Introduction

Commelinaceae Mirb., commonly known as the dayflower and spiderwort family, are the largest family of Commelinales Mirb. ex Bercht. & J. Presl, including four other families: Haemodoraceae, Hanguanaceae, Philydraceae, and Pontederiaceae, [1, 2]. Commelinaceae consist of 41 genera and approximately 730 species, widely distributed in both the Old and New Worlds, except in Europe [2,3,4]. Genus Callisia Loefl. and Tradescantia L. emend. M. Pell. are commonly used as ornamentals, while Commelina L. is used as vegetables and more commonly known as troublesome weeds. The species of Commelinaceae are usually succulent herbs with closed leaf-sheaths, raphide-canals, and three-celled glandular microhairs [3, 4]. Additionally, flowers of Commelinaceae are mainly insect-pollinated, have short blooming times, and lack any kind of nectaries [5, 6]. The flowering unit (inflorescence) of Commelinaceae is a many-branched thyrse, with each branch generally consisting of a many-flowered cincinnus. The cincinni can sometimes be 1-flowered or, more rarely, the whole inflorescence can be reduced to a single flower [4, 7].

Previous classifications of Commelinaceae emphasized floral and anatomical characters. In the first classification, Commelinaceae were divided into two tribes, Commelineae and Tradescantieae, based on the number of stamens and their fertility [8]. Then, Bruckner [9] used flower symmetry, and Pichon [10] used anatomical characters to exclude Cartonema R. Br. from Commelinaceae. In 1966, 15 genera of Commelinaceae were defined using various floral characters [11]. In the current classification, Commelinaceae were divided into two subfamilies, Cartonematoideae (Pichon) Faden ex G. C. Tucker and Commelinoideae Faden & D. R. Hunt, based on the presence of raphide-canals and glandular microhairs [4]. Cartonematoideae consists of two genera (Cartonema and Triceratella Brenan), whereas Commelinoideae includes 39 genera, divided into two tribes, Commelineae (Meisn.) Faden & D. R. Hunt and Tradescantieae (Meisn.) Faden & D. R. Hunt, based on palynological characters. The latter tribe was arranged into seven subtribes based on morphological and cytological characters [4, 12]. However, it is difficult to interpret relationships among genera due to their morphological variation. The morphology-based phylogeny was highly homoplasy and incongruent with the current classification [13]. In order to clarify the relationships within Commelinaceae, several phylogenetic studies have been conducted [14,15,16,17,18,19,20]. Based solely on the plastidial rbcL marker, Cartonema was recovered in a basal clade, and both Commelineae and Tradescantieae were monophyletic, except for the position of Palisota Rchb., which had low support values [15]. Furthermore, the plastidial ndhF suggested that subtribe Tradescantiinae was paraphyletic, whereas Thyrsantheminae and Dichorisandrinae were polyphyletic [16]. Combined data of nuclear 5S NTS and plastid trnL-F regions resulted in a well-supported relationship between Commelineae and Tradescantieae. However, the position of Palisota and Spatholirion Ridl. were ambiguous [17].

Chloroplast genome or plastid genome (cpDNA) is highly conserved and has a typical quadripartite structure containing a large single copy (LSC) and a small single copy (SSC) separated by two inverted repeats (IRs). The size of cpDNA ranges from 19,400 bp (Cytinus hypocistis) to 242,575 bp (Pelargonium transvaalense) and generally contains 120–130 genes, which perform important roles in photosynthesis, translation, and transcription [21, 22]. The rapid development of next-generation sequencing (NGS) has enabled many studies with high-quality complete plastid genomes with raw reads at low costs. Due to its conserved characteristics, chloroplast protein-coding genes were used to reconstruct the phylogenetic relationships in other monocot groups [23,24,25]. Furthermore, these data are useful to infer biogeography, molecular evolution, and age estimation [26,27,28]. The aims of this study are to 1) explore the genome evolution in Commelinaceae subfamily Commelinoideae through analyses of sequence variation, and gene content and order; 2) find latent phylogenetically informative genes through high nucleotide diversity; 3) reconstruct the phylogenetic relationships among members of Commelinoideae with other monocot groups using 77 chloroplast protein-coding genes data, especially the relationships among the six subtribes of Tradescantieae.

Results

Chloroplast genome assembly and annotation

We completed 15 new plastid genomes in this study listed in Table 1 through 9 to 21 million raw reads for each species (Fig. S1, Table S1). A total of 16 plastid genomes, including Belosynapsis ciliata, exhibit the typical quadripartite structure containing LSC and SSC regions separated by two inverted repeats (Fig. 1). Plastid genome sequences of Murdannia edulis and B. ciliata are over 170 kb in length whereas that of Commelina communis is 160,116 bp in length (Table 1). In addition, M. edulis has the lowest GC content (34.4%), whereas Palisota barteri has the highest GC content (36.2%) (Table 1). The highest length difference (about 8801 bp) was observed in the LSC region, between B. ciliata and C. communis. GC content in the SSC region was about 3.4% between Dichorisandra thyrsiflora and M. edulis (Table 1). Plastid genomes of Commelinoideae have 131 genes, of which 111 are unique, and 20 are duplicated in the IR regions (Table 2), except for the rpl22 gene, which was not duplicated in tribe Tradescantieae. There are 77 protein-coding genes (CDS), 30 transfer RNA (tRNA) genes and four ribosomal RNA (rRNA) genes in examined Commelinoideae taxa (Table 2). In these genes, three CDS (rps12, clpP, and ycf3) have two introns, while nine CDS (atpF, ndhA, ndhB, petB, petD, rpl2, rpl16, rpoC1, and rps16) and six tRNA (trnK-UUU, trnG-UCC, trnL-UAA, trnV-UAC, trnI-GAU, and trnA-UGC) have one intron (Table 2). The rps12 gene was trans-spliced, which has the 5′ exon in the LSC region and the 3′ exon and intron in the IR regions. Three pseudogenes (accD, rpoA, and ycf15) were identified from all Commelinoideae species, one (ycf15) of which was duplicated in the IR regions (Table 2). These three genes contained several internal stop codons due to insertions and deletions, thus are identified as pseudogenes. Also, we identified ndhB as a pseudogene in two species (Pollia japonica and Rhopalephora scaberrima) due to point mutation.

Table 1 Comparison of the features of plastomes from 16 genera of Commelinaceae
Fig. 1
figure 1

Representative chloroplast genome of Commelinaceae. The colored boxes represent conserved chloroplast genes. Genes shown inside the circle are transcribed clockwise, whereas genes outside the circle are transcribed counter-clockwise. The small grey bar graphs inner circle shows the GC contents

Table 2 Gene composition within chloroplast genomes of Commelinaceae species

Comparative chloroplast genome structure and nucleotide diversity

The aligned data of whole plastid genomes showed high similarities in coding genes, and high variations in non-coding genes (Fig. 2). We found several genome structure variations among Commelinoideae species. M. edulis and Streptolirion volubile had one inversion from rbcL to psaI intergenetic spacer (approximately 3 kb) and petN to trnE-UUC (approximately 2.8 kb), respectively. Amischotolype hispida and B. ciliata had two large inversions from trnV-UAC to rbcL and psbJ to petD about approximately 5 kb and 16 kb, respectively. The IR-SSC boundary was similar among species of Commelinoideae (Fig. 3). All plastid genomes have an incompletely duplicated ycf1 gene in the IRB-SSC junctions. We also found an expansion of IR regions in tribe Commelineae, which resulted in the duplication of the rpl22 genes (Fig. 3).

Fig. 2
figure 2

Plots of percent sequence identity of the chloroplast genomes of 16 Commelinaceae species with Hanguana malayana as a reference. The percentage of sequence identities was estimated, and the plots were visualized in mVISTA

Fig. 3
figure 3

Comparisons of LSC, SSC, and IR regions boundaries between 16 Commelinaceae species

We analyzed nucleotide divergences of CDS, tRNA, and rRNA to explain variant characteristics among the 16 Commelinoideae plastid genomes (Fig. 4, Table S3). Nucleotide diversity (Pi) for each CDS ranges from 0.00427 (psbL) to 0.09543 (ycf1) with an average of 0.03473. Nine CDS (rps3, ndhG, ndhD, ccsA, rps15, rpl32, ndhF, matK, and ycf1) have remarkably high values (Pi > 0.05) and seven CDS (psbL, rpl23, rps19, ndhB, rpl2, rps7, rps12) have low values (Pi < 0.01; Fig. 4). Compared with tribe Tradescantieae, Commelineae have higher values in 59 out of 77 CDS (Fig. 4). The rpl22 gene has the highest difference of values between Commelineae (Pi = 0.01499) and Tradescantieae (Pi = 0.04655). In the tRNA and rRNA regions, Pi values range from 0 (trnT-UGU, trnH-GUG, trnV-GAC, and trnI-GAU) to 0.02697 (trnQ-UUG), with an average of 0.006. Commelineae has the highest value in the trnL-UAA (Pi = 0.02941), while Tradescantieae has no value in this gene. We tried to find latent phylogenetically informative genes for the Commelinoideae by checking individual CDS with high values (Pi > 0.045) and over 500 bp length. Ten CDS (ndhH, rpoC2, ndhA, rps3, ndhG, ndhD, ccsA, ndhF, matK, and ycf1) were checked with a ML analysis and compared positions among 16 genera of Commelinoideae (Fig. 5). Four CDS (ndhH, rpoC2, matK, and ycf1) have similar topology in Commelinoideae even though the other monocot groups were unclear.

Fig. 4
figure 4

Nucleotide diversity (Pi) values in protein-coding genes, tRNA, and rRNA in 16 Commelinaceae species. The dashed lines are the borders of the LSC, IR and SSC regions

Fig. 5
figure 5

The Maximum Likelihood tree of 42 monocots inferred from 77 chloroplast protein-coding genes. Numbers indicate support (maximum parsimony bootstrap (PBP)/maximum likelihood bootstrap (MBP)/posterior probability (PP)). Only support under PBP = 90/MBP = 100/PP = 1.00 is shown. The dashes “-” indicate incongruence between MP and ML/BI trees

Phylogenetic analysis

The aligned 77 chloroplast protein-coding genes had 65,481 bp, of which 16,380 were parsimony informative. The MP analysis produced single most-parsimonious tree (tree length = 72,586, CI = 0.488, RI = 0.626). The tree topologies of the MP, ML, and BI analyzes were found to be congruent with 100% bootstrap (PBP, MBP) values and 1.00 Bayesian posterior probabilities (PP) in almost all nodes, except for Palisota, which was unresolved in MP analysis (not shown) (Fig. 5). The result suggested that Palisota was sister to the group consisting of the rest of Commelinoideae (Fig. 5). In Tradescantieae, Streptoliriinae was positioned at the basal node. Then, Dichorisandrinae divided into two clades ((Dichorisandra, Siderasis), (Cochliostema, Geogenanthus)) with relatively low support values in both MP and ML analysis (PBP = 77, MBP = 84, PP = 1) (Fig. 5). Among the remaining three subtribes, where two clades ((Coleotrypinae and Cyanotinae), (Tradescantiinae)) were formed with high support values (PBP = 100, MBP = 100, PP = 1), respectively (Fig. 5).

Discussion

Chloroplast genome structure

In this study, we completed 15 new plastid genomes of Commelinoideae taxa (Table 1). Plastid genomes have typical quadripartite structures, including LSC, SSC and two IR regions. Plastid genomes of Commelinoideae have variable total length and GC content. The LSC and SSC regions are relatively longer and higher AT-content than the IR region (Table 1). The functions of AT-rich sequences in the plastid genome were known as enhancing gene transfer success by making stable transcripts [29]. However, AT-rich sequences caused structural variations like inversions by their weak hydrogen bonding. In this study, we identified small to large inversions in four species (Fig. 2). There is one inversion in M. edulis and S. volubile, and two inversions in A. hispida and B. ciliata (Fig. 2). Inversions are known as common genomic rearrangement events and provide informative infrageneric relationships. In previous studies, inversions were caused by microhomology-driven recombination via short repeats and suggested the monophyly of tribe Desmodieae (Fabaceae) [30]. Our results also suggest that both Amischotolype and Belosynapsis have two large inversions in the same loci and formed a clade sister to subtribe Dichorisandrinae (Fig. 5).

We identified an IR expansion in members of Commelineae (Commelina, Murdannia, Pollia, and Rhopalephora). Four species have one more rpl22 gene, which is duplicated in the terminal IR regions (Fig. 3). Although IR expansion affected gene composition, the IR region’s total length is similar among 16 Commelinoideae species. IR expansion and contraction are important events in several families. In Ranunculaceae, IR expansion was detected as a synapomorphy of tribe Anemoneae [31]. Likewise, IR expansion lent further support to the relationship between two subfamilies Ehrhartoideae and Pooideae (Poaceae) [32]. This event also may be phylogenetically informative in Commelinoideae since only members of tribe Commelineae sharing this genome variation (Fig. 5).

Within Commelinoideae plastid genomes, three protein-coding genes (accD, rpoA, and ycf15) were classified as pseudogenes (Fig. S2). The ycf15 gene has several abnormal stop codons caused by insertions and deletions (indel) of bases similar to other monocots. We also identified that all examined species have indels at the frontal part of the accD gene (until 400 bp) and the terminal part of the rpoA gene (after 700 bp; Fig. S2). The accD gene, encoding the beta-carboxyl transferase subunit of acetyl-CoA carboxylase, is found in most flowering plants and synthesizes fatty acids within the chloroplast. It was suggested as an essential gene associated with maintaining chloroplast structure [33]. However, it was reported as a gene loss or pseudogenization in Acoraceae and Poaceae [34, 35]. Recent studies suggested that the accD gene was found to be nuclear originated in several eudicots [36, 37]. The rpoA gene, which encodes the alpha subunit of RNA polymerase, is also found in most flowering plants but was recorded to having been lost in the chloroplast genome of mosses [38]. In one species, Physcomitrella patens (Funariaceae), the rpoA gene was transferred to the nucleus [39]. We need further studies to confirm whether these two genes have been transferred to the nucleus or not in Commelinaceae. We identified that the pseudogened accD and rpoA only appeared in Commelinoideae among Commelinales. It might be a specific character of gene composition in Commelinales. We also found a point mutated base in the third codon of the ndhB gene in P. japonica and R. scaberrima, which formed a clade in this study (Fig. 5).

We measured the nucleotide diversity of CDS, tRNA, and rRNA to identify the genetic divergence between 16 Commelinoideae plastid genomes. We found that the CDS in the IR regions have lower nucleotide diversity than that of the LSC and SSC regions (Fig. 4). This result has also been identified in the other monocots [40,41,42]. It is possibly attributed to a copy correction of the IR regions via gene conversion [43]. Especially, we can see this result in the rpl22 gene. Only Commelineae species present a duplicated rpl22 gene due to the above-mentioned IR expansion, while the remaining 12 taxa have one gene in the LSC or LSC-IR junction (Fig. 3). Difference of nucleotide diversity in this gene between Commelineae (Pi = 0.015) and Tradescantieae (Pi = 0.0466) is 0.0316. It might be phylogenetically useful information for Tradescantieae only.

Implications of plastomes data for phylogenetic reconstructions

The first phylogenetic analysis of Commelinaceae based on rbcL marker revealed a relationship of 32 species representing 30 genera of Commelinaceae [15]. Cartonematoideae was in a basal clade, sister to Commelinoideae and all remaining species [15]. Aside from Palisota, Commelinoideae was divided into two tribes, Commelineae and Tradescantieae, with low bootstrap support values due to insufficient information [15]. Although several phylogenetic studies were conducted, the relationships between the genera of Commelinaceae have remained unresolved. The position of Palisota had been problematic, being recovered as: 1) sister to all genera of Commelinoideae with high bootstrap values [15]; 2) low bootstrap support value with other members of Tradescantieae [16]; or 3) sister to tribe Commelineae [19]. Subtribe Streptoliriinae was recovered as sister to tribe Commelineae in the trnL-trnF analysis [17]. Finally, subtribe Dichorisandrinae seemed polyphyletic in the previous studies [15, 16, 19, 44]. These results are most likely due to limited taxon sampling and/or used few informative genetic markers. The aligned 77 chloroplast coding genes in this study suggest a more well-supported relationship between the genera (Fig. 5). We identified that Commelinoideae divided into two clades, tribe Commelineae and Tradescantieae, with high support values (Fig. 5). However, Palisota, which belongs to Tradescantieae in the current classification [3], is recovered by us as sister to tribe Commelineae (Fig. 5). The ML and BI results present high support values, even though this relationship is unresolved in MP (data not shown). Compared with the current classification, it seems like that subsidiary cells in the stomata and exine morphology are homoplastic to divide two tribes in Commelinoideae [3]. In the Commelinaceae, Palisota is unusual genus for its unique morphological characters like a fleshy berry as a fruit, stamen and staminode arrangement, complex reproductive system, and a basic number of chromosome (x = 20) [13, 45]. Zygomorphic androecium character places Palisota within the Commelineae clade in the morphological cladistic analysis [13]. However, this character is also homoplastic within Commelinoideae. Further research is needed to suggest appropriate characters for Palisota. The four species of Commelineae sampled by us are recovered with a relationship similar to previous studies [15]: (Murdannia, (Commelina, (Pollia, Rhopalephora))). Within Tradescantieae, Streptoliriinae diverged first, followed by Dichorisandrinae divided into two clades with relatively low support values (PBP = 77/MBP = 84/PP = 1) (Fig. 5). The clade composed by Coleotrypinae and Cyanotinae is recovered following the diversion of subtribe Dichorisandrinae, which is sister to Tradescantiinae sensu Pellegrini, [46]. Interestingly, the Asian and African subtribe Coleotrypinae and Cyanotinae were nested well within the New World subtribes (Fig. 5). This result is similar to previous studies and support the hypothesis that one shift from the Old World to the New World followed by dispersal back to the Old World [15, 16].

Conclusions

Our study revealed genome structural characteristics, nucleotide diversity, improved relationships between genera using 15 newly complete chloroplast genomes of Commelinoideae. Compared with other Commelinales, we found two characteristic pseudogenes in all members of Commelinoideae, which might be a synapomorphy within the order. We also reconstruct the phylogenetic relationships using 77 chloroplast protein-coding genes. Although not being able to address the Commelinaceae as a whole, due to not sampling of subfamily Cartonematoideae, we have been able to recover well-supported relationships for the taxa of Commelinoideae, especially between the subtribes of Tradescantieae. One interesting result was that Palisota (subtribe Palisotinae) is more closely related to tribe Commelineae than the remaining members of tribe Tradescantieae. In the current classification, Palisota is a member of Tradescantieae according to the number of subsidiary cells in stomata and pollen exine lacking spines [3]. However, it seems like that these characters are homoplastic, so we need a further study to suggest appropriate characters for two tribes in Commelinoideae. We resolved the ambiguous position of Streptoliriinae which was placed with Commelineae group [17]. Also, Dichorisandrinae was monophyletic in this study which was polyphyletic in the previous studies [15, 16, 19]. Four genes (ndhH, rpoC2, matK, and ycf1) are congruent with the tree estimated from the 77 protein-coding genes. These genes will be helpful to reconstruct relationships of the whole Commelinaceae in the future. Future studies might use the information of chloroplast genomes, relationship between genera to define new classification of Commelinaceae that we provided in this study. These data will make sure the historical biogeography and genome evolution of Commelinaceae.

Materials and methods

Taxon sampling and DNA extraction

Fresh leaf samples were collected in the field and dried directly with silica gel in room temperature until DNA extraction (Table 1). The samples covered four out of 14 genera in tribe Commelineae and 11 out of 25 genera, including six subtribes of tribe Tradescantieae. We prepared the voucher specimens for all used samples and deposited them in the Gachon University Herbarium (GCU) with their accession numbers. We used a modified CTAB method to extract total DNA [47] and checked quality using a spectrophotometer (Biospec-nano; Shimadzu) and assessed by agarose gel electrophoresis.

Genome sequencing, assembly, and annotation

Next-generation sequencing (NGS) was conducted using the Illumina MiSeq sequencing system (Illumina, Seoul, Korea). We imported NGS raw data and trimmed the ends limited to a 5% error probability to remove poor quality reads using Geneious prime 2020.1.2 [48]. Then, we performed ‘map to reference’ using the Hanguana malayana chloroplast genome (GenBank accession = NC_029962.1) as a reference to isolate cpDNA reads. De novo assembly was implemented to reassemble reads using Geneious prime 2020.1.2 [48]. We used newly generated sequences as a reference to reassemble raw reads. We repeated this step until quadripartite structures were completed. Gaps were filled by Sanger sequencing using specific primers. Gene content and order were annotated using H. malayana as a reference using 80% similarity to identify genes in Geneious. All tRNAs were checked by tRNAScan-SE [49] with default search mode. Illustrations of plastomes were produced using OGDraw [50].

Comparative genome analysis

We compared genome structure, size, gene content across all 16 species including B. ciliata (GenBank accession = MK133255.1), subtribe Cyanotinae. The GC content was calculated and compared using Geneious. The whole chloroplast genome sequences of Commelinoideae species were aligned using MUSCLE embedded in Geneious and visualized using LAGAN mode in mVISTA [51, 52]. For the mVISTA plot, we used the annotated cpDNA of H. malayana as a reference. We also examined the nucleotide diversity (Pi) of chloroplast protein-coding genes, transfer RNA genes and ribosomal RNA genes among the 16 Commelinoideae species through a sliding window analysis using DnaSP v. 6.0 [53]. For the sequence divergence analysis, we applied the window size of 100 bp with a 25 bp step size. The IR and SC boundaries of the 16 Commelinoideae species were compared and illustrated using IRscope [54].

Phylogenetic analysis

A total of 42 chloroplast genome sequences (including 15 new chloroplast genomes of Commelinoideae) were used (Table S2). We extracted 77 protein-coding genes and aligned them using the MUSCLE embedded in Geneious prime 2020.1.2 [48]. For the data set, Acorus calamus (Acoraceae) was designated as an outgroup. We performed maximum parsimony (MP), maximum likelihood (ML), and Bayesian inference (BI) to infer relationships of Commelinoideae and related taxa. The MP analyses were carried out in PAUP* v4.0a [55] with all characters equally weighted and unordered. Gaps were treated as missing data. Searches of 1000 random taxon addition replicates used tree-bisection-reconnection (TBR) branch swapping, and MulTrees permitted ten trees to be held at each step. Bootstrap analyses (PBP, parsimony bootstrap percentages, 1000 pseudoreplicates) were conducted to examine internal support with the same parameters. We used jModelTest version 2.1.7 [56, 57] to find the best model with Akaike’s information criterion (AIC) before running the ML and BI analyses. The GTR + I + G was the best model for the concatenated data sets. We used the IQ-TREE web server (http://iqtree.cibiv.univie.ac.at/) to make the ML searches [58]. Support value (MBP, mean bootstrap percentage) was calculated with 1000 replicates of ultrafast bootstrap [59]. MrBayes v3.2.7 [60] was used for BI analyses. Two simultaneous runs were performed starting from random trees for at least 1,000,000 generations. One tree was sampled every 1000 generations. In total, 25% of trees were discarded as burn-in samples. The remaining trees were used to construct a 50% majority-rule consensus tree, with the proportion bifurcations found in this consensus tree given as posterior probability (PP) to estimate the robustness of half of the BI tree. The effective sample size values (ESS) were then checked for model parameters (at least 200). The phylogenetic trees were edited using FigTree v1.4.4 program [61].