Search
2022 Volume 2
Article Contents
REVIEW   Open Access    

The genomics of ornamental plants: current status and opportunities

More Information
  • With the rapid development of sequencing technologies, followed by the reduction of sequencing cost, numerous ornamental plants have been sequenced, resulting in their genomic studies shifting from gene cloning and marker development to whole genome profiling. A profound understanding of genome structure and function at the whole genome level can not only help to modify ornamental traits, such as fragrance, color and flower shape, through genetic engineering, but also infer the genetic relationship and evolutionary history of ornamental plants via comparative genomics analysis. In this paper, we review the current situation of sequencing strategies and the application of genomics to study the origin and evolution of ornamental plants. We highlight challenges of ornamental plant genomic research. The use of cutting-edge technologies, such as genomics, gene editing and molecular design polymerization breeding, can facilitate our understanding of genetic regulation mechanisms and the germplasm innovation of important traits in ornamental plants. The results can be expected to significantly increase the breeding efficiency of ornamental plants.
  • 加载中
  • Supplemental Table S1 List of the representative public accessible genomes of ornamental plant and their database construction status.
    Supplemental Table S2 List of the other public accessible genomes of ornamental plant and their database construction status.
    Supplemental Table S3 The genome-sequenced ornamental species belong to 33 families.
    Supplemental Table S4 Comparison of different sequencing technologies.
    Supplemental Table S5 Overview of important genetically modified ornamental plants.
    Supplemental Table S6 List of current GWAS analysis in ornamental plant.
    Supplemental Fig. S1 Annual increase in the genome-sequenced method.
  • [1]

    Rand ES. 1866. Garden Flowers: How to Cultivate Them. A Treatise on the Culture of Hardy Ornamental Trees, Shrubs, Annuals, Herbaceous and Bedding Plants. Boston: J. E. Tilton and Company https://doi.org/10.5962/bhl.title.19601

    [2]

    Kapoor M. 2017. Managing ambient air quality using ornamental plants-an alternative approach. Universal Journal of Plant Science 5:1−9

    doi: 10.13189/ujps.2017.050101

    CrossRef   Google Scholar

    [3]

    Oloyede F. 2012. Survey of ornamental ferns, their morphology and uses for environmental protection, improvement and management. Ife Journal of Science 14:245−52

    Google Scholar

    [4]

    Guimarães R, Barros L, Carvalho AM, Ferreira ICFR. 2010. Studies on chemical constituents and bioactivity of Rosa micrantha: An alternative antioxidants source for food, pharmaceutical, or cosmetic applications. Journal of Agricultural and Food Chemistry 58:6277−84

    doi: 10.1021/jf101394w

    CrossRef   Google Scholar

    [5]

    Lubbe A, Verpoorte R. 2011. Cultivation of medicinal and aromatic plants for specialty industrial materials. Industrial Crops and Products 34:785−801

    doi: 10.1016/j.indcrop.2011.01.019

    CrossRef   Google Scholar

    [6]

    Mlcek J, Rop O. 2011. Fresh edible flowers of ornamental plants – A new source of nutraceutical foods. Trends in Food Science & Technology 22:561−69

    doi: 10.1016/j.jpgs.2011.04.006

    CrossRef   Google Scholar

    [7]

    Zhang Q, Chen W, Sun L, Zhao F, Huang B, et al. 2012. The genome of Prunus mume. Nature Communications 3:1318

    doi: 10.1038/ncomms2290

    CrossRef   Google Scholar

    [8]

    Saint-Oyant LH, Ruttink T, Hamama L, Kirov I, Lakhwani D, et al. 2018. A high-quality genome sequence of Rosa chinensis to elucidate ornamental traits. Nature Plants 4:473−84

    doi: 10.1038/s41477-018-0166-1

    CrossRef   Google Scholar

    [9]

    Raymond O, Gouzy J, Just J, Badouin H, Verdenaud M, et al. 2018. The Rosa genome provides new insights into the domestication of modern roses. Nature Genetics 50:772−77

    doi: 10.1038/s41588-018-0110-3

    CrossRef   Google Scholar

    [10]

    Song C, Liu Y, Song A, Dong G, Zhao H, et al. 2018. The Chrysanthemum nankingense genome provides insights into the evolution and diversification of chrysanthemum flowers and medicinal traits. Molecular Plant 11:1482−91

    doi: 10.1016/j.molp.2018.10.003

    CrossRef   Google Scholar

    [11]

    Chao YT, Chen WC, Chen CY, Ho HY, Yeh CH, et al. 2018. Chromosome-level assembly, genetic and physical mapping of Phalaenopsis aphrodite genome provides new insights into species adaptation and resources for orchid breeding. Plant Biotechnology Journal 16:2027−41

    doi: 10.1111/pbi.12936

    CrossRef   Google Scholar

    [12]

    Lv S, Cheng S, Wang Z, Li S, Jin X, et al. 2020. Draft genome of the famous ornamental plant Paeonia suffruticosa. Ecology and Evolution 10:4518−30

    doi: 10.1002/ece3.5965

    CrossRef   Google Scholar

    [13]

    Maxam AM, Gilbert W. 1977. A new method for sequencing DNA. PNAS 74:560−64

    doi: 10.1073/pnas.74.2.560

    CrossRef   Google Scholar

    [14]

    Sanger F, Nicklen S, Coulson AR. 1977. DNA sequencing with chain-terminating inhibitors. PNAS 74:5463−67

    doi: 10.1073/pnas.74.12.5463

    CrossRef   Google Scholar

    [15]

    The Arabidopsis Genome Initiative. 2000. Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature 408:796−815

    doi: 10.1038/35048692

    CrossRef   Google Scholar

    [16]

    International Rice Genome Sequencing Project, Sasaki T . 2005. The map-based sequence of the rice genome. Nature 436:793−800

    doi: 10.1038/nature03895

    CrossRef   Google Scholar

    [17]

    Schnable PS, Ware D, Fulton RS, Stein JC, Wei F, et al. 2009. The B73 maize genome: complexity, diversity, and dynamics. Science 326:1112−15

    doi: 10.1126/science.1178534

    CrossRef   Google Scholar

    [18]

    The International Peach Genome Initiative, Verde I, Abbott AG, Scalabrin S, Jung S, et al. 2013. The high-quality draft genome of peach (Prunus persica) identifies unique patterns of genetic diversity, domestication and genome evolution. Nature Genetics 45:487−94

    doi: 10.1038/ng.2586

    CrossRef   Google Scholar

    [19]

    van Dijk EL, Jaszczyszyn Y, Naquin D, Thermes C. 2018. The Third Revolution in Sequencing Technology. Trends in Genetics 34:666−81

    doi: 10.1016/j.tig.2018.05.008

    CrossRef   Google Scholar

    [20]

    Li C, Lin F, An D, Wang W, Huang R. 2017. Genome sequencing and assembly by long reads in plants. Genes 9:6

    doi: 10.3390/genes9010006

    CrossRef   Google Scholar

    [21]

    Schadt EE, Turner S, Kasarskis A. 2010. A window into third-generation sequencing. Human Molecular Genetics 19:R227−R240

    doi: 10.1093/hmg/ddq416

    CrossRef   Google Scholar

    [22]

    Rang FJ, Kloosterman WP, de Ridder J. 2018. From squiggle to basepair: computational approaches for improving nanopore sequencing read accuracy. Genome Biology 19:90

    doi: 10.1186/s13059-018-1462-9

    CrossRef   Google Scholar

    [23]

    Lu H, Giordano F, Ning Z. 2016. a3 Oxford nanopore MinION sequencing and genome assembly. Genomics Proteomics & Bioinformatics 14:265−79

    doi: 10.1016/j.gpb.2016.05.004

    CrossRef   Google Scholar

    [24]

    Schmidt MHW, Vogel A, Denton AK, Istace B, Wormit A, et al. 2017. De Novo assembly of a new Solanum pennellii accession using nanopore sequencing. The Plant Cell 29:2336−48

    doi: 10.1105/tpc.17.00521

    CrossRef   Google Scholar

    [25]

    Kumar KR, Cowley MJ, Davis RL. 2019. Next-generation sequencing and emerging technologies. Seminars in Thrombosis and Hemostasis 45:661−73

    doi: 10.1055/s-0039-1688446

    CrossRef   Google Scholar

    [26]

    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

    [27]

    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

    [28]

    Yang F, Nie S, Liu H, Shi T, Tian X, et al. 2020. Chromosome-level genome assembly of a parent species of widely cultivated azaleas. Nature Communications 11:5269

    doi: 10.1038/s41467-020-18771-4

    CrossRef   Google Scholar

    [29]

    Yang X, Tue Y, Li H, Ding W, Chen G, et al. 2018. The chromosome-level quality genome provides insights into the evolution of the biosynthesis genes for aroma compounds of Osmanthus fragrans. Horticulture Research 5:72

    doi: 10.1038/s41438-018-0108-0

    CrossRef   Google Scholar

    [30]

    Shang J, Tian J, Cheng H, Yan Q, Li L, et al. 2020. The chromosome-level wintersweet (Chimonanthus praecox) genome provides insights into floral scent biosynthesis and flowering in winter. Genome Biology 21:200

    doi: 10.1186/s13059-020-02088-y

    CrossRef   Google Scholar

    [31]

    Lv Q, Qiu J, Liu J, Li Z, Zhang W, et al. 2020. The Chimonanthus salicifolius genome provides insight into magnoliid evolution and flavonoid biosynthesis. Plant Journal 103:1910−23

    doi: 10.1111/tpj.14874

    CrossRef   Google Scholar

    [32]

    Xu W, Zhang Q, Yuan W, Xu F, Muhammad Aslam M, et al. 2020. The genome evolution and low-phosphorus adaptation in white lupin. Nature Communications 11:1069

    doi: 10.1038/s41467-020-14891-z

    CrossRef   Google Scholar

    [33]

    Hufnagel B, Marques A, Soriano A, Marquès L, Divol F, et al. 2020. High-quality genome sequence of white lupin provides insight into soil exploration and seed quality. Nature Communications 11:492

    doi: 10.1038/s41467-019-14197-9

    CrossRef   Google Scholar

    [34]

    Huala E, Dickerman AW, Garcia-Hernandez M, Weems D, Reiser L, et al. 2001. The Arabidopsis Information Resource (TAIR): a comprehensive database and web-based information retrieval, analysis, and visualization system for a model plant. Nucleic Acids Research 29:102−5

    doi: 10.1093/nar/29.1.102

    CrossRef   Google Scholar

    [35]

    Zhang G, Xu Q, Bian C, Tsai WC, Yeh CM, et al. 2016. The Dendrobium catenatum Lindl. genome sequence provides insights into polysaccharide synthase, floral development and adaptive evolution. Scientific Reports 6:19029

    doi: 10.1038/srep19029

    CrossRef   Google Scholar

    [36]

    Silva-Junior OB, Grattapaglia D, Novaes E, Collevatti RG. 2018. Genome assembly of the Pink Ipê (Handroanthus impetiginosus, Bignoniaceae), a highly valued, ecologically keystone Neotropical timber forest tree. GigaScience 7:gix125

    doi: 10.1093/gigascience/gix125

    CrossRef   Google Scholar

    [37]

    Yuan ZH, Fang YM, Zhang TK, Fei ZJ, Han FM, et al. 2018. The pomegranate (Punica granatum L.) genome provides insights into fruit quality and ovule developmental biology. Plant Biotechnology Journal 16:1363−74

    doi: 10.1111/pbi.12875

    CrossRef   Google Scholar

    [38]

    Li H, Yang X, Zhang Y, Gao Z, Liang Y, et al. 2021. Nelumbo genome database, an integrative resource for gene expression and variants of Nelumbo nucifera. Scientific Data 8:38

    doi: 10.1038/s41597-021-00828-8

    CrossRef   Google Scholar

    [39]

    Hoshino A, Jayakumar V, Nitasaka E, Toyoda A, Noguchi H, et al. 2016. Genome sequence and analysis of the Japanese morning glory Ipomoea nil. Nature Communications 7:13295

    doi: 10.1038/ncomms13295

    CrossRef   Google Scholar

    [40]

    Yagi M, Kosugi S, Hirakawa H, Ohmiya A, Tanase K, et al. 2014. Sequence Analysis of the Genome of Carnation (Dianthus caryophyllus L.). DNA Research 21:231−41

    doi: 10.1093/dnares/dst053

    CrossRef   Google Scholar

    [41]

    Jung S, Lee T, Cheng C, Buble K, Zheng P, et al. 2019. 15 years of GDR: New data and functionality in the Genome Database for Rosaceae. Nucleic Acids Research 47:D1137−D1145

    doi: 10.1093/nar/gky1000

    CrossRef   Google Scholar

    [42]

    Goodstein DM, Shu SQ, Howson R, Neupane R, Hayes RD, et al. 2012. Phytozome: a comparative platform for green plant genomics. Nucleic Acids Research 40:D1178−D1186

    doi: 10.1093/nar/gkr944

    CrossRef   Google Scholar

    [43]

    Lee TH, Tang H, Wang X, Paterson AH. 2013. PGDD: a database of gene and genome duplication in plants. Nucleic Acids Research 41:D1152−D1158

    doi: 10.1093/nar/gks1104

    CrossRef   Google Scholar

    [44]

    Proost S, Van Bel M, Sterck L, Billiau K, Van Parys T, et al. 2009. PLAZA: a comparative genomics resource to study gene and genome evolution in plants. The Plant Cell 21:3718−31

    doi: 10.1105/tpc.109.071506

    CrossRef   Google Scholar

    [45]

    Kersey PJ, Allen JE, Christensen M, Davis P, Falin LJ, et al. 2014. Ensembl Genomes 2013: scaling up access to genome-wide data. Nucleic Acids Research 42:D546−D552

    doi: 10.1093/nar/gkt979

    CrossRef   Google Scholar

    [46]

    Chen F, Dong W, Zhang J, Guo X, Chen J, et al. 2018. The sequenced angiosperm genomes and genome databases. Frontiers in Plant Science 9:418

    doi: 10.3389/fpls.2018.00418

    CrossRef   Google Scholar

    [47]

    Li Y, Li L, Ding W, Li H, Shi T, et al. 2020. Genome-wide identification of Osmanthus fragrans bHLH transcription factors and their expression analysis in response to abiotic stress. Environmental and Experimental Botany 172:103990

    doi: 10.1016/j.envexpbot.2020.103990

    CrossRef   Google Scholar

    [48]

    Zhang Q, Zhang H, Sun L, Fan G, Ye M, et al. 2018. The genetic architecture of floral traits in the woody plant Prunus mume. Nature Communications 9:1702

    doi: 10.1038/s41467-018-04093-z

    CrossRef   Google Scholar

    [49]

    Guo L, Winzer T, Yang X, Li Y, Ning Z, et al. 2018. The opium poppy genome and morphinan production. Science 362:343

    doi: 10.1126/science.aat4096

    CrossRef   Google Scholar

    [50]

    Otto SP. 2007. The evolutionary consequences of polyploidy. Cell 131:452−62

    doi: 10.1016/j.cell.2007.10.022

    CrossRef   Google Scholar

    [51]

    Freeling M. 2009. Bias in plant gene content following different sorts of duplication: tandem, whole-genome, segmental, or by transposition. Annual Review of Plant Biology 60:433−53

    doi: 10.1146/annurev.arplant.043008.092122

    CrossRef   Google Scholar

    [52]

    Melters DP, Bradnam KR, Young HA, Telis N, May MR, et al. 2013. Comparative analysis of tandem repeats from hundreds of species reveals unique insights into centromere evolution. Genome Biology 14:R10

    doi: 10.1186/gb-2013-14-1-r10

    CrossRef   Google Scholar

    [53]

    Huang CY, Chen P, Huang MD, Tsou CH, Jane WN, et al. 2013. Tandem oleosin genes in a cluster acquired in Brassicaceae created tapetosomes and conferred additive benefit of pollen vigor. PNAS 110:14480−85

    doi: 10.1073/pnas.1305299110

    CrossRef   Google Scholar

    [54]

    Zhang R, Xue C, Liu G, Liu X, Zhang M, et al. 2017. Segmental duplication of chromosome 11 and its implications for cell division and genome-wide expression in rice. Scientific Reports 7:2689

    doi: 10.1038/s41598-017-02796-9

    CrossRef   Google Scholar

    [55]

    Zhang Z, Yuan L, Liu X, Chen X, Wang X. 2018. Evolution analysis of Dof transcription factor family and their expression in response to multiple abiotic stresses in Malus domestica. Gene 639:137−48

    doi: 10.1016/j.gene.2017.09.039

    CrossRef   Google Scholar

    [56]

    Zou Z, Zhu J, Zhang X. 2019. Genome-wide identification and characterization of the Dof gene family in cassava (Manihot esculenta). Gene 687:298−307

    doi: 10.1016/j.gene.2018.11.053

    CrossRef   Google Scholar

    [57]

    Nan H, Ludlow RA, Lu M, An H. 2021. Genome-wide analysis of Dof genes and their response to abiotic stress in rose (Rosa chinensis). Frontiers in Genetics 12:53873

    doi: 10.3389/fgene.2021.538733

    CrossRef   Google Scholar

    [58]

    Zhang S, Liu L, Yang R, Wang X. 2020. Genome size evolution mediated by Gypsy retrotransposons in brassicaceae. Genomics Proteomics & Bioinformatics 18:321−32

    doi: 10.1016/j.gpb.2018.07.009

    CrossRef   Google Scholar

    [59]

    Hsu C, Chen SY, Lai PH, Hsiao YY, Tsai WC, et al. 2020. Identification of high-copy number long terminal repeat retrotransposons and their expansion in Phalaenopsis orchids. BMC Genomics 21:807

    doi: 10.1186/s12864-020-07221-6

    CrossRef   Google Scholar

    [60]

    De Smet R, Van de Peer Y. 2012. Redundancy and rewiring of genetic networks following genome-wide duplication events. Current Opinion in Plant Biology 15:168−76

    doi: 10.1016/j.pbi.2012.01.003

    CrossRef   Google Scholar

    [61]

    Kim YM, Kim S, Koo N, Shin AY, Yeom SI, et al. 2017. Genome analysis of Hibiscus syriacus provides insights of polyploidization and indeterminate flowering in woody plants. DNA Research 24:71−80

    doi: 10.1093/dnares/dsw049

    CrossRef   Google Scholar

    [62]

    Badouin H, Gouzy J, Grassa CJ, Murat F, Staton SE, et al. 2017. The sunflower genome provides insights into oil metabolism, flowering and Asterid evolution. Nature 546:148−52

    doi: 10.1038/nature22380

    CrossRef   Google Scholar

    [63]

    Zhang L, Chen F, Zhang X, Li Z, Zhao Y, et al. 2020. The water lily genome and the early evolution of flowering plants. Nature 577:79

    doi: 10.1038/s41586-019-1852-5

    CrossRef   Google Scholar

    [64]

    Qin G, Xu C, Ming R, Tang H, Guyot R, et al. 2017. The pomegranate (Punica granatum L.) genome and the genomics of punicalagin biosynthesis. The Plant Journal 91:1108−28

    doi: 10.1111/tpj.13625

    CrossRef   Google Scholar

    [65]

    Wai CM, Weise SE, Ozersky P, Mockler TC, Michael TP, et al. 2019. Time of day and network reprogramming during drought induced CAM photosynthesis in Sedum album. Plos Genetics 15:e1008209

    doi: 10.1371/journal.pgen.1008209

    CrossRef   Google Scholar

    [66]

    Sjödin P, Hedman H, Shavorskaya O, Finet C, Lascoux M, et al. 2007. Recent degeneration of an old duplicated flowering time gene in Brassica nigra. Heredity 98:375−84

    doi: 10.1038/sj.hdy.6800951

    CrossRef   Google Scholar

    [67]

    Salman-Minkov A, Sabath N, Mayrose I. 2016. Whole-genome duplication as a key factor in crop domestication. Nature Plants 2:16115

    doi: 10.1038/nplants.2016.115

    CrossRef   Google Scholar

    [68]

    Seelanan T, Brubaker CL, Stewart JM, Craven LA, Wendel JF. 1999. Molecular systematics of Australian Gossypium section Grandicalyx (Malvaceae). Systematic Botany 24:183−208

    doi: 10.2307/2419548

    CrossRef   Google Scholar

    [69]

    Caicedo AL, Schaal BA. 2004. Population structure and phylogeography of Solanum pimpinellifolium inferred from a nuclear gene. Molecular Ecology 13:1871−82

    doi: 10.1111/j.1365-294X.2004.02191.x

    CrossRef   Google Scholar

    [70]

    Matsuoka Y, Vigouroux Y, Goodman MM, Sanchez GJ, Buckler E, et al. 2002. A single domestication for maize shown by multilocus microsatellite genotyping. PNAS 99:6080−4

    doi: 10.1073/pnas.052125199

    CrossRef   Google Scholar

    [71]

    Li C, Zhou A, Sang T. 2006. Rice domestication by reducing shattering. Science 311:1936−39

    doi: 10.1126/science.1123604

    CrossRef   Google Scholar

    [72]

    Vigouroux Y, Mitchell S, Matsuoka Y, Hamblin M, Kresovich S, et al. 2005. An analysis of genetic diversity across the maize genome using microsatellites. Genetics 169:1617−30

    doi: 10.1534/genetics.104.032086

    CrossRef   Google Scholar

    [73]

    Faust M, Timon B. 1995. Origin and dissemination of peach. In Horticultural Reviews, ed. Janick J, vol 17. New York: John Wiley & Sons. pp. 331−79 https://doi.org/10.1002/9780470650585.ch10

    [74]

    Yi X, Yu X, Chen J, Zhang M, Liu S, et al. 2020. The genome of Chinese flowering cherry (Cerasus serrulata) provides new insights into Cerasus species. Horticulture Research 7:165

    doi: 10.1038/s41438-020-00382-1

    CrossRef   Google Scholar

    [75]

    Huang L, Yang M, Li L, Li H, Yang D, et al. 2018. Whole genome re-sequencing reveals evolutionary patterns of sacred lotus (Nelumbo nucifera). Journal of Integrative Plant Biology 60:2−15

    doi: 10.1111/jipb.12606

    CrossRef   Google Scholar

    [76]

    Liu Z, Zhu H, Zhou J, Jiang S, Wang Y, et al. 2020. Resequencing of 296 cultivated and wild lotus accessions unravels its evolution and breeding history. The Plant Journal 104:1673−84

    doi: 10.1111/tpj.15029

    CrossRef   Google Scholar

    [77]

    Huang X, Liu J, Zhu W, Atzberger C, Liu Q. 2019. The optimal threshold and vegetation index time series for retrieving crop phenology based on a modified dynamic threshold method. Remote Sensing 11:2725

    doi: 10.3390/rs11232725

    CrossRef   Google Scholar

    [78]

    da Silveira Falavigna V, Guitton B, Costes E, Andrés F. 2019. I want to (Bud) break free: the potential role of DAM and SVP-like genes in regulating dormancy cycle in temperate fruit trees. Frontiers in Plant Science 9:1900

    doi: 10.3389/fpls.2018.01990

    CrossRef   Google Scholar

    [79]

    Zhang Y, Yu D, Liu C, Gai S. 2018. Dynamic of carbohydrate metabolism and the related genes highlights PPP pathway activation during chilling induced bud dormancy release in tree peony (Paeonia suffruticosa). Scientia Horticulturae 242:36−43

    doi: 10.1016/j.scienta.2018.07.022

    CrossRef   Google Scholar

    [80]

    Zhong W, Gao Z, Zhuang W, Shi T, Zhang Z, et al. 2013. Genome-wide expression profiles of seasonal bud dormancy at four critical stages in Japanese apricot. Plant Molecular Biology 83:247−64

    doi: 10.1007/s11103-013-0086-4

    CrossRef   Google Scholar

    [81]

    de la Fuente L, Conesa A, Lloret A, Badenes ML, Ríos G. 2015. Genome-wide changes in histone H3 lysine 27 trimethylation associated with bud dormancy release in peach. Tree Genetics & Genomes 11:45

    doi: 10.1007/s11295-015-0869-7

    CrossRef   Google Scholar

    [82]

    Wisniewski M, Norelli J, Bassett C, Artlip T, Macarisin D. 2011. Ectopic expression of a novel peach (Prunus persica) CBF transcription factor in apple (Malus × domestica) results in short-day induced dormancy and increased cold hardiness. Planta 233:971−83

    doi: 10.1007/s00425-011-1358-3

    CrossRef   Google Scholar

    [83]

    Yamane H, Kashiwa Y, Ooka T, Tao R, Yonemori K. 2008. Suppression subtractive hybridization and differential screening reveals endodormancy-associated expression of an SVP/AGL24-type MADS-box gene in lateral vegetative buds of Japanese apricot. Journal of the American Society for Horticultural Science 133:708−16

    doi: 10.21273/JASHS.133.5.708

    CrossRef   Google Scholar

    [84]

    Sasaki R, Yamane H, Ooka T, Jotatsu H, Kitamura Y, et al. 2011. Functional and expressional analyses of PmDAM genes associated with endodormancy in Japanese apricot. Plant Physiology 157:485−97

    doi: 10.1104/pp.111.181982

    CrossRef   Google Scholar

    [85]

    Fornara F, de Montaigu A, Coupland G. 2010. SnapShot: Control of flowering in Arabidopsis. Cell 141:550

    doi: 10.1016/j.cell.2010.04.024

    CrossRef   Google Scholar

    [86]

    Fujiwara S, Nakagawa M, Oda A, Kato K, Mizoguchi T. 2010. Photoperiod pathway regulates expression of MAF5 and FLC that encode MADS-box transcription factors of the FLC family in Arabidopsis. Plant Biotechnology 27:447−54

    doi: 10.5511/plantbiotechnology.10.0823b

    CrossRef   Google Scholar

    [87]

    Wang B, Zhang Q, Wang L, Duan K, Pan A, et al. 2011. The AGL6-like Gene CpAGL6, a Potential Regulator of Floral Time and Organ Identity in Wintersweet (Chimonanthus praecox). Journal of Plant Growth Regulation 30:343−52

    doi: 10.1007/s00344-011-9196-x

    CrossRef   Google Scholar

    [88]

    Schaffer R, Ramsay N, Samach A, Corden S, Putterill J, et al. 1998. The late elongated hypocotyl mutation of Arabidopsis disrupts circadian rhythms and the photoperiodic control of flowering. Cell 93:1219−29

    doi: 10.1016/S0092-8674(00)81465-8

    CrossRef   Google Scholar

    [89]

    Hung FY, Chen FF, Li CL, Chen C, Chen JH, et al. 2019. The LDL1/2-HDA6 histone modification complex interacts with TOC1 and regulates the core circadian clock components in Arabidopsis. Frontiers in Plant Science 10:233

    doi: 10.3389/fpls.2019.00233

    CrossRef   Google Scholar

    [90]

    Li J, Pan B, Niu L, Chen M, Tang M, et al. 2018. Gibberellin Inhibits Floral Initiation in the Perennial Woody Plant Jatropha curcas. Journal of Plant Growth Regulation 37:999−1006

    doi: 10.1007/s00344-018-9797-8

    CrossRef   Google Scholar

    [91]

    Zentella R, Zhang Z, Park M, Thomas SG, Endo A, et al. 2007. Global analysis of DELLA direct targets in early gibberellin signaling in Arabidopsis. The Plant Cell 19:3037−57

    doi: 10.1105/tpc.107.054999

    CrossRef   Google Scholar

    [92]

    Lu J, Yang W, Zhang Q. 2015. Genome-wide Identification and Characterization of the DELLA Subfamily in Prunus mume. Journal of the American Society for Horticultural Science 140:223−32

    doi: 10.21273/jashs.140.3.223

    CrossRef   Google Scholar

    [93]

    Lee JH, Yoo SJ, Park SH, Hwang I, Lee JS, et al. 2007. Role of SVP in the control of flowering time by ambient temperature in Arabidopsis. Genes & Development 21:397−402

    doi: 10.1101/gad.1518407

    CrossRef   Google Scholar

    [94]

    Wang J. 2014. Regulation of flowering time by the miR156-mediated age pathway. Journal of Experimental Botany 65:4723−30

    doi: 10.1093/jxb/eru246

    CrossRef   Google Scholar

    [95]

    Wei Q, Ma C, Xu Y, Wang T, Chen Y, et al. 2017. Control of chrysanthemum flowering through integration with an aging pathway. Nature Communications 8:829

    doi: 10.1038/s41467-017-00812-0

    CrossRef   Google Scholar

    [96]

    Wang S, Beruto M, Xue J, Zhu F, Liu C, et al. 2015. Molecular cloning and potential function prediction of homologous SOC1 genes in tree peony. Plant Cell Reports 34:1459−71

    doi: 10.1007/s00299-015-1800-2

    CrossRef   Google Scholar

    [97]

    Shibuya K. 2018. Molecular aspects of flower senescence and strategies to improve flower longevity. Breeding Science 68:99−108

    doi: 10.1270/jsbbs.17081

    CrossRef   Google Scholar

    [98]

    Wojciechowska N, Sobieszczuk-Nowicka E, Bagniewska-Zadworna A. 2018. Plant organ senescence - regulation by manifold pathways. Plant Biology 20:167−81

    doi: 10.1111/plb.12672

    CrossRef   Google Scholar

    [99]

    Pattyn J, Vaughan-Hirsch J, Van de Poel B. 2021. The regulation of ethylene biosynthesis: a complex multilevel control circuitry. New Phytologist 229:770−82

    doi: 10.1111/nph.16873

    CrossRef   Google Scholar

    [100]

    Yangkhamman P, Tanase K, Ichimura K, Fukai S. 2007. Depression of enzyme activities and gene expression of ACC synthase and ACC oxidase in cut carnation flowers under high-temperature conditions. Plant Growth Regulation 53:155−62

    doi: 10.1007/s10725-007-9213-z

    CrossRef   Google Scholar

    [101]

    Rice LJ, Soós V, Ascough GD, Balázs E, Ördög V, et al. 2013. Ethylene- and dark-induced flower abscission in potted Plectranthus: Sensitivity, prevention by 1-MCP, and expression of ethylene biosynthetic genes. South African Journal of Botany 87:39−47

    doi: 10.1016/j.sajb.2013.03.011

    CrossRef   Google Scholar

    [102]

    Huang LC, Lai UL, Yang SF, Chu MJ, Kuo CI, et al. 2007. Delayed flower senescence of Petunia hybrida plants transformed with antisense broccoli ACC synthase and ACC oxidase genes. Postharvest Biology and Technology 46:47−53

    doi: 10.1016/j.postharvbio.2007.03.015

    CrossRef   Google Scholar

    [103]

    Shibuya K, Watanabe K, Ono M. 2018. CRISPR/Cas9-mediated mutagenesis of the EPHEMERAL1 locus that regulates petal senescence in Japanese morning glory. Plant Physiology and Biochemistry 131:53−57

    doi: 10.1016/j.plaphy.2018.04.036

    CrossRef   Google Scholar

    [104]

    Zhou L, Zhang C, Fu J, Liu M, Zhang Y, et al. 2013. Molecular characterization and expression of ethylene biosynthetic genes during cut flower development in tree peony (Paeonia suffruticosa) in response to ethylene and functional analysis of PsACS1 in Arabidopsis thaliana. Journal of Plant Growth Regulation 32:362−75

    doi: 10.1007/s00344-012-9306-4

    CrossRef   Google Scholar

    [105]

    Satoh S. 2011. Ethylene production and petal wilting during senescence of cut carnation (Dianthus caryophyllus) flowers and prolonging their vase life by genetic transformation. Journal of the Japanese Society for Horticultural Science 80:127−35

    doi: 10.2503/jjshs1.80.127

    CrossRef   Google Scholar

    [106]

    Kinouchi T, Endo R, Yamashita A, Satoh S. 2006. Transformation of carnation with genes related to ethylene production and perception: towards generation of potted carnations with a longer display time. Plant Cell, Tissue and Organ Culture 86:27

    doi: 10.1007/s11240-006-9093-3

    CrossRef   Google Scholar

    [107]

    Woltering EJ, Van Doorn WG. 1988. Role of ethylene in senescence of petals - morphological and taxonomical relationships. Journal of Experimental Botany 39:1605−16

    doi: 10.1093/jxb/39.11.1605

    CrossRef   Google Scholar

    [108]

    Shahri W, Tahir I. 2011. Flower Senescence-Strategies and Some Associated Events. The Botanical Review 77:152−84

    doi: 10.1007/s12229-011-9063-2

    CrossRef   Google Scholar

    [109]

    Kumar M, Singh VP, Arora A, Singh N. 2014. The role of abscisic acid (ABA) in ethylene insensitive Gladiolus (Gladiolus grandiflora Hort.) flower senescence. Acta Physiologiae Plantarum 36:151−59

    doi: 10.1007/s11738-013-1395-6

    CrossRef   Google Scholar

    [110]

    Arrom L, Munné-Bosch S. 2012. Hormonal changes during flower development in floral tissues of Lilium. Planta 236:343−54

    doi: 10.1007/s00425-012-1615-0

    CrossRef   Google Scholar

    [111]

    Iqbal N, Khan NA, Ferrante A, Trivellini A, Francini A, et al. 2017. Ethylene role in plant growth, development and senescence: interaction with other phytohormones. Frontiers in Plant Science 8:475

    doi: 10.3389/fpls.2017.00475

    CrossRef   Google Scholar

    [112]

    Marciniak K, Kućko A, Wilmowicz E, Świdziński M, Przedniczek K, et al. 2018. Gibberellic acid affects the functioning of the flower abscission zone in Lupinus luteus via cooperation with the ethylene precursor independently of abscisic acid. Journal of Plant Physiology 229:170−74

    doi: 10.1016/j.jplph.2018.07.014

    CrossRef   Google Scholar

    [113]

    Lü P, Zhang C, Liu J, Liu X, Jiang G, et al. 2014. RhHB1 mediates the antagonism of gibberellins to ABA and ethylene during rose (Rosa hybrida) petal senescence. The Plant Journal 78:578−90

    doi: 10.1111/tpj.12494

    CrossRef   Google Scholar

    [114]

    Manohar M, Wang D, Manosalva PM, Choi HW, Kombrink E, Klessig DF. 2017. Members of the abscisic acid co-receptor PP2C protein family mediate salicylic acid-abscisic acid crosstalk. Plant Direct 1:e00020

    doi: 10.1002/pld3.20

    CrossRef   Google Scholar

    [115]

    Hermann K, Kuhlemeier C. 2011. The genetic architecture of natural variation in flower morphology. Current Opinion in Plant Biology 14:60−65

    doi: 10.1016/j.pbi.2010.09.012

    CrossRef   Google Scholar

    [116]

    Coito JL, Silva H, Ramos MJN, Montez M, Cunha J, et al. 2018. Vitis Flower Sex Specification Acts Downstream and Independently of the ABCDE Model Genes. Frontiers in Plant Science 9:1029

    doi: 10.3389/fpls.2018.01029

    CrossRef   Google Scholar

    [117]

    Coen ES, Meyerowitz EM. 1991. The war of the whorls - genetic interactions controlling flower development. Nature 353:31−37

    doi: 10.1038/353031a0

    CrossRef   Google Scholar

    [118]

    Theißen G. 2001. Development of floral organ identity: stories from the MADS house. Current Opinion in Plant Biology 4:75−85

    doi: 10.1016/S1369-5266(00)00139-4

    CrossRef   Google Scholar

    [119]

    Theißen G, Saedler H. 2001. Floral quartets. Nature 409:469−71

    doi: 10.1038/35054172

    CrossRef   Google Scholar

    [120]

    Tsai WC, Kuoh CS, Chuang MH, Chen WH, Chen HH. 2004. Four DEF-Like MADS box genes displayed distinct floral morphogenetic roles in Phaldenopsis orchid. Plant and Cell Physiology 45:831−44

    doi: 10.1093/pcp/pch095

    CrossRef   Google Scholar

    [121]

    Ishimori M, Kawabata S. 2014. Conservation and diversification of floral homeotic MADS-box genes in Eustoma grandiflorum. Journal of the Japanese Society for Horticultural Science 83:172−80

    doi: 10.2503/jjshs1.ch-098

    CrossRef   Google Scholar

    [122]

    Sasaki K, Yamaguchi H, Nakayama M, Aida R, Ohtsubo N. 2014. Co-modification of class B genes TfDEF and TfGLO in Torenia fournieri Lind. alters both flower morphology and inflorescence architecture. Plant Molecular Biology 86:319−34

    doi: 10.1007/s11103-014-0231-8

    CrossRef   Google Scholar

    [123]

    Sawettalake N, Bunnag S, Wang Y, Shen L, Yu H. 2017. DOAP1 promotes flowering in the orchid Dendrobium chao praya smile. Frontiers in plant science 8:400

    doi: 10.3389/fpls.2017.00400

    CrossRef   Google Scholar

    [124]

    Shchennikova AV, Shulga OA, Skryabin KG. 2018. Ectopic expression of the homeotic MADS-Box gene HAM31 (Helianthus annuus L.) in transgenic plants Nicotiana tabacum L. affect the gynoecium identity. Doklady Biochemistry and Biophysics 483:363−68

    doi: 10.1134/S1607672918060182

    CrossRef   Google Scholar

    [125]

    Wang Q, Dan N, Zhang X, Lin S, Bao M, et al. 2020. Identification, Characterization and Functional Analysis of C-Class Genes Associated with Double Flower Trait in Carnation (Dianthus caryphyllus L.). Plants-Basel 9:87

    doi: 10.3390/plants9010087

    CrossRef   Google Scholar

    [126]

    Chen YY, Lee PF, Hsiao YY, Wu WL, Pan ZJ, et al. 2012. C- and D-class MADS-Box Genes from Phalaenopsis equestris (Orchidaceae) Display Functions in Gynostemium and Ovule Development. Plant and Cell Physiology 53:1053−67

    doi: 10.1093/pcp/pcs048

    CrossRef   Google Scholar

    [127]

    Xu Y, Zhang L, Xie H, Zhang YQ, Oliveira MM, Ma RC. 2008. Expression analysis and genetic mapping of three SEPALLATA-like genes from peach (Prunus persica (L.) Batsch). Tree Genetics & Genomes 4:693−703

    doi: 10.1007/s11295-008-0143-3

    CrossRef   Google Scholar

    [128]

    Sasaki K, Yoshioka S, Aida R, Ohtsubo N. 2021. Production of petaloid phenotype in the reproductive organs of compound flowerheads by the co-suppression of class-C genes in hexaploid Chrysanthemum morifolium. Planta 253:100

    doi: 10.1007/s00425-021-03605-4

    CrossRef   Google Scholar

    [129]

    Pan ZJ, Chen YY, Du JS, Chen YY, Chung MC, et al. 2014. Flower development of Phalaenopsis orchid involves functionally divergent SEPALLATA-like genes. New Phytologist 202:1024−42

    doi: 10.1111/nph.12723

    CrossRef   Google Scholar

    [130]

    Ishii HS, Harder LD. 2006. The size of individual Delphinium flowers and the opportunity for geitonogamous pollination. Functional Ecology 20:1115−23

    doi: 10.1111/j.1365-2435.2006.01181.x

    CrossRef   Google Scholar

    [131]

    Chen QXC, Guo Y, Warner RM. 2019. Identification of Quantitative Trait Loci for Component Traits of Flowering Capacity Across Temperature in Petunia. G3 Genes Genomes Genetics 9:3601−10

    doi: 10.1534/g3.119.400653

    CrossRef   Google Scholar

    [132]

    Galliot C, Hoballah ME, Kuhlemeier C, Stuurman J. 2006. Genetics of flower size and nectar volume in Petunia pollination syndromes. Planta 225:203−12

    doi: 10.1007/s00425-006-0342-9

    CrossRef   Google Scholar

    [133]

    Reinert S, Gao Q, Ferguson B, Portlas ZM, Prasifka JR, et al. 2020. Seed and floret size parameters of sunflower are determined by partially overlapping sets of quantitative trait loci with epistatic interactions. Molecular Genetics and Genomics 295:143−54

    doi: 10.1007/s00438-019-01610-7

    CrossRef   Google Scholar

    [134]

    Moyers BT, Owens GL, Baute GJ, Rieseberg LH. 2017. The genetic architecture of UV floral patterning in sunflower. Annals of Botany 120:39−50

    doi: 10.1093/aob/mcx038

    CrossRef   Google Scholar

    [135]

    Zhu R, Gao Y, Zhang Q. 2014. Quantitative trait locus mapping of floral and related traits using an F-2 population of Aquilegia. Plant Breeding 133:153−61

    doi: 10.1111/pbr.12128

    CrossRef   Google Scholar

    [136]

    Han Y, Tang A, Wan H, Zhang T, Cheng T, et al. 2018. An APETALA2 Homolog, RcAP2, regulates the number of rose petals derived from stamens and response to temperature fluctuations. Frontiers in Plant Science 9:481

    doi: 10.3389/fpls.2018.00481

    CrossRef   Google Scholar

    [137]

    Bourke PM, Gitonga VW, Voorrips RE, Visser RGF, Krens FA, et al. 2018. Multi-environment QTL analysis of plant and flower morphological traits in tetraploid rose. Theoretical and Applied Genetics 131:2055−69

    doi: 10.1007/s00122-018-3132-4

    CrossRef   Google Scholar

    [138]

    Meng G, Zhu G, Fang W, Chen C, Wang X, et al. 2019. Identification of loci for single/double flower trait by combining genome-wide association analysis and bulked segregant analysis in peach (Prunus persica). Plant Breeding 138:360−67

    doi: 10.1111/pbr.12673

    CrossRef   Google Scholar

    [139]

    Lucibelli F, Valoroso MC, Aceto S. 2020. Radial or Bilateral? The molecular basis of floral symmetry Genes 11:395

    doi: 10.3390/genes11040395

    CrossRef   Google Scholar

    [140]

    Garcês HMP, Spencer VMR, Kim M. 2016. Control of floret symmetry by RAY3, SvDIV1B, and SvRAD in the capitulum of Senecio vulgaris. Plant Physiology 171:2055−68

    doi: 10.1104/pp.16.00395

    CrossRef   Google Scholar

    [141]

    Feng X, Zhao Z, Tian Z, Xu S, Luo Y, et al. 2006. Control of petal shape and floral zygomorphy in Lotus japonicus. PNAS 103:4970−75

    doi: 10.1073/pnas.0600681103

    CrossRef   Google Scholar

    [142]

    Wang Z, Luo Y, Li X, Wang L, Xu S, et al. 2008. Genetic control of floral zygomorphy in pea (Pisum sativum L.). PNAS 105:10414−19

    doi: 10.1073/pnas.0803291105

    CrossRef   Google Scholar

    [143]

    Hileman LC. 2014. Trends in flower symmetry evolution revealed through phylogenetic and developmental genetic advances. Philosophical Transactions of the Royal Society B-Biological Sciences 369:20130348

    doi: 10.1098/rstb.2013.0348

    CrossRef   Google Scholar

    [144]

    Sengupta A, Hileman LC. 2018. Novel Traits, Flower Symmetry, and Transcriptional Autoregulation: New Hypotheses From Bioinformatic and Experimental Data. Frontiers in Plant Science 9:1561

    doi: 10.3389/fpls.2018.01561

    CrossRef   Google Scholar

    [145]

    Raimundo J, Sobral R, Bailey P, Azevedo H, Galego L, et al. 2013. A subcellular tug of war involving three MYB-like proteins underlies a molecular antagonism in Antirrhinum flower asymmetry. The Plant Journal 75:527−38

    doi: 10.1111/tpj.12225

    CrossRef   Google Scholar

    [146]

    Yuan C, Huang D, Yang Y, Sun M, Cheng TR, et al. 2020. CmCYC2-like transcription factors may interact with each other or bind to the promoter to regulate floral symmetry development in Chrysanthemum morifolium. Plant Molecular Biology 103:159−71

    doi: 10.1007/s11103-020-00981-5

    CrossRef   Google Scholar

    [147]

    Specht CD, Howarth DG. 2015. Adaptation in flower form: a comparative evodevo approach. New Phytologist 206:74−90

    doi: 10.1111/nph.13198

    CrossRef   Google Scholar

    [148]

    Moyroud E, Glover BJ. 2017. The evolution of diverse floral morphologies. Current Biology 27:R941−R951

    doi: 10.1016/j.cub.2017.06.053

    CrossRef   Google Scholar

    [149]

    Jasinski S, Vialette-Guiraud ACM, Scutt CP. 2010. The evolutionary-developmental analysis of plant microRNAs. Philosophical Transactions of the Royal Society B-Biological Sciences 365:469−76

    doi: 10.1098/rstb.2009.0246

    CrossRef   Google Scholar

    [150]

    Aida M, Ishida T, Fukaki H, Fujisawa H, Tasaka M. 1997. Genes involved in organ separation in Arabidopsis: An analysis of the cup-shaped cotyledon mutant. The Plant Cell 9:841−57

    doi: 10.1105/tpc.9.6.841

    CrossRef   Google Scholar

    [151]

    Mallory AC, Dugas DV, Bartel DP, Bartel B. 2004. MicroRNA regulation of NAC-domain targets is required for proper formation and separation of adjacent embryonic, vegetative, and floral organs. Current Biology 14:1035−46

    doi: 10.1016/j.cub.2004.06.022

    CrossRef   Google Scholar

    [152]

    Weir I, Lu J, Cook H, Causier B, Schwarz-Sommer Z, et al. 2004. CUPULIFORMIS establishes lateral organ boundaries in Antirrhinum. Development 131:915−22

    doi: 10.1242/dev.00993

    CrossRef   Google Scholar

    [153]

    Zhong J, Powell S, Preston JC. 2016. Organ boundary NAC-domain transcription factors are implicated in the evolution of petal fusion. Plant Biology 18:893−902

    doi: 10.1111/plb.12493

    CrossRef   Google Scholar

    [154]

    Laufs P, Peaucelle A, Morin H, Traas J. 2004. MicroRNA regulation of the CUC genes is required for boundary size control in Arabidopsis meristems. Development 131:4311−22

    doi: 10.1242/dev.01320

    CrossRef   Google Scholar

    [155]

    Martins AE, Camargo MGG, Morellato LPC. 2021. Flowering phenology and the influence of seasonality in flower conspicuousness for bees. Frontiers in Plant Science 11:594538

    doi: 10.3389/fpls.2020.594538

    CrossRef   Google Scholar

    [156]

    Holton TA, Tanaka Y. 1994. Blue roses — a pigment of our imagination. Trends in Biotechnology 12:40−42

    doi: 10.1016/0167-7799(94)90097-3

    CrossRef   Google Scholar

    [157]

    Hoshino A, Mizuno T, Shimizu K, Mori S, Fukada-Tanaka S, et al. 2019. Generation of yellow flowers of the japanese morning glory by engineering its flavonoid biosynthetic pathway toward aurones. Plant and Cell Physiology 60:1871−79

    doi: 10.1093/pcp/pcz101

    CrossRef   Google Scholar

    [158]

    Johnson ET, Yi HK, Shin BC, Oh BJ, Cheong HS, et al. 1999. Cymbidium hybrida dihydroflavonol 4-reductase does not efficiently reduce dihydrokaempferol to produce orange pelargonidin-type anthocyanins. Plant Journal 19:81−85

    doi: 10.1046/j.1365-313X.1999.00502.x

    CrossRef   Google Scholar

    [159]

    Grotewold E. 2006. The genetics and biochemistry of floral pigments. Annual Review of Plant Biology 57:761−80

    doi: 10.1146/annurev.arplant.57.032905.105248

    CrossRef   Google Scholar

    [160]

    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

    [161]

    Xu W, Dubos C, Lepiniec L. 2015. Transcriptional control of flavonoid biosynthesis by MYB-bHLH-WDR complexes. Trends in Plant Science 20:176−85

    doi: 10.1016/j.tplants.2014.12.001

    CrossRef   Google Scholar

    [162]

    Zhao D, Tao J. 2015. Recent advances on the development and regulation of flower color in ornamental plants. Frontiers in Plant Science 6:261

    doi: 10.3389/fpls.2015.00261

    CrossRef   Google Scholar

    [163]

    Lou Q, Liu Y, Qi Y, Jiao S, Tian F, et al. 2014. Transcriptome sequencing and metabolite analysis reveals the role of delphinidin metabolism in flower colour in grape hyacinth. Journal of Experimental Botany 65:3157−64

    doi: 10.1093/jxb/eru168

    CrossRef   Google Scholar

    [164]

    Hong Y, Tang X, Huang H, Zhang Y, Dai S. 2015. Transcriptomic analyses reveal species-specific light-induced anthocyanin biosynthesis in chrysanthemum. BMC Genomics 16:202

    doi: 10.1186/s12864-015-1428-1

    CrossRef   Google Scholar

    [165]

    Chen S, Li C, Zhu X, Deng Y, Sun W, et al. 2012. The identification of flavonoids and the expression of genes of anthocyanin biosynthesis in the chrysanthemum flowers. Biologia Plantarum 56:458−64

    doi: 10.1007/s10535-012-0069-3

    CrossRef   Google Scholar

    [166]

    Brugliera F, Tao G, Tems U, Kalc G, Mouradova E, et al. 2013. Violet/blue Chrysanthemums-metabolic engineering of the anthocyanin biosynthetic pathway results in novel petal colors. Plant & Cell Physiology 54:1696−710

    doi: 10.1093/pcp/pct110

    CrossRef   Google Scholar

    [167]

    Christaki E, Bonos E, Giannenas I, Florou-Paneri P. 2013. Functional properties of carotenoids originating from algae. Journal of the Science of Food and Agriculture 93:5−11

    doi: 10.1002/jsfa.5902

    CrossRef   Google Scholar

    [168]

    Yamamizo C, Kishimoto S, Ohmiya A. 2010. Carotenoid composition and carotenogenic gene expression during Ipomoea petal development. Journal of Experimental Botany 61:709−19

    doi: 10.1093/jxb/erp335

    CrossRef   Google Scholar

    [169]

    Yamagishi M, Kishimoto S, Nakayama M. 2010. Carotenoid composition and changes in expression of carotenoid biosynthetic genes in tepals of Asiatic hybrid lily. Plant Breeding 129:100−7

    doi: 10.1111/j.1439-0523.2009.01656.x

    CrossRef   Google Scholar

    [170]

    Chiou CY, Pan HA, Chuang YN, Yeh KW. 2010. Differential expression of carotenoid-related genes determines diversified carotenoid coloration in floral tissues of Oncidium cultivars. Planta 232:937−48

    doi: 10.1007/s00425-010-1222-x

    CrossRef   Google Scholar

    [171]

    Sun T, Li L. 2020. Toward the 'golden' era: The status in uncovering the regulatory control of carotenoid accumulation in plants. Plant Science 290:110331

    doi: 10.1016/j.plantsci.2019.110331

    CrossRef   Google Scholar

    [172]

    Misawa N, Yamano S, Linden H, de Felipe MR, Lucas M, et al. 1993. Functional expression of the erwinia-uredovora carotenoid biosynthesis gene crtl in transgenic plants showing an increase of beta-carotene biosynthesis activity and resistance to the bleaching herbicide norflurazon. The Plant Journal 4:833−40

    doi: 10.1046/j.1365-313X.1993.04050833.x

    CrossRef   Google Scholar

    [173]

    Suzuki S, Nishihara M, Nakatsuka T, Misawa N, Ogiwara I, et al. 2007. Flower color alteration in Lotus japonicus by modification of the carotenoid biosynthetic pathway. Plant Cell Reports 26:951−59

    doi: 10.1007/s00299-006-0302-7

    CrossRef   Google Scholar

    [174]

    Li X, Tang D, Du H, Shi Y. 2018. Transcriptome sequencing and biochemical analysis of perianths and coronas reveal flower color formation in Narcissus pseudonarcissus. International Journal of Molecular Sciences 19:4006

    doi: 10.3390/ijms19124006

    CrossRef   Google Scholar

    [175]

    Li L, Ye J, Li H, Shi Q. 2020. Characterization of metabolites and transcripts involved in flower pigmentation in Primula vulgaris. Frontiers in Plant Science 11:572517

    doi: 10.3389/fpls.2020.572517

    CrossRef   Google Scholar

    [176]

    Iijima L, Kishimoto S, Ohmiya A, Yagi M, Okamoto E, et al. 2020. Esterified carotenoids are synthesized in petals of carnation (Dianthus caryophyllus) and accumulate in differentiated chromoplasts. Scientific Reports 10:15256

    doi: 10.1038/s41598-020-72078-4

    CrossRef   Google Scholar

    [177]

    Meng Y, Wang Z, Wang Y, Wang C, Zhu B, et al. 2019. The MYB Activator WHITE PETAL1 Associates with MtTT8 and MtWD40-1 to Regulate Carotenoid-Derived Flower Pigmentation in Medicago truncatula. The Plant Cell 31:2751−67

    doi: 10.1105/tpc.19.00480

    CrossRef   Google Scholar

    [178]

    Li BJ, Zheng BQ, Wang JY, Tsai WC, Lu HC, et al. 2020. New insight into the molecular mechanism of colour differentiation among floral segments in orchids. Communications Biology 3:89

    doi: 10.1038/s42003-020-0821-8

    CrossRef   Google Scholar

    [179]

    Pavokovic D, Krsnik-Rasol M. 2011. Complex Biochemistry and Biotechnological Production of Betalains. Food Technology and Biotechnology 49:145−55

    Google Scholar

    [180]

    Nishihara M, Nakatsuka T. 2010. Genetic engineering of novel flower colors in floricultural plants: recent advances via transgenic approaches. In Protocols for In Vitro Propagation of Ornamentals Plants, Methods in Molecular Biology (Methods and Protocols), ed. Jain SM, Ochatt SJ. vol 589. London: Humana Press, Springer Science+Business Media. pp. 325−47 https://doi.org/10.1007/978-1-60327-114-1_29

    [181]

    Henarejos-Escudero P, Guadarrama-Flores B, García-Carmona F, Gandía-Herrero F. 2018. Digestive glands extraction and precise pigment analysis support the exclusion of the carnivorous plant Dionaea muscipula Ellis from the Caryophyllales order. Plant Science 274:342−48

    doi: 10.1016/j.plantsci.2018.06.013

    CrossRef   Google Scholar

    [182]

    Gandía-Herrero F, García-Carmona F. 2020. The dawn of betalains. New Phytologist 227:664−66

    doi: 10.1111/nph.16295

    CrossRef   Google Scholar

    [183]

    Polturak G, Aharoni A. 2018. "La vie en rose'': biosynthesis, sources, and applications of betalain pigments. Molecular Plant 11:7−22

    doi: 10.1016/j.molp.2017.10.008

    CrossRef   Google Scholar

    [184]

    Hatlestad GJ, Sunnadeniya RM, Akhavan NA, Gonzalez A, Goldman IL, et al. 2012. The beet R locus encodes a new cytochrome P450 required for red betalain production. Nature Genetics 44:816−20

    doi: 10.1038/ng.2297

    CrossRef   Google Scholar

    [185]

    Hatlestad GJ, Akhavan NA, Sunnadeniya RM, Elam L, Cargile S, et al. 2015. The beet Y locus encodes an anthocyanin MYB-like protein that activates the betalain red pigment pathway. Nature Genetics 47:92−96

    doi: 10.1038/ng.3163

    CrossRef   Google Scholar

    [186]

    Wu Y, Xu J, Han X, Qiao G, Yang K, et al. 2020. Comparative transcriptome analysis combining SMRT- and illumina-based RNA-Seq identifies potential candidate genes involved in betalain biosynthesis in Pitaya fruit. International Journal of Molecular Sciences 21:3288

    doi: 10.3390/ijms21093288

    CrossRef   Google Scholar

    [187]

    Zhou Z, Gao H, Ming J, Ding Z, Lin X, et al. 2020. Combined Transcriptome and Metabolome analysis of Pitaya fruit unveiled the mechanisms underlying Peel and pulp color formation. BMC Genomics 21:734

    doi: 10.1186/s12864-020-07133-5

    CrossRef   Google Scholar

    [188]

    Suzuki M, Miyahara T, Tokumoto H, Hakamatsuka T, Goda Y, et al. 2014. Transposon-mediated mutation of CYP76AD3 affects betalain synthesis and produces variegated flowers in four o'clock (Mirabilis jalapa). Journal of Plant Physiology 171:1586−90

    doi: 10.1016/j.jplph.2014.07.010

    CrossRef   Google Scholar

    [189]

    Sasaki N, Abe Y, Goda Y, Adachi T, Kasahara K, et al. 2009. Detection of DOPA 4,5-dioxygenase (DOD) activity using recombinant protein prepared from Escherichia coli cells harboring cDNA encoding DOD from Mirabilis jalapa. Plant and Cell Physiology 50:1012−16

    doi: 10.1093/pcp/pcp053

    CrossRef   Google Scholar

    [190]

    Ramya M, An HR, Baek YS, Reddy KE, Park PH. 2018. Orchid floral volatiles: Biosynthesis genes and transcriptional regulations. Scientia Horticulturae 235:62−69

    doi: 10.1016/j.scienta.2017.12.049

    CrossRef   Google Scholar

    [191]

    Knudsen JT, Eriksson R, Gershenzon J, Ståhl B. 2006. Diversity and distribution of floral scent. The Botanical Review 72:1−120

    doi: 10.1663/0006-8101(2006)72[1:DADOFS]2.0.CO;2

    CrossRef   Google Scholar

    [192]

    Dudareva N, Pichersky E. 2000. Biochemical and molecular genetic aspects of floral scents. Plant Physiology 122:627−33

    doi: 10.1104/pp.122.3.627

    CrossRef   Google Scholar

    [193]

    Hsiao YY, Tsai WC, Kuoh CS, Huang TH, Wang HC, et al. 2006. Comparison of transcripts in Phalaenopsis bellina and Phalaenopsis equestris (Orchidaceae) flowers to deduce monoterpene biosynthesis pathway. BMC Plant Biology 6:14

    doi: 10.1186/1471-2229-6-14

    CrossRef   Google Scholar

    [194]

    Hsiao YY, Jeng MF, Tsai WC, Chuang YC, Li CY, et al. 2008. A novel homodimeric geranyl diphosphate synthase from the orchid Phalaenopsis bellina lacking a DD(X)2-4D motif. The Plant Journal 55:719−33

    doi: 10.1111/j.1365-313X.2008.03547.x

    CrossRef   Google Scholar

    [195]

    Chan W-S, Abdullah JO, Namasivayam P. 2011. Isolation, cloning and characterization of fragrance-related transcripts from Vanda Mimi Palmer. Scientia Horticulturae 127:388−97

    doi: 10.1016/j.scienta.2010.09.024

    CrossRef   Google Scholar

    [196]

    Zhang Y, Li C, Wang S, Yuan M, Li B, et al. 2021. Transcriptome and volatile compounds profiling analyses provide insights into the molecular mechanism underlying the floral fragrance of tree peony. Industrial Crops and Products 162:113286

    doi: 10.1016/j.indcrop.2021.113286

    CrossRef   Google Scholar

    [197]

    Boatright J, Negre F, Chen X, Kish CM, Wood B, et al. 2004. Understanding in vivo benzenoid metabolism in petunia petal tissue. Plant Physiology 135:1993−2011

    doi: 10.1104/pp.104.045468

    CrossRef   Google Scholar

    [198]

    Zhang T, Bao F, Yang Y, Hu L, Ding A, et al. 2020. A comparative analysis of floral scent compounds in intraspecific cultivars of Prunus mume with Different Corolla Colours. Molecules 25:11

    doi: 10.3390/molecules25010145

    CrossRef   Google Scholar

    [199]

    Verdonk JC, Haring MA, van Tunen AJ, Schuurink RC. 2005. ODORANT1 regulates fragrance biosynthesis in petunia flowers. Plant Cell 17:1612−24

    doi: 10.1105/tpc.104.028837

    CrossRef   Google Scholar

    [200]

    Tóth G, Barabás C, Tóth A, Kéry Á, Béni S, et al. 2016. Characterization of antioxidant phenolics in Syringa vulgaris L. flowers and fruits by HPLC-DAD-ESI-MS. Biomedical Chromatography 30:923−32

    doi: 10.1002/bmc.3630

    CrossRef   Google Scholar

    [201]

    Dhandapani S, Jin JJ, Sridhar V, Sarojam R, Chua NH, Jang IC. 2017. Integrated metabolome and transcriptome analysis of Magnolia champaca identifies biosynthetic pathways for floral volatile organic compounds. BMC Genomics 18:463

    doi: 10.1186/s12864-017-3846-8

    CrossRef   Google Scholar

    [202]

    Dexter R, Qualley A, Kish CM, Ma CJ, Koeduka T, et al. 2007. Characterization of a petunia acetyltransferase involved in the biosynthesis of the floral volatile isoeugenol. The Plant Journal 49:265−75

    doi: 10.1111/j.1365-313X.2006.02954.x

    CrossRef   Google Scholar

    [203]

    Underwood BA, Tieman DM, Shibuya K, Dexter RJ, Loucas HM, et al. 2005. Ethylene-regulated floral volatile synthesis in petunia corollas. Plant Physiology 138:255−66

    doi: 10.1104/pp.104.051144

    CrossRef   Google Scholar

    [204]

    Kaminaga Y, Schnepp J, Peel G, Kish CM, Ben-Nissan G, et al. 2006. Plant phenylacetaldehyde synthase is a bifunctional homotetrameric enzyme that catalyzes phenylalanine decarboxylation and oxidation. Journal of Biological Chemistry 281:23357−66

    doi: 10.1074/jbc.M602708200

    CrossRef   Google Scholar

    [205]

    Orlova I, Marshall-Colón A, Schnepp J, Wood B, Varbanova M, et al. 2006. Reduction of benzenoid synthesis in petunia flowers reveals multiple pathways to benzoic acid and enhancement in auxin transport. The Plant Cell 18:3458−75

    doi: 10.1105/tpc.106.046227

    CrossRef   Google Scholar

    [206]

    Negre F, Kish CM, Boatright J, Underwood B, Shibuya K, et al. 2003. Regulation of methylbenzoate emission after pollination in snapdragon and petunia flowers. The Plant Cell 15:2992−3006

    doi: 10.1105/tpc.016766

    CrossRef   Google Scholar

    [207]

    Dexter RJ, Verdonk JC, Underwood BA, Shibuya K, Schmelz EA, et al. 2008. Tissue-specific PhBPBT expression is differentially regulated in response to endogenous ethylene. Journal of Experimental Botany 59:609−18

    doi: 10.1093/jxb/erm337

    CrossRef   Google Scholar

    [208]

    Spitzer-Rimon B, Farhi M, Albo B, Cna'ani A, Ben Zvi MM, et al. 2012. The R2R3-MYB-like regulatory factor EOBI, acting downstream of EOBII, regulates scent production by activating ODO1 and structural scent-related genes in Petunia. Plant Cell 24:5089−105

    doi: 10.1105/tpc.112.105247

    CrossRef   Google Scholar

    [209]

    Bao F, Zhang T, Ding A, Ding A, Yang W, et al. 2020. Metabolic, enzymatic activity, and transcriptomic analysis reveals the mechanism underlying the lack of characteristic floral scent in apricot mei varieties. Frontiers in Plant Science 11:57498

    doi: 10.3389/fpls.2020.574982

    CrossRef   Google Scholar

    [210]

    Schade F, Legge RL, Thompson JE. 2001. Fragrance volatiles of developing and senescing carnation flowers. Phytochemistry 56:703−10

    doi: 10.1016/S0031-9422(00)00483-0

    CrossRef   Google Scholar

    [211]

    Jemia MB, Senatore F, Bruno M, Bancheva S. 2015. Components from the Essential oil of Centaurea aeolica Guss. and C-diluta Aiton from Sicily, Italy. Records of Natural Products 9:580−85

    Google Scholar

    [212]

    Gong W, Xu S, Liu Y, Wang C, Martin K, et al. 2019. Chemical composition of floral scents from three Plumeria rubra L. (Apocynaceae) forms linked to petal color proprieties. Biochemical Systematics and Ecology 85:54−59

    doi: 10.1016/j.bse.2019.05.005

    CrossRef   Google Scholar

    [213]

    Dudareva N, Murfitt LM, Mann CJ, Gorenstein N, Kolosova N, et al. 2000. Developmental regulation of methyl benzoate biosynthesis and emission in snapdragon flowers. The Plant Cell 12:949−61

    doi: 10.1105/tpc.12.6.949

    CrossRef   Google Scholar

    [214]

    Ramya M, Kwon OK, An HR, Park PM, Baek YS, et al. 2017. Floral scent: Regulation and role of MYB transcription factors. Phytochemistry Letters 19:114−20

    doi: 10.1016/j.phytol.2016.12.015

    CrossRef   Google Scholar

    [215]

    Schwab W, Davidovich-Rikanati R, Lewinsohn E. 2008. Biosynthesis of plant-derived flavor compounds. The Plant Journal 54:712−32

    doi: 10.1111/j.1365-313X.2008.03446.x

    CrossRef   Google Scholar

    [216]

    Shinoyama H, Mochizuki A, Nomura Y, Kamada H. 2008. Environmental risk assessment of genetically modified chrysanthemums containing a modified cry1Ab gene from Bacillus thuringiensis. Plant Biotechnology 25:17−29

    doi: 10.5511/plantbiotechnology.25.17

    CrossRef   Google Scholar

    [217]

    Takatsu Y, Nishizawa Y, Hibi T, Akutsu K. 1999. Transgenic chrysanthemum (Dendranthema grandiflorum (Ramat.) Kitamura) expressing a rice chitinase gene shows enhanced resistance to gray mold (Botrytis cinerea). Scientia Horticulturae 82:113−23

    doi: 10.1016/S0304-4238(99)00034-5

    CrossRef   Google Scholar

    [218]

    Kumar S, Raj SK, Sharma AK, Varma HN. 2012. Genetic transformation and development of Cucumber mosaic virus resistant transgenic plants of Chrysanthemum morifolium cv. Kundan. Scientia Horticulturae 134:40−45

    doi: 10.1016/j.scienta.2011.10.019

    CrossRef   Google Scholar

    [219]

    Núñez de Cáceres González FF, Davey MR, Cancho Sanchez E, Wilson ZA. 2015. Conferred resistance to Botrytis cinerea in Lilium by overexpression of the RCH10 chitinase gene. Plant Cell Reports 34:1201−9

    doi: 10.1007/s00299-015-1778-9

    CrossRef   Google Scholar

    [220]

    Azadi P, Otang NV, Supaporn H, Khan RS, Chin DP, et al. 2011. Increased resistance to cucumber mosaic virus (CMV) in Lilium transformed with a defective CMV replicase gene. Biotechnology Letters 33:1249−55

    doi: 10.1007/s10529-011-0550-7

    CrossRef   Google Scholar

    [221]

    Kamo K, Thilmony R, Bauchan G. 2019. Transgenic Lilium longiflorum plants containing the bar-uidA gene controlled by the rice RPC1, Agrobacterium rolD, mas2, and CaMV 35S promoters. Plant Cell Tissue and Organ Culture 136:303−12

    doi: 10.1007/s11240-018-1515-5

    CrossRef   Google Scholar

    [222]

    Vieira P, Wantoch S, Lilley CJ, Chitwood DJ, Atkinson HJ, et al. 2015. Expression of a cystatin transgene can confer resistance to root lesion nematodes in Lilium longiflorum cv. 'Nellie White'. Transgenic Research 24:421−32

    doi: 10.1007/s11248-014-9848-2

    CrossRef   Google Scholar

    [223]

    Kamo K, Gera A, Cohen J, Hammond J, Blowers A, et al. 2005. Transgenic Gladiolus plants transformed with the bean yellow mosaic virus coat-protein gene in either sense or antisense orientation. Plant Cell Reports 23:54−63

    doi: 10.1007/s00299-004-0888-6

    CrossRef   Google Scholar

    [224]

    Kamo K, Jordan R, Guaragna MA, Hsu HT, Ueng P. 2010. Resistance to Cucumber mosaic virus in Gladiolus plants transformed with either a defective replicase or coat protein subgroup II gene from Cucumber mosaic virus. Plant Cell Reports 29:695−704

    doi: 10.1007/s00299-010-0855-3

    CrossRef   Google Scholar

    [225]

    Kamo K, Aebig J, Guaragna MA, James C, Hsu HT, et al. 2012. Gladiolus plants transformed with single-chain variable fragment antibodies to Cucumber mosaic virus. Plant Cell, Tissue and Organ Culture 110:13−21

    doi: 10.1007/s11240-012-0124-y

    CrossRef   Google Scholar

    [226]

    Chang C, Chen YC, Hsu YH, Wu JT, Hu CC, et al. 2005. Transgenic resistance to Cymbidium mosaic virus in Dendrobium expressing the viral capsid protein gene. Transgenic Research 14:41−6

    doi: 10.1007/s11248-004-2373-y

    CrossRef   Google Scholar

    [227]

    Kuehnle AR, Sugii N. 1992. Transformation of Dendrobium orchid using particle bombardment of protocorms. Plant Cell Reports 11:484−88

    doi: 10.1007/BF00232696

    CrossRef   Google Scholar

    [228]

    Liao LJ, Pan IC, Chan YL, Hsu YH, Chen WH, et al. 2004. Transgene silencing in Phalaenopsis expressing the coat protein of Cymbidium Mosaic Virus is a manifestation of RNA-mediated resistance. Molecular Breeding 13:229−42

    doi: 10.1023/B:MOLB.0000022527.68551.30

    CrossRef   Google Scholar

    [229]

    Gargul JM, Mibus H, Serek M. 2015. Manipulation of MKS1 gene expression affects Kalanchoë blossfeldiana and Petunia hybrida phenotypes. Plant Biotechnology Journal 13:51−61

    doi: 10.1111/pbi.12234

    CrossRef   Google Scholar

    [230]

    Li X, Gasic K, Cammue B, Broekaert W, Korban SS. 2003. Transgenic rose lines harboring an antimicrobial protein gene, Ace-AMP1, demonstrate enhanced resistance to powdery mildew (Sphaerotheca pannosa). Planta 218:226−32

    doi: 10.1007/s00425-003-1093-5

    CrossRef   Google Scholar

    [231]

    Ramakrishna A, Gill SS. 2018. Metabolic adaptations in plants during abiotic stress. Boca Raton: CRC Press. 442 pp. https://doi.org/10.1201/b22206

    [232]

    He X, Wang C, Wang H, Li L, Wang C. 2020. The function of MAPK cascades in response to various stresses in horticultural plants. Frontiers in Plant Science 11:12

    doi: 10.3389/fpls.2020.00952

    CrossRef   Google Scholar

    [233]

    Song AP, Hu YH, Ding L, Zhang X, Li PL, et al. 2018. Comprehensive analysis of mitogen-activated protein kinase cascades in chrysanthemum. PeerJ 6:e5037

    doi: 10.7717/peerj.5037

    CrossRef   Google Scholar

    [234]

    Hollender CA, Pascal T, Tabb A, Hadiarto T, Srinivasan C, et al. 2018. Loss of a highly conserved sterile alpha motif domain gene (WEEP) results in pendulous branch growth in peach trees. PNAS 115:E4690−E4699

    doi: 10.1073/pnas.1800138115

    CrossRef   Google Scholar

    [235]

    Zhuo X, Zheng T, Zhang Z, Li S, Zhang Y, et al. 2021. Bulked segregant RNA sequencing (BSR-seq) identifies a novel allele associated with weeping traits in prunus mume. Frontiers of Agricultural Science and Engineering 8:196−214

    doi: 10.15302/j-fase-2020379

    CrossRef   Google Scholar

    [236]

    Hollender CA, Hadiarto T, Srinivasan C, Scorza R, Dardick C. 2016. A brachytic dwarfism trait (dw) in peach trees is caused by a nonsense mutation within the gibberellic acid receptor PpeGID1c. New Phytologist 210:227−39

    doi: 10.1111/nph.13772

    CrossRef   Google Scholar

    [237]

    Lu Z, Niu L, Chagné D, Cui G, Pan L, et al. 2016. Fine mapping of the temperature-sensitive semi-dwarf (Tssd) locus regulating the internode length in peach (Prunus persica). Molecular Breeding 36:20

    doi: 10.1007/s11032-016-0442-6

    CrossRef   Google Scholar

    [238]

    Islam MA, Lütken H, Haugslien S, Blystad DR, Torre S, et al. 2013. Overexpression of the AtSHI gene in poinsettia, Euphorbia pulcherrima, results in compact plants. Plos One 8:e53377

    doi: 10.1371/journal.pone.0053377

    CrossRef   Google Scholar

    [239]

    Koike Y, Hoshino Y, Mii M, Nakano M. 2003. Horticultural characterization of Angelonia salicariifolia plants transformed with wild-type strains of Agrobacterium rhizogenes. Plant Cell Reports 21:981−87

    doi: 10.1007/s00299-003-0608-7

    CrossRef   Google Scholar

    [240]

    Pérez de la Torre MC, Fernández P, Greppi JA, Coviella MA, Fernández MN, et al. 2018. Transformation of Mecardonia (Plantaginaceae) with wild-type Agrobacterium rhizogenes efficiently improves compact growth, branching and flower related ornamental traits. Scientia Horticulturae 234:300−11

    doi: 10.1016/j.scienta.2018.02.047

    CrossRef   Google Scholar

    [241]

    Bonnafous F, Fievet G, Blanchet N, Boniface MC, Carrère S, et al. 2018. Comparison of GWAS models to identify non-additive genetic control of flowering time in sunflower hybrids. Theoretical and Applied Genetics 131:319−32

    doi: 10.1007/s00122-017-3003-4

    CrossRef   Google Scholar

    [242]

    Kao CH. 2000. On the differences between maximum likelihood and regression interval mapping in the analysis of quantitative trait loci. Genetics 156:855−65

    doi: 10.1093/genetics/156.2.855

    CrossRef   Google Scholar

    [243]

    Paterson AH, Lander ES, Hewitt JD, Peterson S, Lincoln SE, et al. 1988. Resolution of quantitative traits into Mendelian factors by using a complete linkage map of restriction fragment length polymorphisms. Nature 335:721−26

    doi: 10.1038/335721a0

    CrossRef   Google Scholar

    [244]

    Zhang J, Zhang Q, Cheng T, Yang W, Pan H, et al. 2015. High-density genetic map construction and identification of a locus controlling weeping trait in an ornamental woody plant (Prunus mume Sieb. et Zucc). DNA Research 22:183−91

    doi: 10.1093/dnares/dsv003

    CrossRef   Google Scholar

    [245]

    Li J, Xu Y, Wang Z. 2019. Construction of a high-density genetic map by RNA sequencing and eQTL analysis for stem length and diameter in Dendrobium (Dendrobium nobile × Dendrobium wardianum). Industrial Crops and Products 128:48−54

    doi: 10.1016/j.indcrop.2018.10.073

    CrossRef   Google Scholar

    [246]

    Li S, Lv S, Yu K, Wang Z, Li Y, et al. 2019. Construction of a high-density genetic map of tree peony (Paeonia suffruticosa Andr. Moutan) using restriction site associated DNA sequencing (RADseq) approach. Tree Genetics & Genomes 15:63

    doi: 10.1007/s11295-019-1367-0

    CrossRef   Google Scholar

    [247]

    Zhang Q, Li L, VanBuren R, Liu Y, Yang M, et al. 2014. Optimization of linkage mapping strategy and construction of a high-density American lotus linkage map. BMC Genomics 15:372

    doi: 10.1186/1471-2164-15-372

    CrossRef   Google Scholar

    [248]

    Otagaki S, Ogawa Y, Hibrand-Saint Oyant L, Foucher F, Kawamura K, et al. 2015. Genotype of FLOWERING LOCUS T homologue contributes to flowering time differences in wild and cultivated roses. Plant Biology 17:808−15

    doi: 10.1111/plb.12299

    CrossRef   Google Scholar

    [249]

    Alqudah AM, Sallam A, Baenziger PS, Börner A. 2020. GWAS: Fast-forwarding gene identification and characterization in temperate Cereals: lessons from Barley - A review. Journal of Advanced Research 22:119−35

    doi: 10.1016/j.jare.2019.10.013

    CrossRef   Google Scholar

    [250]

    Michelmore RW, Paran I, Kesseli RV. 1991. Identification of markers linked to disease-resistance genes by bulked segregant analysis: a rapid method to detect markers in specific genomic regions by using segregating populations. PNAS 88:9828−32

    doi: 10.1073/pnas.88.21.9828

    CrossRef   Google Scholar

    [251]

    Zou C, Wang P, Xu Y. 2016. Bulked sample analysis in genetics, genomics and crop improvement. Plant Biotechnology Journal 14:1941−55

    doi: 10.1111/pbi.12559

    CrossRef   Google Scholar

    [252]

    Dougherty L, Singh R, Brown S, Dardick C, Xu K. 2018. Exploring DNA variant segregation types in pooled genome sequencing enables effective mapping of weeping trait in Malus. Journal of Experimental Botany 69:1499−516

    doi: 10.1093/jxb/erx490

    CrossRef   Google Scholar

    [253]

    Dommes AB, Gross T, Herbert DB, Kivivirta KI, Becker A. 2019. Virus-induced gene silencing: empowering genetics in non-model organisms. Journal of Experimental Botany 70:757−70

    doi: 10.1093/jxb/ery411

    CrossRef   Google Scholar

    [254]

    Becker A, Lange M. 2010. VIGS - genomics goes functional. Trends in Plant Science 15:1−4

    doi: 10.1016/j.tplants.2009.09.002

    CrossRef   Google Scholar

    [255]

    Lu J, Sun J, Jiang A, Bai M, Fan C, et al. 2020. Alternate expression of CONSTANS-LIKE 4 in short days and CONSTANS in long days facilitates day-neutral response in Rosa chinensis. Journal of Experimental Botany 71:4057−68

    doi: 10.1093/jxb/eraa161

    CrossRef   Google Scholar

    [256]

    Lai PH, Huang LM, Pan ZJ, Jane WN, Chung MC, et al. 2020. PeERF1, a SHINE-Like transcription factor, is involved in nanoridge development on lip epidermis of Phalaenopsis flowers. Frontiers in Plant Science 10:1709

    doi: 10.3389/fpls.2019.01709

    CrossRef   Google Scholar

    [257]

    Gaj T, Gersbach CA, Barbas CF. 2013. ZFN, TALEN, and CRISPR/Cas-based methods for genome engineering. Trends in Biotechnology 31:397−405

    doi: 10.1016/j.tibtech.2013.04.004

    CrossRef   Google Scholar

    [258]

    Bibikova M, Golic M, Golic KG, Carroll D. 2002. Targeted chromosomal cleavage and mutagenesis in Drosophila using zinc-finger nucleases. Genetics 161:1169−75

    doi: 10.1093/genetics/161.3.1169

    CrossRef   Google Scholar

    [259]

    Christian M, Cermak T, Doyle EL, Schmidt C, Zhang F, et al. 2010. Targeting DNA double-strand breaks with TAL effector nucleases. Genetics 186:757−61

    doi: 10.1534/genetics.110.120717

    CrossRef   Google Scholar

    [260]

    Cong L, Ran FA, Cox D, Lin S, Barretto R, et al. 2013. Multiplex Genome Engineering Using CRISPR/Cas Systems. Science 339:819−23

    doi: 10.1126/science.1231143

    CrossRef   Google Scholar

    [261]

    Yin K, Gao C, Qiu J. 2017. Progress and prospects in plant genome editing. Nature Plants 3:17107

    doi: 10.1038/nplants.2017.107

    CrossRef   Google Scholar

    [262]

    Fichtner F, Urrea Castellanos R, Ülker B. 2014. Precision genetic modifications: a new era in molecular biology and crop improvement. Planta 239:921−39

    doi: 10.1007/s00425-014-2029-y

    CrossRef   Google Scholar

    [263]

    Chen K, Gao C. 2014. Targeted genome modification technologies and their applications in crop improvements. Plant Cell Reports 33:575−83

    doi: 10.1007/s00299-013-1539-6

    CrossRef   Google Scholar

    [264]

    Watanabe K, Kobayashi A, Endo M, Sage-Ono K, Toki S, et al. 2017. CRISPR/Cas9-mediated mutagenesis of the dihydroflavonol-4-reductase-B (DFR-B) locus in the Japanese morning glory Ipomoea (Pharbitis) nil. Scientific Reports 7:10028

    doi: 10.1038/s41598-017-10715-1

    CrossRef   Google Scholar

    [265]

    Yan R, Wang Z, Ren Y, Li H, Liu N, et al. 2019. Establishment of Efficient Genetic Transformation Systems and Application of CRISPR/Cas9 Genome Editing Technology in Lilium pumilum DC. Fisch. and Lilium longiflorum White Heaven. International Journal of Molecular Sciences 20:2920

    doi: 10.3390/ijms20122920

    CrossRef   Google Scholar

    [266]

    Cheng S, Melkonian M, Smith SA, Brockington S, Archibald JM, et al. 2018. 10KP: A phylodiverse genome sequencing plan. Gigascience 7:giy013

    doi: 10.1093/gigascience/giy013

    CrossRef   Google Scholar

    [267]

    Chen F, Song Y, Li X, Chen J, Mo L, et al. 2019. Genome sequences of horticultural plants: past, present, and future. Horticulture Research 6:112

    doi: 10.1038/s41438-019-0195-6

    CrossRef   Google Scholar

    [268]

    Li Z, Barker MS. 2020. Inferring putative ancient whole-genome duplications in the 1000 Plants (1KP) initiative: access to gene family phylogenies and age distributions. Gigascience 9:giaa004

    doi: 10.1093/gigascience/giaa004

    CrossRef   Google Scholar

    [269]

    Zheng T, Li P, Li L, Zhang Q. 2021. Research advances in and prospects of ornamental plant genomics. Horticulture Research 8:65

    doi: 10.1038/s41438-021-00499-x

    CrossRef   Google Scholar

    [270]

    Jung H, Winefield C, Bombarely A, Prentis P, Waterhouse P. 2019. Tools and Strategies for Long-Read Sequencing and De Novo Assembly of Plant Genomes. Trends in Plant Science 24:700−24

    doi: 10.1016/j.tplants.2019.05.003

    CrossRef   Google Scholar

    [271]

    Boutigny AL, Dohin N, Pornin D, Rolland M. 2020. Overview and detectability of the genetic modifications in ornamental plants. Horticulture Research 7:11

    doi: 10.1038/s41438-019-0232-5

    CrossRef   Google Scholar

    [272]

    Kumar K, Gambhir G, Dass A, Tripathi AK, Singh A, et al. 2020. Genetically modified crops: current status and future prospects. Planta 251:91

    doi: 10.1007/s00425-020-03372-8

    CrossRef   Google Scholar

    [273]

    Wenger AM, Peluso P, Rowell WJ, Chang PC, Hall RJ, et al. 2019. Accurate circular consensus long-read sequencing improves variant detection and assembly of a human genome. Nature Biotechnology 37:1155

    doi: 10.1038/s41587-019-0217-9

    CrossRef   Google Scholar

    [274]

    Cheng HY, Concepcion GT, Feng X, Zhang H, Li H. 2021. Haplotype-resolved de novo assembly using phased assembly graphs with hifiasm. Nature Methods 18:170

    doi: 10.1038/s41592-020-01056-5

    CrossRef   Google Scholar

    [275]

    Nurk S, Walenz BP, Rhie A, Vollger MR, Logsdon GA, et al. 2020. HiCanu: accurate assembly of segmental duplications, satellites, and allelic variants from high-fidelity long reads. Genome Research 30:1291−305

    doi: 10.1101/gr.263566.120

    CrossRef   Google Scholar

    [276]

    Chen H, Zeng Y, Yang Y, Huang L, Tang B, et al. 2020. Allele-aware chromosome-level genome assembly and efficient transgene-free genome editing for the autotetraploid cultivated alfalfa. Nature Communications 11:2494

    doi: 10.1038/s41467-020-16338-x

    CrossRef   Google Scholar

    [277]

    Shimomura K, Fukino N, Sugiyama M, Kawazu Y, Sakata Y, et al. 2017. Quantitative trait locus analysis of cucumber fruit morphological traits based on image analysis. Euphytica 213:138

    doi: 10.1007/s10681-017-1926-0

    CrossRef   Google Scholar

    [278]

    Ye M, Zhu X, Gao P, Jiang L, Wu R. 2020. Identification of quantitative trait loci for altitude adaptation of tree leaf shape with Populus szechuanica in the Qinghai-Tibetan Plateau. Frontiers in Plant Science 11:632

    doi: 10.3389/fpls.2020.00632

    CrossRef   Google Scholar

    [279]

    Cao Y, Zhu X, Wu R, Sun L. 2019. A Geometric Morphometrics-Based Mapping Model of Leaf Shape Evolution, In: Evolution, Origin of Life, Concepts and Methods, eds. Pontarotti P. Switzerland: Springer, Cham. pp. 161−77 https://doi.org/10.1007/978-3-030-30363-1_8

    [280]

    Sun L, Wang J, Zhu X, Jiang L, Gosik K, et al. 2018. HpQTL: a geometric morphometric platform to compute the genetic architecture of heterophylly. Briefings in Bioinformatics 19:603−12

    doi: 10.1093/bib/bbx011

    CrossRef   Google Scholar

    [281]

    Cameron DE, Bashor CJ, Collins JJ. 2014. A brief history of synthetic biology. Nature Reviews Microbiology 12:381−90

    doi: 10.1038/nrmicro3239

    CrossRef   Google Scholar

    [282]

    Singh A, Walker KT, Ledesma-Amaro R, Ellis T. 2020. Engineering bacterial cellulose by synthetic biology. International Journal of Molecular Sciences 21:9185

    doi: 10.3390/ijms21239185

    CrossRef   Google Scholar

    [283]

    Gibson DG, Glass JI, Lartigue C, Noskov VN, Chuang RY, et al. 2010. Creation of a bacterial cell controlled by a chemically synthesized genome. Science 329:52−56

    doi: 10.1126/science.1190719

    CrossRef   Google Scholar

    [284]

    Long C, Ni Y, Zhang X, Xin T, Long B. 2015. Biodiversity of Chinese Ornamentals. Acta Horticulturae 1087:209−20

    doi: 10.17660/actahortic.2015.1087.25

    CrossRef   Google Scholar

    [285]

    Zhang L. 2019. Advance of horticultural plant genomes. Horticultural Plant Journal 5:229−30

    doi: 10.1016/j.hpj.2019.12.002

    CrossRef   Google Scholar

  • Cite this article

    Li M, Wen Z, Meng J, Cheng T, Zhang Q, et al. 2022. The genomics of ornamental plants: current status and opportunities. Ornamental Plant Research 2:6 doi: 10.48130/OPR-2022-0006
    Li M, Wen Z, Meng J, Cheng T, Zhang Q, et al. 2022. The genomics of ornamental plants: current status and opportunities. Ornamental Plant Research 2:6 doi: 10.48130/OPR-2022-0006

Figures(2)  /  Tables(2)

Article Metrics

Article views(6008) PDF downloads(1330)

REVIEW   Open Access    

The genomics of ornamental plants: current status and opportunities

Ornamental Plant Research  2 Article number: 6  (2022)  |  Cite this article

Abstract: With the rapid development of sequencing technologies, followed by the reduction of sequencing cost, numerous ornamental plants have been sequenced, resulting in their genomic studies shifting from gene cloning and marker development to whole genome profiling. A profound understanding of genome structure and function at the whole genome level can not only help to modify ornamental traits, such as fragrance, color and flower shape, through genetic engineering, but also infer the genetic relationship and evolutionary history of ornamental plants via comparative genomics analysis. In this paper, we review the current situation of sequencing strategies and the application of genomics to study the origin and evolution of ornamental plants. We highlight challenges of ornamental plant genomic research. The use of cutting-edge technologies, such as genomics, gene editing and molecular design polymerization breeding, can facilitate our understanding of genetic regulation mechanisms and the germplasm innovation of important traits in ornamental plants. The results can be expected to significantly increase the breeding efficiency of ornamental plants.

    • Ornamental plants, including annual herbs, perennial herbs, ornamental shrubs, and trees[1], play an important role in the construction of human settlements by promoting urban and rural greening, regulating air humidity, absorbing harmful gases and dust, and preventing soil erosion[2,3]. The extracts of ornamental plants can also be used for cosmetics, foods, fine chemicals and other industries[46]. The breeding objectives of ornamental plants are focused on new shapes, new colors and new fragrances. Conventional breeding techniques, such as cross and mutation breeding, have produced a large number of new cultivars of ornamental plants. However, these traditional methods are time-consuming and laborious, the compatibility of distant hybridization is poor, and the results are unpredictable. Genomics is a new branch of biology, which takes DNA sequencing as its core technology, combined with genetics, molecular biology, and bioinformatics to accurately interpret genome information. The rapid development of genomics has pushed the breeding work to new heights. Through the precise location of functional genes and genetic transformation, we can potentially accelerate the breeding process and obtain new cultivars with favorable features.

      The completion of the whole genome sequence of Prunus mume[7] fuels the development of ornamental plant genomics. Currently, more than 72 species of ornamental plants have been sequenced, including Rosa chinensis[8,9], Chrysanthemum nankingense[10], Phalaenopsis aphrodite[11], Paeonia suffruticosa[12] and other important ornamental plants, thanks to the rapid development of sequencing technology and the reduction in cost. Through the mapping and cloning of functional genes, and the study of gene expression and regulatory networks at the genome-wide level, researchers can systematically analyze the genetic mechanisms underlying important traits of ornamental plants. By analyzing genomic variation in plant populations, we can infer the population structure, population history, and environmental adaptability of ornamental plants.

      In this paper, we review the recent progress in genome sequencing of important ornamental plants, including the sharing of genomic data and its application in ornamental character mapping and variety classification, and discuss the application prospects of new technologies. We pinpoint several challenges faced by ornamental plant breeding. We limit our review to the genomics of floral traits.

    • To date, at least 72 species of ornamental plants have been sequenced (Supplemental Table S1, Table S2), among which perennial herbs, shrubs, trees, and annual herbs account for 36%, 24%, 24%, and 17% (Table 1, Fig. 1), respectively. Sequenced species include 10 from Rosaceae, five from Orchidaceae, five from Asteraceae and six from Fabaceae (Table 2, Supplemental Table S3, Fig 2). Illumina sequencing (second-generation) has been the main sequencing technique of ornamental plants over the past decade. Yet, since 2015, third-generation sequencing technology has been applied, which markably reduces sequencing cost and improves sequencing quality (Supplemental Fig S1).

      Table 1.  Representative species of different tpye of ornamental plants.

      Representative species
      Perennial herbaceous flowers Catharanthus roseus, Chrysanthemum nankingense, Nelumbo nucifera
      Annual herbaceous flowers Helianthus annuus, Ipomoea nil
      Flowering trees Osmanthus fragrans, Prunus mume, Prunus yedoensis
      Flowering shrubs Lavandula angustifolia, Paeonia suffruticosa, Rosa chinensis

      Figure 1. 

      The reported 72 ornamental plants species fall into 4 categories.

      Table 2.  Representative species of different family of ornamental plants.

      Representative species
      Rosaceae Prunus mume, Prunus yedoensis, Rosa chinensis
      Fabaceae Ammopiptanthus nanus, Mimosa pudica
      Asteraceae Chrysanthemum lavandulifolium, Chrysanthemum nankingense, Helianthus annuus
      Orchidaceae Dendrobium catenatum, Phalaenopsis aphrodite

      Figure 2. 

      The genome-sequenced ornamental species belong to 33 families.

    • In 1977, Maxam & Gilbert[13] reported a sequencing method by chemical degradation. In the same year, Sanger et al.[14] developed a method of DNA sequencing based on the selective incorporation of chain-terminating dideoxynucleotides by DNA polymerase during in vitro DNA replication. The application of these technologies implies the birth of a new generation of sequencing. The earliest plants sequenced, such as Arabidopsis thaliana[15], Oryza sativa[16], Zea mays[17], Prunus persica[18] adopt the method of first-generation sequencing. However, high cost, time-consuming and low-throughput limit its application, so second-generation sequencing with high throughput and low cost arises at a significant moment[19]. So far, second-generation sequencing represented by Illumina is still a popular method of sequencing. Among the sequenced ornamental plants, 36 species have adopted the second-generation sequencing strategy (Supplemental Table S1, Table S2). However, their limited ability to resolve large complex genomes derived from complex whole-genome duplications leads to many problems. First of all, short read-lengths make it hard to distinguish a large number of repeated sequences, it often leads to incomplete assembly, which brings great difficulties to the subsequent gene annotation and other bioinformatic analysis[20]. Secondly, second-generation sequencing methods depend on PCR amplification during library preparation on the flow cells[19], PCR is a bias-prone process that leads to GC bais. Under the requirement of high throughput, low cost and long read lengths, third-generation sequencing technology is gradually developing (Supplemental Table S4).

      HeliScope is the first commercial single molecule sequencing platform, as the method was relatively slow, expensive, and produced short reads, it did not prove viable[19]. Single-Molecule Real-Time Sequencing (SMRT) and Oxford Nanopore Technologies (ONT) are two commercially successful third-generation sequencing techniques, both of which can produce genome assemblies of high quality without PCR amplification. Single-Molecule Real-Time Sequencing (SMRT) was developed by Pacific Biosciences, and takes advantage of the natural process of DNA replication, using zero-mode waveguide (ZMW) technology and gamma-labeled phosphonucleotides for direct observation of DNA synthesis on single DNA molecules in real time[21]. Single-Molecule Real-Time Sequencing offers longer read lengths than the second-generation sequencing technologies, making it well-suited for genome, transcriptome, and epigenetics research[21]. In 2014, Oxford Nanopore Technologies (ONT) released the MinION as a Nanopore sequencing device, it is much smaller than the current NGS platforms. MinION directly observe DNA or RNA bases by means of pores that are embedded in a membrane separating two compartments[22]. Although the MinION is easy to carry and can produce ultra-long reading length[23], because of the lower read accuracy and high cost, the application of Oxford nano-hole sequencing in plants is still in its infancy[19,24]. Although third-generation sequencing has the advantage of longer read length, its error rate is significantly higher than that of second-generation sequencing, which can be combined to make up for their respective shortcomings, which is especially beneficial to assemble the genome from scratch. This method is called hybrid sequencing[25]. Hybrid sequencing is widely used in plant sequencing. Among the sequenced ornamental plants, four species use Illumina+Nanopore strategy for genome sequencing, 31 species use Illumina+PacBio strategy and one species adopt BGISEQ-500+PacBio strategy (Supplemental Table S1, Table S2).

      Considering species with complexity and high repeat sequence, it is very difficult to assemble a complete genome only by sequence reading[20], many researchers use Hi-C data to assist genome assembly. High-throughput chromosome conformation capture (Hi-C) is a technology that can study the three-dimensional architecture of whole genomes. By coupling proximity-based ligation with large parallel sequencing, Hi-C allows unbiased identification of chromatin interactions across an entire genome[26]. The many advantages of Hi-C technology have attracted the wide attention of researchers, the assembled genomes of Rhododendron williamsianum[27], Rhododendron simsii[28], Rosa chinensis[9], Osmanthus fragrans[29], Chimonanthus praecox[30], Chimonanthus salicifolius[31] are assisted by Hi-C technology. At present, two sequenced white lupine genomes provide us with a good example. The genomes using Hi-C data not only have higher assembly integrity, but also have better chromosome collinearity[32,33].

    • The Arabidopsis Information Resource (TAIR, http://www.arabidopsis.org) was launched in 2001, and is the first angiosperm genome database[34]. With the development of the ornamental plant genome sequencing project, various ornamental plant genome databases have been established. The ornamental plant genome database can be further classified as four different types: family database[18,35]; single species database[18,33,35]; clade-oriented database[36,37] and comprehensive database (Supplemental Table S1, Table S2).

      The single species database of individual species not only provides the updated genome sequence, but also provides a series of bioinformatics tools for online analysis. For example, the Nelumbo Genome Database (NGD, http://nelumbo.biocloud.net)[38] contains the most updated lotus genome assembly and information on both gene expression in different tissues and coexpression networks. In addition, genetic variants and phenotypes of 88 key lotus cultivars were integrated. GBrowse provided by NGD can easily view genes, DNA sequences, amino acids, SNPs and Indels. Moreover, the website also provides tools such as blast, blat, and Primer for sequence analysis and search. The database is helpful to the study of population genetics and molecular breeding of lotus, as well as a comprehensive understanding of the genetic variation characteristics of the Nelumbo nucifera genome. The history of cultivation and domestication of lotus can be traced back to the data in order to cultivate new varieties with high yield and high quality and to improve the level and speed of breeding techniques. Unfortunately, among the sequenced ornamental plants, only a few plants such as Helianthus annuus[38], Ipomoea nil[39] and Dianthus caryophyllus[40] have an established database, and the database of many species is far less perfect than that of the Nelumbo genome database.

      Compared with the single species database, the database integrated according to families/clade-oriented pays more attention to the integration and unification of different plant data. The Genome Database for Rosaceae (GDR, https://www.rosaceae.org)[41] first established in 2003, after decades of development, has significantly expanded data and functions centering on Rosaceae plant interaction. GDR contains detailed data on published trait loci in Rosaceae, based on these trait terms, the GDR team developed the Rosaceae Trait Ontology that is tightly linked to the Plant Trait Ontology (TO). Except for this, QTL in GDR also includes aliases, curator-assigned QTL labels, published symbols, trait names, taxa, trait descriptions, screening methods, map positions, associated markers, statistical values, datasets, contact information and references. At the same time, the website supports viewing and statistics of these data according to genus units. The Breeding Information Management System (BIMS) is a Tripal module developed by the GDR team, it provides breeding teams with tools to store, manage, archive and analyze their private or public breeding data, and other breeders can also use BIMS to query and download public breeding data. At present, GDR includes genome assembly and annotation data of five ornamental plants, including Cerasus × yedoensis, Malus baccata, Rosa multiflora, Rosa chinensis, Prunus persica, as well as 24 crops, including apricot, apple and strawberry.

      A family database or clade-oriented database has the disadvantage of incomplete data. At present, GDR lacks the genome data of Prunus mume, Dryas drummondii and Prunus yedoensis. OrchidBase lacks the genome data of Dendrobium officinale and Phalaenopsis Aphrodite. The pomegranate genome lacks two cultivars, 'Tunisia' and 'Dabenzi', in the Hardwood Genomics Project. Many comprehensive plant databases have been established, and the more representative ones include Phytozome[42], PGDD[43], PLAZA[44], and Ensemblplants[45], containing most of the published plant genome information.

      The genome data of ornamental plants are growing rapidly in quantity and complexity, and its main function has evolved from data storage to online analysis[46]. The excellent genome database will provide great convenience for the study of structural genomics, comparative genomics and functional genomics of ornamental plants.

    • In recent years, a large number of plant genome sequencing and resequencing projects have promoted the development of comparative genomics. By comparing the genome sequences of different related species, we can identify the contraction and expansion of gene families[47], estimate the divergence time of species[48], identify gene fusion and gene cluster[10,49] and Analysis system evolution[8]. The study of comparative genomics of ornamental plants will help to speed up the basic research of their biology and improve the efficiency of the development and utilization of wild related species resources.

    • A variety of evidence shows that gene replication can promote the rapid recombination of plant genome, lose a large number of genes and increase structural variation, which is extremely important for plant evolution[50]. The study of gene replication in ornamental plants has important guiding significance for their genetics and breeding.

      Gene duplication includes tandem, tetraploid, segmental and transpositional[51]. Tandem repeats are highly prevalent at centromeres[52], to create a cluster of homologous genes with similar sequence and function, which are arranged in series on the chromosome in a head-to-tail manner[53]. There is evidence that tandem duplication leads to widespread TS and CYP gene clusters in the chrysanthemum genome, which greatly promotes the diversity of terpenoids in chrysanthemum[10]. Segmental duplication is the most important structural variation on chromosomes[54]. Repetitive region caused by segmental duplication is usually the repetition of all genes in a large region, rather than the duplication of single or several genes. For individual gene families, fragment replication also plays a major role in species-specific expansion[5557]. DNA binding with one figure (Dof) proteins play important roles in various biological processes and defense regulatory networks in plants. Dof genes of many plants came primarily from segmental and duplication events[55,56]. A total of 24 full-length RchDof genes were identified in Rosa chinensis, seven pairs of RchDof duplicated genes (12 RchDof genes) were generated by segmental duplication[57]. Transposable elements (TEs) is an important part of the plant genome. They can replicate independently and insert replicative fragments into many different loci on other DNA sequences in the same cell[58]. The long terminal repeat (LTR) retrotransposon in Phalaenopsis equestris were identified and six high-copy retrotransposons were found[59]. Among them, Gypsy1, Gypsy2, Gypsy3, and Orchid-rt1 have amplified within Phalaenopsis orchids concomitant with the expanded genome sizes, and Orchid-rt1 and Gypsy1 may experience various events of mutation and homologous recombination. Polyploid or whole genome duplication is a common phenomenon in plants, which promotes the evolution and diversity of organisms[60]. WGD events played an important role in regulating the evolution of chrysanthemum. It was found that many ornamental and medicinal traits of chrysanthemum were related to gene replication[10]. In addition, gene replication is closely related to flower fragrance[30], flower shape[61], flowering[62], flower development[10,63], resistance[64] and circadian rhythm[65].

      Gene duplication usually results in three different situations: becoming a pseudogene; acquiring a novel function; or subfunctionalization[66]. However, random mutations may not be helpful to human social production. Some studies have suggested that the polyploidy formed by whole genome replication leads to a greater phenotypic breadth in which natural and artificial selection can work, thus making successful domestication possible[67]. Through the domestication of ornamental plants, the variation traits that meet people's needs have been continuously accumulated and strengthened, and the wild varieties have finally formed high-quality modern cultivated species.

    • Domestication and evolution have profoundly shaped the genomic structure and genetic diversity of crops today. The in-depth study and exploration of the genetic mechanism of crop domestication can help us to understand and identify the wild ancestral species of ornamental plants[68]; determine the time and place of the domestication[69]; clarifiying whether it comes from a single domestication event or multiple domestication events[69,70]; understanding the differences in morphological, physiological and biochemical characteristics between ornamental plants and their ancestral species, as well as the genetic basis for such differences[71]; evaluating the changes and causes of the genetic structure of crop populations compared with their ancestral species[72].

      Rosaceae contains nearly 3,000 species, many of which have been domesticated by humans thousands of years ago. The whole genome sequence of Prunus mume was assembled and nine ancestral chromosomes of Rosaceae were reconstructed[7]. It is speculated that they evolved from the same ancestor[7]. Based on the resequencing of Prunus mume, the phylogenetic tree of Prunus mume was constructed by using the core genes of 13 species of Prunus and three related species of Rosaceae, and the evolutionary history of Prunus was reconstructed. The results show that the divergence times between Prunus mume and other Prunus species as ~3.8 Mya and that between wild Prunus mume and cultivated Prunus mume is ~2.2 Mya. It is concluded that divergent selection may have played a role in the differentiation of these two subspecies long before the domestication of Prunus mume[48]. Four thousand to five thousand years ago, peach experienced a primitive domestication bottleneck in China, which showed a significant decline in the diversity of wild Prunus davidiana and the domesticated Asian cultivar[73]. The cultivation and domestication path of peach was analyzed based on the whole genome of peach and 14 other species of Prunus resequenced. It is considered that the effect of the second bottleneck is clearly reflected in the reduction of nucleotide diversity observed during the transfer from eastern to western cultivar. Compared with eastern and wild relatives, these bottlenecks seem to lead to a serious loss of diversity of western varieties, as well as obvious deficiencies of rare varieties and relatively slow LD decay[18]. The de novo chromosome-level reference genome of Cerasus serrulata was assembled[74]. Through 656 single-copy orthologs, the authors constructed a phylogenetic tree containing 10 species. It was found that Cerasus serrulata diverged from Prunus yedoensis about 17.34 million years ago, while Cerasus serrulata and Cerasus avium diverged from 21.44 million years ago. These results confirm that Cerasus serrulata are more closely related to Prunus yedoensis.

      Lotus is an important aquatic plant, which has many values such as medicine, edible and ornamental. Nineteen individuals of lotus, including three cultivated temperate lotus (rhizome lotus, seed lotus and flower lotus), one wild temperate lotus (wild lotus), one tropical lotus (Thai lotus) and one outer group (Nelumbo lutea). Through the analysis of genetic diversity, it was confirmed that wild Thai lotus showed greater differentiation and higher genomic diversity than cultivated lotus[75]. A recent re-sequencing study of 296 lotus germplasm resources further confirmed this conclusion. The analysis of population structure and genomic diversity of three types of cultivated lotus: rhizome, flower, and seed lotuses showed that the seed and rhizome lotus groups had not originated from a single source, but had a more complex multi-source[76].

    • With economic development, people have put forward more requirements for the color, fragrance and shape of ornamental plants. In recent decades, breeders are committed to identifying important ornamental gene loci and regulation patterns. Thanks to the high-quality reference genome, once the ornamental major genes are located, the corresponding molecular markers can be developed, and the mutant plants can be obtained by genetic transformation (Supplemental Table S5). These genes related to important traits are reviewed below.

    • Plant phenology, including flowering phenology, is an important parameter for crop growth monitoring, yield prediction and growth simulation[77]. Understanding the phenology of ornamental plants, especially flowering phenology, can help producers determine the time of sowing, hybridization and harvest. Therefore, many studies have been devoted to analysis of the molecular mechanisms of bud dormancy, flowering, flower development and flowering regulation.

      Dormancy is an adaptive process that enables plants to withstand difficult environmental conditions in winter temperate climate[78]. The study of the molecular mechanism of dormancy is of great significance for perennial temperate ornamental plants. However, compared with other crops, the molecular mechanism of ornamental plants affecting dormancy is still in the primary stage, and the dormancy studies of many important ornamental plants remain in the stage of physiological experiments. At present, only Paeonia suffruticosa[79], Prunus mume[80], Prunus persica[81] and a few others have isolated dormancy-related genes. The verification of gene function is not enough due to the lack of genetic transformation systems. The CBF gene of Prunus persica was isolated and overexpressed in Malus × domestica. Compared with the untransformed apple lines, the transgenic plants with the PpCBF1 gene showed growth cessation and leaf senescence under short-day treatment. This is the first example of short-day light-induced dormancy caused by structural overexpression of the CBF gene[82]. MADS-box gene (PmDAM6) was found expressed in the lateral buds of Prunus mume[83]. Overexpression of PmDAM6 in transgenic poplars showed that the plants exhibited growth cessation and terminal bud set when the stem tips of the control plants continued to grow, indicating that PmDAM6 has the function of growth inhibition[84].

      After the release of dormancy, flowering is a key event in the life cycle of seed plants, and there are six main ways to regulate flowering, including vernalization and autonomous pathways, photoperiod pathway and the circadian clock, gibberellin pathway, ambient temperature pathway, age pathway and meristem responses[85].The mechanisms of the six pathways have been well studied, many genes associated with them have been located in ornamental plants. Among them, a large number of genes have been verified. FLC, a MADS-box protein, is a key inhibitor in vernalization and autonomous pathways[86]. An AGL6-like gene CpAGL6, was cloned from Chimonanthus praecox. The transgenic experiment showed that the ectopic expression of CpAGL6 inhibited the vegetative growth of Arabidopsis thaliana and led to early flowering, which was mainly related to the inhibition of floral inhibitor FLC and floral promoters AP1 and FT[87]. In 1998, the LHY gene was confirmed to be related to the photoperiod pathway and the circadian clock of Arabidopsis thaliana[88]. LHY, TOC1 and CCA1 constitute the core negative feedback loop of the plant circadian clock[89]. A LHY/CCA1-like gene was identified in chrysanthemum and named CsLHY, CsLHY was fused with a gene encoding a short transcriptional repressor domain (SRDX) and constitutively expressed in chrysanthemum. The results showed that although the flowering transition was insensitive to photoperiod in CsLHY-SRDX transgenic plants, the further development of capitulum was blocked and no flowering was observed[90]. Gibberellin (GA) is an important plant hormone, which can promote flowering in annual and biennial plants, and may inhibit flowering in woody perennials[91]. Della protein is a repressor directly acting on the downstream of GA receptor, regulating plant growth and development induced by GA[91].Two Della homologous genes PmDELLA1 and PmDELLA2 in Prunus mume are overexpressed in Arabidopsis thaliana, resulting in plant dwarfing and flowering delay, indicating that these two Della proteins are negative regulators of GA signal in Prunus mume[92]. SVP gene plays an important role in the control of flowering time by ambient temperature[93]. A SVP homologous gene LaSVP, in Lavandula angustifolia was confirmed. The expression of LaSVP in Arabidopsis thaliana delayed flowering and affected the floral dosage-dependent in a dosage-dependent manner. MicroRNA156-mediated spl transcription factors can ensure flowering time under non-inductive conditions, which is an important age pathway[94]. It is found that CmNF-YB8 gene can regulate the early transformation of chrysanthemum from juvenile to adult, as well as early flowering, regardless of day-length conditions. In this process, members of the cmo-miR156 and SPL families, the core components of the age pathway, are involved[95]. During floral induction, the apical meristem changes from vegetative meristem to inflorescence meristem, which is related to the increased expression of SOC1 encoding MADS-box transcription factors[85]. Six homologous SOC1 genes were isolated from four tree peony cultivar, including PrSOC1, PdSOC1, PsSOC1, PsSOC1-1, PsSOC1-2 and PsSOC1-3. The ectopic expression of PsSOC1 in tobacco showed that the high expression of PsSOC1 in transgenic tobacco not only promoted plant growth, but also advanced flowering time[96]. It should be pointed out that in the process of flowering induction, several genes play a role in several tissues more than once, and these genes converge into a complex regulatory network of flower development.

      The longevity of cut flowers is an important trait that determines the sale of ornamental plants. Consumers appreciate long-lasting flowers, and the sales industry hopes to reduce the deterioration of flower quality in the distribution chain[97]. Senescence, as the last stage of flower development, determines the lifespan of ornamental plants. It has been found that the occurrence of aging is regulated by a number of phytohormones, including ethylene, abscisic acid (ABA), jasmonic acid (JA) and polyamine (PA), and cytokinin (CTK), which can inhibit aging[98]. The pathway of ethylene synthesis is relatively simple, in which ACC serves as the unique precursor, ACC synthase (ACS) and ACC oxidase (ACO) are key enzymes[99]. Many studies on the molecular mechanism of ethylene-mediated flower senescence focus on these three key enzymes. At present, several studies have confirmed that ACC synthase and ACC oxidase are closely related to the senescence of ornamental plants[100,101], furthermore, the regulation mechanism of ethylene on flower senescence has been well verified in many plants, such as Petunia hybrida[102], Ipomoea nil[103], Paeonia suffruticosa[104], and Dianthus caryophyllus[105,106]. However, many ornamental plants are not regulated by ethylene, ethylene has little effect on petal senescence in lilies, tulips, chrysanthemums and gladiolus[107]. ABA is considered to be an important regulator of petal senescence in ethylene-insensitive ornamental plants[108], and the effect of abscisic acid on senescence has been proved in many ethylene-insensitive ornamental plants such as gladiolus[109] and lily[110]. It should be noted that it is not a single phytohormones that regulates senescence in ornamental plants. Multiple hormones antagonize each other and jointly regulate the senescence process of plants[111]. At present, there have been many studies on the molecular mechanism of the antagonistic role of plant hormones such as gibberellin[112,113] and salicylic acid[114] in the process of petal senescence.

    • Flowers are unique reproductive organs of angiosperms. For the wild species, the evolutionary direction of plants has always been to promote the reproductive success of plants by changing the shape, color, smell and reward to attracting pollinators[115]. However, the wild species do not fully meet people's needs, and the customer's demand for purposeful and unique horticultural plants urges scientists to constantly improve the shape of flowers. The morphology of ornamental plants mainly includes floral organ structure, flower branch growth state, inflorescence type and plant morphology. Improving floral size, double flowered, number of flower branches and inflorescence type is an important direction of morphological improvement.

      In angiosperms, the basic developmental system of floral organs is explained by the ABCDE model[116]. In this model, the genes related to flower development are divided into five categories, namely A, B, C, D and E, in which A- and E-class determine sepal development, A-, B- and E-class determine petal development, B-, C- and E-class determine stamen development, C- and E-class determine carpels development and D- and E-class determine ovule development. Plants such as Antirrhinum majus, Petunia hybrida and Arabidopsis thaliana have made important contributions to the proposal and perfection of flower development models[117119], its own genes related to flower development have also been well studied. A large number of genes related to flower development are located in model plants which provides great convenience for the subsequent study of genes related to other ornamental plants. Taking Antirrhinum majus as an example, after the B-class genes DEF and GLO were first located in Antirrhinum majus, its homologous genes were also cloned in Phalaenopsis Aphrodite[120], Eustoma grandiflorum[121], Torenia fournieri[122] and other plants. At present, A-[123], B-[124], C-[125,126], D-[126], and E-[87,127] class genes in ornamental plants has been verified by transgenic plants. In chrysanthemum, 14 C-class genes were cloned, of which seven belong to the CAG1 gene and seven belong to the CAG2 gene[128]. The chimeric repressors silencing technique was used to knock out C-class in chrysanthemum to form a multiple-petal phenotype. The expression of CAG1s and CAG2s chimeric repressors led to the morphological changes of pistil and stamens forming petaloid organs in disk florets[128]. In orchid, four SEP-like genes were cloned and divided into PeSEP1/3 and PeSEP2/4. Transcriptome data showed that all PeSEP genes were expressed in all flower organs. When PeSEP3 was silenced by VIGS, the sepal became leaf-like organs, and the characteristics of the epidermis and the content of anthocyanin and chlorophyll changed, while the silencing of PeSEP2 had little effect on flower phenotype[129].

      The floral size and the number of flowers may be the most easily observed traits, which directly affect the visual effect of the whole plant. At the same time, the size of flowers and the number of flowers also have an important effect on plant pollination[130]. However, there are few studies on the gene mapping of these traits in ornamental plants, only in Petunia hybrida[131,132], Helianthus annuus[133,134] and Aquilegia[135].

      The perianth, including the calyx and corolla, are the most prominent parts of a flower. The corolla of most plants has a beautiful shape and bright colors, which is the main ornamental part. A small number of plants, such as Strelitzia reginae, Begonia fuchsioides belong to calyx ornamental plants. The study of the number of petals (including double flower and other traits) and the number of calyx is of great importance to improve the ornamental quality of flowers. Through GWAS, QTL and other methods, scientists have mapped a number of genes related to the number of petals in Dianthus caryphyllus[125], Rosa chinensis[8,136,137], Prunus mume[48], Prunus persica[138] and other plants, and genetic transformation verification is also under way. An APETALA2 homologous gene RcAP2 from rose was overexpressed in Arabidopsis thaliana, increasing the number of petals. In addition, after silencing the RcAP2 gene in 'Old Blush', the petal number decreased significantly[136].

      The corolla of ornamental plants has a variety of shapes. Except for a few asymmetric flowers, most of the corolla are classified as radially symmetrical, Cruciform, Caryophyllaceous, Rosaceous, Campanulate, Tubular, Infundibuliform, Hypocrateriform, Urceolate, Rotate, and bilaterally symmetrical Butterfly-like or Papilionaceous, Ligulate, Bilipped or Bilabiate corolla. The molecular basis of flora symmetry, including related processes such as development, life cycle, and metabolism, is regulated by a specific transcription factor (TF)[139]. In core dicotyledonous plants, CYC2 is the main regulator of flower symmetry, which can independently control the bilaterally symmetrical traits of flowers[140142]. Some studies have shown that the CYC2 gene is at least involved in the evolutionary transformation of the symmetry of Brassicales, Malpighiales, Dipsacales, Asterales and Lamiales[143]. In addition, some genes such as RAD and DIV belonging to the MYB family also control symmetry[144]. In Antirrhinum majus, the antagonistic action that RAD has over DIV by competing for the DRIF proteins preventing the formation of the DRIF-DIV complex and, consequently, in the establishment of an asymmetric pattern of gene activity in the Antirrhinum flower meristem[145]. Six CmCYC2 genes in chrysanthemum were identified[146]. It was found that the ectopic expression of the CmCYC2 gene in Arabidopsistcp1 mutant changed the symmetry and flowering time of Arabidopsis thaliana[146].

      Organ fusion is another way of flower shape change, which can lead to greater morphological diversity of closely related species[147]. Fusion occurs in two ways: when the organ primordium cannot be separated from the floral meristem, or at the later stage of the individual primordium or even the full formation of the organ, the 'fusion program' was initiated to enable the establishment of a permanent connection between them[148]. The genesis of fusion organs is thought to be mediated by NAC transcription factors from the NAM/CUC3 subfamily and microRNA from the miR164 family, which mediates the formation of boundaries during meristem development[149151]. This conclusion has been further confirmed in model plants such as Antirrhinum majus[152], Petunia hybrida[153], Arabidopsis thaliana[154] and so on.

    • Flower color is the most important feature of plants to attract pollinators[155]. At the same time, the bright color of flowers is also an important factor to attract consumers. Although natural ornamental plants have plentiful colors, the colors of some important ornamental plants are limited. For example, roses and chrysanthemums lack blue varieties[156], mei and morning glory lack bright yellow[157], Cymbidium hybrida lack orange and brick-red[158]. Therefore, flower color improvement has always been an important goal for breeders. The main substances that determine flower color are the following three kinds: flavonoids, carotenoids and betaine[159].

      The most important substance of flavonoids is anthocyanins, which determine the pink, red, violet and blue of flowers[160]. It is the main contributor to color designation. It has been found that the key structural genes that catalyze the early and later stages of anthocyanin biosynthesis include phenylalanine ammonia lyase (PAL), cinnamate-4-hydroxylase (C4H), 4-coumarate: CoA ligase (4CL)chalcone synthase (CHS), chalcone isomerase (CHI), flavanone-3-hydroxylase (F3H), flavonoid 3'-hydroxylase (F3'H) and flavonoid 3',5'-hydroxylase (F3'5'H), flavonol synthase (FLS), UDP-glucose: flavonoid glucosyltransferase (UFGT) and methyl transferase (MT)[159,161163]. Many of these genes have been identified in ornamental plants. In chrysanthemum, seven structural genes CHS, F3H, F30H, DFR, ANS, 3GT and 3MT regulated by transcription factors CmMYB5-1, CmMYB6, CmbHLH24 and CmMYB7-1 were identified as key genes for anthocyanin biosynthesis[164,165]. In cultivated chrysanthemum, due to the lack of delphinidin-based anthocyanins, no violet/blue chrysanthemum was found. F3′5′H gene was introduced into eight chrysanthemum lines by Agrobacterium tumefaciens transformation, and it was found that the petal color of all cultivars turned blue due to the accumulation of delphinidin[166].

      Carotenoids are isoprenoid molecules related to plant colors such as yellow, and orange to red[167] and accumulate in nonferrous bodies in the form of fat-soluble products. Structural genes in the carotenoid biosynthesis pathway have been well described in a variety of plant model systems[168170]. Phytoene synthase (PSY), phytoene desaturase (PDS), ζ-carotene isomerase (Z-ISO), carotene isomerase (CRTISO), geranylgeranyl pyrophosphate synthase (GGPPS) are important enzymes involved in the synthesis of carotenoids[171]. The crtW gene encodes β-carotene ketolase (4,4'-b-oxygenase), which can guide the synthesis of pink to red carotenoid pigments. This gene is commonly found in bacteria, fungi and unicellular algae[172]. The crtW gene was isolated from marine bacteria and modified the carotenoid biosynthesis pathway of lotus. In transgenic plants, the color of petals changed from light yellow to dark yellow or orange[173]. In addition, carotenoid biosynthesis genes in many ornamental plants such as Narcissus pseudonarcissus[174], Primula vulgaris[175] and Dianthus caryophyllus[176] have also been identified. In addition, there is evidence that the biosynthesis pathway of flavonoids and carotenoids may be regulated by MYB transcription factors[175,177,178].

      Betaine is a nitrogenous compound whose colors range from red-violet betacyanins to yellow betaxanthins[179]. Similar to flavonoids, betaine accumulates in the vacuoles. At present, only in Caryophyllales (except for Caryophyllaceae and Molluginaceae) are stained with betaine[180,181]. Compared with other major types of plant pigments, the study on betaine biosynthesis lags seriously[180], the core biosynthetic pathway of these pigments has not been fully elucidated until recently[182,183]. At present, the research on the regulation and synthesis of betaine genes is mainly focused on crops such as Beta vulgaris[184,185] and Hylocereusundatus[186,187]. Only Mirabilis jalapa[188,189] has been researched in ornamental plants. The key enzyme gene in betaine biosynthesis was isolated from Mirabilis jalapa: MjDOD, and transformed it into E. coli by constructing a vector containing MjDOD gene, which induced its expression, and successfully obtained betaine[189].

    • Flower fragrance is an important trait of ornamental plants. The fragrance of plants can make people feel good, and the flower volatile compounds related to flower aroma are also widely used in perfumes, medicine and condiments among other things. At the same time, flower fragrance can also attract pollinators and promote flower fertilization[190]. More than 1,700 flower volatile organic compounds have been identified in 1,000 species of angiosperms[191], which come from a few synthetic metabolic pathways, including terpenoids, phenylpropanoids/benzenoids, and fatty acid derived biosynthesis pathways.

      Terpenoids, especially monoterpenes such as linalool, limonene, myrcene, and trans-b-ocimene, but also some sesquiterpenes such as farnesene, nerolidol, and caryophyllene, are common components of floral fragrance[192]. In orchids, the PbGDPS gene was identified in Phalaenopsis bellina, which can produce both precursor of monoterpenes (GDP) and the precursor of sesquiterpenes (FDP)[193,194]. Chan et al. identified the transcripts related to fragance in Vanda Mimi Palmer and found that the clones of 32 transcripts were related to sesquiterpene synthase, germacrene D synthase and tyrosine decarboxylase[195]. In peony, four wild tree peony species were profiled as integrative volatile, including Paeonia ostii, Paeonia rockii, Paeonia delavayi and Paeonia lutea. A total of 67 floral volatiles were identified and the terpenoids in Paeonia ostii and Paeonia rockii were the most abundant. Transcriptome sequencing showed that there were 17,967 DEGs, of which 116 were related to the accumulation of terpenoids. Among them, 1-deoxy-D-xylulose 5-phosphate synthase, geranyl pyrophosphate synthase, farnesyl pyrophosphate synthase and terpene synthase may be the main regulatory enzymes of Paeonia ostii and Paeonia rockii terpene biosynthesis[196].

      Phenylpropanoids/benzenoids are the second largest group of flower fragrance components[191]. The biosynthesis of these compounds is regulated not only by the time and space of flower development, but also by the rhythmic and the biosynthesis of precursors[197]. Phenylpropanoids/benzenoids are the main sources of fragrance in plants such as Prunus mume[198], Petunia hybrida[199], Syringa vulgaris[200] and Magnolia champaca[201]. At present, many coding enzyme genes involved in this pathway have been cloned and verified. Taking Petunia hybrida as an example, Ph BSMT1, Ph BSMT2, Ph BSMT3, Ph IGS1, Ph EGS1, Ph PAAS, Ph BPBT, Ph CFAT, Ph BSMT1 and Ph BSMT2 are known to catalyze the formation of benzoic acid and salicylic acid, which are the components of many plant fragrances[197,202207]. At the same time, the transcription factors ODO1[199], EOBI and EOBII[208] of phenylpropanoid-related compounds have been proven by genetic transformation experiments. In Prunus mume, benzyl alcohol and benzyl acetate were found to be the main sources of the aroma of Prunus mume. No benzyl alcohol or benzaldehyde reductase (BAR) activity was detected in the fragrant variety 'Fenghou'. It is inferred that the lack of benzyl alcohol synthesis of the 'Fenghou' variety is due to the low activity of PmBAR1-Fen and the low expression of PmBAR3[209].

      The proportion of fatty acid derivatives in plant volatiles is lower than that of terpenoids, phenylpropanoids / benzenoids. It is found that aliphatic compounds play an important role in the aroma sources of Dianthus caryophyllus[210], Centaurea aeolica[211], Plumeria rubra[212] and Antirrhinum majus[213]. Fatty acid derivatives in ornamental plants are mainly synthesized by C18 fatty acids, including linolenic and linoleic acid. The biosynthesis of fatty acid derivatives begins with stereospecific oxygenation catalyzed by the lipoxygenase (LOX) pathway[214]. At present, the studies on fatty acid derivatives are mainly focused on genes involved in the lipoxygenase pathway[215], but there are few studies on expression in flower organs.

    • Stress generally refers to biotic or abiotic stresses. Abiotic stresses including diseases, pests and weeds, and abiotic stresses include cold, high temperature, drought, waterlogging, and salinization. The earths' climate is diverse and complex, and the vegetation in different regions face different stresses, which pose a severe challenge to the production of ornamental plants.

      At present, the main strategies for crop disease control are still highly dependent on chemical pesticides. However, all kinds of pesticides usually cause direct or indirect harm to human beings and the natural environment, and it is a strong desire of consumers to move towards an environmentally friendly and healthy production system. In this context, the use of gene editing technology to obtain disease-resistant varieties accurately and efficiently has become the primary breeding strategy. A variety of ornamental plants have acquired the ability to resist diseases and insect pests through transgenic technology. In chrysanthemum, insect-resistant transgenic chrysanthemum plants were obtained by transferring the Bacillus thuringiensiscol1Ab gene[216]. Chrysanthemum containing the rice chitinase gene was transformed by Agrobacterium tumefaciens. It was found that their resistance to Botrytis cinerea was enhanced[217]. Transformants of chrysanthemums with resistance to cucumber mosaic virus can produce better flowers when attacked by the virus[218]. In lily, in addition to the study of anti-fungal[219] and anti-viral[220], there are also studies on nematode resistance[221,222]. Genetic transformation of disease resistance was also studied in Gladiolus[223225], Orchidaceae[226228], Petunia hybrida[229], Rosa chinensis[230] and other ornamental plants.

      Different climatic conditions put forward different requirements for the survival of ornamental plants, such as cold tolerance in northern areas and drought tolerance in desert areas. In order to ensure ornamental plants have a broader living space, it is imperative to improve plant resistance. In many plants/organisms, some genes have been found which encode and synthesize these stress protective compounds, which are mainly divided into three categories: 1) genes that encode the synthesis of osmolytes such as mannitol, glycine betaine, proline, heat shock proteins; 2) genes responsible for ion and water uptake and transport like aquaporins and ion transporters; and 3) genes regulating transcriptional controls and signal transduction mechanisms, for example MAPK and DREBI[231]. A heat shock protein synthesis gene RcHSP17.8 from Rosa chinensis was obtained and transformed into Arabidopsis thaliana by Agrobacterium tumefaciens transformation. It was found that these transgenic plants were more tolerant to to high temperature, salt, osmotic and drought stress[232]. The salt tolerance of three chrysanthemum varieties with wild chrysanthemum were compared through physiological experiments. The results showed that the salt tolerance of Chrysanthemum lavandulifolium and 'Jinba' were better than that of the other two varieties, while 'Xueshan' was salt sensitive. Based on the differential expression analysis of genes, it was found that the genes related to signal transduction, ion transport, proline biosynthesis, reactive oxygen species scavenging systems and flavonoid biosynthesis pathway were related to salt tolerance of chrysanthemum. Activation of mitogen-activated protein kinases (MAPKs) is an important pathway in eukaryotic signaling events, which plays a key role in plant defense response and growth and development[232]. Studies have shown that CmMKK4-CmMPK13 and CmMKK2-CmMPK4 may be involved in the regulation of salt tolerance in chrysanthemum, and the relationship between CmMKK9 and CmMPK6 and temperature stress in chrysanthemum[233].

    • Plant architecture generally refers to the morphology of the aboveground parts of higher plants, which are mainly determined by factors affecting branch branches, plant height and inflorescence morphology. At present, in the field of ornamental plants, the research on plant architecture mainly has two purposes: to make the plant shape more beautiful, to meet people's aesthetic needs (Weeping Prunus mume, turtuosa Prunus mume, Styphnolobium japonicum), and to make plants more adapted to the development of the horticulture industry, and to reduce management costs (dwarfing plants and compact branches). The weeping trait is very popular due to their beautiful shape and common appreciation of flowers and trees. Prunus persica, Prunus mume, Cerasus subhirtella and other Rosaceae plants as well as Ulmus pumila, Betula pendula have the phenomenon of hanging branches. The weeping tree phenotype in Prunus persica is located in the LG2 linkage group and is controlled by the Ppa013325 gene. It may be a potential plant gene regulator of gravity perception or response[234]. In the study of Prunus mume, the major QTL loci related to weeping traits have been mapped within a 0.29 Mb region on chromosome 7, and the only specific expression of Pm024213 in buds and stems was detected in PmWEEP interval. Functional annotation and membrane structure prediction showed that the gene was a transmembrane protein located in the chloroplast and containing thioredoxin domain[235]. In recent years, miniaturized and compact potted flowers and bonsai are becoming more and more popular, as the stalks of many ornamental plants are easily damaged, the dwarfing technology of ornamental plants shows more and more vitality. Dw/dw[236] and Tssd/tssd[237] have been identified to be related to the dwarfing of peach. In addition, SHI, rol and other genes have been well transformed in ornamental plants such as Euphorbia pulcherrima[238], Angelonia salicariifolia[239], and Mecardonia procumbens[240].

    • Ornamental plants are of great commercial value to the horticultural industry, and people have been committed to their ornamental genetic improvement for a long time, but the traditional breeding methods often have a long cycle and low breeding efficiency. Although researchers have obtained a large number of candidate genes by means of GWAS and genetic linkage analysis, only some of them have been verified by transgenic experiments. With the development of biotechnology, gene editing technology can accurately edit the DNA sequence and shorten the breeding cycle, which plays a great role in the breeding of ornamental plants. This part provides reference for modern ornamental plant breeding through an overview of general biotechnology.

    • Quantitative traits are continuous phenotypic variations, which cannot be strictly classified. Quantitative traits controlled by polygenes and are easily affected by the environment. Many ornamental traits of plants, including flowering time[241], double flower[8] and flower color[48], belong to quantitative traits. At present, several methods such as quantitative trait loci mapping (QTL), genome wide association study (GWAS), bulked segregant analysis (BSA), and other methods have been developed for the study of these quantitative traits.

      Quantitative trait locus analysis (QTL) is a widely used tool to analyze the genetic basis of complex traits based on the construction of genetic maps. The basic idea of QTL is to use the linkage information between markers and QTL, to detect the linkage degree between molecular markers and QTL by maximum likelihood method or regression analysis, so as to locate QTL location and estimate QTL effect[242]. Paterson et al. reported the first work on plant QTL[243]. Since then, QTL mapping has gradually become an important tool for functional gene mapping. After more than 30 years of development, the QTL location method has developed from single marker analysis to multi-interval mapping, and from static QTL location to dynamic QTL location. In the pas decade, combined with the continuous development of high-throughput sequencing technology, marker resources have become more abundant, and the construction of high-density genetic maps of some ornamental plants[244247] has become possible. On this basis, more and more major QTL of important ornamental traits of ornamental plants have been cloned[127,248], and considerable progress has been made in the study of QTL of ornamental plants. The F1 hybrid population 'Liuban' × 'Fentaichuizhi' of Prunus mume was selected as experimental material, and a large-scale molecular marker was developed and constructed with a high-density genetic linkage map of Prunus mume using the SLAF-seq technique, and then QTL mapping analysis was carried out on 15 important ornamental traits, such as growth, plant architecture and flower related quantitative. A total of 66 QTLs loci were detected, and 58 possible candidate genes were screened using the genome annotation information of Prunus mume[244].

      The genome wide association study (GWAS) is based on linkage disequilibrium. By analyzing the segregation characteristics of high-density molecular markers of a large number of individuals, researchers can select molecular markers associated with phenotypic variation of complex traits, and then analyze the genetic effects of these molecular markers on phenotypes[249]. This method directly uses the natural population, which can save a lot of time in constructing the population. Besides that, the method has large recombination variation and high mapping accuracy, and can identify many multi-allele/gene loci that have not been found in QTL mapping[249]. Although GWAS was originally mainly used for the analysis of complex genetic diseases in humans, over the past decade it has been gradually extended to plants, including some ornamental plants (Supplemental Table S6). In Prunus mume, researchers have identified significant quantitative trait loci through GWAS technology. The candidate regions associated with traits including petal color, stigma color, calyx color, bud color, staminal filament color, wood color, petal number, pistil character, bud aperture, and branching phenotype were identified. It is the first time the genetic structure of floral size, color, and structure, in terms of the number of loci, their genomic distribution and the magnitude and pattern of their effects in a woody plants has been clarified[48]. At present, researchers have mapped more than a hundred loci (Supplemental Table S6) in seven species of ornamental plants by using GWAS technology, which shows that the application of GWAS has greatly promoted the research of alleles of important traits in ornamental plants.

      Bulked segregant analysis (BSA) is a mapping method for rapidly functional gene mining using extreme phenotypic individuals. The main idea is to sequence two groups of individuals with opposite extreme phenotypes in the segregated population, and to compare whether there is a significant difference in allele frequencies at polymorphic sites between the two groups[250]. Compared with traditional genetic analysis methods, BSA only needs to identify individuals with different extreme phenotypes in the target population, regardless of the accuracy of other individuals, which greatly reduces the scale and cost by simplifying the program[251]. Because BSA does not need a large population, it is favored in the study of plant type traits of woody plants, such as the weeping trait of Malus halliana[252], Prunus persica[234] and Prunus mume[235], the dwarf trait of Prunus persica[234] and gene mining and marker development. Twentyt individuals each of weeping and upright F1 generation of 'Liuban' × 'Fantaichuihzi' were selected. Using the strategy of RNA-seq based on BSA, five QTL related to the weeping trait were detected on chromosome 7 and combined with WGCNA to identify a core candidate gene PmUGT72B3. The Pm024074 (PmUGT72B3) allele encodes a protein containing conserved coniferyl-alcohol glucosyltransferase, which is a key protein regulating lignin and IAA. Now, BSA analysis has become a powerful tool for mapping functional genes in addition to QTL and GWAS.

    • Virus-induced gene silencing (VIGS) is a technique based on RNA interference, which uses the instant knock-down of the target gene expression of the modified plant virus genome[253]. Compared with traditional gene function research methods, such as transgene, gene knock-down and antisense inhibition, VIGS has the advantage of a short test cycle, being independent of transgenic operationand has low cost and high throughput[254]. The photoperiod flowering responses of plants can be divided into three types: long-day (LD); short-day (SD); and day-neutral (DN). Under the condition of SD and LD, the expression of RCCO and RcCOL4 in rose was reduced by VIGS, which successfully delayed the flowering time[255]. The expression of PeERF1 was down regulated in orchids using CymMV-based virus-induced gene silencing (VIGS), and it was found that the nanoridges of the silenced plants reduced. It is proved that PeERF1, as a SHN ortholog transporter, participates in the morphological formation of lip epidermis at the end of flowering by regulating the development of nanoridges structure of Phalaenopsis[256].

    • The development of gene editing technology is the demand of modern plant breeding, which rapidly modifies the plant genome in an accurate and predictable way[257]. Many gene editing techniques have been developed, including Zinc-finger nucleases (ZFNs)[258], transcription activator-like effector nucleases (TALENs)[259] and clustered regulatory interspaced short palindromic repeat (CRISPR)[260]. Some studies have shown that CRISPR/Cas9 is obviously superior to TALENs and ZFNs in terms of design, construction difficulty, time, cost and modification efficiency. CRISPR has gradually become the first choice for plant gene editing[261263].

      Dihydrofavonol-4-reductase-B (DFR-B) is an important enzyme in the anthocyanin biosynthesis pathway. In Ipomoea nil, DFR is present as a small, tandemly arrayed three-gene family (DFR-A, DFR-B and DFR-C), in which DFR-B is the major gene that controls flower color. The DFR-B gene was removed from Ipomoea nil with CRISPR/Cas9. Among the 32 transgenic plants, 24 (75%) had low anthocyanin content whose performance is white[264].This is the first report to use CRISPR/Cas9 technology on flower color research. In lily, CRISPR/Cas9 was used to knock out the LpPDS gene of two kinds of lily and obtained completely albino, pale yellow and albino–green chimeric mutants[265]. CRISPR/Cas9 will significantly promote the development of improved ornamental horticultural crops and bring innovative solutions to sustainable and competitive production of novel ornamental plants.

    • It is predicted that nearly 600 species of flowering plants will be sequenced by July 2021 (www.plabipd.de/plant_genomes_pa.ep). Many research groups have proposed to sequence more plants[266268]. With the development of sequencing technology and the reduction of sequencing cost, plant genome data will show explosive exponential growth in the future. As such, comparative genomics and functional genomics, which are based on whole genome data, will develop more rapidly and infiltrate into all aspects of plant genetic research. Although the development of plant genomics technology has promoted the research of ornamental plants, there are many problems in genome sequencing, data utilization and achievement transformation.

      First, the sequencing and assembly of genomes is still a difficult problem. Many ornamental plants have experienced a long history of evolution and domestication, especially plant with homologous polyploids or extremely large genome sizes[269]. For genome assembly based on short reading (75_700 bp) data, repeat content, and high heterozygosity sequences are usually not well solved, resulting in the formation of chimeric sequences and fragmentation contigs[270]. Even if third-generation sequencing technology with longer read-lengths is used, the integrity of genome assembly of some ornamental plants is only about 80%[62]. Secondly, it is difficult to map genes related to key traits from a large amount of genome data in the process of molecular breeding of ornamental plants, which is largely due to the lack of suitable phenotypic data. The gene mapping of many crops only needs to obtain simple phenotypic data such as plant height, fruit weight and leaf area, while the ever-changing morphology of ornamental plants and the twisted branches of turtuosa Prunus mume and complex flower shapes of chrysanthemum not only make the measurement work extremely difficult, but also the effective estimation of the recorded character data is low, resulting in weak correspondence between the trait data and the allelic variation of candidate genes in the germplasm set. Thirdly, many ornamental plants (especially woody plants) lack effective transformation systems, so the new genes cannot be transformed into plants for cultivar creation. Finally, whether transgenic ornamental plants meet safety and health needs is also one of the problems facing ornamental plant breeders, although at least 50 species of ornamental plants have been transformed[271], few GM ornamental cultivar have passed field trials and obtained regulatory approval. In the biotechnology/GM crop database approved by (ISAAA), an international agricultural biotechnology application service, only three ornamental plants are listed: petunia, rose, and carnation[271]. This may be due to the potential toxicity and allergies to humans, potential environmental risks (such as opportunities for gene flow, adverse effects on non-target organisms, resistance evolution of weeds and insects). Genetically modified crops that are widely used to carry foreign genes face obstacles to promotion[272]. Although these problems have existed in the many years of development of ornamental plant genomics, researchers continue to use new methods and new ideas to overcome these difficulties, such as gene editing technology, with the rapid development of this technology.

      In the aspect of sequencing, DNA sequencing technology in the past produced either highly accurate short reads or less-accurate long reads. The long high-fidelity (HiFi) readings recently launched by PacBio broke this balance. The accuracy of HiFi readings generated consensus sequence from multiple passes of a single template molecule as high as 99.8%, and can produce readings with an average length of 13.5 kb[273]. After the progress of third-generation sequencing technology, the innovation of a genome assembly algorithm followed. HiFiasm[274], a new de novo assembler that strives to preserve the contiguity of all haplotypes. And HiCanu[275] is a modification of the Canu assembler which can accurate assemble segmental duplications, satellites, and allelic variants from high-fidelity long reads. The advantages of these new techniques have been shown in challenging sequencing tasks such as hexaploid California redwood[274] and autotetraploid cultivated alfalfa[276]. On this basis, the gaps, collapsed regions and unassigned sequences in the plant genome will be less, which will provide a guarantee for the comprehensiveness and accuracy of phenotypic analysis in the future. In the mapping of complex traits, researchers are no longer satisfied with the location of simple one-dimensional traits such as plant height and fruit diameter, but the combination of image analysis and QTL mapping to study complex phenotypic traits has gradually become a research hotspot. Many scholars have studied the gene mapping of plant morphology by using the morphometric model based on elliptic Fourier, and have made many gratifying achievements in the location of fruit shape[277] and leaf shape[278,279]. And a mature geometric morphometry platform has been established[280]. Unfortunately, there are few studies on the location of characters of ornamental plants, and the transformation of three-dimensional flower structure into available one-dimensional data is still one of the difficulties that need to be overcome.

      During the 1990's, with the revolution of genomics and the rise of systems biology, biosynthesis emerged[281]. Synthetic biology is the science of how to design and synthesize organisms. Specifically, synthetic biology is an advanced form of gene editing. It applies the principles of modularity and engineering design to synthesize cells by changing their DNA[282]. The study of synthetic biology was first carried out on bacterial cells such as Mycoplasma mycoides[283], and now it has been gradually extended to eukaryotes, including plants. We believe that with the analysis of more gene regulatory networks and metabolic networks of ornamental traits and the progress of biological gene synthesis, ornamental plants designed according to the needs of consumers will emerge in the near future. China has a long history of ornamental plant cultivation and is rich in ornamental plant resources, which is known as the 'mother of gardens'. It is estimated that there are at least 31,000 species of vascular plants in China, of which more than 6,000 species have ornamental value. Most of the ornamental plants, such as rhododendron, rose, camellia and mei, originated and are cultivated in China[284].

      The rapid development of genomics has greatly promoted the research of horticultural plants[285]. In recent years, with the continuous reduction of the cost of genome sequencing technology, more and more ornamental plants have their own genome data, but it is still a small portion of ornamental plant germplasm resources. It has become an arduous task for modern ornamental plant researchers to sequence ornamental plants in an orderly and reasonable manner, to carry out in-depth mining of important genetic resources, and to protect and utilize them. Therefore, breeders need to make full use of genomic theories and methods (including molecular markers, whole genome selection, genome editing and synthetic biology) to alleviate the bottleneck of existing variety innovation and utilization, analyze the genetic regulation mechanism of important ornamental characteristics, establish an efficient biological breeding technology system, and carry out variety creation.

      • This study was supported by the National Natural Science Foundation of China (No. 31870689), Forestry and Grassland Science and Technology Innovation Youth Top Talent Project of China (No.2020132608), and the National Key Research and Development Program of China (2018YFD1000401).

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

      • Copyright: © 2022 by the author(s). Published by 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 (2)  Table (2) References (285)
  • About this article
    Cite this article
    Li M, Wen Z, Meng J, Cheng T, Zhang Q, et al. 2022. The genomics of ornamental plants: current status and opportunities. Ornamental Plant Research 2:6 doi: 10.48130/OPR-2022-0006
    Li M, Wen Z, Meng J, Cheng T, Zhang Q, et al. 2022. The genomics of ornamental plants: current status and opportunities. Ornamental Plant Research 2:6 doi: 10.48130/OPR-2022-0006

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return