Search
2022 Volume 2
Article Contents
ARTICLE   Open Access    

Chromosome-level genome assembly and annotation of the native Chinese wild blueberry Vaccinium bracteatum

  • # These authors contributed equally: Lu Yang, Minghui Li

More Information
  • Vaccinium bracteatum Thunb., an important native Chinese wild blueberry species, is widely used as a rootstock and in blueberry cultivar breeding, as well as in traditional medicine and local food. We report here the genome sequence of V. bracteatum using a combination of Oxford Nanopore Technologies long-read and Illumina HiSeq short-read sequencing technologies to obtain 65.30 Gb of clean data, achieving 114.60-fold genome coverage. The assembled genome has a total sequence length of 569.81 Mb and consists of 36,756 predicted genes. Repetitive DNA sequences represent 57.78% of the genome sequence. Comparative genomic analysis revealed that a total of 336 gene families had expanded and that 298 candidate genes had undergone positive selection during evolution in V. bracteatum. The divergence of V. bracteatum from the related Rhododendron williamsianum and Rhododendron delavayi occurred ~13−85 million years ago. The genome sequence of V. bracteatum allowed us to identify some important genes associated with traits involved in fruit development, such as flavonoid biosynthesis, sugar and acid metabolism, MYB transcription factor gene expression, and hormone regulation. The differential expression patterns of genes encoding flavonoid biosynthesis enzymes and MYB transcription factors might explain the high flavonoid content of V. bracteatum. This chromosome-level genome assembly provides reference sequences for the identification and characterization of genes important in the improvement of blueberry and related research.
  • 加载中
  • Supplementary Table S1 Sequencing data used for genome assembly and annotation of V. bracteatum.
    Supplementary Table S2 Raw data and clean data statistics.
    Supplementary Table S3 Genome completeness evaluation based on Reads alignment, CEGs and BUSCOs.
    Supplementary Table S4 Statistics of Hi-C assembly data.
    Supplementary Table S5 Statistics of repeat sequences.
    Supplementary Table S6 Statistical results of coding gene quantity prediction based on Ab initio, Homology-based and RNA-Seq.
    Supplementary Table S7 Prediction results of coding genes after EVM integration.
    Supplementary Table S8 Noncoding RNA statistics and pseudogene prediction.
    Supplementary Table S9 Gene function annotation statistics.
    Supplementary Table S10 Gene family clustering statistical information of V. bracteatum and other six species.
    Supplementary Table S11 Enrichment analysis of the gene family of Vaccinium bracteatum.
    Supplementary Table S12 Annotation of expansion and contraction gene families among seven species.
    Supplementary Table S13 A total of 298 candidate genes with positive selection in Vaccinium bracteatum.
    Supplementary Table S14 Annotation of positively selected genes in Vaccinium bracteatum.
    Supplementary Table S15 Number statistics of positively selected genes in terms by GO analysis.
    Supplementary Table S16 Annotation of positively selected genes with KEGG analysis.
    Supplementary Fig. S1 K-mer distribution map The k-mer analysis (k=19) of Vaccinium bracteatum genome characteristics.
    Supplementary Fig. S2 Interactive thermogram of Hi-C assembly chromosome Note: LG01-LG12 stands for Lachesis group 01-12; abscissa and ordinate represent the order of each bin on the corresponding chromosome group.
    Supplementary Fig. S3 Distribution of integrated genes derived from three prediction methods.
    Supplementary Figs S4 Statistical chart of functional annotation classification of KOG Note: abscissa is the content of each KOG classification, and the ordinate is the number of genes.
    Supplementary Figs S5 KEGG path annotation diagram.
    Supplementary Figs S6 GO secondary node annotation classification statistical chart Note: The left side of the ordinate is the percentage of gene number, and the right side is the number of genes.
    Supplementary Fig. S7 GO and KEGG enrichment analysis of Vaccinium bracteatum specific gene family.
    Supplementary Fig. S8 The Ks distribution map within and between species.
    Supplementary Fig. S9 Principle component analysis of 12 samples to assess data quality.
    Supplementary Fig. S10 Number of significantly downregulated (down) and upregulated (up) genes in the two comparisons.
    Supplementary Fig. S11 Gene ontology (GO) analysis of Differentially expressed genes (DEGs) in the comparisons of pink fruit vs. green fruit (a) , and blue fruit vs. pink fruit (b) .
    Supplementary Fig. S12 Heat map diagram of the expression of differently expressed genes (DEGs) related to carbohydrate metabolism in fruit at three stages.
  • [1]

    Xu Y, Fan M, Zhou S, Wang L, Qian H, et al. 2017. Effect of Vaccinium bracteatum Thunb. leaf pigment on the thermal, pasting, and textural properties and microstructure characterization of rice starch. Food chemistry 228:435−40

    doi: 10.1016/j.foodchem.2017.02.041

    CrossRef   Google Scholar

    [2]

    Wang L, Zhang Y, Xu M, Wang Y, Cheng S, et al. 2013. Anti-diabetic activity of Vaccinium bracteatum Thunb. leaves’ polysaccharide in STZ-induced diabetic mice. International journal of biological macromolecules 61:317−21

    doi: 10.1016/j.ijbiomac.2013.07.028

    CrossRef   Google Scholar

    [3]

    Wang L, Jiang T, Zhang H, Yao H. 2008. Study on the extraction of black pigment from Vaccinium bracteatum Thunb. leaves by enzyme and its stability. Science and Technology of Food Industry 29:224−226+258

    doi: 10.13386/j.issn1002-0306.2008.10.055

    CrossRef   Google Scholar

    [4]

    Fan M, Lian W, Li T, Fan Y, Rao Z, et al. 2020. Metabolomics approach reveals discriminatory metabolites associating with the blue pigments from Vaccinium bracteatum thunb. leaves at different growth stages. Industrial Crops and Products 147:112252

    doi: 10.1016/j.indcrop.2020.112252

    CrossRef   Google Scholar

    [5]

    Zhang J, Chu C, Li X, Yao S, Yan B, et al. 2014. Isolation and identification of antioxidant compounds in Vaccinium bracteatum Thunb. by UHPLC-Q-TOF LC/MS and their kidney damage protection. Journal of Functional Foods 11:62−70

    doi: 10.1016/j.jff.2014.09.005

    CrossRef   Google Scholar

    [6]

    Ren Y, Ke C, Tang C, Yao S, Ye Y. 2017. Divaccinosides A–D, four rare iridoid glucosidic truxillate esters from the leaves of Vaccinium bracteatum. Tetrahedron Letters 58(24):2385−8

    doi: 10.1016/j.tetlet.2017.05.013

    CrossRef   Google Scholar

    [7]

    Zhao J, Wu Y, Niu X, Zhang Y, Xu X, et al. 2017. Content determination of vaccinoside in leaves of Vaccinium bracteatum Thunb. by HPLC. Shanghai Journal of Traditional Chinese Medicine 51:100−2

    doi: 10.16305/j.1007-1334.2017.10.027

    CrossRef   Google Scholar

    [8]

    Fan M, Li T, Li Y, Qian H, Zhang H, et al. 2021. Vaccinium bracteatum Thunb. as a promising resource of bioactive compounds with health benefits: An updated review. Food Chemistry 356:129738

    doi: 10.1016/j.foodchem.2021.129738

    CrossRef   Google Scholar

    [9]

    Polashock J, Zelzion E, Fajardo D, Zalapa J, Georgi L, et al. 2014. The American cranberry: first insights into the whole genome of a species adapted to bog habitat. BMC Plant Biology 14:165

    doi: 10.1186/1471-2229-14-165

    CrossRef   Google Scholar

    [10]

    Diaz-Garcia L, Garcia-Ortega LF, González-Rodríguez M, Delaye L, Iorizzo M, et al. 2021. Chromosome-Level Genome Assembly of the American Cranberry (Vaccinium macrocarpon Ait. ) and Its Wild Relative Vaccinium microcarpum. Frontiers in Plant Science 12:633310

    doi: 10.3389/fpls.2021.633310

    CrossRef   Google Scholar

    [11]

    Wu C, Deng C, Hilario E, Albert NW, Lafferty D, et al. 2022. A chromosome-scale assembly of the bilberry genome identifies a complex locus controlling berry anthocyanin composition. Molecular Ecology Resources 22:345−60

    doi: 10.1111/1755-0998.13467

    CrossRef   Google Scholar

    [12]

    Tsuda H, Kunitake H, Yamasaki M, Komatsu H, Yoshioka K. 2013. Production of intersectional hybrids between colchicine-induced tetraploid shashanbo (Vaccinium bracteatum) and highbush blueberry ‘Spartan’. Journal of the American Society for Horticultural Science 138:317−24

    doi: 10.21273/JASHS.138.4.317

    CrossRef   Google Scholar

    [13]

    Costich DE, Ortiz R, Meagher TR, Bruederle LP, Vorsa N. 1993. Determination of ploidy level and nuclear DNA content in blueberry by flow cytometry. Theoretical and Applied Genetics 86:1001−6

    doi: 10.1007/BF00211053

    CrossRef   Google Scholar

    [14]

    Li X, Sun H, Pei J, Dong Y, Wang F, et al. 2012. De novo sequencing and comparative analysis of the blueberry transcriptome to discover putative genes related to antioxidants. Gene 511:54−61

    doi: 10.1016/j.gene.2012.09.021

    CrossRef   Google Scholar

    [15]

    Gupta V, Estrada AD, Blakley I, Reid R, Patel K, et al. 2015. RNA-Seq analysis and annotation of a draft blueberry genome assembly identifies candidate genes involved in fruit ripening, biosynthesis of bioactive compounds, and stage-specific alternative splicing. Gigascience 4:5

    doi: 10.1186/s13742-015-0046-9

    CrossRef   Google Scholar

    [16]

    Colle M, Leisner CP, Wai CM, Ou S, Bird KA, et al. 2019. Haplotype-phased genome and evolution of phytonutrient pathways of tetraploid blueberry. GigaScience 8:giz012

    doi: 10.1093/gigascience/giz012

    CrossRef   Google Scholar

    [17]

    Yu J, Hulse-Kemp AM, Babiker E, Staton M. 2021. High-quality reference genome and annotation aids understanding of berry development for evergreen blueberry (Vaccinium darrowii). Horticulture Research 8:228

    doi: 10.1038/s41438-021-00641-9

    CrossRef   Google Scholar

    [18]

    Wang P, Luo Y, Huang J, Gao S, Zhu G, et al. 2020. The genome evolution and domestication of tropical fruit mango. Genome Biology 21:60

    doi: 10.1186/s13059-020-01959-8

    CrossRef   Google Scholar

    [19]

    Rose JP, Kleist TJ, Löfstrand SD, Drew BT, Schönenberger J, et al. 2018. Phylogeny, historical biogeography, and diversification of angiosperm order Ericales suggest ancient Neotropical and East Asian connections. Molecular Phylogenetics and Evolution 122:59−79

    doi: 10.1016/j.ympev.2018.01.014

    CrossRef   Google Scholar

    [20]

    Soza VL, Lindsley D, Waalkes A, Ramage E, Patwardhan RP, et al. 2019. The Rhododendron genome and chromosomal organization provide insight into shared whole-genome duplications across the heath family (Ericaceae). Genome Biology and Evolution 11:3353−71

    doi: 10.1093/gbe/evz245

    CrossRef   Google Scholar

    [21]

    Peng Y, Lin-Wang K, Cooney JM, Wang T, Espley RV, et al. 2019. Differential regulation of the anthocyanin profile in purple kiwifruit (Actinidia species). Horticulture Research 6:3

    doi: 10.1038/s41438-018-0076-4

    CrossRef   Google Scholar

    [22]

    Dong J, Cao L, Zhang X, Zhang W, Yang T, et al. 2021. An R2R3-MYB transcription Factor RmMYB108 responds to chilling stress of Rosa multiflora and conferred cold tolerance of Arabidopsis. Frontiers in Plant Science 12:696919

    doi: 10.3389/fpls.2021.696919

    CrossRef   Google Scholar

    [23]

    Chen Y, Yang X, Li W, Zhao S. 2020. Knockdown of the DUF647 family memberRUS4 impairs stamen development and pollen maturation in Arabidopsis. Plant Science 301:110645

    doi: 10.1016/j.plantsci.2020.110645

    CrossRef   Google Scholar

    [24]

    Cheng H, Han L, Yang C, Wu X, Zhong N, et al. 2016. The cotton MYB108 forms a positive feedback regulation loop with CML11 and participates in the defense response against Verticillium dahliae infection. Journal of Experimental Botany 67:1935−50

    doi: 10.1093/jxb/erw016

    CrossRef   Google Scholar

    [25]

    Wei Z, Hu K, Zhao D, Tang J, Huang Z, et al. 2020. MYB44 competitively inhibits the formation of the MYB340-bHLH2-NAC56 complex to regulate anthocyanin biosynthesis in purple-fleshed sweet potato. BMC Plant Biology 20:258

    doi: 10.1186/s12870-020-02451-y

    CrossRef   Google Scholar

    [26]

    El-Sharkawy I, Liang D, Xu K. 2015. Transcriptome analysis of an apple (Malus × domestica) yellow fruit somatic mutation identifies a gene network module highly associated with anthocyanin and epigenetic regulation. Journal of Experimental Botany 66:7359−76

    doi: 10.1093/jxb/erv433

    CrossRef   Google Scholar

    [27]

    Lin Q, Wang C, Dong W, Jiang Q, Wang D, et al. 2015. Transcriptome and metabolome analyses of sugar and organic acid metabolism in Ponkan (Citrus reticulata) fruit during fruit maturation. Gene 554:64−74

    doi: 10.1016/j.gene.2014.10.025

    CrossRef   Google Scholar

    [28]

    Guo S, Sun H, Zhang H, Liu J, Ren Y, et al. 2015. Comparative Transcriptome Analysis of Cultivated and Wild Watermelon during Fruit Development. PloS One 10:e0130267

    doi: 10.1371/journal.pone.0130267

    CrossRef   Google Scholar

    [29]

    Rahim MA, Robin AHK, Natarajan S, Jung HJ, Lee J, et al. 2018. Identification and Characterization of Anthocyanin Biosynthesis-Related Genes in Kohlrabi. Applied Biochemistry and Biotechnology 184:1120−41

    doi: 10.1007/s12010-017-2613-2

    CrossRef   Google Scholar

    [30]

    Li Y, Nie P, Zhang H, Wang L, Wang H, et al. 2017. Dynamic changes of anthocyanin accumulation and endogenous hormone contents in blueberry. Journal of Beijing Forestry University 39:64−71

    doi: 10.13332/j.1000-1522.20160283

    CrossRef   Google Scholar

    [31]

    Primetta AK, Karppinen K, Riihinen KR, Jaakola L. 2015. Metabolic and molecular analyses of white mutant Vaccinium berries show down-regulation of MYBPA1-type R2R3 MYB regulatory factor. Planta 242:631−43

    doi: 10.1007/s00425-015-2363-8

    CrossRef   Google Scholar

    [32]

    Lin Y, Wang Y, Li B, Tan H, Li D, et al. 2018. Comparative transcriptome analysis of genes involved in anthocyanin synthesis in blueberry. Plant Physiology and Biochemistry 127:561−72

    doi: 10.1016/j.plaphy.2018.04.034

    CrossRef   Google Scholar

    [33]

    Rogers SO, Bendich AJ. 1985. Extraction of DNA from milligram amounts of fresh, herbarium and mummified plant tissues. Plant Molecular Biology 5:69−76

    doi: 10.1007/BF00020088

    CrossRef   Google Scholar

    [34]

    Jiang S, An H, Xu F, Zhang X. 2020. Chromosome-level genome assembly and annotation of the loquat (Eriobotrya japonica) genome. GigaScience 9:giaa015

    doi: 10.1093/gigascience/giaa015

    CrossRef   Google Scholar

    [35]

    Li R, Fan W, Tian G, Zhu H, He L, et al. 2010. The sequence and de novo assembly of the giant panda genome. Nature 463:311−17

    doi: 10.1038/nature08696

    CrossRef   Google Scholar

    [36]

    Koren S, Walenz BP, Berlin K, Miller JR, Bergman NH, et al. 2017. Canu: scalable and accurate long-read assembly via adaptive k-mer weighting and repeat separation. Genome Research 27:722−36

    doi: 10.1101/gr.215087.116

    CrossRef   Google Scholar

    [37]

    Vaser R, Sović I, Nagarajan N, Šikić M. 2017. Fast and accurate de novo genome assembly from long uncorrected reads. Genome Res 27:737−46

    doi: 10.1101/gr.214270.116

    CrossRef   Google Scholar

    [38]

    Walker BJ, Abeel T, Shea T, Priest M, Abouelliel A, et al. 2014. Pilon: an integrated tool for comprehensive microbial variant detection and genome assembly improvement. PLoS One 9:e112963

    doi: 10.1371/journal.pone.0112963

    CrossRef   Google Scholar

    [39]

    Lieberman-Aiden E, van Berkum NL, Williams L, Imakaev M, Ragoczy T, et al. 2009. Comprehensive mapping of long-range interactions reveals folding principles of the human genome. Science 326:289−93

    doi: 10.1126/science.1181369

    CrossRef   Google Scholar

    [40]

    Rao SSP, Huntley MH, Durand NC, Stamenova EK, Bochkov ID, et al. 2014. A 3D map of the human genome at kilobase resolution reveals principles of chromatin looping. Cell 159:1665−80

    doi: 10.1016/j.cell.2014.11.021

    CrossRef   Google Scholar

    [41]

    Burton JN, Adey A, Patwardhan RP, Qiu R, Kitzman JO, et al. 2013. Chromosome-scale scaffolding of de novo genome assemblies based on chromatin interactions. Nature Biotechnology 31:1119−25

    doi: 10.1038/nbt.2727

    CrossRef   Google Scholar

    [42]

    Li H, Durbin R. 2009. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics 25:1754−60

    doi: 10.1093/bioinformatics/btp324

    CrossRef   Google Scholar

    [43]

    Parra G, Bradnam K, Korf I. 2007. CEGMA: a pipeline to accurately annotate core genes in eukaryotic genomes. Bioinformatics 23:1061−67

    doi: 10.1093/bioinformatics/btm071

    CrossRef   Google Scholar

    [44]

    Simão FA, Waterhouse RM, Ioannidis P, Kriventseva EV, Zdobnov EM. 2015. BUSCO: assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics 31:3210−12

    doi: 10.1093/bioinformatics/btv351

    CrossRef   Google Scholar

    [45]

    Price AL, Jones NC, Pevzner PA. 2005. De novo identification of repeat families in large genomes. Bioinformatics 21:i351−i358

    doi: 10.1093/bioinformatics/bti1018

    CrossRef   Google Scholar

    [46]

    Xu Z, Wang H. 2007. LTR_FINDER: an efficient tool for the prediction of full-length LTR retrotransposons. Nucleic Acids Research 35:W265−W268

    doi: 10.1093/nar/gkm286

    CrossRef   Google Scholar

    [47]

    Hoede C, Arnoux S, Moisset M, Chaumier T, Inizan O, et al. 2014. PASTEC: an automatic transposable element classification tool. PloS One 9:e91929

    doi: 10.1371/journal.pone.0091929

    CrossRef   Google Scholar

    [48]

    Jurka J, Kapitonov VV, Pavlicek A, Klonowski P, Kohany O, et al. 2005. Repbase Update, a database of eukaryotic repetitive elements. Cytogenetic and Genome Research 110:462−7

    doi: 10.1159/000084979

    CrossRef   Google Scholar

    [49]

    Tarailo-Graovac M, Chen N. 2009. Using RepeatMasker to identify repetitive elements in genomic sequences. Current Protocols in Bioinformatics 25:4.10.1−4.10.14

    doi: 10.1002/0471250953.bi0410s25

    CrossRef   Google Scholar

    [50]

    Lowe TM, Eddy SR. 1997. tRNAscan-SE: a program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Research 25:955−64

    doi: 10.1093/nar/25.5.955

    CrossRef   Google Scholar

    [51]

    Kent WJ. 2002. BLAT — the BLAST-like alignment tool. Genome Research 12:656−64

    doi: 10.1101/gr.229202

    CrossRef   Google Scholar

    [52]

    She R, Chu JS, Wang K, Pei J, Chen N. 2009. GenBlastA: enabling BLAST to identify homologous gene sequences. Genome Research 19:143−49

    doi: 10.1101/gr.082081.108

    CrossRef   Google Scholar

    [53]

    Birney E, Clamp M, Durbin R. 2004. GeneWise and Genomewise. Genome Research 14:988−95

    doi: 10.1101/gr.1865504

    CrossRef   Google Scholar

    [54]

    Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. 1990. Basic local alignment search tool. Journal of Molecular Biology 215:403−10

    doi: 10.1016/S0022-2836(05)80360-2

    CrossRef   Google Scholar

    [55]

    Dimmer EC, Huntley RP, Alam-Faruque Y, Sawford T, O'Donovan C, et al. 2012. The UniProt-GO Annotation database in 2011. Nucleic Acids Research 40:D565−D570

    doi: 10.1093/nar/gkr1048

    CrossRef   Google Scholar

    [56]

    Emms DM, Kelly S. 2019. OrthoFinder: phylogenetic orthology inference for comparative genomics. Genome Biology 20:238

    doi: 10.1186/s13059-019-1832-y

    CrossRef   Google Scholar

    [57]

    Mi H, Muruganujan A, Ebert D, Huang X, Thomas PD. 2019. PANTHER version 14: more genomes, a new PANTHER GO-slim and improvements in enrichment analysis tools. Nucleic Acids Research 47:D419−D426

    doi: 10.1093/nar/gky1038

    CrossRef   Google Scholar

    [58]

    Nguyen LT, Schmidt HA, von Haeseler A, Minh BQ. 2015. IQ-TREE: a fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Molecular Biology and Evolution 32:268−74

    doi: 10.1093/molbev/msu300

    CrossRef   Google Scholar

    [59]

    Katoh K, Asimenos G, Toh H. 2009. Multiple alignment of DNA sequences with MAFFT. In Bioinformatics for DNA Sequence Analysis. Methods in Molecular Biology, eds. Posada D. (eds) vol. 537: XIV, 354. New York: Humana Press. pp. 39−64 https://doi.org/10.1007/978-1-59745-251-9_3

    [60]

    Talavera G, Castresana J. 2007. Improvement of phylogenies after removing divergent and ambiguously aligned blocks from protein sequence alignments. Systematic Biology 56:564−77

    doi: 10.1080/10635150701472164

    CrossRef   Google Scholar

    [61]

    Kalyaanamoorthy S, Minh BQ, Wong TKF, von Haeseler A, Jermiin LS. 2017. ModelFinder: fast model selection for accurate phylogenetic estimates. Nature Methods 14:587−89

    doi: 10.1038/nmeth.4285

    CrossRef   Google Scholar

    [62]

    Yang Z. 1997. PAML: a program package for phylogenetic analysis by maximum likelihood. Bioinformatics 13:555−56

    doi: 10.1093/bioinformatics/13.5.555

    CrossRef   Google Scholar

    [63]

    Han MV, Thomas GW, Lugo-Martinez J, Hahn MW. 2013. Estimating gene gain and loss rates in the presence of error in genome assembly and annotation using CAFE 3. Molecular Biology and Evolution 30:1987−97

    doi: 10.1093/molbev/mst100

    CrossRef   Google Scholar

    [64]

    Yang Z. 2007. PAML 4: phylogenetic analysis by maximum likelihood. Molecular Biology and Evolution 24:1586−91

    doi: 10.1093/molbev/msm088

    CrossRef   Google Scholar

    [65]

    Zwaenepoel A, Van de Peer Y. 2019. wgd—simple command line tools for the analysis of ancient whole-genome duplications. Bioinformatics (Oxford, England) 35:2153−55

    doi: 10.1093/bioinformatics/bty915

    CrossRef   Google Scholar

    [66]

    Buchfink B, Xie C, Huson DH. 2015. Fast and sensitive protein alignment using DIAMOND. Nature Methods 12:59−60

    doi: 10.1038/nmeth.3176

    CrossRef   Google Scholar

    [67]

    Wang Y, Tang H, Debarry JD, Tan X, Li J, et al. 2012. MCScanX: a toolkit for detection and evolutionary analysis of gene synteny and collinearity. Nucleic Acids Research 40:e49

    doi: 10.1093/nar/gkr1293

    CrossRef   Google Scholar

    [68]

    Tang H, Krishnakuar V, Li J. 2015. jcvi: JCVI utility libraries. Zenodo. http://doi.org/10.5281/zenodo.31631

    [69]

    Xu Y, Bi C, Wu G, Wei S, Dai X, et al. 2016. VGSC: A web-based vector graph toolkit of genome synteny and collinearity. BioMed Research International 2016:7823429

    doi: 10.1155/2016/7823429

    CrossRef   Google Scholar

  • Cite this article

    Yang L, Li M, Shen M, Bu S, Zhu B, et al. 2022. Chromosome-level genome assembly and annotation of the native Chinese wild blueberry Vaccinium bracteatum. Fruit Research 2:8 doi: 10.48130/FruRes-2022-0008
    Yang L, Li M, Shen M, Bu S, Zhu B, et al. 2022. Chromosome-level genome assembly and annotation of the native Chinese wild blueberry Vaccinium bracteatum. Fruit Research 2:8 doi: 10.48130/FruRes-2022-0008

Figures(8)  /  Tables(1)

Article Metrics

Article views(5009) PDF downloads(618)

ARTICLE   Open Access    

Chromosome-level genome assembly and annotation of the native Chinese wild blueberry Vaccinium bracteatum

Fruit Research  2 Article number: 8  (2022)  |  Cite this article

Abstract: Vaccinium bracteatum Thunb., an important native Chinese wild blueberry species, is widely used as a rootstock and in blueberry cultivar breeding, as well as in traditional medicine and local food. We report here the genome sequence of V. bracteatum using a combination of Oxford Nanopore Technologies long-read and Illumina HiSeq short-read sequencing technologies to obtain 65.30 Gb of clean data, achieving 114.60-fold genome coverage. The assembled genome has a total sequence length of 569.81 Mb and consists of 36,756 predicted genes. Repetitive DNA sequences represent 57.78% of the genome sequence. Comparative genomic analysis revealed that a total of 336 gene families had expanded and that 298 candidate genes had undergone positive selection during evolution in V. bracteatum. The divergence of V. bracteatum from the related Rhododendron williamsianum and Rhododendron delavayi occurred ~13−85 million years ago. The genome sequence of V. bracteatum allowed us to identify some important genes associated with traits involved in fruit development, such as flavonoid biosynthesis, sugar and acid metabolism, MYB transcription factor gene expression, and hormone regulation. The differential expression patterns of genes encoding flavonoid biosynthesis enzymes and MYB transcription factors might explain the high flavonoid content of V. bracteatum. This chromosome-level genome assembly provides reference sequences for the identification and characterization of genes important in the improvement of blueberry and related research.

    • Vaccinium bracteatum Thunb. (known as 'sea bilberry', 'oriental blueberry', or 'Nan zhu' in China) is a wild blueberry, widely distributed in East Asia, especially in China, Japan, and Korea[1]. V. bracteatum is a traditional medicinal plant, recorded in the Compendium of Materia Medica. Many studies have reported the health benefits of extracts from V. bracteatum leaves or fruit[24]. In eastern coastal regions of China, the pigment from V. bracteatum leaves is used to dye rice to produce 'Wu Mi Fan', a well-known local traditional food, dating back 1,000 years[4]. Studies have shown that V. bracteatum leaves contain a number of phytochemical compounds, such as flavonoids[4,5], polysaccharides[2], iridoid glycosides[6], vaccinoside[7], free amino acids, and organic acids[1,6,8].

      Chromosome-level genome assembly of some vacciniums have been reported such as cranberry[9,10] and bilberry[11]. V. bracteatum can also be used as a rootstock to enhance the adaptation of cultivated blueberry. However, little research has been reported on V. bracteatum due to the lack of genomic information. Here, we report on the sequences of the whole-genome assembly and of the transcriptome of V. bracteatum. Our results provide key insights into the transcriptional regulation of flavonoid biosynthesis genes in V. bracteatum.

    • For V. bracteatum genome sequencing, a total of 26.51 Gb of filtered, high-quality reads was obtained to construct a library with average insert length not exceeding 350 bp, followed by sequencing on the Illumina HiSeq sequencing platform. The total sequencing coverage depth was about 45×. The Q20 and Q30 read outputs of sequenced data were at least 97.64% and 93.49%, respectively (Supplementary Table S1).

      Karyotype analysis confirmed that V. bracteatum had a diploid genome (2n = 2x = 24). After contaminants and extranuclear DNA content were taken into account, the genome size of V. bracteatum, estimated by K-mer analysis, was 579.42 Mb. The repetitive sequence content was ~42.72%, and the estimated heterozygosity was about 1.10%. In addition, the GC content of the genome of this species was 38.63%. Therefore, the genome of V. bracteatum is a heterozygous complex genome (Table 1 & Supplementary Fig. S1).

      Table 1.  Major indicators of V. bracteatum genome of contig-level and chromosome-scale assembly.

      Assembly featureStatistic
      Contig-level assemblyChromosome-
      scale assembly
      Estimated genome size (by k-mer analysis) (Mb)579.42
      Repetitive sequence content42.72%57.78%
      GC content (estimation)38.63%
      Estimated heterozygosity1.10%
      Assembled genome size (Mb)569.81
      Contig number1,3841,430
      Contig N50 (Mb)1.981.87
      Contig N90 (Mb)0.300.26
      Contig max (Mb)9.429.42
      GC content (Nanopore)38.32%38.32%
      Assembly % of genome98.33%
      Scaffold number973
      Scaffold length (Mb)569.86
      Contig length (Mb)569.81
      Scaffold N50 (Mb)43.77
      Scaffold N90 (Mb)39.17
      Scaffold max (Mb)50.70
      Gap total length (Mb)0.05
    • Using a third generation Nanopore platform sequencer with a sequencing depth of 114.60×, a total of 76.41 Gb of raw data was measured, with 65.3 Gb of clean data being obtained after filtering out adapter sequences, short fragments (length < 2,000 bp) and low-quality sequences. The clean data contained a total of 3,651,468 reads, the average read length was 17,882 bp, the N50 read length was 18,829 bp, and the longest read length was 304,639 bp (Supplementary Table S1 & S2). After further trimming, the genome sequence was 569.81 Mb, and the longest contig was 9.42 Mb, with the contig N50 and N90 being 1.98 Mb and 0.30 Mb, respectively. The GC content of the genome was 38.32% (Table 1). In the evaluation of assembly results, the proportions of properly mapped reads, conserved essential genes (CEGs), and complete Benchmarking Universal Single-Copy Orthologs (BUSCOs) were 90.51%, 94.10%, and 90.95%, respectively (Supplementary Table S3). These results indicated that the genome assembly was of high quality and adequate coverage.

      The genome assembly results from the high-throughput chromosome conformation capture technique (Hi-C) showing that a total of 560.32 Mb sequences were associated with 12 chromosomes, accounting for 98.33% of the genome; of the sequences located on the chromosomes, the total length of the sequences determined was 532.95 Mb, accounting for 95.12% of the total length of the chromosomal sequences located (Table 1). The detailed distribution of each chromosome sequence is exhibited in Supplementary Table S4. Based on the statistics of the genome sequences assembled by Hi-C, the contig N50 and scaffold N50 of the final genome were 1.87 Mb and 43.77 Mb, respectively. All of the genome statistics following Hi-C assembly are shown in Table 1.

      The results of the chromosome heat map of the Hi-C genome assembly revealed 12 distinct chromosomal groups. Within each group, the interaction intensity at the diagonal position (between adjacent sequences) was higher than that at the non-diagonal position. The results showed that the effect of genome assembly was fine, which is consistent with the principle of Hi-C assisted genome assembly (Supplementary Fig. S2). Summaries of the de novo assembly and sequencing analysis of the V. bracteatum genome can be found in Fig. 1, with the distribution of the GC content, gene density, and number of repeat sequences per Mb being shown in Fig. 1b, c and d, respectively.

      Figure 1. 

      Summary of the de novo genome assembly and sequencing analysis of V. bracteatum. (a) Morphological features of the fruit of V. bracteatum. (b) Circos plot showing GC content, gene density, repeat coverage, collinearity between chromosomes of V. bracteatum. I, chromosome number; II, GC content; III, gene density; IV, numbers of repeat sequences per megabase pair; V, paralogous relationships between chromosomes.

    • For repeat sequence annotation, a total of 329.23 Mb of sequences were identified, accounting for 57.78% of the genome size (Supplementary Table S5). Of these repeat sequences, 48.28% (275.11 Mb) and 5.26% (29.95 Mb) were predicted as Class I transposons and Class II retrotransposons, respectively. Of the Class I transposons, Sukkula (LARD) and unknown-type transposons were the most and least frequent, accounting for 22.09% (125.89 Mb) and 0.05% (260.81 Kb), respectively. Of the Class II transposons, terminal inverted repeat (TIR) mobile elements were the most abundant, accounting for 4.41% (25.13 Mb). In addition, potential host genes and Simple Sequence Repeats (SSRs) accounted for 1.56% (8.89 Mb) and 0.04% (232.80 Kb), respectively (Supplementary Table S5).

      To predict the number of coding genes, three different strategies were adopted, namely ab initio, homology-based, and RNA-Seq prediction (Supplementary Table S6). The statistical results of gene information combined the three prediction methods after EVM integration showed that the number of coding genes was 36,756, of which 19,247 coding genes were shared by the three prediction methods (Supplementary Fig. S3). Among the 36,756 genes, the total gene length and average gene length were 194,291,038 bp and 5,285.97 bp, respectively. The total exon length and the average exon length per gene were 51,917,599 bp and 1,412.49 bp, respectively. The total number of exons and the average number of exons per gene were 187,829 and 5.11, respectively (Supplementary Table S7).

      By prediction of non-coding RNAs (ncRNAs), a total of 96 miRNAs, 218 rRNAs, and 376 tRNAs were identified, belonging to 18, 4, and 23 families, respectively. Pseudogenes have sequences similar to functional genes, while losing their original functions due to mutations such as insertion and deletion. Here, pseudogene was obtained using the predicted protein sequence and BlAST alignment to search homologous gene sequence in the genome, and then GeneWise software was employed to find the immature termination codon and frameshift mutation in the gene sequence. Pseudogene prediction results showed that the total number of pseudogenes predicted was 4,218. Total pseudogene length in the genome was 9,251,782 bp, whereas the average pseudogene length was 2,193.40 bp (Supplementary Table S8).

      For gene function annotation, the predicted gene sequences were screened against Non-Redundant protein (NR), Eukaryotic Orthologous Groups (KOG), Gene Ontology (GO), Kyoto Encyclopedia of Genes and Genomes (KEGG), Translated EMBL (TrEMBL), and other functional databases by BLAST v2.2.31 (set cutoff as e-value 1e-5). The gene function annotation analyses reported here were carried out using KEGG, KOG, and GO. The results showed that 29,890 genes, accounting for 81.32% of the total gene number, could be functionally annotated from the KOG, KEGG, and GO databases (Supplementary Table S9), and the results of the analyses are shown in Supplementary Figs S4, S5, and S6, respectively.

    • To investigate the species-specific and common gene families, conserved putative genes from seven different species (V. bracteatum, Rhododendron occidentalis, Rhododendron williamsianum, Rhododendron delavayi, Fragaria vesca, Vitis vinifera, and Solanum lycopersicum) were used to identify gene family clusters. A total of 30,442 gene families were identified. The seven species shared 5,929 common gene families (Fig. 2a & Supplementary Table S10). Venn diagram analysis of the gene families revealed that 7,831 clusters were observed in V. bracteatum, R. delavayi, R. williamsianum, V. vinifera, and R. occidentalis, and 1,725 appeared to be lineage specific to V. bracteatum, whereas 14,299 were common to V. bracteatum and R. delavayi, 12,492 with R. williamsianum, 11,738 with V. vinifera, and 10,667 with R. occidentalis (Fig. 2b). GO and KEGG enrichment analysis was carried out to analyze gene families in V. bracteatum (Supplementary Table S11). Based on the GO database, most of the genes annotated to the three gene ontology categories, namely molecular function, cellular component, and biological process, were enriched most with respect to cellular process, plastid, and binding, respectively. For KEGG enrichment analysis, phenylpropanoid biosynthesis and protein processing in the endoplasmic reticulum (ER) were the dominant pathway terms which contained the most genes (Supplementary Fig. S7).

      Figure 2. 

      Gene family analysis among the seven plant species. (a) Gene family clustering petal map. The middle circle is the number of gene families common to all species, and the edge is the number of gene families unique to each species. (b) Gene family clustering Venn diagram. The number below the species name is the total number of gene families, and the corresponding number of genes are in parentheses; the number in the Venn diagram is the number of gene families.

    • The 2,883 single-copy gene sequences were used to construct the phylogenetic tree, using IQ-TREE v1.6.11 software, to show relationships among the aforementioned seven species. The results indicated that V. bracteatum showed a close relationship with the species from the same Ericaceae family, R. williamsianum and R. delavayi (Fig. 3a). To estimate the dates of divergence among these plant species, the RelTime model was used. The results indicated that the time of divergence between R. williamsianum and R. delavayi, which was used as a calibration point, was ~6−10 million years ago (Mya), and the divergence time of V. bracteatum was ~13−85 Mya, whereas V. bracteatum and S. lycopersicum diverged from a common ancestor ~36−119 Mya (Fig. 3a).

      Figure 3. 

      Evolution of V. bracteatum analysis. (a) Constructed phylogenetic tree and divergence time estimation (MYA, millions of years ago). Pa, Paleogene; Ne, Neogene. (b) The contraction and expansion of gene family of above mentioned seven species. +, no. of gene families expanded on the node; −, no. of gene families contracted on the node. The pie charts show the proportion of corresponding branch contraction and expansion gene families. (c) The Ks distribution map within and between species.

      Combining the results of the phylogenetic tree with divergence time and gene family clustering, the contraction and expansion of gene families in V. bracteatum, relative to its ancestors, were predicted using CAFE v4.2 software. A total of 336 gene families were shown to have expanded in V. bracteatum, whereas only nine gene families had undergone contraction (Fig. 3b & Supplementary Table S12).

      Ks values (number of synonymous substitutions per synonymous site), calculated for homologous gene pairs between two species or within the species itself, can be used to estimate the timing of large-scale duplications. The distribution of Ks values between V. bracteatum paralogous pairs displayed only one peak, at 0.67, which indicated a recent whole-genome duplication (WGD) event in the evolution of V. bracteatum. The distribution of Ks values between V. bracteatum and F. vesca exhibited only one peak, which suggested they shared the same ancient WGT event. Compared with F. vesca and R. occidentalis, the low Ks values found in V. bracteatum suggested that the divergence of V. bracteatum occurred later than with these other species (Fig. 3c & Supplementary Fig. S8). Fossil records were downloaded from the TIMETREE website (https://www.timetree.org) and used to calibrate the results. The divergence time of V. bracteatum and F. vesca was set to 121 Mya, with the Ks value of V. bracteatum vs. F. vesca being 1.25. Based on the generally accepted evolutionary rate, the WGD event of V. bracteatum occurred at approximately 64 Mya. The WGD event occurred at 11 Mya base on the divergence time of V. bracteatum and R. williamsianum or R. delavayi.

    • Chromosome evolution between the genomes of V. bracteatum and those of the related species R. occidentalis, R. williamsianum, and F. vesca was evaluated using MCScan algorithms. The occurrence of large-scale chromosomal fragment rearrangements was found among V. bracteatum, F. vesca, R. occidentalis and R. williamsianum, including inversions and translocations (Fig. 4a). Also, compared with the groupings of V. bracteatum vs. F. vesca, and V. bracteatum vs. R. occidentalis, there were fewer scattered points in V. bracteatum vs. R. williamsianum, suggesting a closer relationship between V. bracteatum and R. williamsianum (Fig. 4b). Overall, these findings offer new insights into the evolution of V. bracteatum chromosomes.

      Figure 4. 

      The chromosomal collinearity among V. bracteatum, R. williamsianum, R. occidentalis, and F. vesca. (a) The chromosome map in four species. (b) The inter-genomic comparison in point graph form.

    • The Ka/Ks ratios of the single-copy V. bracteatum genes were evaluated with those from the other six species (V. vinifera, R. delavayi, R. williamsianum, S. lycopersicum, F. vesca, and R. occidentalis) to identify positively selected genes (Ka/Ks > 1). A total of 298 candidate genes in V. bracteatum exhibited positive selection (Supplementary Table S13), and the annotation of these positively selected genes can be found in Supplementary Table S14. For GO analysis, 179 of the positively selected genes were enriched in three categories namely biological process (the terms of metabolic process, cellular process and response to stimulus account for the top three), molecular function (the terms of cell part, cell and organelle account for the top three), and cellular component (the terms of catalytic activity, binding and transporter activity account for the top three) (Fig. 5a & Supplementary Table S15). For KEGG analysis, most of the enriched genes were associated with pathways dealing with mismatch repair and peroxisome function (Fig. 5b & Supplementary Table S16). For example, EVM0010597 was the positively selected gene related to ABC transporter which may play a role in transmembrane transport of sugars and cell metabolites (Fig. 5b & Supplementary Table S14).

      Figure 5. 

      Gene ontology (GO) and KEGG enrichment analysis of positively selected genes in V. bracteatum by clusterProfile. (a) GO enrichment analysis. The abscissa in the column chart represent the number of genes under the GO term, and the color is the corresponding P value. (b) The KEGG enrichment analysis. The abscissa in the plot represents proportion of the number of positively selected genes under the pathway to the total positively selected genes. The size of the point represents the number of genes enriched on this pathway, and the color is the corresponding P value.

    • Compared with the flavonoid-related genes of Arabidopsis, the greatest number of flavonoid -related genes was found in V. bracteatum. The genome assembly allowed us to locate all the structural genes of the anthocyanin biosynthesis pathway. From the genome sequences, a total of 48 genes, mainly encoding these structural genes, were identified by screening. Of these genes, nine were located on chromosome 2, namely two flavonol synthase genes (FLS), four chalcone synthase genes (CHS), one leucoanthocyanidin reductase gene (LAR), one gene encoding a cytochrome P450 enzyme, and one chalcone and stilbene synthase family protein gene (Fig. 6).

      Figure 6. 

      Flavonoid related genes located on the chromosomes. LG01, LG02, ... and LG12 indicate chromosome 1, 2, ... and 12, respectively.

    • To verify the classification of MYB transcription factors in V. bracteatum, we constructed a phylogenetic tree by aligning the MYB proteins. Fourteen out of 44 MYB genes exhibited significantly different expression during fruit development, which indicated that most MYB-related genes in V. bracteatum may not be related to fruit development, with the exceptions of MYB110_5 and MYB_5 (Fig. 7). Interestingly, no differentially expressed genes (DEGs) were found between green fruit and leaf, which indicated that the MYB family may not contribute to unripe fruit development.

      Figure 7. 

      Phylogenetic tree and heatmap of MYB transcription factor in the V. bracteatum genome.

    • Transcriptomes from different tissues, including leaf, green fruit, pink fruit, and blue fruit, were mapped to the assembled genome. The transcripts among the various tissue samples were studied using principal component analysis (PCA), and expression was based on fragments per kilobase of transcript per million mapped reads (FPKM). Samples were grouped using data for all 39,867 transcripts analyzed by the Cufflinks tool. Our PCA analysis indicated that PCA1 (39.29%) and PCA2 (18.67%) best described the sources of variance among the different samples (Supplementary Fig. S9). Generally, replicate samples belonging to a given developmental stage clustered more closely together than did replicate samples belonging to different developmental stages (Supplementary Fig. S9).

      The DEGs were filtered using the cuffdiff tool with a false discovery rate (FDR) < 0.05. Ultimately, we obtained 5,508 DEGs from green fruit to pink fruit consisting of 2,857 down-regulated genes in the pink fruit and 2,651 up-regulated genes. The number of DEGs in the comparison between blue fruit and pink fruit was 4,563, including 2,315 down-regulated in the blue fruit and 2,248 up-regulated genes (Supplementary Fig. S10).

      To explore the DEGs between the early stages (from green to pink fruit) and the late stages of fruit development (from pink to blue fruit), GO term enrichment analysis was performed (Supplementary Fig. S11). GO analysis found that many of the DEGs were involved in the 'oxidation-reduction process' in the biological process category. Eleven genes belonged to this term, including the flavonoid biosynthesis genes dihydroflavonol 4-reductase (DFR), flavonoid 3′,5′-hydroxylase (F3'5'H), flavonol synthase (FLS), flavanone 3-hydroxylase (F3H), cinnamic acid 4-hydroxylase (C4H), and anthocyanidin reductase (ANR). Many DEGs detected were involved in the 'integral component of membrane' term in the cellular component category, including DFR, F3'5'H, CHS, flavanone 3′-monooxygenase (F3'H), and C4H. The term 'flavonoid biosynthetic process' was enriched in the late stage, but not in the early stage.

      The expression of genes CHS, C4H, and F3'5'H increased from the green fruit stage to the blue fruit stage, whereas that of genes FLS and ANR decreased. Expression of most of the flavonoid biosynthesis-related genes was downregulated in the comparison of green fruit vs. leaf, except for ANR and FLS; the expression of these genes was also higher in pink fruit but lower in blue fruits (Fig. 8).

      Figure 8. 

      Heat map diagram of the expression of differently expressed genes (DEGs) related to flavonoid biosynthesis in the comparisons of green fruit vs. leaf, pink fruit vs. leaf, and blue fruit vs. leaf. Green represents downregulated and red represents upregulated. anthocyanidin reductase (ANR); cinnamic acid 4-hydroxylase (C4H); chalcone isomerase (CHI); chalcone synthase (CHS); dihydroflavonol 4-reductase (DFR); flavonoid 3'-hydroxylase (F3’H); flavonoid 3',5’-hydroxylase (F3’5’H); flavanone 3-hydroxylase (F3H); flavonol synthase (FLS); leucoanthocyanidin reductase (LAR).

    • V. bracteatum is a member of the Ericaceae and is a typical diploid plant, with 12 different chromosomes and a very small genome[12]. In the current study, we report the first de novo assembly of the V. bracteatum genome, through a combination of ONT long-read and Illumina HiSeq short-read sequencing technologies. Due to the complexity of the blueberry genetic diversity, there are many difficulties in the study of blueberry biological characteristics, especially its genome[13,14]. For example, a draft genome for a wild diploid species V. corymbosum (2n = 2x =24) of blueberry was previously assembled consisting of a large number of scaffolds (total of 13,757; N50 of ~145 kb), a high percentage of gaps (~27.35%) in a ~393.16 Mb assembly[15]. The first chromosome-scale genome assembly of the tetraploid highbush blueberry (V. corymbosum cv. Draper) (2n = 4x = 48) consisted of 48 pseudomolecules with ~1.68 Gb of assembled sequences[16]. V. darrowii Camp (2n = 2x = 24) of blueberry is scaffolded into 24 chromosomes with ~1.06 Gb[17]. The genome of V. bracteatum is very small and the diploid may be easier to be homozygous when comparing to multiploidy, so that obtaining the whole-genome information for V. bracteatum for molecular biology research could provide guidance and reference for the larger and more complex genome of the cultivated blueberry.

      Based on the whole-genome sequencing data, comparative genomic analysis was performed between V. bracteatum and six other related plant species. Similar comparative genomics analysis studies have been reported for other plant species. For example, compared with the ratio of gene expansion to contraction in sweet orange (0.6) and longan (0.4), the ratio in the mango genome was 4.5[18]. The highest ratio (expansion of 336 vs. contraction of 9 gene families) among the seven species studied was in the V. bracteatum genome, reflecting a relatively recent occurrence of the WGD event in the V. bracteatum genome.

      PANTHER annotation results showed that the MYB gene family (OG0000036), F-box gene family (OG0000011) and the LRR receptor-like serine/threonine protein kinase gene family (OG0000010) belonged to a large extended gene family (Supplementary Table S12). GO and KEGG analysis of the expanded gene family showed that the gene family associated with oxidoreductase (GO) and flavonoid synthesis-related enzyme (KEGG) activities showed strong gene expansion in V. bracteatum (Supplementary Table S8). These results were in agreement with an earlier investigation in blueberry, where several genes encoding key biosynthetic steps in many antioxidant pathways were enriched with tandem gene duplications, and expanded gene families were involved in the biosynthesis of anthocyanins. Compared with the six other plant species, 298 positively selected genes were identified in the current study in V. bracteatum (Supplementary Table S13). These positively selected, expanded genes offered valuable insights into the formation of phenotypic characteristics and evolution of V. bracteatum.

      A previous study had demonstrated that Rhododendron and Vaccinium represent species-rich genera within the Ericaceae, which had diverged from one another ~77 Mya[19]. Compared with R. delavayi and R. williamsianum, collinearity analysis in the present study showed few scattered points in the V. bracteatum/R. williamsianum comparison in a scatter diagram, suggesting a close relationship between these latter two species. The evolutionary analysis suggested that V. bracteatum and the other two species, R. delavayi and R. williamsianum, may have diverged ~13-85 Mya (Figs 3a & 4). Previous evidence had shown that the two shared WGDs represented by similar Ks values in the Rhododendron and Vaccinium genomes represent two ancient shared WGDs, originating from a common ancestor of the Ericaceae, which can be traced back to a common ancestor of the Ericales[20]. In our current study, the low Ks value of 0.67 found in V. bracteatum, compared with F. vesca and R. occidentalis, suggested that the divergence of V. bracteatum occurred later than for the other two species (Fig. 3c). The genomic data from the current study will provide valuable reference material for understanding the expression and regulation of important agronomic traits in V. bracteatum and related species.

      Flavonoid-related genes were located on the genome of V. bracteatum. These genes did not cluster together, indicating that V. bracteatum may have undergone several WGD events. Chalcone synthase (CHS) is a key enzyme in the flavonoid biosynthesis pathway, and eight genes encoding CHS were detected in the V. bracteatum genome, more than was reported from any of the other six species. Nine genes encoding F3'5'H were detected in the V. bracteatum genome. FLS is an important enzyme necessary for flavonol biosynthesis, and seven FLS genes were detected in the V. bracteatum genome. The up-regulated expression of many structural genes involved in flavonoid biosynthesis during fruit ripening suggested that flavonoids play an important role during fruit development.

      Compared with unripe, green fruit, some MYB transcription factor genes exhibited greater transcript abundance in developing pink and blue fruits, indicating that these genes may be involved in controlling pathways of flavonoid biosynthesis during V. bracteatum fruit maturation. MYB110, MYB108, and MYB44 were particularly highly expressed during fruit development, which indicated that these genes may play an important role in regulating flavonoid biosynthesis during fruit maturation. In kiwifruit[21], the lack of MYB110 expression is responsible for the total absence of anthocyanins in the fruits, as MYB110 promotes the transcription of the F3'H and F3'5'H genes. MYB108 is involved in regulating various biosynthetic pathways in different plant species. In Rosa multiflora, MYB108 expression was induced by chilling stress[22], and MYB108 expression was required for jasmonic acid-mediated stamen and pollen maturation in Arabidopsis[23], while overexpression of MYB108 in Arabidopsis thaliana conferred improved tolerance to the Verticillium dahlia infection[24]. On the other hand, MYB44 acts as a repressor of anthocyanin biosynthesis in sweet potato[25], suggesting that MYB44 may have multiple functions in plants. MYB10 and MYB_3 were downregulated in V. bracteatum during fruit development, which indicated that these genes may inhibit the biosynthesis of anthocyanins. On the other hand, MYB10 expression increased during the biosynthesis of anthocyanidins in apple[26].

      Differential expression patterns of sucrose synthase genes during V. bracteatum fruit development indicated that they may play various roles in fruit development. Sugar transporters have been proved to regulate intercellular sugar transport in the phloem to ripening fruit. In the current study, expression of three of the AST (encoding aspartate aminotransferase) homologs increased during the early stage of V. bracteatum fruit development, with no significant change being observed during the late stage. The expression of PEPC (encoding phosphoenolpyruvate carboxylase 4) was downregulated during the early stage of fruit development, a finding which was consistent with the results from a previous study[27]. We found that expression of IDH (encoding isocitrate dehydrogenase) was upregulated early in fruit development, whereas that of GDH (encoding glutamate dehydrogenase) was upregulated at the late stage. It seemed that most of the genes related to high fruit acidity were highly expressed early in fruit development in V. bracteatum, whereas the genes related to fruit sugar levels were highly expressed late in fruit development (Supplementary Fig. S12).

      The genes associated with ethylene biosynthesis or signaling were highly expressed in the early stage rather than the late stage of fruit development. Two ACC oxidase and one ethylene receptor (ETR) genes were upregulated at the early stage. Ethylene has been reported to negatively regulate anthocyanin content[28]. The expression of NCED5 (encoding 9-cis-epoxycarotenoid dioxygenase 5) was upregulated in both the early and late stages of fruit development. NCEDs catalyze the first step of abscisic acid (ABA) biosynthesis from carotenoids. High expression of NCEDs was found in purple-skinned apple compared with non-purple fruit[29]. The expression of ethylene- or ABA-related genes indicated that the increasing content of anthocyanin during fruit maturation may be related to the regulation of phytohormone concentrations.

      Hormones affect fruit development not only by the interaction between different hormones but also by the interaction between the hormone and sugars. For example, ABA, the auxin indole-3-acetic acid (IAA), and ethylene work together to regulate the development of blueberry fruit, while gibberellins (GA) and IAA can promote the absorption of sugar at the early stage of fruit development[30], whereas ABA and ethylene enhance it at the late stage. In our current study, ABA and ethylene may also increase the anthocyanin content. The blueberry color change during ripening is caused by changes in the anthocyanin content. The expression of some flavonoid biosynthesis structural genes such as phenylalanine ammonia-lyase (PAL), CHS, CHI, F3H, and F3'H, as well that of some transcription factor genes[31], increased during fruit development[32]. These results were in agreement with the findings of our own research.

      In conclusion, we present here a chromosome-level genome sequence of the wild blueberry species V. bracteatum. This first genome assembly from wild blueberry is expected to advance our understanding of the evolutionary history of blueberry and of the gene expression changes which occur during fruit development. The genome sequence will provide fundamental genomic resources for blueberry improvement. Our phylogenetic analysis of MYB transcriptional factors in wild blueberry has already led to the discovery of several novel MYBs, as well as providing evidence to suggest that MYB110, MYB108, and MYB44 may play an important role in fruit maturation. From the transcriptome, levels of CHS, C4H, F3'5'H, sucrose synthases, sugar transporters, ACC oxidase, and ETR were shown to increase at the late stage of fruit development. In this way, the V. bracteatum genome will serve as an important resource for the development of genomics-assisted selection to achieve blueberry improvement, particularly for traits related to the efficiency of flavonoid production and with stress tolerance.

    • Young leaves of V. bracteatum were collected on 15 April 2019 in the low mountain and hill region of southern Anhui Province, China (30°51' N, 118°23' E) for genome sequencing, and were immediately frozen in liquid nitrogen and stored at −80 °C until DNA and RNA were extracted. The fruits were manually picked at three developmental stages: the expansion phase with green fruit (Stage 1), the color accumulation phase with pink-red fruit (Stage 2), and the ripening phase with blue fruit (Stage 3). Three biological replicates were conducted for each stage, with 10 representative fruits being sampled for each replicate at each stage. Fruit samples were frozen with liquid nitrogen, then stored at −80 °C until RNA was extracted. The RNA samples were used for transcriptome-based gene prediction.

    • From young leaves, total genomic DNA was extracted according to the CTAB protocol[33]. After that, the genomic DNA was randomly disrupted into 350 bp length. Through the steps of terminal repair, the addition of A adaptors and linkers, target fragment selection and PCR amplification, a small-fragment sequencing library was established. Subsequently, the library was quality inspected and chip fixed, and then subjected to PE 150 (paired-end 150-bp) sequencing using the Illumina HiSeq sequencer (Illumina, San Diego, CA, USA). Sequence data were characterized and filtered to obtain clean reads, which were used for genome size evaluation, genome assembly, GC content statistics, heterozygosity rate statistics, and post genome assembly assessment. K-mer is an oligonucleotide sequence of length k extracted from the sliding windows of sequencing data. Under the premise of a uniform distribution of sequencing reads, the following formula is as follow: Genomic size = total number of bases/average sequencing depth = total K-mer/median K-mer depth[34,35]. To evaluate genome size, heterozygosity and the repeat sequence ratio, the distribution map of K-mers (K = 19) was constructed based on the 350-bp library data.

    • The processes of sample quality detection, library construction, quality detection and sequencing were performed here according to the manufacturer's instructions (Oxford Nanopore Technologies, ONT, Canada). Clean data of Nanopore's third-generation sequencing were corrected with the software of Canu[36], and then assembled with Wtdbg2 software based on the corrected data. After that, the third-generation sequencing data were used for three rounds of correction through Racon software[37], and then the second-generation data is used for three rounds of correction through Pilon software[38].

      Hi-C technology is used to assist genome assembly. Genomic DNA from the fresh leaves of V. bracteatum was prepared and in situ Hi-C was used for Hi-C library construction and sequencing, mainly including cell cross-linking, endonuclease digestion, terminal repair, cyclization, DNA purification and capture, and computer sequencing through the Illumina platform with reading length of PE150[39,40].

      After sequencing, quality control, and Hi-C library quality evaluation, Hi-C-assisted genome assembly was carried out. The separation into groups, sequencing, and orientation of genomic sequences are carried out by using LACHESIS software, and the assembly results were evaluated[41].

    • The evaluation of assembly results includes three aspects: the ratio to the Illumine sequencing reads, core gene integrity and BUSCO evaluation. First, BWA software was used for reading comparison rate of second generation sequencing[42]. It compares the short sequences obtained from the Illumina HiSeq sequencing platform with the reference genome. By statistical comparison rate, the integrity of the assembled genome can be evaluated. Second, CEGMA v2.5 database contains 458 conserved core genes in eukaryotes, which can be used to evaluate the integrity of the final genome assembly[43]. Third, the database of BUSCO V5.0 was performed to evaluate the integrity of the genome assembly, the parameter of BUSCO was –evalue 1e-05 (E-value cutoff for BLAST searches)[44].

    • Genome annotation analysis included repeat sequence annotation, coding gene prediction, non-coding RNA prediction, pseudogene annotation, and gene function annotation. For repeat sequence annotation, LTR_ FINDER and RepeatScout software were used to construct the repeat sequence database of the V. bracteatum genome based on the principle of structural prediction and de novo prediction[45,46]. PASTEClassifier was used for classifying all isolated sequences which later mapped to the Repbase database using RepeatMasker software[4749].

      Homologous proteins from four plant genomes (Arabidopsis thaliana, Oryza sativa ssp. japonica, Rhododendron_delavayi, and V._corymbosum) were downloaded from the Ensembl and NCBI databases. EVM v1.1.1 was used to integrate the predicted results. For non-coding RNA (ncRNA) prediction, different strategies have been adopted to predict the structural characteristics of different ncRNAs. Based on the Rfam database, the whole genome alignment was carried out with BLASTN to identify micro RNAs (miRNAs) and ribosomal RNA (rRNA). tRNAscan-SE was used to identify tRNA[50]. For pseudogene annotation, the predicted protein sequences were aligned by GenBLASTA software to find homologous gene sequences in the genome[51,52], and then GeneWise software was used to find immature termination codons and frameshift mutations in gene sequences, which could generate pseudogenes[53]. For gene function annotation, the predicted gene sequences were compared with the selected GO function database with BLAST v2.2.31 (-evalue 1e-5), in order to annotate and analyze the gene function[54,55].

    • The protein sequences of V. bracteatum and the six related species (R. occidentalis, R. williamsianum, R. delavayi, F. vesca, V. vinifera, and S. lycopersicum) were used for gene family clustering. Orthofinder v2.4 software was used to classify the protein sequences of the above- mentioned seven species[56], and the PANTHER v15 database was used to annotate the gene families obtained[57]. Finally, GO and KEGG enrichment analyses were carried out for the unique gene families of V. bracteatum .

    • The phylogenetic tree was constructed from single-copy gene sequences using IQ-TREE v1.6.11 software[58]. Specifically, each single-copy gene family sequence was compared using MAFFT v7.205[59], and then Gblocks v0.91b was used to remove those regions with poor sequence alignment or large differences[60]. Finally, all the well-aligned gene family sequences of each species (V. bracteatum, R. occidentalis, R. williamsianum, R. delavayi, F. vesca, V. vinifera, and S. lycopersicum) were connected end-to-end to obtain a supergene, and then the model detection tool ModelFinder, provided by IQ-TREE, was used for model detection to obtain the best model, JTT + F + I + G4, which was used for constructing the phylogenetic tree (using the maximum likelihood (ML) method, and 1000 bootstrap replicates)[61].

      Grape (V. vinifera) was designated as the outlier group of the phylogenetic tree. The MCMCTREE program of the software PAML v4.9i was applied to estimate the divergence time[62]. In detail, the fossil time of R. williamsianum vs. R. delavayi (6.3−9.7 Mya), V. vinifera vs. R. williamsianum (111−131 Mya), and R. occidentalis vs. F. vesca (26−60 Mya) were obtained from the Timetree website (www.timetree.org), which was used to correct the fossil age obtained by the software, based on an algorithm. Then, the two parameters of 'gradient' and 'Hessian' required for divergence time estimation were obtained by the mcmctree ctl file in the PAML package. Finally, the ML method was used to estimate the divergence time by the correlated molecular clock and JC69 model, and two replicate calculations were carried out to assess for consistency.

    • Combining the results of the phylogenetic tree with those from divergence time and gene family clustering analysis, the number of gene family members of the ancestor of each branch was estimated through the birth–mortality model by the CAFE v4.2 software, so as to predict the contraction and expansion of the gene family relative to the ancestor[63]. The criteria for defining whether there was significant expansion or contraction was that the family-wide P-values and viterbi P-values were less than 0.05. For the extracted expansion and contraction of the gene families of V. bracteatum, PANTHER annotation was carried out first, and then GO and KEGG enrichment analysis was carried out using clusterProfile software.

    • The CODEML model in the PAML package is mainly used for positive selection analysis. The single-copy gene families among V. vinifera, R. delavayi, R. williamsianum, V. bracteatum, S. lycopersicum, F. vesca and R. occidentalis were obtained, and then the protein sequences of each gene family were aligned by MAFFT (parameter: -- localpair -- maxiterate 1,000 input_file). The sequences were reversed into codon alignment sequences by PAL2NAL. Finally, CodeML was used (F3 × 4 model of codon frequencies), based on the Branch-site model, Model A (assuming ω > 1) and the null model (no site is allowed with ω > 1), with likelihood ratio tests (LRTs) being performed by the 'chi2' program under the PAML package, and significant differences were identified (P value < 0.05)[64]. Then, the Bayesian empirical Bayes method (BEB) was used to obtain the a posteriori probability of the positive selection site (usually greater than 0.95, which is considered to be the significant positive selection site), and finally the significant positively selected genes were identified. The positively selected genes were then subjected to GO and KEGG functional annotation.

    • Whole-genome duplication (WGD) is a process of genome duplication. We chose the Ks method with the wgd v1.1.1 software[65]. In the study of functional genomics, it is generally believed that collinear genes tend to have the same biological functions. In order to analyze the chromosomal collinearity between V. bracteatum and other related species (R. occidentalis, R. williamsianum, and F. vesca), the gene sequences of pairs of species were compared with diamond v0.9.29.130 software to determine similar gene pairs (e < 1E-5, C score > 0.5, in which the C score value was filtered by JCVI software)[66]. Then, using the gff3 file, the adjacent location on the chromosome of similar gene pairs was identified. This process was mainly carried out through MCScanx (parameter-m 15)[67], until all the genes in the collinear blocks could be obtained. The linear graph patterns and dot matrix forms of the collinearity relationship of each species were drawn by JCVI v0.9.13 and VGSC software, respectively[68,69].

    • Total RNA was extracted from the 12 samples, namely leaf, green fruit, pink fruit, and blue fruit, which each tissue having three replicates. All the samples using the mirVana miRNA Isolation Kit (Ambion) following the manufacturer's protocol. Sequencing libraries were generated using NEBNext UltraTM RNA Library Prep Kit for Illumina (New England Biolabs, Ipswich, MA, USA) following the manufacturer's instructions, and index codes were added to attribute sequences to each sample. The clustering of the index-coded samples was performed on a cBot Cluster Generation System, using TruSeq PE Cluster Kit v4-cBot-HS (Illumina) according to the manufacturer's instructions. After cluster generation, the library preparations were sequenced on an Illumina platform and paired-end reads were generated at BIOMARKER (Beijing, China).

    • Sequencing data used in this study are available in the NCBI Sequence Read Archive (SRA) database under the following accession numbers: BioProject PRJNA794927 (genomic data of native Chinese wild blueberry and RNA-Seq data derived from four different tissues).

      • This work was supported by Anhui Provincial Natural Science Foundation (2008085QC129 and 2008085MC77), the 17th Huo Yingdong Education Fund Project (171022) and and Key Research and Development Program in Anhui Province (202204c06020029).

      • The authors declare that they have no conflict of interest.

      • Copyright: 2022 by the author(s). Exclusive Licensee Maximum Academic Press, Fayetteville, GA. This article is an open access article distributed under Creative Commons Attribution License (CC BY 4.0), visit https://creativecommons.org/licenses/by/4.0/.
    Figure (8)  Table (1) References (69)
  • About this article
    Cite this article
    Yang L, Li M, Shen M, Bu S, Zhu B, et al. 2022. Chromosome-level genome assembly and annotation of the native Chinese wild blueberry Vaccinium bracteatum. Fruit Research 2:8 doi: 10.48130/FruRes-2022-0008
    Yang L, Li M, Shen M, Bu S, Zhu B, et al. 2022. Chromosome-level genome assembly and annotation of the native Chinese wild blueberry Vaccinium bracteatum. Fruit Research 2:8 doi: 10.48130/FruRes-2022-0008

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint