Next Article in Journal
A COL5A2 In-Frame Deletion in a Chihuahua with Ehlers-Danlos Syndrome
Previous Article in Journal
In Silico Analysis of Seven PCR Markers Developed from the CHD1, NIPBL and SPIN Genes Followed by Laboratory Testing Shows How to Reliably Determine the Sex of Musophagiformes Species
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Complete Chloroplast Genome Sequence of Triosteum sinuatum, Insights into Comparative Chloroplast Genomics, Divergence Time Estimation and Phylogenetic Relationships among Dipsacales

1
State Key Laboratory of Plateau Ecology and Agriculture, Qinghai University, Xining 810008, China
2
College of Eco-Environmental Engineering, Qinghai University, Xining 810008, China
3
Key Laboratory of Adaptation and Evolution of Plateau Biota, Northwest Institute of Plateau Biology, Chinese Academy of Sciences, Xining 810001, China
4
Department of Geological Engineering, Qinghai University, Xining 810016, China
5
Department of Botany, Women University of AJK, Bagh 12500, Pakistan
*
Author to whom correspondence should be addressed.
Submission received: 12 April 2022 / Revised: 6 May 2022 / Accepted: 17 May 2022 / Published: 23 May 2022
(This article belongs to the Section Plant Genetics and Genomics)

Abstract

:
Triosteum himalayanum, Triosteum pinnatifidum (Triosteum L., Caprifoliaceae, Dipsacales) are widely distributed in China while Triosteum sinuatum mainly occurrs in northeast China. Few reports have been determined on the genus Triosteum. In the present research, we sequenced 2 chloroplast genomes of Triosteum and analyzed 18 chloroplast genomes, trying to explore the sequence variations and phylogeny of genus Triosteum in the order Dipsacales. The chloroplast genomes of the genus Triosteum ranged from 154,579 bp to 157,178 bp, consisting of 132 genes (86 protein-coding genes, 38 transfer RNA genes, and 8 ribosomal RNA genes). Comparative analyses and phylogenetic analysis supported the division of Dipsacales into two clades, Adoxaceae and six other families. Among the six families, a clade of Valerianaceae+Dipsacaceae was recovered as a sister to a clade of Morinaceae+Linnaeaceae. A closer relationship of T. himalayanum and T. pinnatifidum among three species was revealed. Our research supported that Lonicera ferdinandi and Triosteum was closely related. Zabelia had a closer relationship with Linnaea borealis and Dipelta than Morinaceae. The divergence between T. sinuatum and two other species in Triosteum was dated to 13.4 mya.

1. Introduction

Chloroplasts (cp) are involved in photosynthesis and provide energy for green plants [1,2]. The double-stranded and circular chloroplast genome is composed of two short inverted repeat regions (IRa and IRb) which are separated by a large single copy (LSC) region and a small single copy (SSC) region [3,4,5]. The sequence of chloroplast genome is highly conserved and similar in the order of Dipsacales. Due to its short length, conserved structure and type of genes, together with few repeat regions, the complete chloroplast genome can provide abundant information for species phylogeny and taxonomy. Hence, the complete chloroplast genome is widely used in plant phylogeny and evolution study [6,7,8,9,10].
Triosteum Linn. is perennial medicinal herb, belonging to family Caprifoliaceae (Dipsacales). Some of the Triosteum species can be used as medicine. Previously, several study were conducted to determine the phylogeny and evolutionary lineage in Dipsacales [4,7,11,12,13,14]. The order Dipsacales is widely considered to consist of Caprifoliaceae s.l. and Adoxaceae according to the APG IV and some other previous studies [11,12,13]. Diervillaceae, Dipsacaceae, Linnaeaceae, Morinaceae, Caprifolieae and Valerianaceae belong to the large clade Caprifoliaceae s.l., comprising 41 genera (including Acanthocalyx, Dipsacus, Patrinia, Linnaea, Weigela, Triosteum, Morina, Lonicera, Triplostegia, etc.). Additionally, Adoxaceae includes five genera (Adoxa, Sambucus, Sinadoxa, Tetradoxa and Viburnum).
However, the phylogentic evolution of Triosteum within Dipsacales was poorly invested. In this study, we focused on three species of the genus Triosteum L., including T. himalayanum, T. pinnatifidum, T. sinuatum. These perennial herbs are widely distributed in China as well in Japan and Russia (T. pinnatifidum, T. sinuatum) [15,16]. Studies of chemical constituents have been conducted on T. himalayanum and T. pinnatifidum. Additionally, few evolutionary relationships between Triosteum and other genera have been revealed [17,18,19,20]. Gould and Donoghue (2000) used sequences of ITS and GBSSI to reveal the biogeographic pattern of T. himalayanum and T. pinnatifidum in China and T. sinuatum in Japan. According to the results of the previous studies, Triosteum was close to Lonicera in order Dipsacales. Based on the phylogeographic structure of T. himalayanum localities, the phylogeny of haplotypes and palaeodistributional reconstruction, both westwards and northwards expansion of this species was postulated [15]. For T. pinnatifidum, the genetic structure, unimodal mismatch distribution, star-like network, together with demographic history analysis supported the “dispersal into the Qinghai-Tibetan Plateau” hypothesis [21].
Here, the whole chloroplast (cp) genome sequences of T. pinnatifidum, T. sinuatum and Linnaea borealis were assembled and annotated. The cp genome of T. sinuatum was utilized for the first time for the phylogenetic relationship and tracing of evolutionary lineage in Triosteum. A total of 20 chloroplast genome sequences of Dipsacales were generated, and six sequences from Apiales and Lamiales were used as outgroup taxa. Seventeen cp genome sequences in Dipsacales were downloaded from NCBI database. The sequence variations and detailed interrelationship of 20 species in the order Dipsacales were explored. The present study was carried out to identify the variation of hotspot regions in chloroplast genomes, analyze the simple sequence repeat (SSR) diversity, and calculate relative synonymous codon usage (RSCU) frequency. The phylogenetic relationship between Triosteum and other lineages in the order Dipsacales was reconstructed, and divergence time was estimated based on cp genome.

2. Materials and Methods

2.1. Sampling, DNA Isolation, Sequencing and Assembly

Fresh leaves of T. pinnatifidum were collected in Tibet. Leaves of Linnaea borealis and T. sinuatum were collected in Jilin Province. All the specimens were tagged, dried, poisoned and pasted on standard herbarium sheets. The voucher specimens are deposited in the Herbarium of Northwest Institute of Plateau Biology (HNWP), Xining, Qinghai, China for future reference. The genomic DNA of T. pinnatifidum was extracted from fresh leaves using the cetyltrimethyl ammonium bromide (CTAB) method (Doyle, 1987) with some modifications [22]. The Illumina sequencing of T. sinuatum and L. borealis was carried out by the Genepioneer Biotechnologies (Nanjing, China). Assembly was performed with SPAdes v3.10.1, SPACE v2.0 and Gapfiller v2.1.1 [23]. We rearrange the corrected pseudo genome in order and obtained a complete circular sequence. In addition, 23 complete sequences of chloroplast genomes were selected and downloaded from GeneBank after sequences check, comprising 17 of Dipsacales, 3 of Apiales and 3 of Lamiales (Table 1).

2.2. Genome Annotation

We annotated the genome sequence using GeSeq (https://chlorobox.mpimp-golm.mpg.de/geseq.html (accessed on 10 January 2021)) and used the available sequences downloaded from GeneBank as a reference [24]. The manual correction of annotation was performed with Sequin v.15.50 and ApE [25]. The complete chloroplast genome sequences were submitted to GeneBank. Circular plastid genome maps were obtained from OGDRAW (https://chlorobox.mpimp-golm.mpg.de/OGDraw.html (accessed on 12 January 2021)) [26].

2.3. Genome Structure and Comparative Analysis

Structure and gene contents among the plastomes were manually compared with Notepad++ v7.5.9. Alignments of 20 complete chloroplast genome sequences were visually compared using the Shuffle-LAGAN Model of mVISTA [27], which uses T. sinuatum as a reference. Boundaries between IR and SC regions of these species were also visualized with IRscope (https://irscope.shinyapps.io/irapp/ (accessed on 14 April 2021)) [28].

2.4. Distribution of Simple Sequence Repeats (SSRs)

Chloroplast simple sequence repeat (cpSSR) was widely used to study the phylogeny of plants and genetic structure for its high variability. MISA_web (https://webblast.ipk-gatersleben.de/misa/ (accessed on 24 January 2021)) was used to detect the SSRs in T. sinuatum with the following parameters: 10 repeat units for mononucleotide SSRs, 5 repeat units for dinucleotide SSRs, 4 repeat units each for trinucleotide, and 3 repeat units each for tetra-, penta-, and hexanucleotide SSRs [29].

2.5. Codon Usage Pattern

We used the CodonW software (John Peden, http://www.molbiol.ox.ac.uk/cu (accessed on 24 May 2021), version 1.4.2) to calculate the number of amino acid codon in chloroplast genomes of these twenty species [30]. The FASTA files containing of 20 chloroplast genome sequences were downloaded from NCBI before RSCU analysis. The results were recorded by column graph with R language after running program.

2.6. Phylogenetic Analysis

To reveal the phylogenetic position of the genus Triosteum, a total of 26 species were selected. Out of 26 species, the ingroup contains 20 species that belong to the order Dipsacales. Clustalw v2.1 was used for genome alignment of all species. Iqtree v1.6.12 was used to find the best substitution model and constructed maximum likelihood (ML) tree with the use of GTR+F+G4 model [31,32]. The visualization of the ML tree was completed with Figtree v1.4.4 (http://tree.bio.ed.ac.uk/software/fgtree (accessed on 23 March 2022)).

2.7. Divergence Time Estimation

BEAST v1.8.0 was used to estimate the divergence time of the genus Triosteum, the and GTP+G+I model was chosen as a substitution model [33]. The analysis was calculated on the basis of two fossils of the genus Dipelta and the genus Viburnum. Tree prior was determined as the Yule process and the clock model was a lognormal relaxed clock model. The three calibration points were the genus Dipelta with lognormal prior (mean = 36 Ma, SD = 1.0), the genus Viburnum with lognormal prior (mean = 89.3 Ma, SD = 1.0) and the order Dipsacales with normal prior (mean = 103 Ma, SD = 1.0). A total of 10,000,000 generations were run, and the Tracer was used to determine the correctness. Finally, with 25% burn-in, a maximum clade credibility tree was conducted using TreeAnnotator [33].

3. Results and Discussion

3.1. Plastome Features and Gene Content

Genome maps of T. sinuatum in Dipsacales were drawn with ODDRAW (Figure 1). The chloroplast genome showed a typical double-stranded circular structure, length of T. himalayanum, T. pinnatifidum and T. sinuatum was 154,579 bp, 154,896 bp and 157,178 bp, respectively. Each chloroplast genome was composed of a large single copy region (89,007–90,758 bp) and a small single copy region (18,656–120,543 bp) separated by two inverted repeat regions (22,673–23,882 bp). Protein-coding genes orf42, orf188 and tRNA gene trnP-GGG were lost in T. pinnatifidum. In particular, trnP-UGG could not be found in T. himalayanum.
Twenty complete chloroplast genomes were different in lengths, ranging from 151,267 bp (Patrinia scabra) to 161,576 bp (Linnaea borealis) (Table 1). Except for Linnaea borealis, the lengths of the other 19 sequences were less than 160 kb, which mainly reflected in the LSC and IR regions. The LSC length in Dipsacales ranged from 86,340 bp (Adoxa moschatellina) to 90,750 bp (Acanthocalyx alba), and the IR region ranged from 22,673 bp (T. pinnatifidum) to 29,210 bp (Linnaea borealis). GC contents of the complete chloroplast genomes are similar, which are approximately 38%. The structure characteristics are similar to the chloroplast genomes of most angiosperms [34,35,36]. Most of plastomes encoded about 132 genes, including 86 protein coding genes, 38 tRNA genes, and 8 rRNA genes (Table 2). lhbA only existed as a protein coding gene in Adoxaceae.
Nine out of 86 protein coding genes were for large subunits of ribosome (rpl33, rpl20, rpl36, rpl14, rpl16, rpl22, rpl2, rpl23, rpl32), 12 were for small subunits of ribosome (rps16, rps2, rps14, rps4, rps18, rps12, rps11, rps8, rps3, rps19, rps7, rps15), 4 were for RNA polymerase (rpoC2, rpoC1, rpoB, rpoA), 20 were for photosystem (psaB, psaA, psaI, psaJ, psaC, psbA, psbK, psbI, psbM, psbD, psbC, psbZ, psbJ, psbL, psbF, psbE, psbB, psbT, psbN, psbH) and 6 were for ATP synthase (atpA, atpFa, atpH, atpI, atpE, atpB). Among all these genes, 13 genes contained one intron, including 9 protein coding genes (atpF, ndhA, ndhB, petB, petD, rpl16, rpl2, rpoC1, rps16) and 4 tRNA genes, while one gene (rps12) contained two introns (Table 2). Fifteen genes (all four rRNAs genes, seven tRNAs genes, and four protein coding genes) were duplicated in the IR regions (Table 2).

3.2. Comparative Analyses

The sequence of T. sinuatum was used as a reference to assess similarity and differences among 20 species of Dipsacales. The alignment of sequences was visualized with an mVISTA plot (Figure S1). Alignments of 20 species showed that the whole chloroplast genome is highly conserved. The sequences in the single copy regions were more divergent than those in inverted repeat regions. In addition, more variation was found in non-coding regions than in coding regions. The results of the present study are in accordance with previous research by Cheng et al., 2020, Alzahrani et al., 2021 and Wang et al., 2021. [34,37,38]. The sequences of T. sinuatum, T. himalayanum and T. pinnatifidum showed high similarity but exhibited marked differences in the 133k region and the region of accD. There was greater differentiation where five plastid genomes lost the gene ndhF (Adoxa moschatellina, Sinadoxa corydalifolia, Sambucus williamsii, Viburnum odoratissimum and Viburnum dilatatum). This may considered be as evidence that the genera Sambucus and Viburnum should be placed in Adoxaceae from Caprifoliaceae.
The sizes of chloroplast genomes in Dipsacales species are similar but different for the main factors, the expansion and contraction of inverted repeat regions [39,40,41]. The border structure of 19 plastid genomes in the order Dipsacales was compared to analyse how expansion and contraction in IR regions affect the length of chloroplast genome. Detailed comparisons of LSC, SSC and inverted repeat regions are shown in Figure 2. The rpl23 gene of 12 species was located in IRb, extending into the LSC region by about 43–179 bp. By contrast, for Weigela florida and Triplostegia glandulifera, the rpl23 gene was found in the LSC region. Additionally, the IR/SC boundary in three plastomes (Sambucus williamsii, Viburnum dilatatum and Viburnum odoratissimum) was located on rps19. The trnN gene of 18 species excluding T. pinnatifidum and Weigela florida resided in the IRb and IRa region. In addition, the ndhF gene was positioned within the SSC region similarly. The nad5 gene appeared only on the SSC region of Sambucus williamsii and Viburnum odoratissimum, which was 36 bp and 95 bp from the boundary between the IRb/SSC. On the other hand, ycf2 was mainly found in the IRa/LSC region, while the trnH gene was inserted into the fragment distributed in the LSC region. The ycf1 genes of 14 species are mainly located in the SSC region. However, in Linnaea borealis and the family Adoxaceae (except Adoxa moschatellina), the ycf1 gene expanded to IRa region for 5033 bp to 5730 bp.

3.3. SSR Analysis of Triosteum sinuatum Maxim. Chloroplast Genomes

Chloroplast simple sequence repeats (SSRs) are crucial in phylogenetic analyses and are considered to be useful in identification of species [42,43,44]. Numerous SSRs were detected in 26 chloroplast genome sequences and their amounts ranged from 36 to 77. In total, 76 SSRs of six types (mononucleotide, dinucleotide, trinucleotide, tetranucleotide, pentanucleotide and hexanucleotide repeats) were found in T. sinuatum. The most plentiful type of SSRs was mononucleotide repeats with A/T types (63.2%), followed by dinucleotide (10.5%), tetranucleotide (10.5%), trinucleotide (7.9%), pentanucleotide (5.3%), hexanucleotide (2.6%) repeats (Figure S2a). Mononucleotide repeats, which were largest in number, consisted of A (41.7%) and T (58.3%) motifs, while dinucleotide repeats, which were second largest in number, consisted of AT (37.5%) and TA (62.5%) motifs (Figure S2b). Our results match with many previous studies [39,45,46]. Further analysis showed that the most SSRs were detected in the LSC region (77.6%), with a small portion located in the SSC (14.5%) and IR (7.9%) regions (Figure S2b). In addition, 80.2%, 13.2%, and 6.6% were, respectively, distributed in the intergenic spacer (IGS), protein coding genes (CDS) and introns (Figure S2d). SSRs detected in our study will contribute to providing data for molecular marker development in future evolutionary studies.

3.4. Codon Usage Pattern

Codon usage preference is a common phenomenon in nature, mainly determined by the dynamic balance of gene mutation and natural selection [47]. In addition, it is also related to gene coding structure and function, gene expression and other factors in the evolution process [37]. Natural selection usually causes organisms to prefer to use optimal codons, and mutation will lead to the existence of some non-optional codons. Different species or different genes of the same species have different preferences for codon use after long-term evolution [48,49]. The codon W was used to calculate the relative synonymous codon usage (RSCU) of 20 species. Additionally, the frequency of used codons for amino acids is presented in Figure S3. A codon with an RSCU value more than 1 (RSCU > 1) was considered as a preferred codon because its investigated usage frequency is more than the expected one [50]. AGA is the most frequently used codon (RSCU > 1) in the chloroplast genomes among 20 species. Meanwhile, the lowest frequency codon is CGC (RSCU < 0.51), which encodes the same amino acid (Arginine) as AGA. There is difference among 20 species, but high conservatism was showed in codon usage. In addition, the analysis of the preference of target gene and receptor genomic codon by bioinformatics is an important supplement to phylogenetic analysis.

3.5. Phylogenetic Relationships

The phylogenetic relationship between Triosteum and other genera in the order Dipsacales was constructed on the base of cp genomes of 20 ingroup and 6 outgroup species (Apiales and Lamiales) using Maximum likelihood method (Figure 3). The current study presents highly resolved phylogenies of Dipsacales based on complete plastome sampling of all families in Dipsacales. The present study strongly supports the previous research which also divided 20 Dipsacales species into two caldes [4,7,51,52,53]. One clade is Adoxaceae, and the other clade consisted of six major lineages with a highly supported topology of (Diervillaceae, (Caprifoliaceae, ((Valerianaceae, Dipsacaceae), (Morinaceae, Linnaeaceae)))). Moreover, our results indicate that the family Dipsacaceae (Triplostegia glandulifera, Dipsacus japonicus) is a sister to the family Valerianaceae. The family Morinaceae was placed as a sister to the family Linnaeaceae. However, in our study, Zabelia dielsii showed close relationships with Linnaea and Dipelta with high support rather than with Morinaceae, which is in conflict with a previous study [4,51]. According to the ML tree, three species of Triosteum clustered in one clade with a 100% bootstrap value. Among the three species of Triosteum, T. pinnatifidum and T. himalayanum are supposed to have a closer relationship. The topology result was consistent with the studies of Donoghue et al. (2003) and Liu et al. (2019) [11,54]. Here, complete chloroplast genome provided a stronger support. Lonicera ferdinandi and the three species in the genus Triosteum clustered into a clade, suggesting a very close phylogenetic relationship between the genus Lonicera and the genus Triosteum. Phylogenetic analysis in our study served as a baseline for further phylogeny studies on Triosteum and Dipsacales.

3.6. Divergence Time Estimation

The result of divergence time estimation of 26 species based on chloroplast sequences is shown in Figure 4. The divergence times estimated in our study are more recent than those in previous study by Fan et al. (2018) and Wang et al. (2019) [4,7]. The age of the crown node of Dipsacales was estimated to be upper Cretaceous, approximately 95.6 million years ago (mya). Morinaceae, Dipsacaceae, Valerianaceae, Linnaeaceae and Caprifoliaceae separated from Diervillaceae at about 51.6 mya. The separation between Caprifoliaceae and the clade Morinaceae+Dipsacaceae+Valerianaceae+Linnaeaceae occurred at 51.1 mya. The divergence between Morinaceae and Linnaeaceae was dated to 41.9 mya in the Eocene, which was later than that of the previous study by Wang et al. (2019, in the upper Cretaceous) [4]. These differences between current and previous studies are likely due to the influence of the amount of data and species sampling. The speciation of most of the species investigated here occurred after the beginning of Miocene. Triosteum and Lonicera separated at the time 20.1 mya. The divergence between T. sinuatum and other two Triosteum species was dated to 13.4 mya. The divergence between T. himalayanum and T. pinnatifidum was dated back to 7.6 mya, suggesting that these two species experienced divergence for a long time.

4. Conclusions

In this study, we reported the complete chloroplast genome of Triosteum of the order Dipsacales based on the Illumina platform. The full length of three species in Triosteum ranged from 154,513 to 157,178 bp. For 20 species in the order Dipsacales, our study investigated the variation of sequences and SSRs, compared genome structure, evaluated Codon usage bias, reconstructed phylogenetic relationships and estimated divergence times. A high similarity of sequences was shown within 20 Dipsacales species, but sequences were significantly different in 133k region and region of accD. Within Dipsacales, two major clades were clearly defined: Adoxaceae and another six families with high support. T. pinnatifidum and T. himalayanum are supposed to have a closer relationship than T. sinuatum in the genus. The divergence between T. sinuatum and other two Triosteum species was dated to 13.4 mya. The diversification of Dipsacales occurred at about 95.6 mya in the Cretaceous. The present research is the first documented report in China that assembled the choloplast sequence of T. sinuatum, contributing to further study. We provided stronger support for future studies on molecular identification and the better understanding of phylogeny in Triosteum and Dipsacales.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/genes13050933/s1, Figure S1: The chloroplast genome of 26 species were compared using the mVISTA program; Figure S2: Analysis of simple sequence repeat (SSR) in twenty-six cp genomes; Figure S3: Codon content of 20 amino acid and stop codons in all protein-coding genes of the twenty cp genomes.

Author Contributions

Conceptualization, H.L. and Q.G.; software, W.L. and X.L.; validation, S.C.; formal analysis, B.X.; investigation, H.L. and G.Z.; resources, H.L.; data curation, Q.G.; writing—original draft preparation, H.L. and J.F.; writing—review and editing, H.L., I.A. and D.Z.; visualization, Q.X.; supervision, Q.G.; project administration, H.L.; funding acquisition, H.L. All authors have read and agreed to the published version of the manuscript.

Funding

Natural Science Foundation of Qinghai Province, China (No. 2019-ZJ-951Q).

Data Availability Statement

The chloroplast genome sequences of Triosteum sinuatum were submitted to GenBank (MW526077).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dong, F.; Lin, Z.; Lin, J.; Ming, R.; Zhang, W. Chloroplast Genome of Rambutan and Comparative Analyses in Sapindaceae. Plants 2021, 10, 283. [Google Scholar] [CrossRef]
  2. Zha, X.; Wang, X.; Li, J.; Gao, F.; Zhou, Y. Complete chloroplast genome of Sophora alopecuroides (Papilionoideae): Molecular structures, comparative genome analysis and phylogenetic analysis. J. Genet. 2020, 99, 13. [Google Scholar] [CrossRef]
  3. Bai, G.Q.; Zhou, T.; Zhao, J.X.; Li, W.M.; Han, G.J.; Li, S.F. The complete chloroplast genome of Kolkwitzia amabilis (Caprifoliaceae), an endangered horticultural plant in China. Mitochondrial DNA 2017, 28, 296–297. [Google Scholar] [CrossRef]
  4. Wang, H.X.; Liu, H.; Moore, M.J.; Landrein, S.; Liu, B.; Zhu, Z.X.; Wang, H.F. Plastid phylogenomic insights into the evolution of the Caprifoliaceae s.l. (Dipsacales). Mol. Phylogenet. Evol. 2019, 142, 106641. [Google Scholar] [CrossRef] [PubMed]
  5. Wei, L.; Zhang, C.; Guo, X.; Liu, Q.; Wang, K. Complete chloroplast genome of Camellia japonica genome structures, comparative and phylogenetic analysis. PLoS ONE 2019, 14, e0216645. [Google Scholar] [CrossRef] [Green Version]
  6. Duan, Y.; Zhang, K. Comparative analysis and phylogenetic evolution of the complete chloroplast genome of Ammopiptanthus. Acta Bot. Boreal.-Occident. Sin. 2020, 40, 1323–1332. [Google Scholar] [CrossRef]
  7. Fan, W.B.; Wu, Y.; Yang, J.; Shahzad, K.; Li, Z.H. Comparative Chloroplast Genomics of Dipsacales Species: Insights into Sequence Variation, Adaptive Evolution, and Phylogenetic Relationships. Front. Plant Sci. 2018, 9, 689. [Google Scholar] [CrossRef] [Green Version]
  8. Wang, Y.; Gong, X.; Hu, C.; Hao, G. Phylogeography of an alpine species Primula secundiflora inferred from the chloroplast DNA sequence variation. J. Syst. Evol. 2008, 46, 13–22. [Google Scholar] [CrossRef]
  9. Zhang, L.; Cheng, Y.; Jiang, Z.; Liu, L.; Cai, J. Structure and Phylogeny of Chloroplast Genomes and Spermatophyte Flora in Chinese Theaceae. J. Northwest For. Univ. 2020, 35, 47–53. [Google Scholar] [CrossRef]
  10. Zhao, Y.; Yang, Z.; Zhao, Y.; Li, X.; Zhao, Z.; Zhao, G. Chloroplast genome structural characteristics and phylogenetic relationships of Oleaceae. Chin. Bull. Bot. 2019, 54, 441–454. [Google Scholar]
  11. Donoghue, M.J.; Bell, C.D.; Winkworth, R.C. The Evolution of Reproductive Characters in Dipsacales. Int. J. Plant Sci. 2003, 164, S453–S464. [Google Scholar] [CrossRef]
  12. Xiang, C.L.; Dong, H.J.; Landrein, S.; Zhao, F.; Yu, W.B.; Soltis, D.E.; Soltis, P.S.; Backlund, A.; Wang, H.F.; Li, D.Z.; et al. Revisiting the phylogeny of Dipsacales: New insights from phylogenomic analyses of complete plastomic sequences. J. Syst. Evol. 2019, 58, 103–117. [Google Scholar] [CrossRef]
  13. Zhang, W.H.; Chen, Z.D.; Li, J.H.; Chen, H.B.; Tang, Y.C. Phylogeny of the Dipsacales, s.l. based on chloroplast trnL-F and ndhF sequences. Mol. Phylogenet. Evol. 2003, 26, 176–189. [Google Scholar] [CrossRef]
  14. Landrein, S.; Prenner, G.; Chase, M.W.; Clarkson, J.J. Abelia and relatives: Phylogenetics of Linnaeeae (DipsacalesCaprifoliaceae s.l.) and a new interpretation of their inflorescence morphology. Bot. J. Linn. Soc. 2012, 169, 692–713. [Google Scholar] [CrossRef] [Green Version]
  15. Liu, H.R.; Gao, Q.B.; Zhang, F.Q.; Khan, G.; Chen, S.L. Westwards and northwards dispersal of Triosteum himalayanum (Caprifoliaceae) from the Hengduan Mountains region based on chloroplast DNA phylogeography. PeerJ 2018, 6, e4748. [Google Scholar] [CrossRef] [Green Version]
  16. Wu, Z.Y.; Raven, P.H.; Hong, D.Y. Caprifoliaceae . In Flora of China; Yang, Q., Landrein, S., Osborne, J., Borosova, R., Eds.; Science Press: Beijing, China, 2011; Volume 19, pp. 616–617. [Google Scholar]
  17. Chai, X.; Su, Y.; Yan, S.; Huang, X. Chemical Constituents of the Roots of Triosteum pinnatifidum. Chem. Nat. Compd. 2014, 50, 1149–1152. [Google Scholar] [CrossRef]
  18. Chai, X.; Su, Y.-F.; Zheng, Y.-H.; Yan, S.-L.; Zhang, X.; Gao, X.-M. Iridoids from the roots of Triosteum pinnatifidum. Biochem. Syst. Ecol. 2010, 38, 210–212. [Google Scholar] [CrossRef]
  19. Ding, H.; Chen, Y.; Zhao, J.; Song, Q.; Gao, K. Chemical constituents from the aerial parts of Triosteum pinnatifidum. Chem. Nat. Compd. 2013, 49, 95–96. [Google Scholar] [CrossRef]
  20. Li, Z. Studies on the chemical constituents of Triosteum himalayanum. Master Thesis, Lanzhou University, Lanzhou, China, 2009. [Google Scholar]
  21. Liu, H.R.; Khan, G.; Gao, Q.; Zhang, F.; Liu, W.; Wang, Y.; Fang, J.; Chen, S.; Afridi, S.G. Dispersal into the Qinghai-Tibet plateau: Evidence from the genetic structure and demography of the alpine plant Triosteum pinnatifidum. PeerJ 2022, 10, e1275. [Google Scholar] [CrossRef]
  22. 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]
  23. Bankevich, A.; Nurk, S.; Antipov, D.; Gurevich, A.A.; Dvorkin, M.; Kulikov, A.S.; Lesin, V.M.; Nikolenko, S.I.; Pham, S.; Prjibelski, A.D.; et al. SPAdes: A new genome assembly algorithm and its applications to single-cell sequencing. J. Comput. Biol. 2012, 19, 455–477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Kang, S.-J.; Park, H.-S.; Koo, H.J.; Park, J.Y.; Lee, D.Y.; Kang, K.B.; Han, S.I.; Sung, S.H.; Yang, T.J. The complete chloroplast genome sequence of Korean Lonicera japonica and intra-species diversity. Mitochondrial DNA Part B 2018, 3, 941–942. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Hu, H.; Liu, J.; An, J.; Wang, M.; Wang, Q. Characterization of the complete chloroplast genome of Lonicera macranthoides. Mitochondrial DNA Part B 2018, 3, 1002–1003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Li, Y.; Wang, J.; Li, P.; Wang, H. The characteristics of the chloroplast genome of the Michelia martini. Mitochondrial DNA Part B 2020, 5, 2725–2726. [Google Scholar] [CrossRef] [PubMed]
  27. 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]
  28. Amiryousefi, A.; Hyvonen, J.; Poczai, P. IRscope: An online program to visualize the junction sites of chloroplast genomes. Bioinformatics 2018, 34, 3030–3031. [Google Scholar] [CrossRef]
  29. Yin, X.; Wen, Q.; Wang, J.; Li, T.; Ye, J.; Xu, L. Characterization of microsatellites in complete chloroplast genome of the genus Camellia and marker development. Mol. Plant Breed. 2018, 16, 6761–6769. [Google Scholar]
  30. Yang, Y.; Jiao, J.; Fan, X.; Zhang, Y.; Jiang, J.; Li, M.; Liu, C. Complete Chloroplast Genome Sequence and Characteristics Analysis of Vitis ficifolia. Acta Hortic. Sin. 2019, 46, 635–648. [Google Scholar] [CrossRef]
  31. Thompson, J.D.; Higgins, D.G.; Gibson, T.J. CLUSTAL W: Improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Res. 1994, 22, 4673–4680. [Google Scholar] [CrossRef] [Green Version]
  32. Nguyen, L.T.; Schmidt, H.A.; von Haeseler, A.; Minh, B.Q. IQ-TREE: A fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol. Biol. Evol. 2015, 32, 268–274. [Google Scholar] [CrossRef]
  33. Drummond, A.J.; Suchard, M.A.; Xie, D.; Rambaut, A. Bayesian phylogenetics with BEAUti and the BEAST 1.7. Mol. Biol. Evol. 2012, 29, 1969–1973. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Cheng, Y.; Yang, Y.; Fu, X.; Liu, L.; Jiang, Z.; Cai, J. Plastid genomes of Elaeagnus mollis: Comparative and phylogenetic analyses. J. Genet. 2020, 99, 85. [Google Scholar] [CrossRef] [PubMed]
  35. Huang, J.; Zhang, C.; Zhao, X.; Fei, Z.; Wan, K.; Zhang, Z.; Pang, X.; Yin, X.; Bai, Y.; Sun, X.; et al. The jujube genome provides insights into genome evolution and the domestication of sweetness/acidity taste in fruit trees. PLoS Genet. 2016, 12, e1006433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Liu, M.L.; Fan, W.B.; Wang, N.; Dong, P.B.; Zhang, T.T.; Yue, M.; Li, Z.H. Evolutionary Analysis of Plastid Genomes of Seven Lonicera, L. Species: Implications for Sequence Divergence and Phylogenetic Relationships. Int. J. Mol. Sci. 2018, 19, 4039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Alzahrani, D.A. Complete Chloroplast Genome of Abutilon fruticosum: Genome Structure, Comparative and Phylogenetic Analysis. Plants 2021, 10, 270. [Google Scholar] [CrossRef] [PubMed]
  38. Wang, Y.; Wang, S.; Liu, Y.; Yuan, Q.; Sun, J.; Guo, L. Chloroplast genome variation and phylogenetic relationships of Atractylodes species. BMC Genom. 2021, 22, 103. [Google Scholar] [CrossRef] [PubMed]
  39. Alzahrani, D.A.; Yaradua, S.S.; Albokhari, E.J.; Abba, A. Complete chloroplast genome sequence of Barleria prionitis, comparative chloroplast genomics and phylogenetic relationships among Acanthoideae. BMC Genom. 2020, 21, 393. [Google Scholar] [CrossRef]
  40. Li, Y.; Dong, Y.; Liu, Y.; Yu, X.; Yang, M.; Huang, Y. Comparative Analyses of Euonymus Chloroplast Genomes: Genetic Structure, Screening for Loci with Suitable Polymorphism, Positive Selection Genes, and Phylogenetic Relationships within Celastrineae. Front. Plant Sci. 2020, 11, 593984. [Google Scholar] [CrossRef]
  41. Raman, G.; Park, K.T.; Kim, J.H.; Park, S. Characteristics of the completed chloroplast genome sequence of Xanthium spinosum: Comparative analyses, identification of mutational hotspots and phylogenetic implications. BMC Genom. 2020, 21, 855. [Google Scholar] [CrossRef]
  42. Chincoya, D.A.; Sanchez-Flores, A.; Estrada, K.; Diaz-Velasquez, C.E.; Gonzalez-Rodriguez, A.; Vaca-Paniagua, F.; Davila, P.; Arias, S.; Solorzano, S. Identification of High Molecular Variation Loci in Complete Chloroplast Genomes of Mammillaria (Cactaceae, Caryophyllales). Genes 2020, 11, 830. [Google Scholar] [CrossRef]
  43. Guo, Y.Y.; Yang, J.X.; Li, H.K.; Zhao, H.S. Chloroplast Genomes of Two Species of Cypripedium: Expanded Genome Size and Proliferation of AT-Biased Repeat Sequences. Front. Plant Sci. 2021, 12, 609729. [Google Scholar] [CrossRef] [PubMed]
  44. Ran, H.; Liu, Y.; Wu, C.; Cao, Y. Phylogenetic and Comparative Analyses of Complete Chloroplast Genomes of Chinese Viburnum and Sambucus (Adoxaceae). Plants 2020, 9, 1143. [Google Scholar] [CrossRef] [PubMed]
  45. Gu, L.; Wu, Q.; Yi, Y.; Yu, Z. The complete chloroplast genome of Lonicera hypoglauca Miq (Caprifoliaceae: Dipsacales) from Guangxi, China. Mitochondrial DNA Part B 2021, 6, 450–451. [Google Scholar] [CrossRef] [PubMed]
  46. Rajesh, M.K.; Gangaraj, K.P.; Prabhudas, S.K.; Prasad, T.S.K. The complete chloroplast genome data of Areca catechu (Arecaceae). Data Brief 2020, 33, 106444. [Google Scholar] [CrossRef]
  47. Parvathy, S.T.; Udayasuriyan, V.; Bhadana, V. Codon usage bias. Mol. Biol. Rep. 2022, 49, 539–565. [Google Scholar] [CrossRef]
  48. Randall, T.A.; Dwyer, R.A.; Huitema, E.; Beyer, K.; Cvitanich, C.; Kelkar, H.; Fong, A.V.A.; Gates, K.; Roberts, S.; Yatzkan, E.; et al. Large-Scale gene discovery in the oomycete Phytophthora infestans reveals likely components of phytopathogenicity shared with true fungi. Mol. Plant-Microbe Interact. 2005, 18, 229. [Google Scholar] [CrossRef] [Green Version]
  49. Downie, S.R.; Jansen, R.K. 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, 336–351. [Google Scholar] [CrossRef]
  50. Zhang, R.; Zhang, L.; Wang, W.; Zhang, Z.; Du, H.; Qu, Z.; Li, X.Q.; Xiang, H. Differences in Codon Usage Bias between Photosynthesis-Related Genes and Genetic System-Related Genes of Chloroplast Genomes in Cultivated and Wild Solanum Species. Int. J. Mol. Sci. 2018, 19, 3142. [Google Scholar] [CrossRef] [Green Version]
  51. Jacobs, B.; Pyck, N.; Smets, E. Phylogeny of the Linnaea clade: Are Abelia and Zabelia closely related? Mol. Phylogenet. Evol. 2010, 57, 741–752. [Google Scholar] [CrossRef]
  52. Winkworth, R.C.; Bell, C.D.; Donoghue, M.J. Mitochondrial sequence data and Dipsacales phylogeny: Mixed models, partitioned Bayesian analyses, and model selection. Mol. Phylogenet. Evol. 2008, 46, 830–843. [Google Scholar] [CrossRef]
  53. Bell, C.D.; Donoghue, M.J. Dating the Dipsacales: Comparing models, genes and evolutionary implications. Am. J. Bot. 2005, 92, 284–296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Liu, H.R.; Fang, J.; Xia, M.; Xiao, Q.; Zhang, D.; Xie, J.; Chen, S. The complete chloroplast genome of Triosteum himalayanum (Caprifoliaceae), a perennial alpine herb. Mitochondrial DNA Part B 2019, 4, 4194–4195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Gene map of the Triosteum sinuatum chloroplast genome. Genes on the inside of the circle are transcribed in the clockwise direction, and genes outside the circle are transcribed in the counterclockwise directions. Genes with different functions are shown in different color blocks. The inverted repeat (IRa and IRb) regions are indicated by the thick lines. LSC and SSC indicate large and small single copy, respectively.
Figure 1. Gene map of the Triosteum sinuatum chloroplast genome. Genes on the inside of the circle are transcribed in the clockwise direction, and genes outside the circle are transcribed in the counterclockwise directions. Genes with different functions are shown in different color blocks. The inverted repeat (IRa and IRb) regions are indicated by the thick lines. LSC and SSC indicate large and small single copy, respectively.
Genes 13 00933 g001
Figure 2. Comparisons of LSC, SSC and IR region boundaries among the chloroplast genomes of 20 species in Dipsacales.
Figure 2. Comparisons of LSC, SSC and IR region boundaries among the chloroplast genomes of 20 species in Dipsacales.
Genes 13 00933 g002
Figure 3. ML tree of 26 species. Twenty species are within Dipsacales.
Figure 3. ML tree of 26 species. Twenty species are within Dipsacales.
Genes 13 00933 g003
Figure 4. Divergence times analysis based on the chloroplast plastome.
Figure 4. Divergence times analysis based on the chloroplast plastome.
Genes 13 00933 g004
Table 1. General characteristics of 20 chloroplast genomes in Dipsacales.
Table 1. General characteristics of 20 chloroplast genomes in Dipsacales.
FamilySpeciesGenebankSize (bp)LSC (bp)SSC (bp)IR (bp)Total GC (%)Total GenesProtein GenestRNA GenesrRNA Genes
MorinaceaeMorina longifoliaNC045046157,14990,29218,67924,08938.6012883378
Acanthocalyx albaNC045055157,99990,75019,41323,91838.3612883378
DipsacaceaeTriplostegia glanduliferaNC045051157,56090,23921,09923,11137.9312883378
Dipsacus japonicusMN524645154,70988,60518,53823,78338.9912883378
AdoxaceaeAdoxa moschatellinaNC034792157,23886,34018,67426,11237.7912881398
Sinadoxa corydalifoliaNC032040157,07486,43018,66825,98837.7613183408
Sambucus williamsiiNC033878158,30586,81018,93326,28137.9313284408
Viburnum odoratissimumMN894600158,75787,35018,40926,49938.1013287378
Viburnum dilatatumMW346666158,39287,07018,24226,54038.1013083398
ValerianaceaePatrinia scabraNC045048151,26787,26918,02022,98938.5812883378
Valeriana officinalisNC045052151,50587,61916,28823,79938.4512883378
CaprifoliaceaeTriosteum pinnatifidumNC037952154,89689,00720,54322,67338.5412883378
Triosteum himalayanumNC045219154,57989,15718,68223,37038.3813286388
Triosteum sinuatumMW526077157,17890,75818,65623,88238.3113586418
Lonicera ferdinandiNC040963154,51388,55418,58923,68538.4412883378
LinnaeaceaeLinnaea borealisMN548092161,57685,60917,54729,21038.2613489378
Zabelia dielsiiNC046599155,58489,65219,06223,43538.3912883378
Dipelta yunnanensisNC042201155,35989,46719,09823,39738.4512883378
Dipelta floribundaNC037955155,37089,43119,05223,37838.3812883378
DiervillaceaeWeigela floridaMN524626157,96690,14421,48623,16838.0112883378
Abbreviations: GC, GC content.
Table 2. List of genes within chloroplast genomes of 20 species.
Table 2. List of genes within chloroplast genomes of 20 species.
CategoryGroup of GenesName of Genes
Transcription and TranslationLarge subunit of ribosomerpl33, rpl20, rpl36, rpl14, rpl16 a, rpl22, rpl2 a, rpl23, rpl32
Small subunit of ribosomerps16, rps2, rps14, rps4, rps18, rps12 b,*, rps11, rps8, rps3, rps19, rps7*, rps15
RNA polymeraserpoC2, rpoC1 a, rpoB, rpoA
Ribosomal RNAsrrn16*, rrn23*, rrn4.5*, rrn5*
Transfer RNAstrnH-GUG, trnK-UUU a, trnQ-UUG, trnS-GCU, trnG-GCC a, trnR-UCU, trnC-GCA, trnD-GUC, trnY-GUA, trnE-UUC, trnT-GGU, trnS-UGA, trnG-UCC, trnfM-CAU, trnS-GGA, trnT-UGU, trnL-UAAa, trnF-GAA, trnV-UAC a,*, trnM-CAU, trnW-CCA, trnP-UGG, trnI-CAU *, trnL-CAA *, trnV-GAC *, trnI-GAU *, trnA-UGC *, trnR-ACG *, trnN-GUU *, trnL-UAG
PhotosynthesisPhotosystem IpsaB, psaA, psaI, psaJ, psaC
Photosystem IIpsbA, psbK, psbI, psbM, psbD,
psbC, psbZ, psbJ, psbL, psbF, psbE, psbB, psbT, psbN, psbH
NADH dehydrogenasendhJ, ndhK, ndhC, ndhB a,*, ndhF, ndhD, ndhE, ndhG, ndhI, ndhA a, ndhH
Cytochrome b6/f complexpetN, petA, petL, petG, petB a, petD a
ATP synthaseatpA, atpF a, atpH, atpI, atpE, atpB
Rubisco large subunitrbcL
ATP-dependent
protease subunit p
clpP a
Other genesChloroplast envelope
membrane protein
cemA
MaturasematK
c-typeccsA
Subunit
Acetyl- CoA-Carboxylate
accD
Genes of unknown functionConserved ORFsycf1, ycf2 *, ycf3, ycf4
a Gene with one intron; b gene with two introns; * gene with copies.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, H.; Liu, W.; Ahmad, I.; Xiao, Q.; Li, X.; Zhang, D.; Fang, J.; Zhang, G.; Xu, B.; Gao, Q.; et al. Complete Chloroplast Genome Sequence of Triosteum sinuatum, Insights into Comparative Chloroplast Genomics, Divergence Time Estimation and Phylogenetic Relationships among Dipsacales. Genes 2022, 13, 933. https://0-doi-org.brum.beds.ac.uk/10.3390/genes13050933

AMA Style

Liu H, Liu W, Ahmad I, Xiao Q, Li X, Zhang D, Fang J, Zhang G, Xu B, Gao Q, et al. Complete Chloroplast Genome Sequence of Triosteum sinuatum, Insights into Comparative Chloroplast Genomics, Divergence Time Estimation and Phylogenetic Relationships among Dipsacales. Genes. 2022; 13(5):933. https://0-doi-org.brum.beds.ac.uk/10.3390/genes13050933

Chicago/Turabian Style

Liu, HaiRui, WenHui Liu, Israr Ahmad, QingMeng Xiao, XuMin Li, DeJun Zhang, Jie Fang, GuoFan Zhang, Bin Xu, QingBo Gao, and et al. 2022. "Complete Chloroplast Genome Sequence of Triosteum sinuatum, Insights into Comparative Chloroplast Genomics, Divergence Time Estimation and Phylogenetic Relationships among Dipsacales" Genes 13, no. 5: 933. https://0-doi-org.brum.beds.ac.uk/10.3390/genes13050933

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