Next Article in Journal
Phloroglucinol Mediated Plant Regeneration of Ornithogalum dubium as the Sole “Hormone-Like Supplement” in Plant Tissue Culture Long-Term Experiments
Previous Article in Journal
Enhanced UV-Reflection Facilitated a Shift in the Pollination System of the Red Poppy, Papaver rhoeas (Papaveraceae)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Complete Chloroplast Genome Characterization of Oxalis Corniculata and Its Comparison with Related Species from Family Oxalidaceae

1
Department of Botany, Garden Campus, Abdul Wali Khan University, Mardan 23200, Pakistan
2
Natural and Medical Sciences Research Center, University of Nizwa, Nizwa 616, Oman
3
School of Applied Biosciences, Kyungpook National University, Daegu 41566, Korea
*
Authors to whom correspondence should be addressed.
Plants 2020, 9(8), 928; https://doi.org/10.3390/plants9080928
Submission received: 27 May 2020 / Revised: 14 July 2020 / Accepted: 17 July 2020 / Published: 23 July 2020
(This article belongs to the Section Plant Genetics, Genomics and Biotechnology)

Abstract

:
Oxalis corniculata L. (family Oxalidaceae) is a small creeper wood sorrel plant that grows well in moist climates. Despite being medicinally important, little is known about the genomics of this species. Here, we determined the complete chloroplast genome sequence of O. corniculata for the first time and compared it with other members of family Oxalidaceae. The genome was 152,189 bp in size and comprised of a pair of 25,387 bp inverted repeats (IR) that separated a large 83,427 bp single copy region (LSC) and a small 16,990 bp single copy region (SSC). The chloroplast genome of O. corniculata contains 131 genes with 83 protein coding genes, 40 tRNA genes, and 8 rRNA genes. The analysis revealed 46 microsatellites, of which 6 were present in coding sequences (CDS) regions, 34 in the LSC, 8 in the SSC, and 2 in the single IR region. Twelve palindromic repeats, 30 forward repeats, and 32 tandem repeats were also detected. Chloroplast genome comparisons revealed an overall high degree of sequence similarity between O. corniculata and O. drummondii and some divergence in the intergenic spacers of related species in Oxalidaceae. Furthermore, the seven most divergent genes (ccsA, clpP, rps8, rps15, rpl22, matK, and ycf1) among genomes were observed. Phylogenomic characterization on the basis of 60 shared genes revealed that O. corniculata is closely related to O. drummondii. The complete O. corniculata genome sequenced in the present study is a valuable resource for investigating the population and evolutionary genetics of family Oxalidaceae and can be used to identify related species.

1. Introduction

The largest genus of family Oxalidaceae is Oxalis L., which is distributed mostly in Southern Africa and South America, and comprises of more than 500 species. About one-half of the total species (>200 spp) growing in Southern Africa share a bulbous or tuberous in herbaceous taxa [1,2,3]. A huge morphological variation have been observed among approximately 250 species in South America, where this genus seems to have originated and diversified [2,4]. However, Oxalis sections Corniculatae DC. consist of creeping herbs, and many of the species grows in the temperate and humid areas of the Americas [5]. A cytogenetic study showed that Corniculatae tends to be categorized into two sub-groups: (1) a large number of diploid and polyploid species with a base chromosome number of x = 6, symmetrical karyotypes, and medium and small chromosomes size, and (2) a minor number of diploid species having x = 5, more asymmetrical karyotypes, and average to large chromosomes [3,6,7]. The taxonomy has been suffering with similarities in phenotypes across different species.
Among species from genus Oxalis, Oxalis corniculata L. is a small creeping herb that is adaptive to moist conditions with yellow flowers and cylindrical fruits [4]. The origin of O. corniculata L. has been unknown until now. O. corniculata was first described by Carl Linnaeus from the Mediterranean region and various others suggested that this region is the native range [4,8]. Currently, among vascular plant species, O. corniculata has the third largest distribution [9]. The early spread and global distribution of O. corniculata has contributed to the difficulty in identifying its native range [4]. It is highly persistent in the horticulture industry and commonly found in gardens and as a hitchhiker in plant pots in gardens and nurseries [10]. Additionally, O. corniculata had medicinal values to treat various infectious diseases [11]. It has been known as antifungal, antibacterial, anti-inflammatory, anthelmintic, antidiuretic, astringent, emmenagogue, depurative, febrifuge, lithontriptic, stomachic, and styptic [12]. It has sticky seeds, explosive capsules, persistent flowers, and a brief life cycle, which make it an effective colonizer and constant weed. O. corniculata is not a strong competitor; however, its abundance and enormous range increase its overall impact as a weed.
Chloroplasts (cp) are important plant organelles that carry out photosynthesis and the biosynthesis of amino acid, fatty acids, and pigments [13,14]. Similarly, cp genomes have proven to be a valuable resource for species identification, plant phylogenetics, population genetics, and genetic engineering. Its DNA is inherited maternally in not all but in the majority of angiosperm species [15]. As a result of its inheritance approach, cp DNA plays a significant role in population genetics and molecular evolution. Therefore, its DNA can not only be used for species discrimination but can also be used solved many questions related to taxonomy [15,16]. Chloroplasts contain their own independent genomes and genetic systems, and DNA replication and transmission to daughter organelles result in a cytoplasmic inheritance of characteristics associated with primary events in photosynthesis [17,18]. The cp genome structure of angiosperms is circular—about 120–217 kb in size—and contain a quadripartite conformation [19,20]. It has small single copy (SSC) and large single copy (LSC) regions that are generally segregated by double copies of an inverted repeat region (IRa and IRb) [20]. In terms of gene order and content, the cp genome is usually conserved in various families of the angiosperm, such as Campanulaceae, Fabaceae, Geraniaceae, and Oleaceae [21,22,23]. Due to its conserved structure, small size, and recombination-free nature, the chloroplast genome is broadly used in plant phylogenetic analysis [24,25]. Similarly, the cp genome has a highly conserved structure that simplifies sequencing and primer designing, and chloroplast DNA can be used for the identification of plants as a barcode [26,27].
Molecular phylogenetic tools were broadly used to assess previous taxonomy and evolutionary processes. The number of chromosomes and contents of DNA (cytogenetic data) are usually associated with phylogenetic trees in order to understand the karyological differences involved in the group diversification [28,29]. Molecular phylogenomic studies using cp genome sequence data from the genes and slowly evolving inverted repeat regions have been applied to reveal the deep-level evolutionary relationships of plant taxa [30], producing robust phylogenies that are corroborated by sequence data from mitochondrial and nuclear genomes [31]. With the advancement of different genomic methods and tools, next-generation technologies have provided the rapid sequencing of different cp genomes from both angiosperms and gymnosperms in recent years, which enabled the confirmation of evolutionary relationships and detailed phylogenetic classifications to be conducted at the group, family, genus, and even species levels in plants [32,33]. Therefore, cp genome-scale data have increasingly been used to infer phylogenetic relationships at high taxonomic levels, and even at lower levels, great progress has been made [27,34,35,36].
The phylogenetic analysis of various Oxalis taxa has been reported previously by using nuclear Internal Transcribed Spacer (ITS) regions [37]; however, more detailed insight is still missing, keeping in view the complex taxonomy of O. corniculata. It is variable cytologically and genetically, but it is also phenotypically plastic [38,39]. Its taxonomy is complicated by the description of many subspecific taxa and other species now considered to be synonyms [5]. Furthermore, there has been taxonomic confusion of O. corniculata with closely related species, such as O. stricta and O. dilleniid [4]. Similarly, there is little information available on their genetic structure, especially their chloroplast genomes or their detailed phylogenetic placement. Hence, the current study was performed with the aim to sequence and analyze the complete chloroplast genome of O. corniculata and compare it with related species from the family Oxilidaceae (Oxalis drummondii, Averrhoa carambola and Cephalotus follicularis). We also aimed to elucidate and compare the global pattern of structural variation in the cp genome of O. corniculata and related three species. In addition, we compared the IR region contraction and expansion, intron contents, regions of high sequence divergence, and phylogenomic of O. corniculata with related species cp genomes to reveal more insight regarding the comparative genome architecture.

2. Results

2.1. Chloroplast Genome Structure of O. corniculata

The chloroplast genome of O. corniculata is 152,189 bp and displays a distinctive quadripartite structure with a pair of 25,387 bp inverted repeats (IRs) that separate 83,427 bp single copy regions and 16,990 bp single copy regions (Figure 1). The total GC content is 36.7% with uneven distribution across the whole genome. The GC contents of IRs are higher (42.6%) than the large single copy and small single copy regions (34.4% and 30.3%, respectively). Furthermore, the O. corniculata cp genome consists of 131 genes and among all these genes, 82 are protein-coding genes, 40 are tRNA, and 8 are rRNA (Figure 1; Table 1). The protein-coding genes present in the O. corniculata cp genome included nine genes for large ribosomal proteins (rpl2, 14, 16, 20, 22, 23, 32, 33, 36), 11 genes for small ribosomal proteins (rps2, 3, 4, 7, 8, 11, 12, 14, 15, 18, 19), five genes for photosystem I (psaA, B, C, I, J), 15 genes for photosystem II (psbA, B, C, D, E, F, H, I, J, K, L, M, T, Z) and six genes (atpA, B, E, F, H, I) for ATP synthesis and the electron transport chain (Figure 1, Table 2). Similarly, 17 genes contain introns (11 protein-coding genes and 6 tRNA genes), of which three comprised of two introns (rps12, clpP, ycf3), while the rest have a single intron (Table 3). The small ribosomal protein 12 gene (rps12) is trans-spliced with a single intron. Its 5’ exon is located in the LSC region, while is 3’ exon is located in the IRb region and duplicated in the IRa region (Figure 1). The largest intron was present in trnK-UUU (2558 bp), whilst trnL-UAA contains the smallest intron (492 bp) (Table 3). The protein-coding region accounts for 52%, while the tRNA and rRNA regions constitute 1.99% and 5.94%, respectively, in the cp genome. The length of the protein-coding region is about 79,239 bp while those of tRNA and rRNA are 3042 bp and 9048 bp, respectively. Similarly, the rps16 gene, which is found in most angiosperm plastid genomes, is absent in O. corniculata. Furthermore, infA is also absent in the O. corniculata cp genome.

2.2. Comparative Analysis of O. corniculata Chloroplast Genome with Related Species

The O. corniculata chloroplast genome was compared with three already sequenced genomes from family Oxalidaceae i.e., O. drummondii, A. carambola, and C. follicularis (Table 1). Variations were observed in cp genomes where C. follicularis has the smallest cp genome, 142,706 bp, whilst the A. carambola cp genome was the largest, 155,965 bp, amongst analyzed species. We also compared the O. corniculata cp genome for pairwise sequence divergence (Table S1). Results showed that O. corniculata exhibited the lowest pairwise sequence divergence as compared to O. drummondii (0.044) and A. carambola (0.057; Table S1). Similarly, the whole cp genomes of O. corniculata were compared to identify sequence divergence via mVISTA (Figure 2). Results showed that the coding regions of all cp genomes are conserved compare to non-coding regions, whilst non-coding regions showed a higher divergence rate than the coding regions. The most divergence was observed in intergenic spaces. The matK gene exhibited a high degree of divergence in all genomes, but it was higher in A. carambola and C. follicularis as compared to others. Similarly, rpoC1, rpoC2, and rpoB exhibited more divergence in four cp genomes with the highest in C. follicularis. The region between rpoB and psbD genes showed a high degree of divergence in all species in the LSC. The ndhK and rpl16 genes showed significant divergence compared to the ycf2 gene with lesser values. The ndhB is less divergent in all species except C. follicularis, where it showed high divergence. Similarly, the ndhF gene is highly divergent in all cp genomes, while it is absent in the C. follicularis cp genome. In the SSC, the region between ndhG and ndhH is highly divergent and the cp genome of C. follicularis lack most of the NadH oxidoreductase genes. Furthermore, about 63 protein-coding gene sequences were compared to obtain the average pairwise sequence distance among these species. The results showed that a majority of the genes maintained low levels of average sequence divergence. A relatively lower sequence identity was observed between the chloroplast genomes of O. corniculata with related species, especially in the ccsA, clpP, rps8, rps15, rpl22, matK, and ycf1 genes (Figure 3).

2.3. Expansion and Contraction of IR Regions

A detailed comparison was performed of the four junctions (JLA, JLB, JSA, and JSB) amongst IRs (IRa and IRb) and the LSC/SSC regions for the O. corniculata, O. drummondii, A. carambola, and C. follicularis (Figure 4). We carefully analyzed and compared the exact IR border positions and the adjacent genes among these cp genomes. The results revealed that at the LSC/IRb (JLB) border rps19 gene is present 22 bp away from the JLB junction and located in the LSC region in O. corniculata. However, in O. drummondii, the rps19 gene is located 16 bp away, while in A. carambola, it is 1 bp away from the JLB junction. On the other hand, in C. follicularis, rps19 is located in the JLB junction and extended 67 bp in the IRb region. The ycf1 gene is located 45 bp away from IRb/SSC (JSB) in O. corniculata, while in O. drummondii and A. carambola, the ycf1 gene is located 1 bp away in the IRb region. However, in C. follicularis, the ycf1 gene is absent in the IRb region and located 976 bp apart from the JSA junction and located in the IRa region. Similarly, the ndhF gene is located in the (JSA) border and extended 23 bp and 20 bp in the IRa region in both O. corniculata and O. drummondii genomes, respectively, revealing the expansion of the IR region. In A. carambola, the ndhF gene is located 55 bp away from the JSA junction in the SSC region. However, the ndhF gene is completely absent in the C. follicularis cp genome. Similarly, the rpl2 gene is located 53 bp away from the IRa/LSC (JLA) junction in both O. corniculata and O. drummondii, while in A. carambola and C. follicularis, it is 78 bp and 124 bp away from the JLA junction, respectively (Figure 4).

2.4. Repeat Sequence Analysis

We investigated the repeat sequences of the O. corniculata chloroplast genome with the related species. The results revealed that O. corniculata contains 12 palindromic, 30 forward, and 32 tandem repeats, O. drummondii contains 19 palindromic, 20 forward, and 23 tandem repeats, A. carambola contains 15 palindromic, 18 forward, and 27 tandem repeats, and C. follicularis contain s20 palindromic, 30 forward, and 32 tandem repeats (Figure 5A). In O. corniculata, out of these repeats, the sizes of 9 palindromic repeats were 30–44 bp, while the sizes of 4 repeats were 45–59 bp. Likewise, the size of 28 and 3 tandem repeats were 15–29 bp and 30–44 bp, respectively, whereas the size of 21 forward repeats was found to be about 30–44 bp (Figure 5B–D). Amongst all these cp genomes 74 repeats (the highest) were detected in both A. carambola and O. corniculata. In all types of repeats, tandem repeats are the highest in number in all the cp genomes, followed by forward and palindromic repeats.

2.5. Simple Sequence Repeat (SSR) Analysis

In simple sequence repeats (SSRs) analyses, a total of 46 SSRs were detected in the O. corniculata genome; among them, 42 are mononucleotide repeats, 3 are trinucleotide repeats, and 1 is a pentanucleotides repeat (Figure 6). There are no dinucleotides, tetranucleotides, and hexanucleotides in the O. corniculata genome. In O. corniculata, 13% SSRs are present in the CDS region, 73.9% is present in the LSC region, 15.2% is present in the SSC region, and 2.1% is present in the IR region (Figure 6B–E). Similarly, the highest numbers of SSRs in the other three species are located in the intergenic regions; i.e., O. drummondii (70%), A. carambola (87.6%), and C. follicularis (78.5%) followed by the LSC region—that is, 67.5%, 84.6%, and 73.2%, respectively (Figure 6B–E). On the other hand, in the cp genome of O. drummondii, a total of 40 SSRs were found, of which 37 are mono and 3 are trinucleotide, while di, tetra, Penta and hexanucleotides were not detected. In A. carambola and C. follicularis, 56 and 49 are mononucleotide repeats (Figure 6F). A. carambola have 5 trinucleotides, 3 penta, and 1 hexanucleotides repeat, while di and tetranucleotides repeats were missing. Similarly, in C. follicularis, 4 tri, 1 penta, and 2 hexanucleotides repeats were found, while di and tetranucleotides repeats were absent in this cp genome. Among the four cp genomes, A. carambola has a high number of SSRs; i.e., 56.

2.6. Phylogenetic Analysis

For the phylogenetic analysis of O. corniculata, we have downloaded about 191 genomes from the 20 families mentioned in the Materials and Methods section. We inferred the phylogenetic position of O. corniculata on the basis of 60 shared genes among these genomes. The study revealed that O. corniculata forms a single clade with O. drummondii and A. carambola in the family Oxalidaceae (Figure 7). These results also showed that O. corniculata is closer to O. drummondii than A. carambola, which is a different genus. Furthermore, the phylogenetic tree also inferred that the Oxalidaceae family is close to Cephalotaceae and Celastraceae with high bootstrap support (100%), followed by Zygophyllaceae and Euphorbiaceae. The phylogenetic trees in this study also exhibited that Rosaceae is highly interlinked to Moraceae. Similarly, the phylogenetic trees also indicate the close relationship of Fabaceae with Apodanthaceae.

3. Discussion

In case of genetic and evolutionary relationship assessments among plant species, chloroplast DNA sequences have been extensively used [40,41,42]. The complete chloroplast genome sequences provided sufficient information to reconstruct both current and prehistoric diversifications [43]. The powerful and flexible nature of Next Generation Sequencing (NGS) has permeated many areas of study, enabling the development of a broad range of applications that have transformed study designs capable of unlocking information of the genome, transcriptome, and epigenome of any organism [44]. In the current study, we have sequenced the complete genome of O. corniculata chloroplast for the first time. The results revealed that the chloroplast genome size of O. corniculata is in line with the chloroplast of those flowering plants, which ranges from 125,373 bp to 176,045 bp in Cuscuta exaltata and Vaccinium macrocarpon, respectively [45,46]. The CG content of O. corniculata is 36.7% (Table 1), which is slightly lower than C. follicularis and Paeonia obovata (38.43%) [47]. The GC content in the IR region is higher (42.6%) than that of the LSC and SSC regions. As a result of the presence of the rich GC nucleotide, higher GC content was present in the IR region of rRNA genes such as rrn5, rrn4.5, rrn23, and rrn16, which is consistent with what has been investigated in other cp genomes [48,49,50].
In most angiosperms, it is believed that the gene(s) of the chloroplast genome and their organization are extremely conserved [51]. In correlation, we detected 131 genes in the cp genome of O. corniculata while other studies also show that many angiosperms have retained these genes [52,53]. With the increasing number of chloroplast genome sequences, the diverse organization of the chloroplast genome is becoming more evident, as demonstrated by genome rearrangement and gene losses in the chloroplast genomes of Oxiladaceae. For example, the rps16 gene, which is found in most angiosperm plastid genomes, has been lost in O. corniculata. Similar results have been reported in various cp genomes previously [54]. Furthermore, in O. corniculata, the cp genome the infA gene was lost, as reported previously by various researchers, the infA gene has been independently lost multiple times from angiosperms and especially in most Rosids [32,51]. Moreover, we found 11 protein-coding genes and 6 tRNAs genes containing introns in the O. corniculata cp genome. Among them, three genes, clpP, rps12, and ycf3 have two introns, while the others have one intron. Similar results were also reported previously in the Manihot esculenta chloroplast genome [32] and Oresitrophe chloroplast genome [55]. In this study, genes ccsA, clpP, rps8, rps15, rpl22, matK, and ycf1 were found to have high evolution rates among the four cp genomes (Figure 3), which agreed with earlier reports of Cuenoud et al. [41]. Similar results of these genes were reported previously among 17 vascular plants and Panax species [56].
In the terrestrial plants, the cp genome is very conserved structurally, and the large inverted repeats (IRs) junction is not essential to the function of the cp genome [57]. It is believed that IRs are the most conserved region due to which the rate of natural nucleotide substitution in IRs is lesser as compared to single copy regions, and the variation in IR/LSC and IR/SSC boundaries is the key reason for the size variation among the cp genomes of different groups. The variation in size among four genomes was exhibited by the slight expansion of the IRb (JLB border) in C. follicularis compared to O. corniculata (Figure 4). These results are in agreement with previous work where IRs are one of the efficient tools for conformational reorganizations within the plastids genomes and are regularly subjected to expansion, contraction, or even complete loss [20]. Similarly, previous results showed that contractions and expansions of the IR regions triggered the diversification of size among the cp genomes [58].
The study of different repeats (palindromic, forward, and tandem) in our sequenced cp genome showed variation in the number of repeats, which is similar to other species previously studied [59]. In all types of repeats, tandem repeats were found more than palindromic and forward repeats in four cp genomes; these results are consistent with previous reports of Teucrium and Commiphora species [60,61], as well as S. miltiorrhiza [62]. Similarly, simple sequence repeats (SSRs) usually have a higher rate of mutation compared with other neutral regions of DNA due to slipped strand mispairing. In genetic studies, due to the haploid and nonparental inheritance nature of cp SSRs, they are commonly used for the assessment of population structure as molecular markers [63,64]. In this study, we comparatively studied the ideal SSRs among the four species O. corniculata, O. drummondii, A. carambola, and C. follicularis (Figure 6). The largest number of SSRs was found in A. carambola, followed by C. follicularis. Mononucleotide repeats were found to be the most common type of SSR in all four species; the A or T mononucleotide repeats are most abundant SSRs in O. corniculata (Figure 6), which is congruent to the previous result that SSRs in the chloroplast genome are commonly composed of A or T repeats and rarely G or C repeats [62,65].
Recently, cp genomes information has provided a large amount of data for improving phylogenetic resolution. Chloroplast genome sequences have been widely used for the reconstruction of phylogenetic relationships among plant lineages [66,67,68,69]. The phylogenetic evaluation of plant species might not be easy to resolve evolutionary relationships, specifically at taxonomic levels while using a small number of loci [70,71]. Previous phylogenetic studies based on the complete cp genomes and shared genes have been used to explain problematic phylogenetic relations among nearly associated species [34,68] and to increase our concept related to evolutionary relations of angiosperms [72,73]. Phylogenetic relationships of O. corniculata were inferred by using 60 shared genes datasets by using the ML method. The results showed that O. cornicualata form a single clade with O. drummondii. Similarly, the phylogenetic tree also inferred that Oxalidaceae family is close to Cephalotaceae and Celastraceae with high bootstrap support (100%), followed by Zygophyllaceae and Euphorbiaceae (Figure 7). Further cp genomes from the family Oxalidaceae should be explored to determine the phylogenetic position of O. corniculata within the section Corniculatae.

4. Materials and Methods

4.1. Chloroplast DNA Extraction, Sequencing, and Assembly

Young and immature leaves of O. corniculata were ground into fine powder in liquid nitrogen, and pure DNA was isolated through a DNeasy Plant Mini Kit (Qiagen, Valencia, CA, USA). The resultant chloroplast DNA, by using an Illumina HiSeq-2000 platform (San Diego, CA, USA) at Macrogen (Seoul, Korea) was sequenced. A total of 43,453,336 raw reads were generated for O. corniculata, and CLC Genomics Workbench v7.0 (CLC Bio, Aarhus, Denmark) was used to trim and filter reads for the de novo genome assembly. Trimmomatic 0.36 was used for filtering the reads and trailing and leading nucleotide with a Phred score of <20 or when the Phred score dropped below 20 on implementing a 4-bp sliding-window approach. Similarly, reads of <50 bp were discarded after quality filtering and adaptor trimming. The first assembly was formed using SPADES v3.9.0, with an additional switchover to SOAP denovo v2.04. The resulting contigs were compared against the chloroplast genomes of O. drummondii using BLASTN with an E-value cut-off of 1 × 10−5. The uncertain regions in these genomes, such as IR junction’s region, were chosen from the already published genome mentioned above to adjust the sequence length using the iteration method and by employing the Geneious v11.1.2 software [74]. The chloroplast genome sequence of O. corniculata has been submitted to GenBank (accession number: MN998500).

4.2. Genome Annotation

The Dual Organellar Genome Annotator (DOGMA) [75] was used to annotate the cp genomes of the sequenced species and through BLASTX, the number and position of ribosomal RNAs, transfer RNAs, and other coding genes present in chloroplast genomes were identified and analyzed, while BLASTN tRNAscan-SE version 1.21 was used for tRNA annotation [76] software. Furthermore, Geneious (v11.0) and tRNAscan-SE [76] were used for manual adjustment to compare with the reference genomes reported previously. Similarly, the start and stop codon and intron boundaries were also manually adjusted and compared with the reference chloroplast genome already published. Additionally, by using Organellar Genome DRAW (OGDRAW) [77], the structural characteristics of chloroplast genomes of O. corniculata were demonstrated. Beside this, to determine the relative synonymous codon usage and deviations in synonymous codon usage by avoiding the effect of amino acid composition, MEGA6 software [78] was used.

4.3. Characterization of Repetitive Sequences and SSR

REPuter software [79] was used to determine the repetitive sequences (palindromic, reverse and direct repeats) within these four cp genomes (O. corniculata, O. drummondii, A. carambola, and C. follicularis). Subsequent settings were used for repeat identification through REPuter: (1) a minimum repeat size of 30 bp, (2) ≥90% sequence identity, and (3) a Hamming distance of 1. Tandem Repeats Finder version 4.07 b was used to find tandem repeats by using default settings [80]. The MIcroSAtellite (MISA) identification tool was used for the microsatellite analysis of O. corniculata and another three species’ (O. drummondii, A. carambola, and C. follicularis) cp genomes [81]. The parameters such as unit_size and min_repeats were defined as follows: 1–10, 2–8, 3–4, 4–4, 5–3, and 6–3; the smallest distance between two SSRs was set to 100 bp. The following conditions were set for parametric significance: 10 or more repeats of one base, 6 or more repeats of two bases, 5 or more repeats of three bases, 5 or more repeats of four bases, 4 or more repeats of five bases, and 4 or more repeats of six bases.

4.4. Sequence Divergence and Phylogenetic Analysis

In the O. corniculata chloroplast genome, the average pairwise sequence divergence with three related species (O. drummondii, A. carambola, and C. follicularis) from the family Oxalidaceae was determined. After a comparison of gene order and multiple sequence alignment, comparative sequence analysis was used to recognize missing and unclear gene annotations. For whole genome alignment, MAFFT version 7.222 [82], with default parameters were used, and pairwise sequence divergence was calculated by the use of the selected Kimura’s two-parameter (K2P) model [83]. MEGA 6 software [78] was used to evaluate the relative synonymous codon usage by avoiding the effect of amino acid composition. Finally, the divergence of the new O. corniculata cp genomes from related species of family Oxalidaceae was determined using mVISTA [84] in Shuffle-LAGAN mode and by employing the genome of new O. corniculata as a reference. To resolve the phylogenetic position of O. corniculata within the family Oxalidaceae and to check the relationship of 20 families (Fabaceae, Apodanthaceae, Zygophyllaceae, Cephalotaceae, Oxalidaceae, Celastraceae, Euphorbiaceae, Malpighiaceae, Chrysobalanaceae, Violaceae, Passifloraceae, Salicaceae, Cucurbitaceae, Fagaceae, Juglandaceae, Betulaceae, Elaeagnaceae, Ulmaceae, Cannabaceae, Moraceae, Rosaceae) in monophyletic clade rosids, about 60 share genes from 191 cp genomes were downloaded from the National Center for Biotechnology Information (NCBI) database. For the alignment of 60 shared genes, MAFFT version 7.222 [82] with default parameters was used. The maximum likelihood (ML) method was adopted to infer the phylogenetic trees with MEGA 6 [78], and parameters were adjusted with a BIONJ tree with 1000 bootstrap replicates using the Kimura two-parameter model with gamma-distributed rate heterogeneity and invariant sites.

5. Conclusions

The current findings reveal detailed understandings of the complete cp genome of O. corniculata for the first time through sequencing on Illumina HiSeq-2000 platform. The gene order and gene structure of O. corniculata was found to be similar with three related species from the family Oxalidaceae. Through detailed bioinformatic analysis and comparative assessments, we retrieved essential genetic features such as repetitive sequences, SSRs, codon usage, IR contraction and expansion, sequence divergence, and phylogenomic placement. Whole cp genome comparisons revealed an overall high degree of sequence similarity between O. corniculata and O. drummondii and some divergence in the intergenic spacers of other species. No major structural rearrangement in these four cp genomes was observed. Phylogenomic analyses of the complete plastid genomes revealed that O. corniculata is closely related to O. drummondii. A current plastome genomic dataset and the detailed analysis of O. corniculata and related species and their comparative analysis provide a powerful genetic resource for the future molecular phylogeny, evolution, population genetics, and biological functions of genus Oxalis.

Supplementary Materials

The following are available online at https://www.mdpi.com/2223-7747/9/8/928/s1, Table S1. Average pairwise sequence distance of O. corniculata with related species cp genomes.

Author Contributions

L., S.A. performed experiments; A.L.K., S.A. and R.J. wrote the original draft and bioinformatics analysis: A.L.K., I.-J.L. supervision and arranging resources. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Agenda Program (Project No. PJ015026022020) Rural Development Administration, Republic of Korea.

Conflicts of Interest

The authors have declared that no competing interests exist.

Availability of Data and Materials

All data generated or analyzed during this study are included in this published article.

References

  1. Denton, M.F. Monograph of Oxalis, Section Ionoxalis (Oxalidaceae) in North America; Michigan State University: East Lansing, MI, USA, 1973. [Google Scholar]
  2. Lourteig, A. Oxalis, L. subg é nero Monoxalis (Small) Lourteig, Oxalis y Trifidus Lourteig; Herbarium Bradeanum: Rio de Janeiro, Brazil, 2000. [Google Scholar]
  3. De Azkue, D. Chromosome diversity of South American Oxalis (Oxalidaceae). Bot. J. Linn. Soc. 2000, 132, 143–152. [Google Scholar] [CrossRef]
  4. Groom, Q.J.; Van der Straeten, J.; Hoste, I. The origin of Oxalis corniculata L. PeerJ 2019, 7, e6384. [Google Scholar] [CrossRef] [Green Version]
  5. Lourteig, A. Oxalidaceae extra-austroamericanae. II. Oxalis, L. Sectio Corniculatae DC. Phytologia 1979, 42, 57–198. [Google Scholar]
  6. Marks, G. Chromosome numbers in the genus Oxalis. New Phytol. 1956, 55, 120–129. [Google Scholar] [CrossRef]
  7. Naranjo, C. Estudios citotaxonomicos y evolutivos en especies herbaceas sudamericanas de Oxalis (Oxalidaceae) I. Bol. Soc. Argent. Bot 1982, 20, 183–200. [Google Scholar]
  8. Groom, Q.; Hoste, I.; Janssens, S. Observación confirmada de Oxalis dillenii en España. Collect. Bot. 2017, 36, 4. [Google Scholar] [CrossRef] [Green Version]
  9. Pyšek, P.; Pergl, J.; Essl, F.; Lenzner, B.; Dawson, W.; Kreft, H.; Weigelt, P.; Winter, M.; Kartesz, J.; Nishino, M.; et al. Naturalized alien flora of the world. Preslia 2017, 89, 203–274. [Google Scholar] [CrossRef]
  10. Marble, S.C.; Andrew, K.K.; Gitta, H.; Drew, M.; Annette, C. Efficacy and estimated annual cost of common weed control methods in landscape planting beds. HortTechnology 2017, 27, 199–211. [Google Scholar] [CrossRef]
  11. Rehman, A.; Rehman, A.; Ahmad, I. Antibacterial, antifungal, and insecticidal potentials of Oxalis corniculata and its isolated compounds. Int. J. Anal. Chem. 2015, 2015, 1–5. [Google Scholar] [CrossRef] [Green Version]
  12. Duke, J.A.; Ayensu, E.S. Medicinal Plants of China; Reference Publications: Algonac, MI, USA, 1985; Volume 2. [Google Scholar]
  13. Neuhaus, H.; Emes, M. Nonphotosynthetic metabolism in plastids. Ann. Rev. Plant Biol. 2000, 51, 111–140. [Google Scholar] [CrossRef]
  14. Rodríguez-Ezpeleta, N.; Brinkmann, H.; Burey, S.C.; Roure, B.; Burger, G.; Löffelhardt, W.; Bohnert, H.J.; Philippe, H.; Lang, B.F. Monophyly of primary photosynthetic eukaryotes: Green plants, red algae, and glaucophytes. Curr. Biol. 2005, 15, 1325–1330. [Google Scholar] [CrossRef] [Green Version]
  15. McCauley, D.E.; Sundby, A.K.; Bailey, M.F.; Welch, M.E. Inheritance of chloroplast DNA is not strictly maternal in Silene vulgaris (Caryophyllaceae): Evidence from experimental crosses and natural populations. Am. J. Bot. 2007, 94, 1333–1337. [Google Scholar] [CrossRef] [Green Version]
  16. Reboud, X.; Zeyl, C. Organelle inheritance in plants. Heredity 1994, 72, 132–140. [Google Scholar] [CrossRef] [Green Version]
  17. Allen, J.F. Why chloroplasts and mitochondria retain their own genomes and genetic systems: Colocation for redox regulation of gene expression. Proc. Natl. Acad. Sci. USA 2015, 112, 10231–10238. [Google Scholar] [CrossRef] [Green Version]
  18. Olmstead, R.G.; Palmer, J.D. Chloroplast DNA systematics: A review of methods and data analysis. Am. J. Bot. 1994, 81, 1205–1224. [Google Scholar] [CrossRef]
  19. Chumley, T.W.; Palmer, J.D.; Mower, J.P.; Fourcade, H.M.; Calie, P.J.; Boore, J.L.; Jansen, R.K. The complete chloroplast genome sequence of Pelargonium × hortorum: Organization and evolution of the largest and most highly rearranged chloroplast genome of land plants. Mol. Biol. Evol. 2006, 23, 2175–2190. [Google Scholar] [CrossRef]
  20. Wicke, S.; Schneeweiss, G.M.; de Pamphilis, 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] [Green Version]
  21. Lee, H.L.; Jansen, R.K.; Chumley, T.W.; Kim, K.J. Gene relocations within chloroplast genomes of Jasminum and Menodora (Oleaceae) are due to multiple, overlapping inversions. Mol. Biol. Evol. 2007, 24, 1161–1180. [Google Scholar] [CrossRef] [Green Version]
  22. Greiner, S.; Wang, X.; Herrmann, R.G.; Rauwolf, U.; Mayer, K.; Haberer, G.; Meurer, J. The complete nucleotide sequences of the 5 genetically distinct plastid genomes of Oenothera, subsection Oenothera: II. A microevolutionary view using bioinformatics and formal genetic data. Mol. Biol. Evol. 2008, 25, 2019–2030. [Google Scholar] [CrossRef]
  23. Frailey, D.C.; Chaluvadi, S.R.; Vaughn, J.N.; Coatney, C.G.; Bennetzen, J.L. Gene loss and genome rearrangement in the plastids of five Hemiparasites in the family Orobanchaceae. BMC Plant. Biol. 2018, 18, 30. [Google Scholar] [CrossRef] [Green Version]
  24. Barrett, C.F.; Freudenstein, J.V.; Li, J.; Mayfield-Jones, D.R.; Perez, L.; Pires, J.C.; Santos, C. Investigating the path of plastid genome degradation in an early-transitional clade of heterotrophic orchids, and implications for heterotrophic angiosperms. Mol. Biol. Evol. 2014, 31, 3095–3112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Fan, W.-B.; Wu, Y.; Yang, J.; Shahzad, K.; Li, Z.-H. Comparative Chloroplast Genomics of Dipsacales Species: Insights Into Sequence Variation, Adaptive Evolution, and Phylogenetic Relationships. Front. Plant Sci. 2018, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Shaw, J.; Lickey, E.B.; Beck, J.T.; Farmer, S.B.; Liu, W.; Miller, J.; Siripun, K.C.; Winder, C.T.; Schilling, E.E.; Small, R.L. The tortoise and the hare II: Relative utility of 21 noncoding chloroplast DNA sequences for phylogenetic analysis. Am. J. Bot. 2005, 92, 142–166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. 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: The tortoise and the hare IV. Am. J. Bot. 2014, 101, 1987–2004. [Google Scholar] [CrossRef] [Green Version]
  28. Koehler, S.; Cabral, J.S.; Whitten, W.M.; Williams, N.H.; Singer, R.B.; Neubig, K.M.; Guerra, M.; Souza, A.P.; Amaral, M.D.C.E. Molecular phylogeny of the neotropical genus Christensonella (Orchidaceae, Maxillariinae): Species delimitation and insights into chromosome evolution. Ann. Bot. 2008, 102, 491–507. [Google Scholar] [CrossRef] [Green Version]
  29. Stuessy, T.F.; Blöch, C.; Villaseñor, J.L.; Rebernig, C.A.; Weiss-Schneeweiss, H. Phylogenetic analyses of DNA sequences with chromosomal and morphological data confirm and refine sectional and series classification within Melampodium (Asteraceae, Millerieae). Taxon 2011, 60, 436–449. [Google Scholar] [CrossRef]
  30. Ruhfel, B.R.; Gitzendanner, M.A.; Soltis, P.S.; Soltis, D.E.; Burleigh, J.G. From algae to angiosperms–inferring the phylogeny of green plants (Viridiplantae) from 360 plastid genomes. BMC Evol. Biol. 2014, 14, 23. [Google Scholar] [CrossRef] [Green Version]
  31. Soltis, D.E.; Smith, S.A.; Cellinese, N.; Wurdack, K.J.; Tank, D.C.; Brockington, S.F.; Refulio-Rodriguez, N.F.; Walker, J.B.; Moore, M.J.; Carlsward, B.S.; et al. Angiosperm phylogeny: 17 genes, 640 taxa. Am. J. Bot. 2011, 98, 704–730. [Google Scholar] [CrossRef] [Green Version]
  32. Jansen, R.K.; Cai, Z.; Raubeson, L.A.; Daniell, H.; Depamphilis, C.W.; Leebens-Mack, J.; Müller, K.F.; Guisinger-Bellian, M.; Haberle, R.C.; Hansen, A.K.; et al. Analysis of 81 genes from 64 plastid genomes resolves relationships in angiosperms and identifies genome-scale evolutionary patterns. Proc. Natl. Acad. Sci. USA 2007, 104, 19369–19374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Parks, M.; Cronn, R.; Liston, A. Increasing phylogenetic resolution at low taxonomic levels using massively parallel sequencing of chloroplast genomes. BMC Biol. 2009, 7, 84. [Google Scholar] [CrossRef] [Green Version]
  34. Carbonell-Caballero, J.; Alonso, R.; Ibañez, V.; Terol, J.; Talon, M.; Dopazo, J. A Phylogenetic analysis of 34 chloroplast genomes elucidates the relationships between wild and domestic species within the genus citrus. Mol. Biol. Evol. 2015, 32, 2015–2035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Moore, M.J.; Bell, C.D.; Soltis, P.S.; Soltis, D.E. Using plastid genome-scale data to resolve enigmatic relationships among basal angiosperms. Proc. Natl. Acad. Sci. USA 2007, 104, 19363–19368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Asaf, S.; Khan, A.L.; Khan, A.; Al-Harrasi, A. Unraveling the chloroplast genomes of two prosopis species to identify its genomic information, comparative analyses and phylogenetic relationship. Int. J. Mol. Sci. 2020, 21, 3280. [Google Scholar] [CrossRef] [PubMed]
  37. Oberlander, K.C.; Dreyer, L.L.; Bellstedt, D.U. Molecular phylogenetics and origins of southern African Oxalis. TAXON 2011, 60, 1667–1677. [Google Scholar] [CrossRef]
  38. PM, M. Cytology of Oxalidaceae. Cytologia 1958, 23, 200–210. [Google Scholar]
  39. Vaio, M.; Gardner, A.; Emshwiller, E.; Guerra, M. Molecular phylogeny and chromosome evolution among the creeping herbaceous Oxalis species of sections Corniculatae and Ripariae (Oxalidaceae). Mol. Phyl. Evol. 2013, 68, 199–211. [Google Scholar] [CrossRef]
  40. Ohnishi, O.; Matsuoka, Y. Search for the wild ancestor of buckwheat II. Taxonomy of Fagopyrum (Polygonaceae) species based on morphology, isozymes and cpDNA variability. Genes Genet. Syst. 1996, 71, 383–390. [Google Scholar] [CrossRef] [Green Version]
  41. Cuénoud, P.; Savolainen, V.; Chatrou, L.W.; Powell, M.; Grayer, R.J.; Chase, M.W. Molecular phylogenetics of Caryophyllales based on nuclear 18S rDNA and plastid rbcL, atpB, and matK DNA sequences. Am. J. Bot. 2002, 89, 132–144. [Google Scholar] [CrossRef]
  42. Yamane, K.; Yasui, Y.; Ohnishi, O. Intraspecific cpDNA variations of diploid and tetraploid perennial buckwheat, Fagopyrum cymosum (Polygonaceae). Am. J. Bot. 2003, 90, 339–346. [Google Scholar] [CrossRef]
  43. Cui, Y.; Qin, S.; Jiang, P. Chloroplast transformation of Platymonas (Tetraselmis) subcordiformis with the bar gene as selectable marker. PLoS ONE 2014, 9, e98607. [Google Scholar] [CrossRef] [Green Version]
  44. McPherson, H.; Van der Merwe, M.; Delaney, S.K.; Edwards, M.A.; Henry, R.J.; McIntosh, E.; Rymer, P.D.; Milner, M.L.; Siow, J.; Rossetto, M. Capturing chloroplast variation for molecular ecology studies: A simple next generation sequencing approach applied to a rainforest tree. BMC Ecol. 2013, 13, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. McNeal, J.R.; Kuehl, J.V.; Boore, J.L.; de Pamphilis, C.W. Complete plastid genome sequences suggest strong selection for retention of photosynthetic genes in the parasitic plant genus Cuscuta. BMC Plant Biol. 2007, 7, 57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Fajardo, D.; Senalik, D.; Ames, M.; Zhu, H.; Steffan, S.A.; Harbut, R.; Polashock, J.; Vorsa, N.; Gillespie, E.; Kron, K. Complete plastid genome sequence of Vaccinium macrocarpon: Structure, gene content, and rearrangements revealed by next generation sequencing. Tree Genet. Genomes 2013, 9, 489–498. [Google Scholar] [CrossRef]
  47. Dong, W.; Xu, C.; Cheng, T.; Zhou, S. Complete chloroplast genome of Sedum sarmentosum and chloroplast genome evolution in Saxifragales. PLoS ONE 2013, 8, e77965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Curci, P.L.; De Paola, D.; Danzi, D.; Vendramin, G.G.; Sonnante, G. Complete chloroplast genome of the multifunctional crop globe artichoke and comparison with other Asteraceae. PLoS ONE 2015, 10, e0120589. [Google Scholar] [CrossRef]
  49. Yang, J.-B.; Yang, S.-X.; Li, H.-T.; Yang, J.; Li, D.-Z. Comparative chloroplast genomes of Camellia species. PLoS ONE 2013, 8, e73053. [Google Scholar] [CrossRef] [Green Version]
  50. Asaf, S.; Khan, A.L.; Khan, A.; Khan, G.; Lee, I.-J.; Al-Harrasi, A. Expanded inverted repeat region with large scale inversion in the first complete plastid genome sequence of Plantago ovata. Sci. Rep. 2020, 10, 1–16. [Google Scholar] [CrossRef] [Green Version]
  51. Jansen, R.K.; Saski, C.; Lee, S.-B.; Hansen, A.K.; Daniell, H. Complete plastid genome sequences of three rosids (Castanea, Prunus, Theobroma): Evidence for at least two independent transfers of rpl22 to the nucleus. Mol. Biol. Evol. 2011, 28, 835–847. [Google Scholar] [CrossRef] [Green Version]
  52. Nie, X.; Lv, S.; Zhang, Y.; Du, X.; Wang, L.; Biradar, S.S.; Tan, X.; Wan, F.; Weining, S. Complete chloroplast genome sequence of a major invasive species, crofton weed (Ageratina adenophora). PLoS ONE 2012, 7, e36869. [Google Scholar] [CrossRef] [Green Version]
  53. Kim, K.-J.; Lee, H.-L. Complete chloroplast genome sequences from Korean ginseng (Panax schinseng Nees) and comparative analysis of sequence evolution among 17 vascular plants. DNA Res. 2004, 11, 247–261. [Google Scholar] [CrossRef]
  54. Jansen, R.K.; Wojciechowski, M.F.; Sanniyasi, E.; Lee, S.-B.; Daniell, H. Complete plastid genome sequence of the chickpea (Cicer arietinum) and the phylogenetic distribution of rps12 and clpP intron losses among legumes (Leguminosae). Mol. Phylogenetics Evol. 2008, 48, 1204–1217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Liu, L.; Wang, Y.; He, P.; Li, P.; Lee, J.; Soltis, D.E.; Fu, C. Chloroplast genome analyses and genomic resource development for epilithic sister genera Oresitrophe and Mukdenia (Saxifragaceae), using genome skimming data. BMC Genom. 2018, 19, 235. [Google Scholar] [CrossRef] [PubMed]
  56. Wu, C.-S.; Chaw, S.-M. Evolutionary stasis in cycad plastomes and the first case of plastome GC-biased gene conversion. Genome Biol. Evol. 2015, 7, 2000–2009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Goulding, S.E.; Wolfe, K.; Olmstead, R.; Morden, C. Ebb and flow of the chloroplast inverted repeat. Mol. Gen. Genet. MGG 1996, 252, 195–206. [Google Scholar] [CrossRef] [PubMed]
  58. Kim, Y.-K.; Park, C.-W.; Kim, K.-J. Complete chloroplast DNA sequence from a Korean endemic genus, Megaleranthis saniculifolia, and its evolutionary implications. Mol. Cells 2009, 27, 365. [Google Scholar] [CrossRef]
  59. Asaf, S.; Khan, A.L.; Khan, M.A.; Waqas, M.; Kang, S.-M.; Yun, B.-W.; Lee, I.-J. Chloroplast genomes of Arabidopsis halleri ssp. gemmifera and Arabidopsis lyrata ssp. petraea: Structures and comparative analysis. Sci. Rep. 2017, 7, 1–15. [Google Scholar] [CrossRef] [Green Version]
  60. 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] [Green Version]
  61. Khan, A.; Asaf, S.; Khan, A.L.; Khan, A.; Al-Harrasi, A.; Al-Sudairy, O.; AbdulKareem, N.M.; Al-Saady, N.; Al-Rawahi, A. Complete chloroplast genomes of medicinally important Teucrium species and comparative analyses with related species from Lamiaceae. PeerJ 2019, 7, e7260. [Google Scholar] [CrossRef] [Green Version]
  62. Qian, J.; Song, J.; Gao, H.; Zhu, Y.; Xu, J.; Pang, X.; Yao, H.; Sun, C.; Li, X.E.; Li, C.; et al. The Complete Chloroplast Genome Sequence of the Medicinal Plant Salvia miltiorrhiza. PLoS ONE 2013, 8, e57607. [Google Scholar] [CrossRef]
  63. Leclercq, S.; Rivals, E.; Jarne, P. Detecting microsatellites within genomes: Significant variation among algorithms. BMC Bioinf. 2007, 8, 125. [Google Scholar] [CrossRef] [Green Version]
  64. Echt, C.S.; Deverno, L.L.; Anzidei, M.; Vendramin, G.G. Chloroplast microsatellites reveal population genetic diversity in red pine, Pinus resinosa Ait. Mol. Ecol. 1998, 7, 307–316. [Google Scholar] [CrossRef] [Green Version]
  65. Kuang, D.-Y.; Wu, H.; Wang, Y.-L.; Gao, L.-M.; Zhang, S.-Z.; Lu, L. Complete chloroplast genome sequence of Magnolia kwangsiensis (Magnoliaceae): Implication for DNA barcoding and population genetics. Genome 2011, 54, 663–673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Burke, S.V.; Lin, C.-S.; Wysocki, W.P.; Clark, L.G.; Duvall, M.R. Phylogenomics and Plastome Evolution of Tropical Forest Grasses (Leptaspis, Streptochaeta: Poaceae). Front. Plant Sci. 2016, 7, 1993. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Dong, W.; Xu, C.; Li, W.; Xie, X.; Lu, Y.; Liu, Y.; Jin, X.; Suo, Z. Phylogenetic resolution in Juglans based on complete chloroplast genomes and nuclear DNA sequences. Front. Plant Sci. 2017, 8. [Google Scholar] [CrossRef] [Green Version]
  68. 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, 1–10. [Google Scholar] [CrossRef] [Green Version]
  69. Sun, L.; Fang, L.; Zhang, Z.; Chang, X.; Penny, D.; Zhong, B. Chloroplast phylogenomic inference of green algae relationships. Sci. Rep. 2016, 6, 20528. [Google Scholar] [CrossRef] [Green Version]
  70. Majure, L.C.; Puente, R.; Griffith, M.P.; Judd, W.S.; Soltis, P.S.; Soltis, D.E. Phylogeny of Opuntia s.s. (Cactaceae): Clade delineation, geographic origins, and reticulate evolution. Am. J. Bot. 2012, 99, 847–864. [Google Scholar] [CrossRef] [Green Version]
  71. Hilu, K.W.; Alice, L.A. A phylogeny of Chloridoideae (Poaceae) based on matK sequences. Syst. Bot. 2001, 26, 386–405. [Google Scholar]
  72. Goremykin, V.V.; Nikiforova, S.V.; Biggs, P.J.; Zhong, B.; Delange, P.; Martin, W.; Woetzel, S.; Atherton, R.A.; McLenachan, P.A.; Lockhart, P.J. The evolutionary root of flowering plants. Syst. Biol. 2013, 62, 50–61. [Google Scholar] [CrossRef] [Green Version]
  73. Luo, Y.; Ma, P.-F.; Li, H.-T.; Yang, J.-B.; Wang, H.; Li, D.-Z. Plastid Phylogenomic Analyses Resolve Tofieldiaceae as the Root of the Early Diverging Monocot Order Alismatales. Genome Biol. Evol. 2016, 8, 932–945. [Google Scholar] [CrossRef] [Green Version]
  74. 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] [PubMed]
  75. 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]
  76. Schattner, P.; Brooks, A.N.; Lowe, T.M. The tRNAscan-SE, snoscan and snoGPS web servers for the detection of tRNAs and snoRNAs. Nucleic Acids Res. 2005, 33, W686–W689. [Google Scholar] [CrossRef] [PubMed]
  77. 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. Gen. 2007, 52, 267–274. [Google Scholar] [CrossRef] [PubMed]
  78. Kumar, S.; Nei, M.; Dudley, J.; Tamura, K. MEGA: A biologist-centric software for evolutionary analysis of DNA and protein sequences. Brief. Bioinf. 2008, 9, 299–306. [Google Scholar] [CrossRef] [Green Version]
  79. Kurtz, S.; Choudhuri, J.V.; Ohlebusch, E.; Schleiermacher, C.; Stoye, J.; Giegerich, R. REPuter: The manifold applications of repeat analysis on a genomic scale. Nucl. Acids Res. 2001, 29, 4633–4642. [Google Scholar] [CrossRef] [Green Version]
  80. Benson, G. Tandem repeats finder: A program to analyze DNA sequences. Nucl. Acids Res. 1999, 27, 573–580. [Google Scholar] [CrossRef] [Green Version]
  81. 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] [Green Version]
  82. 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]
  83. Kimura, M. A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. J. Mol. Evol. 1980, 16, 111–120. [Google Scholar] [CrossRef]
  84. Frazer, K.A.; Pachter, L.; Poliakov, A.; Rubin, E.M.; Dubchak, I. VISTA: Computational tools for comparative genomics. Nucl. Acids Res. 2004, 32, W273–W279. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Gene map of the O. corniculata cp genome. Thick lines in the red area indicate the extent of the inverted repeat regions (IRa and IRb; 25,387 bp), which separate the genome into small (SSC; 16,990 bp) and large (LSC; 83,427 bp) single copy regions. Genes drawn inside the circle are transcribed clockwise, and those outsides are transcribed counter clockwise. Genes belonging to different functional groups are color-coded. The dark gray in the inner circle corresponds to the GC content, and the light gray corresponds to the AT content.
Figure 1. Gene map of the O. corniculata cp genome. Thick lines in the red area indicate the extent of the inverted repeat regions (IRa and IRb; 25,387 bp), which separate the genome into small (SSC; 16,990 bp) and large (LSC; 83,427 bp) single copy regions. Genes drawn inside the circle are transcribed clockwise, and those outsides are transcribed counter clockwise. Genes belonging to different functional groups are color-coded. The dark gray in the inner circle corresponds to the GC content, and the light gray corresponds to the AT content.
Plants 09 00928 g001
Figure 2. Visual alignment of plastid genomes of O. corniculata and three other members (O. drummondii, A. carambola, and C. follicularis) from the family Oxalidaceae. VISTA-based identity plot showing sequence identity among three species, using O. corniculata as a reference. The vertical scale indicates percent identity, ranging from 50% to 100%. The horizontal axis indicates the coordinates within the chloroplast genome. Arrows indicate the annotated genes and their transcription direction. The thick black lines show the inverted repeats (IRs).
Figure 2. Visual alignment of plastid genomes of O. corniculata and three other members (O. drummondii, A. carambola, and C. follicularis) from the family Oxalidaceae. VISTA-based identity plot showing sequence identity among three species, using O. corniculata as a reference. The vertical scale indicates percent identity, ranging from 50% to 100%. The horizontal axis indicates the coordinates within the chloroplast genome. Arrows indicate the annotated genes and their transcription direction. The thick black lines show the inverted repeats (IRs).
Plants 09 00928 g002
Figure 3. Pairwise sequence distance of 63 genes from of O. corniculata and three related species (O. drummondii, A. carambola, and C. follicularis) from family Oxidalaceae.
Figure 3. Pairwise sequence distance of 63 genes from of O. corniculata and three related species (O. drummondii, A. carambola, and C. follicularis) from family Oxidalaceae.
Plants 09 00928 g003
Figure 4. Distance between adjacent genes and junctions of the small single-copy (SSC), large single-copy (LSC), and two inverted repeat (IR) regions of O. corniculata with related species cp genomes. Boxes above and below the main line indicate the adjacent border genes. The figure is not to scale regarding sequence length, and it only shows relative changes at or near the IR/SC borders.
Figure 4. Distance between adjacent genes and junctions of the small single-copy (SSC), large single-copy (LSC), and two inverted repeat (IR) regions of O. corniculata with related species cp genomes. Boxes above and below the main line indicate the adjacent border genes. The figure is not to scale regarding sequence length, and it only shows relative changes at or near the IR/SC borders.
Plants 09 00928 g004
Figure 5. Analysis and graphical representation of repeated sequences in the four Oxalidaceae cp genomes. (A) Totals numbers of three repeat types; (B) Number of palindromic repeats by length; (C) Number of tandem repeats by length; (D) Number of forward repeats by length.
Figure 5. Analysis and graphical representation of repeated sequences in the four Oxalidaceae cp genomes. (A) Totals numbers of three repeat types; (B) Number of palindromic repeats by length; (C) Number of tandem repeats by length; (D) Number of forward repeats by length.
Plants 09 00928 g005
Figure 6. Analysis and graphical representation of simple sequence repeats (SSRs). Analysis of simple sequence repeats in the four Oxalidaceae chloroplast genomes (A), SSR numbers detected in the four species’ LSC regions (B), SSR numbers detected in the four species’ SSC regions (C), SSR numbers detected in the four species’ IR regions (D), Frequency of identified SSRs in the coding sequences (CDS) region (E) and frequency of identified SSR motifs in different repeat class types (F).
Figure 6. Analysis and graphical representation of simple sequence repeats (SSRs). Analysis of simple sequence repeats in the four Oxalidaceae chloroplast genomes (A), SSR numbers detected in the four species’ LSC regions (B), SSR numbers detected in the four species’ SSC regions (C), SSR numbers detected in the four species’ IR regions (D), Frequency of identified SSRs in the coding sequences (CDS) region (E) and frequency of identified SSR motifs in different repeat class types (F).
Plants 09 00928 g006
Figure 7. Phylogenetic trees of O. corniculata with related species. The 60 shared genes dataset was analyzed by using maximum likelihood (ML). Numbers above the branches represent bootstrap values in the ML. The red star represents the position of O. corniculata (MN998500).
Figure 7. Phylogenetic trees of O. corniculata with related species. The 60 shared genes dataset was analyzed by using maximum likelihood (ML). Numbers above the branches represent bootstrap values in the ML. The red star represents the position of O. corniculata (MN998500).
Plants 09 00928 g007
Table 1. Summary of complete chloroplast genome of O. corniculata and its comparison with related species from family Oxalidaceae.
Table 1. Summary of complete chloroplast genome of O. corniculata and its comparison with related species from family Oxalidaceae.
A. carambolaC. follicularisO. drummondiiO. corniculata
Size (bp)155,965142,706152,112152,189
Overall GC contents36.537.136.536.7
LSC size in bp86,21881,07183,34183,427
SSC size in bp17,49787391691516,990
IR size in bp25,62625,94925,42925,387
Protein coding regions size in bp78,52866,67277,82679,239
tRNA size in bp2790279027903042
rRNA size in bp9046905090469048
Number of genes131122127131
Number of protein coding genes83718283
Number of rRNA8888
Number of tRNA37373740
Table 2. Genes in the sequenced O. corniculata chloroplast genome.
Table 2. Genes in the sequenced O. corniculata chloroplast genome.
CategoryGroup of GenesName of Genes
Self-replicationLarge subunit of ribosomal proteinsrpl2 *, rpl14, rpl16, rpl20, rpl22, rpl23 *, rpl33, rpl36
Small subunit of ribosomal proteinsrps2, rps3, rps4, rps7 *, rps8, rps11, rps12 *, rps14, rps15, rps18, rps19
DNA dependent RNA polymeraserpoA, rpoB, rpoC1, rpoC2
rRNA genesrrn4.5, 5, 16, 23
tRNA genestrnA-UGC *, trnC-GCA, trnD-GUC, trnE-UUC , trnF-GAA, trnfM-CAU, trnG-GCC *, trnH-GUG, trnI-GAU *, trnK-UUU, trnL-CAA *, trnL-UAA, trnL-UAG, trnM-CAU *, trnN-GUU *, trnP-UGG, trnQ-UUG, trnR-ACG *, trnR-UCU, trnS-GCU, trnS-GGA, trnS-UGA, trnT-GGU, trnT-UGU *, trnV-GAC *, trnW-CCA, trnY-GUA
PhotosynthesisPhotosystem IpsaA, B, C, I, J
Photosystem IIpsbA, psbB, psbC, psbD, psbE, psbF, psbH, psbI, psbJ, psbK, psbL, psbM, psbT, psbZ
NadH oxidoreductasendhA, ndhB *, ndhC, ndhD, ndhE, ndhF, ndhG, ndhH, ndhI, ndhJ, ndhK
Cytochrome b6/f complexpetA, petB, petD, pet G, petL, petN
ATP synthaseatpA, atpB, atpE, atpF, atpH, atpI
RubiscorbcL
Other genesMaturasematK
ProteaseclpP
Envelop membrane proteincemA
Subunit Acetyl- CoA-CarboxylateaccD
c-type cytochrome synthesis geneccsA
UnknownConserved Open reading framesycf1 *, 2 *, 3, 4,
* Duplicated genes.
Table 3. The genes with introns in the four species chloroplast genome and the length of exons and introns.
Table 3. The genes with introns in the four species chloroplast genome and the length of exons and introns.
GeneLocationExon I (bp)Intron 1 (bp)Exon II (bp)Intron II (bp)Exon III (bp)
O. cO. dA. cC. fO. cO. dA. cC. fO. cO. dA. cC. fO. cO. dA. cC. fO. cO. dA. cC. f
atpFLSC144145145145717714714717411410410410
petBLSC6666746746785793645642642642
PetDLSC8877746778707721645475473473
rpl2 *IR391391391391658661659671434434434525
rps16LSC----4040----935929----224227
rpoC1LSC4304324534597277297517591634162916081617
rps12 *IR/LSC 391 527 232 434 25
clpPLSC71716969808812833833289289291291 603612568 228228228
ndhASSC557557559--107710671037--541541557--
ndhB*IR777777777--685685685--756756756--
ycf3LSC124124126126718712718714228230228228684678685678155153153150
trnA-UGC *IR3838383879979876381135353535
trnI –GAU *IR3737373792592493294635353535
trnL-UAALSC3535353549749750749250505050
trnK -UUULSC37343735251925152545255835373537
trnG-GCCLSC71717171----------------
trnV-UACLSC----3535----617618----3939
O. c = Oxalis corniculata, O. d = Oxalis drummondii, A. c = Averrhoa carambola, C. f = Cephalotus follicularis.

Share and Cite

MDPI and ACS Style

Lubna; Asaf, S.; Jan, R.; Khan, A.L.; Lee, I.-J. Complete Chloroplast Genome Characterization of Oxalis Corniculata and Its Comparison with Related Species from Family Oxalidaceae. Plants 2020, 9, 928. https://doi.org/10.3390/plants9080928

AMA Style

Lubna, Asaf S, Jan R, Khan AL, Lee I-J. Complete Chloroplast Genome Characterization of Oxalis Corniculata and Its Comparison with Related Species from Family Oxalidaceae. Plants. 2020; 9(8):928. https://doi.org/10.3390/plants9080928

Chicago/Turabian Style

Lubna, Sajjad Asaf, Rahmatullah Jan, Abdul Latif Khan, and In-Jung Lee. 2020. "Complete Chloroplast Genome Characterization of Oxalis Corniculata and Its Comparison with Related Species from Family Oxalidaceae" Plants 9, no. 8: 928. https://doi.org/10.3390/plants9080928

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