Research Article
Print
Research Article
Complete chloroplast genome sequence of Rhododendron mariesii and comparative genomics of related species in the family Ericaeae
expand article infoZhiliang Li, Zhiwei Huang, Xuchun Wan, Jiaojun Yu, Hongjin Dong, Jialiang Zhang, Chunyu Zhang§, Shuzhen Wang
‡ Huanggang Normal University, Huanggang, China
§ Huazhong Agricultural University, Wuhan, China
Open Access

Abstract

Rhododendron mariesii Hemsley et Wilson, 1907, a typical member of the family Ericaeae, possesses valuable medicinal and horticultural properties. In this research, the complete chloroplast (cp) genome of R. mariesii was sequenced and assembled, which proved to be a typical quadripartite structure with the length of 203,480 bp. In particular, the lengths of the large single copy region (LSC), small single copy region (SSC), and inverted repeat regions (IR) were 113,715 bp, 7,953 bp, and 40,918 bp, respectively. Among the 151 unique genes, 98 were protein-coding genes, 8 were tRNA genes, and 45 were rRNA genes. The structural characteristics of the R. mariesii cp genome was similar to other angiosperms. Leucine was the most representative amino acid, while cysteine was the lowest representative. Totally, 30 codons showed obvious codon usage bias, and most were A/U-ending codons. Six highly variable regions were observed, such as trnK-pafI and atpE-rpoB, which could serve as potential markers for future barcoding and phylogenetic research of R. mariesii species. Coding regions were more conserved than non-coding regions. Expansion and contraction in the IR region might be the main length variation in R. mariesii and related Ericaeae species. Maximum-likelihood (ML) phylogenetic analysis revealed that R. mariesii was relatively closed to the R. simsii Planchon, 1853 and R. pulchrum Sweet,1831. This research will supply rich genetic resource for R. mariesii and related species of the Ericaeae.

Keywords

chloroplast genome, comparative genomics, conservation genetics, phylogeny, Rhododendron mariesii

Introduction

Rhododendron mariesii Hemsley et Wilson, 1907, a typical member of the family Ericaeae, is mainly distributed in central China (Wang et al. 2018). Well known for leaf shape and bright-colored flowers, R. mariesii possesses valuable medicinal and horticultural properties (Wang et al. 2018). The deciduous species R. mariesii attracted great interest of Rhododendron breeders and geneticists. Furthermore, R. mariesii is very important in the woodland flora of the Dabie Mountains, and plays critical roles in ecological balance (Wang et al. 2018). Recently, Rhododendron-based ecological tourism, habitat fragmentation, and human activities have exerted significant effects towards natural growth of the wild Rhododendron population (Wang et al. 2019). Therefore, the research on population genetics and ecological conservation of wild R. mariesii is vital and necessary. However, limited genome information is available for R. mariesii, which has largely hindered corresponding genetic and molecular research.

In higher plants, the majority of plastomes are circular and quadripartite architecture consisting of two inverted repeat regions (IRa and IRb), a large single-copy region (LSC), and a small single-copy region (SSC) (Daniell et al. 2016; Hu et al. 2020; Abdullah et al. 2021). As maternally inherited organelle, the angiosperm plastome has a relatively conserved gene content and stable structure, which offers genetic markers sufficient for genome-wide evolutionary investigation at various taxonomic levels (Asaf et al. 2016; Zhang et al. 2017; Givnish et al. 2018). In plants, the size of cp genomes varied from 107 to 280 kb, containing approximately 130 genes related to photo synthesis and carbon fixation (Daniell et al. 2016; Rossini et al. 2021). The substitution rate of cp genome is lower than that of the nuclear genome, and 115–165 kb in cp genome is highly evolutionarily conserved (Smith 2015). However, specific genes exhibit accelerated evolution rates, such as ycf1, matK, and rbcL, which often serve as DNA barcoding (Dong et al. 2015; Wambugu et al. 2015; Zhang et al. 2017).

Next generation sequencing (NGS) has greatly increased the availability of genome data for non-species model, which facilitates the comparative cp genomics and phylogenetic studies at interspecific level (Santos and Almeida 2019; Pervez et al. 2022). In this research, the cpDNA of R. mariesii was assembled and annotated, SSR loci were characterized, comparative genomics and phylogenetic studies were also performed, hoping to benefit the studies of population evolution and conservation genetics of R. mariesii and related species.

Material and methods

Materials sampling and DNA extraction

Young and disease-free leaves of wild R. mariesii were sampled from the Dabie Mountains (central China, 29°16.13'N, 115°27.07'E, 1,005 m), dried in silica, and stored at -20 °C until further usage. In particular, sample collection was authorized by the Biodiversity Conservation of Huanggang Normal University. The specimens were identified by Hongjin Dong (Huanggang Normal University), who possesses a doctoral degree in botany. All materials were well conserved in the Huanggang Normal University Herbarium (Hubei province, China). Total genomic DNA was extracted and purified from fresh leaves according to Wang et al. (2019). Subsequently, the quality of total genome DNA was verified in 1% agarose gel stained by GelRed and quantified by spectophotometer (NanoDrop 1000, Thermofisher Scientific, USA).

Genome sequencing, assembly, and annotation

Nextera DNA library preparation kit was used to construct the paired-end Illumina libraries. These libraries were sequenced on Illumina NovaSeq6000 Sequencing System (Illumina, Hayward, CA) in a paired-end run (500 cycles, 1 × 250 pb). After trimming adapter sequences and removing low-quality sequences, raw data was filtered by SOAPnuke software (Chen et al. 2018). Then, the high-quality reads were de novo assembled by GetOrganelle pipeline (Jin et al. 2020). BOWTIE2 were used to validate the assembled sequence error of R. mariesii cp genome through mapping raw sequencing reads to the assembled plastome (Hanussek et al. 2021). Online program Organelle Genome DRAW (OGDRAW) was used to draw the physical map of R. mariesiicp genome (Greiner et al. 2019). Furthermore, gene annotation and analysis were carried out with DOGMA and CPGAVAS softwares, respectively (Liu et al. 2012). The final annotations were also manually verified by Geneious (ver.8.0.2) (Yu et al. 2022). The cp genome data had been submitted to the National Center for Biotechnology Information (NCBI) database (https://www.ncbi.nlm.nih.gov/).

Codon usage and nucleotide diversity analysis

Codon usage frequency was analyzed by CodonW software (https://sourceforge.net/projects/codonw/). Particularly, all protein coding genes were used for analysis. Relative synonymous codon usage (RSCU) analysis was carried out to measure codon usage bias (Rossini et al. 2021). The RSCU referred to the ratio of observed frequency of codons to frequency expected in regarding to the equal usage of synonymous codons for a certain amino acid (Rossini et al. 2021). In particular, RSCU value more than 1 means a preferred codon, otherwise the value less than and equal to 1 are considered as no codon usage bias (Morton 2022).

In total, eleven full chloroplast genomes of genus Rhododendron were downloaded from NCBI database: R. molle Siebold et Zuccarini, 1846; R. griersonianum Balfour filius et Forrest, 1919; R. pulchrum (Sweet) George Don, 1834; R. henanense Fang, 1983; R. micranthum Maximowicz, 1870; R. delavayi Franchet, 1886; R. concinnum Hemsley, 1890; R. simsii Planchon, 1876; R. platypodum Diels, 1990; R. datiandingense Feng, 1996; and R. kawakamii Hayata, 1911. Unique genes of these ten downloaded and the newly assemble R. mariesii cp genomes were extracted with PHYLOSUITE v1.2.2 and aligned by Windows version of MAFFT software, then nucleotide diversity (Pi) was calculated for each unique gene with DNASP ver 6.12.03 (Rossini et al. 2021).

Simple sequence repeats (SSR) analysis

MISA software (MicroSAtellite identification tool v2.1, http://pgrc.ipk-gatersleben.de/misa) was used to identify SSR motifs. Minimum number of tandem repeat units were set as follows: five repeat units for tri-, tetra-, penta-, and hexanucleotide SSRs; six repeat units for di-nucleotide SSRs; 10 repeat units for mono-nucleotide SSRs. The maximal number of bases interrupting two SSRs in a compound microsatellite was 100 bp.

Phylogenetic analysis

Through searing NCBI database, 21 cp genomes of Ericaceae species were found and downloaded: 12 species of Rhododendron; two species of Vaccinium Linnaeus, 1753; Arbutus unedo Sims, 1822; Hemitomes congestum Asa Gray, 1858; Allotropa virgata Torrey et Gray, 1868, Monotropa hypopitys Linnaeus, 1753; Pityopus californicus (Eastwood) H.F.Copeland, 1935; and 2 species of Gaultheria Kalm, 1753. Together with the newly assembled R. mariesii cp genome, these 22 cp genomes were used to construct phylogeny tree. These cp genomes were initially aligned with MAFFT for phylogenetic analysis (Yu et al. 2020). RAxML (version 8.2.8 for Windows) was used to run maximum likelihood (ML) analysis with a bootstrap value of 1000 (Alexandros 2014). FIGTREE v1.4 was used to visualize and adjust the ML trees (Yu et al. 2022). In particular, cp genome of Pyrola rotundifolia Benth. (1840) played the roles of an out-group.

Comparative analysis of genome structure

The structural characteristics of cp genomes, containing newly assembled R. mariesii and 10 cp genomes of the genus Rhododendron (R. delavayi, R. henanense, R. micranthum, R. concinnum, R. griersonianum, R. simsii, R. kawakamii, R. molle, R. platypodum, and R. datiandingense) were compared and analyzed with mVISTA online tool (using Shuffle-LAGAN alignment program). In particular, the annotated cp genome of R. mariesii served as a reference against the other cp genome. Genome alignments, including rearrangements or inversions, was detected with MAUVE (Darling et al. 2004). For investigating whether expansion or contraction occurred in R. mariesii cp genome, the IR/LSC and IR/SSC junction regions were compared with IRscope software (Amiryousefi et al. 2018).

Results

General features of R. mariesii chloroplast genome

In total, 19,498,900 reads were obtained from NovaSeq paired-end run. After stringent quality assessment and filtering, 19,309,162 clean reads (2.891 Gb) with an average of 149 bp read length were obtained. The percentage of clean reads was 99.03%, and the clean bases were 2,891,089,781 bp. In particular, GC content was 39.52%. In addition, Q20 (a base with quality value greater than 20) and Q30 (a base with quality value greater than 20) values were 97.28% and 92.34%, respectively. The size of R. mariesii cp genome is 203,480 bp. Moreover, typical quadripartite structure was observed, as a large single-copy (LSC) region (113,715 bp) and a small single-copy (SSC) region (7,953 bp) were separated by a pair identical inverted repeat regions (IRs) (40,918 bp) (Fig. 1).

Figure 1.

The chloroplast genome map of R. mariesii. Thick lines represented LSC, SSC, and IR regions, respectively. Genes shown inside circle were transcribed counterclockwise, and the outside outer circle were transcribed clockwise. Different gene groups were represented by different colors.

In total, 151 genes were successfully annotated, including 98 protein-coding genes, 45 tRNA genes, and 8 rRNA genes. The lengths of CDS, rRNA, tRNA, intergenic regions, and introns were 65,889 bp (32.38%), 8,998 bp (4.42%), 3,449 bp (1.7%), 45,409 bp (22.32%), and 80,033 bp (39.33%), respectively. The GC content of CDS, rRNA, tRNA, intergenic regions, and intron were 37.67%, 54.87%, 51.49%, 32.06%, and 33.75%, respectively. A set of 55 photosynthesis-related genes were found, containing six subunits of ATP synthase (atpA, atpB, atpE, atpF, atpH, and atpI), seven subunits of photosystem I, 17 subunits of photosystem II, 17 subunits of NADH-dehydrogenase, seven subunits of cytochrome b/f complex, and one subunit of rubisco (rbcL) (Table 1). Considering the self replication, 11 genes were large subunits of ribosome, four genes were DNA-dependent RNA polymerase, one gene was translational initiation factor, eight genes were rRNA genes, 45 genes were tRNA genes, and 18 genes were small subunit of ribosome (Table 1). The other genes were related to acetyl-CoA carboxylase, c-type cytochrome synthesis gene, envelope membrane protein, and maturase (Table 1). In addition, there were three conserved open reading frames, including one ycf3 and two ycf4. Totally, 16 genes contained introns, containing trnK-UUU, ycf3, trnL-UAA, trnV-UAC, ropB, atpF, trnS-CGA, accD, rpl16, ndhB, trnE-UUC, trnA-UGC, ndhA, trnA-UGC, and trnE-UUC (Table 2). Besides ycf3 and accD genes (three exons and two introns), the other 14 genes all had two exons and one intron.

Table 1.

Gene content of R. mariesii chloroplast genome. The duplicated genes were included into brackets.

Category of genes Group of genes Name of genes
Genes for photosynthesis Subunits of ATP synthase atpA, atpB, atpE, atpF, atpH, atpI
Subunits of photosystem I psaA, psaB, psaC (2×), psaI (2×), psaJ
Subunits of photosystem II psbA, psbB, psbC, psbD, psbE, psbF, psbH, psbI, psbJ (3×), psbK, psbL, psbM, psbN, psbT, psbZ
Subunits of NADH-dehydrogenase ndhA (2×), ndhB, ndhC, ndhD (2×), ndhE (2×), ndhF, ndhG (2×), ndhH (2×), ndhI (2×), ndhJ, ndhK
Subunits of cytochrome b/f complex petA (2×), petB, petD, petG, petL, petN
Subunit of rubisco rbcL
Self replication Large subunit of ribosome rpl2, rpl14, rpl16, rpl20, rpl22 (3×), rpl32 (2×), rpl33, rpl36
DNA dependent RNA polymerase rpoA, rpoB, rpoC1, rpoC2
Translational initiation factor infA
Ribosomal RNA genes rrn5S (2×), rrn16S (4×), rrn23S (2×)
Transfer RNA genes trnK-UUU, trnH-GUC, trnS-GGA, trnT-UGU , trnT-GGU, trnL-UAA, trnL-CAA, trnL-UAG (2×), trnM-CAU (12×), trnF-GAA, trnV-UAC, trnV-GAC (2×), trnR-UCU, trnR-ACG (2×), trnS-CGA, trnS-GCU, trnS-UGA, trnQ-UUG, trnW-CCA, trnP-UGG, trnC-GCA, trnD-GUC, trnY-GUA, trnE-UUC (2×), trnA-UGC (2×), trnN-GUU, trnA-UGC, trnE-UUC (2×)
Small subunit of ribosome rps2, rps3 (3×), rps4, rps7, rps8, rps11, rps14, rps15 (3×), rps16, rps18 (2×), rps19 (3×)
Other genes Subunit of acetyl-CoA-carboxylase accD
c-type cytochrom synthesis gene ccsA (2×)
Envelop membrane protein cemA (2×)
Maturase matK
Unkown function Conserved open reading frames ycf3, ycf4 (2×)
Table 2.

The characteristics list of genes possessing introns.

Gene Strand Start End ExonI IntronI ExonII IntronII ExonIII
trnK-UUU - 1,834 4,404 37 2499 35
ycf3 - 6,794 8,753 124 711 232 742 151
trnL-UAA + 11,313 11,909 35 512 50
trnV-UAC - 15,031 15,692 39 588 35
rpoB + 21,836 25,719 3,169 677 38
atpF + 35,554 36,820 161 700 406
trnS-CGA - 38,853 39,609 31 666 60
accD + 55,356 56,894 571 159 150 54 605
rpl16 - 59,253 167,784 9 108,121 402
ndhB - 101,387 103,550 721 685 758
trnE-UUC + 112,521 113,535 32 943 40
trnA-UGC + 113,600 114,490 37 818 36
ndhA + 126,079 128,272 563 1090 541
ndhA - 181,938 184,131 563 1090 541
trnA-UGC - 195,720 196,610 37 818 36
trnE-UUC - 196,675 197,689 32 943 40

Codon usage analysis and nucleotide diversity analysis

In the R. mariesii chloroplast genome, the protein-coding regions presented 40,013 codons (Table 3). Particularly, leucine (Leu) was the main amino acid (10.477%), followed by isoleucine (Ile, 8.972%) and glycine (Gly, 7.148%) (Fig. 2). In particular, cysteine (Cys) and tryptophan (Trp) were the lowest representative amino acids, accounting for 1.180% and 1.869%, respectively. According to RSCU values, a total of 30 codons showed obvious codon usage bias, as RSCU value were more than 1 (Table 3). Except Leu codon (UUG), all the other 29 codons were A/U-ending. For the 34 codons with RSCU values less than 1, 31 were C/G-ending, while 3 were A/U-ending.

Table 3.

The relative synonymous codon usage in R. mariesii cp genome.

Amino acid Codon No RSCU The codon frequency per amino acid(%) Amino acid Codon No RSCU The codon frequency per amino acid(%)
Ala GCA 703 1.17 29.16 Pro CCA 509 1.22 30.55
GCC 345 0.57 14.31 CCC 294 0.71 17.65
GCG 275 0.46 11.41 CCG 202 0.48 12.12
GCU 1088 1.81 45.13 CCU 661 1.59 39.67
Cys UGC 118 0.5 24.99 Gln CAA 1053 1.59 79.65
UGU 354 1.5 74.97 CAG 269 0.41 20.35
Asp GAC 273 0.39 19.57 Arg AGA 653 1.64 27.37
GAU 1122 1.61 80.44 AGG 179 0.45 7.5
Glu GAA 1395 1.54 77.2 CGA 626 1.57 26.24
GAG 412 0.46 22.8 CGC 147 0.37 6.16
Phe UUC 753 0.64 32.19 CGG 158 0.4 6.62
UUU 1586 1.36 67.8 CGU 623 1.57 26.11
Gly GGA 1079 1.51 37.73 Ser AGC 188 0.4 6.61
GGC 327 0.46 11.43 AGU 580 1.22 20.41
GGG 465 0.65 16.26 UCA 507 1.07 17.84
GGU 989 1.38 34.58 UCC 415 0.88 14.6
His CAC 223 0.47 23.28 UCG 240 0.51 8.44
CAU 735 1.53 76.73 UCU 912 1.93 32.09
Ile AUA 1125 0.94 31.34 Thr ACA 640 1.22 30.39
AUC 674 0.56 18.78 ACC 411 0.78 19.52
AUU 1791 1.5 49.89 ACG 213 0.4 10.11
Lys AAA 1631 1.54 77.11 ACU 842 1.6 39.98
AAG 484 0.46 22.88 Val GUA 845 1.44 36.11
Leu CUA 518 0.74 12.36 GUC 302 0.52 12.91
CUC 251 0.36 5.99 GUG 319 0.55 13.63
CUG 248 0.35 5.92 GUU 874 1.49 37.35
CUU 889 1.27 21.21 Trp UGG 748 1 100.02
UUA 1475 2.11 35.18 Tyr UAC 306 0.41 20.32
UUG 811 1.16 19.35 UAU 1200 1.59 79.68
Met AUG 954 1 100.01 Stop* UAA 133 1.22 40.69
Asn AAC 382 0.46 22.78 UAG 88 0.81 26.92
AAU 1295 1.54 77.22 UGA 106 0.97 32.42
Figure 2.

Occurrence percentage of amino acids in R. mariesii chloroplast genome.

Nucleotide diversity analysis showed that sequence level of divergence existed between different Rhododendron cp genomes. Pi values for each gene region varied from 0 to 0.06896. High level of genetic variation mainly existed in SSC region (Pi = 0.01723), followed by LSC (Pi = 0.00697) and IR (Pi = 0.001224) regions (Fig. 3). In total, six gene regions showed high levels of nucleotide diversity (Pi > 0.02), containing trnI-GAU (Pi = 0.06896), trnG-UCC (Pi = 0.06721), rps3 (Pi = 0.04509), rps12 (Pi = 0.03947), trnV-UAC (Pi = 0.03622), and trnK-UUU (Pi = 0.02554).

Figure 3.

The nucleotide diversity (Pi) of 11 Rhododendron species chloroplast genomes. X-axis presented the position of aligned chloroplast genomes, and Y-axis referred to Pi value. Below the X-axis, large single-copy (LSC), small single-copy (SSC), as well as inverted repeat (IR) regions were displayed with arrow bars.

SSR analysis of R. mariesii plastome

A set of 70 SSRs were identified from R. mariesii cp genome, and 5 SSRs were present in compound formation. Particularly, 65 SSRs (92.86%) were mononucleotide motifs, 2 were dinucleotide motifs (2.86%), 2 were trinucleotide motifs (2.86%), and 1 were hexanucleotide repeats (1.43%) (Table 4). Dominant mononucleotide repeats were A/T (91.429%), while C/G repeats accounted for 1.429%. In related to dinucleotide motifs, only AT/AT type was found (2.857%). For trinucleotide motifs, only one (AAG/CTT)5 and one (AAT/ATT) 11 motif were found. For hexanucleotide repeats, one (AAGGGT/ACCCTT)5 was found, accounting for 1.429%.

Table 4.

The frequency of each type of microsatellite in R. mariesii cp genome.

Repeats 5 6 7 8 9 10 11 12 13 14 15 total Percentage
A/T - - - - - 32 14 9 7 2 64 91.429%
C/G - - - - - 1 1 1.429%
AT/AT - 2 2 2.857%
AAG/CTT 1 1 1.429%
AAT/ATT 1 1 1.429%
AAGGGT/ACCCTT 1 1 1.429%

Mononucleotide A/T repeats with repeat numbers of 10–14 were the most abundant. Meanwhile, (C/G)n microsatellites were all repeated 15 times. In relation to dinucleotide repeats, the identified SSRs all have 7 repeat motifs. Regarding to trinucleotide motifs, AAG/CTT and AAT/ATT microsatellites repeated 5 and 11 times, respectively. The hexanucleotide motif AAGGGT/ACCCTT repeated 5 times. Totally, 34 SSRs were present in the intergenic spacer region, accounting for 41.43%. Moreover, 28 SSRs were present in rpl16 gene. All the remaining 13 microsatellites were found in ccsA, cemA, ndhA, rpoA, rpoC2, rps7, rps8, and trnL-UAA genes.

Phylogenetic analysis

For clarifying the phylogenetic location of R. mariesii among the Ericaeae, complete plastomes of R. mariesii and other 21 species in the Ericaeae with fully sequenced chloroplast genomes were used in reconstructing phylogenetic relationships. The phylogenetic tree revealed that R. mariesii had a close genetic relationship with R. simsii and R. pulchrum (Fig. 4). In particular, all these 22 taxa belonging to Ericaeae were grouped into one clade and clustered into two subclades. Topological structure was almost consistent with the previously published phylogeny (Liu et al. 2021). A. unedo (NCBI: JQ067650), H. congestum (NCBI: NC_035581), A. virgate (NCBI: NC_035580), M. hypopitys (NCBI: NC_029704), P. californicus (NCBI: NC_035584), Vaccinium oldhamii Miquel, 1866 (NCBI: MK049537), Vaccinium macrocarpon Aiton, 1789 (NCBI: JQ757046), Gaultheria griffithiana Wight, 1847 (NCBI: MW528025), Gaultheria fragrantissima Wallich, 1820 (NCBI: MW563322), and P. rotundifolia (NCBI: KU833271), were relatively distant related with R. mariesii.

Figure 4.

The maximum-likelihood phylogenetic tree for R. mariesii. Numbers on each node referred to bootstrap support value.

Comparative plastome sequence divergence and hotspot regions

Structural characteristics of 11 Rhododendron cp genomes were investigated with mVISTA software, containing the newly assembled R. mariesii cp genome and 10 download cp genome of the Rhododendron genus. In particular, the annotated R. mariesii cp genome served as a reference. Relatively high similarity was detected among these 11 Rhododendron species. Coding regions were more conserved than non-coding regions (CNS in Fig. 5). The LSC and SSC regions were relatively more stable than IR regions. Among these coding regions, rpoB, rpoC2, rps8, petD, rpl23, rpl22, and ndhF were relatively divergent because of intron regions. In R. mariesii plastid genome, highly variable regions mostly existed in the intergenic spacer, such as trnK-pafI, atpE-rpoB, trnT-rpl14, rpoA-psbJ, rpl20-trnE, ndhI-rps19, and rpl16-ndhI. Compared with intergenic spacer, protein coding regions were highly conserved, such as rps4, ndhJ, ndhK, rpcL, rps2, atpI, psaA, psaB, psbB, cemA, and petA. No rearrangements and inversions occurred in these 11 cp genomes of Rhododendron species.

Figure 5.

Comparison of cp genomes with R. mariesii annotation serving as the reference. Vertical scale indicated the percentage of identity (50–100%), and horizontal axis was coordinates within cp genome. The genome regions were color-coded as exons, introns, and conserved non-coding sequences, respectively.

Particularly, lengths of the IR regions of 6 cp genomes ranged from 14,194 bp (R. mariesii cp genome) to 47,467 (R. griersonianum cp genome) (Fig. 6). Expansion and contraction existed in these cp genomes. In R. griersonianum cp genome, JLB line (line between LSC and IRb) was located between genes ycf15 and trnR, while ycf15 was located in the LSC region with 166 bp extending to IRb region. In R. mariesii cp genome, JLB line was located between trnV and rrn16 (476 bp extending to LSC region). In R. micranthum and R. henanense cp genomes, the JLB lines were located between rps12 and trnV, while trnV was located in IRb region with 913 bp and 935 bp extending to LSC region, respectively. However, JLB line was located between rps7 and trnI, and trnI was located in IRb region with 911 bp extending to LSC region in R. concinnum. Except for R. mariesii, ndhF was located in SSC region with 296 bp–314 bp extending to IRb region in the five other cp genomes (Fig. 6). Meanwhile, rps15 was located in SSC region of R. mariesii cp genome. The JSA line (line between SSC and IRa) was located between ndhF and rpl32, and ndhF was distributed in SSC region with 54 bp, 53 bp, 67 bp, 54 bp, and 37 bp to IRa region in R. griersonianum, R. concinnum, R. micranthum, R. henanense, and R. delavayi cp genomes, respectively. Besides ndhH, rps15 was also located in SSC region with 114 bp extending to IRa region in R. mariesii cp genome. Furthermore, JLA line (line between IRa and LSC) was located between trnV and psbA in R. concinnum, R. micranthum, and R. henanense. However, JLA line was located between trnR and trnH in R. griersonianum cp genome. In R. mariesii cp genome, rrn16 and trnV were located besides the JLA line.

Figure 6.

The comparison of LSC, SSC, and IR regional boundaries of cp genome between R. mariesii and related taxa. JLB, JSB, JSA, and JLA respected “junction line between LSC and IRb”, “junction line between IRb and SSC”, “junction line between SSC and IRa” , as well as “junction line between IRa and LSC”, respectively.

Discussion

The chloroplast genome is the main organelle for plant transforming light energy into chemical energy (Zhang et al. 2018). Plastome genome is useful for comparative genomic research and phylogenomic analyses due to polymorphic regions generated through genomic expansion, inversion, contraction, and gene rearrangement (Sanitá Lima et al. 2016; Kahraman et al. 2019; Li et al. 2019; Li et al. 2020; Liu et al. 2020; Wang et al. 2020). The single circular cp genome structure of R. mariesii was the same as other species belonging to the Ericaeae with a typical quadripartite structure and similar GC content unevenly distributed across the cp genome (Liu et al. 2021; Xu et al. 2022). The GC content of R. mariesii cp genome (39.52%) was slightly larger than that of Myracrodruon urundeuva Allemão, 1862 (37.8%) (Rossini et al. 2021). Relative to both LSC (35.85%) and SSC (36.49%) regions, the GC content in IR region (30.48%) is lower. However, the GC content in IR region was larger than LSC and SSC regions in cp genome Xanthium spinosum Linnaeus, 1753 (Raman et al. 2020). The length of total genome size and each region were similar to other plant cp genomes, such as R. molle, Rubus species (Rosaceae), and rubber dandelion (Taraxacum kok-saghyz Rodin) (Zhang et al. 2017; Xu et al. 2022; Yu et al. 2022).

The size of R. mariesii cp genome (203,480 bp) was larger than that of R. pulchrum (146,941 bp), R. simsii (152,214 bp), R. molle (197,877 bp), R. delavayi (193,798 bp), and R. platypodum (201,047 bp), but smaller than that of R. kawakamii (230,777 bp), R. micranthum (207,233 bp), R. henanense (208,015 bp), R. griersonianum (206,467 bp), R. concinnum (207,236 bp), and R. datiandingense (207,311 bp). Totally, 151 genes existed in R. mariesii cp genome, which were more than that of R. molle (149 genes) and R. pulchrum (73 genes) (Shen et al. 2020; Xu et al. 2022). In particular, protein-coding genes accounted for 64.901% in R. mariesii cp genome, which were lower than that of R. molle cp genome (65.101%) but higher than that of R. pulchrum cp genome (57.534%) (Shen et al. 2020; Xu et al. 2022). Moreover, 45, 44, and 29 tRNA genes were found in cp genomes of R. mariesii, R. molle, and R. pulchrum, respectively. In both R. mariesii and R. molle cp genomes, eight rRNA genes were annotated, but only two were found in R. pulchrum cp genome (Shen et al. 2020; Xu et al. 2022).

Besides genes involved in photosynthesis transforming light energy into chemical energy, other genes also existed in R. mariesii cp genome. For example, accD gene, encoding plastid beta carboxyl transferase subunit of acetyl-CoA carboxylase (ACCase) important for plant growth (leaf growth, leaf longevity, fatty acid biosynthesis, and embryo development), has been reported to be involved in the adaptation to specific ecological niches during radiation of dicotyledonous plants (Hu et al. 2015). In R. mariesii cp genome, one copy of accD gene was also found. Codons coding Leu (10.477%), Ile (8.972%), and Gly (7.148%) were dominant, while Cys (1.180%) and Trp (1.869%) were the least, which were the same as that of M. urundeuva cp genome (Rossini et al. 2021). Codon bias, an efficient mechanism of translation influenced by natural selection and mutation pressure, takes place if synonymous codons are used at different frequencies (Zhang et al. 2022). A total of 30 codons showed codon usage bias, and most were A/U-ending codons, which were the same as that observed in M. urundeuva and Solanum (Zhang et al. 2018a; Rossini et al. 2021). In total, six gene regions showed high levels of nucleotide diversity (Pi > 0.02), containing trnI-GAU, trnG-UCC, rps3, rps12, trnV-UAC, and trnK-UUU, serving as the first candidate for developing molecular markers to identify Rhododendron species.

A total of 70 SSRs were identified from R. mariesii cp genome, more than that of M. urundeuva (36 SSRs), Spondias bahiensis P. Carvalho, 2015 (53 SSRs) and Mangifera indica Wallich, 1847 (57 SSRs), but fewer than that of Syringa pinnatifolias Hemsley, 1906 (253 SSRs) (Jo et al. 2017; Santos and Almeida 2019). Variation in the number and type of microsatellites might play important roles in plastome organization. The main motifs were A/T repeats (91.429%), which was the same with that of M. urundeuva, S. bahiensis, and M. indica (Jo et al. 2017; Santos and Almeida 2019; Rossini et al. 2021). However, no correlation was found between large repeat regions and rearrangement endpoints, which was similar with Liu et al. (2013). Very limited tandem (G/C)n-containing microsatellites were observed, which might be due to the low content of G and C bases in chloroplast genome. Molecular markers developed for the intergenic regions could be used for phylogenetic, phylogeographic, and barcoding studies of Rhododendron species.

Non-coding regions often mutate relatively faster than coding regions (Yu et al. 2022). In R. mariesii cp genome, coding regions were more conserved than the non-coding regions. Relatively high similarity was detected among these Rhododendron cp genomes, but expansion and contraction also existed in IR regions, which might be the dominant reason for variation in cp genome size. Obvious differences were found in cp IR boundary regions, containing gene contents and locations. However, IR regions were least divergent, which were mainly due to the presence of four highly conserved rRNA sequences in X. spinosum (Raman et al. 2020). Furthermore, LSC and SSC regions were relatively more stable than IR regions in R. mariesii cp genome. These genetic variations may significantly facilitate R. mariesii adapting to the changes of survival conditions. According to neutral theory, nucleotide substitution in non-coding regions (intergenic spacer, intron region, and pseudogenes) are neutral or near-neutral, which could not be affected by natural selection (Akashi et al. 2012). Therefore, evolutionary history of R. mariesii could be well calculated from the rate of molecular evolution in non-coding region.

This research aimed to expand the molecular genetic resources available for R. mariesii through high-throughput sequencing and cp genome assembly. The R. mariesii cp genome sequence could be used in distinguishing and resolving phylogenetic relationships within Ericaeae lineage. Moreover, this research will be vital for further genetic analysis on R. mariesii and other species in the Ericaeae family.

Conflicts of interests

The authors declare that they have no competing interests.

Data availability statement

The cp genome of R. mariesii was submitted to GenBank database under the accession number of OM161981.

Acknowledgements

This work was supported by grant from Scientific and Technological Research Project of Hubei Provincial Department of Education (B2022204) and Open fund of Hubei Key Laboratory of Economic Forest Germplasm Improvement and Resources Comprehensive Utilition (202303202).

References

  • Abdullah , Henriquez CL, Mehmood F, Hayat A, Sammad A, Waseem S, Waheed MT, Matthews PJ, Croat TB, Poczai P, Ahmed I (2021) Chloroplast genome evolution in the Dracunculus clade (Aroideae, Araceae). Genomics 113(1): 183–192. https://doi.org/10.1016/j.ygeno.2020.12.016
  • Asaf S, Khan AL, Khan AR, Waqas M, Kang SM, Khan MA, Lee SM, Lee IJ (2016) Complete chloroplast genome of Nicotiana otophora and its comparison with related species. Frontiers in Plant Science 7: 843. https://doi.org/10.3389/fpls.2016.00843
  • Darling AC, Mau B, Blattner FR, Perna NT (2004) Mauve: multiple alignment of conserved genomic sequence with rearrangements. Genome Research 14: 1394–1403. https://doi.org/10.1101/gr.2289704
  • Dong W, Xu C, Cheng T, Lin K, Zhou S (2013) Sequencing angiosperm plastidgenomes made easy: a complete set of universal primers and a case study on thephylogeny of Saxifragales. Genome Biology and Evolution 5: 989–997. https://doi.org/10.1093/gbe/evt063
  • Dong W, Xu C, Li C, Sun J, Zuo Y, Shi S, Cheng T, Guo J, Zhou S (2015) ycf1, the most promising plastid DNA barcode of land plants. Scientific Reports 5: 8348. https://doi.org/10.1038/srep08348
  • Greiner S, Lehwark P, Bock R (2019) Organellar genome DRAW(OGDRAW) version 1.3.1: expanded toolkit for the graphical visualization of organellar genomes. Nucleic Acids Research 47(W1): W59–W64. https://doi.org/10.1093/nar/gkz238
  • Givnish TJ, Zuluaga A, Spalink D, Soto Gomez M, Lam VKY, Saarela JM, Sass C, Iles WJD, de Sousa DJL, Leebens-Mack J, Chris Pires J, Zomlefer WB, Gandolfo MA, Davis JI, Stevenson DW, de Pamphilis C, Specht CD, Graham SW, Barrett CF, Ané C (2018) Monocot plastid phylogenomics, timeline, net rates of species diversification, the power of multi-gene analyses, and a functional model for the origin of monocots. American Journal of Botany 105: 1888–1910. https://doi.org/10.1002/ajb2.1178
  • Hanussek M, Bartusch F, Krüger J (2021) Performance and scaling behavior of bioinformatic applications in virtualization environments to create awareness for the efficient use of compute resources. PLoS Computational Biology 17(7): e1009244. https://doi.org/10.1371/journal.pcbi.1009244
  • Hu G, Cheng L, Huang W, Cao Q, Zhou L, Jia W, Lan Y (2020) Chloroplast genomes of seven species of Coryloideae (Betulaceae): structures and comparative analysis. Genome 63(7): 1–12. https://doi.org/10.1139/gen-2019-0153
  • Hu S, Sablok G, Wang B, Qu D, Barbaro E, Viola R, Li M, Varotto C (2015) Plastome organization and evolution of chloroplast genes in Cardamine species adapted to contrasting habitats. BMC Genomics 16: 306. https://doi.org/10.1186/s12864-015-1498-0
  • Jin JJ, Yu WB, Yang JB, Song Y, de Pamphilis CW, Yi TS, Li DZ (2020) GetOrganelle: a fast and versatile toolkit for accurate de novo assembly of organelle genomes. Genome Biology 21(1): 241. https://doi.org/10.1186/s13059-020-02154-5
  • Jo S, Kim HW, Kim YK, Sohn JY, Cheon SH, Kim KJ (2017) The complete plastome sequences of Mangifera indica L. (Anacardiaceae). Mitochondrial DNA Part B-Resources 2(2): 698–700. https://doi.org/10.1080/23802359.2017.1390407
  • Kahraman K, Lucas SJ (2019) Comparison of different annotation tools for characterization of the complete chloroplast genome of Corylus avellana cv Tombul. BMC Genomics 20(1): 874. https://doi.org/10.1186/s12864-019-6253-5
  • Li X, Zuo Y, Zhu X, Liao S, Ma J (2019) Complete chloroplast genomes and comparative analysis of sequences evolution among seven Aristolochia (Aristolochiaceae) medicinal species. International Journal of Molecular Sciences 20(5): 1045. https://doi.org/10.3390/ijms20051045
  • Liu C, Shi L, Zhu Y, Chen H, Zhang J, Lin X, Guan X (2012) CpGAVAS, an integrated web server for the annotation, visualization, analysis, and genbank submission of completely sequenced chloroplast genome sequences. BMC Genomics 13(1): 715. https://doi.org/10.1186/1471-2164-13-715
  • Liu Q, Li X, Li M, Xu W, Schwarzacher T, Heslop-Harrison JS (2020) Comparative chloroplast genome analyses of Avena: insights into evolutionary dynamics and phylogeny. BMC Plant Biology 20: 406. https://doi.org/10.1186/s12870-020-02621-y
  • Liu Y, Li Q, Wang L, Wu L, Huang Y, Zhang J, Song Y, Liao J (2021) The complete chloroplast genome of Rhododendron molle and its phylogenetic positionwithin Ericaceae. Mitochondrial DNA Part B-Resources 6(9): 2587–2588. https://doi.org/10.1080/23802359.2021.1959458
  • Liu Y, Huo N, Dong L, Wang Y, Zhang S, Young HA, Feng X, Gu YQ (2013) Complete chloroplast genome sequences of Mongolia medicine Artemisia frigida and phylogenetic relationships with other plants. PLoS ONE 8(2): e57533. https://doi.org/10.1371/journal.pone.0057533
  • Morton BR (2022) Context-dependent mutation dynamics, not selection, explains the codon usage bias of most angiosperm chloroplast genes. Journal of Molecular Evolution 90: 17–29. https://doi.org/10.1007/s00239-021-10038-w
  • Pervez MT, Hasnain MJU, Abbas SH, Moustafa MF, Aslam N, Shah SSM (2022) A comprehensive review of performance of next-generation sequencing platforms. Biomed Research International 2022: 3457806. https://doi.org/10.1155/2022/3457806
  • Raman G, Park KT, Kim JH, Park S (2020) Characteristics of the completed chloroplast genome sequence ofxanthium spinosum: comparative analyses, identification of mutational hotspots and phylogenetic implications. BMC Genomics 21(1): 855. https://doi.org/10.1186/s12864-020-07219-0
  • Rossini BC, de Moraes MLT, Marino CL (2021) Complete chloroplast genome of Myracrodruon urundeuva and its phylogenetics relationships in Anacardiaceae family. Physiology and Molecular Biology of Plants 27(4): 801–814. https://doi.org/10.1007/s12298-021-00989-1
  • Santos V, Almeida C (2019) The complete chloroplast genome sequences of three Spondias species reveal close relationship among the species. Genetics and Molecular Biology 42(1): 132–138. https://doi.org/10.1590/1678-4685-gmb-2017-0265
  • Sanitá Lima M, Woods LC, Cartwright MW, Smith DR (2016) The (in)complete organelle genome: exploring the use and nonuse of available technologies for characterizing mitochondrial and plastid chromosomes. Molecular Ecology Resources 16: 1279–1286. https://doi.org/10.1111/1755-0998.12585
  • Shen J, Li X, Zhu X, Huang X, Jin S (2020) The complete plastid genome of Rhododendron pulchrum and comparative genetic analysis of Ericaceae species. Forests 11(2): 158. https://doi.org/10.3390/f11020158
  • Wambugu PW, Brozynska M, Furtado A, Waters DL, Henry RJ (2015) Relationships of wild and domesticated rices (Oryza AA genome species) based upon whole chloroplast genome sequences. Scientific Reports 5: 13957. https://doi.org/10.1038/srep13957
  • Wang S, Li Z, Guo X, Fang Y, Xiang J, Jin W (2018) Comparative analysis of microsatellite, SNP, and Indel markers in four Rhododendron species based on RNA-seq. Breeding Science 68(5): 536–544. https://doi.org/10.1270/jsbbs.18092
  • Wang S, Jin Z, Luo Y, Li Z, Fang Y, Xiang J, Jin W (2019) Genetic diversity and population structure of Rhododendron simsii (Ericaceae) as reveled by microsatellite markers. Nordic Journal of Botany 37(4): 1–10. https://doi.org/10.1111/njb.02251
  • Wang X, Rhein HS, Jenkins J, Schmutz J, Grimwood J, Grauke LJ, Randall JJ (2020) Chloroplast genome sequences of Carya illinoinensis from two distinct geographic populations. Tree Genetics and Genomes 16(4): 48. https://doi.org/10.1007/s11295-020-01436-0
  • Yu J, Dong H, Xiang J, Xiao Y, Fang Y (2020) Complete chloroplast genome sequence and phylogenetic analysis of Magnolia pilocarpa, a highly ornamental species endemic in central China. Mitochondrial DNA Part B-Resources 5(1): 720–722. https://doi.org/10.1080/23802359.2020.1714511
  • Zhang Y, Iaffaldano BJ, Zhuang X, Cardina J, Cornish K (2017) Chloroplast genome resources and molecular markers differentiate rubber dandelion species from weedy relatives. BMC Plant Biology 17: 34. https://doi.org/10.1186/s12870-016-0967-1
  • Zhang R, Zhang L, Wang W, Zhang Z, Du H, Qu Z, Li XQ, Xiang H (2018) Differences in codon usage bias between photosynthesis-related genes and genetic system-related genes of chloroplast genomes in cultivated and wild Solanum species. International Journal of Molecular Sciences19: 3142. https://doi.org/10.3390/ijms19103142
  • Zhang J, Huang H, Qu C, Meng X, Meng F, Yao X, Wu J, Guo X, Han B, Xing S (2021) Comprehensive analysis of chloroplast genome of Albizia julibrissin durazz. (Leguminosae sp.). Planta 255(1): 26. https://doi.org/10.1007/s00425-021-03812-z

Supplementary material

Supplementary material 1 

Taxonomic and accession information on cp genomes downloaded from NCBI database

Zhiliang Li, Zhiwei Huang, Xuchun Wan, Jiaojun Yu, Hongjin Dong, Jialiang Zhang, Chunyu Zhang, Shuzhen Wang

Data type: wps

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0/). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (56.50 kb)
login to comment