Next Article in Journal
Trends of Peatland Research Based on Topic Modeling: Toward Sustainable Management under Climate Change
Next Article in Special Issue
The Complete Plastid Genome Sequences of the Belian (Eusideroxylon zwageri): Comparative Analysis and Phylogenetic Relationships with Other Magnoliids
Previous Article in Journal
Trade-Off between Hydraulic Safety and Efficiency in Plant Xylem and Its Influencing Factors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Complete Chloroplast Genome Structural Characterization and Comparative Analysis of Viburnum japonicum (Adoxaceae)

Zhejiang Academy of Forestry, Hangzhou 310023, China
*
Author to whom correspondence should be addressed.
Forests 2023, 14(9), 1819; https://doi.org/10.3390/f14091819
Submission received: 3 August 2023 / Revised: 27 August 2023 / Accepted: 4 September 2023 / Published: 6 September 2023

Abstract

:
Viburnum japonicum (Thunb.) Sprengel is an endangered species endemic to coastal regions of eastern Asia (China, Japan, and South Korea). However, its systematic position has been controversial. In this study, we present the complete chloroplast (cp) genome of V. japonicum (GenBank OP644292) sequenced using the Illumina NovaSeq platform. The cp genome has a total length of 158,606 bp and a G+C contents of 38.08%. It consists of a large single-copy (LSC) region, a small single-copy (SSC) region, and a pair of inverted repeats (IRs) regions measuring 87,060 bp, 18,510 bp, and 26,516 bp, respectively. A total of 131 genes were identified, including 87 protein-coding genes, 36 transfer RNA genes (tRNAs), and 8 ribosomal RNA genes(rRNAs). Additionally, a total of 44 dispersed repeats were detected, including three types: forward, palindromic, and reverse. Among the 38 SSR loci that were discovered, the majority were mononucleotide loci composed of A/T. Furthermore, we found that 15 genes (accD, atpF, ndhA, ndhB, petB, petD, rpl16, rpl2, rpoC1, rps12, rps16, trnA-UGC, trnI, trnK-UUU, and trnL-UAA) contain one intron, while clpP and ycf3 have two introns. The relative synonymous codon usage (RSCU) analysis detected 31 high-frequency codons, where A/U bases accounted for 93.55% of the total, indicating an asymmetry in chloroplast gene and a presence for A/U bases. Comparative analysis of genome structure and sequences data of V. japonicum chloroplast genomes in comparison with other closely related species demonstrated a high level of conservation in their structure and organization. Furthermore, three mutation hotspots (psbH, rps19, and trnL) were identified, which could be valuable for future phylogenetic and population genetic research. Phylogenetic analysis revealed that the two accessions of V. japonicum are closely related to a group of V. setigerum, V. erosum, and V. fordiae within the Viburnum genus. In conclusion, this study provides important insights for accurately identifying and understanding the phylogeny of Viburnum species through the complete cp genome sequencing of V. japonicum.

1. Introduction

Viburnum japonicum (Thunb.) Sprengel 1825, also known as Japanese viburnum, belongs to the Adoxaceae family and is a valuable species for ornamental and economic purposes. For example, chavicol extracted from V. japonicum and its related effective compounds can significantly inhibit drosophila larvae and adults, making it a promising source of biopesticides [1]. It is easily distinguished from other recorded Viburnum species due to its rounded and evergreen shrubs, broad leaves, and glabrous texture. Previously, it was believed to be distributed only along the coast of Japan, mainly Honshu, Kyushu, and Ryukyu Islands. However, in 1994, Chinese plant taxonomists discovered the species on the east coast of Zhejiang Province. Since then, it has been recognized as a plant species with extremely small populations (PSESP) and included in Zhejiang’s list of the key protected wild plants. In 2003, V. japonicum was also observed in South Korea on Gageo-do Island [2], indicating its potential to spread beyond its native range. In recent years, research on V. japonicum has primarily focused on physiological resistance [3,4,5], breeding techniques [6,7], molecular marker development [8] and population genetics [9]. However, limited genomic exploration has been conducted, with only a few brief reports from South Korean scholars regarding chloroplast [10,11,12], and no further genomic information was accessible. The taxonomic status and phylogenetic position of V. japonicum remain controversial due to inadequate species sampling within the Viburnum genus.
The chloroplast genome of plants plays a crucial role in elucidating chloroplast functions, species identification, population genetics, phylogenetics, and conservation of plant resources [13,14,15,16]. In this study, we conducted a comprehensive analysis of the complete cp genome sequence of V. japonicum using second-generation sequencing technology. This analysis involved annotating functional genes, analyzing codon preference, examining simple sequence repeats (SSR), and comparing chloroplast genomes. The main objective was to deepen our understanding of the genomic information specific to this species. Consequently, the findings of this study will serve as a reliable reference genome for future investigations focused on molecular evidence and phylogenetic relationships among Asian Viburnum species.

2. Materials and Methods

2.1. Plant Material, DNA Extraction, and Sequencing

One individual of V. japonicum (Figure 1) was collected, which was deposited at the Arboretum of Zhejiang Academy of Forestry, Hangzhou, Zhejiang Province, China (30°12′ N, 120°01′ E) with a voucher number of HZ2022-05 (contact: Hong Zhu, Email: zh1107401987@126.com). The total genomic DNA was extracted from fresh leaves of the collected V. japonicum using a modified CTAB method [17]. Paired-end (PE) reads of 2 × 150 bp were conducted on an Illumina NovaSeq 6000 platform (Illumina, San Diego, CA, USA) at Personalbio Biotech (Shanghai, China).

2.2. Genome Assembly and Annotation

The raw data obtained underwent a filtering process where low-quality reads were removed, and high-quality reads were evaluated using the FastQC v0.120 software [18] (http://www.bioinformatics.babraham.ac.uk/projects/fastqc accessed on 2 August 2023). The filtering criteria were as follows: (1) Removing the sequencing linker and primer sequence from the reads; (2) Discarding reads with an average quality value lower than Q5; (3) Filtering out any reads with a length of 50 bp or less. Based on these cleaning reads, the cp genome of V. japonicum was assembled via Fast-Plast v1.2.9 software [19] (https://github.com/mrmckain/Fast-Plast accessed on 2 August 2023). The genome was annotated using GeSeq (https://chlorobox.mpimp-golm.mpg.de/geseq.html accessed on 2 August 2023). The annotated V. erosum (MN218778) chloroplast genome was used as a reference. The OGDRAW program (https://chloro-box.mpimpgolm.mpg.de/index.html accessed on 2 August 2023) was used to draw a detailed physical map of the V. japonicum complete cp genome. The CPGView tool [20] (http://www.1kmpg.cn/cpgview accessed on 2 August 2023) was utilized to visualize the structures of splicing genes within the chloroplast genome (Figures S1 and S2). The final annotated genome sequence of V. japonicum was submitted to the GenBank database and assigned the accession number OP644292.

2.3. Repeat Sequence Analysis

The MISA software [21] (https://webblast.ipk-gatersleben.de/misa/ accessed on 2 August 2023) was utilized to identify SSR sites in the chloroplast genome of V. japonicum. Dispersed repeats of four types (forward repeats (F), reverse repeats (R), complement repeats (C), and palindromic repeats (P)) were detected in the chloroplast whole genome using the REPute v2.5 software (https://bibiserv.cebitec.uni-bielefeld.de/reputer accessed on 2 August 2023). The parameters were set as follows: minimum length of 30 bp and hamming distance of 3.

2.4. Relative Condon Usage Analysis

The analysis of relative synonymous codon usage (RSCU) in the chloroplast genome coding genes was conducted using the CodonW v1.4.2 software [22] (https://www.softpedia.com/get/Science-CAD/CodonW.shtml accessed on 2 August 2023), with default parameter settings.

2.5. Genome Comparison Analysis

The differences in boundary sequences were visualized using the IRscope v2.01 software (https://irscope.shinyapps.io/irapp/ accessed on 2 August 2023). Homologous gene sequences of different species were compared globally using MAFFT software v 7.520 [23], and nucleotide variability (π) values were calculated for each gene using DnaSP v5 software [24], and scatter fold plots were plotted.

2.6. Phylogenetic Analysis

To clarify the phylogenetic position of V. japonicum within the genus Viburnum, 17 additional complete cp genome sequences (Dipsacus japonicus MZ934745, Sambucus chinensis MW455170, and S. williamsii KX510276 as outgroups) were aligned using the program MAFFT v7.520. The software RAxML v8.2.12 [25] was used for phylogenetic tree construction based on the maximum likelihood (ML) algorithm under the parameter model of GTRGAMMA and 1000 bootstrap replicates.

3. Results and Discussion

3.1. Genome Sequencing Assembly and Characterization

A total of 23,763,912 raw read and 3,588,350,712 raw bases were gained. The proportion of Q20 and Q30 was 97.45% and 92.92%, respectively. As a result, we obtained a scaffold with its high quality of assembled 155,827 bp in length. Finally, the complete cp genome of V. japonicum was successfully assembled with an average read coverage depth of 650X (Figure S3) and a typical circular structure of 158,606 bp in length, consisting of four subregions: a large single-copy (LSC) of 87,060 bp, a small single-copy (SSC) region of 18,510 bp, and a pair of inverted repeat regions (IRs), each of 26,516 bp (Figure 2). The overall nucleotide composition of the cp genome sequence was: A (30.62%), T (31.29%), C (19.38%), and G (18.71%), and the total G + C contents of the cp genome were 38.08%; LSC (36.42%), SSC (31.97%), and IRs (42.97%) (Table 1).
A total of 131 genes were annotated, including 87 protein-coding genes (PCGs), 8 ribosomal RNA genes (rRNAs), and 36 transfer RNA genes (tRNAs). Of these genes, 15 (accD, atpF, ndhA, ndhB, petB, petD, rpl16, rpl2, rpoC1, rps12, rps16, trnA-UGC, trnI, trnK-UUU, and trnL-UAA) contained one intron, while another two genes (clpP and ycf3) had two introns (Figure S1). The functions of the major genes were broadly divided into three categories: photosynthesis-related genes, self-replication-related genes, and other genes. Photosynthesis- and self-replication-related genes constituted most of the chloroplast genome (Table 2).

3.2. SSR Analysis and Repeat Sequence Statistics

According to MISA analysis, a total of 38 SSRs loci were detected in the V. japonicum chloroplast genome. Among them, there were 36 mononucleotide loci and two dinucleotide loci, and no other types of SSR were found. Among the mononucleotide loci, the most frequent simple repeat sequence was A/T, accounting for 97.22%, while only one locus was composed of C/G. As for the dinucleotide loci, they were all composed of AT/AT, indicating a preference for base composition in SSR genes (Figure 3).
In the V. japonicum chloroplast genome, a total of 44 dispersed repeats were detected, involving 3 types, including 18 of F type, 23 of P type, and 3 of R type. However, the C type was not found (Figure 4). The length of these dispersed repeats ranged from 30 to 64 base pairs, with the majority of longer repeats found in the ycf gene and the intergenic spacer (IGS).

3.3. Relative Codon Preference Analysis

In the chloroplast genome of V. japonicum, there are a total of 26,248 codons, encoding 21 amino acids. In terms of the number of codons used, the codons encoding leucine (Leu), arginine (Arg), and serine (Ser) are the most diverse, with six codon types each. On the other hand, the codons encoding tryptophan (Trp) are the least diverse, with only one codon (UGG). Leucine (Leu) has the highest frequency of usage (10.5%, 2746 codons), while the codons UAG and UGA, which serve as termination codons, are the least used. Among the 65 codons used to encode the 21 amino acids, 31 codons exhibit preference (RSCU > 1), and only one codon for Trp does not show codon preference (UGG). Additionally, 33 codons encoding amino acids are used with lower frequency (RSCU < 1) (Figure 5 and Table 3).

3.4. Comparative Genomic Analysis

To analyze whether there is a contraction or expansion phenomenon in the boundaries of the chloroplast genome of V. japonicum, we compared the distribution of boundary genes in seven species of the Viburnum genus, including V. japonicum. The overall gene differences among the six species of the Viburnum genus are small, around 158,000. The JLB boundary region is relatively conserved among the six Adoxaceae species, falling within the rps19 gene. The JSB boundary region of V. opulus, V. fordiae, and V. japonicum is located between the trnH and ndhF genes, while the other two species of Viburnum do not have the ndhF gene in the IRb and SSC regions. The JSA boundary falls within the ycf1 gene, with the ycf1 gene of V. japonicum spanning 5036 bp in the IRa region, while the ycf1 genes of the other five species span over 5600 bp in the IRa region. The bases of the ycf1 gene in the IRa region are relatively longer in length. The JLA boundary region falls between the rpl2 and trnH genes. Except for V. japonicum, the JLA boundary of the other six Viburnum species is relatively close to the trnH gene, with distances of 6, 0, 2, 0, 3, and 0 bp, respectively. However, the JLA boundary of V. fordiae is 88 bp away from the trnH gene (Figure 6).
DnaSP software v5.0 was used to analyze nucleotide polymorphism in the chloroplast genomes of V. japonicum and its closely related species. The results showed that their nucleotide variability (π) ranged from 0 to 0.016, with an average value of 0.003. Among them, the psbH, rps19, and trnL regions exhibited relatively high variation, with all π values exceeding 0.01. The highest polymorphism was observed in the trnL region located in the IR region (π = 0.016), while the other two regions originated from the LSC region (Figure 7).

3.5. Phylogenetic Analysis of Viburnum Species

The ML tree showed that all the Viburnum species clustered into two clades (Clade A and Clade B) by high bootstrap support. The two accessions of V. japonicum were a sister group to a small clade of three Viburnum species including V. setigerum, V. erosum, and V. fordiae, and were placed in a monophyletic group with two other Viburnum species (V. luzonicum and V. melanocarpum) in Clade A, while the remaining six Viburnum species were placed in a sister Clade B (Figure 8).

4. Discussion

Using Illumina high-throughput sequencing technology, we obtained the complete chloroplast sequence of V. japonicum. The results showed that the complete sequence length of the chloroplast genome of V. japonicum (158,606 bp) closely resembles recently reported data for other Viburnum members, such as V. erosum (158,624 bp) [12], and exhibits near-identity with an individual from South Korea (158,614 bp) [10]. The complete chloroplast genome displayed a distinct quadripartite structure with a conserved arrangement within the chloroplast genome of V. japonicum, showing adherence to the typical length range (120–160 kb) observed in angiosperms [26] as well as the total GC content found in Adoxaceae [27]. These findings provide compelling evidence for the highly conserved and slow evolution of the chloroplast genome in this genus. This inference is further validated by thorough analyses of contraction and expansion at the inverted repeat region boundaries (Figure 6) as well as comprehensive whole-genome alignments. When comparing with the article published by South Korean scholars Cho et al. [10], this study supplements and improves the analysis conducted in their previous article. Our results had a more comprehensive analysis in terms of repeat sequence, SSR analysis, RCSU analysis, IR boundaries, and sliding window analysis. As a result, a more comprehensive understanding of their chloroplast genome has been achieved. Compared to the Internal Transcribed Spacer (ITS), chloroplast genes, and intergenic spacer regions, the chloroplast genome provides a greater wealth of informative data.
The number and types of SSRs are species-dependent and can be utilized for species identification, phylogenetic analysis, and molecular-assisted breeding [28]. In this study, 38 SSR loci were identified from the chloroplast genes of V. japonicum. These loci primarily consisted of the A/T form, which were comparatively fewer in number than in other Viburnum species, such as V. henryi and V. nervoum [27]. In the arrangement of the entire SSR, A and T bases occurred significantly more frequently than G and C, aligning with the higher proportion of A/T in the complete genome. In the genome, AT base pairs are less stable than GC base pairs [29], thus SSRs with more A/T repeats could lead to more mutation sites. Consequently, these SSR loci can be further developed as molecular markers for genetic variation and species identification purposes within the Viburnum genus and even the Caprifoliaceae family. In addition, V. japonicum exhibits three types of SSR repeats and lacks the C-type, which is similar to species from the same genus. The biased usage of codons is a result of the long-term adaptive evolution of species to the environment and is influenced by multiple factors. The RSCU value reflects the codon usage pattern of different genes, with higher values indicating higher codon usage frequency. Among the 31 high-frequency codons, those ending with A/U accounted for 93.55%, consistent with the codon usage bias in other plant chloroplast genomes, suggesting the presence of asymmetry in chloroplast gene codons and a preference for A/U bases.
Comparative analysis of the complete chloroplast (cp) genome sequences has proven to be a reliable tool for supporting the study of plant phylogeny and taxonomy [30]. Korean botanists have proposed a close relationship between V. japonicum and V. utile [10], while others have suggested a sister relationship with V. erosum [11,12,31]. However, the inconsistencies and limitations in these research conclusions can be attributed to insufficient sampling of species within the Viburnum genus. To clarify the phylogenetic relationships of V. japonicum with other Viburnum species, this study constructed a phylogenetic tree using published chloroplast genome data of 15 Viburnum species, with three species serving as outgroups. Based on the support values of the phylogenetic tree, our study divided the Viburnum genus into two clades. V. japonicum was placed in Clade A and exhibited a relatively close evolutionary relationship with V. setigerum, V. erosum, and V. fordiae. The results above indicate that despite a complete cp genome, species in young evolutionary lineages may still be indistinguishable. Therefore, the conclusions of the phylogenetic analysis should take into account specific characteristics of the nuclear genome.

5. Conclusions

In this study, we sequenced, characterized, and analyzed the second complete chloroplast genome sequence of V. japonicum. In conjunction with the published chloroplast genome sequences of other Viburnum species, we conducted comparative genomic analysis and constructed a phylogenetic tree to elucidate the systematic relationships between V. japonicum and other species within the genus. This study aimed to provide more comprehensive and in-depth information. It establishes a foundation for species identification in the genus Viburnum and may also contribute to the development of new conservation and management strategies to support efforts in species conservation.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/f14091819/s1, Figure S1: The structures of 14 cis-splicing genes of the V. japonicum plastome; Figure S2: The trans-splicing gene of the V. japonicum plastome; Figure S3: The read coverage depth map of V. japonicum.

Author Contributions

H.Z. and J.L. collected the sample. H.Z. analyzed data and wrote the manuscript. H.L., M.G. and C.Y. designed the study and revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Zhejiang Provincial Scientific Research Institute special project (Grant nos. 2022F1068-3 and 2021F1065-6).

Data Availability Statement

The genome sequence data that support the findings of this study are openly available in GenBank of NCBI at https://www.ncbi.nlm.nih.gov under the accession no. OP644292. The associated BioProject, SRA, and Bio-Sample numbers are PRJNA916291, SRS16920082, and SAMN32422573, respectively.

Acknowledgments

We thank Shanghai Personalbio Technology Co., Ltd. (Shanghai, China) and Genepioneer Biotechnologies Co., Ltd. (Nanjing, China) for providing technical assistance and valuable tools for data analysis and visualization.

Conflicts of Interest

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

References

  1. Ohigashi, H.; Koshimizu, K. Chavicol, as a Larva-growth inhibitor, from Viburnum japonicum spreng. Agr. Biol. Chem. 1976, 40, 2283–2287. [Google Scholar] [CrossRef]
  2. Hong, H.H.; Im, H.T. Viburnum japonicum (Caprifoliaceae): An unrecorded species in Korea. Korean J. Plant Taxon. 2003, 33, 271–277. [Google Scholar] [CrossRef]
  3. Li, R.J.; Yue, C.L.; Li, H.P.; Wang, J.; Guo, L.; Chen, X.Z.; Yang, L. Effects of drought stress on the physiological and biochemical characteristics of Viburnum japonicum seedlings. J. Northwest For. Univ. 2018, 33, 56–61. [Google Scholar]
  4. Xu, Q.R.; Gu, J.C.; Li, H.P.; Wang, J.; Fang, Y.Y.; Zhu, H.; Yue, C.L. Effect of salt spray stress on photosynthetic physiology of Viburnum japonicum with extremely small population. J. Trop. Subtrop. Bot. 2023, 31, 241–248. [Google Scholar]
  5. Li, R.J.; Chen, X.Z.; Yue, C.L.; Li, H.P.; Wang, J.; Guo, L.; Yang, L. Effects of drought stress on the photosynthetic characteristics of Viburnum japonicum seedlings. ACTA Ecol. Sin. 2018, 38, 2041–2047. [Google Scholar]
  6. Chen, X.Z.; Zhu, C.F.; Diao, S.; Cai, Y.K.; Lv, H.F. Effects of different harvesting times and storage methods on the seedling emergence of Viburnum japonicum. J. Changshu Inst. Technol. Nat. Sci. 2020, 34, 97–101. [Google Scholar]
  7. Wang, B.D.; Chen, X.Z.; Li, H.P.; Yang, L.; Guo, L.; Wang, J.; Yue, C.L. Study on softwood cutting propagation of Viburnum japonicum. J. Zhejiang Forest. Sci. Technol. 2018, 38, 9–15. [Google Scholar]
  8. Liu, J.; Yue, C.L.; Zhu, H.; Li, H.P.; Li, H.B. Analysis of SSR loci and development of polymorphic SSR makers in Viburnum japonicum transcriptome. Mol. Plant Breed. 2022, 20, 7183–7192. [Google Scholar]
  9. Zhu, H.; Liu, J.; Gao, M.; Yue, C.; Li, H. Population genetic assessment of Viburnum japonicum in China using ddRAD-seq. Front. Genet. 2023, 14, 1150437. [Google Scholar] [CrossRef]
  10. Cho, W.B.; Han, E.K.; Choi, H.J.; Lee, J.H. The complete chloroplast genome sequence of Viburnum Japonicum (Adoxaceae), an evergreen broad-leaved shrub. Mitochondrial DNA B Resour. 2018, 3, 458–459. [Google Scholar] [CrossRef]
  11. Choi, Y.G.; Yun, N.; Park, J.; Xi, H.; Min, J.; Kim, Y.; Oh, S.H. The second complete chloroplast genome sequence of the Viburnum erosum (Adoxaceae) showed a low level of intra-species variations. Mitochondrial DNA B Resour. 2020, 5, 271–272. [Google Scholar] [CrossRef]
  12. Park, J.; Choi, Y.G.; Yun, N.; Xi, H.; Min, J.; Kim, Y.; Oh, S.H. The complete chloroplast genome sequence of Viburnum erosum (Adoxaceae). Mitochondrial DNA B Resour. 2019, 4, 3278–3279. [Google Scholar] [CrossRef] [PubMed]
  13. Celiński, K.; Sokołowska, J.; Fuchs, H.; Maděra, P.; Wiland-Szymańska, J. Characterization of the complete chloroplast genome sequence of the socotra dragon‘s blood tree (Dracaena cinnabari balf.). Forests 2022, 13, 932. [Google Scholar] [CrossRef]
  14. Tao, L.; Duan, H.; Tao, K.; Luo, Y.; Li, Q.; Li, L. Complete chloroplast genome structural characterization of two Phalaenopsis (Orchidaceae) species and comparative analysis with their alliance. BMC Genom. 2023, 24, 359. [Google Scholar] [CrossRef]
  15. Luo, C.; Huang, W.; Yer, H.; Kamuda, T.; Li, X.; Li, Y.; Rong, Y.; Yan, B.; Wen, Y.; Wang, Q.; et al. Complete chloroplast genomes and comparative analyses of three ornamental impatiens species. Front. Genet. 2022, 13, 816123. [Google Scholar] [CrossRef]
  16. Alawfi, M.S.; Alzahrani, D.A.; Albokhari, E.J. Complete chloroplast genome sequences of two Ehretia trees (Ehretia cymosa and Ehretia obtusifolia): Genome structures and phylogenetic analysis. Forests 2023, 14, 1486. [Google Scholar] [CrossRef]
  17. Yang, J.B.; Li, D.Z.; Li, H.T. Highly effective sequencing whole chloroplast genomes of angiosperms by nine novel Uni-versal primer pairs. Mol. Ecol. Resour. 2014, 14, 1024–1031. [Google Scholar]
  18. Andrews, S. Babraham Bioinformatics FastQC: A Quality Control Tool for High Throughput Sequence Data. Available online: http://www.bioinformatics.babraham.ac.uk/projects/fastqc (accessed on 2 August 2023).
  19. Mckain, M.R.; Wilson, M. Fast-Plast: Rapid De Novo Assembly and Finishing for Whole Chloroplast Genomes. 2017. Available online: https://github.com/mrmckain/Fast-Plast (accessed on 2 August 2023).
  20. Liu, S.; Ni, Y.; Li, J.; Zhang, X.; Yang, H.; Chen, H.; Liu, C. CPGView: A package for visualizing detailed chloroplast genome structures. Mol. Ecol. Resour. 2023, 23, 694–704. [Google Scholar] [CrossRef]
  21. 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]
  22. Sharp, M.; Tuohy, T.M.F.; Mosurski, K.R. Codon usage in yeast: Cluster analysis dearly differentiates highly and lowly expressed genes. Nucleic Acids Res. 1986, 14, 5125–5143. [Google Scholar] [CrossRef]
  23. Katoh, K.; Rozewicki, J.; Yamada, K.D. MAFFT online service: Multiple sequence alignment, interactive sequence choice and visualization. Brief. Bioinform. 2019, 20, 1160–1166. [Google Scholar] [CrossRef]
  24. Librado, P.; Rozas, J. DnaSP v5: A software for comprehensive analysis of DNA polymorphism data. Bioinformatics 2009, 25, 1451–1452. [Google Scholar] [CrossRef] [PubMed]
  25. Stamatakis, A. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 2014, 30, 1312–1313. [Google Scholar] [CrossRef] [PubMed]
  26. 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]
  27. Ran, H.; Liu, Y.; Wu, C.; Cao, Y. Phylogenetic and comparative analyses of complete chloroplast genomes of Chinese Viburnum and Sambucus (Adoxaceae). Plants 2020, 9, 1143. [Google Scholar] [CrossRef]
  28. Yang, A.H.; Zhang, J.J.; Yao, X.H.; Huang, H.W. Chloroplast microsatellite markers in Liriodendron tulipifera (Magnoliaceae) and cross-species amplification in L. chinense. Am. J. Bot. 2011, 98, e123–e126. [Google Scholar] [CrossRef]
  29. Wang, A.H.J.; Hakoshima, T.; van der Marel, G.; van Boom, J.H.; Rich, A. AT base pairs are less stable than GC base pairs in Z-DNA: The crystal structure of d (m5CGTAm5CG). Cell 1984, 37, 321–331. [Google Scholar] [CrossRef]
  30. Li, Y.; Sylvester, S.P.; Li, M.; Zhang, C.; Li, X.; Duan, Y.; Wang, X. The complete plastid genome of Magnolia zenii and genetic comparison to Magnoliaceae species. Molecules 2019, 24, 261. [Google Scholar] [CrossRef]
  31. Gu, L.; Wang, G.; Weng, Q. Characterization of the complete chloroplast genome of China Viburnum burejaeticum Regel et Herd and intra-species diversity. Mitochondrial DNA B Resour. 2021, 6, 1353–1354. [Google Scholar] [CrossRef]
Figure 1. Plant morphological characteristics of V. japonicum (Thunb.) Sprengel. (A) Front and back views of fresh leaves of this species attached with a scale bar. (B) Tiny, white, and strongly scented flowers produced a compound umbel-like inflorescence. (C) Young infructescence in late spring. (D) Mature infructescence with bright red appearance. (E) Understory seedlings. Plant photos by Dr. Hong Zhu without any copyright issues.
Figure 1. Plant morphological characteristics of V. japonicum (Thunb.) Sprengel. (A) Front and back views of fresh leaves of this species attached with a scale bar. (B) Tiny, white, and strongly scented flowers produced a compound umbel-like inflorescence. (C) Young infructescence in late spring. (D) Mature infructescence with bright red appearance. (E) Understory seedlings. Plant photos by Dr. Hong Zhu without any copyright issues.
Forests 14 01819 g001
Figure 2. Chloroplast genome map of V. japonicum. The genes shown outside of the circle are transcribed clockwise, while those inside are transcribed anticlockwise. Genes belonging to different functional groups are color-coded. The dashed area in the inner circle indicates the GC content of the cp genome, and the lighter gray and darker gray in the inner circle indicate the A + T and G + C contents of the genome. LSC: large single-copy; SSC: small single-copy; IR: inverted repeat.
Figure 2. Chloroplast genome map of V. japonicum. The genes shown outside of the circle are transcribed clockwise, while those inside are transcribed anticlockwise. Genes belonging to different functional groups are color-coded. The dashed area in the inner circle indicates the GC content of the cp genome, and the lighter gray and darker gray in the inner circle indicate the A + T and G + C contents of the genome. LSC: large single-copy; SSC: small single-copy; IR: inverted repeat.
Forests 14 01819 g002
Figure 3. SSR type and numbers of chloroplast genome in V. japonicum.
Figure 3. SSR type and numbers of chloroplast genome in V. japonicum.
Forests 14 01819 g003
Figure 4. Type and number of long repeats in V. japonicum chloroplast genome. F: Forward repeats; P: palindromic repeats; R: reverse repeats; C: complement repeats.
Figure 4. Type and number of long repeats in V. japonicum chloroplast genome. F: Forward repeats; P: palindromic repeats; R: reverse repeats; C: complement repeats.
Forests 14 01819 g004
Figure 5. Relative synonymous codon usage (RSCU) analysis of the chloroplast genome of V. japonicum.
Figure 5. Relative synonymous codon usage (RSCU) analysis of the chloroplast genome of V. japonicum.
Forests 14 01819 g005
Figure 6. Comparison of IR boundaries among Viburnum species.
Figure 6. Comparison of IR boundaries among Viburnum species.
Forests 14 01819 g006
Figure 7. Sliding window analysis for the nucleotide diversity (π) of V. japonicum.
Figure 7. Sliding window analysis for the nucleotide diversity (π) of V. japonicum.
Forests 14 01819 g007
Figure 8. Phylogenetic relationship between newly sequenced V. japonicum and other 17 representative species within the family Adoxaceae based on complete cp genome analysis. The tree was constructed using the maximum likelihood (ML) method by RAxML v8.2.12, with bootstrap support values indicated on each branch. Dipsacus japonicus MZ934745, Sambucus chinensis MW455170, and S. williamsii KX510276 were chosen as the outgroup. The newly identified genome of V. japonicum was represented in red bold. The GenBank accession number was listed after the species name.
Figure 8. Phylogenetic relationship between newly sequenced V. japonicum and other 17 representative species within the family Adoxaceae based on complete cp genome analysis. The tree was constructed using the maximum likelihood (ML) method by RAxML v8.2.12, with bootstrap support values indicated on each branch. Dipsacus japonicus MZ934745, Sambucus chinensis MW455170, and S. williamsii KX510276 were chosen as the outgroup. The newly identified genome of V. japonicum was represented in red bold. The GenBank accession number was listed after the species name.
Forests 14 01819 g008
Table 1. Base composition of the chloroplast genome of V. japonicum (Thunb.) Sprengel.
Table 1. Base composition of the chloroplast genome of V. japonicum (Thunb.) Sprengel.
RegionPercentage of Bases
AT (%)GC (%)A (%)T (%)G (%)C (%)
Total61.9138.0830.6231.2918.7119.38
IRa57.0342.9728.6128.4222.2520.72
IRb57.0342.9728.4228.6120.7222.25
SSC69.0331.9733.9734.0615.2216.75
LSC63.5836.4231.1832.4017.7618.66
Table 2. Annotation table of functional genes in the chloroplast genome of V. japonicum.
Table 2. Annotation table of functional genes in the chloroplast genome of V. japonicum.
CategoriesGroup of GenesName of Genes
PhotosynthesisSubunits of photosystem IpsaA, psaB, psaC, psaI, psaJ
Subunits of photosystem IIpsbA, psbB, psbC, psbD, psbE, psbF, psbH, psbI, psbJ, psbK, psbL, psbM, psbN, psbT, psbZ
Subunits of NADH dehydrogenasendhA *, ndhB * (2), ndhC, ndhD, ndhE, ndhF, ndhG, ndhH, ndhI, ndhJ, ndhK
Subunits of cytochrome b/f complexpetA, petB *, petD *, petG, petL, petN
Subunits of ATP synthaseatpA, atpB, atpE, atpF *, atpH, atpI
Large subunit of rubiscorbcL
Self-replicationProteins of large ribosomal subunitrpl14, rpl16 *, rpl2 * (2), rpl20, rpl22, rpl23 (2), rpl32, rpl33, rpl36
Proteins of small ribosomal subunitrps11, rps12 ** (2), rps14, rps15, rps16 *, rps18, rps19, rps2, rps3, rps4, rps7 (2), rps8
Subunits of RNA polymeraserpoA, rpoB, rpoC1 *, rpoC2
Ribosomal RNAsrrn16 (2), rrn23 (2), rrn4.5 (2), rrn5 (2)
Transfer RNAstrnA-UGC * (2), trnC-GCA, trnD-GUC, trnE-UUC, trnF-GAA, trnG-GCC, trnH-GUG, trnI * (2), trnK-UUU *, trnL-CAA (2), trnL-UAA *, trnL-UAG, trnM-CAU (4), trnN-GUU (2), trnP-UGG, trnQ-UUG, trnR-ACG (2), trnR-UCU, trnS-CGA *, trnS-GCU, trnS-GGA, trnS-UGA, trnT-GGU, trnT-UGU, trnV-GAC (2), trnW-CCA, trnY-GUA
Other genesMaturasematK
ProteaseclpP **
Envelope membrane proteincemA
Acetyl-CoA carboxylaseaccD *
c-type cytochrome synthesis geneccsA
Translation initiation factorinfA
Genes of unknown functionConserved hypothetical chloroplast ORFycf1, ycf15 (2), ycf2 (2), ycf3 **, ycf4
* Genes containing one intron, ** genes containing two introns, (2): gene with two copies.
Table 3. Relative synonymous codon usage (RSCU) in V. japonicum chloroplast genome.
Table 3. Relative synonymous codon usage (RSCU) in V. japonicum chloroplast genome.
Amino AcidCodonNumberRSCUAmino AcidCodonNumberRSCU
Termination codonUAA401.40MetAUG6192.00
UAG230.80GUG10.00
UGA230.80AsnAAC2840.45
AlaGCA3901.10AAU9681.55
GCC2230.63ProCCA3271.19
GCG1610.46CCC2080.76
GCU6391.81CCG1530.56
CysUGC760.52CCU4111.50
UGU2151.48GlnCAA7051.50
AspGAC2090.38CAG2320.50
GAU8941.62ArgAGA4841.81
GluGAA10271.49AGG1850.69
GAG3540.51CGA3641.36
PheUUC5430.74CGC1070.40
UUU9161.26CGG1220.46
GlyGGA7041.57CGU3411.28
GGC1960.44SerAGC1290.38
GGG3280.73AGU4001.18
GGU5651.26UCA4161.22
HisCAC1380.43UCC3260.96
CAU5031.57UCG1840.54
IleAUA6850.93UCU5841.72
AUC4770.65ThrACA4021.20
AUU10531.43ACC2530.76
LysAAA10281.48ACG1520.45
AAG3570.52ACU5311.59
LeuCUA3800.83ValGUA5231.48
CUC1950.43GUC1780.50
CUG1750.38GUG2140.60
CUU5971.30GUU5001.41
UUA8111.77TrpUGG4681.00
UUG5881.28TyrUAC2030.42
UAU7611.58
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhu, H.; Liu, J.; Li, H.; Yue, C.; Gao, M. Complete Chloroplast Genome Structural Characterization and Comparative Analysis of Viburnum japonicum (Adoxaceae). Forests 2023, 14, 1819. https://doi.org/10.3390/f14091819

AMA Style

Zhu H, Liu J, Li H, Yue C, Gao M. Complete Chloroplast Genome Structural Characterization and Comparative Analysis of Viburnum japonicum (Adoxaceae). Forests. 2023; 14(9):1819. https://doi.org/10.3390/f14091819

Chicago/Turabian Style

Zhu, Hong, Juan Liu, Hepeng Li, Chunlei Yue, and Meirong Gao. 2023. "Complete Chloroplast Genome Structural Characterization and Comparative Analysis of Viburnum japonicum (Adoxaceae)" Forests 14, no. 9: 1819. https://doi.org/10.3390/f14091819

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