNA sequences, and its biogeographic implications. J. Syst. Evol. 2009, 47, 402–415. [Google Scholar] [CrossRef]
  50. Huang, J.L.; Sun, G.L.; Zhang, D.M. Molecular evolution and phylogeny of the angiosperm ycf2 gene. J. Syst. Evol. 2010, 48, 240–248. [Google Scholar] [CrossRef]
  51. Nakaji, M.; Tanaka, N.; Sugawara, T. A molecular phylogenetic study of Lonicera L. (Caprifoliaceae) in Japan based on chloroplast DNA sequences. APG 2015, 66, 137–151. [Google Scholar]
  52. Liu, P.L.; Wen, J.; Duan, L.; Arslan, E.; Ertuğrul, K.; Chang, Z.Y. Hedysarum L. (Fabaceae: Hedysareae) is not monophyletic–evidence from phylogenetic analyses based on five nuclear and five plastid sequences. PLoS ONE 2017, 12, e0170596. [Google Scholar] [CrossRef] [Green Version]
  53. Machado, L.O.; Vieira, L.N.; Stefenon, V.M.; Faoro, H.; Pedrosa, F.O.; Guerra, M.P.; Nodari, R.O. Molecular relationships of Campomanesia xanthocarpa within Myrtaceae based on the complete plastome sequence and on the plastid ycf2 gene. Genet. Mol. Biol. 2020, 43, e20180377. [Google Scholar] [CrossRef]
  54. Wang, P.L.; Pu, F.D.; Ma, J.S. Pollen morphology of the genus Ligusticum from China and its systematic significance. Acta Phytotax. Sin. 1991, 29, 235–245. [Google Scholar]
  55. Pimenov, M.G.; Kljuykov, E.V.; Ostroumova, T.A. Himalayan species of Selinum L. s.l. (Umbelliferae). The genus Oreocome Edgew. Willdenowia 2001, 31, 101–124. [Google Scholar] [CrossRef]
  56. Pimenov, M.G. Updated checklist of Chinese Umbelliferae: Nomenclature, synonymy, typification, distribution. Turczaninowia 2017, 20, 106–239. [Google Scholar]
  57. Doyle, J.J.; Doyle, J.L. A rapid DNA isolation procedure for small quantities of fresh leaf tissue. Phytochem. Bull. 1987, 19, 11–15. [Google Scholar]
  58. Chen, S.; Zhou, Y.; Chen, Y.; Gu, J. Fastp: An ultra-fast all-in-one FASTQ preprocessor. Bioinformatics 2018, 34, i884–i890. [Google Scholar] [CrossRef] [PubMed]
  59. Dierckxsens, N.; Mardulyn, P.; Smits, G. NOVOPlasty: De novo assembly of organelle genomes from whole genome data. Nucleic Acids Res. 2017, 45, e18. [Google Scholar]
  60. Jin, J.J.; Yu, W.B.; Yang, J.B.; Song, Y.; de Pamphilis, C.W.; Yi, T.S.; Li, D.Z. GetOrganelle: A fast and versatile toolkit for accurate de novo assembly of organelle genomes. Genome Biol. 2020, 21, 241. [Google Scholar] [CrossRef]
  61. Shi, L.; Chen, H.; Jiang, M.; Wang, L.; Wu, X.; Huang, L.; Liu, C. CPGAVAS2, an integrated plastome sequence annotator and analyzer. Nucleic Acids Res. 2019, 47, W65–W73. [Google Scholar] [CrossRef]
  62. Kearse, M.; Moir, R.; Wilson, A.; Stones-Havas, S.; Cheung, M.; Sturrock, S.; Buxton, S.; Cooper, A.; Markowitz, S.; Duran, C.; et al. Geneious basic: An integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 2012, 28, 1647–1649. [Google Scholar] [CrossRef] [Green Version]
  63. Lohse, M.; Drechsel, O.; Bock, R. OrganellarGenomeDRAW (OGDRAW): A tool for the easy generation of high-quality custom graphical maps of plastid and mitochondrial genomes. Curr. Genet. 2007, 52, 267–274. [Google Scholar] [CrossRef]
  64. Mower, J.P. The PREP suite: Predictive RNA editors for plant mitochondrial genes, chloroplast genes, and user-defined alignments. Nucleic Acids Res. 2009, 37, W253–W259. [Google Scholar] [CrossRef]
  65. Darling, A.C.E.; Mau, B.; Blattner, F.R.; Perna, N.T. Mauve: Multiple alignment of conserved genomic sequence with rearrangements. Genome Res. 2004, 14, 1394–1403. [Google Scholar] [CrossRef]
  66. Frazer, K.A.; Pachter, L.; Poliakov, A.; Rubin, E.M.; Dubchak, I. VISTA: Computational tools for comparative genomics. Nucleic Acids Res. 2004, 32, W273–W279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Librado, P.; Rozas, J. DnaSP v5: A software for comprehensive analysis of DNA polymorphism data. Bioinformatics 2009, 25, 1451–1452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Stamatakis, A. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 2014, 30, 1312–1313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Ronquist, F.; Teslenko, M.; van der Mark, P.; Ayres, D.L.; Darling, A.; Höhna, S.; Larget, B.; Liu, L.; Suchard, M.A.; Huelsenbech, J.P. MrBayes 3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 2012, 61, 539–542. [Google Scholar] [CrossRef] [Green Version]
  71. Posada, D.; Crandall, K.A. Modeltest: Testing the model of DNA substitution. Bioinformatics 1998, 14, 817–818. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Morphology of P. nanum (A,B) and P. violaceum (C,D).
Figure 1. Morphology of P. nanum (A,B) and P. violaceum (C,D).
Ijms 24 00382 g001
Figure 2. Whole plastome map of Peucedanum nanum and P. violaceum. Genes shown outside of the circle were transcribed counterclockwise, while those presented inside were transcribed clockwise. The dark gray area of the inner circle shows the GC content of the plastome.
Figure 2. Whole plastome map of Peucedanum nanum and P. violaceum. Genes shown outside of the circle were transcribed counterclockwise, while those presented inside were transcribed clockwise. The dark gray area of the inner circle shows the GC content of the plastome.
Ijms 24 00382 g002
Figure 3. The RSCU values of 79 protein-coding regions for thirteen plastomes. Red represents higher RSCU values, while blue indicates lower RSCU values.
Figure 3. The RSCU values of 79 protein-coding regions for thirteen plastomes. Red represents higher RSCU values, while blue indicates lower RSCU values.
Ijms 24 00382 g003
Figure 4. The different repeat types identified in thirteen plastomes.
Figure 4. The different repeat types identified in thirteen plastomes.
Ijms 24 00382 g004
Figure 5. Comparison of IR borders among thirteen plastomes.
Figure 5. Comparison of IR borders among thirteen plastomes.
Ijms 24 00382 g005
Figure 6. Mauve alignment of thirteen plastomes.
Figure 6. Mauve alignment of thirteen plastomes.
Ijms 24 00382 g006
Figure 7. mVISTA alignment for thirteen plastomes.
Figure 7. mVISTA alignment for thirteen plastomes.
Ijms 24 00382 g007
Figure 8. The nucleotide diversity (Pi) values among thirteen plastomes. (A) protein-coding regions; (B) non-coding and intron regions.
Figure 8. The nucleotide diversity (Pi) values among thirteen plastomes. (A) protein-coding regions; (B) non-coding and intron regions.
Ijms 24 00382 g008
Figure 9. Phylogenetic tree inferred from maximum likelihood (ML) and Bayesian inference (BI) analyses based on plastome data. The numbers indicate Bayesian posterior probabilities (PP) and maximum likelihood bootstrap values (BS), respectively.
Figure 9. Phylogenetic tree inferred from maximum likelihood (ML) and Bayesian inference (BI) analyses based on plastome data. The numbers indicate Bayesian posterior probabilities (PP) and maximum likelihood bootstrap values (BS), respectively.
Ijms 24 00382 g009
Table 1. Comparison of plastome features among Peucedanum nanum, P. violaceum, and Ligusticopsis species.
Table 1. Comparison of plastome features among Peucedanum nanum, P. violaceum, and Ligusticopsis species.
TaxonTotal Length
(bp)
LSC
(bp)
SSC
(bp)
IR
(bp)
Total GC Content
(%)
Total Genes
(Unique)
Protein Coding Genes
(Unique)
rRNA Genes
(Unique)
tRNA Genes
(Unique)
L. brachyloba148,63392,26517,58819,39037.40%11379430
L. capillacea147,80891,90717,50319,19937.50%11379430
L. daucoides148,07891,66617,58219,41537.40%11379430
L. hispida147,79791,84617,62719,16237.40%11379430
L. integrifolia148,19692,30517,57519,15837.50%11379430
L. involucrata147,75291,78217,56019,20537.40%11379430
L. modesta148,13392,24717,56819,15937.50%11379430
L. oliveriana148,17592,27317,53419,18437.50%11379430
L. rechingeriana148,52591,81317,65419,52937.30%11379430
L. scapiformis148,10792,21417,58119,15637.50%11379430
L. wallichiana148,59492,28117,56719,37337.40%11379430
P. nanum146,90091,48017,30819,05637.50%11379430
P. violaceum148,19091,81116,33520,02237.50%11379430
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, C.; Deng, J.; Zhou, R.; Song, B.; Zhou, S.; He, X. Plastid Phylogenomics Provide Evidence to Accept Two New Members of Ligusticopsis (Apiaceae, Angiosperms). Int. J. Mol. Sci. 2023, 24, 382. https://doi.org/10.3390/ijms24010382

AMA Style

Liu C, Deng J, Zhou R, Song B, Zhou S, He X. Plastid Phylogenomics Provide Evidence to Accept Two New Members of Ligusticopsis (Apiaceae, Angiosperms). International Journal of Molecular Sciences. 2023; 24(1):382. https://doi.org/10.3390/ijms24010382

Chicago/Turabian Style

Liu, Changkun, Jiaojiao Deng, Renxiu Zhou, Boni Song, Songdong Zhou, and Xingjin He. 2023. "Plastid Phylogenomics Provide Evidence to Accept Two New Members of Ligusticopsis (Apiaceae, Angiosperms)" International Journal of Molecular Sciences 24, no. 1: 382. https://doi.org/10.3390/ijms24010382

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop