Next Article in Journal
Rice RBH1 Encoding A Pectate Lyase is Critical for Apical Panicle Development
Next Article in Special Issue
New Insight into Taxonomy of European Mountain Pines, Pinus mugo Complex, Based on Complete Chloroplast Genomes Sequencing
Previous Article in Journal
Extensive Analysis of miRNA Trimming and Tailing Indicates that AGO1 Has a Complex Role in miRNA Turnover
Previous Article in Special Issue
A Small Number of Gametophytes with Gametangia and Stunted Sporophytes of Antrophyum obovatum Baker (Pteridaceae): The Suppression of Functional Sporophyte Production by Prezygotic and Postzygotic Sterility
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Complete Chloroplast Genome of Abutilon fruticosum: Genome Structure, Comparative and Phylogenetic Analysis

by
Dhafer A. Alzahrani
Department of Biological Sciences, Faculty of Sciences, King Abdulaziz University, P.O. Box 80203, Jeddah 21589, Saudi Arabia
Plants 2021, 10(2), 270; https://doi.org/10.3390/plants10020270
Submission received: 24 November 2020 / Revised: 23 January 2021 / Accepted: 24 January 2021 / Published: 30 January 2021
(This article belongs to the Special Issue Plant Evolution, Systematics, and Chloroplast Genome)

Abstract

:
Abutilon fruticosum is one of the endemic plants with high medicinal and economic value in Saudi Arabia and belongs to the family Malvaceae. However, the plastome sequence and phylogenetic position have not been reported until this study. In this research, the complete chloroplast genome of A. fruticosum was sequenced and assembled, and comparative and phylogenetic analyses within the Malvaceae family were conducted. The chloroplast genome (cp genome) has a circular and quadripartite structure with a total length of 160,357 bp and contains 114 unique genes (80 protein-coding genes, 30 tRNA genes and 4 rRNA genes). The repeat analyses indicate that all the types of repeats (palindromic, complement, forward and reverse) were present in the genome, with palindromic occurring more frequently. A total number of 212 microsatellites were identified in the plastome, of which the majority are mononucleotides. Comparative analyses with other species of Malvaceae indicate a high level of resemblance in gene content and structural organization and a significant level of variation in the position of genes in single copy and inverted repeat borders. The analyses also reveal variable hotspots in the genomes that can serve as barcodes and tools for inferring phylogenetic relationships in the family: the regions include trnH-psbA, trnK-rps16, psbI-trnS, atpH-atpI, trnT-trnL, matK, ycf1 and ndhH. Phylogenetic analysis indicates that A. fruticosum is closely related to Althaea officinalis, which disagrees with the previous systematic position of the species. This study provides insights into the systematic position of A. fruticosum and valuable resources for further phylogenetic and evolutionary studies of the species and the Malvaceae family to resolve ambiguous issues within the taxa.

1. Introduction

The genus Abutilon Mill. [1,2], whose members are widely distributed in tropical and subtropical regions [3], is considered as one of the largest genera of Malvaceae [4,5], with ca. 200 accepted species in all continents except Antarctica [3]. The systematic position of some of the taxa in the genus is still not clear [6]; hence, it is the most difficult genera in the Malvaceae with a need for critical systematic studies. The genus is distinguished from sister taxa by the presence of an endoglossum and dorsal dehiscence and a lack of an epicalyx [7]. Members of the genus received a large amount of attention due to their medicinal and economic value [8]. In addition, parts of the plant including the flower, bark, fruit and seeds are reported to contain some phytoconstituents that are responsible for their biological activity [9]. The plants contain no toxins; therefore, many researchers are focusing on them [10,11,12]. Abutilon fruticosum is reported to have medicinal values; all parts of the plant are used in the treatment of various ailments including ulcers, leprosy, inflammation of the bladder, piles, bronchitis, rheumatism and jaundice [10,13,14]. The fiber from the plant is used as a substitute of jute [8]. Despite its importance, the phylogenetic position of the genus is still not clear, and its complete chloroplast genome had not yet been reported until this study. The phylogenetic position of the species within the genus Abutilon and the family Malvaceae has not been reported. According to the available literature, there has been no study that tried to address the phylogenetic position of the species at any taxonomic level where the species belong. As the plant is of medicinal and economic importance, there is a need to report its plastome sequence for conservation purposes, for authentication and to resolve its systematic position.
The family Malvaceae (commonly known as mallows), to which Abutilon belongs, is one of the largest families of angiosperm. It comprises ca. 4225 identified species distributed in 244 genera [15,16]. Until this study, only a few complete cp genomes of genera of the family Malvaceae had been known: Gossypium, Abelmoschus, Hibiscus, Firmiana, Bombax, Theobroma, Craigia and Talipariti.
Genetic information is a reliable source of understanding evolutionary relationships among species at various taxonomic levels (categories, ranks). The genetic information in the chloroplast genome contains sufficient information for comparative analysis and studies of species diversification because of the presence of functional genes which have a vital role in plant cells [17]. The chloroplast organelle takes part in carbon fixation and photosynthesis in plants [18]. Among the three genomes present in plants, the chloroplast is the most conserved. In general, chloroplast DNA has a double-stranded, circular and, typically, four-segment structure which includes a large single copy, a small single copy and a pair of inverted repeats [19]. However, recently, Oldenburg and Bendich [20] demonstrated that some plants’ ptDNA is of a linear structure. The cp genome is uniparentally inherited and also non-recombinant, and these characteristics make it highly conserved in structure and content [21]. However different kinds of mutations do occur [22], which, as a result, lead to sequence divergence among species and could be used to study evolutionary relationships in plants [23]. Despite the importance of the plastome in modern taxonomy, only chloroplast genomes of few genera in the whole Malvaceae family have been reported.
Here, the whole genome data of Abutilon fruticosum were obtained for the first time using Illumina sequencing technology, and the complete chloroplast genome was assembled using NOVOPlasty3.8.1. The features of the cp genomes were analyzed and compared with other Malvaceae species to provide resources for identification and evolutionary, phylogenetic and population genetics studies of the taxon.

2. Results and Discussion

2.1. Characteristics of A. fruticosum Chloroplast Genome

Studies have shown that the complete chloroplast genome of angiosperms is highly conserved in content and structural organization; however, contraction and expansion do occur [24,25]. The complete plastome sequence of A. fruticosum has a circular and quadripartite structure with a total length of 160,357 bp. The plastome has four distinct regions which are a small single copy (SSC), a large single copy (LSC) and a pair of inverted repeats (IRa and IRb) which separates the SSC and LSC (Figure 1; GeneBank MT772391). The gene coding region is 81,205 bp in length which constitutes 50.64% of the genome, and the remaining 69,517 bp is the non-coding region which includes introns and intergenic spacers (43.35%). The length of the SSC, LCS, IRa and IRb is 20,031, 89,034, 25,646 and 25,646 bp, respectively. The LSC and SSC regions possess a GC content of 34.76 % and 31.97 %, respectively, while the inverted repeats IRa and IRb have 42.9 % (Table 1). The percentage of GC in the inverted repeat regions is found to be higher than the large and small single regions.
The complete chloroplast genome of A. fruticosum contained a total of 133 genes, where 114 genes out of the 133 are unique and are present in the single copy regions; 18 genes are duplicated in the inverted repeat region which includes 7 protein-coding genes, 4 rRNAs and 7 tRNAs. There are 80 protein-coding genes, 4 rRNAs and 30 tRNAs in the plastome (Table 2 and Figure 1). The inverted repeat region contained seven protein-coding genes, seven tRNA and four rRNA, while in the single copy region, the LSC contained 62 protein-coding genes and 22 tRNA genes; the rest of the 12 protein-coding genes and 1 tRNA are located within the SSC region. Almost all the protein-coding genes start with the ATG codon that codes for methionine, whereas some of the genes contained alternative start codons such as ATC, GTG and ACG; this is common in most chloroplast genomes of flowering plants (angiosperms) [26,27,28].
The A. fruticosum chloroplast genome is found to contain an intron in some of the coding genes, such as in other chloroplast genomes of flowering plants [26,27]. Among the 114 coding genes in A. fruticosum, 17 contain introns (Table 3). Out of the 17 genes with an intron, 11 are protein-coding genes and six are tRNAs. The LSC region contains introns in 11 genes and the IR region contains introns in 5 genes, while the SSC region contains introns in only 1 gene. Two genes, ycf3 and clpP, possess two introns and the other 15 genes have only one intron. trnK-UUU has the longest intron, while accD has the shortest intron (Table 3).
Codon usage compares the frequencies of each codon that codes for a particular amino acid [29]. Codons are used in transmitting genetic information because they are the building blocks of proteins [30]. Codon usage is a factor shaping the evolution of chloroplast genomes because of bias in mutation [28], and it varies across different species [31]. The frequency of the codon present in the chloroplast genome was computed using the nucleotide sequence of protein-coding genes and tRNA genes 84,048 bp. The relative synonymous codon usage (RSCU) of the genes in the genome is presented in Table 4. The results show that genes in the plastome are encoded by 27,967 codons. Codons that code for leucine appear more frequently in the genome 2957 (10.57%) (Figure 2). Meanwhile, codons coding for cysteine are the least with 325 (1.16%) in the genome. Guanine and cytosine endings are found to be more frequent than their counterparts adenine and thymine; this is not the case in other plastome sequences [32,33,34]. The result of the analysis (Table 4) shows that codon usage bias is low in the chloroplast genome of A. fruticosum. The RSCU values of 30 codons were greater than 1 and all of them have an A/T ending, while for 31 codons, the values were less than 1 and are all of the G/C ending. Only two amino acids, tryptophan and methionine, have an RSCU value of 1 and therefore they are the only amino acids with no codon bias.
RNA editing is a set of processes including the insertion, deletion and modification of nucleotides that alters the DNA-encoded sequence [35], which is a way to create transcript and protein diversity [36]. Some chloroplast RNA editing sites are preserved in plants [37]. The program PREP suite was used to predict the RNA editing sites in the chloroplast genome of A. fruticosum. The first nucleotide of the codon was used in all the analyses. The result of the analysis shows that most of the conversions in the codons are from serine to leucine (Table 5).
Generally, 50 editing sites in the genome were revealed which were distributed within 22 protein-coding genes. The gene ndhB has the highest number of editing sites with 12 sites, and this is consistent with previous studies [38,39,40]. One gene with eight editing sites is ndhD and other genes with a high number of editing sites are ndhF and rpoB having four and matK with three editing sites. The genes accD, atpA, ndhA, ndhG, rpoA, rpoC1, rpoC2 and rps2 have two editing sites.
The following genes: atpF, atpI, ccsA, clpP, petB, psbF, rpl20, rps8 and rps14, with one editing site, have the lowest number of editing sites. Conversions of proline to serine were observed, which involve the change of amino acids in the RNA editing site from a nonpolar to a polar group. Genes such as atpB, petD, petG, petL, psaB, psaI, psbB, psbE, psbL, rpl2, rpl23, rps16 and ycf3 do not possess predicted RNA sites in their first codon.

2.2. Repeat Analyses

2.2.1. Long Repeats

The program REPuter was used to identify long repeat sequences present in the A. fruticosum chloroplast genome. It was discovered from the results that all four types of long repeats (palindromic, forward, reverse and complement) were present in the plastome of A. fruticosum (Table 6). The analysis showed 22 palindromic repeats, 21 forward repeats, 5 reverse repeats and 1 complement repeat (Table 6). In total, there were 49 long repeats in the chloroplast genome of A. fruticosum. The majority of the repeats were between 20 and 29 bp (87.75%) in size, followed by 30–39 bp (8.16%) and 50–59 bp (4.08%) long repeats. In the first location, the intergenic spacer harbored 61.22% of the repeats. The tRNA contained four repeats (8.16%), and eight repeats (16.32%) were located in the protein-coding genes. The length of repeated sequences in the A. fruticosum chloroplast genome ranged from 10 to 69 bp, analogously to the other angiosperm plants [41,42,43]. I compared the frequency of repeats among four Malvaceae cp genomes and found that all the types of repeats (palindromic, forward, reverse and complement) were present in all genomes (Figure 3). Malva parviflora has the highest number of palindromic repeats (25), while Sida szechuensis has the lowest with 17. A. fruticosum and M. parviflora have the same number of forward repeats—21 for each of them. T. populnea has the highest number of reverse repeats (9), while M. parviflora has the lowest (3). Complement repeats were found to be the least numerous types of repeat across the genome in A. fruticosum, in S. szechuensis and in M. parviflora, occurring once. In the plastome of T. populnea, there were three complement repeats.

2.2.2. Simple Sequence Repeats (SSRs)

There were short repeats of nucleotide series (1–6 bp) that were dispersed through the whole genome called microsatellites (SSRs). These short repeats in the plastid genome were passed from a single parent. As a result, they are used as molecular indicators in developmental studies such as genetic diversity and also contribute to the recognition of species [44,45,46]. A total of 212 microsatellites were found in the chloroplast genome of A. fruticosum in this study (Table 7). The majority of SSRs in the cp genome are mononucleotides (88.88), where poly-A (polyadenine) and poly-T (polythymine) are dominant (Figure 4). Poly-A constituted 45.06%, whereas poly-T constituted 41.97%. This is consistent with previous studies [47]. Among the dinucleotide repeats, only AG/CT and AT/AT were found in the cp genome. Taking into account the complementarity of series, only one trinucleotide (AAT/ATT), five tetranucleotides (AAAG/CTTT, AAAT/ATTT, AACT/AGTT, AATC/ATTG and AATG/ATTC) and only one pentanucleotide (AAAGT/ACTTT) were present in the cp genome (Figure 4). The intergenic/non-coding regions harbored most of the microsatellites (75.92%) (Figure 5).
The rate of occurrence of SSRs among the plastomes of the five species of Malvaceae was compared (Figure 6); the comparison indicates a high number of mononucleotides across all the plastomes. E. attenuatus and A. paniculata had the highest number of mononucleotides with 107 and 104, respectively. Pentanucleotides were not found in the plastome of B. prionitis, E. attenuatus, A. knappiae, B. ciliaris and R. breedlovei, while hexanucleotides were only present in B. prionitis, R. breedlovei and A. knappiae.

2.3. Comparative Analysis of Plastomes of Malvaceae Species

To examine the degree of divergence in the chloroplast genome of the six species of Malvaceae, comparative analysis was conducted using the mVISTA program to align the sequences using the annotation of A. fruticosum as a reference. The alignment showed that the genomes are highly conserved with some degree of variation. The coding regions are more conserved than the non-coding regions and the inverted repeat regions are more conserved than the single copy regions (Figure 7). This was reported in the chloroplast genomes of some genera in previous studies [47,48]. The most divergent non-coding regions among the six cp genomes are trnH-psbA, trnK-rps16, psbI-trnS, atpH-atpI, trnT-trnL, ndhC-trnV, accD-psaI, petA-psbJ, atpB-rbcL, rps12 and trnL-rpl32. A slightly lower level of variability was observed in the following genes: matK, ycf1, ndhH, ycf2 and accD. These regions can be used as a source of potential barcodes for identification/authentication of Malvaceae species as well as resources for inferring phylogenetic relationships of the family.
Generally, angiosperms retain the structure and size of the chloroplast genome [46]; however, due to evolutionary events such as an expansion and contraction in the genome, slight variations in the size and location of the boundaries of inverted repeats and single copy regions do occur [49,50]. I compared IR–LSC and IR–SSC boundaries of six cp genomes of Malvaceae (Abutilon fruticosum, Althaea officinalis, Abelmoschus esculentus, Malva parviflora, Sida szechuensis, Thespesia populnea) (Figure 8). The length of the six cp genomes ranged from 158,412 (M. parviflora) to 163,121 bp (A. esculentus). The genes rps19, rpl2 and trnH were located at the junctions LSC–IR and SSC–IR of the compared cp genomes with the exception of A. esculentus. The cp genome of A. fruticosum is different from the other cp genomes by having the ndhF gene in the reverse strand and in the junction of SSC and IRa. The ycf1 gene is located on the SSC–IRa border in the A. esculentus, M. Parviflora, S. szechuensis and T. populnea cp genomes and extends into IRa with 959 bp in A. esculentus. The cp genome of A. esculentus is unique by having the rpl16 gene on the LSC–IRb border and the rps3 gene on the Ira–LSC border. The cp genome of S. szechuensis has the smallest IR region, at 25,288 bp, while A. esculentus has the longest, at 28,009 bp. The ndhF gene is found on the IRb–SSC border of A. esculentus, S. szechuensis and T. populnea.

2.4. Divergence of Protein-Coding Gene Sequences

The rates of synonymous (dS) and nonsynonymous (dN) substitutions and the dN/dS ratio were calculated using DNAsp among the plastome of six species of Malvaceae to detect whether the 80 shared protein-coding genes were under selective pressure. The results show that the dN/dS ratio is less than 1 in almost all of the paired genes except petD of A. fruticosum vs. T. populnea, psaI of A. fruticosum vs. S. szechuensis and rps12 of A. fruticosum vs. T. populnea, A. fruticosum vs. S. szechuensis and A. fruticosum vs. T. populnea (Figure 9). This indicates that the majority of the genes were under negative selection, and only three of them underwent positive selection. The synonymous (dS) values range from 0.01 to 0.16 in all the genes (Figure 9). Some of the genes including infA, petG, petN, psaJ, psbA, psbZ, psbF, psbH, psbI, psbL and rps7 showed that no nonsynonymous changes occur in the plastome of the paired species of Malvaceae.

2.5. Phylogenetic Analysis

A complete chloroplast genome is a good resource for inferring evolutionary and phylogenetic relationships [51,52,53]. Many researchers have used plastome sequences to resolve phylogenetic relationships at various taxonomic levels [54,55]. To understand the evolutionary relationship of Malvoideae, Malvaceae and the phylogenetic position of A. fruticosum in the family, the complete plastome sequences of 10 species belonging to Malvoideae were downloaded from the GenBank database. In addition, two species, C. yunnanensis (Tilioideae, Malvaceae) and Bombax ceiba (Bombacoideae, Malvaceae), used as an outgroup, were also downloaded from GenBank. The downloaded cp genomes and the plastome of A. fruticosum were aligned using MAFFT. The phylogenetic tree was constructed using the Bayesian inference approach. The results reveal (Figure 10) that the species belonging to the subfamily Malvoideae are in one clade (monophyletic) with highly strong support, with a posterior probability (PP) value of (1.00). This is congruent with previous studies using molecular and morphological data [56,57,58]. The tree showed four distinct clades: a first clade containing Abutilon and Altheae and a second clade including Malvea species and being sister to a large clade containing two clades (Hibisceae and Gossypieae). A similar tree was obtained in a previous study using ITS [59] with slight variation. The species A. fruticosum is closely related and sister to A. officinalis. This result is incongruent with the earlier systematic position of A. fruticosum and S. szechuensis. Previous studies [60] reported that two species are sister taxa. In a recent classification, the subfamily Malvoideae [61] was divided into four tribes, namely, Malveae, Hibisceae, Gossypieae and Kydieae. Traditionally, Abutilon was placed in Malveae together with Malva and Sida by various researchers [62,63]. Later, Hutchinson [64] restructured the traditional classification using morphological data, particularly the ovule positions and their number. He proposed an introduction of such tribes as Abutileae (comprising two subtribes Abutilinae and Sidinae), Malveae, Malopeae and Hibisceae. Traditionally, Bentham, Hooker and Schumann classified Abutilon (tribe Malveae, subtribe Abutilinae), Malva (tribe Malveae, subtribe Eumalvinae), Sida (tribe Malveae, subtribe Sidinae) and Altheae (tribe Malveae, subtribe Eumalvnae); Hutchinson, later revised Abutilon (tribe Abutileae, subtribe Abutilinae), Malva and Altheae (tribe Malveae, subtribe Malvinae). Here, my results disagree with all the previous tribal positions of the genera. The tree showed that Abutilon is closely related to Altheae (with strong support) and Sida, which was reported as a sister to Abutilon, is in a different clade. Additionally, Malva and Altheae are also in different clades but were included in the same subtribe by previous classification. Based on the result in this study, I proposed the exclusion of Altheae from the tribe Malvae and its placement in Abututileae. Comparative analysis in this study (Figure 6 and Figure 7) also showed high similarity between cp genomes of Abutilon and Altheae. More sequenced chloroplast genomes of the representatives of the subfamily Malvoideae and phylogenetic analysis based on them would still be useful to establish the final systematic position of the genera within it.

3. Materials and Methods

3.1. Plant Material and DNA Extraction

Leaf material of Abutilon fruticosum was collected during field research in Jeddah, Saudi Arabia. Total genomic DNA was extracted from the samples using the Qiagen genomic DNA extraction kit according to the manufacturer’s protocol.

3.2. Library Construction, Sequencing and Assembly of the Chloroplast Genome

A total amount of 1.0 μg DNA was used as an input material for the DNA sample preparations. Sequencing libraries were generated using the NEBNext DNA Library Prep Kit for Illumina following the manufacturer’s recommendations. The genomic DNA was randomly fragmented into 350 bp long sequences. The raw reads were filtered to get the clean reads (5 Gb) using PRINSEQlite v0.20.4 [65] and were subjected to de novo assembly using NOVOPlasty2.7.2 [66] with kmer (K-mer= 31–33) to assemble the complete chloroplast genome from the whole genome sequence. One contig containing the complete chloroplast genome sequence was generated. The chloroplast genome sequence of A. fruticosum has been submitted to GenBank (accession number: MT772391)

3.3. Gene Annotation

Genes were annotated using DOGMA (Dual Organellar GenoMe Annotator, University of Texas at Austin, Austin, TX, USA) [67]. The positions of start and stop codons were adjusted manually. tRNA genes were identified by the trnAscan-SE server (http://lowelab.ucsc.edu/tRNAscan-SE/) [68]. The circular chloroplast genome maps were drawn using OGDRAW (Organellar Genome DRAW) [69].

3.4. Sequence Analysis

The relative synonymous codon usage values (RSCU), base composition and codon usage were computed using MEGA 6.0. Possible RNA editing sites present in the protein-coding genes of the cp genome of Malvaceae species were determined using PREP suite [35] with 0.8 as the cutoff value.

3.5. Repeat Analysis

Simple sequence repeats (SSRs) were identified in the Abutilon fruticosum chloroplast genome using the online software MIcroSAtellite (MISA) [70] with the following parameters: eight, five, four and three repeat units for mononucleotides, dinucleotides, trinucleotides and tetra-, penta- and hexanucleotide SSR motifs, respectively. For analysis of long repeats (palindromic, forward, reverse and complement), the program REPuter (https://bibiserv.cebitec.uni-bielefeld.de/reputer) [71] with default parameters was used to identify the size and location of the repeats in the genome.

3.6. Genome Comparison

The complete chloroplast genomes of six species of Malvaceae were compared with the program mVISTA [72] using the annotation of A. fruticosum as a reference in the Shuffle-LAGAN mode [73]. The border regions between the large single copy (LSC) and inverted repeat (IR) and small single copy (SSC) and inverted repeat (IR) junctions were compared using an IR scope.

3.7. Characterization of Substitution Rate

DNAsp v5.10.01 [74] was used to analyze synonymous (dS) and nonsynonymous (dN) substitution rates and the dN/dS ratio to detect the genes that are under selection pressure; the chloroplast genome of A. fruticosum was compared with the cp genome of M. parviflora, S. szchuensis, T. populnea and A. officinalis.

3.8. Phylogenetic Analysis

The complete chloroplast genomes of eleven Malvoideae and two species, Craigia yunnanensis (Tilioideae) and Bombax ceiba (Bombacoideae), were downloaded from GenBank. The downloaded sequences were aligned with the sequenced cp genome of A. fruticosum using MAFFT v.7 [75]. The data were analyzed with the Bayesian inference approach using MrBayes version 3.2.6 [76]. jModelTest version 3.7 [77] was used to select the suitable model.

Funding

This research was funded by the DEANSHIP OF SCIENTIFIC RESEARCH (DSR), KING ABDULAZIZ UNIVERSITY, JEDDAH, grant number DF-295-130-1441” and “The APC was funded by D. A. A.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are openly available in GenBank of NCBI at https://www.ncbi.nlm.nih.gov, reference number (A. fruticosum, MT772391).

Conflicts of Interest

The author declares no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Miller, P. The Gardeners Dictionary 3:23; John and James Rivington: London, UK, 1731. [Google Scholar] [CrossRef] [Green Version]
  2. Sweet, R. Hortus Britannicus, 1st ed.; J. Ridgway: London, UK, 1826. [Google Scholar]
  3. Sivarajan, V.V.; Pradeep, A.K. Malvaceae of Southern Peninsular India: A Taxonomic Monograph; Daya Publ. House: Delhi, India, 1996; pp. 201–204. [Google Scholar]
  4. Kearney, T.H. A tentative key to the South American species of Abutilon. Leafl. West. Bot. 1958, 8, 201–216. [Google Scholar]
  5. Fryxell, P.A. The American genera of Malvaceae-II. Brittonia 1997, 49, 204–269. [Google Scholar] [CrossRef]
  6. Fryxell, P.A. An Abutilon nomenclator (Malvaceae). Lundellia 2002, 5, 79–118. [Google Scholar] [CrossRef]
  7. Esteves, G.L.; Krapovickas, A. New species of Abutilon (Malvaceae) from Sao Paulo State, Brazil. Kew Bull. 2002, 57, 479–482. [Google Scholar] [CrossRef]
  8. Husain, S.A.; Baquar, S.R. Biosystematic studies in the genus Abutilon from Pakistan. Phyton 1974, 15, 219–234. [Google Scholar]
  9. Patel, M.K.; Rajput, A.P. Therapeutic significance of Abutilon indicum: An overview. Am. J. Pharm. Tech. Res. 2013, 4, 20–35. [Google Scholar]
  10. Khadabadi, S.S.; Bhajipale, N.S. A review on some important medicinal plants of Abutilon spp. Res. J. Pharm. Biol. Chem. Sci. 2010, 1, 718–729. [Google Scholar]
  11. Pingale, S.S.; Virkar, P.S. Evaluation of acute toxicity for Abutilon indicum. Pharm. Lett. 2011, 3, 37–42. [Google Scholar]
  12. Bano, I.; Deora, G.S. Studies on micro morphological taxonomic variations in Abutilon species of Indian Thar Desert. IOSR J. Pharm. Biol. Sci. 2017, 12, 60–68. [Google Scholar]
  13. Baquar, S.R. Medicinal and Poisonous Plants of Pakistan; Printas: Karachi, Pakistan, 1989. [Google Scholar]
  14. Ramar, K.; Ayyadurai, V. The present investigation deals with in-vitro Callus induction and plant regeneration of Abutilon indicum (L.). J. Pharmacogn. Phytochem. 2015, 3, 248–251. [Google Scholar]
  15. Christenhusz, M.J.M.; Byng, J.W. The number of known plants species in the world and its annual increase. Phytotaxa 2016, 261, 201–217. [Google Scholar] [CrossRef] [Green Version]
  16. Xu, Z.; Deng, M. Malvaceae, Identification and Control of Common Weeds, 1st ed.; Springer: Dordrecht, The Netherlands, 2017; pp. 717–735. [Google Scholar]
  17. Grevich, J.J.; Daniell, H. Chloroplast Genetic Engineering: Recent Advances and Future Perspectives. Crit. Rev. Plant Sci. 2005, 24, 83–107. [Google Scholar] [CrossRef]
  18. Neuhaus, H.; Emes, M. Nonphotosynthetic metabolism in plastids. Ann. Rev. Plant Biol. 2000, 51, 111–140. [Google Scholar] [CrossRef] [PubMed]
  19. Wicke, S.; Schneeweiss, G.M.; Depamphilis, C.W.; Kai, F.M.; 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] [Green Version]
  20. Oldenburg, D.J.; Bendich, A.J. The linear plastid chromosomes of maize: Terminal sequences, structures and implications for DNA replication. Curr. Genet. 2016, 62, 431–442. [Google Scholar] [CrossRef]
  21. Shaw, J.; Shafer, H.L.; Leonard, O.R.; Kovach, M.J.; Schorr, M.; Morris, A.B. Chloroplast DNA sequence utility for the lowest phylogenetic and phylogeographic inferences in angiosperms: Thetortoise and the hare IV. Am. J. Bot. 2014, 101, 1987–2004. [Google Scholar] [CrossRef] [Green Version]
  22. Guisinger, M.M.; Chumley, T.W.; Kuehl, J.V.; Boore, J.L.; Jansen, R.K. Implications of the plastid genome sequence of Typha (Typhaceae, poales) for understanding genome evolution in poaceae. J. Mol. Evol. 2010, 70, 149–166. [Google Scholar] [CrossRef] [Green Version]
  23. Yang, J.B.; Tang, M.; Li, H.T.; Zhang, Z.R.; Li, D.Z. Complete chloroplast genome of the genus Cymbidium: Lights into the species identification, phylogenetic implications and population genetic analysis. BMC Evol. Biol. 2013, 13. [Google Scholar] [CrossRef] [Green Version]
  24. Chen, H.; Shao, J.; Zhang, H.; Jiang, M.; Huang, L.; Zhang, Z.; Yang, D.; He, M.; Ronaghi, M.; Luo, X.; et al. Sequencing and analysis of Strobilanthes cusia (Nees) Kuntze chloroplast Genome revealed the rare simultaneous contraction and expansion of the inverted repeat region in Angiosperm. Front. Plant Sci. 2018, 9, 324. [Google Scholar] [CrossRef] [Green Version]
  25. Chang, C.C.; Lin, H.C.; Lin, I.P.; Chow, T.Y.; Chen, H.H.; Chen, W.H.; Cheng, C.H.; Lin, C.Y.; Liu, S.M.; Chang, C.C.; et al. 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, 279–291. [Google Scholar] [CrossRef] [Green Version]
  26. Raman, G.; Park, S. The complete chloroplast genome sequence of Ampelopsis: Gene organization, comparative analysis, and phylogenetic relationships to other angiosperms. Front. Plant Sci. 2016, 341, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Park, I.; Kim, W.J.; Yeo, S.-M.; Choi, G.; Kang, Y.-M.; Piao, R.; Moon, B.C. The complete chloroplast genome sequences of Fritillaria ussuriensis maxim. In addition, Fritillaria cirrhosa D. Don, and comparative analysis with other Fritillaria species. Molecules. 2017, 282, 22. [Google Scholar]
  28. Li, B.; Lin, F.; Huang, P.; Guo, W.; Zheng, Y. Complete chloroplast genome sequence of Decaisnea insignis: Genome organization, genomic resources and comparative analysis. Sci. Rep. 2017, 7, 10073. [Google Scholar] [CrossRef] [PubMed]
  29. Campbell, W.H.; Gowri, G. Codon usage in higher plants, green algae, and cyanobacteria. Plant Physiol. 1990, 92, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Liu, Q.; Dou, S.; Ji, Z.; Xue, Q. Synonymous codon usage and gene function are strongly related in Oryza sativa. Biosystems 2005, 80, 123–131. [Google Scholar] [CrossRef]
  31. Srivastava, D.; Shanker, A. Identification of simple sequence repeats in chloroplast genomes of Magnoliids through bioinformatics approach. Interdiscip. Sci. Comput. Life Sci. 2016, 8, 327–336. [Google Scholar] [CrossRef]
  32. Zhou, J.; Chen, X.; Cui, Y.; Sun, W.; Li, Y.; Wang, Y. Molecular structure and phylogenetic analyses of complete chloroplast genomes of two Aristolochia medicinal species. Int. J. Mol. Sci. 2017, 18, 1839. [Google Scholar] [CrossRef]
  33. Jiang, D.; Zhao, Z.; Zhang, T.; Zhong, W.; Liu, C.; Yuan, Q.; Huang, L. The chloroplast genome sequence of Scutellaria baicalensis provides insight into intraspecific and interspecific chloroplast genome diversity in Scutellaria. Gene 2017, 8, 227. [Google Scholar] [CrossRef]
  34. Zhou, J.; Cui, Y.; Chen, X.; Li, Y.; Xu, Z.; Duan, B.; Li, Y.; Song, J.; Yao, H. Complete chloroplast genomes of Papaver rhoeas and Papaver orientale: Molecular structures, comparative analysis, and phylogenetic analysis. Molecules 2018, 23, 437. [Google Scholar] [CrossRef] [Green Version]
  35. Mower, J.P. The PREP suite: Predictive RNA editors for plant mitochondrial genes, chloroplast genes and user-defined alignments. Nucleic Acids Res. 2009, 37, 253–259. [Google Scholar] [CrossRef]
  36. Bundschuh, R.; Altmuller, J.; Becker, C.; Nurnberg, P.; Gott, J.M. Complete characterization of the edited transcriptome of the mitochondrion of Physarum polycephalum using deep sequencing of RNA. Nucleic Acids Res. 2011, 39, 6044–6055. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Zeng, W.H.; Liao, S.C.; Chang, C.C. Identification of RNA editing sites in chloroplast transcripts of Phalaenopsis Aphrodite and comparative analysis with those of other seed plants. Plant Cell Physiol. 2007, 48, 362–368. [Google Scholar] [CrossRef] [Green Version]
  38. Wang, W.; Yu, H.; Wang, J.; Lei, W.; Gao, J.; Qiu, X.; Wang, J. The complete chloroplast genome sequences of the medicinal plant Forsythia suspense (oleaceae). Int. J. Mol. Sci. 2017, 18, 2288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Kumbhar, F.; Nie, X.; Xing, G.; Zhao, X.; Lin, Y.; Wang, S.; Weining, S. Identification and characterisation of RNA editing sites in chloroplast transcripts of einkorn wheat (Triticum monococcum). Ann. Appl. Biol. 2018, 172, 197–207. [Google Scholar] [CrossRef]
  40. Park, M.; Park, H.; Lee, H.; Lee, B.-H.; Lee, J. The complete plastome sequence of an antarctic Bryophyte Sanionia uncinata (Hedw.) loeske. Int. J. Mol. Sci. 2018, 19, 709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Li, Y.G.; Xu, W.Q.; Zou, W.T.; Jiang, D.Y.; Liu, X.H. Complete chloroplast genome sequences of two endangered Phoebe (Lauraceae) species. Bot. Stud. 2017, 58, 37. [Google Scholar] [CrossRef]
  42. Greiner, S.; Wang, X.; Rauwolf, U.; Silber, M.V.; Mayer, K.; Meurer, J.; Haberer, G.; Herrmann, R.G. 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, 2366–2378. [Google Scholar] [CrossRef] [Green Version]
  43. Song, Y.; Wang, S.; Ding, Y.; Xu, J.; Li, M.F.; Zhu, S.; Chen, N. Chloroplast Genomic Resource of Paris for Species Discrimination. Sci. Rep. 2017, 7, 3427. [Google Scholar] [CrossRef] [Green Version]
  44. Bryan, G.J.; McNicol, J.W.; Meyer, R.C.; Ramsay, G.; De Jong, W.S. Polymorphic simple sequence repeat markers in chloroplast genomes of Solanaceous plants. Theor. Appl. Genet. 1999, 99, 859–867. [Google Scholar] [CrossRef]
  45. Provan, J. Novel chloroplast microsatellites reveal cytoplasmic variation in Arabidopsis thaliana. Mol. Ecol. 2000, 9, 2183–2185. [Google Scholar] [CrossRef]
  46. Ebert, D.; Peakall, R. Chloroplast simple sequence repeats (cpSSRs): Technical resources and recommendations for expanding cpSSR discovery and applications to a wide array of plant species. Mol. Ecol. Resour. 2009, 9, 673–690. [Google Scholar] [CrossRef] [PubMed]
  47. Jacinta, N.M.; Xiang, D.; Jia-Xin, Y.; Elijah, M.M.; Vincent, O.W.; Millicent, A.O.; Josphat, K.S.; Paulm, M.M.; Guang-Wan, H. Complete chloroplast genome of Chlorophytum comosum and Chlorophytum gallabatense: Geome structures, comparative and phylogenetic analysis. Plants 2020, 9, 296. [Google Scholar]
  48. Dhafer, A.A.; Samaila, S.Y.; Enas, J.A.; Abidina, A. Complete chloroplast genome sequence of Barleria prionitis, Comparative chloroplast genomics and phylogenetic relationships among Acanthoideae. BMC Genom. 2020. [Google Scholar] [CrossRef]
  49. Philippe, H.; Delsuc, F.; Brinkmann, H.; Lartillot, N. Phylogenomics, Annual Review of Ecology. Evol. Syst. 2005, 36, 541–562. [Google Scholar] [CrossRef]
  50. Raubeson, L.A.; Peery, R.; Chumley, T.W.; Dziubek, C.; Fourcade, H.M.; Boorem, J.L.; Jansen, R.K. Comparative chloroplast genomics: Analyses including new sequences from the angiosperms Nupharadvena and Ranunculus macranthus. BMC Genom. 2007, 8, 174–201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Borsch, T.; Quandt, D. Mutational dynamics and phylogenetic utility of noncoding chloroplast DNA. Plant Syst. Evol. 2009, 282, 169–199. [Google Scholar] [CrossRef]
  52. Dong, W.P.; Liu, J.; Yu, J.; Wang, L.; Zhou, S.L. Highly variable chloroplast markers for evaluating plant phylogeny at low taxonomic levels and for DNA barcoding. PLoS ONE 2012, 7, e35071. [Google Scholar] [CrossRef]
  53. Tong, W.; Kim, T.S.; Park, Y.J. Rice chloroplast genome variation architecture and phylogenetic dissection in diverse Oryza species assessed by whole-genome resequencing. Rice 2016, 9, 57. [Google Scholar] [CrossRef] [Green Version]
  54. Dong, W.-P.; Liu, H.; Xu, C.; Zuo, Y.; Chen, Z.; Zhou, S. A chloroplast genomic strategy for designing taxon specific DNA mini-barcodes: A case study on ginsengs. BMC Genet. 2014, 15, 138. [Google Scholar] [CrossRef] [Green Version]
  55. Du, Y.-P.; Bi, Y.; Yang, F.-P.; Zhang, M.-F.; Chen, X.-Q.; Xue, J.; Zhang, X.-H. Complete chloroplast genome sequences of Lilium: Insights into evolutionary dynamics and phylogenetic analyses. Sci. Rep. 2017, 7, 5751. [Google Scholar] [CrossRef] [Green Version]
  56. Judd, W.S.; Manchester, S.R. Circumscription of Malvaceae (Malvales) as determined by a preliminary cladistic analysis of morphological, anatomical, palynological, and chemical characters. Brittonia 1997, 49, 384–405. [Google Scholar] [CrossRef]
  57. Alverson, W.S.; Whitlock, B.A.; Nyffeler, R.; Bayer, C.; Baum, D.A. Phylogeny of the core Malvales: Evidence from ndhF sequence data. Am. J. Bot. 1999, 86, 1474–1486. [Google Scholar] [CrossRef] [PubMed]
  58. Bayer, C.; Fay, M.F.; Bruijn, A.Y.D.; Savolainen, V.; Morton, C.M.; Kubitzki, K.; Alverson, W.S.; Chase, M.W. Support for an expanded family concept of Malvaceae within a recircumscribed order Malvales: A combined analysis of plastid atpB and rbcL DNA sequences. Bot. J. Linn. Soc. 1999, 129, 267–303. [Google Scholar] [CrossRef]
  59. Jennifer, A.T.; Javier, F.A.; Steven, J.W.; John, C.L.; Tracey, A.B.D.; Beryl, B.S. Phylogenetic relationships within the tribe Malveae (Malvaceae, subfamily Malvoideae) as inferred from ITS sequence data. Am. J. Bot. 2005, 92, 602. [Google Scholar]
  60. La Duke, J.C.; Doebley, J. A chloroplast DNA based phylogeny of the Malvaceae. Syst. Bot. 1995, 20, 259–271. [Google Scholar] [CrossRef]
  61. Bayer, C.; Kubitzki, K. Malvaceae. In Flowering Plants, Dicotyledons: Malvales, Capparales, and Nonbetalain Caryophyllales; Kubitzki, K., Bayer, C., Eds.; Springer: Berlin, Germany, 2003; pp. 225–311. [Google Scholar]
  62. Bentham, G.; Hooker, J.D. Malvaceae. In Genera Plantarum; Bentham, G., Hooker, J.D., Eds.; Reeve & Co.: London, UK, 1862; pp. 195–213. [Google Scholar]
  63. Schumann, K. Malvaceae. In Die naturlichen Pflanzenfamilien; Wilhelm Engelmann: Leipzig, Germany, 1890; pp. 30–53. [Google Scholar]
  64. Hutchinson, J. The Genera of Flowering Plants; Clarendon Press: Oxford, UK, 1967. [Google Scholar]
  65. Schmieder, R.; Edwards, R. Quality control and preprocessingof metagenomic datasets. Bioinformatics 2011, 27, 863–864. [Google Scholar] [CrossRef] [Green Version]
  66. Dierckxsens, N.; Mardulyn, P.; Smits, G. NOVOPlasty: De novo assembly of organelle genomes from whole genome data. Nucleic Acids Res. 2016, 45, e18. [Google Scholar]
  67. Wyman, S.K.; Jansen, R.K.; Boore, J.L. Automatic annotation of organellar genomes with DOGMA. Bioinformatics 2004, 20, 3252–3255. [Google Scholar] [CrossRef] [Green Version]
  68. Schattner, P.; Brooks, A.N.; Lowe, T.M. The tRNA scan-SE, snoscan and snoGPS web servers for the detection of tRNAs and snoRNAs. Nucleic Acids Res. 2005, 33, 686–689. [Google Scholar] [CrossRef]
  69. Lohse, M.; Drechsel, O.; Bock, R. OrganellarGenomeDRAW (OGDRAW): A tool for the easy generation of high-quality custom graphical maps of plastid and mitochondrial genomes. Curr. Genet. 2007, 52, 267–274. [Google Scholar] [CrossRef]
  70. Thiel, T.; Michalek, W.; Varshney, R.; 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, 411–422. [Google Scholar] [CrossRef] [PubMed]
  71. 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] [PubMed] [Green Version]
  72. Mayor, C.; Brudno, M.; Schwartz, J.R.; Poliakov, A.; Rubin, E.M.; Frazer, K.A.; Pachter, L.S.; Dubchak, I. VISTA: Visualizing global DNA sequence alignments of arbitrary length. Bioinformatics 2000, 16, 1046–1047. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Frazer, K.A.; Pachter, L.; Poliakov, A.; Rubin, E.M.; Dubchak, I. VISTA: Computational tools for comparative genomics. Nucleic Acids Res. 2004, 32, 273–279. [Google Scholar] [CrossRef] [PubMed]
  74. Librado, P.; Rozas, J. DnaSP v5: A software for comprehensive analysis of DNA polymorphism data. Bioinformatics 2009, 25, 1451–1452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. 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] [Green Version]
  76. Fredrik, R.; Maxim, T.; Paul, V.M.; Daniel, L.A.; Aaron, D.; Sebastian, H.; Bret, L.; Liang, L.; Mar, A.S.; John, P.H. MrBayes 3.2: Efficient Bayesian Phylogenetic Inference and Model Choice Across a Large Model Space. Syst. Biol. 2012, 61, 539–542. [Google Scholar]
  77. Posada, D. jModelTest: Phylogenetic model averaging. Mol. Biol. Evol. 2008, 25, 1253–1259. [Google Scholar] [CrossRef]
Figure 1. Gene map of the A. fruticosum chloroplast genome. Genes outside the circles are transcribed in the counter-clockwise direction and those inside in the clockwise direction. Known functional genes are indicated by colored bar. The GC and AT contents are denoted by the dark gray and light gray colors in the inner circle, respectively. LSC indicates large single copy; SSC indicates small single copy; and IR indicates inverted repeat.
Figure 1. Gene map of the A. fruticosum chloroplast genome. Genes outside the circles are transcribed in the counter-clockwise direction and those inside in the clockwise direction. Known functional genes are indicated by colored bar. The GC and AT contents are denoted by the dark gray and light gray colors in the inner circle, respectively. LSC indicates large single copy; SSC indicates small single copy; and IR indicates inverted repeat.
Plants 10 00270 g001
Figure 2. Amino acid frequencies in A. fruticosum chloroplast genome protein-coding sequences.
Figure 2. Amino acid frequencies in A. fruticosum chloroplast genome protein-coding sequences.
Plants 10 00270 g002
Figure 3. Number of different repeats in chloroplast genomes of five species of Malvaceae. P = palindromic, F = forward, R = reverse and C = complement.
Figure 3. Number of different repeats in chloroplast genomes of five species of Malvaceae. P = palindromic, F = forward, R = reverse and C = complement.
Plants 10 00270 g003
Figure 4. Frequency of different simple sequence repeat (SSR) motifs in different repeat types in the A. fruticosum chloroplast genome.
Figure 4. Frequency of different simple sequence repeat (SSR) motifs in different repeat types in the A. fruticosum chloroplast genome.
Plants 10 00270 g004
Figure 5. Number of SSR types in the complete cp genome, and protein-coding and non-coding sequences in A. fruticosum.
Figure 5. Number of SSR types in the complete cp genome, and protein-coding and non-coding sequences in A. fruticosum.
Plants 10 00270 g005
Figure 6. Number of different SSR types in the chloroplast genome of five Malvaceae.
Figure 6. Number of different SSR types in the chloroplast genome of five Malvaceae.
Plants 10 00270 g006
Figure 7. Sequence alignment of six chloroplast genomes of representatives of the Malvaceae family performed with mVISTA using annotation of A. fruticosum as a reference. The top arrow shows transcription direction, blue color indicates protein-coding sequences, pink color shows conserved non-coding sequences (CNS) and light green indicates tRNAs and rRNAs. The x-axis represents the coordinates in the cp genome, while the y-axis represents percentage identity within 50–100%.
Figure 7. Sequence alignment of six chloroplast genomes of representatives of the Malvaceae family performed with mVISTA using annotation of A. fruticosum as a reference. The top arrow shows transcription direction, blue color indicates protein-coding sequences, pink color shows conserved non-coding sequences (CNS) and light green indicates tRNAs and rRNAs. The x-axis represents the coordinates in the cp genome, while the y-axis represents percentage identity within 50–100%.
Plants 10 00270 g007
Figure 8. Comparison of the borders of the IR, SSC and LSC regions among six chloroplast genomes of Malvaceae (JLB: juction of LSC and IRB; JSB: junction of SSC and IRB; JSA: juction SSC and IRA; JLA: junction LSC and IRA).
Figure 8. Comparison of the borders of the IR, SSC and LSC regions among six chloroplast genomes of Malvaceae (JLB: juction of LSC and IRB; JSB: junction of SSC and IRB; JSA: juction SSC and IRA; JLA: junction LSC and IRA).
Plants 10 00270 g008
Figure 9. The synonymous (dS) and nonsynonymous (dN)/dS ratio values of 80 protein-coding genes from five Malvaceae cp genomes (A.f: A. fruticosum; A.o: A. officinalis; M.P: M. parviflora; T.p: T. populnea; S.s: S. szechuensis).
Figure 9. The synonymous (dS) and nonsynonymous (dN)/dS ratio values of 80 protein-coding genes from five Malvaceae cp genomes (A.f: A. fruticosum; A.o: A. officinalis; M.P: M. parviflora; T.p: T. populnea; S.s: S. szechuensis).
Plants 10 00270 g009aPlants 10 00270 g009b
Figure 10. Phylogenetic tree reconstruction of 14 taxa based on the complete chloroplast genomes using the Bayesian inference (BI) method showing relationships within the 14 species of Malvaceae. The numbers in the branch nodes represent posterior probability (PP).
Figure 10. Phylogenetic tree reconstruction of 14 taxa based on the complete chloroplast genomes using the Bayesian inference (BI) method showing relationships within the 14 species of Malvaceae. The numbers in the branch nodes represent posterior probability (PP).
Plants 10 00270 g010
Table 1. Base composition in the A. fruticosum chloroplast genome.
Table 1. Base composition in the A. fruticosum chloroplast genome.
Region T(U) (%)C (%)A (%)G (%)Total (bp)
cp genome 31.1318.2631.918.69160,357
LSC 33.3817.9431.8416.8289,032
SSC 33.8815.0334.5716.9420,031
IRA 28.5122.2628.5920.6325,646
IRB 28.5820.6128.5322.2625,646
1st Position31.2418.3431.8118.5953,453
2nd Position31.0718.2532.0218.6453,452
3rd Position31.0918.1831.8618.8553,452
Table 2. Genes present in the chloroplast genome of A. fruticosum.
Table 2. Genes present in the chloroplast genome of A. fruticosum.
CategoryGroup of GenesName of Genes
RNA genesRibosomal RNA genes (rRNA)rrn5, rrn4.5, rrn16, rrn23
Transfer RNA genes (tRNA)trnH-GUG, trnK-UUU +, trnQ-UUG, trnS-GCU, trnG-UCC, trnR-UCU, trnC-GCA, trnD-GUC, trnY-GUA, trnE-UUC, trnM-CAU, trnT-GGU, trnS-UGA, trnG-GCC, trnfM-CAU, trnS-GGA, trnT-UGU, trnL-UAA +, trnF-GAA, trnV-UAC +, trnW-CCA, trnP-UGG, trnP-GGG, trnL-CAA a, trnV-GAC a, trnI-GAU +,a, trnA-UGC +,a, trnR-ACG a, trnN-GUU a, trnL-UAG.
Ribosomal proteinsSmall subunit of ribosomerps2, rps3, rps4, rps7 a, rps8, rps11, rps12 a, rps14, rps15, rps16 +, rps18, rps19
TranscriptiongenesLarge subunit of ribosomerpl2 +,a, rpl14, rpl16, rpl20, rpl22, rpl23a, rpl32, rpl33, rpl36.
DNA-dependent RNA polymeraserpoA, rpoB, rpoC1 +, rpoC2
Protein genesPhotosystem IpsaA, psaB, psaC, psaI, psaJ, ycf3 ++
Photosystem IIpsbA, psbB, psbC, psbD, psbE, psbF, psbH, psbI, psbJ, psbK, psbL, psbM, psbN, psbT, psbZ
Subunit of cytochromepetA, petB, petD, petG, petL, petN
Subunit of synthaseatpA, atpB, atpE, atpF +, atpH, atpI
Large subunit of rubiscorbcL
NADH dehydrogenasendhA +, ndhB +,a, ndhC, ndhD, ndhE, ndhF, ndhG, ndhH, ndhI, ndhJ, ndhK
ATP-dependent protease subunit PclpP ++
Chloroplast envelope membrane proteincemA
Other genesMaturasematK
Subunit acetyl-coA carboxylaseaccD
C-type cytochrome synthesisccsA
Translational initiation factorinfA
Hypothetical proteinsycf2 a, ycf4, ycf15 a
Component of TIC complexycf1 a
+ Gene with one intron, ++ gene with two introns and a gene with copies.
Table 3. The introns in the genes of the A. fruticosum plastome.
Table 3. The introns in the genes of the A. fruticosum plastome.
GeneLocationExon I (bp)Intron I (bp)Exon II (bp)Intron II (bp)Exon III (bp)
rps16LSC22488935
atpFLSC410790158
rpoC1LSC1634759434
ycf3LSC152773227763125
ndhKLSC7167155
accDLSC4461487
clpPLSC22764029391470
rpI2IR434682392
ycf15IR13124125
ndhBIR755684776
ndhASSC5511124539
trnK-UUULSC34257436
trnG-UCCLSC3178960
trnL-UAALSC3656049
trnV-UACLSC3659037
trnI-GAUIR4195834
trnA-UGCIR3780034
Table 4. Codon–anticodon recognition patterns and codon usage of the A. fruticosum chloroplast genome.
Table 4. Codon–anticodon recognition patterns and codon usage of the A. fruticosum chloroplast genome.
CodonAmino AcidRSCUtRNACodonAmino AcidRSCUtRNA
UUUPhe1.27trnF-GAAUAUTyr1.57trnY-GUA
UUCPhe0.73 UACTyr0.43
UUALeu1.85trnL-UAAUAAStop1.35
UUGLeu1.24trnL-CAAUAGStop1.06
CUULeu1.22trnL-UAGCAUHis1.49trnH-GUG
CUCLeu0.46 CACHis0.51
CUALeu0.83 CAAGln1.5trnQ-UUG
CUGLeu0.41 CAGGln0.5
AUUIle1.46trnI-GAUAAUAsn1.52trnN-GUU
AUCIle0.62 AACAsn0.48
AUAIle0.93trnI-CAUAAALys1.48trnK-UUU
AUGMet1trnM-CAUAAGLys0.52
GUUVal1.45trnV-GACGAUAsp1.61trnD-GUC
GUCVal0.53 GACAsp0.39
GUAVal1.46 GAAGlu1.45trnE-UUC
GUGVal0.56trnV-UACGAGGlu0.55
UCUSer1.65trnS-GGAUGUCys1.5trnC-GCA
UCCSer0.98 UGCCys0.5
UCASer1.22 UGAStop0.59
UCGSer0.57trnS-UGAUGGTrp1trnW-CCA
CCUPro1.47trnP-UGGCGUArg1.2trnR-ACG
CCCPro0.77 CGCArg0.49trnR-UCU
CCAPro1.15 CGAArg1.35
CCGPro0.62 CGGArg0.52
ACUThr1.57 AGAArg1.71
ACCThr0.78 AGGArg0.73
ACAThr1.16trnT-GGUAGUSer1.14trnS-GCU
ACGThr0.49trnT-UGUAGCSer0.43
GCUAla1.76trnA-UGCGGUGly1.28
GCCAla0.7 GGCGly0.46
GCAAla1.03 GGAGly1.51
GCGAla0.52 GGGGly0.75trnG-UCC
Table 5. Predicted RNA editing sites in the A. fruticosum chloroplast genome.
Table 5. Predicted RNA editing sites in the A. fruticosum chloroplast genome.
GeneNucleotide PositionAmino Acid PositionCodonAmino AcidScore
accD854285TCG => TTGS => L0.8
1463488CCT => CTTP => L1
atpA914305TCA => TTAS => L1
1148383TCA => TTAAS => L1
atpF9231CCA => CTAP => L0.86
atpI629210TCA => TTAS => L1
ccsA662221ACT => ATTT => I0.86
clpP559187CAT => TATH => Y1
matK457153CAT => TATH => Y1
634212CAT => TATH => Y1
1237413CAC => TACH => Y1
ndhA341114TCG => TTGS => L1
566189TCA => TTAS => L1
ndhB14950TCA => TTAS => L1
467156CCA => CTAP => L1
542181ACG => ATGT => M1
586196CAT => TATH => Y1
611204TCA => TTAS => L0.8
737246CCA => CTAP => L1
746249TCT => TTTS => F1
830277TCAG => TTGS => L1
836279TCA => TTAS => L1
1255419CAT => TATH => Y1
1291431CTC => CTAL => F1
1481494CCA => CTAP => L1
ndhD21ACG => ATGT => M1
269ACA => ATAT => I1
4716TCT => TTTS => F0.8
383128TCA => TTAS => L1
568190CCT => TCTP => S1
674225TCG => TTGS => L1
878293TCA => TTAS => L1
1298433TCA => TTAS => L0.8
ndhF29097TCA => TTAS => L1
1549517CTT => TTTL => F1
1826609ACA => ATAT => I0.8
1892631GCG => GTGA => V0.8
ndhG16656CAT => TATH => Y0.8
314105ACA => ATAT => I0.8
petB425142CGG => TGGR => W1
psbF7726TCT => TTTS => F1
rpl20308103TCA => TTAS => L0.86
rpoA329110GCC => GTCA => V0.86
830277TCA => TTAS => L1
rpoB338113TCT => TTTS => F1
551184TCA => TTAS => L1
566189TCG => TTGS => L1
2426809TCA => TTAS => L0.86
rpoC14114TCA => TTAS => L1
1273425CCG => TCGP => S0.86
rpoC22296766CGG => TGGR => W1
31881063CCC => CTCP => L0.86
rps224883TCG => TTGS => L1
325109CCC => TCCP => S1
rps821773CAT => TATH => Y1
rps1414950TCA => TTAS => L1
Table 6. Repeat sequences present in the A. fruticosum chloroplast genome.
Table 6. Repeat sequences present in the A. fruticosum chloroplast genome.
SNRepeat SizeRepeat Position 1Repeat TypeRepeat LocationRepeat Position 2Repeat Location 2E-Value
15833,156FIGS33,213IGS8.71 × 10−26
2500PIGS89,084IGS5.71 × 10−21
33644,187FIGS44,205IGS1.53 × 10−12
436103,838PIGS123,909ndhA-Intron1.53 × 10−12
536123,909FndhA-Intron145,617IGS1.53 × 10−12
63064,915FIGS64,930IGS6.27 × 10−9
7298535PtrnS-GCU47,062IGS2.51 × 10−8
8266039PIGS6039IGS1.61 × 10−6
92610,662PIGS10,662IGS1.61 × 10−6
102696,437Fycf296,455ycf21.61 × 10−6
112696,437Pycf2153,010ycf21.61 × 10−6
122696,455Pycf2153,028ycf21.61 × 10−6
1326153,010Fycf2153,028ycf21.61 × 10−6
142553,987PIGS58,381IGS6.42 × 10−6
152554,230RIGS54,230IGS6.42 × 10−6
162419,010FrpoC219,034rpoC22.57 × 10−6
172440,578FpsaB42,802psaA2.57 × 10−5
182448,652FIGS48,672IGS2.57 × 10−5
192348,902RIGS48,902IGS1.03 × 10−4
202388,769Frps1988,792rps191.03 × 10−4
2123112,986FIGS113,018IGS1.03 × 10−4
2223112,986PIGS136,450IGS1.03 × 10−4
2323113,018PIGS136,482IGS1.03 × 10−4
2423136,450FIGS136,482IGS1.03 × 10−4
252210,362PIGS10,388IGS4.11 × 10−4
262238,111PIGS38,111IGS4.11 × 10−4
2722113,764FIGS113,785IGS4.11 × 10−4
2822113,764PIGS135,684IGS4.11 × 10−4
2922113,785PIGS135,705IGS4.11 × 10−4
3022135,684FIGS135,705IGS4.11 × 10−4
31218540FtrnS-GCU37,036trnS-UGA1.64 × 10−3
32219201FIGS9220IGS1.64 × 10−3
332110,136FIGS10,157IGS1.64 × 10−3
342110,491RtrnR-UCU10,491trnR-UCU1.64 × 10−3
352137,036PtrnS-UGA47,065IGS1.64 × 10−3
362143,822RIGS43,822IGS1.64 × 10−3
372178,782PpsbN78,810psbN1.64 × 10−3
382196,428Fycf296,482ycf21.64 × 10−3
392196,428Pycf2152,988ycf21.64 × 10−3
402196,482Pycf2153,042ycf21.64 × 10−3
4121152,988Fycf2153,042ycf21.64 × 10−3
4220402PIGS402IGS6.58 × 10−3
43205288FIGS5307IGS6.58 × 10−3
442010,128CIGS82,502IGS6.58 × 10−3
452014,151PIGS54,042IGS6.58 × 10−3
462050,997PtrnF-GAA55,404IGS6.58 × 10−3
472053,309RndhC53,309ndhC6.58 × 10−3
482055,405PIGS108,890ycf26.58 × 10−3
492055,405FIGS140,581ycf26.58 × 10−3
Table 7. Simple sequence repeats in the chloroplast genome of A. fruticosum.
Table 7. Simple sequence repeats in the chloroplast genome of A. fruticosum.
RepeatLength (bp)NumberStart Position
A8404686; 5481; 5865; 6926; 7897; 12,271; 13,473; 13,702; 15,318; 19,663; 22,147; 23,354; 30,009; 30,421; 31,419; 49,433; 50,206; 51,576; 54,400; 69,630; 74,881; 76,238; 82,296; 84,962; 85,726; 115,568; 117,338; 118,072; 120,375; 120,736; 128,085; 130,563;131,446; 131,580; 135,564; 135,895; 140,843; 141,163; 145,401; 148,617
9124151; 8835; 16,318; 23,595; 67,970; 80,923; 87,139; 94,543; 117,759; 119,376; 125,042; 125,285
10115166; 7193; 28,188; 37,725; 48,400; 74,710; 75,051; 76,083; 115,690; 128,759; 131,209
11552,345; 82,469; 116,213; 117,460; 118,643
123320; 2221; 65,189
14114,115
151145,189
C8127,432
10114,953
11130,215
T828116; 4444; 9679; 28,812; 32,051; 36,848; 37,359; 58,437; 61,319; 63,814; 67,383; 67,842; 68,540; 71,227; 79,674; 81,661; 88,975; 100,748; 103,964; 108,202; 108,522; 113,470; 113,801; 115,834; 126,161; 127,095; 130,350; 131,860
9132643; 6292; 13,241; 20,064; 31,032; 43,959; 51,110; 54,805; 71,700; 82,737; 87,273; 123,499; 154,821
10155982; 9102; 10,369; 10,642; 12,750; 14,567; 17,590; 30,779; 34,153; 49,053; 58,034; 64,821; 66,148; 81,282; 84,685
11619,520; 21,952; 29,839; 70,361; 75,959; 85,189
12427,254; 54,171; 57,834; 63,361
151104,169
1618658
AT5320,902; 53,741; 63,067
6128,039
TA5210,115; 33,119
TC5165,402
CT5117,231
AAT4113,837
TTA41160,252
AATA3113,014
AGAA31115,404
ATCA31126,521
GAAT31118,455
TAGT3162,483
TTTA3172,098
ACTTT31139,519
TAAAG31109,838
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Alzahrani, D.A. Complete Chloroplast Genome of Abutilon fruticosum: Genome Structure, Comparative and Phylogenetic Analysis. Plants 2021, 10, 270. https://doi.org/10.3390/plants10020270

AMA Style

Alzahrani DA. Complete Chloroplast Genome of Abutilon fruticosum: Genome Structure, Comparative and Phylogenetic Analysis. Plants. 2021; 10(2):270. https://doi.org/10.3390/plants10020270

Chicago/Turabian Style

Alzahrani, Dhafer A. 2021. "Complete Chloroplast Genome of Abutilon fruticosum: Genome Structure, Comparative and Phylogenetic Analysis" Plants 10, no. 2: 270. https://doi.org/10.3390/plants10020270

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