Skip to main content

Effects of geographic isolation on the Bulbophyllum chloroplast genomes

Abstract

Background

Because chloroplast (cp) genome has more conserved structures than nuclear genome and mitochondrial genome, it is a useful tool in estimating the phylogenetic relationships of plants. With a series of researches for cp genomes, there have been comprehensive understandings about the cp genome features. The genus Bulbophyllum widely distributed in Asia, South America, Australia and other places. Therefore, it is an excellent type genus for studying the effects of geographic isolation.

Results

In this study, the cp genomes of nine Bulbophyllum orchids were newly sequenced and assembled using the next-generation sequencing technology. Based on 19 Asian (AN) and eight South American (SA) Bulbophyllum orchids, the cp genome features of AN clade and SA clade were compared. Comparative analysis showed that there were considerable differences in overall cp genome features between two clades in three aspects, including basic cp genome features, SSC/IRB junctions (JSBs) and mutational hotspots. The phylogenetic analysis and divergence time estimation results showed that the AN clade has diverged from the SA clade in the late Oligocene (21.50–30.12 mya). After estimating the occurrence rates of the insertions and deletions (InDels), we found that the change trends of cp genome structures between two clades were different under geographic isolation. Finally, we compared selective pressures on cp genes and found that long-term geographic isolation made AN and SA Bulbophyllum cp genes evolved variably.

Conclusion

The results revealed that the overall structural characteristics of Bulbophyllum cp genomes diverged during the long-term geographic isolation, and the crassulacean acid metabolism (CAM) pathway may play an important role in the Bulbophyllum species evolution.

Peer Review reports

Background

Chloroplasts (cp) are semi-autonomous organelles with their own unique genomes. The cp genomes of terrestrial plants are relatively conservative, with sizes ranging from 120 to 170 kbp, and can be divided into four regions: large single copy (LSC) region, small single copy (SSC) region and two inverted repeat (IRB and IRA) regions [1, 2]. Compared with nuclear genome and mitochondrial genome, cp genome is characterized by conserved structures, which made it a useful tool in estimating the phylogenetic relationships of plants [3, 4]. Although the cp gene content and composition are highly conserved, e.g., containing about 120 genes, mainly consisting of three categories: self-replication genes, photosynthesis genes, and RNA genes [5], the cp genomes have undergone structural changes in the long evolutionary process [6], such as IR contraction [7], inversion [8, 9], loss of genes and introns [10].

In recent years, as the sequencing costs continue to decrease, more and more complete cp genomes were sequenced and assembled, which is helpful for the plant phylogenetic studies. Orchids have become the focus of phylogenetic studies due to rich species, wide distributions and epiphytic habits. Based on 124 orchid cp genomes, Kim et al. (2020) cleared out the relationships of five subfamilies within Orchidaceae (Apostasioideae, Vanilloideae, Cypripedioideae, Orchidoideae and Epidendroideae) [4]. Moreover, there have been comprehensive understandings about the orchid cp genome features. For example, (i) cp genome substitution rates are considerably lower in IR regions than in SC regions [11, 12]; (ii) the repeat sequences played an important role in cp genome rearrangements [13,14,15]; (iii) mutational hotspots are genus specific [16]. However, the effects of geographic isolation on the cp genomes have not been investigated.

The genus Bulbophyllum Thou. belongs to the family Orchidaceae Juss., mainly distributed in Asia, South America, Australia and other places [17]. It is one of the largest genera of flowering plants, comprising about 2,000 species, which are predominantly restricted to rainforest habitats [18]. And it is characterized by epiphytic growth, CAM metabolism and fleshy flowers. There are about 105 species in China, which are widely distributed in Guizhou, Guangxi, Yunnan and Fujian provinces [19]. The wide distributions make the genus Bulbophyllum an excellent mode for studying the effects of geographic isolation on the cp genomes. There have been previous studies on the Bulbophyllum phylogenetic relationships. Gamisch et al. (2015) used internal transcribed spacer (ITS) sequences and four short cp genome fragments to study the phylogenetic relationships and estimate divergence times of Madagascar Bulbophyllum [20]. Gamisch et al. (2019) analyzed the evolutionary relationships of Bulbophyllum between different continents with ITS sequences, showing that Asian (AN) clade first separated from other Bulbophyllum orchids, and then South American (SA) clade separated from African Bulbophyllum [18]. Therefore, the Bulbophyllum is an excellent type genus to study the effects of geographical isolation on cp genomes. 

In this study, the cp genomes of nine AN Bulbophyllum orchids were newly sequenced and assembled using the next-generation sequencing technology. The comprehensive comparative analyses were performed, based on 27 complete cp genomes of AN and SA Bulbophyllum. The aims of this study were: (1) to uncover the differences in cp genome features between AN clade and SA clade; (2) to assess the effects of long-term geographic isolation on cp genome structures; (3) to investigate the direct causes of the variable selective pressures on cp genes. We believed that this study will give us a comprehensive understanding of the effects of geographic isolation on the complete cp genomes.

Results

Chloroplast genome features

The nine newly sequenced Bulbophyllum cp genomes were circular with the typical quadripartite structure, including a large single copy (LSC), a small single copy (SSC) and a pair of inverted repeats (i.e., IRA and IRB) (Fig. 1). We combined the published cp genomes of ten AN and eight SA Bulbophyllum orchids with this study’s nine AN Bulbophyllum orchids to compare the basic cp genome features of two clades. As shown in Table 1, the AN cp genome sizes were ranged from 144,380 bp to 158,524 bp, with GC contents of 36.66–38.04%. The cp genomes were variable in LSC and SSC regions, with 77,088 to 86,200 bp and 12,140 to 18,632 bp, while conserved in IR regions, with sizes ranging from 25,824 to 26,919 bp. The AN cp genome GC contents varied slightly in LSC (34.20–36.05%), SSC (29.00–30.93%) and IR (43.03–43.38%) regions. Notably, except for IR regions, the GC contents of AN clade were higher than SA clade (Mann–Whitney two-sided, P < 0.05) [17], which indicated that there were significant cp genome differences between two clades. Moreover, the Bulbophyllum cp genomes were well-conserved in gene number and gene order, with 74 protein-coding genes, 30 tRNA genes and 4 rRNA genes.

Fig. 1
figure 1

Cp genomes map of Bulbophyllum orchids. The lengths of the cp genomes are shown in the circle. Cp genes at four junctions and SSC regions are colored in

Table 1 The features of Bulbophyllum cp genomes

The levels of the nine newly sequenced cp genome sequence variability were estimated and plotted using mVISTA, with Dendrobium huoshanense as the reference (Fig. S1). As expected, the non-coding regions (intergenic spacers and introns) were more divergent than coding regions, and the SC regions exhibited higher divergence levels than IR regions. Comparative cp genomic analysis of Bulbophyllum revealed a minor organizational variation, which covered an inversion of rpl32.

Repeat and SSR analysis

The Longer dispersed repeats (LDRs) were detected using REPuter and shown in Fig. S2A. The results showed that, (i) the LDRs distribution densities in SSC region were higher than LSC and IR regions (Mann–Whitney two-sided, P < 0.05). (ii) The LDRs in SSC regions were highly varied between two clades (T-test for Independent Samples: P < 0.05). The analysis of simple sequence repeats (SSRs) also revealed the identical trends (Fig. S2B). Considering the higher LDRs and SSRs distribution densities with the lower GC contents in SSC regions, we conducted correlation tests to see if there were correlative relationships among them. Indeed, significant negative correlations were detected among them (LDRs vs GC contents: Pearson's r = -0.529; SSRs vs GC contents: Pearson's r = -0.977, P < 0.05), which indicated that low GC contents may lead to the SSC regions more variable.

Comparison of sequences flanking IR/SC junctions between Asian and South American clades

The sequences flanking IR/SC junctions in cp genomes were compared. As shown in Fig. 1 and Fig. S3, the Bulbophyllum IR/SC junctions were highly conserved. The SSC/IRA junctions (JSAs) was located in the 5’ end of ycf1, which resulted in a pseudogene fragment Ψycf1 (33 to 1,434 bp) in the SSC/IRB junctions (JSBs). Meanwhile, the LSC/IRB junctions (JLBs) were located in rpl22, which leaded Ψrpl22 (10 to 79 bp) in the LSC/IRA junctions (JLAs). Remarkably, IRB expanded progressively from JSBs to ndhF. The main differences between the two clades were in JSB, which can be divided into three types, (A) a relatively long Ψycf1 (796 to 1,434 bp) with a ΨndhF (B. forrestii, B. leopardinum, B. odoratissimum, B. orientale, B. hirtum, B. hirundinis, B. reptans, B. menghaiense, B. andersonii, B. affine, B. pectinatum, B. funingense, B. pentaneurum, B. disciflorum, B. gedangense, B. lingii, B. steyermarkii); (B) a short Ψycf1 (156 to 360 bp) with a ΨndhF, including B. mentosum and B. epiphytum; (C) no ΨndhF, including B. kwangtungense, B. inconspicuum, B. pingnanense, B. regnellii, B. exaltatum, B. granulosum, B. plumosum and B. weddellii. However, there was no type B in AN clade. Meanwhile, there were more type C in SA clade. These results revealed that the evolution of IR/SC junctions between two clades were diversified.

Diverse mutational hotspots between the Asian and South American clades

We extracted a total of 72 intergenic and intronic loci (conserved sites > 100 bp) from two clades and calculated their sequence variabilities (SVs), with D. huoshanense as the reference. As shown in Fig. 2, the SVs in SC regions were significantly higher than IR regions (Mann–Whitney two-sided, P < 0.05). In the AN clade, the top-ten loci (SV = 0.4309 to 0.6177) mainly located in LSC region (eight in LSC and two in SSC) (Fig. 2A). And in the SA clade, only two mutational hotspots were identical with the AN clade (ccsA-ndhD and clpP-ex1-psbB) (Fig. 2B), which suggested that the evolution of non-coding regions between two clades were disproportional.

Fig. 2
figure 2

Sequence variabilities of non-coding loci among AN and SA clades. Dashed lines indicate top ten hotspots. Blue points indicate common top ten hotspots of two clades

Geographic isolation verification

The aligned cp genomes and ITS sequences were connected to test hybridization. The ten AN Bulbophyllum orchids were set as the BAN branch, the seven SA Bulbophyllum orchids were set as the BSO branch, two D. huoshanense species were set as the DHU branch, four D. moniliforme species were set as the DMO branch and five D. officinale species were set as the DOF branch (Fig. S4). The JML results showed that no hybridization occurred between BAN branch and BSO branch. These results proved that there was no gene flow between two clades, which may due to the long periods of geographic isolation.

Phylogenetic analysis and divergence time estimation

The phylogenetic relationships among 27 Bulbophyllum species were reconstructed by 61 angiosperm cp genomes (Table S1). The ML tree exhibited identical topology with the BI tree. As shown in the Fig. 3B, the phylogenetic trees revealed a monophyletic relationship for the orchids with strongly support (bootstrap values = 100/100). Meanwhile, the Bulbophyllum achieved a monophyly with 100/100 bootstrap values, and was sister to Dendrobium. As expected, the 27 Bulbophyllum orchids were divided into the AN clade and SA clade. These results were consistent with previous studies [4, 18, 25].

Fig. 3
figure 3

Chronogram and Phylogenetic relationships of 61 angiosperms. a Molecular dating results of 61 angiosperms by BEAST2; b ML tree topology with ML and BI bootstrap values. The first one represents ML bootstrap value and the second one represents BI bootstrap value. * represents 100 bootstrap value

ITS sequences and three cp genes (psbA, matK and rbcL) were used to reconstructed the phylogenetic relationships of 27 orchids (including Bulbophyllum, Dendrobium and Phalaenopsis) (Fig. S5). The results supported that 27 Bulbophyllum orchids were strictly divided into two clades. In addition, we found that neither the ITS tree nor the ITS + psbA + matK + rbcL tree was comparable to cp genome tree in bootstrap values. Obviously, the cp genomes have a better performance in studying plant phylogenetic relationships.

The divergence times of 61 orchids were estimated using BEAST2. The topologies of the BEAST tree were almost consistent to ML and BI trees except for B. affine and B. reptans, which may be because reconstructing phylogenetic relationships at the angiosperm-scale hid SNPs that can distinguish between close species (Fig. 3A). In this study, Bulbophyllum was inferred to have diverged from Dendrobium at 36.77 (32.15–41.39) million years ago (mya), and the Bulbophyllum AN clade was inferred to have diverged from SA clade at 25.81 (21.5–30.12) mya. Notably, these results were slightly higher than previous studies [18, 25], mainly because cp genome sequences are longer than ITS sequences and other short fragments.

Cp genome structural variation analysis

The insertions and deletions (InDels) were identified in 27 Bulbophyllum cp genomes, with D. huoshanense as outgroup. There were more deletions (5,314 to 14,498 bp) than insertions (4,216 to 10,652 bp) in Bulbophyllum (Fig. 4A). In addition, the distribution of InDels were different between AN and SA clades. For example, (i) the InDels lengths were varied between two clades. The total InDels lengths of AN clade (5,540 to 10,652 bp and 5,314 to 14,498 bp) were higher than SA clade (4,216 to 6,801 bp and 8,466 to 12,642 bp) (Fig. 4A). (ii) Except for partial AN Bulbophyllum species, the InDels distribution densities in SSC region were higher than in LSC and IR regions (Fig. 4B), which indicated that SSC region structure was diversified among two clades. (iii) In addition to InDels distribution densities in different regions, the occurrence rates of InDels between AN and SA clades were also inconsistent (Fig. 4C). The InDels occurrence rates of AN clade (473.01 and 246.90 bp/myr) were higher than SA clade (127.38 and 213.95 bp/myr). These results revealed that the evolution of cp genome structures between two clades were asymmetry.

Fig. 4
figure 4

InDels analysis results of Bulbophyllum orchids. a Total InDels lengths with reference D. huoshanense; b InDels distribution densities in different regions of Bulbophyllum orchids; c InDels occurrence rates of different branches in Bulbophyllum

Evolution analysis of protein-coding genes

The synonymous (dS) and non-synonymous (dN) substitution rates for each protein-coding genes were estimated using codonML, with D. huoshanense as reference. The values of dS (0–0.1973) were higher than dN (0–0.0451) in Bulbophyllum. Although most of ω values between two clades were similar (Fig. 5A), there were some significantly different protein-coding genes, such as atpA, psbZ, rpl16 and atpE (Fig. 5B, Table S2). These genes, which mainly contained photosynthesis and self-replication functions, were mostly distributed in SC regions (14 in SC regions, two in IR regions) (Table S3). These results indicated that a part of protein-coding genes between two clades were under different selective pressures.

Fig. 5
figure 5

Selective pressure analysis results in Bulbophyllum. a Scatter diagram for dN and dS of 63 cp protein-coding genes from 27 Bulbophyllum orchids with reference D. huoshanense; b Scatter diagram for dN and dS of atpA, psbZ, rpl16 and atpE; c ML tree with positively selective genes in Bulbophyllum branch, AN branch and SA branch, respectively

The Branch model of codeML were used to investigate the possible role of positive selection in driving protein-coding gene evolution among three branches (Bulbophyllum branch, AN branch and SA branch). The comparison showed that a part of genes was under different selective pressures. For example, there were six genes (psbB, psbF, rbcL, rpl23, ycf3 and ycf2) were found with positive selection in the Bulbophyllum branch; and five genes (matK, psbJ, ycf2, psbZ and psbF) were found in the AN branch; and three genes (rpl2, rpl36 and ycf2) were found in the SA branch (Fig. 5C). Most of them were in the LSC region (Table S4). The positively selective protein-coding genes between AN and SA branches were different, which suggested that different environments diversified the evolution of the protein-coding genes.

Discussion

Chloroplast genomes of Bulbophyllum Asian and South American clades were different

Comparative cp genome analyses have revealed that cp genomes are conserved in genome sizes, structures, and gene contents [26,27,28]. Because evolutionary rates of cp genomes are slower than nuclear and mitochondrial genomes [3, 29], complete cp genomes are useful resources for phylogenetic analysis and species identification [30,31,32]. With the rapid development of cp genome researches, there have been comprehensive understandings about the cp genome characteristics. For example, (i) cp genome substitution rates are considerably lower in IR regions than in SC regions [3, 11, 12]; (ii) the repeat sequences played an important role in rearrangements of cp genomes [13,14,15]; (iii) mutational hotspots are genus specific [33, 34]. However, the differences of the cp genomes in different continents have not been analyzed. The genus Bulbophyllum, which is widely distributed in Asia, south America and other places [17, 18], is an excellent source for the comparison of cp genome features, cp genome junctions and mutational hotspots.

We compared the overall structures and variations of Bulbophyllum cp genomes. All results showed that there were considerable differences in overall cp genome features between AN and SA clades. Firstly, they have different cp genome features. For example, (i) the overall, LSC and SSC GC contents of AN clade are higher than SA clade; (ii) part of the rpl32 orientations in AN clade were different from SA clade, which caused by an inversion; (iii) except LSC LDRs and IR SSRs distribution densities, the LDRs and SSRs distribution densities in SA clade were higher than in AN clade. Secondly, JSBs between two clades varied significantly. AN and SA clades had different types of JSBs. There was no type B JSBs in AN clade and there were more type C JSBs in SA clade. Thirdly, the distributions and kinds of mutational hotspots between two clades were significantly different. Niu et al. (2017) found specific barcodes searched by 8–10 closely related cp genomes from specific genus can uncover their cp genome features [33]. Therefore, the mutational hotspots of two clades estimated in this study can reflect the variations of non-coding loci. However, there were only two hotspots shared between two clades, which indicated that the AN and SA clades had discordant evolution in non-coding regions. In conclusion, AN and SA clades had considerable differences in overall cp genome features.

The cp genome structures between Asian and South American clades were different under geographic isolation

Hybridization, a driver of plant evolution, is common among terrestrial orchids [35]. Although dust-like seeds have the capacity to travel long distance [36,37,38], the results of this study showed that Bulbophyllum AN clade has diverged completely from SA clade in the late Oligocene (21.50–30.12 mya) [18]. Because the results of the JML software have ruled out the possibility of hybridization between the two clades [39], Bulbophyllum AN and SA clades have evolved independently for at least 25.81 million years. Our subsequent analyses revealed significant cp genome structural variations between two clades.

The insertions and deletions (InDels) are important structural variations of cp genomes [32]. In order to evaluate the structural variations of the cp genomes during the evolution of Bulbophyllum, we calculated the total InDels lengths of each Bulbophyllum by the aligned complete cp genome sequences, with D. huoshanense as outgroup. There were 5,314 to 14,498 bp deletions and 4,216 to 10,652 bp insertions in Bulbophyllum. Obviously, the lengths of deletions were longer than insertions, which is consistent with previous studies [16, 40]. To obtain more detailed information about InDels, we calculated the distribution densities of InDels in different cp genome regions. The distribution densities of insertions in LSC, IR, and SSC regions were 22.7–48.9 bp/kbp, 4.8–46.1 bp/kbp, and 29.4–286.9 bp/kbp, respectively. And the deletions distribution densities in LSC, IR and SSC regions were 37.8–130.5 bp/kbp, 16.2–92.3 bp/kbp, and 24–389.1 bp/kbp, respectively. The results revealed that the distributions of InDels are dependent on their locations in cp genomes [16, 41, 42], e.g., associated with low GC contents.

To estimate the pace of structural variations in the Bulbophyllum cp genomes, we dated the occurrence rates of InDels by located them at the BEAST tree branches. The results showed that, during the past ~ 10.96 million years since the late Eocene, the insertions and deletions occurred at the speeds of 67.15 bp/myr and 194.07 bp/myr. Then, AN clade and SA clade diverged at ~ 25.81 mya. Our results further revealed that InDels remarkably varied across different lineages [40]. Nevertheless, the most InDels occurrence rates of AN Bulbophyllum were higher than SA Bulbophyllum. We speculated that there were two main reasons for the higher occurrence rates of InDels in the AN clade: (1) the results of Gamisch et al. (2021) indicated that Asia was the original continent of the Bulbophyllum orchids [43]; (2) Of all the continents where Bulbophyllum orchids grow, Asia has the largest suitable area for speciation and niche expansion of the genus Bulbophyllum [43]. In conclusion, the cp genome structures between two clades were different under geographic isolation. And for further study, InDels may be an important indicator of species evolution.

Crassulacean acid metabolism (CAM) pathway may be the inducement of cp gene differentiation

Crassulacean acid metabolism (CAM) is the typical plant photosynthetic pathway that improves water use efficiency of plants [44]. About 10% of orchid species have been confirmed to be CAM plants [45,46,47], and 50% of Orchidaceae species were anticipated to be CAM plants [48]. CAM has evolved multiple independent times in the Orchidaceae [49]. Phylogenetic analyses of the neotropical orchids indicated that CAM has evolved independently at least 10 times [48]. For example, CAM in Dendrobium, the sister genus of Bulbophyllum, has arisen for at least eight times [49]. The Bulbophyllum orchids are typical CAM orchids. Although the C3 pathway is the main photosynthetic pathway of Bulbophyllum presently, the CAM pathway plays an important role in adapting to new ecological niches and increasing species diversity [43].

In order to evaluate the effects of CAM pathway on cp gene evolution in the genus Bulbophyllum, Selective pressure analysis of functional genes has been employed [50,51,52]. Non-synonymous (dN), synonymous (dS) nucleotide substitution rates and non-synonymous/synonymous rate ratio (ω = dN/dS) are widely used as indicators of adaptive evolution [40]. The values of dN and dS on all protein-coding genes between two clades were compared, and we found three photosynthesis genes, ten self-replication genes and three other genes, which have significantly different ω values. The three photosynthesis genes (psbZ, atpE and atpA) encode fractions of the photosystems and the thylakoid ATP synthase complex. In addition, positively selective genes (ω values > 1) through the Bulbophyllum evolution were screened. There were six, five and three positively selective cp genes in the Bulbophyllum branch, AN branch and SA branch, respectively. And these genes were mostly distributed in the LSC regions, which is consistent with previous studies [40]. Because there were many photosynthesis genes, which are components of Photosystem II, among two clades under different selective pressures, the CAM pathway may play an important role in adaptive evolution of Bulbophyllum. In addition, Hu et al. (2022) assessed the origin and loss of CAM in the Bulbophyllum since middle Miocene using two cp genes (matK and psbA) and two nuclear sequences (ITS and Xdh) [53]. They found that although CAM increased the adaptability of the Bulbophyllum plants to specific conditions, it was less competitive than C3 in the long term. Predictably, the role of CAM in the species evolution of Bulbophyllum can be further explored using the complete cp genomes.

Before the ancestors of the Bulbophyllum diverged from Dendrobium at late Eocene, CAM and epiphytism in orchids are likely to have originated about 65 mya [54, 55]. Then, some of the Bulbophyllum photosynthetic genes (psbB, psbF and rbcL) had evolved to adapt to the changeable environment and expand habitat region. At late Oligocene, AN clade diverged from other Bulbophyllum clade, through which SA clade had no gene flow with AN clade. The cp genomes of two clades evolved to different features under different environments. In this process, the CAM pathway affects the selective pressures of photosynthetic genes. Therefore, in subsequent studies, CAM studies and cp genome studies will be more closely linked to explore species evolution of Bulbophyllum.

Conclusions

In this study, the nine Bulbophyllum cp genomes were newly sequenced and assembled. Comparative analyses showed that there were considerable differences in overall cp genome features between AN and SA clades in three aspects, including basic cp genome features, JSBs and mutational hotspots. Then we found that the cp genome structures between two clades were different under geographic isolation by estimating the occurrence rates of the InDels. Finally, the results showed that long-term geographic isolation put AN and SA Bulbophyllum cp genes under different selective pressures. All results indicated that the overall structural characteristics of Bulbophyllum cp genomes diverged during the long-term geographic isolation, and the CAM pathway may play an important role in the cp genes evolution of Bulbophyllum.

Methods

Plant materials and DNA extraction

In this study, nine medicinal Bulbophyllum orchids sequenced (B. forrestii Seidenf. (voucher specimen: Yang202101), B. leopardinum (Wall.) Lindl. (voucher specimen: Yang202102), B. odoratissimum Lindl. (voucher specimen: Yang202103), B. orientale Seidenf. (voucher specimen: Yang202104), B. hirtum (J. E. Smith) Lindl. (voucher specimen: Yang202105), B. reptans Lindl. (voucher specimen: Yang202106), B. kwangtungense Schltr. (voucher specimen: Yang202107), B. hirundinis (Gagnep.) Seidenf. (voucher specimen: Yang202108) and B. andersonii (Hook. f.) J. J. Smith (voucher specimen: Yang202201)) were collected in Yunnan and Fujian provinces, China, and stored in College of Life Sciences, Nanjing Normal University, Nanjing, China. All the samples were identified by Prof. Dr. Xiaoyu Ding. Total genomic DNA of each sample was extracted from 0.2 g fresh leaves using Dneasy Plant Mini Kits (QIAGEN, Germany). DNA samples that met the quality requirements (A260/280 ratio = 1.8–2.0, A260/230 ratio > 1.7, and DNA concentration > 300 ng/μL) were used for sequencing. In addition, ten published AN Bulbophyllum (B. disciflorum, B. gedangense, B. lingii, B. menghaiense, B. pentaneurum, B. pingnanense, B. inconspicuum, B. affine, B. pectinatum and B. funingense) and eight SA Bulbophyllum (B. epiphytum, B. exaltatum, B. granulosum, B. mentosum, B. plumosum, B. regnellii, B. steyermarkii and B. weddellii) cp genomes were selected for subsequent comparative genome analysis.

DNA sequencing, assembly and annotation

The total genomic DNA of nine Bulbophyllum orchids was sequenced using Illumina Hiseq4000 platform (Illumina Inc, San Diego, CA, USA). Approximately 5.0 Gb of raw data were generated with 150 bp paired-end reads for each sample. The fragments with coverage < 50 × were deleted and filtered paired-end reads were assembled on CLC Genomics Workbench v8.5.1 (CLC Bio, Aarhus, Denmark), with reference Dendrobium huoshanense C. Z. Tang et S. J. Cheng. The four junctions between LSC/SSC and IRs were validated with PCR amplification. The assembled cp genomes were annotated using DOGMA v1.2 and tRNAscan-SE v1.21 [56, 57]. The boundaries of annotated genes were confirmed by multiple sequence alignment.

Comparative analysis of cp genomes

The gene location information was extracted, and the IR/SC junction information was collated with the gene location information. The GC contents of 27 Bulbophyllum orchids were analyzed. The nine newly sequenced Bulbophyllum cp genomes were compared using online mVISTA (https://genome.lbl.gov/vista/mvista/submit.shtml) on LAGAN model, with reference D. huoshanense [58].

Characterization of repeat sequences and simple sequence repeats

The REPuter program (https://bibiserv.cebitec.uni-bielefeld.de/reputer) was used to identify the four different types of repeats (forward, reverse, complement and palindromic repeats), with sequence identity ≥ 90%, minimum repeat size ≥ 30 bp, and Hamming distance = 3 [59]. Simple sequence repeats (SSRs) were discovered using the online program MISA (https://webblast.ipk-gatersleben.de/misa/) with the following parameters: 8 for mononucleotide motifs, 5 for polynucleotide motifs [60]. Then the distribution densities of repeat sequences and SSRs on LSC, SSC and IR regions were compared.

IR/SC junctions map

Refer to the method of Zhu et al. (2018) [42], the IR/SC junctions map of 19 AN Bulbophyllum orchids, eight SA Bulbophyllum orchids and D. huoshanense was drawn.

Mutational hotspots

Based on 19 AN Bulbophyllum cp genomes, the top ten hotspots in AN clade were screened. Based on eight SA Bulbophyllum cp genomes, the top ten mutational hotspots in SA clade were screened. The gene location information was transformed into the non-coding region location information, and the sequence of each non-coding region was extracted (ycf1-rpl32 was taken as a whole fragment due to the loss of ndhF). Each non-coding region was aligned with MEGA 5.0 [61], and gaps at both ends were deleted. SNPs, InDels, and conserved sites were calculated by DnaSP v5 [62]. Non-coding regions, of which conserved sites were less than 100 bp, were removed. Sequence variabilities were calculated as follows: Sequence Variabilities = (SNPs + InDels) / (SNPs + InDels + Conserved sites) × 100% [34, 63].

Hybridization testing

The cp genomes and ITS sequences of 17 Bulbophyllum orchids, two D. huoshanense, four D. moniliforme and five D. officinale were aligned using MAFFT 7.220 [64], and were connected together. Under the rule of AIC (Akaike information criterion), the optimum base substitution model calculated by Modeltest 3.7 was GTR + I + Γ [65]. The species tree was constructed using BEAST2 [66]. The Markov Chain Monte Carlo (MCMC) chains were performed to check convergence with 10,000,000 generations. The probability of hybridization was calculated using JML [67]. The significance level was 0.01.

Phylogenetic analysis

Phylogenetic relationships were analyzed based on 61 complete cp genomes, including one dicotyledonous plant, eight monocotyledonous non-orchids, and 52 orchids (Table S1). The cp genome sequences were aligned using MAFFT 7.220 [64]. The conserved loci were selected by Gblocks v.0.91b with the following parameters: minimum number of sequences for a conserved position: 37, minimum number of sequences for a flank position: 62, maximum number of contiguous non-conserved positions: 8, minimum length of a block: 10, allowed gap positions: none [68]. Under the rule of AIC (Akaike information criterion), the optimum base substitution model calculated by Modeltest 3.7 was GTR + I + Γ [65]. The Maximum Likelihood (ML) phylogenetic tree was constructed using RAxML (version 7.4.2) with 1000 rapid bootstrap inferences [69]. The outgroup was Arabidopsis thaliana. The BI analysis was made by using MrBayes 3.2.7 with 1000,000 generations [70]. Trees were sampled every 1,000 generations, and the first 25% of these were discarded. The remaining trees were used to build the Bayesian tree of posterior probabilities. In addition, ITS, psbA, matK, rbcL and cp genomes of 27 orchids were used to reconstructed ML trees with 1000 rapid bootstrap inferences.

Divergence time estimation

Divergence times were estimated by BEAST2 [66]. The following constraints were used for time calibrations: (1) The root age was set as 194 million years ago (Mya) (prior distribution: normal, mean: 194, sd: 8) [71,72,73]. (2) The stem age of Lemnoideae was set as 62 Mya (prior distribution: normal, mean: 62, sd: 3) [74]. (3) Based on the fossil record of Dendrobium, the Crown age of Dendrobium (the divergence time between the Dendrobium Asian clade and the Australian clade) was set at 21 Mya (prior distribution: normal, mean: 21, sd: 5) [25, 75]. Three independent MCMC were performed to check convergence, each with 10,000,000 generations. LogCombiner was used to integrate the data results of three independent runs and remove the top 10% of unstable data.

Structural variation analysis of Bulbophyllum cp genomes

The cp genome sequences of 27 Bulbophyllum orchids and D. huoshanense were aligned using MAFFT 7.220 [64]. The InDels of each Bulbophyllum cp genome were counted, with D. huoshanense as reference. Then, the common InDels of 27 Bulbophyllum cp genomes were counted to calculate the InDels occurrence rates of AN, SA and Bulbophyllum branches. The above common InDels were excluded to calculate the recent occurrence rates of InDels in each Bulbophyllum cp genome.

Nucleotide substitutions rates and positive selection analysis of protein-coding genes

The molecular evolution of 63 cp protein-coding genes were investigated from 27 Bulbophyllum orchids and D. huoshanense. The values of dN, dS and ω were estimated by the codonML in the PAML software (version 4.4) with reference D. huoshanense [76, 77]. Chloroplast protein-coding genes, whose ω values were different between two clades (T-test for Independent Samples) significantly, were screened. EasyCodeML was used to detect positively selective genes in the Bulbophyllum branch, AN branch and SA branch under Branch model with 5% significance level, respectively [40].

Availability of data and materials

All of the raw sequence reads used in this study have been deposited in NCBI (BioProject accession number: LC642719, LC642720, LC642721, LC642722, LC642723, LC642724, LC642725, LC642726 and LC703293).

Abbreviations

Cp:

Chloroplast

AN:

Asian

SA:

South American

InDels:

Insertions and Deletions

CAM:

Crassulacean Acid Metabolism

LSC:

Large Single Copy

SSC:

Small Single Copy

IR:

Inverted Repeat

ITS:

Internal Transcribed Spacer

LDRs:

Longer Dispersed Repeats

SSRs:

Simple Sequence Repeats

JSAs:

SSC/IRA junctions

JSBs:

SSC/IRB junctions

JLBs:

LSC/IRB junctions

JL As:

LSC/IRA junctions

SVs:

Sequence Variabilities

mya:

Million Years Ago

ML:

Maximum Likelihood

References

  1. Li X, Yang Y, Henry RJ, Rossetto M, Wang Y, Chen S. Plant DNA barcoding: from gene to genome. Biol Rev Camb Philos Soc. 2015;90(1):157–66.

    Article  PubMed  Google Scholar 

  2. Feng S, Zheng K, Jiao K, Cai Y, Chen C, Mao Y, Wang L, Zhan X, Ying Q, Wang H. Complete chloroplast genomes of four Physalis species (Solanaceae): lights into genome structure, comparative analysis, and phylogenetic relationships. BMC Plant Biol. 2020;20(1):242.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Wolfe KH, Li WH, Sharp PM. Rates of nucleotide substitution vary greatly among plant mitochondrial, chloroplast, and nuclear DNAs. Proc Natl Acad Sci U S A. 1987;84(24):9054–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Kim YK, Jo S, Cheon SH, Joo MJ, Hong JR, Kwak M, Kim KJ. Plastome Evolution and Phylogeny of Orchidaceae, With 24 New Sequences. Front Plant Sci. 2020;11:22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Jansen RK, Cai Z, Raubeson LA, Daniell H, Depamphilis CW, Leebens-Mack J, Müller KF, Guisinger-Bellian M, Haberle RC, Hansen AK, Chumley TW, Lee SB, Peery R, McNeal JR, Kuehl JV, Boore JL. Analysis of 81 genes from 64 plastid genomes resolves relationships in angiosperms and identifies genome-scale evolutionary patterns. Proc Natl Acad Sci U S A. 2007;104(49):19369–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Dyall SD, Brown MT, Johnson PJ. Ancient invasions: from endosymbionts to organelles. Science. 2004;304(5668):253–7.

    Article  CAS  PubMed  Google Scholar 

  7. Palmer JD, Nugent JM, Herbon LA. Unusual structure of geranium chloroplast DNA: A triple-sized inverted repeat, extensive gene duplications, multiple inversions, and two repeat families. Proc Natl Acad Sci U S A. 1987;84(3):769–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Kim KJ, Lee HL. Widespread occurrence of small inversions in the chloroplast genomes of land plants. Mol Cells. 2005;19(1):104–13.

    CAS  PubMed  Google Scholar 

  9. Greiner S, Wang X, Rauwolf U, Silber MV, Mayer K, Meurer J, Haberer G, Herrmann RG. The complete nucleotide sequences of the five genetically distinct plastid genomes of Oenothera, subsection Oenothera: I. sequence evaluation and plastome evolution. Nucleic Acids Res. 2008;36(7):2366–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Magee AM, Aspinall S, Rice DW, Cusack BP, Sémon M, Perry AS, Stefanović S, Milbourne D, Barth S, Palmer JD, Gray JC, Kavanagh TA, Wolfe KH. Localized hypermutation and associated gene losses in legume chloroplast genomes. Genome Res. 2010;20(12):1700–10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Smith DR. Mutation rates in plastid genomes: they are lower than you might think. Genome Biol Evol. 2015;7(5):1227–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Wu CS, Chaw SM. Evolutionary Stasis in Cycad Plastomes and the First Case of Plastome GC-Biased Gene Conversion. Genome Biol Evol. 2015;7(7):2000–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Asaf S, Khan AL, Khan AR, Waqas M, Kang SM, Khan MA, Lee SM, Lee IJ. Complete Chloroplast Genome of Nicotiana otophora and its Comparison with Related Species. Front Plant Sci. 2016;7:843.

    Article  PubMed  PubMed Central  Google Scholar 

  14. Choi KS, Chung MG, Park S. The Complete Chloroplast Genome Sequences of Three Veroniceae Species (Plantaginaceae): Comparative Analysis and Highly Divergent Regions. Front Plant Sci. 2016;7:355.

    PubMed  PubMed Central  Google Scholar 

  15. Zhang Y, Du L, Liu A, Chen J, Wu L, Hu W, Zhang W, Kim K, Lee SC, Yang TJ, Wang Y. The Complete Chloroplast Genome Sequences of Five Epimedium Species: Lights into Phylogenetic and Taxonomic Analyses. Front Plant Sci. 2016;7:306.

    PubMed  PubMed Central  Google Scholar 

  16. Zhitao N, Shuying Z, Jiajia P, Ludan L, Jing S, Xiaoyu D. Comparative analysis of Dendrobium plastomes and utility of plastomic mutational hotspots. Sci Rep. 2017;7(1):2073.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Zavala-Páez M, Vieira LDN, de Baura VA, Balsanelli E, de Souza EM, Cevallos MC, Chase MW, Smidt EC. Comparative Plastid Genomics of Neotropical Bulbophyllum (Orchidaceae; Epidendroideae). Front Plant Sci. 2020;11:799.

    Article  PubMed  PubMed Central  Google Scholar 

  18. Gamisch A, Comes HP. Clade-age-dependent diversification under high species turnover shapes species richness disparities among tropical rainforest lineages of Bulbophyllum (Orchidaceae). BMC Evol Biol. 2019;19(1):93.

    Article  PubMed  PubMed Central  Google Scholar 

  19. Ya JD, Guo YJ, Liu C, Cai J, Dong GJ, Jiang H, Li DZ. Bulbophyllum reflexipetalum (Orchidaceae, Epidendroideae, Malaxideae), a new species from Xizang. China PhytoKeys. 2019;130:33–9.

    Article  PubMed  Google Scholar 

  20. Gamisch A, Fischer GA, Comes HP. Multiple independent origins of auto-pollination in tropical orchids (Bulbophyllum) in light of the hypothesis of selfing as an evolutionary dead end. BMC Evol Biol. 2015;15:192.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Tang Y, Yang J, Niu Z, Ding X. The complete chloroplast genome sequence of a traditional Chinese medicine plant Bulbophyllum disciflorum Rolfe (Orchidaceae). Mitochondrial DNA B Resour. 2019;5(1):59–60.

    Article  PubMed  PubMed Central  Google Scholar 

  22. Tang H, Tang L, Shao S, Peng Y, Li L, Luo Y. Chloroplast genomic diversity in Bulbophyllum section Macrocaulia (Orchidaceae, Epidendroideae, Malaxideae): Insights into species divergence and adaptive evolution. Plant Divers. 2021;43(5):350–61.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Zhang S, Zeng MY, Gao XY, Li MH, Chen SP. The complete plastome of Bulbophyllum pingnanense (Orchidaceae: Dendrobiinae). Mitochondrial DNA B Resour. 2021;6(11):3193–5.

    Article  PubMed  PubMed Central  Google Scholar 

  24. Yang J, Zhu Z, Fan Y, Zhu F, Chen Y, Niu Z, Ding X. Comparative plastomic analysis of three Bulbophyllum medicinal plants and its significance in species identification. Acta Pharmaceutica Sinica. 2020;55(11):2736–45.

    Google Scholar 

  25. Xiang X, Mi X, Zhou H, Li J, Chung S, Li D, Huang W, Jin W, Li Z, Huang L, Jin X. Biogeographical diversification of mainland Asian Dendrobium (Orchidaceae) and its implications for the historical dynamics of evergreen broad-leaved forests. J Biogeogr. 2016;43(7):1310–23.

    Article  Google Scholar 

  26. Chang CC, Lin HC, Lin IP, Chow TY, Chen HH, Chen WH, Cheng CH, Lin CY, Liu SM, Chang CC, Chaw SM. The chloroplast genome of Phalaenopsis aphrodite (Orchidaceae): comparative analysis of evolutionary rate with that of grasses and its phylogenetic implications. Mol Biol Evol. 2006;23(2):279–91.

    Article  CAS  PubMed  Google Scholar 

  27. Jheng CF, Chen TC, Lin JY, Chen TC, Wu WL, Chang CC. The comparative chloroplast genomic analysis of photosynthetic orchids and developing DNA markers to distinguish Phalaenopsis orchids. Plant Sci. 2012;190:62–73.

    Article  CAS  PubMed  Google Scholar 

  28. Yang JB, Tang M, Li HT, Zhang ZR, Li DZ. Complete chloroplast genome of the genus Cymbidium: lights into the species identification, phylogenetic implications and population genetic analyses. BMC Evol Biol. 2013;13:84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Drouin G, Daoud H, Xia J. Relative rates of synonymous substitutions in the mitochondrial, chloroplast and nuclear genomes of seed plants. Mol Phylogenet Evol. 2008;49(3):827–31.

    Article  CAS  PubMed  Google Scholar 

  30. Tonti-Filippini J, Nevill PG, Dixon K, Small I. What can we do with 1000 plastid genomes? Plant J. 2017;90(4):808–18.

    Article  CAS  PubMed  Google Scholar 

  31. Carbonell-Caballero J, Alonso R, Ibañez V, Terol J, Talon M, Dopazo J. A Phylogenetic Analysis of 34 Chloroplast Genomes Elucidates the Relationships between Wild and Domestic Species within the Genus Citrus. Mol Biol Evol. 2015;32(8):2015–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Cho KS, Yun BK, Yoon YH, Hong SY, Mekapogu M, Kim KH, Yang TJ. Complete Chloroplast Genome Sequence of Tartary Buckwheat (Fagopyrum tataricum) and Comparative Analysis with Common Buckwheat (F esculentum). PLoS One. 2015;10(5):e0125332.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Niu Z, Xue Q, Zhu S, Sun J, Liu W, Ding X. The Complete Plastome Sequences of Four Orchid Species: Insights into the Evolution of the Orchidaceae and the Utility of Plastomic Mutational Hotspots. Front Plant Sci. 2017;8:715.

    Article  PubMed  PubMed Central  Google Scholar 

  34. 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.

    Article  PubMed  Google Scholar 

  35. Piett S, Hager HA, Gerrard C. Characteristics for evaluating the conservation value of species hybrids. Biodivers Conserv. 2015;24(8):1931–55.

    Article  Google Scholar 

  36. Li X, Ding X, Chu B, Zhou Q, Ding G, Gu S. Genetic diversity analysis and conservation of the endangered Chinese endemic herb Dendrobium officinale Kimura et Migo (Orchidaceae) based on AFLP. Genetica. 2008;133(2):159–66.

    Article  CAS  PubMed  Google Scholar 

  37. Fay MF, Pailler T, Dixon KW. Orchid conservation: making the links. Ann Bot. 2015;116(3):377–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Niu Z, Hou Z, Wang M, Ye M, Zhang B, Xue Q, Liu W, Ding X. A comparative plastomics approach reveals available molecular markers for the phylogeographic study of Dendrobium huoshanense, an endangered orchid with extremely small populations. Ecol Evol. 2020;10(12):5332–42.

    Article  PubMed  PubMed Central  Google Scholar 

  39. Barton NH. The role of hybridization in evolution. Mol Ecol. 2001;10(3):551–68.

    Article  CAS  PubMed  Google Scholar 

  40. Gao LZ, Liu YL, Zhang D, Li W, Gao J, Liu Y, Li K, Shi C, Zhao Y, Zhao YJ, Jiao JY, Mao SY, Gao CW, Eichler EE. Evolution of Oryza chloroplast genomes promoted adaptation to diverse ecological habitats. Commun Biol. 2019;2:278.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Ahmed I, Biggs PJ, Matthews PJ, Collins LJ, Hendy MD, Lockhart PJ. Mutational dynamics of aroid chloroplast genomes. Genome Biol Evol. 2012;4(12):1316–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Zhu S, Niu Z, Xue Q, Wang H, Xie X, Ding X. Accurate authentication of Dendrobium officinale and its closely related species by comparative analysis of complete plastomes. Acta Pharm Sin B. 2018;8(6):969–80.

    Article  PubMed  PubMed Central  Google Scholar 

  43. Gamisch A, Winter K, Fischer GA, Comes HP. Evolution of crassulacean acid metabolism (CAM) as an escape from ecological niche conservatism in Malagasy Bulbophyllum (Orchidaceae). New Phytol. 2021;231(3):1236–48.

    Article  PubMed  Google Scholar 

  44. Gilman IS, Edwards EJ. Crassulacean acid metabolism. Curr Biol. 2020;30(2):R57–62.

    Article  CAS  PubMed  Google Scholar 

  45. Winter K, Wallace BJ, Stocker GC, Roksandic Z. Crassulacean acid metabolism in australian vascular epiphytes and some related species. Oecologia. 1983;57(1–2):129–41.

    Article  PubMed  Google Scholar 

  46. Earnshaw MJ, Winter K, Ziegler H, Stichler W, Cruttwell NE, Kerenga K, Cribb PJ, Wood J, Croft JR, Carver KA, Gunn TC. Altitudinal changes in the incidence of crassulacean acid metabolism in vascular epiphytes and related life forms in Papua New Guinea. Oecologia. 1987;73(4):566–72.

    Article  CAS  PubMed  Google Scholar 

  47. Kluge M, Vinson B, Ziegler H. Ecophysiological Studies on Orchids of Madagascar: Incidence and Plasticity of Crassulacean Acid Metabolism in Species of the Genus Angraecum Bory. Plant Ecol. 1998;135(1):43–57.

    Article  Google Scholar 

  48. Silvera K, Santiago LS, Cushman JC, Winter K. Crassulacean acid metabolism and epiphytism linked to adaptive radiations in the Orchidaceae. Plant Physiol. 2009;149(4):1838–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Li MH, Liu DK, Zhang GQ, Deng H, Tu XD, Wang Y, Lan SR, Liu ZJ. A perspective on crassulacean acid metabolism photosynthesis evolution of orchids on different continents: Dendrobium as a case study. J Exp Bot. 2019;70(22):6611–9.

    Article  CAS  PubMed  Google Scholar 

  50. Thomas JW, Touchman JW, Blakesley RW, Bouffard GG, Beckstrom-Sternberg SM, Margulies EH, Blanchette M, Siepel AC, Thomas PJ, McDowell JC, Maskeri B, Hansen NF, Schwartz MS, Weber RJ, Kent WJ, Karolchik D, Bruen TC, Bevan R, Cutler DJ, Schwartz S, Elnitski L, Idol JR, Prasad AB, Lee-Lin SQ, Maduro VV, Summers TJ, Portnoy ME, Dietrich NL, Akhter N, Ayele K, Benjamin B, Cariaga K, Brinkley CP, Brooks SY, Granite S, Guan X, Gupta J, Haghighi P, Ho SL, Huang MC, Karlins E, Laric PL, Legaspi R, Lim MJ, Maduro QL, Masiello CA, Mastrian SD, McCloskey JC, Pearson R, Stantripop S, Tiongson EE, Tran JT, Tsurgeon C, Vogt JL, Walker MA, Wetherby KD, Wiggins LS, Young AC, Zhang LH, Osoegawa K, Zhu B, Zhao B, Shu CL, De Jong PJ, Lawrence CE, Smit AF, Chakravarti A, Haussler D, Green P, Miller W, Green ED. Comparative analyses of multi-species sequences from targeted genomic regions. Nature. 2003;424(6950):788–93.

    Article  CAS  PubMed  Google Scholar 

  51. Nielsen R, Bustamante C, Clark AG, Glanowski S, Sackton TB, Hubisz MJ, Fledel-Alon A, Tanenbaum DM, Civello D, White TJ, Sninsky J, Adams MD, Cargill M. A scan for positively selected genes in the genomes of humans and chimpanzees. PLoS Biol. 2005;3(6):e170.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Sawyer SL, Wu LI, Emerman M, Malik HS. Positive selection of primate TRIM5alpha identifies a critical species-specific retroviral restriction domain. Proc Natl Acad Sci U S A. 2005;102(8):2832–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Hu AQ, Gale SW, Liu ZJ, Fischer GA, Saunders RMK. Diversification Slowdown in the Cirrhopetalum Alliance (Bulbophyllum, Orchidaceae): Insights From the Evolutionary Dynamics of Crassulacean Acid Metabolism. Front Plant Sci. 2022;13:794171.

    Article  PubMed  PubMed Central  Google Scholar 

  54. Russell KM. On the evolutionary pathways resulting in C4 photosynthesis and Crassulacean acid metabolism (CAM). Adv Ecol Res. 1989;19:57–92.

    Article  Google Scholar 

  55. Ehleringer JR, Monson RK. Evolutionary and ecological aspects of photosynthetic pathway variation. Annu Rev Ecol Syst. 1993;24:411–39.

    Article  Google Scholar 

  56. Wyman SK, Jansen RK, Boore JL. Automatic annotation of organellar genomes with DOGMA. Bioinformatics. 2004;20(17):3252–5.

    Article  CAS  PubMed  Google Scholar 

  57. Lowe TM, Chan PP. tRNAscan-SE On-line: integrating search and context for analysis of transfer RNA genes. Nucleic Acids Res. 2016;44(W1):W54–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Frazer KA, Pachter L, Poliakov A, Rubin EM, Dubchak I. VISTA: computational tools for comparative genomics. Nucleic Acids Res. 2004;32:W273–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Kurtz S, Choudhuri JV, Ohlebusch E, Schleiermacher C, Stoye J, Giegerich R. REPuter: the manifold applications of repeat analysis on a genomic scale. Nucleic Acids Res. 2001;29(22):4633–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Thiel T, Michalek W, Varshney RK, Graner A. Exploiting EST databases for the development and characterization of gene-derived SSR-markers in barley (Hordeum vulgare L.). Theor Appl Genet. 2003;106(3):411–22.

    Article  CAS  PubMed  Google Scholar 

  61. Kumar S, Nei M, Dudley J, Tamura K. MEGA: a biologist-centric software for evolutionary analysis of DNA and protein sequences. Brief Bioinform. 2008;9(4):299–306.

    Article  CAS  PubMed  Google Scholar 

  62. Librado P, Rozas J. DnaSP v5: a software for comprehensive analysis of DNA polymorphism data. Bioinformatics. 2009;25(11):1451–2.

    Article  CAS  PubMed  Google Scholar 

  63. Shaw J, Lickey EB, Beck JT, Farmer SB, Liu W, Miller J, Siripun KC, Winder CT, Schilling EE, Small RL. The tortoise and the hare II: relative utility of 21 noncoding chloroplast DNA sequences for phylogenetic analysis. Am J Bot. 2005;92(1):142–66.

    Article  CAS  PubMed  Google Scholar 

  64. Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 2013;30(4):772–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Posada D, Crandall KA. MODELTEST: testing the model of DNA substitution. Bioinformatics. 1998;14(9):817–8.

    Article  CAS  PubMed  Google Scholar 

  66. Bouckaert R, Vaughan TG, Barido-Sottani J, Duchêne S, Fourment M, Gavryushkina A, Heled J, Jones G, Kühnert D, De Maio N, Matschiner M, Mendes FK, Müller NF, Ogilvie HA, du Plessis L, Popinga A, Rambaut A, Rasmussen D, Siveroni I, Suchard MA, Wu CH, Xie D, Zhang C, Stadler T, Drummond AJ. BEAST 2.5: An advanced software platform for Bayesian evolutionary analysis. PLoS Comput Biol. 2019;15(4):e1006650.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Joly S. JML: testing hybridization from species trees. Mol Ecol Resour. 2012;12(1):179–84.

    Article  PubMed  Google Scholar 

  68. Castresana J. Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol Biol Evol. 2000;17(4):540–52.

    Article  CAS  PubMed  Google Scholar 

  69. Stamatakis A. RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics. 2006;22(21):2688–90.

    Article  CAS  PubMed  Google Scholar 

  70. Ronquist F, Huelsenbeck JP. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics. 2003;19(12):1572–4.

    Article  CAS  PubMed  Google Scholar 

  71. Zhang GQ, Liu KW, Li Z, Lohaus R, Hsiao YY, Niu SC, Wang JY, Lin YC, Xu Q, Chen LJ, Yoshida K, Fujiwara S, Wang ZW, Zhang YQ, Mitsuda N, Wang M, Liu GH, Pecoraro L, Huang HX, Xiao XJ, Lin M, Wu XY, Wu WL, Chen YY, Chang SB, Sakamoto S, Ohme-Takagi M, Yagi M, Zeng SJ, Shen CY, Yeh CM, Luo YB, Tsai WC, Van de Peer Y, Liu ZJ. The Apostasia genome and the evolution of orchids. Nature. 2017;549(7672):379–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Jaillon O, Aury JM, Noel B, Policriti A, Clepet C, Casagrande A, Choisne N, Aubourg S, Vitulo N, Jubin C, Vezzi A, Legeai F, Hugueney P, Dasilva C, Horner D, Mica E, Jublot D, Poulain J, Bruyère C, Billault A, Segurens B, Gouyvenoux M, Ugarte E, Cattonaro F, Anthouard V, Vico V, Del Fabbro C, Alaux M, Di Gaspero G, Dumas V, Felice N, Paillard S, Juman I, Moroldo M, Scalabrin S, Canaguier A, Le Clainche I, Malacrida G, Durand E, Pesole G, Laucou V, Chatelet P, Merdinoglu D, Delledonne M, Pezzotti M, Lecharny A, Scarpelli C, Artiguenave F, Pè ME, Valle G, Morgante M, Caboche M, Adam-Blondon AF, Weissenbach J, Quétier F, Wincker P. French-Italian Public Consortium for Grapevine Genome Characterization. The grapevine genome sequence suggests ancestral hexaploidization in major angiosperm phyla. Nature. 2007;449(7161):463–7.

    Article  CAS  PubMed  Google Scholar 

  73. Ramírez SR, Gravendeel B, Singer RB, Marshall CR, Pierce NE. Dating the origin of the Orchidaceae from a fossil orchid with its pollinator. Nature. 2007;448(7157):1042–5.

    Article  CAS  PubMed  Google Scholar 

  74. Givnish TJ, Spalink D, Ames M, Lyon SP, Hunter SJ, Zuluaga A, Iles WJ, Clements MA, Arroyo MT, Leebens-Mack J, Endara L, Kriebel R, Neubig KM, Whitten WM, Williams NH, Cameron KM. Orchid phylogenomics and multiple drivers of their extraordinary diversification. Proc Biol Sci. 1814;2015(282):20151553.

    Google Scholar 

  75. Conran JG, Bannister JM, Lee DE. Earliest orchid macrofossils: Early Miocene Dendrobium and Earina (Orchidaceae: Epidendroideae) from New Zealand. Am J Bot. 2009;96(2):466–74.

    Article  PubMed  Google Scholar 

  76. Yang Z. Likelihood ratio tests for detecting positive selection and application to primate lysozyme evolution. Mol Biol Evol. 1998;15(5):568–73.

    Article  CAS  PubMed  Google Scholar 

  77. Yang Z. PAML 4: phylogenetic analysis by maximum likelihood. Mol Biol Evol. 2007;24(8):1586–91.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We would like to thank the College of Life Sciences, Nanjing Normal University for supporting this work. We are also grateful to the Jiangsu Provincial Engineering Research Center for Technical Industrialization for Dendrobiums for technical support.

Funding

Our work was funded by the National Natural Science Foundation of China (Grant No. 31900268 and 32070353), Natural Science Foundation of Jiangsu Province (BK20190699), Natural science fund for colleges and universities in Jiangsu Province, China (19KJB180005). None of these funding bodies have any relationship with the publication of this manuscript.

Author information

Authors and Affiliations

Authors

Contributions

D.X.Y., N.Z.T. and Y.J.P. designed the study. Y.J.P., Z.F.W., G.Y.J., Y.W.H., X.Q.Q., W.M.T., W.H.M., X.Q.Y. and L.W. performed the experiments. Y.J.P., Z.F.W., G.Y.J. and Y.W.H. analyzed the data. Y.J.P. wrote the manuscript. All authors approved the final version of the manuscript.

Corresponding authors

Correspondence to Zhitao Niu or Xiaoyu Ding.

Ethics declarations

Ethics approval and consent to participate

The collected Bulbophyllum species are distributed in Yunnan, China. Experimental researches with Bulbophyllum species comply with Nanjing Normal University guidelines (http://bwc.njnu.edu.cn/info/1085/1433.htm), preserving the genetic background of the species used. It does not require ethical approval.

Conflicts of interest

The authors have no conflicts of interest to declare.

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:

 Fig S1. mVISTA analysis results of nine newly sequenced Bulbophyllum orchids with reference D. huoshanense. Chloroplast coding regions are indicated in blue, non-translation regions in cyan, andnon-coding regions in red.

Additional file 2:

 Fig S2. Distribution densities of LDRs and SSRs in 27 Bulbophyllum orchids and D. huoshanense. a Distribution density of LDRs; b Distribution density of SSRs.

Additional file 3:

Fig S3. IR/SC junctions map of 19 AN Bulbophyllum, eight SA Bulbophyllum and D. huoshanense. Yellow represents the rpl22 gene, blue represents the ycf1 gene, red represents the ndhF pseudogene and green represents the psbA gene.

Additional file 4:

Fig S4. The species tree constructed by BEAST2.

Additional file 5:

Fig S5. Phylogenetic relationships of 27 orchids. a ML tree reconstructed by ITS sequences; b MLtree reconstructed by ITS sequences, psbA, matK and rbcL; c ML tree reconstructed by cp genomes. * represents 100 bootstrap value.

Additional file 6:

Table S1. The species information of 61 angiosperms used in the phylogenetic analysis

Additional file 7:

Table S2. dN and dS of 16 screened protein-coding genes in Bulbophyllum.

Additional file 8:

Table S3. The basic information of 16 screened protein-coding genes.

Additional file 9:

Table S4. The basic information of positively selective genes.

Rights and permissions

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, visithttp://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yang, J., Zhang, F., Ge, Y. et al. Effects of geographic isolation on the Bulbophyllum chloroplast genomes. BMC Plant Biol 22, 201 (2022). https://doi.org/10.1186/s12870-022-03592-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12870-022-03592-y

Keywords