Search
2024 Volume 4
Article Contents
REVIEW   Open Access    

The China orchid industry: past and future perspectives

More Information
  • There are nearly 30,000 species of orchids globally, of which over 1,700 species are found in China. Orchids share a profound and intimate connection with Chinese society. With the rapid development of science and technology, China's orchid industry has flourished with many scientific and technological achievements. Here, we summarize the developmental history, current situation, latest research achievements, and industrialization technology of the orchid industry in China, and present a discussion and outlook on the future development direction of orchid research in China. This review unveils new prospects for the high-quality advancement of China's orchid industry.
  • 加载中
  • [1]

    Chen S. 1999. Flora republicae popularis sinicae. Beijing: Science Press.

    [2]

    Wang Z, Lv Q. 2021. Found 31 new species! Important achievements have been made in the special investigation of wild Orchidaceae in China. People's Daily Online. http://finance.people.com.cn/n1/2021/1206/c1004-32301050.html (Accessed 20 October, 2023).

    [3]

    Lin D, Ya J, Schuiteman A, Ma C, Liu C, et al. 2021. Four new species and a new record of Orchidinae (Orchidaceae: Orchideae) from China. Plant Diversity 43:390−400

    doi: 10.1016/j.pld.2021.05.003

    CrossRef   Google Scholar

    [4]

    Ya J, Zhang T, Pandey TR, Liu C, Han Z, et al. 2021. New contributions to Goodyerinae and Dendrobiinae (Orchidaceae) in the flora of China. Plant Diversity 43:362−78

    doi: 10.1016/j.pld.2021.05.006

    CrossRef   Google Scholar

    [5]

    Li J, Ya J, Ye D, Liu C, Liu Q, et al. 2021. Taxonomy notes on Vandeae (Orchidaceae) from China: five new species and two new records. Plant Diversity 43:379−89

    doi: 10.1016/j.pld.2021.01.009

    CrossRef   Google Scholar

    [6]

    Ya J, Lin D, Han Z, Cai L, Zhang Z, et al. 2021. Three new species of Liparis s.l. (Orchidaceae:Malaxideae) from Southwest China based on morphological characters and phylogenetic evidence. Plant Diversity 43:401−08

    doi: 10.1016/j.pld.2021.01.006

    CrossRef   Google Scholar

    [7]

    Zhou Z, Shi R, Zhang Y, Xing X, Jin X. 2021. Orchid conservation in China from 2000 to 2020: achievements and perspectives. Plant Diversity 43:343−49

    doi: 10.1016/j.pld.2021.06.003

    CrossRef   Google Scholar

    [8]

    Animal and Plant Department, National Forestry and Grassland Administration. 2021. Announcement of ministry of agriculture and rural affairs and national forestry and grassland administration (No. 15, 2021) (List of National Key Protected Wild Plants). http://www.forestry.gov.cn/search/272693 (Accessed 20 October, 2023).

    [9]

    Huang W, Fang Z, Zeng S, Zhang J, Wu K, et al. 2012. Molecular cloning and functional analysis of three FLOWERING LOCUS T (FT) homologous genes from Chinese Cymbidium. International Journal of Molecular Sciences 13:11385−98

    doi: 10.3390/ijms130911385

    CrossRef   Google Scholar

    [10]

    Li M, Luo Q, Zhao X, Ding C, Chi Z, et al. 2023. Divergence and relationships of Cymbidium tortisepalum and its kindred germplasms by EST-SSR markers. The Journal of Horticultural Science and Biotechnology 98:483−94

    doi: 10.1080/14620316.2022.2159887

    CrossRef   Google Scholar

    [11]

    Feng S, Lu J, Gao L, Liu J, Wang H. 2014. Molecular phylogeny analysis and species identification of Dendrobium (Orchidaceae) in China. Biochemical Genetics 52:127−36

    doi: 10.1007/s10528-013-9633-6

    CrossRef   Google Scholar

    [12]

    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

    [13]

    Li D, Zhu G. 2022. High-density genetic linkage map construction and QTLs identification associated with four leaf-related traits in lady's slipper Orchids (Paphiopedilum concolor × Paphiopedilum hirsutissimum). Horticulturae 8:842

    doi: 10.3390/horticulturae8090842

    CrossRef   Google Scholar

    [14]

    Yang Z, Zhang X, Zhang M, Zhu G, Lv F. 2006. A study on karyotype of Paphiopedilum villosum. Journal of Northwest A&F University (Natural Science Edition) 34:163−65

    doi: 10.3321/j.issn:1671-9387.2006.11.036

    CrossRef   Google Scholar

    [15]

    Yang Z, Zhu G, Lü F, Zhang X, Wang B. 2006. Studies on the karyotypes of eight species of Paphiopedilum subgenus brachypetalum. Acta Horticulturae Sinica 33:1015−20

    doi: 10.3321/j.issn:0513-353X.2006.05.016

    CrossRef   Google Scholar

    [16]

    Zhu G, Yang Z, Wang B, Lü F, Zhang X. 2011. Karyotypes of 12 species of Paphiopedilum subgenus Paphiopedilum. Journal of Tropical and Subtropical Botany 19:152−58

    doi: 10.3969/j.issn.1005-3395.2011.02.009

    CrossRef   Google Scholar

    [17]

    Zhang D, Zhu G, Ye Q, Chen H. 2013. Cytological observations on chromosome numbers in 50 hybrid cultivars and species of Phalaenopsis. Chinese Journal of Tropical Crops 34:1871−76

    doi: 10.3969/j.issn.1000-2561.2013.10.004

    CrossRef   Google Scholar

    [18]

    Zhu G, Huang S, Bao M, Wang Z, Lv F. 2012. Chromosome ploidy identification of spring Dendrobium cultivars. Abstracts of 2012 Symposium On Chromosome Ploidy Operation and Genetic Improvement of Horticultural Plants. 130−32.

    [19]

    Mo R, Leng Q, Huang M, Pen B, Tang Y, et al. 2009. A karyological study of fourteen species in eleven genera of the Orchidaceae. Acta Botanica Yunnanica 31:504−08

    Google Scholar

    [20]

    Leng Q, Mo R, Luo Y, Huang M, Deng X, et al. 2009. A karyological study of six species of Cleisostoma from Hainan. Journal of Wuhan Botanical Research 27:121−26

    Google Scholar

    [21]

    Zhang Z, He J, Song X, Wu S, Wu W, et al. 2019. Karyotype analysis of two Phalaenopsis (Orchidaceae) species. Molecular Plant Breeding 17:2974−81

    doi: 10.13271/j.mpb.017.002974

    CrossRef   Google Scholar

    [22]

    Lü F, Zhu G, Wang B, Zhang G. 2005. Karyotypes of four Cymbidium sinense cultivars in Guangdong. Journal of Tropical and Subtropical Botany 13:423−28

    doi: 10.3969/j.issn.1005-3395.2005.05.01

    CrossRef   Google Scholar

    [23]

    Zhu J, Liu Y, Zeng R, Li Y, Guo H, et al. 2014. Preliminarily study on formation and cytological mechanism of unreduced male gametes in different ploidy Phalaenopsis. Acta Horticulturae Sinica 41:2132−38

    Google Scholar

    [24]

    Xu S. 2019. Study on polyploid breeding technology of hybrid Cymbidium based on 2n male gamete pathway. Thesis. South China Agricultural University, CN. pp 5−8

    [25]

    Cai J, Liu X, Vanneste K, Proost S, Tsai WC, et al. 2015. The genome sequence of the orchid Phalaenopsis equestris. Nature Genetics 47:65−72

    doi: 10.1038/ng.3149

    CrossRef   Google Scholar

    [26]

    Chao Y, Chen W, Chen C, Ho H, Yeh C, 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

    [27]

    Zhang G, Liu K, Li Z, Lohaus R, Hsiao YY, et al. 2017. The Apostasia genome and the evolution of orchids. Nature 549:379−83

    doi: 10.1038/nature23897

    CrossRef   Google Scholar

    [28]

    Zhang W, Zhang G, Zeng P, Zhang Y, Hu H, et al. 2021. Genome sequence of Apostasia ramifera provides insights into the adaptive evolution in orchids. BMC Genomics 22:536

    doi: 10.1186/s12864-021-07852-3

    CrossRef   Google Scholar

    [29]

    Yan L, Wang X, Liu H, Tian Y, Lian J, et al. 2015. The Genome of Dendrobium officinale illuminates the biology of the important traditional Chinese Orchid herb. Molecular Plant 8:922−34

    doi: 10.1016/j.molp.2014.12.011

    CrossRef   Google Scholar

    [30]

    Niu Z, Zhu F, Fan Y, Li C, Zhang B, et al. 2021. The chromosome-level reference genome assembly for Dendrobium officinale and its utility of functional genomics research and molecular breeding study. Acta Pharmaceutica Sinica B 11:2080−92

    doi: 10.1016/j.apsb.2021.01.019

    CrossRef   Google Scholar

    [31]

    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

    [32]

    Han B, Jing Y, Dai J, Zheng T, Gu F, et al. 2020. A chromosome-level genome assembly of Dendrobium Huoshanense using long reads and Hi-C data. Genome Biology and Evolution 12:2486−90

    doi: 10.1093/gbe/evaa215

    CrossRef   Google Scholar

    [33]

    Zhang Y, Zhang G, Zhang D, Liu X, Xu X, et al. 2021. Chromosome-scale assembly of the Dendrobium chrysotoxum genome enhances the understanding of orchid evolution. Horticulture Research 8:183

    doi: 10.1038/s41438-021-00621-z

    CrossRef   Google Scholar

    [34]

    Xu Q, Niu S, Li K, Zheng P, Zhang X, et al. 2022. Chromosome-scale assembly of the Dendrobium nobile genome provides insights into the molecular mechanism of the biosynthesis of the medicinal active ingredient of Dendrobium. Frontiers in Genetics 13:844622

    doi: 10.3389/fgene.2022.844622

    CrossRef   Google Scholar

    [35]

    Sherpa R, Devadas R, Suprasanna P, Bolbhat SN, Nikam TD. 2022. First De novo whole genome sequencing and assembly of mutant Dendrobium hybrid cultivar 'Emma White'. GigaByte 2022:gigabyte66

    doi: 10.46471/gigabyte.66

    CrossRef   Google Scholar

    [36]

    Yuan Y, Jin X, Liu J, Zhao X, Zhou J, et al. 2018. The Gastrodia elata genome provides insights into plant adaptation to heterotrophy. Nature Communication 9:1615

    doi: 10.1038/s41467-018-03423-5

    CrossRef   Google Scholar

    [37]

    Xu Y, Lei Y, Su Z, Zhao M, Zhang J, et al. 2021. A chromosome-scale Gastrodia elata genome and large-scale comparative genomic analysis indicate convergent evolution by gene loss in mycoheterotrophic and parasitic plants. The Plant Journal 108:1609−23

    doi: 10.1111/tpj.15528

    CrossRef   Google Scholar

    [38]

    Bae EK, An C, Kang MJ, Lee SA, Lee SJ, et al. 2022. Chromosome-level genome assembly of the fully mycoheterotrophic orchid Gastrodia elata. G3 Genes|Genomes|Genetics 12:jkab433

    doi: 10.1093/g3journal/jkab433

    CrossRef   Google Scholar

    [39]

    Jiang Y, Hu X, Yuan Y, Guo X, Chase MW, et al. 2022. The Gastrodia menghaiensis (Orchidaceae) genome provides new insights of orchid mycorrhizal interactions. BMC Plant Biology 22:179

    doi: 10.1186/s12870-022-03573-1

    CrossRef   Google Scholar

    [40]

    Jiang L, Lin M, Wang H, Song H, Zhang L, et al. 2022. Haplotype-resolved genome assembly of Bletilla striata (Thunb.) Reichb.f. to elucidate medicinal value. The Plant Journal 111:1340−53

    doi: 10.1111/tpj.15892

    CrossRef   Google Scholar

    [41]

    Wang J, Xie J, Chen H, Qiu X, Cui H, et al. 2022. A draft genome of the medicinal plant Cremastra appendiculata (D. Don) provides insights into the colchicine biosynthetic pathway. Communications Biology 5:1294

    doi: 10.1038/s42003-022-04229-4

    CrossRef   Google Scholar

    [42]

    Yang F, Gao J, Wei Y, Ren R, Zhang G, et al. 2021. The genome of Cymbidium sinense revealed the evolution of orchid traits. Plant Biotechnology Journal 19:2501−16

    doi: 10.1111/pbi.13676

    CrossRef   Google Scholar

    [43]

    Ai Y, Li Z, Sun W, Chen J, Zhang D, et al. 2021. The Cymbidium genome reveals the evolution of unique morphological traits. Horticulture Research 8:255

    doi: 10.1038/s41438-021-00683-z

    CrossRef   Google Scholar

    [44]

    Sun Y, Chen G, Huang J, Liu D, Xue F, et al. 2021. The Cymbidium goeringii genome provides insight into organ development and adaptive evolution in orchids. Ornamental Plant Research 1:10

    doi: 10.48130/opr-2021-0010

    CrossRef   Google Scholar

    [45]

    Chung O, Kim J, Bolser D, Kim HM, Jun JH, et al. 2022. A chromosome-scale genome assembly and annotation of the spring orchid (Cymbidium goeringii). Molecular Ecology Resources 22:1168−77

    doi: 10.1111/1755-0998.13537

    CrossRef   Google Scholar

    [46]

    Fan W, He Z, Zhe M, Feng J, Zhang L, et al. 2023. High-quality Cymbidium mannii genome and multifaceted regulation of crassulacean acid metabolism in epiphytes. Plant Communications 4:100564

    doi: 10.1016/j.xplc.2023.100564

    CrossRef   Google Scholar

    [47]

    Li M, Liu K, Li Z, Lu H, Ye Q, et al. 2022. Genomes of leafy and leafless Platanthera orchids illuminate the evolution of mycoheterotrophy. Nature Plants 8:373−88

    doi: 10.1038/s41477-022-01127-9

    CrossRef   Google Scholar

    [48]

    Li X, Luo J, Yan T, Xiang L, Jin F, et al. 2013. Deep sequencing-based analysis of the Cymbidium ensifolium floral transcriptome. PLoS ONE 8:e85480

    doi: 10.1371/journal.pone.0085480

    CrossRef   Google Scholar

    [49]

    Sun Y, Wang G, Li Y, Jiang L, Yang Y, et al. 2016. De novo transcriptome sequencing and comparative analysis to discover genes related to floral development in Cymbidium faberi Rolfe. SpringerPlus 5:1458

    doi: 10.1186/s40064-016-3089-1

    CrossRef   Google Scholar

    [50]

    He C, Liu X, Teixeira Da Silva JA, Liu N, Zhang M, et al. 2020. Transcriptome sequencing and metabolite profiling analyses provide comprehensive insight into molecular mechanisms of flower development in Dendrobium officinale (Orchidaceae). Plant Molecular Biology 104:529−48

    doi: 10.1007/s11103-020-01058-z

    CrossRef   Google Scholar

    [51]

    Yang F, Zhu G, Wang Z, Liu H, Xu Q, et al. 2017. Integrated mRNA and microRNA transcriptome variations in the multi-tepal mutant provide insights into the floral patterning of the orchid Cymbidium goeringii. BMC Genomics 18:367

    doi: 10.1186/s12864-017-3756-9

    CrossRef   Google Scholar

    [52]

    Chen Y, Xu Z, Shen Q, Sun C. 2021. Floral organ-specific proteome profiling of the floral ornamental orchid (Cymbidium goeringii) reveals candidate proteins related to floral organ development. Botanical Studies 62:23

    doi: 10.1186/s40529-021-00330-9

    CrossRef   Google Scholar

    [53]

    Su S, Shao X, Zhu C, Xu J, Lu H, et al. 2018. Transcriptome-wide analysis reveals the origin of peloria in Chinese Cymbidium (Cymbidium sinense). Plant and Cell Physiology 59:2064−74

    doi: 10.1093/pcp/pcy130

    CrossRef   Google Scholar

    [54]

    Bielenberg DG, Wang YE, Li Z, Zhebentyayeva T, Fan S, et al. 2008. Sequencing and annotation of the evergrowing locus in peach [Prunus persica (L.) Batsch] reveals a cluster of six MADS-box transcription factors as candidate genes for regulation of terminal bud formation. Tree Genetics & Genomes 4:495−507

    doi: 10.1007/s11295-007-0126-9

    CrossRef   Google Scholar

    [55]

    Howe GT, Horvath DP, Dharmawardhana P, Priest HD, Mockler TC, et al. 2015. Extensive transcriptome changes during natural onset and release of vegetative bud dormancy in Populus. Frontiers in Plant Science 6:989

    doi: 10.3389/fpls.2015.00989

    CrossRef   Google Scholar

    [56]

    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 intemperate fruit trees. Frontiers in Plant Science 9:1990

    doi: 10.3389/fpls.2018.01990

    CrossRef   Google Scholar

    [57]

    Liu J, Ren M, Chen H, Wu S, Yan H, et al. 2020. Evolution of SHORT VEGETATIVE PHASE (SVP) genes in Rosaceae: implications of lineage-specific gene duplication events and function diversifications with respect to their roles in processes other than bud dormancy. The Plant Genome 13:e20053

    doi: 10.1002/tpg2.20053

    CrossRef   Google Scholar

    [58]

    Zhang J, Shi X, Liu H, Ma G, Zou Q, et al. 2018. Study on the differential accumulation of anthocyanin in different-colored Phalaenopsis based on transcriptomics. Molecular Plant Breeding 16:4530−42

    doi: 10.13271/j.mpb.016.004530

    CrossRef   Google Scholar

    [59]

    Zheng D, Wang Y, Ou Y, Ke Y, Yao Y, et al. 2021. Research advances of genes responsible for flower colors in Orchidaceae. Acta Horticulturae Sinica 48:2057−72

    doi: 10.16420/j.issn.0513-353x.2021-0444

    CrossRef   Google Scholar

    [60]

    Liang D, Yang F, Ye Q, Zhu G. 2019. Preliminary study on the function of NADPH: Protochlorophyllide Oxidoreductase (CsPORB) gene in Cymbidium sinense 'Dharma'. Journal of Tropical and Subtropical Botany 27:285−93

    doi: 10.11926/jtsb.3957

    CrossRef   Google Scholar

    [61]

    Zhu G, Yang F, Shi S, Li D, Wang Z, et al. 2015. Transcriptome characterization of Cymbidium sinense 'Dharma' using 454 pyrosequencing and its application in the identification of genes associated with leaf color variation. PLoS ONE 10:e128592

    doi: 10.1371/journal.pone.0128592

    CrossRef   Google Scholar

    [62]

    Li B, Zheng B, Wang J, Tsai W, Lu H, 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

    [63]

    Gao J, Liang D, Xu Q, Yang F, Zhu G. 2020. Involvement of CsERF2 in leaf variegation of Cymbidium sinense 'Dharma'. Planta 252:29

    doi: 10.1007/s00425-020-03426-x

    CrossRef   Google Scholar

    [64]

    Ai Y, Zheng Q, Wang M, Xiong L, Li P, et al. 2023. Molecular mechanism of different flower color formation of Cymbidium ensifolium. Plant Molecular Biology 113:193−204

    doi: 10.1007/s11103-023-01382-0

    CrossRef   Google Scholar

    [65]

    Zhang Y, Zhou T, Dai Z, Dai X, Li W, et al. 2020. Comparative transcriptomics provides insight into floral color polymorphism in a Pleione limprichtii Orchid population. International Journal of Molecular Sciences 21:247

    doi: 10.3390/ijms21010247

    CrossRef   Google Scholar

    [66]

    Huang LM, Huang H, Chuang YC, Chen WH, Wang CN, et al. 2021. Evolution of terpene synthases in Orchidaceae. International Journal of Molecular Sciences 22:6947

    doi: 10.3390/ijms22136947

    CrossRef   Google Scholar

    [67]

    Zhang Y, Li X, Wang Y, Tian M, Fan M. 2011. Changes in fragrance components of Dendrobium officinale during different flowering stages and in different parts of the flower. Chinese Journal of Agricultural Sciences 1:110−17

    Google Scholar

    [68]

    Zhao C, Yu Z, Teixeira da Silva JA, He C, Wang H, et al. 2020. Functional characterization of a Dendrobium officinale geraniol synthase DoGES1 involved in floral scent formation. International Journal of Molecular Sciences 21:7005

    doi: 10.3390/ijms21197005

    CrossRef   Google Scholar

    [69]

    Yu Z, Zhao C, Zhang G, Teixeira da Silva JA, Duan J. 2020. Genome-wide identification and expression profile of TPS gene family in Dendrobium officinale and the role of DoTPS10 in linalool biosynthesis. International Journal of Molecular Sciences 21:5419

    doi: 10.3390/ijms21155419

    CrossRef   Google Scholar

    [70]

    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

    [71]

    Jiang S, Liang F, Niu S, Zhang Y, Ma J, et al. 2020. Cloning and bioinformatics analysis of SjHMGR gene in Sedirea Japonica. Chinese Journal of Tropical Crops 41:521−28

    doi: 10.3969/j.issn.1000-2561.2020.03.014

    CrossRef   Google Scholar

    [72]

    Chuang Y, Hung Y, Tsai W, Chen W, Chen H. 2018. PbbHLH4 regulates floral monoterpene biosynthesis in Phalaenopsis orchids. Journal of Experimental Botany 69:4363−77

    doi: 10.1093/jxb/ery246

    CrossRef   Google Scholar

    [73]

    Ramya M, Lee SY, An HR, Park PM, Kim NS, et al. 2019. MYB1 transcription factor regulation through floral scent in Cymbidium cultivar 'Sael Bit'. Phytochemistry Letters 32:181−87

    doi: 10.1016/j.phytol.2019.06.007

    CrossRef   Google Scholar

    [74]

    Xu Q, Wang S, Hong H, Zhou Y. 2019. Transcriptomic profiling of the flower scent biosynthesis pathway of Cymbidium faberi Rolfe and functional characterization of its jasmonic acid carboxyl methyltransferase gene. BMC Genomics 20:125

    doi: 10.1186/s12864-019-5501-z

    CrossRef   Google Scholar

    [75]

    Peng PH, Lin CH, Tsai HW, Lin TY. 2014. Cold response in Phalaenopsis aphrodite and characterization of PaCBF1 and PaICE1. Plant and Cell Physiology 55:1623−35

    doi: 10.1093/pcp/pcu093

    CrossRef   Google Scholar

    [76]

    An F, Chan M. 2012. Transcriptome-wide characterization of miRNA-directed and non-miRNA-directed endonucleolytic cleavage using Degradome analysis under low ambient temperature in Phalaenopsis aphrodite subsp. formosana. Plant and Cell Physiology 53:1737−50

    doi: 10.1093/pcp/pcs118

    CrossRef   Google Scholar

    [77]

    Zhang C, Chen J, Huang W, Song X, Niu J. 2021. Transcriptomics and metabolomics reveal purine and phenylpropanoid metabolism response to drought stress in Dendrobium sinense, an endemic orchid species in Hainan Island. Frontiers in Genetics 12:692702

    doi: 10.3389/fgene.2021.692702

    CrossRef   Google Scholar

    [78]

    Zhao D, Shi Y, Senthilkumar HA, Qiao Q, Wang Q, et al. 2019. Enriched networks 'nucleoside/nucleotide and ribonucleoside/ribonucleotide metabolic processes' and 'response to stimulus' potentially conferred to drought adaptation of the epiphytic orchid Dendrobium wangliangii. Physiology and Molecular Biology of Plants 25:31−45

    doi: 10.1007/s12298-018-0607-3

    CrossRef   Google Scholar

    [79]

    Li J, Chen X, Hu X, Ma L, Zhang S. 2018. Comparative physiological and proteomic analyses reveal different adaptive strategies by Cymbidium sinense and C. tracyanum to drought. Planta 247:69−97

    doi: 10.1007/s00425-017-2768-7

    CrossRef   Google Scholar

    [80]

    Ren R, Wei Y, Ahmad S, Jin J, Gao J, et al. 2020. Identification and characterization of NPR1 and PR1 homologs in Cymbidium orchids in response to multiple hormones, salinity and viral stresses. International Journal of Molecular Sciences 21:1977

    doi: 10.3390/ijms21061977

    CrossRef   Google Scholar

    [81]

    Chen J, Lu H, Chen C, Hsu H, Chen H, et al. 2013. The NPR1 ortholog PhaNPR1 is required for the induction of PhaPR1 in Phalaenopsis aphrodite. Botanical Studies 54:31

    doi: 10.1186/1999-3110-54-31

    CrossRef   Google Scholar

    [82]

    Chang L, Chang HH, Chang JC, Lu HC, Wang TT, et al. 2018. Plant A20/AN1 protein serves as the important hub to mediate antiviral immunity. PLos Pathogens 14:e1007288

    doi: 10.1371/journal.ppat.1007288

    CrossRef   Google Scholar

    [83]

    Petchthai U, Yee CSL, Wong SM. 2018. Resistance to CymMV and ORSV in artificial microRNA transgenic Nicotiana benthamiana plants. Scientific Reports 8:9958

    doi: 10.1038/s41598-018-28388-9

    CrossRef   Google Scholar

    [84]

    Kuo SY, Hu CC, Huang YW, Lee CW, Luo MJ, et al. 2021. Argonaute 5 family proteins play crucial roles in the defence against Cymbidium mosaic virus and Odontoglossum ringspot virus in Phalaenopsis aphrodite subsp. formosana. Molecular Plant Pathology 22:627−43

    doi: 10.1111/mpp.13049

    CrossRef   Google Scholar

    [85]

    Zhao X, Liu D, Wang Q, Ke S, Li Y, et al. 2022. Genome-wide identification and expression analysis of the GRAS gene family in Dendrobium chrysotoxum. Frontiers in Plant Science 13:1058287

    doi: 10.3389/fpls.2022.1058287

    CrossRef   Google Scholar

    [86]

    Zeng X, Ling H, Yang J, Li Y, Guo S. 2018. LEA proteins from Gastrodia elata enhance tolerance to low temperature stress in Escherichia coli. Gene 646:136−42

    doi: 10.1016/j.gene.2018.01.002

    CrossRef   Google Scholar

    [87]

    Zhan X, Qian Y, Mao B. 2022. Metabolic profiling of terpene diversity and the response of Prenylsynthase-Terpene Synthase genes during biotic and abiotic stresses in Dendrobium catenatum. International Journal of Molecular Sciences 23:6398

    doi: 10.3390/ijms23126398

    CrossRef   Google Scholar

    [88]

    Zhu M, Wang Q, Tu S, Ke S, Bi Y, et al. 2023. Genome-wide identification analysis of the R2R3-MYB transcription factor family in Cymbidium sinense for insights into drought stress responses. International Journal of Molecular Sciences 24:3235

    doi: 10.3390/ijms24043235

    CrossRef   Google Scholar

    [89]

    Huang H, Wang H, Tong Y, Wang Y. 2020. Insights into the Superoxide Dismutase gene family and its roles in Dendrobium catenatum under abiotic stresses. Plants 9:1452

    doi: 10.3390/plants9111452

    CrossRef   Google Scholar

    [90]

    Wang T, Song Z, Wei L, Li L. 2018. Molecular characterization and expression analysis of WRKY family genes in Dendrobium officinale. Genes & Genomics 40:265−79

    doi: 10.1007/s13258-017-0602-z

    CrossRef   Google Scholar

    [91]

    Zhu G, Yang F, Lv F, Li Z, Chen H, et al. 2020. Research advances in orchid breeding and industrialization technology. Guangdong Agricultural Science 47:218−25

    doi: 10.16768/j.issn.1004-874X.2020.11.024

    CrossRef   Google Scholar

    [92]

    Vilcherrez-Atoche JA, Iiyama CM, Cardoso JC. 2022. Polyploidization in orchids: from cellular changes to breeding applications. Plants 11:469

    doi: 10.3390/plants11040469

    CrossRef   Google Scholar

    [93]

    Zeng R, Zhu J, Xu S, Du G, Guo H, et al. 2020. Unreduced male gamete formation in Cymbidium and its use for developing sexual polyploid cultivars. Frontiers in Plant Science 11:558

    doi: 10.3389/fpls.2020.00558

    CrossRef   Google Scholar

    [94]

    Xie L, Zhou S, Wang M, Zeng R, Guo H, et al. 2017. Creation and micropropagation of polyploids in Cymbidium hybridum. Acta Horticulturae 1167:107−14

    doi: 10.17660/actahortic.2017.1167.16

    CrossRef   Google Scholar

    [95]

    Luo Y, Fang N, Fan R, Huang M. 2022. Research progress of polyploid induction in orchids. Jiangsu Agricultural Science 50:6−13

    Google Scholar

    [96]

    Wang M, Zeng R, Xie L, Gao X, Zhang Z. 2011. In vitro polyploid Induction and its identification in Cymbidium sinense. Chinese Agricultural Science Bulletin 27:132−36

    Google Scholar

    [97]

    Li H, Long C, Zheng S, Li Z. 2005. Polyploid induction of Cymbidium iridioides and its biological characteristics. Acta Horticulturae Sinica 5:853

    doi: 10.3321/j.issn:0513-353X.2005.05.043

    CrossRef   Google Scholar

    [98]

    Zhang X, Gao J. 2020. In vitro tetraploid induction from multigenotype protocorms and tetraploid regeneration in Dendrobium officinale. Plant Cell, Tissue and Organ Culture (PCTOC) 141:289−98

    doi: 10.1007/s11240-020-01786-6

    CrossRef   Google Scholar

    [99]

    Zhang X, Gao J. 2021. Colchicine-induced tetraploidy in Dendrobium cariniferum and its effect on plantlet morphology, anatomy and genome size. Plant Cell, Tissue and Organ Culture (PCTOC) 144:409−20

    doi: 10.1007/s11240-020-01966-4

    CrossRef   Google Scholar

    [100]

    Wu T, Jia R, Yang S, Zhao X, Yu X, et al. 2022. Research advances and prospects on Phalaenopsis polyploid breeding. Acta Horticulturae Sinica 49:448−62

    doi: 10.16420/j.issn.0513-353x.2020-0916

    CrossRef   Google Scholar

    [101]

    Wu Y, Ye R, Lv X, Wang W, Zhang M, et al. 2017. Colchicine induce polyploid from protocorm of Cremastra appendiculata. Plant Physiology Journal 53:407−12

    doi: 10.13592/j.cnki.ppj.2016.0396

    CrossRef   Google Scholar

    [102]

    Cheng F. 2013. Research about rapid propagation conditions optimization and polyploidy induction, identify of Nerviliae fordii (Herba) Schltr. Thesis. Guangzhou University of Chinese Medicine, CN. pp. 26−37.

    [103]

    Luan LQ, Uyen NHP, Ha VTT. 2012. In vitro mutation breeding of Paphiopedilum by ionization radiation. Scientia Horticulturae 144:1−9

    doi: 10.1016/j.scienta.2012.06.028

    CrossRef   Google Scholar

    [104]

    Zhuo Z. 2020. Study on Effect of Heavy Ion Irradiation and Mechanism of Thread Art Mutation on Leaf in Cymbidium. Thesis. South China Agricultural University, CN. pp. 1−28.

    [105]

    Chen X, Peng B, Xu M, Zhao G, Zheng P, et al. 2013. Breeding results and analysis of new Phalaenopsis variety 'Hangdie No. 2'. Satellite Application 11:61−65

    Google Scholar

    [106]

    Xu S, Zhang Y, Liang F, Jiang S, Niu S, et al. 2022. Cloning of PhaSEP3 gene in Phalaenopsis and its expression in floral organ mutants. Acta Agriculturae Zhejiangensis 34:1703−12

    doi: 10.3969/j.issn.1004-1524.2022.08.14

    CrossRef   Google Scholar

    [107]

    Tokuhara K, Mii M. 2001. Induction of embryogenic callus and cell suspension culture from shoot tips excised from flower stalk buds of Phalaenopsis (Orchidaceae). In Vitro Cellular & Developmental Biology – Plant 37:457−61

    doi: 10.1007/s11627-001-0080-4

    CrossRef   Google Scholar

    [108]

    Chen H, Lv F, Li Z, Xiao W. 2022. Research progress on the intergeneric hybridization breeding of Phalaenopsis spp. Journal of China Agricultural University 27:125−35

    doi: 10.11841/j.issn.1007-4333.2022.09.12

    CrossRef   Google Scholar

    [109]

    Xu W, Lin Y, Zhao Z, Zhou Z. 2022. Advances in genetic resources and breeding research of Cymbidium. Acta Horticulturae Sinica 49:2722−42

    doi: 10.16420/j.issn.0513-353x.2021-0943

    CrossRef   Google Scholar

    [110]

    Jiang J. 2020. Cross breeding technology in cattleya alliance. Thesis. South China Agricultural University, CN. pp. 9−43.

    [111]

    Li C, Dong N, Zhao Y, Wu S, Liu Z, et al. 2021. A review for the breeding of orchids: current achievements and prospects. Horticultural Plant Journal 7:380−92

    doi: 10.1016/j.hpj.2021.02.006

    CrossRef   Google Scholar

    [112]

    Cai X, Feng Z, Zhang X, Xu W, Hou B, et al. 2011. Genetic diversity and population structure of an endangered Orchid (Dendrobium loddigesii Rolfe) from China revealed by SRAP markers. Scientia Horticulturae 129:877−81

    doi: 10.1016/j.scienta.2011.06.001

    CrossRef   Google Scholar

    [113]

    Chen X, Guan J, Ding R, Zhang Q, Ling X. 2013. Conservation genetics of the endangered terrestrial orchid Liparis japonica in Northeast China based on AFLP markers. Plant Systematics and Evolution 299:691−98

    doi: 10.1007/s00606-012-0744-z

    CrossRef   Google Scholar

    [114]

    Zhao Y, Tang M, Bi Y. 2017. Nuclear genetic diversity and population structure of a vulnerable and endemic orchid (Cymbidium tortisepalum) in Northwestern Yunnan, China. Scientia Horticulturae 219:22−30

    doi: 10.1016/j.scienta.2017.02.033

    CrossRef   Google Scholar

    [115]

    Wu W, Chung Y, Kuo Y. 2017. Development of SSR markers in Phalaenopsis orchids, their characterization, cross-transferability and application for identification. In Orchid Biotechnology III, eds Chen W, Chen H. pp 91−107. https://doi.org/10.1142/9789813109223_0005

    [116]

    Li X. 2020. Study on technique of molecular marker-assisted flower color selection and identification of flower colour and fragrance gene in Cymbidium. Thesis. South China Agricultural University, CN. pp. 9−49.

    [117]

    Xiao W, Li Z, Chen H, Lv F. 2021. Identification and validation of single-nucleotide polymorphism markers linked to flower ground color in Phalaenopsis by using combined specific-locus amplified fragment sequencing and bulked segregant analysis. Journal of China Agricultural University 26:92−100

    doi: 10.11841/j.issn.1007-4333.2021.09.10

    CrossRef   Google Scholar

    [118]

    Chen H, Tsai W, Lv F, Xiao W, Li Z, et al. 2023. Research progress of genetic maps and QTL mapping in Orchidaceae. Journal of China Agricultural University 28:63−72

    doi: 10.11841/j.issn.1007-4333.2023.06.06

    CrossRef   Google Scholar

    [119]

    Hsu CC, Chen SY, Chiu SY, Lai CY, Lai PH, et al. 2022. High-density genetic map and genome-wide association studies of aesthetic traits in Phalaenopsis orchids. Scientifc Reports 12:3346

    doi: 10.1038/s41598-022-07318-w

    CrossRef   Google Scholar

    [120]

    Lu J, Liu Y, Xu J, Mei Z, Shi Y, et al. 2018. High-density genetic map construction and stem total polysaccharide content-related QTL exploration for chinese endemic Dendrobium (Orchidaceae). Frontiers in Plant Science 9:398

    doi: 10.3389/fpls.2018.00398

    CrossRef   Google Scholar

    [121]

    Prado JR, Segers G, Voelker T, Carson D, Dobert R, et al. 2014. Genetically engineered crops: from idea to product. Annual Review of Plant Biology 65:769−90

    doi: 10.1146/annurev-arplant-050213-040039

    CrossRef   Google Scholar

    [122]

    Hsieh RM, Chen WH, Hsu HM, Lin YS, Tsai WT, et al. 1997. Agrobacterium tumefaciens-mediated transformation of Phlaenopsis orchid. Report of Taiwan Sugar Research Institute 155:41−54

    Google Scholar

    [123]

    Wu R, Kang Y, Wang J, Liu J, Cui B, et al. 2015. Genetic transformation of ACO antisense gene into Dendrobium officinale Kimura et Migo. by Agrobacterium Mediation. Acta Agriculturae Boreali-Sinica 30:17−21

    doi: 10.7668/hbnxb.2015.02.004

    CrossRef   Google Scholar

    [124]

    Li J. 2005. High-efficient plantlets regeneration system and genetic transformation in Phalaenopsis, Oncidium and Cymbidium. Thesis. Nanjing Forestry University, CN. pp. 19−82.

    [125]

    Xie L, Wang F, Zeng R, Guo H, Zhou Y, et al. 2015. Agrobacterium-mediated transformation of Cymbidium sinensis. Chinese Journal of Biotechnology 31:542−51

    doi: 10.13345/j.cjb.140358

    CrossRef   Google Scholar

    [126]

    Zhang L, Chin DP, Mii M. 2010. Agrobacterium-mediated transformation of protocorm-like bodies in Cattleya. Plant Cell, Tissue and Organ Culture (PCTOC) 103:41−47

    doi: 10.1007/s11240-010-9751-3

    CrossRef   Google Scholar

    [127]

    Wang J, Zeng S, Chen Z, Wu K, Zhang J, et al. 2013. Genetic transformation of Doritis pulcherrima (Orchidaceae) via ovary-injection. Chinese Journal of Tropical Crops 34:1498−501

    doi: 10.3969/j.issn.1000-2561.2013.08.016

    CrossRef   Google Scholar

    [128]

    Zhou L. 2009. Exploration on high frequency regeneration system of orchid. Journal of Anhui Agricultural Sciences 35:8854−56

    doi: 10.3969/j.issn.0517-6611.2009.19.024

    CrossRef   Google Scholar

    [129]

    Xu J. 2011. Sudies on LyCYC1 transforming Saintpaulia ionantha and CFL transforming Cyclamen persicum and Cymbidium Goeringii × Cymbidium hybidium. Thesis. Hangzhou Normal University, CN. pp. 20−62.

    [130]

    Chen Z, Ming X, Liu R, Li Y, Chen Z. 2000. The Expression of GUS Gene in Orchid Protocorms After Bombardment. Chinese High Technology Letters 10:94−96

    doi: 10.3321/j.issn:1002-0470.2000.04.025

    CrossRef   Google Scholar

    [131]

    Cao Y, Hu S, Sun X, Lu X, Han Y. 2007. Study on Agrobacterium-mediated transformation of protocorm-like bodies of the hybrid orchid of Dendrobium × Phalaenopsis. Journal of Fujian Forestry Science and Technology 34:27−30

    doi: 10.3969/j.issn.1002-7351.2007.03.007

    CrossRef   Google Scholar

    [132]

    Jia S, Zeng R, Zhang Z, Wei Q, Xie L, et al. 2024. Research advances on molecular breeding technique of orchid. Journal of South China Agricultural University 45:1−14

    Google Scholar

    [133]

    Cao Z. 2018. Study on Agrobacterium-mediated transgenic technology of orchid. Thesis. South China Agricultural University, CN. pp. 21−54.

    [134]

    Wang W, Yang X, Yang J, Cheng Z, Zhu Y, et al. 2011. Study on Agrobacterium tumefaciens-mediated transformation of Cymbidium hybridum. Acta Botanica Boreali-Occidentalia Sinica 31:27−32

    Google Scholar

    [135]

    Jinek M, Chylinski K, Fonfara I, Hauer M, Doudna JA, et al. 2012. A programmable dual-RNA–guided DNA endonuclease in adaptive bacterial immunity. Science 337:816−21

    doi: 10.1126/science.1225829

    CrossRef   Google Scholar

    [136]

    Chen K, Wang Y, Zhang R, Zhang H, Gao C. 2019. CRISPR/Cas genome editing and precision plant breeding in agriculture. Annual Review of Plant Biology 70:667−97

    doi: 10.1146/annurev-arplant-050718-100049

    CrossRef   Google Scholar

    [137]

    Pan C, Qi Y. 2023. CRISPR-Combo-mediated orthogonal genome editing and transcriptional activation for plant breeding. Nature Protocols 18:1760−94

    doi: 10.1038/s41596-023-00823-w

    CrossRef   Google Scholar

    [138]

    Lu Y, Tian Y, Shen R, Yao Q, Wang M, et al. 2020. Targeted, efficient sequence insertion and replacement in rice. Nature Biotechnology 38:1402−07

    doi: 10.1038/s41587-020-0581-5

    CrossRef   Google Scholar

    [139]

    Kui L, Chen H, Zhang W, He S, Xiong Z, et al. 2016. Building a genetic manipulation tool box for orchid biology: identification of constitutive promoters and application of CRISPR/Cas9 in the orchid, Dendrobium officinale. Frontiers in Plant Science 7:2036

    doi: 10.3389/fpls.2016.02036

    CrossRef   Google Scholar

    [140]

    Tong CG, Wu FH, Yuan YH, Chen YR, Lin CS, et al. 2020. High-efficiency CRISPR/Cas-based editing of Phalaenopsis orchid MADS genes. Plant Biotechnology Journal 18:889−91

    doi: 10.1111/pbi.13264

    CrossRef   Google Scholar

    [141]

    Morel GM. 1960. Producing virus-free Cymbidiums. American Orchid Society Bulletin 29:495−97

    Google Scholar

    [142]

    Wang X, Chen J, Liu G, Gu M, Bao C. 1981. Clonal propagation of orchids by means of tissue culture. Acta Phytophysiologia Sinica 2:203−07

    Google Scholar

    [143]

    Zhang Z, Ou X. 1995. Tissue culture of Cymbidium sinense. Acta Horticulture Sinica 22:303−04

    Google Scholar

    [144]

    Wang Y, Yu D. 2005. In vitro culture of Cymbidium goeringii rhizomes. Acta Horticulture Sinica 32:706

    doi: 10.3321/j.issn:0513-353X.2005.04.046

    CrossRef   Google Scholar

    [145]

    Wimber DE. 1965. Additional observations on clonal multiplication of Cymbidiums through culture of shoot meristems. Cymbidium Society News 20:7−10

    Google Scholar

    [146]

    Churchill ME, Ball EA, Arditti J. 1972. Tissue culture of orchids. II. Methods for root tips. American Orchid Society Bulletin 41:726−30

    Google Scholar

    [147]

    Luo H, Chen R. 1997. Tissue culture and rapid propagation of Cymbidium sinense. Plant Physiology Journal 33:436−37

    Google Scholar

    [148]

    Tian M, Wang F, Qian N, Sun A. 1985. In vitro seed germination and Developmental morphology of seedling in Cymbidium ensifolium. Journal of Integrative Plant Biology 27:455−59

    Google Scholar

    [149]

    Zeng S, Wang J, Wu K, Teixeira da Silva JA, Zhang J, et al. 2013. In vitro propagation of Paphiopedilum hangianum Perner & Gruss. Scientia Horticulturae 151:147−56

    doi: 10.1016/j.scienta.2012.10.032

    CrossRef   Google Scholar

    [150]

    Huang S, Guo J, Ye Q, Lin J. 2003. Floral induction and development in Phalaenopsis under different temperatures. Acta Scientiarum Naturalium Universitatis Sunyatseni 42:132−34

    doi: 10.3321/j.issn:0529-6579.2003.04.035

    CrossRef   Google Scholar

    [151]

    Chen J, Lan H, Chen X, Zhang L. 2002. Influence of different cultivation management measures on the growth of Phalaenopsis. Journal of Forestry Engineering 1:29−30

    Google Scholar

    [152]

    Liu X, Ling X, Xiang L, Yu L, Shen H, et al. 2022. Effect of temperature and gibberellin on flowering regulation of Cymbidium goeringii. Acta Agriculturae Zhejiangensis 35:355−63

    doi: 10.3969/j.issn.1004-1524.2023.02.13

    CrossRef   Google Scholar

    [153]

    Zhang Y. 2008. Study in florescence control of Oncidium. Chinese Agricultural Science Bulletin 24:315−18

    Google Scholar

    [154]

    Song Z. 2008. Study on soilless culture and regulation of florescence of Phalaenopsis. Thesis. Agricutural University of Hebei, CN. pp. 1−55.

    [155]

    Huang X. 2013. Research on flowering regulation of Cymbidium goeringii and its physiological characteristics. Thesis. Guangxi Normal University, CN. pp. 1−30.

    [156]

    Pan R, Chen J, Wen Z. 1994. Influence of different potassium levels on growth, development and physiology in Cymbidiun sinense following potassium starvation. Journal of Tropical and Subtropical Botany 2:46−53

    Google Scholar

    [157]

    Cai Y, Guo Y, Hong H. 2010. Effect of N and K fertilization on the growth characteristics, flower quality and chemical componient Cymbidium orchid. Research Report on Agricultural Improvement Farm in Taichung District 109:15−27

    Google Scholar

    [158]

    Li A, Zhang Y, Sun J, Zhang J, Gong Z, et al. 2021. Effects of plant growth regulators and temperature on the rate of double peduncle, flowering period and flower characteristics of Phalaenopsis. Chinese Journal of Tropical Crops 42:732−38

    doi: 10.3969/j.issn.1000-2561.2021.03.018

    CrossRef   Google Scholar

    [159]

    Yin Y, Li J, Guo B, Li L, Ma G, et al. 2022. Exogenous GA3 promotes flowering in Paphiopedilum callosum (Orchidaceae) through bolting and lateral flower development regulation. Horticulture Research 9:uhac091

    doi: 10.1093/hr/uhac091

    CrossRef   Google Scholar

    [160]

    Chen M, Wang H, Zhu G. 2001. Studies on the flowering regulation of Phalaenopsis. Guangdong Agricultural Sciences 4:26−28

    Google Scholar

    [161]

    Xu J. 2017. Effects of paclobutrazolon growth and flowering of Dendrobium to my kids 'smile'. Research Report on Agricultural Improvement Farm in Taichung District 137:13−24

    Google Scholar

    [162]

    Li T, Huang J, Fu Z, Huang Z, Zhang J. 2021. Effects of the exogenous gibberellin and 6-benzylaminopurine on flowering of Cymbidium sinense. Northern Horticulture 21:64−71

    Google Scholar

    [163]

    Yang F, Xu Q, Zhu G. 2015. Effects of plant hormones on the regulation of flowering time of the orchid plant Cymbidium ensifolium. Advances in Ornamental Horticulture of China456−60

    Google Scholar

    [164]

    Wen J, Li W, Ding G, Xiao R. 2012. Occurrence and integrated management of the pests and diseases on orchid. Hunan Forestry Science & Technology 39:54−58

    doi: 10.3969/j.issn.1003-5710.2012.06.015

    CrossRef   Google Scholar

    [165]

    Huang F. 2011. Biological characteristics and chemical control of leaf butt rot of Phalaenopsis. Fujian Journal of Agricultural Sciences 26:808−11

    doi: 10.3969/j.issn.1008-0384.2011.05.026

    CrossRef   Google Scholar

    [166]

    Yuan H. 2017. Evaluationon resistance to stem rotand creation of germplasm resource in Cymbidium. Thesis. South China Agricultural University, CN. pp. 18−55.

    [167]

    Xu B, Zeng S, Song F, Liu H, Feng S, et al. 2014. Study on pathogen identification and biological characteristics of stem rot of Paphiopedilum and its indoor chemical control. Guangdong Agricultural Sciences 41:70−75

    doi: 10.3969/j.issn.1004-874X.2014.14.016

    CrossRef   Google Scholar

    [168]

    Shi Z, Zhang Y, Huang J, Xiang M. 2011. Identification of fusarium diseases on Cymbidium sinense and Phalaenopsis amabilis. Journal of Zhongkai University of Agriculture and Engineering 24:5−8

    doi: 10.3969/j.issn.1674-5663.2011.02.003

    CrossRef   Google Scholar

    [169]

    Li Y, Li X, Chen C, Li H. 2007. Isolation and identification of pathogens causing root rot disease of Cymbidium hybrida. Journal of Henan Agricultural University 41:85−89

    doi: 10.3969/j.issn.1000-2340.2007.01.020

    CrossRef   Google Scholar

    [170]

    Zhang X, Jiao Z, Yang X. 2017. Isolation and identification of the pathogens causing soft rot disease in the Phalaenopsis in the Yili. Contemporary Horticulture 13:5−7

    Google Scholar

    [171]

    Yang H, Zhou Y. 2010. Study on the Antagonistic Activity of Trichoderma viride and Bacillus subtilis against Fusarium acuminatum. Heilongjiang Agricultural Sciences 12:57−59

    doi: 10.3969/j.issn.1002-2767.2010.12.019

    CrossRef   Google Scholar

    [172]

    Zheng P, Liu R, Xu M, Zhao G, Wang T, et al. 2001. Comprehensive survey and research on the two main viruses causing orchid diseases in China. Guangdong Agricultural Sciences 6:37−40

    Google Scholar

    [173]

    Zheng Y, Li Y, Liu Y, Xu X, Ye M, et al. 2010. Identification of impatiens necrosis spot virus from Phalaenopsis amabilis in Yunnan. Acta Horticulturae Sinica 37:313−18

    Google Scholar

    [174]

    Zheng Y, Chen C, Chen Y, Jan FJ. 2008. Identification and characterization of a potyvirus causing chlorotic spots on Phalaenopsis orchids. European Journal of Plant Pathology 121:87−95

    doi: 10.1007/s10658-008-9281-6

    CrossRef   Google Scholar

    [175]

    Liu F, You S, Shen Z, Long L, Li X, et al. 2023. Diversity analysis of endophytic microbes in roots of Dendrobium nobile (Lindl.)(Orchidaceae) on two substrates. Journal of Southern Agriculture 54:1186−97

    doi: 10.3969/j.issn.2095-1191.2023.04.023

    CrossRef   Google Scholar

    [176]

    Wang H, Li C, Song F, Lin A, Meng Y. 2021. Isolation and identification of an endophytic bacterium from Bletilla striata and preliminary study of antibacterial mechanism. Journal of South-Central University for Nationalities (Natural Science Edition) 40:246−51

    Google Scholar

    [177]

    Wu W, Lu B, Zhang X. 2013. Isolation and purification of endophytic bacteria from Cymbidium goeringii. Shanhai Vegetables 2013:78−79

    Google Scholar

    [178]

    Chen Q, Liu B, Guan X, Tang J. 2014. Analysis of Anoectochilus roxburghii root microbial diversity by metagenomic technology. Journal of Agricultural Biotechnology 22:1441−46

    doi: 10.3969/j.issn.1674-7968.2014.11.015

    CrossRef   Google Scholar

    [179]

    Zhao Y, Guo S, Gao W, Du S. 1999. The symbiosis of three endophytic fungi with Cymbidium sp. and its effects on the mineral nutrition absorption. Acta Horticulture Sinica 26:110−15

    doi: 10.3321/j.issn:0513-353X.1999.02.009

    CrossRef   Google Scholar

    [180]

    Chen X, Guo S, Wang C. 2005. Effects of four endophytic fungi on the growth and polysaccharide content of Anoectochilus roxburghii ( Wall.)Lindl. Chinese Pharmaceutical Journal 40:13−16

    Google Scholar

    [181]

    Pan C, Chen R, Ye Q. 1999. Infection characteristics and physiological features of mycorrhizal fungi in roots of Cymbidium sinense and C. ensifolium. Soil and Environmental Sciences 8:3

    Google Scholar

  • Cite this article

    Yang F, Gao J, Li J, Wei Y, Xie Q, et al. 2024. The China orchid industry: past and future perspectives. Ornamental Plant Research 4: e002 doi: 10.48130/opr-0023-0024
    Yang F, Gao J, Li J, Wei Y, Xie Q, et al. 2024. The China orchid industry: past and future perspectives. Ornamental Plant Research 4: e002 doi: 10.48130/opr-0023-0024

Figures(1)  /  Tables(10)

Article Metrics

Article views(2107) PDF downloads(302)

REVIEW   Open Access    

The China orchid industry: past and future perspectives

Ornamental Plant Research  4 Article number: e002  (2024)  |  Cite this article

Abstract: There are nearly 30,000 species of orchids globally, of which over 1,700 species are found in China. Orchids share a profound and intimate connection with Chinese society. With the rapid development of science and technology, China's orchid industry has flourished with many scientific and technological achievements. Here, we summarize the developmental history, current situation, latest research achievements, and industrialization technology of the orchid industry in China, and present a discussion and outlook on the future development direction of orchid research in China. This review unveils new prospects for the high-quality advancement of China's orchid industry.

    • China is one of the pioneers in the cultivation of orchids and currently holds the largest orchid planting area in the world. Aromatic terrestrial orchids, often referred to as traditional Chinese orchids, display unique attributes, such as fragrance, purity, elegance, and harmony. Since the 1980s, following China's economic reforms, tropical orchids, including Phalaenopsis and Dendrobium, are introduced in abundance over time. These orchids become integral components of the floriculture industry in China. Simultaneously thriving, the native Chinese orchids and their tropical counterparts have propelled China to the forefront of orchid cultivation.

      Over the past two decades, the global orchid market has seen robust growth, but there has been a shift in production trends. Potted orchids now dominate over cut orchid flowers in both production and demand. Countries like China, the Netherlands, and the USA have now overtaken Japan and Korea as primary potted orchid producers. China has now emerged as a key player in cut orchid flower production, whereas countries such as the USA and Japan have faced a decline in their production.

      The evolution of the orchid industry sector is closely associated with the technological advancements. Through dedicated efforts of orchid researchers, China's orchid resources have been extensively investigated, and a rigorous orchid conservation system has been put in place. Significant strides have also been made in orchid breeding. New testing guidelines for DUS characteristics of varieties like Oncidium and Cattleya are now available. As of March 2022, there have been 1,212 applications for new orchid plant varieties, with 372 receiving approvals. Over 21 orchid species have their high-quality genomes mapped, and substantial research advancements have been achieved in cytology, physiology, genetics, reproduction, and cultivation of orchids.

    • China has fostered a unique orchid culture, embodying attributes like fragrance, purity, elegance, and harmony. The foundational era of this culture can be traced back to the Spring and Autumn Period (770 BC−221 BC). Notable figures like Confucius (551 BC−479 BC) has lauded orchids as the 'king of fragrance'. Goujian (496 BC−464 BC) has showcased his resilience by planting orchids, while Qu Yuan (340 BC−278 BC) has wore them as a symbol of his integrity. Post the Tang Dynasty (618−907), these plants have gained immense popularity among researchers, rulers, and military leaders.

      Unlike Western preferences, China holds a special reverence for particular fragrant flowers within the Cymbidium genus, commonly referred to as Chinese orchids. This includes species such as Cymbidium goeringii, Cymbidium faberi, Cymbidium sinense, Cymbidium ensifolium, Cymbidium tortisepalum, Cymbidium tortisepalum var. longibracteatum, Cymbidium serratum, Cymbidium kanran and Cymbidium szechuanicum. During the Song Dynasty, Huang Tingjian (1045−1105) was the first to classify Cymbidium goeringii and Cymbidium faberi, noting the difference based on the number of flowers in their inflorescence. Zhao Shigeng's (1233) renowned work, Jinzhang Lanpu, stands as the earliest global record of orchids, documenting various cultivars. Chunlan Tu, the artwork of Zhao Mengjian (1199−1264), is considered the oldest renowned orchid depiction. In the Ming Dynasty, Zhang Yu (1323−1385) extolled the beauty of Cymbidium foliage in his writings, emphasizing their value over the flowers, and detailed the graceful shape of Cymbidium leaves (Fig. 1ae). In his poetry, he mentions that 'Leaves are more valuable than flowers', and expressed his deep appreciation for leaf color and shape variegation. The Qing Dynasty witnesses Bao Yiyun (1708−1778) introducing varied petal types of Chinese orchids in his literary work Yi Lan Zha Ji (Fig. 1fo). This classification was subsequently expanded upon by several researchers (Table 1).

      Figure 1. 

      (a)−(e) Natural mutant of leaf colours and (f)−(o) natural mutant (varieties) of flower types in Chinese orchid, Bar = 1 cm.

      Table 1.  The number of Chinese orchid (Cymbidium) cultivars documented in the references.

      YearCymbidium goeringiiCymbidium faberiCymbidium sinenseCymbidium ensifoliumCymbidium tortisepalumCymbidium
      tortisepalum var.
      longibracteatum
      Cymbidium kanlanCymbidium serratum
      Total
      Authors/
      publications
      1233182139Zhao Shigeng, Jin Zhang Lan Pu
      1865253156Xu Jilou, Lanhui Tongxinlu
      1900221100123Ou Jinche, Linghai Lanyan
      19236640141Wu Enyua, Lanhui Xiaoshi
      199313270144617411429Wu Yingxiang, Chinese Cymbidium
      20101831452341561411681111191,257Chen Xinqi, Complete Book of Chinese Cymbidium and its varieties

      In modern times, coinciding with China's economic reforms, the 1980s marked a revival in the orchid industry. In 1987, the China Orchid Association was founded in Guangdong. This association inaugurated the first China Orchid Expo in 1988 in Guangzhou. Since then, this event has been held 30 times. The appreciation of orchids has evolved over time, with a shift from traditional petal shapes to a more diverse range, including various flower types and leaf color forms. This evolution has significantly boosted China's orchid industry.

      Since the 1980s, tropical orchids including Phalaenopsis, hybrid Cymbidium and Cattleya (commonly known as foreign orchids) have been introduced in large numbers, which have gained significant importance in China's flower industry (Tables 2 & 3). Domestic orchids and foreign orchids are developing hand in hand. The orchid planting scale of China ranks No.1 in the world. According to the Ministry of Agriculture's national flower statistics, in 2017, the country's orchid category expanded over an area of 11,664 hectares, accounting for 9.8% of the country's potted flower planting area. The potted flowers received a sale of 9.02 billion yuan, accounting for 23.6% of the country's potted flower sales.

      Table 2.  Hybrid Cymbidium and Phalaenopsis planting area in China from 2012 to 2020.

      YearHybrid Cymbidium (ha)Phalaenopsis (ha)Total (ha)
      20121,7215092,230
      20131,2946721,966
      20147535061,259
      20157125441,256
      20161,7676102,377
      20171,7065712,277
      20181,8887892,677
      20191,8677272,594
      20201,9048522,756

      Table 3.  Oncidiums and pot orchids planting area in Taiwan, China from 2009 to 2019.

      YearOncidiums (ha)Pot orchids (ha)Total (ha)
      2009213608821
      2010207645852
      2011217726943
      2012243662905
      2013249677926
      2014261691952
      2015261707968
      20162667451,011
      20172497511,000
      2018247750997
      2019237761998
    • China stands out as one of the global leaders in orchid diversity. As documented in the Flora Republicae Popularis Sinicae (1999), China boasts 171 genera and 1,247 orchid species[1]. In 2018, the National Forestry and Grassland Bureau initiated a comprehensive survey of wild Orchidaceae resources. This survey covered key orchid-rich provinces of Yunnan and Guizhou, and extended to regions of Tibet and Fujian[2]. By 2021, surveys span to provinces of Zhejiang and Shaanxi, with the project set to conclude in 2023. Current data reveal investigations of 78,776 quadrats, recording approximately 132,000 orchids from 1,258 species[2]. Notably, around 800 native orchid species are conserved ex-situ in botanical gardens, with a significant percentage protected within national or provincial nature reserves[2]. During these botanical explorations, 31 new species and a new genus, Microtatorchis, are identified (Table 4). Presently, China recognizes 181 genera and 1,708 orchid species, with five new genera and 365 new species described during the last 21 years[7].

      Table 4.  List of orchids newly discovered during the botanical investigation.

      GeneraSpeciesType of discoveryReferences
      HerminiumHerminium lijiangenseNew species[3]
      PeristylusPeristylus fasciculatusNew species[3]
      Peristylus tenuicallusNew recorded species[3]
      PlatantheraPlatanthera milinensisNew species[3]
      PonerorchisPonerorchis gongshanensisNew species[3]
      CheirostylisCheirostylis chuxiongensisNew species[4]
      Cheirostylis yeiNew species[4]
      MyrmechisMyrmechis lingulataNew species[4]
      Myrmechis longiiNew species[4]
      BulbophyllumBulbophyllum ximaenseNew species[4]
      Bulbophyllum xizangenseNew species[4]
      Bulbophyllum retusumNew species[4]
      Bulbophyllum pulcherissimumNew species[4]
      Bulbophyllum frostiiNew recorded species[4]
      Bulbophyllum raskotiiNew recorded species[4]
      Bulbophyllum nematocaulonNew recorded species[4]
      GastrochilusGastrochilus yeiNew species[5]
      Gastrochilus minimusNew species[5]
      LuisiaLuisia simaoensisNew species[5]
      Luisia inconspicuaNew recorded species[5]
      TaeniophyllumTaeniophyllum xizangenseNew species[5]
      TuberolabiumTuberolabium subulatumNew species[5]
      CleisostomaCleisostoma tricornutumNew recorded species[5]
      LiparisLiparis aureolabellaNew species[6]
      Liparis mengziensisNew species[6]
      Liparis bingzhongluoensisNew species[6]

      Stringent conservation measures have been implemented to recognize the significance of Orchidaceae in China. In 2018, key protection initiatives are introduced, with nearly 300 species across 23 genera receiving protection, including 41 primary and 256 secondary protected species (Table 5)[8]. Wild orchids under state protection constitute a significant portion of China's wild plants. As of June 2021, national nature reserves safeguarded around 1,100 species, with provincial reserves and botanical gardens conserving additional species[7].

      Table 5.  Wild orchids under special state protection in China.

      GeneraI Protected speciesII Protected species
      AeridesAerides odorata
      AnoectochilusAll species
      BletillaBletilla striata
      BulbophyllumBulbophyllum rothschildianum
      CalantheCalanthe striata var. sieboldiierCalanthe dulongensis
      ChangnieniaChangnienia amoena
      CorybasCorybas taliensis
      CremastraCremastra appendiculata
      CymbidiumCymbidium insigne and Cymbidium wenshanenseAll species except Cymbidium lancifolium
      CypripediumCypripedium subtropicumAll species except Cypripedium plectrochilum
      DanxiaorchisAll species
      DendrobiumDendrobium flexicaule and Dendrobium huoshanenseAll species
      GastrodiaGastrodia elata and Gastrodia angusta
      GymnadeniaGymnadenia conopsea and Gymnadenia orchidis
      LudisiaLudisia discolor
      PaphiopedilumAll speciesPaphiopedilum hirsutissimum and Paphiopedilum micranthum
      PhaiusPhaius hainanensis and Phaius wenshanensis
      PhalaenopsisPhalaenopsis zhejiangensisPhalaenopsis lobbii, Phalaenopsis wilsonii and Phalaenopsis malipoensis
      PleioneAll species
      RenantheraAll species
      RhynchostylisRhynchostylis retusa
      VandaVanda coerulea
      VanillaVanilla shenzhenica
    • Genetic diversity and relationships among Orchidaceae have been extensively researched using various molecular markers. Studies have delved into the genetic diversity of species like Cymbidium[9,10] and Dendrobium[11]. With the advent of high-throughput sequencing, numerous SNP markers have been identified in orchids. These markers have been instrumental in studying the origin, evolution, and genetic diversity of species such as Dendrobium[12] and Phalaenopsis[13].

    • Chromosomal studies have revealed a rich diversity in orchid species. The chromosome number has been observed to be x = 11, 13, 18, 19, 20, 27, and so on. For example, most Paphiopedilum species have 2n = 26 chromosomes (x = 13). However, the chromosome number of different species in the Paphiopedilum subgenus Paphiopedilum section Barbata varied, with 2n = 32 for Paphiopedilum callosum, 2n = 38 for Paphiopedilum microchilum, Paphiopedilum appletonianum and Paphiopedilum hainanensis[1416]. There are diploids, triploids and tetraploids in the cultivated cultivars of Phalaenopsis, and many aneuploids are produced in the process of cross breeding[17]. Zhu et al.[18] identified the chromosome ploidy of 57 introduced Japanese Dendrobium cultivars and 20 native species by flow cytometry and chromosome compression method. The results have showed that the coincidence rate of the two methods is 86.4%. Among the spring Dendrobium hybrids introduced from Japan, tetraploid cultivars account for 64.9%, triploid cultivars account for 28.1%, the diploid and hexaploid cultivars account for 3.5% respectively, indicating that most of the cultivated hybrid species of Dendrobium were polyploid (96.5%). In the study of chromosomes of several genera of Subtrib. Aeridinae Pfitz. their chromosome bases are all x = 19. Among these, Renanthera coccinea is diploid and hexaploid, Vanda concolor is diploid and tetraploid, Vanda pumila is tetraploid and the rest are diploid[1921]. Most of the Cymbidium species exhibit a chromosome number of 2n = 40, except for Cymbidium lancifolium (2n = 38) and Cymbidium serratum (2n = 41, 43, 60 and 80)[22]. Additionally, the formation of natural 2n gametes in orchids like Phalaenopsis[23] and Cymbidium hybridum[24] has been explored, paving the way for future polyploid breeding endeavors.

    • Leveraging the swift advancements in high-throughput sequencing technology, genomic data of 26 orchids have been disclosed, and notably, Chinese researchers have made significant contributions to 20 of these datasets (Table 6). These research studies indicate that a distinct Whole Genome Duplication (WGD) event takes place in the Orchidaceae lineage around the Cretaceous/Paleogene transition. Subsequently, within a short timeframe, differentiation into five subfamilies transpired, namely Apostasioideae, Cypripedioideae, Vanilloideae, Orchidoideae, and Epidendroideae. Notably, the subfamily Apostasioideae exhibits the highest gene loss, while the subfamily Vanilloideae has the least. Phalaenopsis equestris is the first orchid to be sequenced[25]. Three years after the publication of the Phalaenopsis equestris genome in 2015, researchers in Taiwan sequenced and assembled the Phalaenopsis aphrodite genome, which becomes the first orchid genome to be integrated with SNP-based gene-linkage mapping and validated through physical mapping[26]. In 2017, the publication of the Apostasia shenzhenica genome by Chinese researchers in the journal Nature triggered a new era of discussion about Darwin's conjecture in the scientific community. The successful assembly of Apostasia shenzhenica genome reconstructed the 'genetic toolkit' of ancestral orchids[27], and revealed the species-specific adaptive evolution of orchids, including the epiphytic adaptation, floral differentiation and reversion, and pollen aggregation into pollinia[28].

      Table 6.  Orchid genome published by Chinese scientists.

      SpeciesGenome size (Gb)Chromosome numberContig N50Scaffold N50Protein-coding genesReferences
      Apostasia shenzhenic0.352N = 2X = 680.083.0321,841[27]
      Apostasia ramifera0.370.030.2922,841[28]
      Bletilla striata5.062N = 2X = 322.371.65146.39[40]
      Cymbidium sinense3.522N = 2X = 401.1129 638[42]
      Cymbidium goeringii4.12N = 2X = 401.04209.0429,272[44]
      Cymbidium goeringii3.992N = 2X = 400.38178.229,556[45]
      Cymbidium ensifolium3.622N = 2X = 401.21154.8829,073[43]
      Dendrobium officinale1.352N = 2X = 380.0010.0435,567[29]
      Dendrobium catenatum1.012N = 2X = 380.030.3928,910[31]
      Dendrobium huoshanense1.2852N = 2X = 380.671.7921,070[32]
      Dendrobium chrysotoxum1.372N = 2X = 381.5467.830,044[33]
      Dendrobium officinale1.232N = 2X = 381.1463.0725,894[30]
      Dendrobium nobile1.192N = 2X = 381.6264.4629,476[34]
      Gastrodia elata1.120.074.918,969[37]
      Gastrodia elata1.0432N = 2X = 3621.3321,115[38]
      Gastrodia menghaiensis0.9872N = 2X = 362.376.8214,233[39]
      Phalaenopsis equestris1.0860.020.3629,431[25]
      Phalaenopsis aphrodite1.0252N = 2X = 380.020.9528,902[26]
      Platanthera guangdongensis4.192N = 2X = 421.77192.3524,513[47]
      Platanthera zijinensis4.21.57193.1422,559

      In China, orchids are been revered for millennia for their medicinal properties. They have garnered significant attention, especially those with therapeutic and nutritional benefits. Seven Dendrobium species have undergone whole genome sequencing, includes Dendrobium officinale[29, 30], Dendrobium catenatum[31], Dendrobium huoshanense[32], Dendrobium chrysotoxum[33], Dendrobium nobile[34] and Dendrobium × sp.[35]. Successful assembly of Dendrobium genomes provides an analytical basis for in-depth understanding of the formation mechanism of Dendrobium medicinal components and the regulation of Dendrobium polysaccharide synthesis. Gastrodia elata is a common traditional Chinese medicine used to treat neurological headaches, numbness and convulsions. As a fully mycoheterotrophic orchid, the mechanism of interactions between Gastrodia elata and fungi has been a hot research topic[3638]. Genomic analysis of Gastrodia menghaiensis suggests that the proteome of Gastrodia menghaiensis is the smallest proteome thus far among angiosperms[39]. The Chinese herbal medicine Bletilla striata has the effects of astringency, blood coagulation, subduing swelling and promoting muscle growth. By assembling and analyzing the Bletilla striata genome, transcription factor BsMYB2 was found to regulate the biosynthesis of BSPs[40]. Cremastra appendiculata has the medicinal effects of relieving heat and removing toxins, resolving phlegm and dispersing knots. Based on its genomic information, the researchers have identified 35 key genes in the colchicine biosynthesis pathway, and systematically elucidated the phylogenetic relationship of O-methyltransferase[41].

      Cymbidium species hold the most cultural significance in China. Genomic sequencing has been completed for Cymbidium sinense[42], Cymbidium ensifolium[43], and Cymbidium goeringii[44, 45]. The Cymbidium sinense genome reveals the molecular regulatory mechanisms of important horticultural traits such as flowering time, floral morphology, color and fragrance[42]. Both Cymbidium ensifolium[43] and Cymbidium goeringii genomes shed light on the genetic basis of floral diversity[45]. In 2023, the Cymbidium mannii genome is reported[46], combined with other multi-omics data, the mechanisms of CAM photosynthesis in epiphytic plants has been revealed in this work.

    • Floral pattern is a distinctive feature of orchids, and understanding its formation has been a focal point of research. There are two main directions for research on orchid floral patterns, one direction is to study the flower development mechanism, and the other direction is to study the specific formation mechanism of flower organs. The transcriptome sequencing of Cymbidium ensifolium was conducted to analyze gene expression during the four stages of flower development in order to understand molecular mechanisms during floral organ development. These sequences provide valuable information on the molecular mechanisms of floral development and flowering as the first major genomic resource for the genus Cymbidium[48]. In Cymbidium faberi, 12 TCP transcription factors and 34 MADS-box genes were selected for transcriptome analysis by using vegetative and flower buds[49]. In addition to genus Cymbidium, there is also a report on floral patterning research in genus Dendrobium. In Dendrobium officinale, joint analysis of transcriptome and metabolome found that MIKC-type MADS-box proteins and ARFs, endogenous hormones IAA and ABA are potentially involved in the development of flower organs[50]. Current research on the formation mechanisms of specific floral organs of orchid plants are concentrated in the genus Cymbidium. In Cymbidium goeringii, the transcriptome combined with microRNA analysis was used to screen out two transcription factor/microRNA-based genetic pathways that may be involved in the formation of the multi-petal trait[51], and by using proteome profiling, three transcription factors bHLH13, WRKY33 and VIP1 were identified as candidate regulators related to specific floral organ development[52]. Analysis of the Cymbidium sinense transcriptome shows that interactions between MADS factors play a crucial role in orchid floral zygomorphy and that mutations in these factors may be maintained during artificial selection[53].

    • To regulate the orchid flowering time, it is essential to grasp their flowering traits. Chinese scholars have investigated flowering physiology of Cymbidium, Dendrobium, Oncidium, and Phalaenopsis, identifying the molecular controls of flowering in these orchids. Based on multi omics data analysis and gene screening, flowering promoter such as SOC1, CO/COL, AGL24, AP1, FTIP1 from Dendrobium, MADS2, MADS4, SEP3, OAGL6-1, AP1, MADS11 of Oncidium and FD, MADS7, SEP3 from Phalaenopsis have been identified, heterologous overexpression in model plants have resulted in early-flowering, while overexpression of SVP-like genes from Cymbidium and Phalaenopsis, TFL1-like genes in Dendrobium and Oncidium, COL genes of Cymbidium and Doritaenopsis DhEFL2,3,4 in different heterologous systems resulting in delayed-flowering. Those results indicate that most homologs of the flowering-relevant genes play evolutionarily conserved roles in orchid compared with model plants such as Arabidopsis and Nicotiana benthamiana.

      There also exist many species-specific regulatory patterns. For example, a group of tandem repeats of DAM (Dormancy Associated MADS-box) transcription factors, has been identified in the genomes of woody trees as a potential marker of dormancy[54]. However, the whole genome sequencing of orchids does not find any DAM orthologs, suggesting that orchids work independently of the networks regulated by SVP/StMADS11 in other perennial species, such as kiwifruit, Populus trichocarpa, and Rosaceae species[55,56,57].

    • The color of orchids is mainly composed of three pigments: chlorophyll, flavonoids and carotenoids. The expression of structural genes of the pigment synthesis pathway can affect the synthesis and accumulation of pigment, thereby affecting the color phenotype of orchid plant tissues[58]. The structural genes for anthocyanin synthesis that have been cloned in orchids include CHS, CHI, F3H, F3'5'H, F3'H, DFR, UFGT and ANS[59]. Several structural genes involved in carotenoid synthesis have been cloned in orchids, including PSY, PDS, ISO, NCED, HYB, ZEP, CCD1, and CRTISO[59]. There are some reports on structural genes related to chlorophyll synthesis in genus Cymbidium, involving all known structural genes in the process of chlorophyll synthesis and metabolism. There are some reports on structural genes related to chlorophyll synthesis in orchids, among which PORB and CLH have been cloned[60,61]. Transcription factors affect orchid color changes by regulating the expression of structural genes. Among the transcription factors involved in regulating the anthocyanin biosynthesis pathway, the most studied ones are MYB[62], bHLH[63], WRKY, MADS and ZIP[59]. These transcription factors affect the presentation of orchid flower color by activating or inhibiting the expression levels of structural genes[59]. Compared with structural genes, there are fewer studies on transcription factors in the carotenoid synthesis pathway. Currently, only the R2R3-MYB transcription factor gene RcRCP1 that regulates CBP structural genes has been found in Cattleya[62]. For the chlorophyll metabolite pathway, CsERF2 transcript factor triggers structural changes in chloroplasts by regulating sugar signals, thereby promoting the degradation of chlorophyll[63]. Multi-omics analysis methods can also be used to analyze the formation mechanism of color in orchids. In terms of the mechanism of orchid flower color formation, most of the research ideas adopted are to select petals of different flower color varieties and use transcriptome combined with metabolome to explore key metabolic pathways. The important role of CeMYB104 in the regulation of flower color in Cymbidium ensifolium was identified by joint analysis of transcriptome and metabolome[64]. In Pleione limprichtii, transcriptomic methods were used to analyze the coloring mechanisms of petals of three different colors and found that PlFLS plays a decisive role in whether the petals appear white, PlANS and PlUFGT are related to the accumulation of anthocyanins. And an important candidate transcription factor PlMYB10 was predicted to form an MBW protein complex (MYB, bHLH, and WDR), regulate PlFLS expression[65].

    • Currently, research on molecular regulation of orchid fragrance is mainly focused on the terpene metabolism pathway. In Oncidium, five key TPS genes were screened based on transcriptome data, and the key TPS genes are different in varieties with different aromas[66,67]. Through expression level analysis combined with exogenous application of methyl jasmonate, DoTPS10 and DoGES are involved in the synthesis of linalool and geraniol respectively in Dendrobium officinale[68, 69]. PbGDPS may play an important regulatory role in the synthesis of monoterpenes in Phalaenopsis[70]. In Cymbidium sinense, a key enzyme SjHMGR gene in the MVA pathway of terpenoid biosynthesis, may be closely related to the formation of floral fragrance[71]. In addition, a comparison of transcriptome information from different Phalaenopsis fragrant types reveals that PbbHLH4 is crucial in controlling the biosynthesis of floral monoterpenes in orchids[72]. In Cymbidium, CsMYB1 is involved in regulating the synthesis of phenylpropanoid/benzene compounds in the Cymbidium cultivar 'Seal Bit'[73]. Methyl jasmonate is one of the main aroma components of Cymbidium faberi. The promoters of the CfAOC gene and CfJMT gene are related to the synthesis of methyl jasmonate which is activated by heterologously overexpression of four CfMYB genes, and the content of methyl jasmonate is increased, indicating that CfMYB can regulate the synthesis of methyl jasmonate[74].

    • Resistance is also a very important issue in orchid growth and development. Both abiotic and biotic stresses can have significant adverse effects on the growth and yield of orchids or even cause death. Chinese researchers have extensively studied orchid stress resistance using multi-omics joint analysis, gene function identification, and gene family systematic analysis. Orchids native to the tropics are sensitive to low temperatures, especially during the reproductive growth stage. In the typical tropical orchid Phalaenopsis, researchers not only analyzed the expression patterns of the classic cold-resistant transcription factors CBF and ICE1[75], but also used transcriptome, sRNA combined with degradome to screen out the key cold-resistant genes digalactosyldiacylglycerol synthase 2 (DGD2) and its specific natural antisense transcripts (NATs)[76].

      When considering drought resistance, the integrated data from the transcriptome and metabolome in Dendrobium suggest multiple pathways related to purine metabolism and phenylpropanoid biosynthesis[77]. These pathways along with networks connected to nucleotide processes and stress responses[78], are pivotal to drought adaptation by Dendrobium orchids. Terrestrial and epiphytic orchids employ different drought survival strategies. Proteome analyses indicate that epiphytic orchids are more drought-resistant than their terrestrial counterparts. Epiphytic orchids are superior in sustaining carbon equilibrium and responding to ABA[79].

      Cymbidium mosaic virus (CymMV) and Odotoglossum ringspot virus (ORSV) are global threats to orchid economic stability, intensifying the focus on disease resistance within the industry. Studies in China show that genes such as NPR1[80,81] and plant A20/AN1 proteins[82] boost orchid defenses against CymMV and ORSV. Investigations also suggest that custom-engineered microRNAs (miRNAs) enhance orchid disease specificity[83].

      Beyond identifying key resistance genes in various stress scenarios, exploring gene families comprehensively is critical for understanding the role of stress-resistant gene clusters and future applications. To date, published gene family analyses in orchids encompass families like AGO[84], GRAS[85], LEA[86], prenylsynthase-terpene synthase[87], R2R3-MYB transcription factors[88], SOD[89], and WRKY[90]. These studies typically involve gene family structure analysis, stress-conditioned gene expression analysis, and essential gene functionality identification.

    • The Royal Horticultural Society (RHS) established a global orchid hybrid registry in 1854, with Calanthe Dominyi (Calanthe furcata × Calanthe masuca) being the inaugural registration[91]. With the development of aseptic seeding technology of orchids, the number of new hybrids of orchids has shown an explosive growth, and there are currently more than 170,000 registered orchid hybrids. Among them, we have registered 109 novel hybrids.

      In the agriculture and forestry sectors of China, protection for new orchid varieties has expanded to nine genera, including Paphiopedilum, Pleione, Phalaenopsis, Cattleya, Cymbidium, Cypripedium, Dendrobium, Gastrodia and Vanda (Table 7). As of October 6, 2023, national authorization for new plant variety rights has been granted to 536 orchid varieties. Notably, varieties of Phalaenopsis represent 79.48% of these (sources: www.nybkjfzzx.cn/p_pzbh/sub_gg.aspx?n=21, http://lygc.lknet.ac.cn/s/sqpzsjk.html). Regarding medicinal orchids, of the 30 Dendrobium newcomers, nearly 50% are Dendrobium officinale, while new variety authorizations for medicinal Gastrodia remain absent. Within Cymbidium, selection breeding from 1993 to 2020 resulted in 621 natural variants (data from Orchid Branch of China Flower Association, not yet published).

      Table 7.  Number of orchids granted new plant variety rights in China.

      GenusNumber of varietiesPercentage (%)
      Paphiopedilum00
      Pleione00
      Phalaenopsis42679.48
      Cattleya20.37
      Cymbidium7814.55
      Dendrobium305.60
      Cypripedium00
      Gastrodia00
      Vanda00
      Total536100
    • Polyploid orchids typically exhibit enhanced organ dimensions and augmented resilience to stress, and polyploidization is useful for restoring the fertility of some excellent germplasm resources with low or no fertility[92]. Researchers have identified endopolyploids (spontaneous polyploid occurrences) within diverse cells, tissues, and organs across multiple genera, including Phalaenopsis, Cymbidium, and Cattleya[92]. As orchid tissue culture techniques have reached to advanced levels, there is a growing trend in studies focusing on polyploid stimulation in conjunction with these methods. Over the last two decades in China, orchid polyploid breeding was extended to 13 different genera, with Phalaenopsis, Dendrobium, and Cymbidium receiving more attention (Table 8). The primary approach to inducing polyploidy in orchids involves the use of chemicals such as colchicine, oryzalin, and pendimethalin on various orchid parts, including seeds, protocorms, protocorm-like bodies, cluster buds, rhizomes, and stem explants. By using these chemicals, the rates of polyploid induction ranges from 6.32% to 72.7%. A study shows that the formation rates of 2n male gametes (unreduced gametes) fluctuate between 0.15% and 4.03%, and by utilizing 2n gametes for sexual crossbreeding also presents a potent strategy for procuring polyploid orchids[93].

      Table 8.  Polyploidy induction of orchids.

      GeneraSpeciesType of explantOptimal treatmentAssessment methodInduction rateReferences
      CymbidiumCymbidium hydridumPLBsColchicine 0.1% for 3 dA, C27.6%[94]
      Cymbidium hybridium 'Sunrise'PLBsColchicine 0.05% for 5 dA, D23.7%[95]
      Cymbidium hybridium 'Hongpubu'Cluster budsColchicine 0.05% for 24 hA, B, D28.2%[95]
      Cymbidium sinenseRhizomesColchicine 0.01% for 3 dA, B11.1%[96]
      Five Cymbidium sinense cultivars and four hybrids2n gametesInterspecific hybridization to produce sexual polyploidsA, B, Cseven pairs of crosses produced five triploid and two tetraploid hybrids[93]
      Cymbidium Ruby Shower 'Murasakin Okimi'ProtocormsColchicine 300 mg/L for 15 dA, B30.0%[95]
      Cymbidium Golden Elf 'Sundust'RhizomesOryzalin 0. 002% for 48 h, and then EMS 50 mg /L for 1 monthA, B, D40%[95]
      Cymbidium sinense 'Lvmosu' × Cymbidium hybridum 'Shijieheping' F1 generationProtocormsColchicine 0.03% for 72 hA, C, D36%[95]
      Cymbidium lowianumCluster budsColchicine 0.04% for 72 hA, B, D60%[95]
      Interspecific hybridsRhizomesColchicine 0.1% for 48 hA, B36%[95]
      Cultivar 'Suxinhuang'ProtocormsColchicine 0.005% for 3 dA, B16.7%[95]
      Cymbidium faberiCluster budsColchicine 0.5% for 48 hA, B, D13.3%[95]
      Cymbidium iridioidesCluster budsColchicine 0.05% for 72 hA, B, D74%[97]
      DendrobiumDendrobium officinaleProtocormsOryzalin 14.4 μM for 24 hA, B, C37.4%[98]
      Dendrobium officinalePLBs and cluster budsColchicine 0.09% for 24 hA, B, D48%[95]
      Dendrobium officinaleProtocormsColchicine 0.1% for 15 dA, B, C57.69%[95]
      Dendrobium officinaleSeeds and PLBSColchicine 50 mg/LA, B, C50%[95]
      Dendrobium officinaleProtocormsColchicine 0.6 mg/L for 30 minA, B, C, D16.7%
      [95]
      Dendrobium officinaleProtocormsColchicine 2.0 mg/L for 36 hA, B, C, D20%[95]
      Dendrobium cariniferumProtocormsColchicine 0.05% for 24 hA,B, C, D33.0%[99]
      Dendrobium devonianumCluster budsColchicine 0.03% for 24 hA, B60.0%[95]
      Hybrids of Dendrobium snow flake 'Red star' and Dendrobium white rabbit 'Sakurahine'Cluster budsColchicine 0.06% for 12 hA, B, D69.1%[95]
      Dendrobium wardianumProtocormsColchicine 0.1% for 12 hA, B, D26%[95]
      Dendrobium sineseProtocormsOryzalin 20 mg/L for 4 dA, B, C, D35%[95]
      Dendrobium hybrida 'Sonia'ProtocormsColchicine 0.01% + Oryzalin
      5 mg/L for 8−10d
      A, BAbove 90%[95]
      Dendrobium ochreatumProtocormsColchicine 0.05%−0.1%
      for 2−3 d
      A, B, C, D[95]
      Hybrids of Dendrobium utopia 'Messenger' × Dendrobium Whiterabbit 'Sakurahime'Tube seedlingsColchicine 0.6 mg/L for 24 hC, D62.2%[95]
      PhalaenopsisH-03SeedsColchicine 0.05% for 15dA, B, C, D50.0%[100]
      Phalaenopsis Tailin 'Red Angle'leaves and adventitious budsColchicine 0.01% for 30 dA, B, D33.3%[100]
      Phalaenopsis 'TsueiFoa Lady'Protocormsand PLBsColchicine 0.1% for 7 dA, B, D30%[100]
      Phalaenopsis amabilis
      Phalaenopsis aphrodite
      ProtocormsEndopolyploid. in vitro regenerationA, CA large number of stable polyploid plants were obtained in a short time[100]
      Phalaenopsis aphroditeProtocormsEndopolyploid, Horizontal cuttingPolyploid plants can be produced in large numbers[100]
      Phalaenopsis zhejiangensisSeedsColchicine 0.2% for 1 dA, C, D27.75%[100]
      HybridsCluster budsColchicine 0.05% for 24 hA, B, D20%[100]
      IonopsisIonopsis utricularioidesLeaves and embryoidsColchicine 200 mg/L for 24 hA, C, D8.53%[100]
      ArundinaArundina graminifoliProtocorms0.1% colchicine for 12 hA, B, C, D23.33%[100]
      CremastraCremastra appendiculataProtocormsColchicine 0.05% for 1 dA, B20%[101]
      PhaiusPhaius tankervilleaeProtocormsColchicine 0.02% for 6 dA, B22.5%[100]
      PleionePleione maculaataProtocormsColchicine 0.2% for 60 hA, B25.64%[100]
      NerviliaNervilia fordiiRhizomesColchicine 300 mg/L + DMSO
      10 ml/L + 2.0 mg/LKT for 28 d
      A, B, D50%[102]
      SpathoglottisSpathoglottis plicataSeeds and transverse thin cell layerColchicine 0.3% for 30 dA, B, D27.8%[100]
      OncidiumOncidium flexuosumPLBsColchicine 1,500 mg/L for 54 hA, B, C, D26.67 %[100]
      BletillaBletilla striataPLBsColchicine 0.2% for 36 hA, B, C, D26.7%[100]
      AnoectochilusAnoectochilus roxburghiiStemsPendimethalin 90 mg/L for 48 hA, C50.0%[100]
      Anoectochilus roxburghiiStemsColchicine 0.1% for 48 hA, B, D48%[100]
      Anoectochilus roxburghiiStemsPendimethalin 200 μmol/L
      for 8 d
      A, B, D44.17%[100]
      Anoectochilus roxburghiiStemsColchicine 700 mg/L for 15 hA, B, D53%[100]
      Anoectochilus roxburghiiStemsColchicine 300 mg/L for 13 dA, B72.7%[100]
    • Mutation breeding is the use of various methods to induce plants to produce new traits. It has a unique role in the improvement of ornamental plant varieties, and it can induce acquisition of new genes or new germplasm. Compared with traditional hybridization, mutation breeding has the advantages of higher mutation frequency and rich mutation types within a short time. Mutation breeding in China began in the 1950s, and there are many reports on orchids at present. Except natural variation and artificially induced polyploids,60Co-γ physical radiation is the most reported method, and it has been used in several species of orchids such as Phalaenopsis spp., Epidendreae spp., Paphiopedilum delenatii, Paphiopedilum callosum, Cymbidium goeringii, Cymbidium faberi, Dendrobium crumenatum, Dendrobium Officinale and Dendrobium Sonia'166'[103,104]. The radiation dose ranged from 5 to 40 Gy. Heavy ion radiation and space mutation are also alternative methods to create mutants. The rhizoids of four Cymbidium hybridum were irradiated by 12C6+ heavy ions, and SSR molecular marker detection revealed new polymorphic bands of '17-33' at a dose of 20 Gy[104]. A new Phalaenopsis cutivar 'Hangdie No.2' was selected from space mutation, and it formed a series of excellent characteristics over the original varieties[105].

      In addition, mutants are also easily produced through callus or protocorm propagation during tissue culture. Most of the reported mutant characters are leaf color variation in Cymbidium and Dendrobium, and mutation with flower pattern, flower color and spots found in Phalaenopsis[106]. Phalaenopsis exhibited notable genetic variability, with the occurrence of somaclonal variants spanning from 5.6% to 47.9%[107]. Some red flower Phalaenopsis varieties with Phalaenopsis equestris or Phalaenopsis pulcherrima lineages, such as Phalaenopsis Formosa Rose, Phalaenopsis King Shiang's Beauty, Phalaenopsis King Shiang's Rose, are easy to produce three-lip flower type variation in tissue culture. Waxy flower varieties with spotted, yellow or red flowers are also prone to variation, but large white flower varieties are less susceptible to variation[91]. Therefore, somaclonal variation is one of the effective ways to breed new varieties of Phalaenopsis. In the process of tissue culture of Phalaenopsis Sogo Vivien, we also found some leaf color variation types, and selected a new cultivar 'Jingxiang' with large gold edges on the leaves.

    • Crossbreeding is a widely employed and highly effective method for breeding orchids. Achieving successful inter-species and even inter-genus hybridization is feasible, although orchid seeds pose unique challenges due to lack of endosperm and immature embryos. To overcome these obstacles, they rely on symbiotic fungi for germination in limited rates[91]. The success of orchid hybridization hinges on several crucial factors, including pollen viability, stigma acceptability and ploidy, making the selection of hybrid parents a pivotal task. In the case of Phalaenopsis, we have seen the creation of various intergeneric hybrid combinations, such as those between Phalaenopsis and other genera like Sedirea, Neofinetia, Renanthera, Rhynchonopsis and Vanda, aiming to yield new varieties with delightful fragrances or vibrant hues. The success rate of these crosses ranges from 2.78% to 12.50%[108]. Overcoming incompatibility between genera is possible through techniques like expanding the range of parental hybridization, repeated pollination, cross-pollination, and embryo rescue[108].

      In the realm of Cymbidium orchids, the breeding of Cymbidium hybridium stands out as a prominent achievement in domestic orchid breeding. After years of interspecific hybridization, a diverse array of varieties has emerged. Enhancements in flower color, flower type, fragrance, flowering time, leaf color, and plant shape have been achieved by crossing Cymbidium hybridium with native species of florets in Cymbidium, such as Cymbidium goeringii, Cymbidium kanran, Cymbidium sinense and other indigenous Chinese species[109]. Notably, hybrid combinations with Cymbidium hybridium as the female parent have exhibited very low seed setting rates or no seed at all, while the same hybrids with Cymbidium hybridium as the male parent have displayed varying degrees of seed setting[109]. Meanwhile, the fruit setting rate is as high as 100.0% when the Cattleya hybrid 'KTL4' is the female parent. However, it is only 10.0%. when it is a male parent[110]. These findings underscore that the compatibility of the same orchid species differs when it serves as a male or female parent.

    • Molecular marker-assisted breeding serves as an effective means to expedite plant breeding efforts. Several molecular markers, including SRAP, AFLP, ISSR, SSR, and EST-SSR, are commonly employed in orchid breeding and have found widespread application in plant identification, fingerprinting, core germplasm establishment, genetic diversity assessment, and phylogenetic analysis[111]. This approach is particularly valuable when working with domestic medicinal and endangered wild orchid resources, such as Dendrobium loddigesii[112], Liparis japonica[113] and Cymbidium tortisepalum[114]. Notably, SSR markers were used to identify the genes associated with flower color, shape and resistance of Phalaenopsis, which provided valuable insights for genetic engineering breeding of Phalaenopsis[115]. In a separate study, eight pairs of SSR markers were screened to distinguish accurately the pure yellow-green flowers and varietal flowers in hybrid offspring of Cymbidium hubridum 'Xiao Feng' and Cymbidium sinense 'Wu Zi Cui'[116]. Additionally, specific-locus amplified fragment sequencing and bulked segregant analysis led to the identification of two single-nucleotide polymorphism (SNP) markers linked to flower ground color in Phalaenopsis, achieving an impressive accuracy rate of 93.3%[117].

      Furthermore, the advent of high-throughput sequencing has facilitated the development of numerous SNP markers through simplified genome sequencing and genotyping by sequencing. The construction of high-density genetic linkage maps using polymorphic SNP markers has emerged as a pivotal method for identifying functional genes or markers closely linked to specific traits[12]. Presently, there are 14 reports on genetic maps for orchids, with a primary focus on four genera: Paphiopedilum, Phalaenopsis, Dendrobium and Cymbidium (Table 9). These maps have unveiled QTLs (or eQTLs) and candidate genes associated with flower color, polysaccharide content, as well as traits of leaves and stems. Nevertheless, further refinement and verification of these findings remain necessary.

      Table 9.  Genetic map of orchids.

      GenusMapping
      population
      Population
      type
      Population
      size
      Marker
      type
      Total distance
      of map (cM)
      Average
      distance
      (cM)
      Number of markersQTL or eQTLsReferences
      PaphiopedilumPaphiopedilum concolor × Paphiopedilum hirsutissimumF195SNP1,616.180.198,41012 QTLs linked to leaf length, leaf width, leaf thickness, and leaf number[13]
      PhalaenopsisPhalaenopsis '462' × Phalaenopsis '20'F188AFLPPhal '462': 878.3
      Phal '20': 820.3
      5.0
      6.7
      175
      122
      [118]
      Phalaenopsis aphrodite × Phalaenopsis equestrisF1117SNPl5l92.050.l3113,51710 QTLs associated with four color related traits[119]
      DendrobiumDendrobium Lucky Gal × Dendrobium FantasyF190RAPD6,568.750.11121[118]
      Dendrobium Second Love × Dendrobium Sekand RaveF192SSRDen. Second Love: 571
      Den. Sekand Rave: 566.3
      4.6
      8.5
      124
      67
      [118]
      Dendrobium officinale × Dendrobium hercoglossumF190RAPD
      SRAP
      Den. officinale: 629.4
      Den. ercoglossum: 1,304.6
      11.2
      11.6
      62
      112
      [118]
      Dendrobium officinale × Dendrobium moniliformeF190EST-SSR, SRAP, ISSR, RAPDD. moniliforme: 1,332.6
      D.officinale: 1,425.9
      10.41
      10.41
      226
      220
      [118]
      Dendrobium officinale × Dendrobium aduncumF1140SRAP
      SSR
      1,580.411.89157[118]
      Dendrobium nobile × Dendrobium moniliformF190RAPD
      ISSR
      D. nobile: 1,474
      D. moniliform: 1,326.5
      14.75
      14.88
      116
      117
      [118]
      Dendrobium moniliforme × Dendrobium officinaleF1111SNP2,737.490.328,5735QTL for Polysaccharide content[120]
      Dendrobium nobile × Dendrobium wardianumF1100SNP3,612.120.419,6452 eQTL for stem length and 1 eQTL for stem diameter[12]
      Dendrobium mangosteen × Dendrobium Burana Pink No.2F1190SSR, SRAP, RSAP, ISSR1,4219.56274[118]
      Dendrobium mangosteen × Dendrobium Burana Pink No.2F1190SSR, SRAP, ISSR1,548.99.91230[118]
      CymbidiumCymbidium hybridum 'Yunv' × Cymbidium Sinense 'Huangyehongmo'F194SSR1,608.932.1556[118]
    • Transgenic breeding generates desired traits by transferring exogenous genes or changing the expression characteristics of endogenous genes, which is precise, targeted, and time saving compared to traditional breeding methods[121]. As early as 1997, Hsieh reported the method of transforming Phalaenopsis equestris protocorms mediated by Agrobacterium, successfully introducing the GUS (β-glucuronidase) reporter gene into Phalaenopsis equestris[122]. After that, other reporter genes and functional genes were successively reported in Dendrobium[123], Oncidium[124], Cymbidium[125], Cattleya[126], Doritis[127], and Paphiopedilum[128]. Moreover, there are also genetic transformation studies of interspecific hybrids and intergeneric hybrids. For example, CLF gene was introduced into Cymbidium × Dendrobium[129], and the GUS reporter gene was injected into Arachnis × Vanda[130] and Dendrobium × Phalaenopsis[131].

      To date, more than 70 reports (Table 10) have been presented on the genetic transformation system and transgenic breeding of orchids in China. The molecular breeding for orchids mainly involves the color, flowering period, flower development, disease resistance and cold resistance related research. The explants include protocorms, quasi-protocorms, callus and pollen tube channels. The key transformation techniques included Agrobacterium-mediated transformation and the use of gene guns.

      Table 10.  Transgenic research on orchids.

      SpeciesExplantsGenetic transformationReport geneReferences
      PhalaenopsisPLB (PLBs)AgrobacteriumACS[132]
      PLB (PLBs)AgrobacteriumLFY[132]
      PLB (PLBs)AgrobacteriumGFP[132]
      PLB (PLBs)AgrobacteriumVwF3'5'H, GUS[133]
      PLB (PLBs)AgrobacteriumGFP, Hpt[132]
      PLB (PLBs)AgrobacteriumCymMV-CP[132]
      PLB (PLBs)AgrobacteriumICE1[133]
      PLB (PLBs)AgrobacteriumCAMV, GUS, eGFP[133]
      PLB (PLBs)AgrobacteriumGUS[133]
      PLB (PLBs)AgrobacteriumYUCCA6, GUS, Hpt[133]
      CallusAgrobacteriumLTP[133]
      CallusAgrobacteriumGUS, Hpt[133]
      Pollen tubeAgrobacteriumCbf1[133]
      Immature embryoAgrobacteriumLycB[133]
      LeafAgrobacteriumGAFP-NP1[133]
      OvaryAgrobacteriumGUS, Hpt[133]
      PLB (PLBs)Particle gunGUS[132]
      PLB (PLBs)Particle gunHpt, GUS[133]
      PLB (PLBs)Particle gunPeUFGT3[133]
      DendrobiumPLB (PLBs)AgrobacteriumDOH1[132]
      PLB (PLBs)AgrobacteriumRTACO[133]
      PLB (PLBs)AgrobacteriumACS[132]
      PLB (PLBs)AgrobacteriumaiiA-hacD[133]
      PLB (PLBs)AgrobacteriumGUS[133]
      PLB (PLBs)AgrobacteriumGUS, Hpt[132]
      PLB (PLBs)AgrobacteriumPR1, PR10[133]
      PLB (PLBs)AgrobacteriumCHS, F3'5'H[133]
      PLB (PLBs)AgrobacteriumCyMV-CP,
      ORSV-CP
      [132]
      PLB (PLBs)AgrobacteriumRTACS[133]
      PLB (PLBs)AgrobacteriumNAC[133]
      PLB (PLBs)AgrobacteriumGUS[133]
      CallusAgrobacteriumDcOSEP1[133]
      CallusAgrobacteriumHpt[133]
      CallusAgrobacteriumAFP[133]
      OvaryAgrobacteriumGUS, Hpt[133]
      PLB (PLBs)Particle gunGUS, Hpt[133]
      PLB (PLBs)Particle gunCymMV CP[133]
      PLB (PLBs)Particle gunLFY[133]
      OncidiumPLB (PLBs)Agrobacteriumpflp[133]
      PLB (PLBs)AgrobacteriumACS[133]
      PLB (PLBs)Particle gunCBF3[133]
      PLB (PLBs)AgrobacteriumAtTIP5;1[133]
      PLB (PLBs)AgrobacteriumCyMV-CP ORSV-CP[133]
      CallusAgrobacteriumGAFP-NP1[133]
      CallusAgrobacteriumGUS[133]
      PLB (PLBs)AgrobacteriumGUS[134]
      CymbidiumPLB (PLBs)AgrobacteriumICE1[133]
      PLB (PLBs)AgrobacteriumGAFP[133]
      PLB (PLBs)AgrobacteriumCyMV, ORSV[133]
      PLB (PLBs)AgrobacteriumGUS[132]
      Somatic embryoAgrobacteriumGAFP-NP1[133]
      RhizomeAgrobacteriumGUS[133]
      PLB (PLBs)Particle gunCiDREB1, PeDREB2, CYMV-CP[133]
      DoritisOvaryAgrobacteriumGUS[127]
      CattleyaPLB (PLBs)AgrobacteriumGUS, ORSV[126]
      Cymbidium goeringii × Cymbidium hybridumPLB (PLBs)AgrobacteriumCLF[129]
      Archnis × VandaPLB (PLBs)Particle gunGUS[130]
      Dendrobium × PhalaenopsisPLB (PLBs)AgrobacteriumGUS[131]
    • In the field of molecular biology, significant strides have been made in the Clustered Regularly Interspaced Short Palindromic Repeats (CRISPR)/CRISPR-associated (Cas) system, representing a groundbreaking genome editing technique[135]. Recent years have witnessed remarkable breakthroughs in gene editing technologies based on Crispr/Cas, enabling targeted modifications of gene characteristics in plants. These alterations encompass gene knockout, reduction of gene expression, and the augmentation of gene expression by modifying promoters[136138].

      Nevertheless, the application of the CRISPR/Cas9 method for genetic editing of orchids in China has seen limited exploration. Only a handful of studies have reported successful gene editing in orchids, involving the modification of genes related to the lignin synthesis pathway (C3H, C4H, 4CL, CCR, and IRX) in Dendrobium[139] and the editing of MADS44, MADS36, and MADS8 genes in Phalaenopsis[140].

      The progress of genome editing in ornamental plants faces constraints, including the absence of specialized transformation systems, inefficiencies in existing transformation methods, and limited comprehension of certain trait-regulatory networks. Thus, it becomes imperative to develop suitable expression systems and target sequence configurations tailored to different ornamental plants. Additionally, adherence to legal regulations and the enhancement of target traits is essential during genome editing. Despite the challenges of low transformation and gene editing efficiencies in orchids, the burgeoning orchid market and increased research endeavors are expected to yield more gene editing studies in orchids and other ornamental plants, ultimately resulting in the establishment of efficient genome editing systems for orchids.

    • Orchid seeds present a unique challenge due to their incomplete embryo development, lacking both endosperm and cotyledons, which hinders natural germination. They predominantly rely on symbiotic fungi and are typically propagated through division, a process characterized by its sluggishness. Orchid rapid propagation initiated in the 1960s when Morel utilized stem-tip tissue from Hybrid Cymbidium, cultivating it on a cytokinin-containing medium. This led to the formation of protocorms, which subsequently differentiated into roots and leaves, resulting in the first virus-free orchid plantlets[141]. Today, tissue culture methods have enabled the propagation of over 60 genera and hundreds of orchid species.

      In China, after the 1970s, orchid reproduction and rapid propagation gained prominence, including Cymbidium ensifolium[142], Cymbidium sinense[143], Cymbidium goeringii[144], among others.

      Orchid reproduction and rapid propagation primarily involve aseptic seeding and shoot-tip induction pathways, although other explants such as leaves, stems, root tips, flower stalks, and axillary buds can also be induced. Wimber first successfully induced protocorms from leaves of Cymbidium faberi, demonstrating the feasibility of obtaining regenerated plants using leaf explants[145]. Churchill et al. used root tips as explants to achieve successful plant regeneration[146]. Furthermore, protocorm-like bodies were successfully induced from flower buds of Cymbidium sinense, Dendrobium by Zhang & Ou[143]. However, floral organs as explants generally yield low induction rates and pose a greater challenge.

      The tissue culture-based rapid propagation of orchids necessitates the use of intermediary propagules, primarily rhizomes or PLBs. Orchids like Cymbidium, Dendrobium, Oncidium, Cattleya, and Arundina graminifolia have PLBs as intermediary propagules, and they are considered easily propagated species. In contrast, Phalaenopsis, due to the propensity of PLBs to generate variations, is moderately challenging to propagate. Terrestrial orchids such as Cymbidium sinense[147], Cymbidium ensifolium[148], and Cymbidium goeringii[144] rely on rhizomes as intermediary propagules, which pose challenges for shoot regeneration, categorizing them as difficult-to-propagate species. Zeng et al. successfully established a rapid propagation system for Paphiopedilum hangianum and Paphiopedilum Maudiae through tissue culture, marking a pioneering achievement on a global scale[149].

    • Cultivation facilities have enabled the manual control of orchid flowering periods, a practice extensively employed in orchid varieties such as Phalaenopsis and Dendrobium. Orchid flowering regulation research primarily centers around temperature, light, nutrient levels, and hormone signaling.

      For many orchids, the induction of flower buds require low-temperature exposure and day/night temperature fluctuations. For instance, Phalaenopsis[150] requires low-temperature treatment to initiate flower bud formation. Temperatures below 15 °C reduce flowering rates, while those exceeding 28 °C hinder flowering[151]. In contrast, most Cymbidium orchids, including Cymbidium goeringii[152], form flower buds at high temperatures, but the subsequent flower bud differentiation requires lower temperatures. While Oncidium[153] do not strictly require low temperatures for flower formation, a certain degree of cold exposure enhances flower development.

      Light plays a crucial role in orchid flowering. While orchids are not particularly stringent about day-length requirements, adequate light is essential for nutritional accumulation, while insufficient light can lead to growth retardation and delays flowering. A 7-h photoperiod fosters nutrient growth in Phalaenopsis, a 5-h photoperiod delays flowering, and a 9-h photoperiod enhances Phalaenopsis flower quality[154]. Additional lighting stimulates reproductive growth in Cymbidium goeringii[155] , advancing their flowering time.

      Nitrogen, phosphorus, and potassium are vital elements for orchid growth and development. Elevated nitrogen levels promote vegetative growth, while increased phosphorus and potassium levels enhance reproductive growth. Treatment with 5 mmol/L KCL solution accelerates Cymbidium sinense flower bud emergence by 20 d[156]. Application of 500 mg/L nitrogen fertilizer and 250 mg/L potassium fertilizer increase the number of Cymbidium orchid leaf blades, while 500 mg/L nitrogen and 500 mg/L potassium fertilizer enhance the number of Cymbidium orchid flower stalks[157].

      Exogenous plant hormones are extensively employed to manipulate orchid flowering times. 6-BA stimulates flower bud differentiation, increasing the number of flower stalks and flowers in Phalaenopsis[158]. GA3 promotes flowering in Paphiopedilum callosum[159], serving as a substitute for the low-temperature requirement for flowering in certain Phalaenopsis varieties, enabling flowering at room temperature[160]. PP333 reduces the height of Dendrobium nobile[161], resulting in compact flower arrangements, improved resistance to collapse, and earlier flowering. Combining hormones can synergistically regulate flowering. For instance, combining 6-BA with GA3 can mitigate blooming deformities caused by high GA3 concentrations, promoting flowering in Cymbidium sinense[162] and Cymbidium ensifolium[163].

    • Numerous diseases and pests pose threats to orchids. Therefore, disease identification and technological control are pivotal in orchid care. Common orchid diseases encompass stem rot, white silk disease, brown spot disease, leaf blight, and anthracnose[164,165]. Fungal stem rot and bacterial soft rot, especially, present significant challenges in orchid production. Stem-rot strains of Cymbidium[166], Paphiopedilum[167], Phalaenopsis amabilis[168], and Cymbidium hybridum[169] have been identified to be affected by Fusarium oxysporum and Fusarium solani. Additionally, Fusarium fujikuroi is known to cause P. amabilis flower stem rot[168]. Bacterial soft rot, primarily affecting orchid leaves, can lead to rapid plant death in high-temperature conditions. Chrysanthemum bacteria (Erwinia chrysanthemi) have been identified as the causative agents of soft rot in Phalaenopsis in regions such as Zhejiang, Yili, and Taiwan[170]. Pathogenic microbes responsible for other diseases like anthracnose, white silk disease, blight, gray rot disease, include Colletotrichum gloeosporioides, Atheliarolfsii, Phytophthora parasitica, Phytophthora palmivora, Phytophthora cactorum, and Sclerotina fucheliana. Some biocontrol effects on Fusarium oxysporum have been observed with Trichoderma viride, Trichoderma harzianum, and Trichoderma pseudokoningii[171].

      Orchid viral diseases pose challenges in orchid care, with up to 30 viruses known to infect orchids, including Cymbidium mosaic virus (CymMV), cucumber mosaic virus (CMV), tobacco mosaic virus (TMV), Odontoglossum ring spot virus (ORSV)[172], impatiens necrotic spot virus (INSV)[173], Phalaenopsis chlorosis spot virus (PhCSV)[174], Capsicum chlorosis virus (CaCV), Basella rugose mosaic virus (BaRMV), Rhabdovirus, Potyvirus, and Tospoviruses. Stem tip culture to breed virus-free seedlings is a primary method for controlling viral diseases.

    • Endophytic fungi are vital components of the Orchidaceae plant micro-ecosystem, playing crucial roles in orchid propagation. Orchids require a stable symbiotic relationship with fungi for seedling development, as these fungi provide essential nutrients through mycorrhizal associations, facilitating germination and healthy plant growth. Although China began researching endophytic fungi in Orchidaceae relatively late, approximately 50 orchid species, including Dendrobium[175], Bletilla[176], Cymbidium goeringii[177], and Anoectochilus roxburghii[178], have had their endophytic fungi isolated and identified. These fungi play critical roles in germination, growth, and disease resistance in orchids. For instance, symbiosis with three endophytic fungi significantly increased stem and leaf dry weights in the aseptic seedlings of Cymbidium sp. compared to mineral nutrient enhancement, improving growth by 173.2% to 250.1% and promoting nutrient absorption[179]. Some endophytic fungi can produce active ingredients found in medicinal plants and stimulate their host plants to produce these ingredients, significantly enhancing growth and polysaccharide synthesis in Anoectochilus roxburghii (Wall.) Lindl.[180]. Mycorrhizal fungi isolated from wild Cymbidium roots have been shown to significantly enhance respiration rates, cytochrome C oxidase, and peroxidase activity in Cymbidium sinense and Cymbidium ensifolium, positively affecting plant growth, development, and resistance[181] .

    • Orchids exhibit unique features, including diverse floral structures, vibrant colors, variable plant types, leaf shapes, leaf colors, complex growth habits (epiphytic, overground, and saprophytic), diverse propagation modes (rhizomes, protocorms, clustered buds), different photosynthetic pathways (C3 and CAM pathways), and various flowering characteristics (seasonal flowering, continuous flowering). Numerous challenges and opportunities within the realms of horticulture, biology, genetics, pollination, embryology, microbiology, and molecular biology concerning orchids. These include diverse pseudobulb types (monocotyledonous and compound), non-rigid genus (species) definitions, sexual hybridization polyploidy, underdeveloped embryos (lack of endosperm), dependence on symbiotic bacteria, lengthy juvenile periods, endangered species, medicinal components, and mechanisms of action. Substantial advancements are expected in molecular mechanism analysis and the identification of functional genes related to critical traits.

      Orchids are high-value flowers, particularly the commercial orchid varieties such as Phalaenopsis, Cymbidium, Dendrobium, Cattleya, Paphiopedilum, Oncidium and Vanda. Breeding objectives and perspectives for these orchids are diversifying, with a focus on delicate fragrance, enhanced leaf and flower characteristics, and energy-efficient and emission-reducing attributes in response to carbon emissions and carbon neutrality requirements. Advancements in breeding technology, including molecular marker-assisted breeding, transgenic techniques, gene editing, and molecular design breeding, are poised to usher orchid breeding from systematic breeding (breeding 1.0) and hybrid breeding (breeding 2.0) to molecular breeding (breeding 3.0) and intelligent design breeding (breeding 4.0). With the progression of digital and information-based agricultural technologies, orchid research on resources, breeding, physiology, cultivation, facilities, equipment, and agricultural economics is expected to become more systematic, further expanding the orchid industry.

      Beyond the established commercial orchids, species like Bulbophyllum (1,789 species), Eupatorium (1,125 species), Habenaria (848 species), Maxillaria (552 species), Masdevallia (507 species), Liparis (418 species), Eria (404,000 species), Calanthe (187 species), Coelogyne (182 species), Dracula (111 species), Prosthechea (93 species), Cypripedium (50 species), Lycaste (50 species), Phaius (48 species), Cycnoches (33 species), Tolumnia (31 species), Aerides (25 species), Pleione (20 species), Renanthera (17 species), Erycina (seven species), Rhynchostylis (three species), and Neofinetia (two species), which possess great ornamental or medicinal values, remain largely untapped. These orchids represent a rich source for the development of new commercial flowers, promising to diversify and enrich the future flower market.

      The Chinese have developed the national orchid and influenced Japan, South Korea and other East Asian countries, which have been increasingly accepted by Western countries in recent years. Phalaenopsis is mainly native to tropical and subtropical areas of Asia and Oceania. Currently, three production centers have emerged, including China, Europe and the United States. Cattleya native to tropical regions of central and south America, is mainly produced in Thailand and other Southeast Asian countries. China is going to become a major center of research, development, production and consumption of orchids in the future.

    • The authors confirm contribution to the paper as follows: original manuscript preparation: Yang F, Wong SM, Zhu G; data analysis: Gao J, Li J, Wei Y, Xie Q, Jin J, Lu C, Zhu W. All authors reviewed the results and approved the final version of the manuscript.

    • All data generated or analyzed during this study are included in this published article.

      • This research was funded by grants from 2022 Provincial Rural Revitalization Strategy Special Fund Seed Industry Revitalization Project (2022-NPY-00-039), the National Natural Science Foundation of China (31872151,31672184), Guangzhou Science and Technology Project (2022B03J00703); Innovation Team of Modern Agriculture Industry Technology System in Guangdong Province (2021KJ121); Guangdong Academy of Agricultural Sciences Discipline Team Construction Project (202127TD, R2020PY-JX018, BZ202006).

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

      • Copyright: © 2024 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 (1)  Table (10) References (181)
  • About this article
    Cite this article
    Yang F, Gao J, Li J, Wei Y, Xie Q, et al. 2024. The China orchid industry: past and future perspectives. Ornamental Plant Research 4: e002 doi: 10.48130/opr-0023-0024
    Yang F, Gao J, Li J, Wei Y, Xie Q, et al. 2024. The China orchid industry: past and future perspectives. Ornamental Plant Research 4: e002 doi: 10.48130/opr-0023-0024

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return