Skip to main content

Plastome structure, phylogenomics and evolution of plastid genes in Swertia (Gentianaceae) in the Qing-Tibetan Plateau

Abstract

Background

The genus Swertia is of great medicinal importance and one of the most taxonomically challenging taxa within Gentianaceae, largely due to the morphological similarities of species within this genus and with its closely related genera. Previous molecular studies confirmed its polyphyly but suffered from low phylogenetic resolutions because only limited sequence loci were used. Thus, we conducted the structural, gene evolutionary, and phylogenetic analyses of 11 newly obtained plastomes of Swertia. Our result greatly improved the phylogenetic resolutions in Swertia, shed new light on the plastome evolution and phylogenetic relationships of this genus.

Results

The 11 Swertia plastomes together with the published seven species proved highly similar in overall size, structure, gene order, and content, but revealed some structural variations caused by the expansion and contraction of the IRb region into the LSC region, due to the heterogeneous length of the ψycf1. The gene rps16 was found to be in a state flux with pseudogenes or completely lost. Similar situation was also documented in other genera of Gentianaceae. This might imply loss of the gene in the common ancestor of Gentianaceae. The distribution plot of ENC vs. GC3 showed all these plastomes arranging very close in the Wright line with an expected ENC value (49–52%), suggesting the codon usage of Swertia was mainly constrained by a GC mutation bias. Most of the genes remained under the purifying selection, however, the cemA was identified under positive selection, possibly reflecting an adaptive response to low CO2 atmospheric conditions during the Late Miocene. Our phylogenomic analyses, based on 74 protein-coding genes (CDS), supported the polyphyly of Swertia with its close allies in the subtribe Swertiinae, presumably due to recent rapid radiation. The topology inferred from our phylogenetic analyses partly supported the current taxonomic treatment. Finally, several highly variable loci were identified, which can be used in future phylogenetic studies and accurate identification of medicinal genuineness of Swertia.

Conclusions

Our study confirmed the polyphyly of Swertia and demonstrated the power of plastome phylogenomics in improvement of phylogenetic resolution, thus contributing to a better understanding of the evolutionary history of Swertia.

Peer Review reports

Background

The genus Swertia Linnaeus, belonging to subtribe Swertiinae, family Gentianaceae with reported 168 species, is widely distributed among Asia, Africa, North America, and Europe [1]. Seventy-nine species of Swertia are distributed over China [2], with a diversification center in the Qing-Tibetan Plateau region, especially in south-western China and the Himalaya region [1,2,3]. Ho described Swertia in China and classified them under seven sections [2]. Later, based on their previous studies, Ho and Liu published a worldwide classification of Swertia, comprising 168 species, under three subgenus, 11 sections [1].

Swertia can be recognized with rotated flowers, one or two nectaries per petal lobe, and a variety of different appendages around the nectaries. However, this genus is one of the most taxonomically difficult genera within Gentianaceae, largely due to the morphological similarities (i.e., nectariferous and rotate corolla lobes) of the species in Swertia and with its allies, namely, Halenia, Lomatogonium, Lomatogoniopsis and veratrilla [1, 4,5,6,7]. Over the past decades, continues efforts have been devoted to molecular phylogeny studies of Swertia based on ITS and plastid sequence loci. It has become clear that Swertia was highly polyphyletic, intertwining with other genera in the subtribe Swertiinae on the inferred trees. Some clades were left unresolved and usually with weak node support, e.g., the Comastoma-Lomatogonium-sect. Platynema, Halenia-sect. Swertopsis, most likely due to only a few molecular markers were employed [4,5,6, 8]. More recently, a study based on 76 protein-coding genes generated from plastomes, supported Swertia’s polyphyly with greatly improved resolution when compared with those earlier molecular studies, however, only a few Swertia species were included as the representative of this genus [9].

Species of Swertia have long been used in traditional herbal medicine in Asian countries, including China, India, Korea, and Japan. The crude drug of Swertia is applied to treat a variety of diseases, including hepatic, choleric, inflammatory diseases, and bitter stomachic [10,11,12]. These herbaceous plants are rich in xanthones, flavonoids, iridoid and seco-iridoid glycosides, terpenoids, and alkaloids [13]. Despite the excellent medicinal value and abundant medicinal studies, the accurate identification of genuineness of Swertia species has been proved difficult and still lacks of DNA barcoding method [14]. To this aim, barcoding by a comparative plastome method of Swertia authentication is needed to assure good medicinal quality.

In almost all land plants, plastomes are highly conserved in structural organization, gene arrangement, and gene content [15,16,17]. A typical plastome usually has a quadripartite structure of a large single copy region (LSC) of 82–90 kb, a small single copy (SSC) of 15–20 kb, and two identical copies of inverted regions (IRa & IRb) [17]. In photosynthetic angiosperms, most plastomes contain 110–133 distinct genes, including approximately 80 protein-coding genes (CDS), 30 transfer RNA (tRNA) genes, and four ribosomal RNA (rRNA) genes [17]. Due to their conserved structure, lack of recombination, and uniparental inheritance [15, 18], plastomes data have recently been employed to generate highly resolved phylogenies or to efficiently barcode and identify plant species, especially in taxonomically complex groups [19,20,21,22,23].

Here, we report newly sequenced 11 plastomes of Swertia. Our aims were to: (a) infer the plastome structural evolution and adaptation of Swertia; (b) identify the most variable regions as candidate DNA barcodes for future Swertia species identification; (c) test the phylogeny of Swertia in regard to its current taxonomic treatment.

Results

Plastome features of Swertia

For the 11 newly sequenced Swertia species, Illumina pair end sequencing generated 13,182,812 (S. dichotoma) to 60,069,018 (S. przewalskii) raw reads. After quality filtering, 13,167,950 (S. dichotoma) to 59,482,032 (S. przewalskii) clean reads were obtained. The assembled plastomes of the 11 Swertia species had coverage depth from 760 X (S. dichotoma) to 7916 X (S. cincta) (Additional file 9, Fig. S1). The raw sequencing data were deposited in the NCBI GenBank (Additional file 1, Table S1).

The lengths of the 20 plastomes varied from 149,036 (S. pubescens) to 154,073 bp (S. kouitchensis) (Fig. 1, Table 1). All the 20 Swertia plastomes contained two copies of IR regions, each ranging from 25,331 (S. cincta) to 25,889 bp (S. cordata). The IRs were separated by a LSC region ranging from 83,610 (S. cordata) to 80,432 bp (S. pubescens) and a SSC region ranging from 17,903 (S. hispidicalyx) to 18,547 bp (S. tetraptera). The overall GC content of the 20 Swertia plastomes were similar and ranged from 38% in S. tetraptera to 38.2% in S. cincta, S. macrosperma, and S. pubescens (Table 1).

Fig. 1
figure 1

Plastomes maps of Swertia. Black arrows indicated pseudogenes other than ψycf1 and ψrps19

Table 1 Summary of 20 Swertia plastomes

The 20 Swertia plastomes contained 129–131 genes, composing of 84–86 protein-coding genes (CDS), 8 rRNAs and 37 tRNAs (Fig. 1, Table 1). Twelve CDSs and six tRNAs contained one intron. The clpP and ycf3 contained two introns (Table 2). The intron of trnK-UUU included the coding sequence for the matK. Rps12 was identified as a trans-spliced gene, with the first exon (5′ end of the sequence) in the LSC region (rpl20_clpP), and the remaining two exons (3′ end of the sequence) in the IR region (rps7_trnV-GAC). Ycf1 sequence was duplicated in all the 20 Swertia, with one complete copy (1,014–5,517 bp) located on the SSC-IRa border served as a functional gene, and one incomplete copy (920–1,008 bp) on the IRb-SSC border identified as a pseudogene. The same situation was also found in rps19, one complete copy (279 bp) located on the LSC-IRb border acted as a functional gene, while the incomplete copy (161–303 bp) on the Ira-LSC border identified as a pseudogene. ψycf1 and ψrps19 were resulted from incomplete duplications of their functional copes. Furthermore, rps16 was either identified as a pseudogene (in 16 species) or completely lost (in four species). Also, ψinfA were observed in 11 Swertia plastomes. In addition, ψaccD were found in S. cordata and S. mussotii, ψrpl33 in S. hispidicalyx, and Ψycf2 in S. wolfganiana (Fig. 1, Table 3).

Table 2 Genes annotated in 20 Swertia plastomes
Table 3 Statistics of pseudogenes

The boundaries between IR and SSC/LSC regions in the 20 plastomes were compared (Fig. 2). The LSC/IRb junction was found only within the rps19. The Ira/SSC junction was found within the ycf1. The boundary between IRb and LSC slightly varied: 12 species were within the ndhF gene, and seven species within the overlapping region of the ψycf1 and ndhF, only S. tetraptera was identified within the ψycf1. The Ira/LSC border was either within ψrps19 or adjacent to the trnH-GUG.

Fig. 2
figure 2

Comparison of LSC, IRs and SSC borders among 20 Swertia plastomes

Codon preference analysis

The total sequence sizes of the CDSs for codon analysis were 66,522–79,371 bp in the 20 Swertia plastomes, and they encoded 22,174–26,457 codons (Additional file 3, Table S3). Isoleucine (Ile) was encoded by the highest number of codons (1,901–2,223), whereas cysteine (Cys) was the least (261–308). The relative synonymous codon usage (RSCU) values of all codons in the form of a heatmap were shown in Fig. 3. The red indicated higher RSCU values and the blue indicated lower RSCU values. Thirty codons were used frequently with RSCU > 1. UUA showed a preference in all the 20 plastomes. All the biased 30 codons ended with a purine (A/U). The total GC content (GC, 0.383–0.398), GC content at first codon position (GC1, 0.456–0.463), GC content at second codon position (GC2, 0.384–0.391), and GC content at third codon position (GC3, 0.298–0.313) were less than 0.5, suggesting that the 20 plastomes tended to use A/T bases and A/T–ending codons. The codon adaptation index (CAI) values between 0.162 and 0.167 indicated a slight bias in codon usage in the 20 Swertia species. The values of the effective number of codons (ENC) ranged from 49.71 (S. wolfangiana) to 51.24 (S. cordata) (Table 4, Additional file 4, Table S4). Further, the ENC vs. GC3 plot (Fig. 4) showed the 20 Swertia plastomes clustered just below the Wright Curve[24, 25], and ranging around the expected ENC of 50% GC3.

Fig. 3
figure 3

The RSCU values of all merged CDSs among 20 Swertia plastomes. Color key: the red values indicated higher RSCU values while the blue values indicated lower RSCU values

Table 4 Codon feature of the 20 Swertia plastomes
Fig. 4
figure 4

ENC-plot analysis

Characterization of repeat sequences and simple sequence repeats (SSR)

In our study, a total of 980 repeat sequences were detected in the 20 Swertia plastomes, including four categories: forward, reverse, complement, and palindromic. The numbers of the four repeated types were as follows: 18–23 in forward repeats, 3–11 in reverse, 0–2 in complement, and 17–24 in palindromic (Fig. 5 a, Additional file 5, Table S5). The sizes of repeats ranged from 17 to 509 bp (Fig. 5 b, Table 5, Additional file 5, Table S5). According to the length of repeat sequences, we artificially divided them into six categories: 17–26 bp, 27–36 bp, 37–46 bp, 47–56 bp, 57–66 bp, and > 66 bp. Among them, the majority of repeats (81%) were 17–26 bp long, while only one repeat was found > 66 bp (509 bp forward repeat) in S. cordata. Slightly more than half (53.4%) of the repeat sequences were located in intergenic spacer (IGS).

Fig. 5
figure 5

Analysis of repeat sequencies in the 20 Swertia plastomes. a Total number of four repeat types. b Number of repeats divided by length

Table 5 Repeat sequencies in the 20 Swertia plastomes

A total of 588 simple sequence repeats (SSRs) were identified across the 20 Swertia plastomes (Table 6). The majority (94.4%) of these SSRs were monomers, particularly A/T, and their lengths were mostly in the 12–19 bp range. There were only 2.7% of dimers (AT and TA) and 2.9% of trimers (GAA, TTA, and ATT), no tetramers, hexamers, and pentamers were found.

Table 6 SSRs types in 20 Swertia plastomes

Hypervariable regions

Hypervariable regions were identified by genome-wide comparative analyses with S. erythrostica as the reference (Fig. 6). Our nucleotide diversity analyses included 79 CDSs (protein-coding regions) and 74 IGS and intron regions (non-coding regions) (Additional file 6, Table S6). The pairwise sequence divergence (Pi) values for each CDS ranged from 0.00035 (rpl23) to 0.05872 (rps16) with an average of 0.01271, in which 10 genes (rps16, ccsA, rpl22, ndhF, rps19, matK, infA, rps15, rpl20, rpl32) had remarkably high values (Pi > 0.02207). For the 74 non-coding regions, Pi values ranged from 0.00107 (rpl2 intron) to 0.07895 (ndhC_trnV-UAC) with an average of 0.029679. Again, 10 of the non-coding regions (ndhC_trnV-UAC, petB intron, trnL-UAA intron, rps16_trnK-UUU, trnL-UCC_trnG-UCC, trnG-UCC_trnR-UCU, trnS-GGA_rps4, psbM_trnD-GUC, trnG-GCC_trnfM-CAU) had higher Pi values than the other regions (Fig. 7a, b). The non-coding regions exhibited more variation than the coding regions (Fig. 7c). The SSC region had the highest divergence and the IR regions had the lowest (Fig. 7d).

Fig. 6
figure 6

mVISTA-based sequence identity plot of 20 Swertia plastomes compared using S. erythrosticta as a reference. The top grey arrow shows genes in order (transcription direction) and the position of each gene. A 70% cut-off was used for the plots. The vertical scale represents the percentage of identity between 50 and 100%. Coding and non-coding regions (IGS/intron) are marked in purple and red, respectively, while conserved genes are in blue

Fig. 7
figure 7

The genetic diversity based on Kimura’s two-parameter model. a The Pi values of protein-coding genes. b The Pi values of introns and intergenic regions. c Boxplots of Pi values difference among LSC, IR and SSC regions. d Boxplots of Pi values differences between protein-coding genes and intron & intergenic regions

Phylogenetic analysis

Both ML and BI analyses based on the 74 common CDSs shared between the 35 plastomes (20 Swertia, 10 Swertiinae, four Gentianinae, and one outgroup species) (Additional file 2, Table S2) inferred completely identical topologies with high support values (Fig. 8). The monophyly of both subtribe Swertiinae and subtribe Gentianinae was strongly supported (BS = 100%, PP = 1). Within Swertiinae, Swertia was polyphyletic in relation to the other genera (Comastoma, Lomatogoniopsis, Lomatogonium, Halenia, and Veratrilla) and distributed in four main clades. Besides, Pterygocalyx volubilis and two Gentianopsis consisted of a monophyletic basal clade, being sister to the four clades mentioned above. In Gentianinae lineage, three species of different genera formed a well-supported clade, being sister clade to Swertiinae (BS = 100%, PP = 1). Moreover, while most of these clades were distinct, however, S. kouitchensis, S. punicea, S. franchetiana, and S. mussotii clustered together in clade II with a near-zero branch length. These four species showed very low (0.0000683–0.0006831) pairwise distance values when compared to the overall mean pairwise distance (0.01) among all the 20 Swertia species (Additional file 8, Table S8). The same scenario was also found in clade IV, where S. przewalskii, S. bifolia, S. souliei, and S. wolfangiana clustered together with remarkably low pairwise distance values (0.0000929–0.0007623) (Additional file 8, Table S8). According to the pairwise distance calculation, the largest sequence divergence was found between S. tetraptera and S. pubescens (0.025819), and the lowest was between S. franchetiana and S. mussotii (0.0000683) (Additional file 8, Table S8).

Fig. 8
figure 8

Phylogenic tree derived from 74 CDSs. The black spots indicate phylogenetic support values both from maximum likelihood and Bayesian with 100% bootstrap support (BS) and 1.00 posterior probability (PP) respectively. The double slashes indicate an artificial shortening of this branch, as the outgroup, Coffer arabica, produces a very long branch. The terminals are color-coded, with each color a clade in the phylogenic tree corresponding to a section in Ho TN taxonomic treatment, and a picture of species representing its section and/or series. The species displayed are A, S. erythrosticta; B, S. multicaulis; C, S. cordata; D, S. dichotoma; E, S. przewalskii; F. S. verticillifolia; G, S. mussotii; H, S. cincta. Photographs: Gao Qingbo, Ma Xiaolei, Cao Qian, Zhou Yubi, Xu Bo, Yang Fengmao

Selective pressure in plastid genes

The site-specific model showed none of the 50 CDSs of the 35 Gentianaceae species was under strong positive selection. Though, the cemA (ω = 0.59623) and ycf1 (ω = 0.95807) were found under slightly positive selection. The ω ratios of the most genes were less than 0.2, indicating that they were under a strong purifying selection (Additional file 7, Table S7). The branch-specific model detected that the cemA was under positive selection in most Swertia clades along the phylogenetic tree (Fig. 8, Additional file 10, Fig. S2), including Clade I (ω = 290), Clade III (ω = 35) and Clade IV (ω = 999), but not for Clade II (ω < 1). For the cemA, the two-ratio model (m2) was significantly favoured over the one-ratio model (m0, 2ΔL = 54.96; P < 0.05). However, for the ycf1, both two-ratio model (m2) and free-ratio model (m1) were refused in the LRT test, indicating the same ω ratio of the ycf1 for all branches in the phylogeny (Additional file 7, Table S7).

Discussion

Plastome structural evolution and likely adaptation of Swertia

The plastome sequences newly obtained from the 11 Swertia species, plus previously sequenced eight species (Addition file 1) were very similar in the overall structure, length, gene order, and contents (Table 1 and Fig. 1). However, the pseudogenes, especially ψrps16 and ψinfA, deserved closer scrutiny. The rps16 was in a state flux with pseudogene in 16 species (S. bifolia, S. cordata, S. diluta, S. erythrosticta, S. franchetiana, S. macrosperma, S. przewalskii, S punicea, S. wolfganiana, S. tetraptera, S. kouitchensis, S. leducii, S. multicaulis, S. mussotii, S. verticillifolia, S. souliei) and entirely missing in the rest four species (S. cincta, S. dichotoma, S. pubescens, S. hispidicalyx). The pseudogene rps16 was caused by loss of the second exon and part of the intron. Losses of plastid-encoded rps16 have been reported across a wide range of seed plants [15]. Among Gentianaceae with published plastome sequences, the loss of rps16 has been previously documented for members of variable genera [26,27,28,29,30]. We assumed that the loss of functionality of rps16 might be a synapomorphy shared in the whole family of Gentianaceae, implying an ancient loss in the ancestor of Gentianaceae. The gene loss was likely resulted from the intracellular nuclear-coded rps16 that transferred from mitochondria [31,32,33].

The InfA was identified to be pseudogene in 11 Swertia species. Among them, there was no start codon in 10 species (S. cordata, S. bifolia, S. leducii, S. franchetiana, S. mussotii, S. hispidicalyx, S. verticillifolia, S. souliei, S. wolfganiana, S. erythrosticta), but in S. macrosperma, a 59 bp-long-sequence has been deleted. The infA was also found to be pseudogene in seven Gentiana species [34], a sister clade of Swertia in Gentianaceae. In an intensive comparative research on the infA, the entirely missing and/or pseudogene have been reported occurred across 309 angiosperms, suggesting this gene has been transferred to the nucleus many times in different seed plants [35].

Mutation pressure on DNA sequences and natural selection are the two major factors that have been widely accepted to account for codon usage bias [36]. Our analysis of the 20 Swertia plastomes revealed a codon bias having A or U at the third position (Additional file 3, Fig. 3). In the absence of selection pressures, this A/U bias has increased the RSCU values for A/U ending codons. The fact that chloroplast genomes had similar biases of codon usage suggested that chloroplasts, in general, might have particular characteristics of codon usage [37], our finding was consistent with previous reports from many other chloroplast genomes [38,39,40]. The distribution plot of ENC vs. GC3 for the 20 Swertia plastomes showed all the plastomes lying very close in the Wright line [24] with an expected ENC value (49–52%), suggesting the codon usage of Swertia was mainly constrained by a GC mutation bias but not natural selection (Fig. 4).

Long repeat sequences of plastomes have been reported to play major roles in genomic rearrangements and sequence variations [41, 42]. However, our study showed no correlation between large repeats and rearrangement events. SSRs have been used widely in plant genomic and evolutionary studies because of their high amounts of variability within species (Table 6) [43, 44]. Our study showed the most abundant SSRs were A/T rich mononucleotide, which was consistent with previous reports that SSRs usually consisted of polyA or polyT repeats and rarely contain G or C repeats [45, 46].

Our selective pressure analyses of 50 common CDSs (> 300 bp) extracted from the 35 Gentianaceae plastomes only identified the cemA being positively selected. The cemA was found to encode an envelope membrane protein [47]. More recent researches revealed the disruption of this gene led to increased light sensitivity and impacted the CO2 transportation system in plants [48, 49]. Genes related to a specific environment are normally assumed to be under positive selection [48]. Previous studies suggested that the origin and speciation of Swertiinae occurred within 10 Ma (the Late Miocene), and Swertia within 4 Ma (the Pliocene to the Quaternary), possibly triggered by the uplift of the Tibetan Plateau and the climate change associated [6, 50, 51]. During Mid to Late Miocene, vast areas on earth had experienced a climate cooling transition [52, 53], its driven factor, as believed by many researchers, was the decline in atmospheric CO2 levels, which subsequently led to a transition of the modern ecosystem, e.g., the global radiation of C4 plants [54,55,56,57,58]. Therefore, the positive selection detected in cemA might indicate it occurred in the adaptation and speciation processes in Swertia species during the Late Miocene as a response to low level of atmospheric CO2. Interestingly, more positive selection was detected in species from higher-elevation regions in the Qing-Tibetan Plateau region of China (Clade IV) but no positive selection was found in species from lower-elevation regions (the eastern Qing-Tibetan Plateau and the south western China) (Clade II) (Additional file 10, Fig. S2), suggesting an adaptive response to the low atmospheric CO2 environment of the Qinghai-Tibetan Plateau region. However, more Swertia species are needed to be included to test this hypothesis whether positive selection detected on cemA as a beneficial result in the evolution of Swertia genus. In addition, further studies are needed to uncover whether changes in selection pressure are associated with specific biochemical pathways or functions of the cemA in the adaptation and divergence of Swertia.

Candidate DNA barcodes

DNA barcodes are defined as short DNA sequences with adequate variations to identify species in a given taxonomic group [59]. Variable Loci (e.g., matK, rbcL, trnH-psbA) from plastomes have been widely used to determine the phylogeny of seed plants [59, 60]. Variable cpDNA loci have been employed to uncover phylogeny patterns of Swertia, including trnL-UAA intron, trnL-UAA_trnF-GAA, and trnS-UGA_ycf9 [4], matK and trnL-UAA intron [5], matK [8], matK and rbcL [6]. However, these results were suffered from weak solutions. Furthermore, several clades were left unsolved in previously published phylogenetic trees. Therefore, additional loci are needed to be included to achieve higher phylogenetic reference of the genus Swertia. Moreover, loci used in these previous studies (e.g., matK, trnL-UAA intron, rbcL), are today considered intermediately variable regions [61]. Based on our nucleotide diversity analyses, only trnL-UAA intron had a high Pi value (ranked four), other loci had intermediate to low Pi values (e.g., trnL-UAA_trnF-GAA ranked 19, matK ranked 49, rbcL ranked 111) (Additional file 6, Table S6). Consequently, additional phylogenetically informative loci should be included to increase the phylogenetic resolutions of Swertia. Based on our study, we proposed a set of 10 most divergent loci (trnH-GUG_psbA, ndhC_trnV-UAC, petB intron, trnL-UAA intron, rps16, rps16_trnK-UUU, trnS-UCC_trnG-UCC, trnS-GGA_rps4, psbM_trnD-GUC, ccsA_ndhD) that be helpful for future phylogenetic and barcoding studies of Swertia.

Phylogenetic analysis

Our plastome phylogenetic study did not produce any surprises topologically: the relationship inferred from the 74 common CDSs were consistent with clades resolved in previously published phylogenetic trees, Swertia was polyphyletic with its closely-related allies, namely Veratrilla, Halenia, Comastoma, Lomatogonium, and Lomatogoniopsis [4,5,6, 8, 9]. However, our study has greatly improved the phylogenetic resolution of the inferred clades in comparison with these earlier published trees, with most of the nodes having 100% support. The 20 Swertia species divided into four well-supported clades in the phylogenetic tree. Our molecular phylogeny study agreed with Ho TN taxonomic classification in sect. Swertopsis and sect. Platynema [1], besides, a wide range of incongruences between morphological and molecular data was observed (Fig. 8).

Clade I comprised five species: two species belonging to subgen. Poephila (sect. Poephila and sect. Macranthos), paralleling three species of sect. Platynema, Subgen. Ophilia. Clade I was the sister clade to Lomatogonipsis-Logmatogonium-Comastoma lineage in a larger clade.

Clade II consisted of six species of sect. Swertopsis, subgen. Ophilia, forming a well-supported monophyletic clade. However, S. leducii, too, belonging to the sect. Swertopsis, was separated from Clade II and positioned in a larger group consisted of Lomatogonipsis-Logmatogonium-Comastoma, Clade I, and Clade II.

Clade III comprised two Halenia species, S. dichotoma of sect. Japonicae, subgen. Ophelia, and S. tetraptera. These four species clustered together with a notably higher solutions in internodes (BS 98–100%, 1.00 PP) than in previously published trees [6, 8], supported their close relationship. The taxonomic treatment of S. tetraptera has been in constant debates: Grisebach and Ma segregated it from Swertia as genus Anagallidium [62, 63]. Ho, based on morphology, pollination mode, and molecular evidence, segregated this species from Swertia and elevated it to a separate genus as Sinoswertia, but was not used in the recent account of the genus by Flora of China [2, 64]. Our phylogenetic study supported S. tetraptera’ close relationship with S. dichotoma (sect. Japonicae). Furthermore, these two Swertia species are more closely related to Halenia than with any other Swertia species. The genus Halenia is distributed in Asian mountains but with its diversification in south America [4, 5, 64], yet we only included two QTP-distributed Halenia species in this study, further studies are needed to discriminate the relationship between Swertia and Halenia.

Lastly, clade IV consisted of two paralleled groups: in one group, five species belonging to two different sections (sect. Apterae and sect. Apterae) of subgen. Swertia clustered together, being sister with S. cordata, a species from a different subgenus (sect. Ophelia, Subgen. Ophilia).

Our topology inferred from phylogenetic analyses included two clades that contained remarkably short terminal branches: S. kouitchensis-S. punicea-S. fracnetiana-S. mussotii in clade II and S. przewalskii-S. bifolia-S. souliei-S. wolfganiana in clade IV. This result indicated the emergence of these Swertia species might have occurred only recently, as confirmed by their extremely small genetic distance values, and also supported by our previous divergence time analysis based on rbcL and matK loci: most of the extant Swertia species were evolved during the Pliocene to the late Pleistocene period (4–0.1 Ma) [6]. Recent rapid radiations were detected in numerous studies, and can be explained by climate oscillations in the Quaternary [65,66,67,68,69]. Recently diverged species often have incomplete reproductive barriers and subsequently experience hybridization and/or introgression, especially where their habitats overlap [70]. On-going hybridization and introgression that accompanied recent rapid radiation have been detected in numerous studies, both plants and animals [71,72,73,74,75]. This could be the same scenario in Swertia and its allies in our study. Rapid radiations are among the most notorious phylogenetic problems, because retention and stochastic sorting of ancestral polymorphisms can yield incongruence between taxonomic circumscription and molecular data [76], as we observed in the case of Swertia and its closely related taxa, as well as in other rapidly diverged species [71,72,73,74,75]. To uncover the real evolutionary history of Swertia, we will use multi-locus molecular sequences that have evolved independently and population genomic approaches in our future study.

Conclusion

This work was the first to report full sequence data and features of the whole plastomes of 11 Swertia species. The comparison of these plastomes revealed high similarities in terms of the overall structure, long repeat sequence, SSR, and codon usage. However, expansion and contraction in the IRb-LSC region were detected, due to the heterogeneous length of the ψycf1. Our phylogenetic analyses based on 74 CDSs supported the polyphyly of Swertia, was consistent with previous molecular studies, but with significant improved solutions. Furthermore, our phylogenetic study yielded extremely short branches, indicating a recent rapid radiation, hypothetically triggered by the Quaternary climate fluctuations. Our result inferred from the molecular data only partly supported the current taxonomic treatment. Furthermore, the cemA was identified under positive selection in most of the Swertia species, possibly reflecting an adaptive response to low CO2 atmospheric conditions during the Late Miocene. We proposed a set of 10 most variable loci (trnH-GUG_psbA, ndhC_trnV-UAC, petB intron, trnL-UAA intron, rps16, rps16_trnK-UUU, trnS-UCC_trnG-UCC, trnS-GGA_rps4, psbM_trnD-GUC, ccsA_ndhD) as markers for future phylogenetic and taxonomic studies of Swertia. Overall, our results demonstrated the power of plastome phylogenomic to improve phylogenetic resolution and contribute to a better understanding of plastid gene evolution in Swertia.

Methods

Sample collection, plastome sequencing, assembly, and annotation

We collected plant specimens in the Qing-Tibetan Plateau region (QTP) in China. Fresh leaves were collected in the field and dried in silica gel for the future DNA extraction. Total genomic DNA was extracted with the CTAB protocol [77]. Voucher specimens were deposited at the Herbarium of Northwest Institute of Plateau Biology, China Academy of Science (Xining, China) (Additional file 1, Table S1).

We sequenced the complete plastomes of the 11 Swertia species with an Illumina Hiseq 2500 sequencing system following instructions of DNA Nextera XT Sample Prep Kit (Illumina ™) at Novegene Co., Ltd. in Wuhan, China. Genomic DNA was fragmented randomly and then the required length of DNA fragments was obtained by electrophoresis. Adapters were ligated to DNA fragments followed by cluster preparation and sequencing. A paired-end library (150 × 2) was constructed with an insert size of 350 base pairs (bp), and then 150 bp paired reads were sequenced using the Illumina Hiseq 2500 sequencing system. The total number, length and quality of reads sequenced for each Swertia sample was analyzed with FastQC V. 0.11.9 [78]. Low-quality reads were filtered (> 50% of bases with a quality score ≤ 20 and ≥ 1% of missing bases) using SOAPnuke 2.1.0 [79]. High quality clean reads were assembled using NOVOPlasty v3.6 [80] with the default parameters. In order to validate the assembled plastome sequence error, the clean reads were mapped to the assembled plastomes using the Circlator V. 1.5.5 program [81]. The coverage depth of plastomes was calculated with the SAMtools V. 1.10 program [82]. Complete plastomes were annotated using the online program GeSeq [83] with two reference plastomes (Swertia mussotii, GenBank: KU641021 and Gentianopsis grandis, GenBank: NC_049879) using the default values to predict genes coding for proteins (CDS), transfer RNAs (tRNA), and ribosomal DNAs (rRNAs). The start and stop codon positions of the open reading frame for all the coding genes were manually checked and adjusted if necessary, with the Sequin program (NCBI). Organellar Genome DRAW [84] was employed to draw gene maps. The raw pair-end reads, the complete plastome sequence and gene annotation of the 11 newly assembled Swertia plastomes were submitted to the NCBI GenBank under BioProject No. PRJNA807092 (Additional file 1, Table S1).

Plastome features

For each plastome, the length and gene were identified; LSC, SSC, IRa, and IRb region were plotted with junction positions being compared; the Guanine-cytosine (GC) content was calculated for each region with Geneious R7 [85].

Codon usage

All CDSs for each Swertia plastome were extracted using Phylosuite v1.2.2 [86]. The amount of codon and relative synonymous codon usage (RSCU) ratio was calculated using Mega X [43]. The RSCU distribution was illustrated in the form of heatmaps using OriginPro2021 software (OriginLab Corporation, Northampton, MA, USA). The level of codon usage bias was determined by calculating ENC, GC1, GC2, GC3, and CAI with CodonW (http://codonw.sourceforge.net/). The ENC-GC3 (ENC vs. GC3) plot was carried out to examine whether the codon usage of a certain gene was affected by mutation or also by other factors such as natural selection. If the corresponding points were distributed around the expectation curve, it was possible to predict the mutation was the only factor affecting codon usage bias, as introduced by Wright [24] originally, and later improved by Liu [25]. The equation of the curve is as follows: pENC = 2.5—s + {29.5 / [s2 + (1 − s)2]}[24, 25]. The ENC vs. GC3 plot was generated by ggplot2 in R v.3.6.3 (https://www.r-project.org/).

Characterization of repeat sequences and simple sequence repeats (SSR)

Forward, reverse, palindromic, and complementary repeats were identified by REPuter online program with the default settings, Maximum Computed Repeats 50 and Minimal Repeat Size 8 [87]. Simple sequence repeats (SSRs) were exploited using the MISA-web program [88] with the following parameters: ≥ 10 repeat units for mononucleotide SSRs, ≥ 6 for dinucleotide, and ≥ 3 for dinucleotide, tetranucleotide, pentanucleotide, and hexanucleotide SSRs. All overlapping repeats were removed from the final results.

Hypervariable regions

For identifying hypervariable regions, the whole plastome alignment of the 20 Swertia plastomes was visualized using the mVISTA [89]. The CDSs, intergenic, and intron-containing regions of the 20 plastomes were extracted and aligned using Phylosuite v1.2.2 program [90]. Subsequently, the genetic divergence for all protein-coding and noncoding (intergenic spacers and introns) regions between the 20 Swertia plastomes was calculated based on nucleotide diversity value (Pi) using DnaSP v5.10.01 [91].

Phylogenetic analysis

Phylogenetic analyses of Swertia were performed using both maximum likelihood (ML) and Bayesian inference (BI) methods based on the 74 CDSs shared between the 20 plastomes of Swertia (include eight published species) and altogether 11 species in subtribe Swertiinae and four species in Gentianaceae, with Coffea arabica (Rubiaceae) as the outgroup (Additional file 2, Table S2) [9, 92,93,94,95,96,97,98,99]. The nucleotide sequences of the common 74 CDSs were extracted from each plastome (with pseudogenes excluded), concatenated, and aligned with Phylosuite v1.2.2 [90], using default settings. The best‐fit models of nucleotide substitution, the GTR + I + G were selected by jModeltest v.2.1.7 [100] with the Akaike Information Criterion (AIC) [101]. The ML analyses were performed with RAXML v.8.2.12 [102]. Search for the best ML tree (the “‐f a” option) was using the GTRGAMMAI substitution model (GTR + I + G) that was selected by jModeltest, then drawing support values using the rapid bootstrap (1000 replicates). The BI analysis was performed for 20 million generations in MrBayes v.3.2.6 [103], with two runs of two independent Markov chain Monte Carlo (MCMC) chains. Each chain started with a random tree, default priors, and sampling trees every 100 generations, and the first 25% generation was discarded as burn‐in. Moreover, the evolution pairwise distance matrices were constructed with MEGA X based on the whole plastome sequences and 74 CDSs among the 20 Swertia species, using the Kimura 2-parameter model [43, 104].

Selective pressure analysis

Selective pressures were analyzed for 50 common CDSs that are longer than 300 bp of 35 Gentianaceae species included in the phylogenetic analysis (see above). Firstly, the ratio (ω = dN/dS) of nonsynonymous (dN) to synonymous nucleotide (dS) substitution rates was calculated using the Codeml program in PAML4.9 with the site-specific model (seqtype = 1, model = 0, NSsites = 0, 1, 2, 3, 7, 8) [105, 106]. The codon frequencies were set by the F3 × 4 model. The likelihood ratio test (LRT) was used to identify positive selected sites in comparisons of M0 vs. M3, M1 vs. M2, M7 vs. M8. The Bayes Empirical Bayes (BEB) method was used to identify codons under positive selection. BEB values higher than 0.95 indicate sites that are potentially under positive selection. Furthermore, for CDSs that was detected under positive selection from the site-specific model, the branch model was used to detect signatures of positive selection along specific lineages by using the three models (one-ratio, free-ratio and two-ratio model). The one-ratio model (m0), assumes the same dN/dS ratio for all branches in the phylogeny. This model was compared to the free-ratio model (m1) that assumes an independent ω ratio for each branch. The one-ratio model (m0) was then compared to the two-ratio model (m2) that assumes the Swertia clades (set as the foreground branch) have ω ratios different from the other branches (set as the background branch). Similarly, the LRT was used to identify positive selected branches in comparisons of m0 vs. m1and m0 vs. m2 [105, 106].

Availability of data and materials

The assembled sequences described in this study have been deposited in the National Center for Biotechnology and Information (NCBI) under the accessions as summarized in Additional File 1, Table S1.

Abbreviations

BEB:

Bayes empirical bayes

BI:

Bayesian inference

bp:

Base pair

BS:

Branch support

CDS:

Protein-coding sequence

CTAB:

Cetyl trimethyl ammonium bromide

GC:

Guanine-cytosine

IRs:

Inverted repeats

ITS:

Internal transcribed spacer

LRT:

Likelihood ratio test

LSC:

Large single copy

MCMC:

Markov chain Monte Carlo

ML:

Maximum likelihood

PP:

Posterior probability

rRNA:

Ribosomal RNA

RSCU:

Relative synonymous codon usage

SSC:

Small single copy

SSR:

Simple sequence repeat

tRNA:

Transfer RNA

References

  1. Ho TN, Liu SW. Swertia Linnaeus. In: Wang J, Luo J, Bu X, editors. A worldwide of Swertia and its allies. Beijing: Science Press; 2015. 85–6.

    Google Scholar 

  2. Ho TN, Barthlomew B, Gilbert M. Swertia. Flora of China, vol. 62. Beijing: Science Press; 1998. p. 344.

    Google Scholar 

  3. Struwe L, Kadereit JW, Klackenberg J, Nilsson S, Thiv M, Von Hagen KB, et al. Systematics, character evolution, and biogeography of Gentianaceae, including a new tribal and subtribal classification. In: Struve L, Albert VA, editors., et al., Gentianaceae : systematics and natural history. Cambridge: Cambridge University Press; 2002. 21–309.

    Google Scholar 

  4. Chassot P, Nemomissa S, Yuan YM, Küpfer P. High paraphyly of Swertia L. (Gentianaceae) in the Gentianella-lineage as revealed by nuclear and chloroplast DNA sequence variation. Plant Syst Evol. 2001;229(1):1–21. https://doi.org/10.1007/s006060170015.

  5. Hagen KB, Kadereit JW. Phylogeny and flower evolution of the Swertiinae (Gentianaceae-Gentianeae): homoplasy and the principle of variable proportions. Syst Bot. 2002;27(3):548–72.

    Google Scholar 

  6. Cao Q, Xu LH, Wang JL, Zhang FQ, Chen SL. Molecular phylogeny of subtribe Swertiinae. BBR. 2021;41(03):408–18.

    Google Scholar 

  7. Sun SS, Fu PC. Study on taxonomy and evolution of Gentianeae (Gentianaceae). Acta Bot Sin. 2019;39(02):363–70.

    Google Scholar 

  8. Xi HC, Sun Y, Xue CY. Molecular phylogeny of Swertiinae (Gentianaceae Gentianeae) based on sequence data of ITS and matK. Plant Divers. 2014;36(002):145–56.

    CAS  Google Scholar 

  9. Zhang X, Sun Y, Landis JB, Lv Z, Shen J, Zhang H, et al. Plastome phylogenomic study of Gentianeae (Gentianaceae): widespread gene tree discordance and its association with evolutionary rate heterogeneity of plastid genes. BMC Plant Biol. 2020;20(1):340. https://doi.org/10.1186/s12870-020-02518-w.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Brahmachari G, Mondal S, Gangopadhyay A, Gorai D, Mukhopadhyay B, Saha S, et al. Swertia (Gentianaceae): chemical and pharmacological aspects. Chem Biodivers. 2004;1(11):1627–1651. https://doi.org/10.1002/cbdv.200490123.

  11. Singh M, Ram M, Mishra Si, Goyal N, Shanker K, Gupta N, et al. Ultra performance liquid chromatography coupled with principal component and cluster analysis of Swertia chirayita for adulteration check. J Pharm Biomed Anal. 2019;164:302–308. https://doi.org/10.1016/j.jpba.2018.10.054.

  12. Kakiuchi N, Iwaki N, Mikage M, Xiao H, Wang Z, Hattori M. Phylogenetic examination of crude drugs derived from Yunnanese Swertia plants. J Nat Med. 2014;68(1):206–10. https://doi.org/10.1007/s11418-013-0776-x.

    Article  PubMed  Google Scholar 

  13. Li J, Zhao YL, Huang HY, Wang YZ. Phytochemistry and pharmacological activities of the genus Swertia (Gentianaceae): a Review. Am J Chin Med. 2017;45(4):667–736. https://doi.org/10.1142/s0192415x17500380.

    Article  CAS  PubMed  Google Scholar 

  14. Liu C, Zhang YX, Liu Y, Chen YL, Fan G, Xiang L, et al. Identification of Tibetan medicine “Dida” of Gentianaceae using DNA barcoding. China J Chin Materia Med. 2016;41(04):567–71.

    CAS  Google Scholar 

  15. Jansen RK, Ruhlman TA. Plastid Genomes of Seed Plants. Genomics of Chloroplasts and Mitochondria. Dordrecht: Springer Netherlands; 2012. 103–126.

  16. Raven JA, Allen JF. Genomics and chloroplast evolution: what did cyanobacteria do for plants? Genome Biol. 2003;4(3):209. https://doi.org/10.1186/gb-2003-4-3-209.

    Article  PubMed  PubMed Central  Google Scholar 

  17. Ravi V, Khurana JP, Tyagi AK, Khurana P. An update on chloroplast genomes. Syst Evol. 2008;271(1):101–22. https://doi.org/10.1007/s00606-007-0608-0.

    Article  CAS  Google Scholar 

  18. Parks M, Cronn R, Liston A. Increasing phylogenetic resolution at low taxonomic levels using massively parallel sequencing of chloroplast genomes. BMC Biol. 2009;7(1):84. https://doi.org/10.1186/1741-7007-7-84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Liu LM, Du XY, Guo C, Li DZ. Resolving robust phylogenetic relationships of core Brassicaceae using genome skimming data. J Syst Evol. 2020;00(0):1–12. https://doi.org/10.1111/jse.12666.

  20. Barrett CF, Specht CD, Leebens MJ, Stevenson DW, Zomlefer WB, Davis JI. Resolving ancient radiations: can complete plastid gene sets elucidate deep relationships among the tropical gingers (Zingiberales)? Ann Bot. 2014;113(1):119–33. https://doi.org/10.1093/aob/mct264.

    Article  PubMed  Google Scholar 

  21. Ma PF, Zhang YX, Zeng CX, Guo ZH, Li DZ. Chloroplast phylogenomic analyses resolve deep-level relationships of an intractable bamboo tribe Arundinarieae (poaceae). Syst Biol. 2014;63(6):933–50. https://doi.org/10.1093/sysbio/syu054.

    Article  PubMed  Google Scholar 

  22. Li HT, Yi TS, Gao LM, Ma PF, Zhang T, Yang JB, et al. Origin of angiosperms and the puzzle of the Jurassic gap. Nat Plants. 2019;5(5):461–70. https://doi.org/10.1038/s41477-019-0421-0.

    Article  PubMed  Google Scholar 

  23. Yu XQ, Gao LM, Soltis DE, Soltis PS, Yang JB, Fang L, et al. Insights into the historical assembly of East Asian subtropical evergreen broadleaved forests revealed by the temporal history of the tea family. New Phytol. 2017;215(3):1235–1248. https://doi.org/10.1111/nph.14683.

  24. Wright F. The ‘effective number of codons’ used in a gene. Gene. 1990;87(1):23–29. https://doi.org/10.1016/0378-1119(90)90491-9.

  25. Liu XE. A more accurate relationship between “effective number of codons” and GC3s under assumptions of no selection. Comput Biol Chem. 2013;42:35–9. https://doi.org/10.1016/j.compbiolchem.2012.11.003.

    Article  CAS  PubMed  Google Scholar 

  26. Dong BR, Zhao ZL, Ni LH, Wu XR, Danzhenhhuoga. Comparative analysis of complete chloroplast genome sequences within Gentianaceae and significance of identifying species. CHM. 2020;51(6):1641–9. https://doi.org/10.7501/j.issn.0253-2670.2020.06.033

  27. Zhang X, Auid O, Sun Y, Landis JB, Lv Z, S J., et al. Plastome phylogenomic study of Gentianeae (Gentianaceae): widespread gene tree discordance and its association with evolutionary rate heterogeneity of plastid genes. BMC Plant Biol. 2020;20:340. https://doi.org/10.1186/s12870-020-02518-w.

  28. Wang JL, Cao Q, Wang KZ, Xing R, Wang L, Zhou DW. Characterization of the complete chloroplast genome of Pterygocalyx volubilis (Gentianaceae). Mitochondrial DNA B. 2019;4(2):2579–2580. https://doi.org/10.1080/23802359.2019.1640644.

  29. Yan JY, Cao Q, Wu Z, Chen SF, Wang JL, Zhou DW, et al. Complete plastome sequence of Eustoma grandiflorum (Gentianaceae), a popular cut flower. Mitochondrial DNA B. 2019;4(2):3163–3164. https://doi.org/10.1080/23802359.2019.1667893.

  30. Ni LH, Zhao ZL, Xu HX, Chen SL, Dorje G. Chloroplast genome structures in Gentiana (Gentianaceae), based on three medicinal alpine plants used in Tibetan herbal medicine. Curr Genet. 2017;63(2):241–52. https://doi.org/10.1007/s00294-016-0631-1.

    Article  CAS  PubMed  Google Scholar 

  31. Ueda M, Nishikawa T, Fujimoto M, Takanashi H, Arimura S, Tsutsumi N, et al. Substitution of the gene for chloroplast RPS16 was assisted by generation of a dual targeting signal. Mol Biol Evol. 2008;25(8):1566–75. https://doi.org/10.1093/molbev/msn102.

    Article  CAS  PubMed  Google Scholar 

  32. Roy S, Ueda M, Kadowaki K, Tsutsumi N. Different status of the gene for ribosomal protein S16 in the chloroplast genome during evolution of the genus Arabidopsis and closely related species. Genes Genet Syst. 2010;85(5):319–26. https://doi.org/10.1266/ggs.85.319.

    Article  CAS  PubMed  Google Scholar 

  33. Schwarz E, Ruhlman T, Sabir J, Hajarah N, Alharbi N, Al Malki A, et al. Plastid genome sequences of legumes reveal parallel inversions and multiple losses of rps16 in papilionoids. J Syst Evol. 2015;53. https://doi.org/10.1111/jse.12179.

  34. Sun SS, Fu PC, Zhou XJ, Cheng YW, Zhang FQ, Chen SL, et al. The complete plastome sequences of seven species in Gentiana sect. Kudoa (Gentianaceae): insights into plastid gene loss and molecular evolution. Front Plant Sci. 2018;9(493). https://doi.org/10.3389/fpls.2018.00493.

  35. Millen RS, Olmstead RG, Adams KL, Palmer JD, Lao NT, Heggie L, et al. Many parallel losses of infA from chloroplast DNA during angiosperm evolution with multiple independent transfers to the nucleus. Plant Cell. 2001;13(3):645–58. https://doi.org/10.1105/tpc.13.3.645.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Pfitzinger H, Guillemaut P, Weil JH, Pillay DT. Adjustment of the tRNA population to the codon usage in chloroplasts. NAR. 1987;15(4):1377–86. https://doi.org/10.1093/nar/15.4.1377.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Liu Q, Xue Q. Comparative studies on codon usage pattern of chloroplasts and their host nuclear genes in four plant species. J Genet. 2005;84(1):55–62. https://doi.org/10.1007/bf02715890.

    Article  CAS  PubMed  Google Scholar 

  38. Ren T, Li ZX, Xie DF, Gui LJ, Peng C, Wen J, et al. Plastomes of eight ligusticum species: characterization, genome evolution, and phylogenetic relationships. BMC Plant Biol. 2020;20(1):519. https://doi.org/10.1186/s12870-020-02696-7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Sablok G, Nayak KC, Vazquez F, Tatarinova TV. Synonymous codon usage, GC3, and evolutionary patterns across plastomes of three pooid model species: emerging grass genome models for monocots. Mol Biotechnol. 2011;49(2):116–28. https://doi.org/10.1007/s12033-011-9383-9.

    Article  CAS  PubMed  Google Scholar 

  40. Chi X, Zhang F, Dong Q, Chen S. Insights into comparative genomics, codon usage bias, and phylogenetic relationship of species from biebersteiniaceae and nitrariaceae based on complete chloroplast genomes. Plants (Basel, Switzerland). 2020;9(11). https://doi.org/10.3390/plants9111605.

  41. Ogihara Y, Terachi T, Sasakuma T. Intramolecular recombination of chloroplast genome mediated by short direct-repeat sequences in wheat species. Proc Natl Acad Sci USA. 1988;85(22):8573–7. https://doi.org/10.1073/pnas.85.22.8573.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Weng ML, Blazier JC, Govindu M, Jansen RK. Reconstruction of the ancestral plastid genome in Geraniaceae reveals a correlation between genome rearrangements, repeats, and nucleotide substitution rates. Mol Biol Evol. 2014;31(3):645–59. https://doi.org/10.1093/molbev/mst257.

    Article  CAS  PubMed  Google Scholar 

  43. Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol. 2018;35(6):1547–9. https://doi.org/10.1093/molbev/msy096.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Qin Z, Wang YP, Wang QM, Li A, Hou FY, Zhang L. Evolution analysis of simple sequence repeats in plant genome. PLoS One. 2015;10(12):e0144108. https://doi.org/10.1371/journal.pone.0144108.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Bi Y, Zhang MF, Xue J, Dong R, Du YP, Zhang XH. Chloroplast genomic resources for phylogeny and DNA barcoding: a case study on Fritillaria. Sci Rep. 2018;8(1):1184. https://doi.org/10.1038/s41598-018-19591-9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Ellis JR, Burke JM. EST-SSRs as a resource for population genetic analyses. Heredity. 2007;99(2):125–32. https://doi.org/10.1038/sj.hdy.6801001.

    Article  CAS  PubMed  Google Scholar 

  47. Sasaki Y, Sekiguchi K, Nagano Y, Matsuno R. Chloroplast envelope protein encoded by chloroplast genome. FEBS Letters. 1993;316(1):93–8. https://doi.org/10.1016/0014-5793(93)81743-J.

    Article  CAS  PubMed  Google Scholar 

  48. Sonoda M, Katoh H, Ohkawa H, Ogawa T, Structure and function of cemA homologue (PXCA) in cyanobacteria. The chloroplast: from molecular biology to biotechnology. Dordrecht: Springer; 1997. p. 99–105.

    Google Scholar 

  49. Katoh A, Lee KS, Fukuzawa H, Ohyama K, Ogawa T. cemA homologue essential to CO2 transport in the cyanobacterium Synechocystis PCC6803. Proc Natl Acad Sci USA. 1996;93(9):4006–10. https://doi.org/10.1073/pnas.93.9.4006.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Favre A, Ingo M, Chen CH, Wang JC, Pringle James S, Matuszak Sabine, et al. Out-of-Tibet: the spatio-temporal evolution of Gentiana (Gentianaceae). J Biogeogr. 2016;43(10):1967–1978. https://doi.org/10.1111/jbi.12840.

  51. Yu H, Deane DC, Sui X, Fang S, He F. Testing multiple hypotheses for the high endemic plant diversity of the Tibetan Plateau. Glob Ecol Biogeogr. 2019;28:131–44.

    Article  Google Scholar 

  52. Zachos J, Pagani M, Sloan L, Thomas E, Billups K. Trends, rhythms, and aberrations in global climate 65 Ma to present. Science. 2001;292:686–93. https://doi.org/10.1126/science.1059412.

    Article  CAS  PubMed  Google Scholar 

  53. Zachos JC, Dickens GR, Zeebe RE. An early Cenozoic perspective on greenhouse warming and carbon-cycle dynamics. Nature. 2008;451(7176):279–83. https://doi.org/10.1038/nature06588.

    Article  CAS  PubMed  Google Scholar 

  54. Herbert TD, Lawrence KT, Tzanova A, Peterson LC, CaballeroGill R, Kelly CS. Late Miocene global cooling and the rise of modern ecosystems. Nat Geosci. 2016;9(11):843–7. https://doi.org/10.1038/ngeo2813.

    Article  CAS  Google Scholar 

  55. Cerling TE, Wang Y, Quade J. Expansion of C4 ecosystems as an indicator of global ecological change in the late Miocene. Nature. 1993;361(6410):344–5. https://doi.org/10.1038/361344a0.

    Article  Google Scholar 

  56. Pagani M. Late Miocene atmospheric CO2 concentrations and the expansion of C4 grasses. Science. 1999;285(5429):876–9.

    Article  CAS  Google Scholar 

  57. Herbert TD, Lawrence KT, Tzanova A, Peterson LC, Caballero Gill R, Kelly CS. Late Miocene global cooling and the rise of modern ecosystems. Nat Geosci. 2016;(9):843–7. https://doi.org/10.1038/ngeo2813.

  58. Pound MJ, Haywood AM, Salzmann U, Riding JB. Global vegetation dynamics and latitudinal temperature gradients during the mid to late Miocene (15.97–5.33Ma). Earth Sci Rev. 2012;112(1):1–22. https://doi.org/10.1016/j.earscirev.2012.02.005.

    Article  Google Scholar 

  59. Li X, Yang Y, Henry RJ, Rossetto M, Wang Y, Chen S. Plant DNA barcoding: from gene to genome. Biol Rev. 2015;90(1):157–66. https://doi.org/10.1111/brv.12104.

    Article  PubMed  Google Scholar 

  60. Kress WJ. Plant DNA barcodes: applications today and in the future. J Syst Evol. 2017;55(4):291–307. https://doi.org/10.1111/jse.12254.

    Article  Google Scholar 

  61. Shaw J, Shafer HL, Leonard OR, Kovach MJ, Schorr M, Morris AB. Chloroplast DNA sequence utility for the lowest phylogenetic and phylogeographic inferences in angiosperms: the tortoise and the hare IV. Am J Bot. 2014;101(11):1987–2004. https://doi.org/10.3732/ajb.1400398.

    Article  PubMed  Google Scholar 

  62. Grisebach A. Genera et species Gentianearum adjectis observationibus quibusdam phytogeographicis. Stuttgartiae: J.G. Cotta; 1839.

    Book  Google Scholar 

  63. Marquand CB. The botanical collection made by captain F. Kingdon Ward in the Eastern Himalaya and Tibet in 1924–25. Biol J Linn Soc Lond. 1929;48(321):149–229. https://doi.org/10.1111/j.1095-8339.1929.tb00587.x.

    Article  Google Scholar 

  64. Ho TN, Liu SW, Liu JQ. A new Qinghai-Tibet Plateau endemic genus Sinoswertia and its pollination mode. Plant Divers. 2013;35(3):7.

    Google Scholar 

  65. Hewitt G. The genetic legacy of the quaternary ice ages. Nature. 2000;405(6789):907–13. https://doi.org/10.1038/35016000.

    Article  CAS  PubMed  Google Scholar 

  66. Yu HB, Zhang YL. Advances in phylogeography of alpine plants in the Tibetan plateau and adjacent regions. Acta Bot Sin. 2013;33(06):1268–78.

    Google Scholar 

  67. Ding WN, Ree RH, Spicer RA, Xing YW. Ancient orogenic and monsoon-driven assembly of the world’s richest temperate alpine flora. Science. 2020;369(6503):578–81. https://doi.org/10.1126/science.abb4484.

    Article  CAS  PubMed  Google Scholar 

  68. Gao QB, Zhang DJ, Duan YZ, Zhang FQ, Li YH, Fu PC, et al. Intraspecific divergences of Rhodiola alsia (Crassulaceae) based on plastid DNA and internal transcribed spacer fragments. Bot J Linn Soc. 2012;168(2):204–15. https://doi.org/10.1111/j.1095-8339.2011.01193.x.

    Article  Google Scholar 

  69. Zhang FQ, Gao QB, Zhang DJ, Duan YZ, Li YH, Fu PC, et al. Phylogeography of spiraea alpina (Rosaceae) in the Qinghai-Tibetan plateau inferred from chloroplast DNA sequence variations. J Syst Evol. 2012;50(4):276–83. https://doi.org/10.1111/j.1759-6831.2012.00194.x.

    Article  Google Scholar 

  70. Liu JQ. The integrative species concept”and“pecies on the speciation way. Biodiversity Sci. 2016;24(09):1004–8.

    Article  Google Scholar 

  71. Fontaine MC, Pease JB, Steele A, et al. Extensive introgression in a malaria vector species complex revealed by phylogenomics. Science. 2015;347(621702):1258524.

    Article  Google Scholar 

  72. C. G, Ma PF, Yang GQ, Ye XY, Guo Y, Liu JX, et al. Parallel ddRAD and genome skimming analyses reveal a radiative and reticulate evolutionary history of the temperate bamboos. Syst Biol. 2020. https://doi.org/10.1093/sysbio/syaa076.

  73. Edelman NB, Frandsen PB, Miyagi M, Clavijo B, Davey J, Dikow RB, et al. Genomic architecture and introgression shape a butterfly radiation. Science. 2019;366(6465):594–9. https://doi.org/10.1126/science.aaw2090.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Arnold ML, Buckner CM, Robinson JJ. Pollen-mediated introgression and hybrid speciation in Louisiana irises. PNAS. 1991;88(4):1398–402. https://doi.org/10.1073/pnas.88.4.1398.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Zhang R, Gong X, Folk R. Evidence for continual hybridization rather than hybrid speciation between Ligularia duciformis and L. paradoxa (Asteraceae). PeerJ. 2017;5:e3884. https://doi.org/10.7717/peerj.3884.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Pamilo P, M Nei. Relationships between gene trees and species trees. Mol Biol Evol. 1988;5(5):568–583. https://doi.org/10.1093/oxfordjournals.molbev.a040517.

  77. Doyle JJ, Doyle JL. A rapid DNA isolation procedure for small quantities of fresh leaf tissue. Chem Pharm Bull. 1987;19(1):11–15.

  78. Simon A, FastQC. A quality control tool for high throughput sequence data. 2010. http://www.bioinformatics.babraham.ac.uk/projects/fastqc.

  79. Chen YX, Chen YS, Shi CM, Huang ZB, Zhang Y, Li SK, et al. SOAPnuke: a MapReduce acceleration-supported software for integrated quality control and preprocessing of high-throughput sequencing data. GigaScience. 2018;7(1). https://doi.org/10.1093/gigascience/gix120.

  80. Dierckxsens N, Mardulyn P, Smits G. NOVOPlasty: de novo assembly of organelle genomes from whole genome data. Nucleic Acids Res. 2017;45(4):18. https://doi.org/10.1093/nar/gkw955.

    Article  CAS  Google Scholar 

  81. Hunt M, Silva ND, Otto TD, Parkhill J, Keane JA, Harris SR. Circlator: automated circularization of genome assemblies using long sequencing reads. Genome Biol. 2015;16:294–294. https://doi.org/10.1186/s13059-015-0849-0.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Petr D, James KB, Jennifer L, John M, Valeriu O, Martin OP, et al. Twelve years of SAMtools and BCFtools GigaScience. 2021;10(2). https://doi.org/10.1093/gigascience/giab008.

  83. Tillich M, Lehwark P, Pellizzer T, Ulbricht-Jones ES, Fischer A, Bock R, et al. GeSeq – versatile and accurate annotation of organelle genomes. NAR. 2017;45(W1):W6–11. https://doi.org/10.1093/nar/gkx391.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Greiner S, Lehwark P, Bock R. OrganellarGenomeDRAW (OGDRAW) version 1.3.1: expanded toolkit for the graphical visualization of organellar genomes. NAR. 2019;47(W1):W59-W64. https://doi.org/10.1093/nar/gkz238.

  85. Kearse M, Moir R, Wilson A. Geneious Basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics. 2012;28(12):2. https://doi.org/10.1093/bioinformatics/bts199.

    Article  Google Scholar 

  86. Zhang D, Gao FL, Jakovlić I, Zou H, Zhang J, Li WX, et al. PhyloSuite: an integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Mol Ecol Resour. 2020;20(1):348–55. https://doi.org/10.1111/1755-0998.13096.

    Article  PubMed  Google Scholar 

  87. Kurtz S, Choudhuri JV, Ohlebusch E, Schleiermacher C, Stoye J, Giegerich R. REPuter: the manifold applications of repeat analysis on a genomic scale. NAR. 2001;29(22):4633–42. https://doi.org/10.1093/nar/29.22.4633.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Beier S, Thiel T, Münch T, Scholz U, Mascher M. MISA-web: a web server for microsatellite prediction. Bioinformatics. 2017;33(16):2583–5. https://doi.org/10.1093/bioinformatics/btx198.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Frazer KA, Pachter L, Poliakov A, Rubin EM, Dubchak I. VISTA: computational tools for comparative genomics. NAR. 2004;32(Web Server issue):W273–279. https://doi.org/10.1093/nar/gkh458.

  90. Zhang D, Gao FL, Jakovlić I, Zou H, Zhang J, Li WX, et al. PhyloSuite: an integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Mol Ecol Resour. 2020;20(1):348–55. https://doi.org/10.1111/1755-0998.13096.

    Article  PubMed  Google Scholar 

  91. Rozas J, Sánchez DelBarrio JC, Messeguer X, Rozas R. DnaSP, DNA polymorphism analyses by the coalescent and other methods. Bioinformatics. 2003;19(18):2496–7. https://doi.org/10.1093/bioinformatics/btg359.

    Article  CAS  PubMed  Google Scholar 

  92. Samson N, Bausher MG, Lee SB, Jansen RK, Daniell H. The complete nucleotide sequence of the coffee (Coffea arabica L.) chloroplast genome: organization and implications for biotechnology and phylogenetic relationships amongst angiosperms. Plant Biotechnol J. 2007;5(2):339–353. https://doi.org/10.1111/j.1467-7652.2007.00245.x.

  93. Zhang Yu, Xia MZ, Khan G, Yu JY, Chen SL, Zhang FQ. The complete chloroplast genome of Comastoma falcatum (Gentianaceae). Mitochondrial DNA B. 2019;4(1):987–988. https://doi.org/10.1080/23802359.2019.1581110.

  94. Park J, Yun N, Sang H. The complete chloroplast genome of an endangered species in Korea, Halenia corniculata (L.) Cornaz (Gentianaceae). Mitochondrial DNA B. 2019;4(1):1539–1540. https://doi.org/10.1080/23802359.2019.1601532.

  95. Wang JL, Cao Q, He CC, Ma YL, Li YL, Liu JX, et al. Complete chloroplast genome of Exacum affine (Gentianaceae): the first plastome of the tribe Exaceae in the family Gentianaceae. Mitochondrial DNA B. 2019;4(2):3529–3530. https://doi.org/10.1080/23802359.2019.1676672.

  96. Wang JL, Cao Q, Wang KZ, Xing R, Wang LR, Zhou DW. Characterization of the complete chloroplast genome of pterygocalyx volubilis (Gentianaceae). Mitochondrial DNA B. 2019;4(2):2579–80. https://doi.org/10.1080/23802359.2019.1640644.

    Article  Google Scholar 

  97. Ling LZ. Characterization of the complete chloroplast genome of Gentiana rhodantha (Gentianaceae). Mitochondrial DNA B. 2020;5:902–3. https://doi.org/10.1080/23802359.2020.1718026.

    Article  Google Scholar 

  98. Huang CX, Liu ML, Zhang HJ, Chang L, Wang YC, Yan JX. The complete nucleotide sequence of chloroplast genome of Gentiana apiata (Gentianaceae), an endemic medicinal herb in China. Mitochondrial DNA B. 2019;4(2):2596–7. https://doi.org/10.1080/23802359.2019.1641442.

    Article  Google Scholar 

  99. Huang L, Chen C, Yang Q, Yang W HQ. The complete chloroplast genome of Swertia cordata. Mitochondrial DNA B. 2019;4(2):3818–9.

    Article  Google Scholar 

  100. Darriba D, Taboada GL, Doallo R, Posada D. jModelTest 2: more models, new heuristics and parallel computing. Nat Methods. 2012;9(8):772–772. https://doi.org/10.1038/nmeth.2109.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Posada D, Buckley TR. Model selection and model averaging in phylogenetics: advantages of akaike information criterion and bayesian approaches over likelihood ratio tests. Syst Biol. 2004;53(5):793–808. https://doi.org/10.1080/10635150490522304.

    Article  PubMed  Google Scholar 

  102. Stamatakis A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics. 2014;30(9):1312–3. https://doi.org/10.1093/bioinformatics/btu033.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Ronquist F, Teslenko M, van der Mark P, Ayres DL, Darling A, Höhna S, et al. MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Syst Biol. 2012;61(3):539–542. https://doi.org/10.1093/sysbio/sys029.

  104. Tamura K, Nei M, Kumar S. Prospects for inferring very large phylogenies by using the neighbor-joining method. PNAS. 2004;101:11030–5. https://doi.org/10.1073/pnas.0404206101.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Yang Z. PAML 4: phylogenetic analysis by maximum likelihood. Mol Biol Evol. 2007;24(8):1586–91. https://doi.org/10.1093/molbev/msm088.

    Article  CAS  PubMed  Google Scholar 

  106. Yang Z, Nielsen R. Codon-substitution models for detecting molecular adaptation at individual sites along specific lineages. Mol Biol Evol. 2002;19(6):908–17. https://doi.org/10.1093/oxfordjournals.molbev.a004148.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank Wang Jiuli and Jia Liukun for providing materials. We thank Wang Zhihua, Yu Jingya and Xia Mingze for sharing their bioinformatic programing skills with us. We thank Zhou Yubi, Xu Bo, and Yang Fengmao for sharing photos of Swertia.

Funding

This study was supported by the Second Tibetan Plateau Scientific Expedition and Research (STEP) program (2019QZKK05020102), the Strategic Priority Research Program of Chinese Academy of Sciences (XDA2005020405), the Applied Basic Research Project of Qinghai Province (2019-ZJ-7019), and the Qinghai High-End Innovative Talents Thousands of People Plan which provided general funding for facilities and salaries. These entities had no role in the design sample collection, analysis, and interpretation of data nor in writing the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

C-SL designed the work. CQ, G-QB, M-XL, Z-FQ, XR and C-XF performed the experiments, analyzed the sequence data, and wrote the manuscript. All authors have read and approved the manuscript.

Corresponding author

Correspondence to Shilong Chen.

Ethics declarations

Ethics approval and consent to participate

All the plant materials were sampled from natural populations in the Qing-Tibetan Plateau region of China and no specific permission was needed to collect such samples. This study was conducted in accordance with local legislation and the Convention on the Trade in Endangered Species of Wild Fauna and Flora. Shilong Chen identified specimens of S. bifolia, S. franchetiana, and S. wolfangiana. Qingbo Gao identified S. diluta and S. przewalskii. Qian Cao identified S. dichotoma, S. cincta, S. kouitchensis, S. macrosperma, S. punicea, and S pubescens.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

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

Supplementary Information

Additional file 1:

Table S1. Basic information of 11 newly sequenced Swertia plastomes.

Additional file 2:

Table S2. Information of species included in phylogenetic analyses.

Additional file 3:

Table S3. RSCU values of protein-coding genes of the 20 Swertia plastomes.

Additional file 4:

Table S4. Codon feature of the 20 Swertia plastomes.

Additional file 5:

Table S5. The repeat sequence distribution in the 20 Swertia plastomes.

Additional file 6:

Table S6. Information of the pairwise sequence divergence (Pi) of the 20 Swertia plastomes.

Additional file 7:

Table S7. Results of selective pressure analysis.

Additional file 8:

Table S8. Pairwise genetic distance of the 20 Swertia plastomes.

Additional file 9:

Figure S1. Coverage depth of 11 newly sequenced Swertia plastomes.

Additional file 10:

Figure S2. The ω ratio of cemA along the phylogenetic tree of Gentianaceae.

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

Cao, Q., Gao, Q., Ma, X. et al. Plastome structure, phylogenomics and evolution of plastid genes in Swertia (Gentianaceae) in the Qing-Tibetan Plateau. BMC Plant Biol 22, 195 (2022). https://doi.org/10.1186/s12870-022-03577-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12870-022-03577-x

Keywords