Next Article in Journal
Influence of Site-Specific Conditions on Estimation of Forest above Ground Biomass from Airborne Laser Scanning
Next Article in Special Issue
Comparative Plastome Analyses and Phylogenetic Applications of the Acer Section Platanoidea
Previous Article in Journal
Multifunctionality of Forests: A White Paper on Challenges and Opportunities in China and Germany
Previous Article in Special Issue
Constitutive and Cold Acclimation-Regulated Protein Expression Profiles of Scots Pine Seedlings Reveal Potential for Adaptive Capacity of Geographically Distant Populations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Complete Chloroplast Genome of Michelia shiluensis and a Comparative Analysis with Four Magnoliaceae Species

1
Guangdong Key Laboratory for Innovative Development and Utilization of Forest Plant Germplasm, Guangzhou 510642, China
2
College of Forestry and Landscape Architecture, South China Agricultural University, Guangzhou 510642, China
3
Guangzhou Hengsheng Construction Engineering Co., Ltd., Guangzhou 510000, China
*
Author to whom correspondence should be addressed.
Forests 2020, 11(3), 267; https://doi.org/10.3390/f11030267
Submission received: 7 February 2020 / Revised: 24 February 2020 / Accepted: 25 February 2020 / Published: 27 February 2020
(This article belongs to the Special Issue Genetic and Phenotypic Variation in Tree Crops Biodiversity)

Abstract

:
Michelia shiluensis is a rare and endangered magnolia species found in South China. This species produces beautiful flowers and is thus widely used in landscape gardening. Additionally, its timber is also used for furniture production. As a result of low rates of natural reproduction and increasing levels of human impact, wild M. shiluensis populations have become fragmented. This species is now classified as endangered by the IUCN. In the present study, we characterized the complete chloroplast genome of M. shiluensis and found it to be 160,075 bp in length with two inverted repeat regions (26,587 bp each), a large single-copy region (88,105 bp), and a small copy region (18,796 bp). The genome contained 131 genes, including 86 protein-coding genes, 37 tRNAs, and 8 rRNAs. The guanine-cytosine content represented 39.26% of the overall genome. Comparative analysis revealed high similarity between the M. shiluensis chloroplast genome and those of four closely related species: Michelia odora, Magnolia laevifolia, Magnolia insignis, and Magnolia cathcartii. Phylogenetic analysis shows that M. shiluensis is most closely related to M. odora. The genomic information presented in this study is valuable for further classification, phylogenetic studies, and to support ongoing conservation efforts.

1. Introduction

Michelia shiluensis Chun and Y. F. Wu (Magnoliaceae) is an endangered flowering plant that is sparsely distributed throughout Hainan Province, China [1]. It is characterized by leafy branches and beautiful flowers, and is, therefore, widely used in landscape gardening [2]. This species is also a source of excellent quality wood which is in demand for furniture production [3]. In recent decades, there has been a serious decline in wild populations of this species as a result of the illegal harvesting to supply both the timber and horticultural markets [4]. Moreover, this species naturally has a low seeding rate and its wild populations are declining [5]. Consequently, M. shiluensis is categorized as a Class II National Key Protected Species in China [6] and is considered endangered (EN) by the International Union for Conservation of Nature [7]. Currently, most studies on M. shiluensis have focused on its use in landscape gardening and its protection in China [5]; however, there remains a lack of evolutionary and phylogenetic research.
The chloroplast is an important organelle in plants with its own genome (hereafter, cp genome) and participates in photosynthesis and other functions [8]. The cp genome of most land plants has a circular structure, including four segments: A large single-copy (LSC), a small single-copy (SSC), and two invert repeats (IRs) [9]. Although the cp genome is generally conserved, it has undergone intra- and inter-species rearrangement during evolution [10,11], including IR expansion and contraction. The information obtained from sequence rearrangements can be applied in phylogenetic analyses to solve taxonomic problems, such as low-level classifications, using genome comparison [12,13,14,15,16,17]. In the section Michelia, complete cp genomes have been reported for only Magnolia alba (NC037005), Magnolia laevifolia (NC035956), and Michelia odora (NC023239). Therefore, analysis of the cp genomes of other Michelia plants is necessary because of the similarity of morphology among Magnoliaceae species [18].
In the present study, we characterized the cp genome of M. shiluensis and compared its sequence features with four closely related species (M. odora, M. laevifolia, Magnolia insignis, and Magnolia cathcartii). The phylogenetic relationships among 28 Magnoliaceae species were constructed based on 79 protein-coding gene (PCG) sequences and show that M. shiluensis is most closely related to M. odora.

2. Materials and Methods

2.1. Plant Material and DNA Extraction

Fresh leaves of M. shiluensis were collected in the South China Botanical Garden (113°21′ E, 23°10′ N), China and transported to the laboratory at the South China Agricultural University. Total genomic DNA was isolated from the leaves using the CTAB method [19].

2.2. Genome Sequencing and Annotation

An Illumina shotgun library was established according to the manufacturer’s protocol, and high-throughput sequencing was conducted using the HiSeq X TEN platform (Illumina, San Diego, CA, USA). After filtration using SOAPnuke [20], 4.93 GB of clean data were generated. Filtered reads were assembled de novo using SPAdes (version 3.10.1) [21] by referencing them against the cp genome sequence of M. odora (NC037005.1) using BLAST v2.2.30 (National Center for Biotechnology Information, Bethesda, MD, USA). Gene annotation was performed using GeSeq [22]. The cp genome map was generated using Organellar Genome DRAW (version 1.2) [23]. The annotated sequence was submitted to GenBank (accession number MN418056).

2.3. Sequence and Repeat Analysis

We used the Editseq v7.1.0 [24] software to calculate the guanine-cytosine (GC) content. MEGA v7.0.26 [25] was used to generate the relative synonymous codon usage (RSCU) values based on 79 PCGs. RNA editing sites in PCGs were predicted using the PREP suite [26] with the default settings.
The REPuter [27] online service was used to identify repeats (forward, reverse, complement, and palindromic) in the cp genome with default parameters. Chloroplast simple sequence repeats (cpSSRs) were identified using MISA-web [28] with minimal repeat numbers of 8, 5, 4, 3, 3, and 3 for mono-, di-, tri-, tetra-, penta-, and hexanucleotide repeats, respectively.

2.4. Genome Comparison and Sequence Divergence

Comparisons between five Magnoliaceae cp genomes were visualized using online mVISTA software [29] with the annotation of M. shiluensis as the reference in Shuffle-LAGAN mode. The borders of four different regions among the five cp genomes of Magnoliaceae were visualized using IRscope [30]. The nucleotide diversity (Pi), the rate of nonsynonymous (dN) substitutions, the rate of synonymous (dS) substitutions were determined using DNAsp v6.12.03 [31] to investigate the nucleotide diversity of sequences and genes that are considered to be under selection pressure.

2.5. Phylogenetic Analysis

To research the phylogenetic relationships and allow for comparisons among Magnoliaceae species, a maximum likelihood tree was constructed using RAxML [32], with 1000 bootstrap replicates, based on the PCG sequences found in 28 Magnoliaceae cp genomes. All 28 Magnoliaceae cp genome sequences were downloaded from the NCBI nucleotide database.

3. Result

3.1. Structures and Features of M. shiluensis Chloroplast Genome

The complete cp genome of M. shiluensis was 160,075 bp in length and comprised two IR regions of 26,587 bp each, separated by an LSC region of 88,105 bp, and an SSC region of 18,796 bp. The cp genome had the following base proportions: Adenine (A), 29.99%; thymine (T), 30.75%; cytosine (C), 19.98%; and guanine (G), 19.28%. Therefore, the total GC content was 39.26%. The GC content of the LSC, SSC, and IR regions were 37.95%, 34.28%, and 43.20%, respectively (Table 1).
The cp genome of M. shiluensis contained 113 unique genes, including 79 PCGs, 30 tRNAs, and four rRNAs (Table 1, Figure 1). A total of 58 genes were found to be involved in self-replication, 12 genes encoded small ribosomal subunit proteins, eight genes encoded large ribosomal subunit proteins, 30 genes encoded tRNA, and four genes encoded RNA polymerase subunits. A total of 44 genes were found to be involved in photosynthesis, including six genes for ATP synthase, 11 genes for NADH dehydrogenase, six genes for the cytochrome b/f complex, five genes for photosystem I, 15 genes for photosystem II, and one gene for the large chain of Rubisco. In total, 18 genes were duplicated in the cp genome of M. shiluensis, including seven PCGs, seven tRNA genes, and four rRNA genes, all of which were located in the IR region (Table 2). None of the genes contained stop codons in coding sequences, therefore, no pseudogenes were detected.
A total of 16 genes were found to have introns, including 10 PCGs and six tRNA genes. Of these genes, clpP, trnA-UGC, trnI-GAU, and ycf3 had two introns, whereas atpF, ndhA, ndhB, petB, rpl2, rpoC1, rps12, rps16, trnG-UCC, trnK-UUU, trnL-UAA, and trnV-UAC had one intron. The rps12 gene which encodes the 40S ribosomal protein S12, was trans-spliced, with one exon located in the LSC region, and the other two exons located in the IR region. The largest intron was located in the trnK gene (2490 bp) with the matK gene inside; trnL-UAA had the smallest intron (491 bp) (Table 3).
We compared the basic cp genome features of M. shiluensis with four Magnoliaceae species. The cp genome lengths of M. laevifolia and M. insignis were 45 and 42 bp longer than that of M. shiluensis, respectively, while the cp genome lengths of M. odora and M. cathcartii were five and 125 bp shorter, respectively. Compared with M. shiluensis, the variation in the lengths of the LSC, SSC, and IR regions ranged from 7 to 90, 3 to 14, and 1 to 78 bp, respectively. In addition, the GC content of the whole genome and of each region of M. shiluensis were highly similar to those of the other four species. Moreover, there was no variation with respect to the total number of genes, PCGs, tRNA genes, rRNA genes, and genes with introns (Table 1).

3.2. Codon Usage and RNA Analysis

Based on the PCGs, 22,791 codons were detected (excluding the stop codons) (Table 4). The three most abundant amino acids were leucine (2423 codons), isoleucine (2085 codons), and serine (1719 codons), and the three least abundant amino acids were cysteine (314 codons), tryptophan (427 codons), and methionine (602 codons) (Figure S1). Of the 30 most frequent codons (RSCU > 1), most of them end with A or U, and only the UUG codon ends with G. In contrast, most of the 32 least frequent codons (RSCU < 1) end with C or G. In addition, two codons, AUG and UGG, have no codon bias (RSCU = 1).
PREP suite was used to edit predictions in the genome of M. shiluensis by manipulating the first codon position of the first nucleotide (Table S1). A total of 106 RNA editing sites were detected from the PCGs in M. shiluensis; with the majority of the amino acid conversions involving the conversion of serine to leucine. Most of the RNA editing sites were located on the ndhB gene (14 sites), followed by ndhD (11 sites), and matK (nine sites). Most of the conversions changed from a polar group to a nonpolar group, while only two sites changed from a nonpolar group to a polar group (proline to serine); one of these was located on the psbE gene while the other was located on the ccsA gene.

3.3. Repeat Sequence Analysis

The REPuter results show that the M. shiluensis cp genome contains a total of 49 repeats: 23 palindromic, 18 forward, and eight reverse repeats (Figure 2). The repeat size ranged from 18 to 33 bp. The most abundant repeats were 18 bp (12 sites) followed by 20 bp (10 sites) (Figure S2).
In the first location, 46.9% of repeats were detected in the intergenic spacers (IGSs), while 34.7% were in the PCGs, and 18.4% were in the tRNA genes. Of all the PCGs, the ycf2 gene had five forward repeats and four palindromic repeats and was the gene with the most repeats (Table S2). Comparison of the repeat types with the other four species revealed no substantial variation among the five species (Figure 2). Michelia shiluensis had the highest frequency of palindromic repeats (23), while M. laevifolia had the lowest (21). Magnolia cathcartii, M. odora, and M. shiluensis had the same number of forward repeats (18), while M. shiluensis and M. cathcartii had eight reverse repeats. In addition, only one complement repeat was found in the genomes of M. laevifolia and M. odora, whereas no complements were identified in the cp genomes of M. shiluensis and M. insignis. A total of 141 cpSSRs were found in the cp genome of M. shiluensis (Table S3). The majority of them were mononucleotide repeats (118), followed by tetranucleotide repeats (9), and dinucleotide repeats (8) (Figure 3). No pentanucleotide repeats were detected in the cp genome of M. shiluensis. The longest repeat was 18 bp while the shortest was 8 bp. Noncoding regions, including IGSs (97) and introns (19), contained most of the SSRs while only 25 repeats were located in coding regions, including cemA, ndhD, ndhF, psbC, rpoB, rpoC1, rpoC2, rps19, rps3, ycf1, ycf2, and ycf4 (Table 5). The cpSSRs were mainly distributed in the LSC region (72.34%), followed by the SSC region (17.73%), with just 4.96% in the IR. The cpSSRs in M. shiluensis had base bias towards A-T bases. In total, 113 SSRs had A or T bases, accounting for 80.14% of the total SSRs (Figure 3). Comparison among the five species of Magnoliaceae show high similarity in SSR type and distribution. The variation in the total amount, mono-, di-, tri-, tetra-, penta-, and hexanucleotide repeats among the five species was 5, 4, 0, 2, 0, 1, and 1, respectively (Figure 3). The number of SSRs in the IR were the same among the five species while the counts in different locations and regions were highly conserved.

3.4. Genome Comparison and Sequence Divergence

The mVISTA online software was used to compare the variation in the whole cp genome among the five species (Figure 4). The alignments indicated that the whole cp genome of the five species was highly conserved, especially in the IR region. The noncoding sequences had relatively more divergence than the coding sequences. The noncoding sequences that contained high levels of divergence were rps16-trnQ, atpH-atpI, trnT-psbD, petA-psbJ, and ndhF-trnL. In the coding sequences, only ycf1 show relatively more variation than the other genes. No obvious insertions were found among the five species.
The four junctions in the regions of the cp genomes of the five species were shown using IRscope (Figure 5). There was a conserved structure on each border, and slight distance differences among the five species. Gene rps19 was fully located in the LSC at a distance of 1–6 bp from the LSC/IRb border, while gene rpl2 was fully located in the IRb. The ndhF gene was found in the SSC region and was 61 bp away from the IRb/SSC border in M. odora, M. laevifolia, and M. insignis, while it was 21 bp longer in M. shiluensis, and 7 bp shorter in M. cathcartii. The SSC/IRa border was inside the ycf1 gene in all five species. Compared to M. shiluensis and M. odora, the part of the ycf1 gene in the SSC region of M. laevifolia and M. insignis was almost 20 bp longer, and this resulted in the differences in gene length. However, the ycf1 gene of M. cathcartii was almost 30 bp shorter on both sides of the SSC/IRa border; thus, the ycf1 gene of M. cathcartii was almost 60 bp shorter than those of the other four species. The distance from the trnH gene to the IRa/LSC border was 11 bp in M. shiluensis, M. odora, and M. laevifolia, while it was 16 bp in M. insignis and 9 bp in M. cathcartii. Due to the short length in the IR region, the whole length of the M. cathcartii cp genome was significantly shorter than those of the other four species.
To detect the selective pressures on the PCGs in the M. shiluensis cp genome, the rate of nonsynonymous (dN) substitutions, the rate of synonymous (dS) substitutions, and their ratio (dN/dS) were calculated based on the 79 PCG sequences of the five Magnoliaceae species (Figure 6). Only four genes had a dN/dS ratio greater than 1 (accD in M. insignis vs. M. cathcartii, score 1.14; ndhD in M. shiluensis vs. M. cathcartii, score 1.29; ndhF in M. odora vs. M. cathcartii, score 1.89; and rpoC2 in M. laevifolia vs. M. cathcartii, score 2.50), which indicates that most genes are under the influence of negative selection, while only a few genes are under the influence of positive selection.
In the coding region, the mean Pi in the PCGs was 0.00117 (ranging from 0 to 0.00606); and the mean values of Pi in the LSC, IR, and SSC regions were 0.001192, 0.000186, and 0.001634, respectively (Figure 7). Meanwhile, in the IGSs, the mean Pi value was 0.00295 (ranging from 0 to 0.02416); and the mean Pi value in the LSC, IR, and SSC regions were 0.0297, 0.00045, and 0.00731, respectively. This result indicates that the Pi value in the coding region is lower than that in the IGSs. The results also demonstrate that the IR region is the most conserved region among the five species, followed by the LSC and SSC regions. In total, 20 mutation sites (Pi > 0.005) were identified, including 19 sites in IGSs and one site in a PCG. The mutation sites in IGSs were as follows: trnH-psbA, psbK-psbI, atpA-atpF, rps2-rpoC2, trnT-psbD, ycf3-trnS, ndhJ-ndhK, ndhK-ndhC, accD-psaI, psbL-psbF, petL-petG, trnW-trnP, trnP-psaJ, rpl18-rpl20, ndhF-trnL, ccsA-ndhD, ndhD-psaC, ndhG-ndhI, and ndhI-ndhA. One gene, psaJ, was unique and had a Pi value greater than 0.005.

3.5. Phylogenetic Analysis

To reveal the evolutionary relationships between the investigated species and to enable comparison with traditional phylogenies, a maximum likelihood phylogenetic tree was constructed using RAxML (with 1000 bootstrap replicates) based on the PCG sequences found in 28 Magnoliaceae cp genomes (Figure 8). The phylogenetic tree generated 25 nodes; most of which had 100% bootstrap support. This result strongly supports the notion that M. shiluensis is most closely related to M. odora.

4. Discussion

In this study, we characterized the complete cp genome of M. shiluensis, an endangered and valuable species (Figure 1). By comparing five closely related species, we found that gene content, order, structure, and other features were highly conserved among them (Figure 3, Figure 4 and Figure 5). Michelia shiluensis was shown to be most closely related to M. odora (Figure 8). This finding can help to further our understanding of the characterization of the M. shiluensis cp genome and reveal information concerning the evolution, population genetics, and phylogeny of this species.
Normally, the length of the cp genome in higher plants is in the range of about 120–160 kb, with a stable structure and conserved sequence [8,33]. The cp genome of M. shiluensis displayed a typical quadripartite structure, with an LSC and an SSC which were separated by two IR regions (Figure 1). The whole length of this genome was 160,075 bp, with 39.26% GC content, and containing 113 unique genes and 16 genes with one or two introns (Table 1, Table 2 and Table 3). Among the 26 Magnoliaceae species, the length of the cp genome ranged from 158,177 to 160,183 bp, the GC content ranged from 39.15% to 39.30%, and they collectively contained 112 common genes, including 79 PCGs, 29 tRNA, and four rRNA genes; also, one or two introns were found among these 16 genes. The results for the M. shiluensis cp genome were consistent with a previous analysis of 26 Magnoliaceae species [34], except for the number of genes, one additional tRNA gene (trnV-GAC) was detected in M. shiluensis. Similar to other angiosperms, a high GC content was detected in the IR region of M. shiluensis, which may be a result of the existence of high-GC rRNA sequences [9,35,36,37]. Introns play a vital role in selective gene splicing [38]. However, introns have been lost among some species during their evolution [39,40]. In this study, no introns were lost in the cp genome of M. shiluensis during evolution, which reflects the fact that the cp genome is highly conserved in Magnoliaceae [34].
In total, 22,791 codons were found in the cp genome of M. shiluensis (Table 4), among which, the codons for leucine were the most abundant (10.63%). This result has also been observed in Ailanthus altissima [41] and Justicia flava [42]. Among the preferred codons (RSCU > 1), we found that most of them ended with A or U, except UUG. This is not unique to the M. shiluensis cp genome as similar findings have been observed in Papaver rhoeas and P. orientale [36], Ageratina adenophora [43], and Oryza sativa [44]. Of the PCGs in the M. shiluensis cp genome, 106 possible sites for RNA editing were detected (Table S1). The majority of the amino acid conversion was from serine to leucine, and the ndhB gene accounted for a high number of editable sites (14 of the total 106 sites). Similar results have been obtained for Forsythia suspensa [45] and Sanionia uncinata [46].
Repeat sequences play an important role in genomic structural variation, expansion, and rearrangement [8,40]. Previous research has indicated that most of the repeat sequences are located in the IGS regions followed by the coding regions [47,48]. A similar result was found in this study, with 46.9% of repeats detected in the IGS regions, followed by 34.7% in the coding regions, and the remainder in the tRNAs (Table S2). The cpSSR is an effective marker [49,50] that is widely used in population genetics, biogeographic studies, and phylogenetic evaluation [51,52]. In the cp genome of M. shiluensis, over 80% of the SSRs consisted of A or T bases, and over 80% were mononucleotide repeats. Similar results have been observed in other studies [48,50,53]. The majority of SSRs are found in the SSC and LSC regions [50] and, in this respect, M. shiluensis is no exception (Table 5). These two regions accounted for 90.07% of the SSRs, and only seven SSRs were found in the IR region.
Although the cp genome of angiosperms is relatively conserved in structure and size [54,55], the expansion and contraction of the IR region, as caused by evolutionary events, has resulted in minor changes in the IR boundary and size of the genome [39,56], thus increasing the chloroplast genetic diversity of angiosperms [57,58]. In this study, comparative analysis of five Magnoliaceae species revealed that the IR lengths were similar in M. shiluensis, M. odora, and M. laevifolia (Figure 5). However, the IR region of M. insignis had completely lost 11 bp in the rps12-trnV, rrn23-rrn4.5 IGSs, while M. cathcartii had lost 5 bp in the rpl2 intron, 6 bp in the rps12 intron, 26 bp in ycf1, and 41 bp in the trnN-ndhF IGS. The losses of these bases resulted in differences in the lengths of the IR regions among the five species.
DNA barcoding is a technique that is widely applied in plant identification studies [59,60]. However, only a few regions have been used for the DNA barcoding of Magnoliaceae [61,62,63]. We used mVISTA to compare the genomes of five Magnoliaceae species and revealed that the IR region is more highly conserved than the LSC and SSC regions, and that the coding region was more highly conserved than the noncoding region (Figure 4), consistent with other angiosperms [9,38]. Five regions in the M. shiluensis cp genome had high levels of variation (four on IGSs and one on a PCG). The Pi value was also investigated among the 79 PCGs and 125 IGSs (Figure 7), and only 20 regions were found to have a Pi value greater than 0.005; which confirmed the low base substitution rate in Magnoliaceae [64]. Regions with a high degree of variation can be used to develop high resolution DNA barcoding for identification.
Due to the high morphological similarity among Magnoliaceae species [18,65], there have been some difficulties with respect to the classification of the family. Thus, the classification of Magnoliaceae has always been controversial [66,67,68,69,70,71,72]. The cp genome contains sufficient information and has been shown to be more effective than cpDNA fragments for clarifying low level phylogenetic relationships in plants [53,73]. In this study, the phylogenetic results of 28 Magnoliaceae plants based on PCG sequences revealed that M. shiluensis is most closely related to M. odora (Figure 8), which is consistent with phylogenetic results based on the ndhF sequence [70]. According to traditional morphological classification, M. insignis has been placed in the subgenera Manglietia, and M. alba has been placed in the section Michelia [69,74]. However, the phylogenetic relationship results based on the cp genome show that M. insignis is located in the section Michelia clade and M. alba is located in the subgenera Yulania clade. This result differs from that of traditional morphological classification [69,74] and the results of three nuclear gene sequences [75]. These findings confirm that not even a complete cp genome can distinguish species in young evolutionary lineages, and that phylogenetic conclusions may require consideration of certain features in the nuclear genome [76].

5. Conclusions

The complete cp genome provided by this study can be used for in-depth genetic research on M. shiluensis and Magnoliaceae species in general, and may also play an important role in the development of new conservation and management strategies to ultimately aid species conservation efforts.

Supplementary Materials

Supplementary materials can be found at https://www.mdpi.com/1999-4907/11/3/267/s1. Figure S1, Amino acid frequency among 79 protein-coding genes in the Michelia shiluensis chloroplast genome; Table S1, Possible RNA editing sites in the chloroplast genome of Michelia shiluensis; Table S2, Repeats in the chloroplast genome of Michelia shiluensis; Figure S2, Different lengths of repeats in the Michelia shiluensis cp genome; Table S3, Single sequence repeats in the chloroplast genome of Michelia shiluensis.

Author Contributions

Conceptualization, Y.D.; methodology, Y.D. and Y.H.; software, Y.L. and C.L.; validation, Y.D., Y.L., Y.H., C.L., and X.D.; formal analysis, Y.L. and Y.H.; resources, X.Q. and X.D.; writing—original draft preparation, Y.D.; writing—review and editing, Y.D.; supervision, X.D.; project administration, X.D.; funding acquisition, X.D. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Forestry Public Welfare Industry Research of China (Grant Number: 201404116), the National Natural Science Foundation of China (Grant Number: 31670601), and the Forestry Science and Technology Innovation of Guangdong Province grant programs (Grant Numbers: 2014KJCX006 and 2017KJCX023).

Acknowledgments

We are sincerely grateful to Xiaomei Deng (supervisor to Yanwen Deng) for her support during this study, which included assistance with funding, materials, resources, and consultations with other academics. We are also grateful to Xinsheng Hu for his invaluable guidance during the manuscript preparation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Francisco-Ortega, J.; Wang, F.-G.; Wang, Z.-S.; Xing, F.-W.; Liu, H.; Xu, H.; Xu, W.-X.; Luo, Y.-B.; Song, X.-Q.; Gale, S. Endemic seed plant species from Hainan Island: A checklist. Bot. Rev. 2010, 76, 295–345. [Google Scholar] [CrossRef]
  2. Bao, S.; Zhou, S.; Yu, Y. Comparative anatomy of the leaves in Michelia. Guangxi Zhi Wu 2002, 22, 140–144. [Google Scholar]
  3. Xiong, J.; Wang, L.-J.; Qian, J.; Wang, P.-P.; Wang, X.-J.; Ma, G.-L.; Zeng, H.; Li, J.; Hu, J.-F. Structurally diverse sesquiterpenoids from the endangered ornamental plant Michelia shiluensis. J. Nat. Prod. 2018, 81, 2195–2204. [Google Scholar] [CrossRef]
  4. Chen, B. Forest wild biospecies diversity and its preservation in China. Chin. J. Ecol. 1993, 12, 39–43. [Google Scholar] [CrossRef]
  5. Wei, Y.; Hong, F.; Yuan, L.; Kong, Y.; Shi, Y. Population distribution and age structure characteristics of Michelia shiluensis, an endangered and endemic species in Hainan Island. Chin. J. Trop. Crops 2017, 38, 2280–2284. [Google Scholar] [CrossRef]
  6. State Forestry Administration of China. List of national key protected wild plants (first batch). State Counc. Bull. 1999, 13, 6. [Google Scholar]
  7. Bhandari, M. International Union for Conservation of Nature. In The Wiley-Blackwell Encyclopedia of Globalization; Ritzer, G., Ed.; John Wiley & Sons: Hoboken, NJ, USA, 2012. [Google Scholar] [CrossRef]
  8. Wicke, S.; Schneeweiss, G.M.; dePamphilis, C.W.; Müller, K.F.; Quandt, D. The evolution of the plastid chromosome in land plants: Gene content, gene order, gene function. Plant Mol. Biol. 2011, 76, 273–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Chen, Y.; Hu, N.; Wu, H. Analyzing and characterizing the chloroplast genome of Salix wilsonii. BioMed Res. Int. 2019, 2019, 5190425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Tangphatsornruang, S.; Uthaipaisanwong, P.; Sangsrakru, D.; Chanprasert, J.; Yoocha, T.; Jomchai, N.; Tragoonrung, S. Characterization of the complete chloroplast genome of Hevea brasiliensis reveals genome rearrangement, RNA editing sites and phylogenetic relationships. Gene 2011, 475, 104–112. [Google Scholar] [CrossRef] [PubMed]
  11. Walker, J.F.; Jansen, R.K.; Zanis, M.J.; Emery, N.C. Sources of inversion variation in the small single copy (SSC) region of chloroplast genomes. Am. J. Bot. 2015, 102, 1751–1752. [Google Scholar] [CrossRef] [Green Version]
  12. Johansson, J.T. There large inversions in the chloroplast genomes and one loss of the chloroplast generps16 suggest an early evolutionary split in the genusAdonis (Ranunculaceae). Plant Syst. Evol. 1999, 218, 133–143. [Google Scholar] [CrossRef]
  13. 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] [PubMed] [Green Version]
  14. 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. Phylogenet. Evol. 2008, 48, 1204–1217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Shaw, J.; Lickey, E.B.; Schilling, E.E.; Small, R.L. Comparison of whole chloroplast genome sequences to choose noncoding regions for phylogenetic studies in angiosperms: The tortoise and the hare III. Am. J. Bot. 2007, 94, 275–288. [Google Scholar] [CrossRef] [Green Version]
  16. Mardanov, A.V.; Ravin, N.V.; Kuznetsov, B.B.; Samigullin, T.H.; Antonov, A.S.; Kolganova, T.V.; Skyabin, K.G. Complete sequence of the duckweed (Lemna minor) chloroplast genome: Structural organization and phylogenetic relationships to other angiosperms. J. Mol. Evol. 2008, 66, 555–564. [Google Scholar] [CrossRef]
  17. Park, I.; Kim, W.-J.; Yang, S.; Yeo, S.-M.; Li, H.; Moon, B.C. The complete chloroplast genome sequence of Aconitum coreanum and Aconitum carmichaelii and comparative analysis with other Aconitum species. PLoS ONE 2017, 12, e0184257. [Google Scholar] [CrossRef] [Green Version]
  18. Li, J.; Conran, J. Phylogenetic relationships in Magnoliaceae subfam. Magnolioideae: A morphological cladistic analysis. Plant Syst. Evol. 2003, 242, 33–47. [Google Scholar] [CrossRef]
  19. Doyle, J.J.; Doyle, J.L. A rapid DNA isolation procedure for small quantities of fresh leaf tissue. Phytochem. Bull. 1987, 19, 11–15. [Google Scholar]
  20. Chen, Y.; Shi, C.; Huang, Z.; Zhang, Y.; Li, S.; Li, Y.; Ye, J.; Yu, C.; Li, Z. SOAPnuke: A MapReduce acceleration-supported software for integrated quality control and preprocessing of high-throughput sequencing data. Gigascience 2017, 7, gix120. [Google Scholar] [CrossRef] [Green Version]
  21. Bankevich, A.; Nurk, S.; Antipov, D.; Gurevich, A.A.; Dvorkin, M.; Kulikov, A.S.; Lesin, V.M.; Nikolenko, S.I.; Pham, S.; Prjibelski, A.D.; et al. SPAdes: A new genome assembly algorithm and its applications to single-cell sequencing. J. Comput. Biol. 2012, 19, 455–477. [Google Scholar] [CrossRef] [Green Version]
  22. Tillich, M.; Lehwark, P.; Pellizzer, T.; Ulbricht-Jones, E.S.; Fischer, A.; Bock, R.; Greiner, S. GeSeq-versatile and accurate annotation of organelle genomes. Nucleic Acids Res. 2017, 45, W6–W11. [Google Scholar] [CrossRef] [PubMed]
  23. Lohse, M.; Drechsel, O.; Kahlau, S.; Bock, R. OrganellarGenomeDRAW—A suite of tools for generating physical maps of plastid and mitochondrial genomes and visualizing expression data sets. Nucleic Acids Res. 2013, 41, W575–W581. [Google Scholar] [CrossRef] [PubMed]
  24. Burland, T.G. DNASTAR’s Lasergene sequence analysis software. In Bioinformatics Methods and Protocols; Misener, S., Krawetz, A.K., Eds.; Springer: Basel, Switzerland, 2000; pp. 71–91. [Google Scholar]
  25. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Mower, J.P. The PREP suite: Predictive RNA editors for plant mitochondrial genes, chloroplast genes and user-defined alignments. Nucleic Acids Res. 2009, 37, W253–W259. [Google Scholar] [CrossRef]
  27. Kurtz, S.; Choudhuri, J.V.; Ohlebusch, E.; Schleiermacher, C.; Stoye, J.; Giegerich, R. REPuter: The manifold applications of repeat analysis on a genomic scale. Nucleic Acids Res. 2001, 29, 4633–4642. [Google Scholar] [CrossRef] [Green Version]
  28. 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]
  29. Frazer, K.A.; Pachter, L.; Poliakov, A.; Rubin, E.M.; Dubchak, I. VISTA: Computational tools for comparative genomics. Nucleic Acids Res. 2004, 32, W273–W279. [Google Scholar] [CrossRef]
  30. Amiryousefi, A.; Hyvönen, J.; Poczai, P. IRscope: An online program to visualize the junction sites of chloroplast genomes. Bioinformatics 2018, 34, 3030–3031. [Google Scholar] [CrossRef]
  31. Rozas, J.; Ferrer-Mata, A.; Sánchez-DelBarrio, J.C.; Guirao-Rico, S.; Librado, P.; Ramos-Onsins, S.E.; Sánchez-Gracia, A. DnaSP 6: DNA sequence polymorphism analysis of large data sets. Mol. Biol. Evol. 2017, 34, 3299–3302. [Google Scholar] [CrossRef]
  32. Stamatakis, A. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 2014, 30, 1312–1313. [Google Scholar] [CrossRef]
  33. Palmer, J.D. Comparative organization of chloroplast genomes. Annu. Rev. Genet. 1985, 19, 325–354. [Google Scholar] [CrossRef] [PubMed]
  34. Shen, Y.; Chen, K.; Gu, C.; Zheng, S.; Ma, L. Comparative and phylogenetic analyses of 26 Magnoliaceae species based on complete chloroplast genome sequences. Can. J. For. Res. 2018, 48, 1456–1469. [Google Scholar] [CrossRef]
  35. Asaf, S.; Khan, A.L.; Khan, A.R.; Waqas, M.; Kang, S.-M.; Khan, M.A.; Lee, S.-M.; Lee, I.-J. Complete chloroplast genome of Nicotiana otophora and its comparison with related species. Front. Plant Sci. 2016, 7, 843. [Google Scholar] [CrossRef] [Green Version]
  36. 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] [PubMed] [Green Version]
  37. Meng, J.; Li, X.; Li, H.; Yang, J.; Wang, H.; He, J. Comparative analysis of the complete chloroplast genomes of four Aconitum Medicinal Species. Molecules 2018, 23, 1015. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Zhong, Q.; Yang, S.; Sun, X.; Wang, L.; Li, Y. The complete chloroplast genome of the Jerusalem artichoke (Helianthus tuberosus L.) and an adaptive evolutionary analysis of the ycf2 gene. PeerJ 2019, 7, e7596. [Google Scholar] [CrossRef] [Green Version]
  39. He, L.; Qian, J.; Li, X.; Sun, Z.; Xu, X.; Chen, S. Complete chloroplast genome of medicinal plant Lonicera japonica: Genome rearrangement, intron gain and loss, and implications for phylogenetic studies. Molecules 2017, 22, 249. [Google Scholar] [CrossRef]
  40. Maréchal, A.; Brisson, N. Recombination and the maintenance of plant organelle genome stability. New Phytol. 2010, 186, 299–317. [Google Scholar] [CrossRef]
  41. Saina, J.; Li, Z.-Z.; Gichira, A.; Liao, Y.-Y. The complete chloroplast genome sequence of tree of heaven (Ailanthus altissima (Mill.) (Sapindales: Simaroubaceae), an important pantropical tree. Int. J. Mol. Sci. 2018, 19, 929. [Google Scholar] [CrossRef] [Green Version]
  42. Yaradua, S.S.; Alzahrani, D.A.; Albokhary, E.J.; Abba, A.; Bello, A. Complete chloroplast genome sequence of Justicia flava: Genome comparative analysis and phylogenetic relationships among Acanthaceae. BioMed Res. Int. 2019, 2019, 4370258. [Google Scholar] [CrossRef] [Green Version]
  43. 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]
  44. Liu, Q.; Xue, Q. Comparative studies on codon usage pattern of chloroplasts and their host nuclear genes in four plant species. J. Genet. 2005, 84, 55–62. [Google Scholar] [CrossRef] [PubMed]
  45. Wang, W.; Yu, H.; Wang, J.; Lei, W.; Gao, J.; Qiu, X.; Wang, J. The complete chloroplast genome sequences of the medicinal plant Forsythia suspensa (Oleaceae). Int. J. Mol. Sci. 2017, 18, 2288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. 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]
  47. Yang, Y.; Zhou, T.; Duan, D.; Yang, J.; Feng, L.; Zhao, G. Comparative analysis of the complete chloroplast genomes of five Quercus species. Front. Plant Sci. 2016, 7, 959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Zhou, T.; Chen, C.; Wei, Y.; Chang, Y.; Bai, G.; Li, Z.; Kanwal, N.; Zhao, G. Comparative transcriptome and chloroplast genome analyses of two related Dipteronia species. Front. Plant Sci. 2016, 7, 1512. [Google Scholar] [CrossRef] [Green Version]
  49. Provan, J.; Powell, W.; Hollingsworth, P.M. Chloroplast microsatellites: New tools for studies in plant ecology and evolution. Trends Ecol. Evol. 2001, 16, 142–147. [Google Scholar] [CrossRef]
  50. 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. Res. 2009, 9, 673–690. [Google Scholar] [CrossRef]
  51. Mohammad-Panah, N.; Shabanian, N.; Khadivi, A.; Rahmani, M.-S.; Emami, A. Genetic structure of gall oak (Quercus infectoria) characterized by nuclear and chloroplast SSR markers. Tree Genet. Genomes 2017, 13, 70. [Google Scholar] [CrossRef]
  52. Zeng, J.; Chen, X.; Wu, X.F.; Jiao, F.C.; Xiao, B.G.; Li, Y.P.; Tong, Z.J. Genetic diversity analysis of genus Nicotiana based on SSR markers in chloroplast genome and mitochondria genome. Acta Tabacaria Sin. 2016, 22, 89–97. [Google Scholar] [CrossRef]
  53. 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, 1148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Liu, L.-X.; Li, R.; Worth, J.R.; Li, X.; Li, P.; Cameron, K.M.; Fu, C.-X. The complete chloroplast genome of Chinese bayberry (Morella rubra, Myricaceae): Implications for understanding the evolution of Fagales. Front. Plant Sci. 2017, 8, 968. [Google Scholar] [CrossRef] [PubMed]
  55. Lu, R.-S.; Li, P.; Qiu, Y.-X. The complete chloroplast genomes of three Cardiocrinum (Liliaceae) species: Comparative genomic and phylogenetic analyses. Front. Plant Sci. 2017, 7, 2054. [Google Scholar] [CrossRef] [PubMed]
  56. Raubeson, L.A.; Peery, R.; Chumley, T.W.; Dziubek, C.; Fourcade, H.M.; Boore, J.L.; Jansen, R.K. Comparative chloroplast genomics: Analyses including new sequences from the angiosperms Nuphar advena and Ranunculus macranthus. BMC Genom. 2007, 8, 174. [Google Scholar] [CrossRef] [Green Version]
  57. 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]
  58. Huang, H.; Shi, C.; Liu, Y.; Mao, S.-Y.; Gao, L.-Z. Thirteen Camellia chloroplast genome sequences determined by high-throughput sequencing: Genome structure and phylogenetic relationships. BMC Evol. Biol. 2014, 14, 151. [Google Scholar] [CrossRef] [Green Version]
  59. Zhang, W.; Fan, X.; Zhu, S.; Zhao, H.; Fu, L. Species-specific identification from incomplete sampling: Applying DNA barcodes to monitoring invasive Solanum plants. PLoS ONE 2013, 8, e55927. [Google Scholar] [CrossRef]
  60. Dick, C.W.; Webb, C.O. Plant DNA barcodes, taxonomic management, and species discovery in tropical forests. In DNA Barcodes. Methods in Molecular Biology (Methods and Protocols); Kress, W., Webb, C.O., Eds.; Humana Press: Totowa, NJ, USA, 2012; Volume 858, pp. 379–393. [Google Scholar]
  61. Yu, H.; Wu, K.; Song, J.; Zhu, Y.; Yao, H.; Luo, K.; Dai, Y.; Xu, S.; Lin, Y. Expedient identification of Magnoliaceae species by DNA barcoding. Plant Omics 2014, 7, 47. [Google Scholar]
  62. Huan, H.V.; Trang, H.M.; Toan, N.V. Identification of DNA barcode sequence and genetic relationship among some species of Magnolia Family. Asian J. Plant Sci. 2018, 17, 56–64. [Google Scholar] [CrossRef]
  63. Huan, H.V.; Nguyen, L.T.T.; Quang, N.M. To create DNA barcode data of Magnolia chevalieri (Dandy) V.S. Kumar for identification species and researching genetic diversity. J. For. Sci. Technol. 2019, 18, 3–9. [Google Scholar]
  64. Azuma, H.; Thien, L.B.; Kawano, S. Molecular phylogeny of Magnolia (Magnoliaceae) inferred from cpDNA sequences and evolutionary divergence of the floral scents. J. Plant Res. 1999, 112, 291–306. [Google Scholar] [CrossRef]
  65. Bentley, R.A. Manual of Botany: Including the Structure, Classification, Properties, Uses, and Functions of Plants; J. & A. Churchill: London, UK, 1882. [Google Scholar]
  66. Azuma, H.; García-Franco, J.G.; Rico-Gray, V.; Thien, L.B. Molecular phylogeny of the Magnoliaceae: The biogeography of tropical and temperate disjunctions. Am. J. Bot. 2001, 88, 2275–2285. [Google Scholar] [CrossRef] [PubMed]
  67. Dandy, J.E. A revised survey of the genus Magnolia together with Manglietia and Michelia. In Magnolias; Treseder, N.G., Ed.; Faber & Faber: London, UK, 1978; pp. 27–37. [Google Scholar]
  68. Liang, C.B.; Nooteboom, H.P. Notes on Magnoliaceae III: The Magnoliaceae of China. Ann. Mo. Bot. Gard. 1993, 80, 999–1104. [Google Scholar] [CrossRef]
  69. Figlar, R.B.; Nooteboom, H.P. Notes on Magnoliaceae IV. Blumea 2004, 49, 87–100. [Google Scholar] [CrossRef]
  70. Kim, S.; Park, C.W.; Kim, Y.D.; Suh, Y. Phylogenetic relationships in family Magnoliaceae inferred from ndhF sequences. Am. J. Bot. 2001, 88, 717–728. [Google Scholar] [CrossRef] [PubMed]
  71. Kim, S.; Nooteboom, H.P.; Park, C.-W.; Suh, Y. Taxonomic revision of Magnolia section Maingola (Magnoliaceae). Blumea 2002, 47, 319–339. [Google Scholar]
  72. Nooteboom, H. Different looks at the classification of the Magnoliaceae. In Proceedings of the International Symposium on the Family Magnoliaceae, Guangzhou, China, 18–22 May 1998; Science Press: Beijing, China, 2000; pp. 26–37. [Google Scholar]
  73. Zhao, M.-L.; Song, Y.; Ni, J.; Yao, X.; Tan, Y.-H.; Xu, Z.-F. Comparative chloroplast genomics and phylogenetics of nine Lindera species (Lauraceae). Sci. Rep. 2018, 8, 8844. [Google Scholar] [CrossRef]
  74. Noteboom, H. Notes on Magnoliaceae with a revision of Pachylarnax and Elmerrillia and the Malasian species of Manglietia and Michelia. Blumea 1985, 31, 65–121. [Google Scholar]
  75. Nie, Z.-L.; Wen, J.; Azuma, H.; Qiu, Y.-L.; Sun, H.; Meng, Y.; Sun, W.-B.; Zimmer, E.A. Phylogenetic and biogeographic complexity of Magnoliaceae in the Northern Hemisphere inferred from three nuclear data sets. Mol. Phylogenet. Evol. 2008, 48, 1027–1040. [Google Scholar] [CrossRef]
  76. Ruhsam, M.; Rai, H.S.; Mathews, S.; Ross, T.G.; Graham, S.W.; Raubeson, L.A.; Mei, W.; Thomas, P.I.; Gardner, M.F.; Ennos, R.A. Does complete plastid genome sequencing improve species discrimination and phylogenetic resolution in Araucaria? Mol. Ecol. Resour. 2015, 15, 1067–1078. [Google Scholar] [CrossRef]
Figure 1. Gene map of the Michelia shiluensis chloroplast genome. Genes on the outside of the circle are transcribed counter-clockwise, while genes on the inside are transcribed clockwise. Different colors represent different kinds of functional genes. The guanine-cytosine content is indicated by darker gray and the adenine-thymine content is indicated by light gray.
Figure 1. Gene map of the Michelia shiluensis chloroplast genome. Genes on the outside of the circle are transcribed counter-clockwise, while genes on the inside are transcribed clockwise. Different colors represent different kinds of functional genes. The guanine-cytosine content is indicated by darker gray and the adenine-thymine content is indicated by light gray.
Forests 11 00267 g001
Figure 2. Comparison of repeats among five Magnoliaceae species: Michelia shiluensis, Michelia odora, Magnolia laevifolia, Magnolia insignis, and Magnolia cathcartii. (F: Forward; R: Reverse; C: Complement; and P: Palindromic).
Figure 2. Comparison of repeats among five Magnoliaceae species: Michelia shiluensis, Michelia odora, Magnolia laevifolia, Magnolia insignis, and Magnolia cathcartii. (F: Forward; R: Reverse; C: Complement; and P: Palindromic).
Forests 11 00267 g002
Figure 3. Single sequence repeats (SSRs) in the chloroplast genome of Michelia shiluensis. (a) Comparison of the SSRs among five Magnoliaceae species (M. shiluensis, Michelia odora, Magnolia laevifolia, Magnolia insignis, and Magnolia cathcartii); (b) base composition of SSRs in the cp genome of M. shiluensis.
Figure 3. Single sequence repeats (SSRs) in the chloroplast genome of Michelia shiluensis. (a) Comparison of the SSRs among five Magnoliaceae species (M. shiluensis, Michelia odora, Magnolia laevifolia, Magnolia insignis, and Magnolia cathcartii); (b) base composition of SSRs in the cp genome of M. shiluensis.
Forests 11 00267 g003
Figure 4. Sequence alignment of five whole chloroplast genomes in Magnoliaceae (Michelia shiluensis, Michelia odora, Magnolia laevifolia, Magnolia insignis, and Magnolia cathcartii) using M. shiluensis as a reference in mVISTA.
Figure 4. Sequence alignment of five whole chloroplast genomes in Magnoliaceae (Michelia shiluensis, Michelia odora, Magnolia laevifolia, Magnolia insignis, and Magnolia cathcartii) using M. shiluensis as a reference in mVISTA.
Forests 11 00267 g004
Figure 5. The four junctions of the regions in the chloroplast genomes of the five Magnoliaceae species, determined using IRscope.
Figure 5. The four junctions of the regions in the chloroplast genomes of the five Magnoliaceae species, determined using IRscope.
Forests 11 00267 g005
Figure 6. The synonymous (dS) and nonsynonymous substitutions (dN)/dS ratio values of 79 protein-coding genes from five Magnoliaceae chloroplast genomes (Ms: Michelia shiluensis; Mo: Michelia odora; Ml: Magnolia laevifolia; Mi: Magnolia insignis; Mc: Magnolia cathcartii).
Figure 6. The synonymous (dS) and nonsynonymous substitutions (dN)/dS ratio values of 79 protein-coding genes from five Magnoliaceae chloroplast genomes (Ms: Michelia shiluensis; Mo: Michelia odora; Ml: Magnolia laevifolia; Mi: Magnolia insignis; Mc: Magnolia cathcartii).
Forests 11 00267 g006
Figure 7. Nucleotide diversity (Pi) in the chloroplast genome of five Magnoliaceae species (Michelia shiluensis; Michelia odora; Magnolia laevifolia; Magnolia insignis; and Magnolia cathcartii).
Figure 7. Nucleotide diversity (Pi) in the chloroplast genome of five Magnoliaceae species (Michelia shiluensis; Michelia odora; Magnolia laevifolia; Magnolia insignis; and Magnolia cathcartii).
Forests 11 00267 g007
Figure 8. Maximum likelihood tree with 1000 bootstrap replicates constructed using RAxML based on chloroplast genomes of 28 Magnoliaceae species (26 species of Magnolia and Michelia, and two species of Liriodendron as outgroups). Bootstrap values (%) are shown above branches. Accession numbers: Magnolia alba NC037005, Magnolia liliiflora NC023238, Magnolia denudata NC018357, Magnolia sprengeri NC023242, Magnolia salicifolia NC023240, Magnolia biondii KY085894, Magnolia kobus NC023237, Michelia odora NC023239, Michelia shiluensis MN418056, Magnolia laevifolia NC035956, Magnolia insignis NC035657, Magnolia cathcartii NC023234, Magnolia yunnanensis NC024545, Magnolia sinica NC023241, Magnolia kwangsiensis NC015892, Magnolia conifera NC037001, Magnolia dandyi NC037004, Magnolia aromatica NC037000, Magnolia fordiana MF990562, Magnolia glaucifolia NC037003, Magnolia duclouxii NC037002, Magnolia tripetala NC024027, Magnolia officinalis NC020316, Magnolia grandiflora NC020318, Magnolia pyramidata NC023236, Magnolia dealbata NC023235, Liriodendron chinense NC030504, and Liriodendron tulipifera NC008326.
Figure 8. Maximum likelihood tree with 1000 bootstrap replicates constructed using RAxML based on chloroplast genomes of 28 Magnoliaceae species (26 species of Magnolia and Michelia, and two species of Liriodendron as outgroups). Bootstrap values (%) are shown above branches. Accession numbers: Magnolia alba NC037005, Magnolia liliiflora NC023238, Magnolia denudata NC018357, Magnolia sprengeri NC023242, Magnolia salicifolia NC023240, Magnolia biondii KY085894, Magnolia kobus NC023237, Michelia odora NC023239, Michelia shiluensis MN418056, Magnolia laevifolia NC035956, Magnolia insignis NC035657, Magnolia cathcartii NC023234, Magnolia yunnanensis NC024545, Magnolia sinica NC023241, Magnolia kwangsiensis NC015892, Magnolia conifera NC037001, Magnolia dandyi NC037004, Magnolia aromatica NC037000, Magnolia fordiana MF990562, Magnolia glaucifolia NC037003, Magnolia duclouxii NC037002, Magnolia tripetala NC024027, Magnolia officinalis NC020316, Magnolia grandiflora NC020318, Magnolia pyramidata NC023236, Magnolia dealbata NC023235, Liriodendron chinense NC030504, and Liriodendron tulipifera NC008326.
Forests 11 00267 g008
Table 1. Summary of the chloroplast genomes of Michelia shiluensis and four closely related species (Michelia odora, Magnolia laevifolia, Magnolia insignis, and Magnolia cathcartii).
Table 1. Summary of the chloroplast genomes of Michelia shiluensis and four closely related species (Michelia odora, Magnolia laevifolia, Magnolia insignis, and Magnolia cathcartii).
M. shiluensisM. odoraM. laevifoliaM. insignisM. cathcartii
AccessionMN418056NC023239NC035956NC035657NC023234
Genome
Length (bp)160,075160,070160,120160,117159,950
GC (%)39.2639.2639.2439.2439.22
LSC
length (bp)88,10588,09888,14588,19588,142
GC (%)37.9537.9537.937.9237.91
Length (%)55.0455.0455.0555.0855.11
SSC
length (bp)18,79618,80018,79918,78218,790
GC (%)34.2834.2834.3234.2534.13
Length (%)11.7411.7411.7411.7311.75
IR
length (bp)26,58726,58626,58826,57026,509
GC (%)43.243.243.243.1943.2
Length (%)16.6116.6116.6116.5916.57
No. of Genes (duplicated in IR)
Genes131(18)131(18)131(18)131(18)131(18)
PCGs86(7)86(7)86(7)86(7)86(7)
tRNA37(7)37(7)37(7)37(7)37(7)
rRNA8(4)8(4)8(4)8(4)8(4)
With introns16(5)16(5)16(5)16(5)16(5)
Notes: GC: Guanine-cytosine; LSC: Large single-copy; IR: Invert repeat; SSC: Small single-copy; PCG: Protein-coding gene.
Table 2. List of the annotated genes in the Michelia shiluensis chloroplast genome.
Table 2. List of the annotated genes in the Michelia shiluensis chloroplast genome.
Category of GenesSubcategory of GenesGene Names
Self-replicationrRNA genesrrn 4.5 *rrn5 *rrn16 *rrn23 *
tRNA genestrnA-UGC *trnC-ACAtrnD-GUCtrnE-UUC
trnF-GAAtrnfM-CAUtrnG-GCCtrnG-UCC
trnH-GUGtrnI-CAUtrnI-GAU *trnK-UUU
trnL-CAA *trnL-UAAtrnL-UAGtrnM-CAU *
trnN-GUU *trnP-UGGtrnQ-UUGtrnR-ACG *
trnR-UCUtrnS-GCUtrnS-GGAtrnS-UGA
trnT-GGUtrnT-UGUtrnV-GAC *trnV-UAC
trnW-CCAtrnY-GUA
Small subunit of ribosomerps2rps3rps4rps7 *
rps8rps11rps12 *rps14
rps15rps16rps18rps19
Large subunit of ribosomerpl2 *rpl14rpl16rpl20
rpl23 *rpl32rpl33rpl36
RNA polymerase subunitsrpoArpoBrpoC1rpoC2
PhotosynthesisATP synthase geneatpAatpBatpEatpF
atpHatpI
NADH dehydrogenasendhAndhB *ndhCndhD
ndhEndhFndhGndhH
ndhIndhJndhK
Cytochrome b/f complexpetApetBpetDpetG
petLpetN
Photosystem IpsaApsaBpsaCpsaI
psaJ
Photosystem IIpsbApsbBpsbCpsbD
psbEpsbFpsbHpsbI
psbJpsbKpsbLpsbM
psbNpsbTpsbZ
Large chain of RubiscorbcL
Other genesATP-dependent proteaseclpP
Cytochrome c biogenesisccsA
Acetyl-CoA carboxylaseaccD
Translation initiation factor IF-1infA
Membrane proteincemA
MaturasematK
Unknown functionHypothetical chloroplast reading frameycf1ycf2 *ycf3ycf4
ycf15 *
Note: “*” indicates duplicated genes.
Table 3. Characteristics of the genes that contain introns in the cp genome of Michelia shiluensis.
Table 3. Characteristics of the genes that contain introns in the cp genome of Michelia shiluensis.
GeneLocationExon I (bp)Intron I (bp)Exon II (bp)Intron II (bp)Extron III (bp)
trnK-UUULSC35249037
rps16LSC21782544
trnG-UCCLSC2476848
atpFLSC411706144
rpoC1LSC1624722434
ycf3LSC154729227739126
trnL-UAALSC3549150
trnV-UACLSC375845656539
clpPLSC24663029178169
petBLSC5786641
rpl2IR432658387
rps12IR/LSC114-25536232
ndhBIR756700777
trnI-GAUIR4293635
trnA-UGCIR3879935
ndhASSC5411078551
Table 4. Relative synonymous codon usage (RSCU) in the chloroplast genome of Michelia shiluensis.
Table 4. Relative synonymous codon usage (RSCU) in the chloroplast genome of Michelia shiluensis.
CodonAmino AcidCountRSCUCodonAmino AcidCountRSCU
UUUPhe7141.1UAUTyr6001.47
UUCPhe5860.9UACTyr2140.53
UUALeu6711.66UAA*1470.98
UUGLeu5101.26UAG*1370.91
CUULeu4391.09CAUHis4291.41
CUCLeu2060.51CACHis1780.59
CUALeu4000.99CAAGln5601.41
CUGLeu1970.49CAGGln2370.59
AUUIle8931.28AAUAsn7121.44
AUCIle5450.78AACAsn2780.56
AUAIle6470.93AAALys7771.37
AUGMet6021AAGLys3590.63
GUUVal4741.37GAUAsp6311.54
GUCVal2020.58GACAsp1890.46
GUAVal4871.41GAAGlu7801.42
GUGVal2230.64GAGGlu3180.58
UCUSer4571.6UGUCys2141.36
UCCSer2660.93UGCCys1000.64
UCASer3731.3UGA*1661.11
UCGSer1920.67UGGTrp4271
CCUPro3451.41CGUArg2561.15
CCCPro2130.87CGCArg720.32
CCAPro3071.25CGAArg2891.3
CCGPro1170.48CGGArg1290.58
ACUThr4261.49AGUSer3241.13
ACCThr2370.83AGCSer1070.37
ACAThr3581.26AGAArg4101.84
ACGThr1200.42AGGArg1790.8
GCUAla5371.82GGUGly5421.33
GCCAla1850.63GGCGly1710.42
GCAAla3351.13GGAGly6281.54
GCGAla1260.43GGGGly2910.71
Note: “*” indicates the stop codon.
Table 5. Distribution of single sequence repeats in different locations and regions among five Magnoliaceae species: Michelia shiluensis, Michelia odora, Magnolia laevifolia, Magnolia insignis, and Magnolia cathcartii.
Table 5. Distribution of single sequence repeats in different locations and regions among five Magnoliaceae species: Michelia shiluensis, Michelia odora, Magnolia laevifolia, Magnolia insignis, and Magnolia cathcartii.
SpeciesNumberLocationRegions
LSCIRSSCCDSIntronIGS
M. cathcartii1431037262518100
M. insignis1411027252615100
M. laevifolia141101726271896
M. odora13899725241995
M. shiluensis141102725251997
Notes: LSC: Large single copy; IR: Invert region; SSC: Small single copy; CDS: Coding sequence; IGS: Intergenic spacer.

Share and Cite

MDPI and ACS Style

Deng, Y.; Luo, Y.; He, Y.; Qin, X.; Li, C.; Deng, X. Complete Chloroplast Genome of Michelia shiluensis and a Comparative Analysis with Four Magnoliaceae Species. Forests 2020, 11, 267. https://doi.org/10.3390/f11030267

AMA Style

Deng Y, Luo Y, He Y, Qin X, Li C, Deng X. Complete Chloroplast Genome of Michelia shiluensis and a Comparative Analysis with Four Magnoliaceae Species. Forests. 2020; 11(3):267. https://doi.org/10.3390/f11030267

Chicago/Turabian Style

Deng, Yanwen, Yiyang Luo, Yu He, Xinsheng Qin, Chonggao Li, and Xiaomei Deng. 2020. "Complete Chloroplast Genome of Michelia shiluensis and a Comparative Analysis with Four Magnoliaceae Species" Forests 11, no. 3: 267. https://doi.org/10.3390/f11030267

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