Next Article in Journal
Integrated Transcriptome and Molecular Docking to Identify the Hub Superimposed Attenuation Targets of Curcumin in Breast Cancer Cells
Next Article in Special Issue
CpCAF1 from Chimonanthus praecox Promotes Flowering and Low-Temperature Tolerance When Expressed in Arabidopsis thaliana
Previous Article in Journal
Cytokine-Induced Killer Cells in Combination with Heat Shock Protein 90 Inhibitors Functioning via the Fas/FasL Axis Provides Rationale for a Potential Clinical Benefit in Burkitt’s lymphoma
Previous Article in Special Issue
IRAPs in Combination with Highly Informative ISSRs Confer Effective Potentials for Genetic Diversity and Fidelity Assessment in Rhododendron
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparative Phylogenetic Analysis for Aerides (Aeridinae, Orchidaceae) Based on Six Complete Plastid Genomes

Key Laboratory of National Forestry and Grassland Administration for Orchid Conservation and Utilization at Landscape Architecture and Arts, Fujian Agriculture and Forestry University, Fuzhou 350002, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(15), 12473; https://doi.org/10.3390/ijms241512473
Submission received: 3 June 2023 / Revised: 4 July 2023 / Accepted: 3 August 2023 / Published: 5 August 2023
(This article belongs to the Special Issue Research Advances in Ornamental Plants Breeding and Biotechnology)

Abstract

:
Aerides Lour. (Orchidaceae, Aeridinae) is a group of epiphytic orchids with high ornamental value, mainly distributed in tropical and subtropical forests, that comprises approximately 20 species. The species are of great value in floriculture and garden designing because of their beautiful flower shapes and colors. Although the morphological boundaries of Aerides are clearly defined, the relationship between Aerides and other closely related genera is still ambiguous in terms of phylogeny. To better understand their phylogenetic relationships, this study used next-generation sequencing technology to investigate the phylogeny and DNA barcoding of this taxonomic unit using genetic information from six Aerides plastid genomes. The quadripartite-structure plastomes ranged from 147,244 bp to 148,391 bp and included 120 genes. Among them, 74 were protein coding genes, 38 were tRNA genes and 8 were rRNA genes, while the ndh genes were pseudogenized or lost. Four non-coding mutational hotspots (rpl20–rpl33, psbM, petB, rpoB–trnCGCA, Pi > 0.06) were identified. A total of 71–77 SSRs and 19–46 long repeats (>30 bp) were recognized in Aerides plastomes, which were mostly located in the large single-copy region. Phylogenetic analysis indicated that Aerides was monophylic and sister to Renanthera. Moreover, our results confirmed that six Aerides species can be divided into three major clades. These findings provide assistance for species identification and DNA barcoding investigation in Aerides, as well as contributes to future research on the phylogenomics of Orchidaceae.

1. Introduction

Aerides Lour. (Aeridinae, Vandeae and Orchidaceae) consists of approximately 20 species of epiphytic and lithophytic perennial herbs [1], which are mainly distributed in the tropical and subtropical regions of Asia [2]. The members of Aerides produce a spectacular branched inflorescence containing numerous pinkish and scented flowers [3]. These orchids are used by the orchid breeders to produce numerous artificial hybrids and cultivars. Since the establishment of Aerides by Loureiro in 1790, the genus has undergone multiple intra-generic taxonomic revisions [2,3,4,5]. Lindley (1840) [4] and Pfitzer (1887) [5] introduced five sections of Aerides, including 26 and 15 species, respectively. Christenson (1987) used morphological analysis to subdivide the genus into four clades (Aerides, Falcata, Fieldingia, Rubescens) with a total of 19 species and described the subdivision characteristics of each clade in detail [6]. Although the morphological boundaries of the Aerides have been clarified, the relationship between Aerides and other closely related genera is still ambiguous in terms of phylogeny.
The constant enrichment of molecular evidence for phylogenetic relationships is an ongoing outcome of the advancements in molecular biology. Kocyan (2008) used three markers such as ITS, matK and trnL-F to study the molecular phylogeny of Aerides and showed that the genus was monophyletic and subdivided it into three evolutionary clades (i.e., sect. Fieldingia, sect. Aerides and sect. Crispa) [3]. However, Christenson (1987) found that only some clades were basically consistent. Comparing these clades with the four clades established based on morphology [6], Xiang (2012) found that the Aerides had multiple origins: A. odorata and A. flabellate and Ascocentrum ampullaceum and Vanda are sister groups [7]. In addition, A. thibautiana and A. krabiensis were separated into a monophyletic group. In order to better understand their phylogeny, it is necessary to determine the differences in genetic information among the major lineages of the Aerides.
With the development of next-generation sequencing (NGS) technology, plastid genome data contributed to the deeper comprehension of the phylogenetic relationships among numerous complex orchid taxa [8,9]. Tu (2020) used the plastid genomes of 46 Goodyerinae plants to reveal the molecular location of the Cheirostylis and Goodyera clades within the Goodyerinae [10]. Liu (2020) used plastid genome sequencing to explain phylogenetic relationships among members of the Cleisostoma–Gastrochilus clades in Aeridinae but did not include Aerides members [11].
To comprehensively understand the evolution of the plastid genomes of Aerides, six species were collected in this study. The plastid structure, sequence differences, mutation hotspots and repeat regions were characterized, variable sites were identified, and the phylogenetic relationships were revealed to further elucidate the evolutionary pattern at the plastid genome level of Aerides. Therefore, our study provides valuable information on the phylogenetic relationships and species identification of Aerides.

2. Results

2.1. Characteristics of the Plastome

The genome sizes of the six newly sequenced plastomes of Aerides ranged from 147,244 bp (A. crassifolia) to 148,391 bp (A. quinquevulnera) (Figure 1), which were in the range of common angiosperm plastome sizes. The quadripartite structure with two inverted repeat regions (IRs) exhibited a large single-copy region (LSC) and a small single-copy region (SSC). The size of each region varied across all species (Table 1). For the IR region, the sizes ranged from 25,706 bp (A. odorata) to 25,852 bp (A. rosea). The SSC regions were 11,033 bp (A. rosea) to 11,897 bp (A. crassifolia), and the LSC regions were in the range of 83,913 bp (A. crassifolia) to 85,570 bp (A. falcata). The G/C content was approximately 36.8% (Table 1).
The plastomes of Aerides encoded 120 genes (including repetitive genes), including 74 protein-coding genes, 38 transfer RNA (tRNA) genes and eight ribosomal RNA (rRNA) genes (Table 1). The ndh genes were all pseudogenes with 5–8 members in each plastome (Table 1). The plastomes of A. falcata and A. odorata contained eight (ndhB/C/D/E/G/I/J/K) pseudogenes; A. lawrenceae and A. quinquevulnera possessed seven (ndhB/C/D/E/G/J/K) pseudogenes; A. rosea and A. crassifolia possessed six (ndhB/C/D/G/J/K) and five (ndhB/D/E/G/I) pseudogenes, respectively.
The positions of the IR junctions were well-conserved across the six species of Aerides (Figure 2). At the junction between the LSC and IRb (JLB), the rpl22 genes of the LSC crossed over into IRb. The adjacent regions of LSC and IRa (JLA), which were located in the rps19 genes and psbA genes, were similar in Aerides. The rpl32 and trnN genes were adjacent to the junction between SSC and IRb (JSB), while the trnN and ycf1 genes were adjacent to the junction between SSC and IRa (JSA). The ycf1 genes were complete in the SSC region.

2.2. Repeated Analysis

The number and distribution area of SSRs were analyzed to elucidate allied species or intra-species variations. The long repeats of Aerides plastomes, including the complement (C), forward (F), palindrome (P) and reverse (R) types, were analyzed by the online REPuter program [12] (Figure 3A, Supplementary Table S2). The number of large repeats detected in the six plastomes were 49 (A. falcata, A. odorata, A. quinquevulnera and A. rosea), 50 (A. lawrenceae) and 65 (A. crassifolia), respectively. Except for A. rosea, almost all the repeats ranged from 20 to 39 bp, with the fewest in 40–49 bp. No complement repeats were detected above 40 bp in length, and they were rare even in the smaller size ranges. In the 30–39 bp group, complement and reverse repeats were found in a few species (Figure 3A).
We also analyzed regions in mononucleotide, dinucleotide, trinucleotide, tetranucleotide, pentanucleotide, and hexanucleotide SSR. Aerides had a total of 71 (A. crassifolia)–77 (A. rosea) SSRs (Figure 3B, Supplementary Table S3). Among the SSRs, the mononucleotide repeats were the most abundant. At least 47–53 mononucleotide repeats were found in the six species, followed by the dinucleotides (9–12 repeats), trinucleotides (4–9 repeats), tetranucleotides (3–5 repeats) and pentanucleotide repeats (1–3 repeats). For hexanucleotide repeats, all the species were recorded with 1–2 repeats, except A. rosea, which had no repeats. Most SSRs were located in the LSC region, whereas few SSRs were located in the IR region (Figure 3B, Supplementary Table S3).

2.3. Plastome Sequence Divergence and Barcoding Investigation

A comparative analysis of the complete plastomes can reveal differences between different species (Figure 4 and Figure 5). We found that the plastome sequences of Aerides exhibited a high degree of similarity, and no rearrangement occurred. The intergenic and intragenic regions were found to have the least similarity between plastomes in Aerides (Figure 5), especially in LSC regions (from psb1 to trnGGCU, rpoB to psbD and trnFGAA-trnVUAC) and SSC regions (from rpl32 to ycf1). Given these results, there are many intergenic and intragenic regions to develop DNA barcodes to differentiate Aerides species.
To further analyze the mutational hotspots of Aerides plastomes, we used DnaSP6 to analyze the nucleotide diversity (Pi) for the alignment of complete genome (Figure 6, Supplementary Tables S4 and S5). The nucleotide diversity (Pi) values of the six plastomes ranged from 0 to 0.12333. At the cutoff point of Pi > 0.06, we selected four mutational hotspots (rpl20-rpl33 > psbM > petB > rpoB-trnCGCA) as candidate barcodes. Protein-coding genes were also used in nucleotide diversity analysis. The results showed that at the cutoff point of Pi > 0.03, two coding sequences (rps12 > ycf1) had high nucleotide diversity and were suitable for phylogeny.

2.4. Phylogenetic Analysis

The phylogenetic relationships inferred by ML analysis (IQ-Tree Ultrafast Method) of the complete genomes and 68 protein-coding genes predicted the same topology (Figure 7; Supplementary Figure S1). In general, the phylogenetic relationship within Renanthera was well-resolved (BS ≥ 75%, PP ≥ 0.90). Six species of Aerides formed a monophyletic genus, and they were classified into three major clades. A. rosea formed the first clade of Aerides, and A. crassifolia formed the second clade. A. lawrenceae, A. quinquevulnera, A. falcata and A. odorata formed the third clade. All the branch nodes in the phylogenetic tree were strongly supported by the ML analysis and the BI analysis.

3. Discussion

3.1. The Plastome Characteristics and Structural Evolution

The size of the plastid genome of Orchidaceae has a large variation depending on the different life types, ranging from 19,047 bp (Epipogium roseum) [13] to 212,688 bp (Cypripedium subtropicum) [14]. The plastid genome size of E. roseum is mainly the result of the loss in genes associated with the fungal heterotrophic habitat, while the plastid genome size of C. subtropicum is due to the expansion of non-coding regions [13,14,15]. Currently, the published plastids of Aeridinae members range from 142,859 bp (Schoenorchis seidenfadenii) [11] to 149,689 bp (Thrixspermum tsii) [16]. In this study, we obtained the plastome sequences of six species of Aerides using next-generation sequencing technology. The plastome size ranges from 147,244 bp to 148,391 bp, wherein the structure and gene order are highly conserved. Our result is in line with the sizes of the previously reported orchid plastomes [11,13,14,15,16,17].
In the majority of angiosperms, plastid genomes are typically inherited maternally and exhibit minimal recombination, thereby maintaining a highly conserved structure among closely related species [18]. The previous studies have shown that contraction and expansion of the IR region is a common phenomenon in the evolution process [19], which is the main reason for differences in plastome genome length [20,21]. Such variations can occur at the boundaries of inverted repeats (IRs) and single-copy regions (LSC and SSC), allowing certain genes into IR or SC regions. It is also the main reason for the difference in plastome length [20]. In this study, we observed slight difference in the IR/SC boundary regions of Aerides plastomes. For example, the 3′ end of the rpl22 gene of all six species of Aerides, which should be intact in the LSC region, extended into the IRb region for 30–31 bp, which was found in Renanthera [22]. In addition, the ycf1 gene in the SSC region of other orchids (such as Pholidota [23] and Thuniopsis [24]) was observed crossing over JSA, extending into the IRa region. However, the ycf1 of Aerides was intact in the SSC region. Our study indicates that the plastid genome size and the IR regions are conserved in Aerides compared with other orchids. Therefore, the difference in plastid genome size in Aerides may be due to the presence of indels in the intergenic spacer regions and the loss of pseudogenization in the ndh genes [14,17]. However, the total GC content of Aerides does not exceed 36.8%, which is similar to other members of Aeridinae.
There are 11 ndh genes in the plastids of plants [25], and we detected 5-8 ndh genes in six Areides plants, and all of them were pseudogenized, while the ndh A/F/H genes were completely lost. This is consistent with the plastome study of the subtribe Aeridinae [11,26]. Although the ndh genes were detected in the mitochondrial (mt) genome of some orchids, there is no direct evidence that these genes are associated with the loss of ndh genes in the plastome [27]. Therefore, the mechanism of ndh gene deletion and pseudogenization in orchids needs to be explored in further studies. In addition, studies have suggested that the loss of ndh genes in angiosperms can prevent plants from evolving and diversifying [28,29] and reduce the ecological adaptability of species. The deletion of ndh genes was found in all five subfamilies of Orchidaceae, with epiphytic orchids showing a higher frequency of ndh gene loss than geophytic orchids [15]. However, this does not directly prove that the diversity and adaptability of Orchidaceae are related to the loss of ndh genes.

3.2. The Barcoding Investigation and Phylogenetic Analysis

Nucleotide diversity (Pi) can indicate the degree of variation in the nucleic acid sequences in different species, and the position with higher variability can be used as a molecular marker of population genetics [30,31]. The genetic loci used for DNA barcoding usually contain enough informative loci to effectively define closely related species, which was demonstrated in orchids [32,33,34,35]. In this study, a nucleotide diversity analysis was performed on intact plastids of Aerides, and four highly variable regions were identified. The nucleotide diversity analysis of protein-coding genes identified two highly variable coding genes. The six highly variable regions identified in this study can be used as DNA molecular markers to distinguish Aerides relatives, and the results can be used to develop Aeridinae DNA barcodes. Repeated sequences play an important role in the evolution of species, as well as the inheritance and variation of genes within species [36,37]. These repetitive sequences are widely used in the studies on genetic diversity, population structure, and closely related species identification [38,39,40,41]. In this study, a total of 71–77 SSRs and 19–46 long repeats (>30 bp) were identified from Aerides plasmids, indicating that the plasmid genome of Aerides retained abundant genetic information. Most of the SSRs are mononucleotide repeats in these six Aerides species, which are mostly located in the intergeneric regions of LSC and enriched in non-coding regions. Similar results are found in most angiosperms [42,43,44,45]. The above findings can provide a data basis for further studies on population genetics.
Our results revealed the phylogenetic position of Aerides and the intrageneric relationships. Despite being a valuable and threatened class of orchids, there are few studies on the phylogenetic relationships in Aerides. According to phylogenetic analysis using short gene sequences, Aerides is monophyletic, sisterly to Renanthera, Arachnis and Esmeralda [46]. However, their phylogenetic relationships predicted unstable topology and low support values based on the small amount of short gene sequence data. The role of plastome data in the phylogenetic relationships in the reconstruction of tribes, subtribes and genera in Orchidaceae was also demonstrated [10,11,32]. Our results suggest that the Aerides clade is an independent clade of Aeridinae that is sister to the Renanthera clade. Similar results were found in previous studies on short gene sequences [46]. Nevertheless, the Aerides-Renanthera clade is sister to the Vanda clade, which is inconsistent compared with the previous studies [46]. This study highly signifies the intrageneric consanguinity of Aerides and provides a new perspective for elucidating their relationships.

4. Materials and Methods

4.1. Plant Materials, DNA Extraction and Sequencing

Six Aerides species were selected, including A. falcata, A. crassifolia, A. odorata, A. quinquevulnera, A. rosea and A. lawrenceae. They were introduced and cultivated in the Fujian Agriculture and Forestry University, Fujian province, China. Their voucher information is given in Supplementary Table S1. The total DNA was isolated using a modified CTAB method [47]. Short-insert (500 bp) pair-end (PE) libraries were constructed, and the sequencing was conducted by the Beijing Genomics Institute (Shenzhen, China) on the Illumina HiSeq 2500 platform with a read length of 150 bp. At least 10 Gb of clean data were obtained for each species.

4.2. Plastome Assembly and Annotation

The plastome assembly and annotation were performed following the previously described methods [9]. In short, The paired-end reads were assembled using the GetOrganelle pipeline (https://github.com/Kinggerm/GetOrganelle, accessed on 30 April 2023), and then, the filtered reads were assembled by SPAdes version 3.10 [48]. The published plastome of Phalaenopsis hygrochila (MN124430) was used as a reference for the assembly of plastomes. The gene annotation was carried out using DOGMA [49] based on default parameters and checked through Geneious Prime v2021.1.1 [50]. The circle maps were drawn using OGDRAW [51].

4.3. Genome Comparison and Analysis, IR Border and Divergence Analyses

The plastome genomes across the six species of Aerides were aligned with mVISTA using the LAGAN alignment program [52] with the sequence of A. falcata as a reference. The rearrangements of plastomes were detected and plotted using Mauve of six species [53]. The boundaries between the IRs, SSC and LSC of the plastomes were compared using the online program IRscope (https://irscope.shinyapps.io/irapp/, accessed on 30 April 2023) [54].
To identify the mutational hotspot regions and genes, the plastome sequences were aligned using MAFFT v7 [55]. Then, the nucleotide diversity (Pi) of six plastomes of Aerides was calculated using DnaSP v6.12.03 (DNA Sequences Polymorphism) [56]. Highly mutational hotspot regions were identified through a sliding window strategy. The step size was set to 25 bp, with a 100 bp window length.

4.4. Repeat Sequence Analysis

The online software REPuter (https://bibiserv.cebitec.uni-bielefeld.de/reputer, accessed on 30 April 2023) was used to identify the repeat sequences, including forward, palindrome, reverse, and complement long repeats. The maximum and minimum repeat sizes were set as 50 bp and 20 bp, respectively; while the hamming distance was set to 3 [12]. MISA-web was used to detect simple sequence repeats (SSRs). The parameters were set as a threshold of mononucleotide, dinucleotide, trinucleotide, tetranucleotide, pentanucleotide and hexanucleotide SSRs and minimal repeat numbers of 10, 5, 4,3, 3 and 3, respectively [57].

4.5. Phylogenetic Reconstruction

We used the whole plastome and 68 protein-coding sequences to perform the phylogenetic analysis of 28 species of Orchidaceae. Of these 28 species, 6 Aerides species are our newly sequenced species and the other 22 species of 18 genera are from the complete plastid data publicly available at NCBI. A list of the taxa analyzed with voucher information and GenBank accessions is provided in Supplementary Table S1. The whole plastome sequences were aligned by Geneious Prime v2021.1.1 [50]. A total of 68 protein-coding genes were aligned by PhyloSuite v1.2.2 [58]. The phylogenetic relationships were analyzed by maximum parsimony (MP), maximum likelihood (ML) and Bayesian inference (BI) on the website CIPRES Science Gateway [59]. All characters were equally weighted and unordered, and a heuristic search with 1000 random addition sequence replicates and TBR branch swapping was performed. For ML analysis, the GTRCAT model was specified for all datasets and 1000 repeated self-expanding analyses were performed [60].
The Bayesian analyses were performed with MrBayes v. 3.2.6 [61], and four Markov chains were run for 10,000,000 generations, with one tree sampled every 100 generations. The first 25% of trees were discarded as burn-in samples to ensure that each chain reaches a stable state and the posterior probabilities (PP) were estimated.

5. Conclusions

Our research shows that the overall structure and gene content of the plastomes of six Aerides species are relatively conserved, with only certain differences in genome size, gene content, GC content, repeat sequences and IR boundary, and all ndh genes were lost or pseudogenized. This study provides a reference for developing DNA barcoding in further studies on the Aerides species. The phylogenetic analyses based on the available data identified the genus Aerides as a separate clade of Aeridinae, sister to Renanthera, thereby significantly aiding in reconstructing the phylogenetic connections of Aeridinae. Therefore, our findings offer valuable support for future investigations into the phylogeny and evolution of Aerides and Orchidaceae.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms241512473/s1.

Author Contributions

Conceptualization, methodology, J.C., Z.L. and D.P.; Software, formal analysis, visualization, J.C.; Investigation, resources and data curation, J.C., F.W. and C.Z.; writing—original draft preparation, J.C.; writing—review and editing, Y.Z., S.A., M.L., Z.L. and D.P.; supervision, J.C., Z.L. and D.P.; project administration, D.P.; funding acquisition D.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the following fundings: the Forestry Peak Discipline Construction Project of Fujian Agriculture and Forestry University (No. 72202200205), the Nature Science Foundation of Fujian Province, China (2021J01134) provided support for M.-H.L., and the Innovation and Application Engineering Technology Research Center of Ornamental Plant Germplasm Resources in Fujian Province (No. 115-PTJH16005).

Data Availability Statement

The six plastome sequences are deposited in GenBank of the NCBI repository, accession numbers OR159896 to OR159901.

Acknowledgments

We acknowledge the technical support by lab staff during the conduction of lab experiments, Zhuang Zhao, Gui-Zhen Chen, Jie Huang, Ding-Kun Liu, Kai Zhao, Yang Hao and He-Ming Chen.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chase, M.W.; Cameron, K.M.; Freudenstein, J.V.; Pridgeon, A.M.; Salazar, G.; Van den Berg, C.; Schuiteman, A. An updated classification of Orchidaceae. Bot. J. Linn. Soc. 2015, 177, 151–174. [Google Scholar] [CrossRef]
  2. Pridgeon, A.M.; Cribb, P.J.; Chase, M.W. Genera Orchidacearum Volume 6: Epidendroideae (Part 3); OUP: Oxford, UK, 2014. [Google Scholar]
  3. Kocyan, A.; de Vogel, E.F.; Conti, E.; Gravendeel, B. Molecular phylogeny of Aerides (Orchidaceae) based on one nuclear and two plastid markers: A step forward in understanding the evolution of the Aeridinae. Mol. Phylogenet. Evol. 2008, 48, 422–443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Lindley, J. The Genera and Species of Orchidaceous Plants; Ridgways, Piccadilly: London, UK, 1840. [Google Scholar]
  5. Pfitzer, E.H.H. Entwurf einer Natürlichen Anordnung der Orchideen; Winter’s Universitä Tsbuchhandlung: Heidelberg, Germany, 1887. [Google Scholar]
  6. Christenson, E.A. An infrageneric classification of Holcoglossum Schltr. (Orchidaceae: Sarcanthinae) with a key to the genera of the Aerides-Vanda alliance. Notes R. Bot. Gard. 1987, 44, 249–256. [Google Scholar]
  7. Xiang, X.; Li, D.; Jin, X.; Hu, H.; Zhou, H.; Jin, W.; Lai, Y. Monophyly or paraphyly the taxonomy of Holcoglossum (Aeridinae: Orchidaceae). PLoS ONE 2012, 7, e52050. [Google Scholar] [CrossRef] [Green Version]
  8. Li, Y.X.; Li, Z.H.; Schuiteman, A.; Chase, M.W.; Li, J.W.; Huang, W.C.; Hidayat, A.; Wu, S.S.; Jin, X.H. Phylogenomics of Orchidaceae based on plastid and mitochondrial genomes. Mol. Phylogenet. Evol. 2019, 139, 106540. [Google Scholar] [CrossRef] [PubMed]
  9. Jin, J.-J.; Yu, W.-B.; Yang, J.-B.; Song, Y.; DePamphilis, C.W.; Yi, T.S.; Li, D.Z. GetOrganelle: A fast and versatile toolkit for accurate de novo assembly of organelle genomes. Genome Biol. 2020, 21, 241. [Google Scholar] [CrossRef]
  10. Tu, X.D.; Liu, D.K.; Xu, S.W.; Zhou, C.Y.; Gao, X.Y.; Zeng, M.Y.; Zhang, S.; Chen, J.L.; Ma, L.; Zhou, Z.; et al. Plastid phylogenomics improves resolution of phylogenetic relationship in the Cheirostylis and Goodyera clades of Goodyerinae (Orchidoideae, Orchidaceae). Mol. Phylogenet. Evol. 2021, 164, 107269. [Google Scholar] [CrossRef]
  11. Liu, D.K.; Tu, X.D.; Zhao, Z.; Zeng, M.Y.; Zhang, S.; Ma, L.; Zhang, G.Q.; Wang, M.M.; Liu, Z.J.; Lan, S.R.; et al. Plastid phylogenomic data yield new and robust insights into the phylogeny of Cleisostoma-Gastrochilus clades (Orchidaceae, Aeridinae). Mol. Phylogenet. Evol. 2020, 145, 106729. [Google Scholar] [CrossRef]
  12. Kurtz, S.; Choudhuri, J.V.; Ohlebusch, E.; Schleiermacher, C.; Stoye, J.; Giegerich, R. REPuter: The manifold applications of repeat analysis on a genomic scale. Nucleic Acids Res. 2001, 29, 4633–4642. [Google Scholar] [CrossRef] [Green Version]
  13. Schelkunov, M.I.; Shtratnikova, V.Y.; Nuraliev, M.S.; Selosse, M.A.; Penin, A.A.; Logacheva, M.D. Exploring the limits for reduction of plastid genomes: A case study of the mycoheterotrophic orchids Epipogium aphyllum and Epipogium roseum. Genome Biol. Evol. 2015, 7, 1179–1191. [Google Scholar] [CrossRef] [Green Version]
  14. Guo, Y.Y.; Yang, J.X.; Li, H.K.; Zhao, H.S. Chloroplast Genomes of Two Species of Cypripedium: Expanded Genome Size and Proliferation of AT-Biased Repeat Sequences. Front. Plant Sci. 2021, 12, 609729. [Google Scholar] [CrossRef] [PubMed]
  15. Kim, Y.K.; Jo, S.; Cheon, S.H.; Joo, M.J.; Hong, J.R.; Kwak, M.; Kim, K.J. Plastome Evolution and Phylogeny of Orchidaceae, with 24 New Sequences. Front. Plant Sci. 2020, 11, 22. [Google Scholar] [CrossRef] [Green Version]
  16. Ye, B.J.; Zhang, S.; Tu, X.D.; Liu, D.K.; Li, M.H. The complete plastid genome of Thrixspermum tsii (Orchidaceae, Aeridinae). Mitochondrial DNA Part B 2020, 5, 384–385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Guo, Y.Y.; Yang, J.X.; Bai, M.Z.; Zhang, G.Q.; Liu, Z.J. The chloroplast genome evolution of Venus slipper (Paphiopedilum): IR expansion, SSC contraction, and highly rearranged SSC regions. BMC Plant Biol. 2021, 21, 248. [Google Scholar] [CrossRef]
  18. Wicke, S.; Schneeweiss, G.M.; dePamphilis, C.W.; Müller, K.F.; Quandt, D. The evolution of the plastid chromosome in land plants: Gene content, gene order, gene function. Plant Mol. Biol. 2011, 76, 273–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Huang, H.; Shi, C.; Liu, Y.; Mao, S.Y.; Gao, L.Z. Thirteen Camellia chloroplast genome sequences determined by high-throughput sequencing: Genome structure and phylogenetic relationships. BMC Evol. Biol. 2014, 14, 151. [Google Scholar] [CrossRef] [Green Version]
  20. Raubeson, L.A.; Peery, R.; Chumley, T.W.; Dziubek, C.; Fourcade, H.M.; Boore, J.L.; Jansen, R.K. Comparative chloroplast genomics: Analyses including new sequences from the angiosperms Nuphar advena and Ranunculus macranthus. BMC Genom. 2007, 8, 174. [Google Scholar] [CrossRef] [Green Version]
  21. Dugas, D.V.; Hernandez, D.; Koenen, E.J.; Schwarz, E.; Straub, S.; Hughes, C.E.; Jansen, R.K.; Nageswara-Rao, M.; Staats, M. Mimosoid legume plastome evolution: IR expansion, tandem repeat expansions, and accelerated rate of evolution in clpP. Sci. Rep. 2015, 5, 16958. [Google Scholar] [CrossRef] [Green Version]
  22. Xiao, T.; He, L.; Yue, L.; Zhang, Y.; Lee, S.Y. Comparative phylogenetic analysis of complete plastid genomes of Renanthera (Orchidaceae). Front. Genet. 2022, 13, 998575. [Google Scholar] [CrossRef]
  23. Li, L.; Wang, W.; Zhang, G.; Wu, K.; Fang, L.; Li, M.; Liu, Z.; Zeng, S. Comparative analyses and phylogenetic relationships of thirteen Pholidota species (Orchidaceae) inferred from complete chloroplast genomes. BMC Plant Biol. 2023, 23, 269. [Google Scholar] [CrossRef]
  24. Li, L.; Wu, Q.; Fang, L.; Wu, K.; Li, M.; Zeng, S. Comparative Chloroplast Genomics and Phylogenetic Analysis of Thuniopsis and Closely Related Genera within Coelogyninae (Orchidaceae). Front. Genet. 2022, 13, 850201. [Google Scholar] [CrossRef] [PubMed]
  25. Shikanai, T. Chloroplast NDH: A different enzyme with a structure similar to that of respiratory NADH dehydrogenase. Biochim. Biophys. Acta 2016, 1857, 1015–1022. [Google Scholar] [CrossRef] [PubMed]
  26. Kim, Y.K.; Jo, S.; Cheon, S.H.; Kwak, M.; Kim, Y.D.; Kim, K.J. Plastome evolution and phylogeny of subtribe Aeridinae (Vandeae, Orchidaceae). Mol. Phylogenet. Evol. 2020, 144, 106721. [Google Scholar] [CrossRef]
  27. Lin, C.S.; Chen, J.J.; Huang, Y.T.; Chan, M.T.; Daniell, H.; Chang, W.J.; Hsu, C.T.; Liao, D.C.; Wu, F.H.; Lin, S.Y.; et al. The location and translocation of ndh genes of chloroplast origin in the Orchidaceae family. Sci. Rep. 2015, 5, 9040. [Google Scholar] [CrossRef] [Green Version]
  28. Sabater, B. On the Edge of Dispensability, the Chloroplast ndh Genes. Int. J. Mol. Sci. 2021, 22, 12505. [Google Scholar] [CrossRef]
  29. Sun, Y.; Deng, T.; Zhang, A.; Moore, M.J.; Landis, J.B.; Lin, N.; Zhang, H.; Zhang, X.; Huang, J.; Zhang, X.; et al. Genome Sequencing of the Endangered Kingdonia uniflora (Circaeasteraceae, Ranunculales) Reveals Potential Mechanisms of Evolutionary Specialization. iScience 2020, 23, 101124. [Google Scholar] [CrossRef] [PubMed]
  30. Ding, S.; Dong, X.; Yang, J.; Guo, C.; Cao, B.; Guo, Y.; Hu, G. Complete Chloroplast Genome of Clethra fargesii Franch., an Original Sympetalous Plant from Central China: Comparative Analysis, Adaptive Evolution, and Phylogenetic Relationships. Forests 2021, 12, 441. [Google Scholar] [CrossRef]
  31. Abdullah; Mehmood, F.; Shahzadi, I.; Waseem, S.; Mirza, B.; Ahmed, I.; Waheed, M.T. Chloroplast genome of Hibiscus rosasinensis (Malvaceae): Comparative analyses and identification of mutational hotspots. Genomics 2020, 112, 581–591. [Google Scholar] [CrossRef]
  32. Zhang, L.; Huang, Y.W.; Huang, J.L.; Ya, J.D.; Zhe, M.Q.; Zeng, C.X.; Zhang, Z.R.; Zhang, S.B.; Li, D.Z.; Li, H.T.; et al. DNA barcoding of Cymbidium by genome skimming: Call for next-generation nuclear barcodes. Mol. Ecol. Resour. 2023, 23, 424–439. [Google Scholar] [CrossRef]
  33. Li, Y.; Tong, Y.; Xing, F. DNA Barcoding Evaluation and Its Taxonomic Implications in the Recently Evolved Genus Oberonia Lindl. (Orchidaceae) in China. Front. Plant Sci. 2016, 7, 1791. [Google Scholar] [CrossRef] [Green Version]
  34. 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, 2073. [Google Scholar] [CrossRef] [Green Version]
  35. Smidt, E.C.; Páez, M.Z.; Vieira, L.D.N.; Viruel, J.; de Baura, V.A.; Balsanelli, E.; de Souza, E.M.; Chase, M.W. Characterization of sequence variability hotspots in Cranichideae plastomes (Orchidaceae, Orchidoideae). PLoS ONE 2020, 15, e0227991. [Google Scholar] [CrossRef]
  36. Chen, Y.; Hu, N.; Wu, H. Analyzing and characterizing the chloroplast genome of Salix wilsonii. BioMed Res. Int. 2019, 2019, 5190425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Khan, A.; Asaf, S.; Khan, A.L.; Al-Harrasi, A.; Al-Sudairy, O.; AbdulKareem, N.M.; Khan, A.; Shehzad, T.; Alsaady, N.; Al-Lawati, A.; et al. First complete chloroplast genomics and comparative phylogenetic analysis of Commiphora gileadensis and C. foliacea: Myrrh producing trees. PLoS ONE 2019, 14, e0208511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Yu, J.; Dossa, K.; Wang, L.; Zhang, Y.; Wei, X.; Liao, B.; Zhang, X. PMDBase: A database for studying microsatellite DNA and marker development in plants. Nucleic Acids Res. 2017, 45, D1046–D1053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Singh, R.B.; Mahenderakar, M.D.; Jugran, A.K.; Singh, R.K.; Srivastava, R.K. Assessing genetic diversity and population structure of sugarcane cultivars, progenitor species and genera using microsatellite (SSR) markers. Gene 2020, 753, 144800. [Google Scholar] [CrossRef] [PubMed]
  40. Liu, X.; Xu, D.; Hong, Z.; Zhang, N.; Cui, Z. Comparative and Phylogenetic Analysis of the Complete Chloroplast Genome of Santalum (Santalaceae). Forests 2021, 12, 1303. [Google Scholar] [CrossRef]
  41. Hong, Z.; He, W.; Liu, X.; Tembrock, L.R.; Wu, Z.; Xu, D.; Liao, X. Comparative Analyses of 35 Complete Chloroplast Genomes from the Genus Dalbergia (Fabaceae) and the Identification of DNA Barcodes for Tracking Illegal Logging and Counterfeit Rosewood. Forests 2022, 13, 626. [Google Scholar] [CrossRef]
  42. Li, D.-M.; Zhao, C.-Y.; Liu, X.-F. Complete chloroplast genome sequences of Kaempferia galanga and Kaempferia elegans: Molecular structures and comparative analysis. Molecules 2019, 24, 474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Mo, Z.; Lou, W.; Chen, Y.; Jia, X.; Zhai, M.; Guo, Z.; Xuan, J. The chloroplast genome of Carya illinoinensis: Genome structure, adaptive evolution, and phylogenetic analysis. Forests 2020, 11, 207. [Google Scholar] [CrossRef] [Green Version]
  44. Xu, J.; Liu, C.; Song, Y.; Li, M. Comparative Analysis of the Chloroplast Genome for Four Pennisetum Species: Molecular Structure and Phylogenetic Relationships. Front. Genet. 2021, 12, 687844. [Google Scholar] [CrossRef] [PubMed]
  45. Sun, Y.; Zou, P.; Jiang, N.; Fang, Y.; Liu, G. Comparative analysis of the complete chloroplast genomes of nine Paphiopedilum species. Front. Genet. 2022, 12, 772415. [Google Scholar] [CrossRef] [PubMed]
  46. Zou, L.H.; Huang, J.X.; Zhang, G.Q.; Liu, Z.J.; Zhuang, X.Y. A molecular phylogeny of Aeridinae (Orchidaceae: Epidendroideae) inferred from multiple nuclear and chloroplast regions. Mol. Phylogenet. Evol. 2015, 85, 247–254. [Google Scholar] [CrossRef] [PubMed]
  47. Li, J.; Wang, S.; Yu, J.; Wang, L.; Zhou, S. A Modified CTAB Protocol for Plant DNA Extraction. Chin. Bull. Bot. 2013, 48, 72–78. [Google Scholar]
  48. Bankevich, A.; Nurk, S.; Antipov, D.; Gurevich, A.A.; Dvorkin, M.; Kulikov, A.S.; Lesin, V.M.; Nikolenko, S.I.; Pham, S.; Prjibelski, A.D.; et al. SPAdes: A new genome assembly algorithm and its applications to single-cell sequencing. J. Comput. Biol. J. Comput. Mol. Cell Biol. 2012, 19, 455–477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Wyman, S.K.; Jansen, R.K.; Boore, J.L. Automatic annotation of organellar genomes with DOGMA. Bioinformatics 2004, 20, 3252–3255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Kearse, M.; Moir, R.; Wilson, A.; Stones-Havas, S.; Cheung, M.; Sturrock, S.; Buxton, S.; Cooper, A.; Markowitz, S.; Duran, C.; et al. Geneious Basic: An integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 2012, 28, 1647–1649. [Google Scholar] [CrossRef] [Green Version]
  51. Greiner, S.; Lehwark, P.; Bock, R. OrganellarGenomeDRAW (OGDRAW) version 1.3.1: Expanded toolkit for the graphical visualization of organellar genomes. Nucleic Acids Res. 2019, 47, W59–W64. [Google Scholar] [CrossRef] [Green Version]
  52. Brudno, M.; Malde, S.; Poliakov, A.; Do, C.B.; Couronne, O.; Dubchak, I.; Batzoglou, S. Glocal alignment: Finding rearrangements during alignment. Bioinformatics 2003, 19 (Suppl. S1), i54–i62. [Google Scholar] [CrossRef] [Green Version]
  53. Rissman, A.I.; Mau, B.; Biehl, B.S.; Darling, A.E.; Glasner, J.D.; Perna, N.T. Reordering contigs of draft genomes using the Mauve aligner. Bioinformatics 2009, 25, 2071–2073. [Google Scholar] [CrossRef]
  54. Amiryousefi, A.; Hyvonen, J.; Poczai, P. IRscope: An online program to visualize the junction sites of chloroplast genomes. Bioinformatics 2018, 34, 3030–3031. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Rozas, J.; Ferrer-Mata, A.; Sánchez-DelBarrio, J.C.; Guirao-Rico, S.; Librado, P.; Ramos-Onsins, S.E.; Sánchez-Gracia, A. DnaSP 6: DNA Sequence Polymorphism Analysis of Large Data Sets. Mol. Biol. Evol. 2017, 34, 3299–3302. [Google Scholar] [CrossRef] [PubMed]
  57. Beier, S.; Thiel, T.; Münch, T.; Scholz, U.; Mascher, M. MISA-web: A web server for microsatellite prediction. Bioinformatics 2017, 33, 2583–2585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Zhang, D.; Gao, F.; Jakovlić, I.; Zou, H.; Zhang, J.; Li, W.X.; Wang, G.T. PhyloSuite: An integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Mol. Ecol. Resour. 2020, 20, 348–355. [Google Scholar] [CrossRef] [PubMed]
  59. Miller, M.A.; Pfeiffer, W.; Schwartz, T. Creating the CIPRES Science Gateway for inference of large phylogenetic trees. In Proceedings of the 2010 Gateway Computing Environments Workshop (GCE), New Orleans, LA, USA, 14 November 2010; pp. 1–8. [Google Scholar]
  60. Stamatakis, A.; Hoover, P.; Rougemont, J. A rapid bootstrap algorithm for the RAxML Web servers. Syst. Biol. 2008, 57, 758–771. [Google Scholar] [CrossRef]
  61. Ronquist, F.; Teslenko, M.; Van Der Mark, P.; Ayres, D.L.; Darling, A.; Höhna, S.; Larget, B.; Liu, L.; Suchard, M.A.; Huelsenbeck, J.P. MrBayes 3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 2012, 61, 539–542. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The annotation map of six Aerides plastomes. The darker gray in the inner circle corresponds to the GC content. The IRa and IRb (two inverted repeating regions), LSC (large single-copy region) and SSC (small single-copy region) are indicated outside of the GC content.
Figure 1. The annotation map of six Aerides plastomes. The darker gray in the inner circle corresponds to the GC content. The IRa and IRb (two inverted repeating regions), LSC (large single-copy region) and SSC (small single-copy region) are indicated outside of the GC content.
Ijms 24 12473 g001
Figure 2. Comparison of junctions between the LSC, SSC and IR regions among six newly assembled Aerides plastomes.
Figure 2. Comparison of junctions between the LSC, SSC and IR regions among six newly assembled Aerides plastomes.
Ijms 24 12473 g002
Figure 3. Summary of the simple sequence repeats (SSR) across the Aerides species. (A) Variation in repeat abundance and type in six plastomes. (B) Number of SSRs for each Aerides species by SSR unit size and number of SSRs for each Aerides species by location in IR, LSC and SSC.
Figure 3. Summary of the simple sequence repeats (SSR) across the Aerides species. (A) Variation in repeat abundance and type in six plastomes. (B) Number of SSRs for each Aerides species by SSR unit size and number of SSRs for each Aerides species by location in IR, LSC and SSC.
Ijms 24 12473 g003
Figure 4. Plastome comparison of six species of Aerides using a progressive MAUVE algorithm.
Figure 4. Plastome comparison of six species of Aerides using a progressive MAUVE algorithm.
Ijms 24 12473 g004
Figure 5. Global alignment of six Aerides plastomes using mVISTA with A. falcata as reference. The y-axis shows the coordinates between the plastomes.
Figure 5. Global alignment of six Aerides plastomes using mVISTA with A. falcata as reference. The y-axis shows the coordinates between the plastomes.
Ijms 24 12473 g005
Figure 6. Sliding window test of nucleotide diversity (Pi) in the Aerides plastomes. (A) The nucleotide diversity of the complete plastome and four mutation hotspot regions (Pi > 0.06) were annotated. (B) The nucleotide diversity of 68 protein coding sequences showing two mutation hotspot regions (Pi > 0.03). The window size was set to 100 bp and the sliding window size was 25 bp. x -axis, position of the midpoint of a window; y -axis, Pi values of each window.
Figure 6. Sliding window test of nucleotide diversity (Pi) in the Aerides plastomes. (A) The nucleotide diversity of the complete plastome and four mutation hotspot regions (Pi > 0.06) were annotated. (B) The nucleotide diversity of 68 protein coding sequences showing two mutation hotspot regions (Pi > 0.03). The window size was set to 100 bp and the sliding window size was 25 bp. x -axis, position of the midpoint of a window; y -axis, Pi values of each window.
Ijms 24 12473 g006
Figure 7. Phylogenetic tree of Aerides and the other 16 Aeridinae species based on the complete plastome data. Numbers near the nodes are bootstrap percentages and Bayesian posterior probabilities (BSML left, BSMP middle and PP right). A dash (-) indicates that a node is inconsistent between the topology of the MP/ML trees and the Bayesian tree. An asterisk (*) indicates that the node is 100 bootstrap percentage or 1.00 posterior probability.
Figure 7. Phylogenetic tree of Aerides and the other 16 Aeridinae species based on the complete plastome data. Numbers near the nodes are bootstrap percentages and Bayesian posterior probabilities (BSML left, BSMP middle and PP right). A dash (-) indicates that a node is inconsistent between the topology of the MP/ML trees and the Bayesian tree. An asterisk (*) indicates that the node is 100 bootstrap percentage or 1.00 posterior probability.
Ijms 24 12473 g007
Table 1. Characteristics of the complete plastomes of the Aerides lineages.
Table 1. Characteristics of the complete plastomes of the Aerides lineages.
Species Name Size (bp)GC Content (%)LSC Size in bp (%)IR Size in bp (%)SSC Size in bp (%)Total Number of GeneProtein-Encoding GenetRNA GenerRNA GeneNumber of ndh Fragment
Aeridescrassifolia147,24436.883,913 (57.0)25,717 (34.9)11,897 (8.1)120743886
A. falcata148,34736.885,570 (57.7)25,727 (34.7)11,323 (7.6)120743889
A. lawrenceae148,29036.885,394 (57.6)25,714 (34.7)11,468 (7.7)120743888
A. odorata148,21636.885,293 (57.5)25,706 (34.7)11,511 (7.8)120743889
A. quinquevulnera148,39136.785,420 (57.6)25,773 (34.7)11,425 (7.7)120743888
A. rosea148,11036.785,373 (57.6)25,852 (34.9)11,033 (7.5)120743887
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chen, J.; Wang, F.; Zhou, C.; Ahmad, S.; Zhou, Y.; Li, M.; Liu, Z.; Peng, D. Comparative Phylogenetic Analysis for Aerides (Aeridinae, Orchidaceae) Based on Six Complete Plastid Genomes. Int. J. Mol. Sci. 2023, 24, 12473. https://doi.org/10.3390/ijms241512473

AMA Style

Chen J, Wang F, Zhou C, Ahmad S, Zhou Y, Li M, Liu Z, Peng D. Comparative Phylogenetic Analysis for Aerides (Aeridinae, Orchidaceae) Based on Six Complete Plastid Genomes. International Journal of Molecular Sciences. 2023; 24(15):12473. https://doi.org/10.3390/ijms241512473

Chicago/Turabian Style

Chen, Jinliao, Fei Wang, Chengyuan Zhou, Sagheer Ahmad, Yuzhen Zhou, Minghe Li, Zhongjian Liu, and Donghui Peng. 2023. "Comparative Phylogenetic Analysis for Aerides (Aeridinae, Orchidaceae) Based on Six Complete Plastid Genomes" International Journal of Molecular Sciences 24, no. 15: 12473. https://doi.org/10.3390/ijms241512473

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

Article Metrics

Back to TopTop