Search
2022 Volume 2
Article Contents
REVIEW   Open Access    

Hidden players in the regulation of secondary metabolism in tea plant: focus on non-coding RNAs

More Information
  • Non-coding RNAs (ncRNAs) are functional transcripts with minimal or no protein-coding capacity that comprise a large portion of the plant transcriptome. Among them, the microRNAs (miRNAs), linear long ncRNAs (lncRNAs), and circular long ncRNAs (circRNAs) have been widely proven to play essential regulatory roles in the biosynthesis of secondary metabolites (SMs) by modulating the expression of key synthesis-related genes in plants. Tea boasts numerous characteristic SMs, such as catechins, theanine, caffeine, volatile compounds, etc., which have distinguished health properties and largely determine the pleasant flavor quality. Thus, understanding how the tea plant produces these specialized metabolites is of great research interest. With the innovation and progress of biotechnologies in recent years, significant progress has been made in research on the regulation mechanism of SMs in tea plants at the DNA, mRNA, protein and metabolite levels. The release of the genome sequences of tea plants paves a path for precisely exploring ncRNAs and their functions in tea, and their huge potential for the biosynthesis regulation of SMs has gradually received attention. We herein summarize recent progress on miRNAs, lncRNAs, and circRNAs in tea plants and discuss their regulatory roles in the accumulation of SMs to enlighten the development of novel agronomic tools to enhance the quality of tea.
  • 加载中
  • [1]

    Jia X, Zhang W, Fernie AR, Wen W. 2021. Camellia sinensis (Tea). Trends in Genetics 37:201−2

    doi: 10.1016/j.tig.2020.10.002

    CrossRef   Google Scholar

    [2]

    Liao YY, Zhou XC, Zeng LT. 2021. How does tea (Camellia sinensis) produce specialized metabolites which determine its unique quality and function: a review. Critical Reviews in Food Science and Nutrition 6:3751−67

    doi: 10.1080/10408398.2020.1868970

    CrossRef   Google Scholar

    [3]

    Stamp N. 2003. Out of the quagmire of plant defense hypotheses. The Quarterly Review of Biology 78:23−55

    doi: 10.1086/367580

    CrossRef   Google Scholar

    [4]

    Yue Y, Chu G, Liu X, Tang X, Wang W, et al. 2014. TMDB: a literature-curated database for small molecular compounds found from tea. BMC Plant Biology 14:243

    doi: 10.1186/s12870-014-0243-1

    CrossRef   Google Scholar

    [5]

    Zhang L, Cao QQ, Granato D, Xu YQ, Ho CT. 2020. Association between chemistry and taste of tea: A review. Trends in Food Science & Technology 101:139−49

    doi: 10.1016/j.jpgs.2020.05.015

    CrossRef   Google Scholar

    [6]

    Zeng L, Watanabe N, Yang Z. 2019. Understanding the biosyntheses and stress response mechanisms of aroma compounds in tea (Camellia sinensis) to safely and effectively improve tea aroma. Critical reviews in food science and nutrition 59:2321−34

    doi: 10.1080/10408398.2018.1506907

    CrossRef   Google Scholar

    [7]

    Wan X, Xia T. 2015. Secondary Metabolism of Tea Plant. Beijing, China: Science Press. pp. 39

    [8]

    Ye J, Ye Y, Yin J, Jin J, Liang Y, et al. 2022. Bitterness and astringency of tea leaves and products: Formation mechanism and reducing strategies. Trends in Food Science & Technology 123:130−43

    doi: 10.1016/j.jpgs.2022.02.031

    CrossRef   Google Scholar

    [9]

    Yao S, Liu Y, Zhuang J, Zhao Y, Dai X, et al. 2022. Insights into acylation mechanisms: co-expression of serine carboxypeptidase-like acyltransferases and their non-catalytic companion paralogs. Plant Journal 111:117−33

    doi: 10.1111/tpj.15782

    CrossRef   Google Scholar

    [10]

    Zhao J, Li P, Xia T, Wan X. 2020. Exploring plant metabolic genomics: chemical diversity, metabolic complexity in the biosynthesis and transport of specialized metabolites with the tea plant as a model. Critical Reviews in Biotechnology 40:667−88

    doi: 10.1080/07388551.2020.1752617

    CrossRef   Google Scholar

    [11]

    Zhu B, Guo J, Dong C, Li F, Qiao S, et al. 2021. CsAlaDC and CsTSI work coordinately to determine theanine biosynthesis in tea plants (Camellia sinensis L.) and confer high levels of L-theanine accumulation in a non-tea plant. Plant Biotechnology Journal 19:2395−97

    doi: 10.1111/pbi.13722

    CrossRef   Google Scholar

    [12]

    Yang Z, Baldermann S, Watanabe N. 2013. Recent studies of the volatile compounds in tea. Food Research International 53:585−99

    doi: 10.1016/j.foodres.2013.02.011

    CrossRef   Google Scholar

    [13]

    Chen D, Sun Z, Gao JJ, Peng JK, Wang Z, et al. 2022. Metabolomics combined with proteomics provides a novel interpretation of the compound differences among Chinese tea cultivars (Camellia sinensis var. sinensis) with different manufacturing suitabilities. Food Chemistry 377:131976

    doi: 10.1016/j.foodchem.2021.131976

    CrossRef   Google Scholar

    [14]

    Gong A, Lian S, Wu N, Zhou Y, Zhao S, et al. 2020. Integrated transcriptomics and metabolomics analysis of catechins, caffeine and theanine biosynthesis in tea plant (Camellia sinensis) over the course of seasons. BMC Plant Biology 20:294

    doi: 10.1186/s12870-020-02443-y

    CrossRef   Google Scholar

    [15]

    Zeng L, Zhou X, Liao Y, Yang Z. 2021. Roles of specialized metabolites in biological function and environmental adaptability of tea plant (Camellia sinensis) as a metabolite studying model. Journal of Advanced Research 34:159−71

    doi: 10.1016/j.jare.2020.11.004

    CrossRef   Google Scholar

    [16]

    Chen Y, Wang M, Huang P, Tsao TM, Lin K. 2009. Influence of catechin on precipitation of aluminum hydroxide. Geoderma 152:296−300

    doi: 10.1016/j.geoderma.2009.06.017

    CrossRef   Google Scholar

    [17]

    Wang X, Zhu W, Cheng X, Lu Z, Liu X, et al. 2021. The effects of circadian rhythm on catechin accumulation in tea leaves. Beverage Plant Research 1:8

    doi: 10.48130/bpr-2021-0008

    CrossRef   Google Scholar

    [18]

    Yu X, He C, Li Y, Zhou J, Chen Y, et al. 2021. Effects of different spreading treatments on the formation of aroma quality in green tea. Beverage Plant Research 1:14

    doi: 10.48130/bpr-2021-0014

    CrossRef   Google Scholar

    [19]

    Guo YQ, Chang XJ, Zhu C, Zhang ST, Li XZ, et al. 2019. De novo transcriptome combined with spectrophotometry and gas chromatography-mass spectrometer (GC-MS) reveals differentially expressed genes during accumulation of secondary metabolites in purple-leaf tea (Camellia sinensis cv Hongyafoshou). The Journal of Horticultural Science and Biotechnology 94:349−67

    doi: 10.1080/14620316.2018.1521708

    CrossRef   Google Scholar

    [20]

    Guo Y, Zhu C, Zhao S, Zhang S, Wang W, et al. 2019. De novo transcriptome and phytochemical analyses reveal differentially expressed genes and characteristic secondary metabolites in the original oolong tea (Camellia sinensis) cultivar 'Tieguanyin' compared with cultivar 'Benshan'. BMC Genomics 20:265

    doi: 10.1186/s12864-019-5643-z

    CrossRef   Google Scholar

    [21]

    Yang J, Gu D, Wu S, Zhou X, Chen J, et al. 2021. Feasible strategies for studying the involvement of DNA methylation and histone acetylation in the stress-induced formation of quality-related metabolites in tea (Camellia sinensis). Horticulture Research 8:253

    doi: 10.1038/s41438-021-00679-9

    CrossRef   Google Scholar

    [22]

    Xia E, Tong W, Wu Q, Wei S, Zhao J, et al. 2020. Tea plant genomics: achievements, challenges and perspectives. Horticulture Research 7:7

    doi: 10.1038/s41438-019-0225-4

    CrossRef   Google Scholar

    [23]

    Yu Y, Zhang Y, Chen X, Chen Y. 2019. Plant noncoding RNAs: hidden players in development and stress responses. Annual review of cell and developmental biology 35:407−31

    doi: 10.1146/annurev-cellbio-100818-125218

    CrossRef   Google Scholar

    [24]

    Ma X, Zhao F, Zhou B. 2022. The characters of non-coding RNAs and their biological roles in plant development and abiotic stress response. International Journal of Molecular Sciences 23:4124

    doi: 10.3390/ijms23084124

    CrossRef   Google Scholar

    [25]

    Axtell MJ, Meyers BC. 2018. Revisiting criteria for plant microRNA annotation in the era of big data. The Plant Cell 30:272−84

    doi: 10.1105/tpc.17.00851

    CrossRef   Google Scholar

    [26]

    Song XW, Li Y, Cao XF, Qi YJ. 2019. MicroRNAs and their regulatory roles in plant-environment interactions. Annual Review of Plant Biology 70:489−525

    doi: 10.1146/annurev-arplant-050718-100334

    CrossRef   Google Scholar

    [27]

    He M, Kong X, Jiang Y, Qu H, Zhu H. 2022. MicroRNAs: emerging regulators in horticultural crops. Trends in Plant Science 27:936−51

    doi: 10.1016/j.tplants.2022.03.011

    CrossRef   Google Scholar

    [28]

    Bhogireddy S, Mangrauthia SK, Kumar R, Pandey AK, Singh S, et al. 2021. Regulatory non-coding RNAs: a new frontier in regulation of plant biology. Functional & Integrative Genomics 21:313−30

    doi: 10.1007/s10142-021-00787-8

    CrossRef   Google Scholar

    [29]

    Xu S, Zhou L, Ponnusamy M, Zhang L, Dong Y, et al. 2018. A comprehensive review of circRNA: from purification and identification to disease marker potential. PeerJ 6:e5503

    doi: 10.7717/peerj.5503

    CrossRef   Google Scholar

    [30]

    Jiang S, Cui J, Li X. 2021. MicroRNA-mediated gene regulation of secondary metabolism in plants. Critical Reviews in Plant Sciences 40:459−78

    doi: 10.1080/07352689.2022.2031674

    CrossRef   Google Scholar

    [31]

    Gupta OP, Karkute SG, Banerjee S, Meena NL, Dahuja A. 2017. Contemporary Understanding of miRNA-Based Regulation of Secondary Metabolites Biosynthesis in Plants. Frontiers in Plant Science 8:374

    doi: 10.3389/fpls.2017.00374

    CrossRef   Google Scholar

    [32]

    Krishnatreya DB, Agarwala N, Gill SS, Bandyopadhyay T. 2021. Understanding the role of miRNAs for improvement of tea quality and stress tolerance. Journal of Biotechnology 328:34−46

    doi: 10.1016/j.jbiotec.2020.12.019

    CrossRef   Google Scholar

    [33]

    Samad AFA, Sajad M, Nazaruddin N, Fauzi IA, Murad AMA, et al. 2017. MicroRNA and transcription factor: key players in plant regulatory network. Frontiers in Plant Science 8:565

    doi: 10.3389/fpls.2017.00565

    CrossRef   Google Scholar

    [34]

    Li H, Lin QQ, Yan ML, Wang ML, Wang P, et al. 2021. Relationship between secondary metabolism and miRNA for important flavor compounds in different tissues of tea plant (Camellia sinensis) as revealed by genome-wide miRNA analysis. Journal of Agricultural and Food Chemisrty 69:2001−12

    doi: 10.1021/acs.jafc.0c07440

    CrossRef   Google Scholar

    [35]

    Salmena L, Poliseno L, Tay Y, Kats L, Pandolfi PP. 2011. A ceRNA hypothesis: the Rosetta Stone of a hidden RNA language? Cell 146:353−58

    doi: 10.1016/j.cell.2011.07.014

    CrossRef   Google Scholar

    [36]

    Das A, Mondal TK. 2010. Computational identification of conserved microRNAs and their targets in tea (Camellia sinensis). American Journal of Plant Sciences 01:77−86

    doi: 10.4236/ajps.2010.12010

    CrossRef   Google Scholar

    [37]

    Prabu GR, Mandal AKA. 2010. Computational Identification of miRNAs and Their Target Genes from Expressed Sequence Tags of Tea (Camellia sinensis). Genomics, Proteomics & Bioinformatics 8:113−21

    doi: 10.1016/S1672-0229(10)60012-5

    CrossRef   Google Scholar

    [38]

    Zhu Q, Luo Y. 2013. Identification of miRNAs and their targets in tea (Camellia sinensis). Journal of Zhejiang University SCIENCE B 14:916−23

    doi: 10.1631/jzus.B1300006

    CrossRef   Google Scholar

    [39]

    Mohanpuria P, Yadav SK. 2012. Characterization of novel small RNAs from tea (Camellia sinensis L.). Molecular Biology Reports 39:3977−86

    doi: 10.1007/s11033-011-1178-3

    CrossRef   Google Scholar

    [40]

    Zhang Y, Zhu X, Chen X, Song C, Zou Z, et al. 2014. Identification and characterization of cold-responsive microRNAs in tea plant (Camellia sinensis) and their targets using high-throughput sequencing and degradome analysis. BMC Plant Biology 14:271

    doi: 10.1186/s12870-014-0271-x

    CrossRef   Google Scholar

    [41]

    Zheng C, Zhao L, Wang Y, Shen J, Zhang Y, et al. 2015. Integrated RNA-seq and sRNA-seq analysis identifies chilling and freezing responsive key molecular players and pathways in tea plant (Camellia sinensis). PLoS ONE 10:e125031

    doi: 10.1371/journal.pone.0125031

    CrossRef   Google Scholar

    [42]

    Liu S, Xu Y, Ma J, Wang W, Chen W, et al. 2016. Small RNA and degradome profiling reveals important roles for microRNAs and their targets in tea plant response to drought stress. Physiologia Plantarum 158:435−51

    doi: 10.1111/ppl.12477

    CrossRef   Google Scholar

    [43]

    Guo Y, Zhao S, Zhu C, Chang X, Yue C, et al. 2017. Identification of drought-responsive miRNAs and physiological characterization of tea plant (Camellia sinensis L. ) under drought stress. BMC plant biology 17:211

    doi: 10.1186/s12870-017-1172-6

    CrossRef   Google Scholar

    [44]

    Sun P, Cheng C, Lin Y, Zhu Q, Lin J, et al. 2017. Combined small RNA and degradome sequencing reveals complex microRNA regulation of catechin biosynthesis in tea (Camellia sinensis). PLoS ONE 12:e171173

    doi: 10.1371/journal.pone.0171173

    CrossRef   Google Scholar

    [45]

    Jeyaraj A, Zhang X, Hou Y, Shangguan M, Gajjeraman P, et al. 2017. Genome-wide identification of conserved and novel microRNAs in one bud and two tender leaves of tea plant (Camellia sinensis) by small RNA sequencing, microarray-based hybridization and genome survey scaffold sequences. BMC Plant Biology 17:212

    doi: 10.1186/s12870-017-1169-1

    CrossRef   Google Scholar

    [46]

    Jeyaraj A, Liu S, Zhang X, Zhang R, Shangguan M, et al. 2017. Genome-wide identification of microRNAs responsive to Ectropis oblique feeding in tea plant (Camellia sinensis L.). Scientific Reports 7:13634

    doi: 10.1038/s41598-017-13692-7

    CrossRef   Google Scholar

    [47]

    Jeyaraj A, Wang X, Wang S, Liu S, Zhang R, et al. 2019. Identification of regulatory networks of microRNAs and their targets in response to Colletotrichum gloeosporioides in tea plant (Camellia sinensis L.). Frontiers in Plant Science 10:1096

    doi: 10.3389/fpls.2019.01096

    CrossRef   Google Scholar

    [48]

    Xia E, Zhang H, Sheng J, Li K, Zhang Q, et al. 2017. The tea tree genome provides insights into tea flavor and independent evolution of caffeine biosynthesis. Molecular Plant 10:866−77

    doi: 10.1016/j.molp.2017.04.002

    CrossRef   Google Scholar

    [49]

    Wei C, Yang H, Wang S, Zhao J, Liu C, et al. 2018. Draft genome sequence of Camellia sinensis var. sinensis provides insights into the evolution of the tea genome and tea quality. PNAS 115:E4151−E4158

    doi: 10.1073/pnas.1719622115

    CrossRef   Google Scholar

    [50]

    Zhang Q, Li W, Li K, Nan H, Shi C, et al. 2020. The chromosome-level reference genome of tea tree unveils recent bursts of non-autonomous LTR retrotransposons in driving genome size evolution. Molecular Plant 13:935−38

    doi: 10.1016/j.molp.2020.04.009

    CrossRef   Google Scholar

    [51]

    Chen J, Zheng C, Ma J, Jiang C, Ercisli S, et al. 2020. The chromosome-scale genome reveals the evolution and diversification after the recent tetraploidization event in tea plant. Horticulture Research 7:63

    doi: 10.1038/s41438-020-0288-2

    CrossRef   Google Scholar

    [52]

    Wang X, Feng H, Chang Y, Ma C, Wang L, et al. 2020. Population sequencing enhances understanding of tea plant evolution. Nature Communications 11:4447

    doi: 10.1038/s41467-020-18228-8

    CrossRef   Google Scholar

    [53]

    Xia E, Tong W, Hou Y, An Y, Chen L, et al. 2020. The reference genome of tea plant and resequencing of 81 diverse accessions provide insights into its genome evolution and adaptation. Molecular Plant 13:1013−26

    doi: 10.1016/j.molp.2020.04.010

    CrossRef   Google Scholar

    [54]

    Zhang W, Zhang Y, Qiu H, Guo Y, Wan H, et al. 2020. Genome assembly of wild tea tree DASZ reveals pedigree and selection history of tea varieties. Nature Communications 11:3719

    doi: 10.1038/s41467-020-17498-6

    CrossRef   Google Scholar

    [55]

    Wang P, Yu JJ, Jin S, Chen S, Yue C, et al. 2021. Genetic basis of high aroma and stress tolerance in the oolong tea cultivar genome. Horticulture Research 8:107

    doi: 10.1038/s41438-021-00542-x

    CrossRef   Google Scholar

    [56]

    Zhang X, Chen S, Shi L, Gong D, Zhang S, et al. 2021. Haplotype-resolved genome assembly provides insights into evolutionary history of the tea plant Camellia sinensis. Nature Genetics 53:1250−59

    doi: 10.1038/s41588-021-00895-y

    CrossRef   Google Scholar

    [57]

    Liu L, Chen H, Zhu J, Tao L, Wei C. 2022. miR319a targeting of CsTCP10 plays an important role in defense against gray blight disease in tea plant (Camellia sinensis). Tree Physiology 42:1450−62

    doi: 10.1093/treephys/tpac009

    CrossRef   Google Scholar

    [58]

    Suo A, Lan Z, Lu C, Zhao Z, Pu D, et al. 2021. Characterizing microRNAs and their targets in different organs of Camellia sinensis var assamica. Genomics 113:159−70

    doi: 10.1016/j.ygeno.2020.11.020

    CrossRef   Google Scholar

    [59]

    Jeyaraj A, Elango T, Yu Y, Chen X, Zou Z, et al. 2021. Impact of exogenous caffeine on regulatory networks of microRNAs in response to Colletotrichum gloeosporioides in tea plant. Scientia Horticulturae 279:109914

    doi: 10.1016/j.scienta.2021.109914

    CrossRef   Google Scholar

    [60]

    Wang SS, Liu L, Mi XZ, Zhao SQ, An YL, et al. 2021. Multi-omics analysis to visualize the dynamic roles of defense genes in the response of tea plants to gray blight. Plant Journal 106:862−75

    doi: 10.1111/tpj.15203

    CrossRef   Google Scholar

    [61]

    Zhu C, Zhang S, Zhou C, Chen L, Zaripov T, et al. 2020. Integrated transcriptome, microRNA, and phytochemical analyses reveal roles of phytohormone signal transduction and ABC transporters in flavor formation of oolong tea (Camellia sinensis) during solar withering. Journal of Agricultural and Food Chemistry 68:12749−67

    doi: 10.1021/acs.jafc.0c05750

    CrossRef   Google Scholar

    [62]

    Zhao S, Mi X, Guo R, Xia X, Liu L, et al. 2020. The biosynthesis of main taste compounds is coordinately regulated by miRNAs and phytohormones in tea plant (Camellia sinensis). Journal of Agricultural and Food Chemistry 68:6221−36

    doi: 10.1021/acs.jafc.0c01833

    CrossRef   Google Scholar

    [63]

    Zhao L, Chen C, Wang Y, Shen J, Ding Z. 2019. Conserved microRNA act boldly during sprout development and quality formation in Pingyang Tezaocha (Camellia sinensis). Frontiers in Genetics 10:237

    doi: 10.3389/fgene.2019.00237

    CrossRef   Google Scholar

    [64]

    Liu S, Mi X, Zhang R, An Y, Zhou Q, et al. 2019. Integrated analysis of miRNAs and their targets reveals that miR319c/TCP2 regulates apical bud burst in tea plant (Camellia sinensis). Planta 250:1111−29

    doi: 10.1007/s00425-019-03207-1

    CrossRef   Google Scholar

    [65]

    Zhao S, Wang X, Yan X, Guo L, Mi X, et al. 2018. Revealing of microRNA involved regulatory gene networks on terpenoid biosynthesis in Camellia sinensis in different growing time points. Journal of Agricultural and Food Chemistry 66:12604−16

    doi: 10.1021/acs.jafc.8b05345

    CrossRef   Google Scholar

    [66]

    Zhou C, Zhu C, Fu H, Li X, Chen L, et al. 2019. Genome-wide investigation of superoxide dismutase (SOD) gene family and their regulatory miRNAs reveal the involvement in abiotic stress and hormone response in tea plant (Camellia sinensis). PLoS ONE 14:e223609

    doi: 10.1371/journal.pone.0223609

    CrossRef   Google Scholar

    [67]

    Fan K, Fan D, Ding Z, Su Y, Wang X. 2015. Cs-miR156 is involved in the nitrogen form regulation of catechins accumulation in tea plant (Camellia sinensis L.). Plant Physiology and Biochemistry 97:350−60

    doi: 10.1016/j.plaphy.2015.10.026

    CrossRef   Google Scholar

    [68]

    Sun P, Zhang Z, Zhu Q, Zhang G, Xiang P, et al. 2017. Identification of miRNAs and target genes regulating catechin biosynthesis in tea (Camellia sinensis). Journal of Integrative Agriculture 17:1154−64

    doi: 10.1016/S2095-3119(17)61654-X

    CrossRef   Google Scholar

    [69]

    Tian C, Zhou C, Zhu C, Chen L, Shi B, et al. 2022. Genome-wide investigation of the miR166 family provides new insights into its involvement in the drought stress responses of tea plants (Camellia sinensis (L.) O. Kuntze). Forests 13:628

    doi: 10.3390/f13040628

    CrossRef   Google Scholar

    [70]

    Zhu C, Zhang S, Fu H, Zhou C, Chen L, et al. 2019. Transcriptome and phytochemical analyses provide new insights into long non-coding RNAs modulating characteristic secondary metabolites of oolong tea (Camellia sinensis) in solar-withering. Frontiers in Plant Science 10:1638

    doi: 10.3389/fpls.2019.01638

    CrossRef   Google Scholar

    [71]

    Sharma D, Tiwari M, Pandey A, Bhatia C, Sharma A, et al. 2016. MicroRNA858 is a potential regulator of phenylpropanoid pathway and plant development. Plant Physiology 171:944−59

    doi: 10.1104/pp.15.01831

    CrossRef   Google Scholar

    [72]

    Li F, Wang W, Zhao N, Xiao B, Cao P, et al. 2015. Regulation of nicotine biosynthesis by an endogenous target mimicry of microRNA in tobacco. Plant Physiology 169:1062−71

    doi: 10.1104/pp.15.00649

    CrossRef   Google Scholar

    [73]

    Wang L, Tang X, Zhang S, Xie X, Li M, et al. 2022. Tea GOLDEN2-LIKE genes enhance catechin biosynthesis through activating R2R3-MYB transcription factor. Horticulture Research 9:uhac117

    doi: 10.1093/hr/uhac117

    CrossRef   Google Scholar

    [74]

    Wu L, Huang X, Liu S, Liu J, Guo Y, et al. 2020. Understanding the formation mechanism of oolong tea characteristic non-volatile chemical constitutes during manufacturing processes by using integrated widely-targeted metabolome and DIA proteome analysis. Food Chemistry 310:125941

    doi: 10.1016/j.foodchem.2019.125941

    CrossRef   Google Scholar

    [75]

    Zhu B, Chen L, Lu M, Zhang J, Han J, et al. 2019. Caffeine content and related gene expression: novel insight into caffeine metabolism in Camellia plants containing low, normal, and high caffeine concentrations. Journal of Agricultural and Food Chemistry 67:3400−11

    doi: 10.1021/acs.jafc.9b00240

    CrossRef   Google Scholar

    [76]

    Lin S, Chen Z, Chen T, Deng W, Wan X, et al. 2022. Theanine metabolism and transport in tea plants (Camellia sinensis L.): advances and perspectives. Critical Reviews in Biotechnology 00:1−15

    doi: 10.1080/07388551.2022.2036692

    CrossRef   Google Scholar

    [77]

    Ament K, Kant MR, Sabelis MW, Haring MA, Schuurink RC. 2004. Jasmonic acid is a key regulator of spider mite-induced volatile terpenoid and methyl salicylate emission in tomato. Plant Physiology 135:2025−37

    doi: 10.1104/pp.104.048694

    CrossRef   Google Scholar

    [78]

    Hong G, Xue X, Mao Y, Wang L, Chen X. 2012. Arabidopsis MYC2 Interacts with DELLA Proteins in Regulating Sesquiterpene Synthase Gene Expression. The Plant Cell 24:2635−48

    doi: 10.1105/tpc.112.098749

    CrossRef   Google Scholar

    [79]

    Zhou C, Zhu C, Li X, Chen L, Xie S, et al. 2022. Transcriptome and phytochemical analyses reveal the roles of characteristic metabolites in the taste formation of white tea during withering process. Journal of Integrative Agriculture 21:862−77

    doi: 10.1016/S2095-3119(21)63785-1

    CrossRef   Google Scholar

    [80]

    Wang J, Meng X, Dobrovolskaya OB, Orlov YL, Chen M. 2017. Non-coding RNAs and their roles in stress response in plants. Genomics, Proteomics & Bioinformatics 15:301−12

    doi: 10.1016/j.gpb.2017.01.007

    CrossRef   Google Scholar

    [81]

    Zhang G, Chen D, Zhang T, Duan A, Zhang J, et al. 2018. Transcriptomic and functional analyses unveil the role of long non-coding RNAs in anthocyanin biosynthesis during sea buckthorn fruit ripening. DNA Research 25:465−76

    doi: 10.1093/dnares/dsy017

    CrossRef   Google Scholar

    [82]

    Bordoloi KS, Baruah PM, Das M, Agarwala N. 2022. Unravelling lncRNA mediated gene expression as potential mechanism for regulating secondary metabolism in Citrus limon. Food Bioscience 46:101448

    doi: 10.1016/j.fbio.2021.101448

    CrossRef   Google Scholar

    [83]

    Varshney D, Rawal HC, Dubey H, Bandyopadhyay T, Bera B, et al. 2019. Tissue specific long non-coding RNAs are involved in aroma formation of black tea. Industrial Crops and Products 133:79−89

    doi: 10.1016/j.indcrop.2019.03.020

    CrossRef   Google Scholar

    [84]

    Wan S, Zhang Y, Duan M, Huang L, Wang W, et al. 2020. Integrated Analysis of Long Non-coding RNAs (lncRNAs) and mRNAs Reveals the Regulatory Role of lncRNAs Associated With Salt Resistance in Camellia sinensis. Frontiers in Plant Science 11:218

    doi: 10.3389/fpls.2020.00218

    CrossRef   Google Scholar

    [85]

    Li D, Jiang S, Wen X, Song X, Yang Y, et al. 2021. Sequencing and functional annotation of mRNAs and lncRNAs from tea (Camellia sinensis) leaves during infection by the fungal pathogen Lasiodiplodia theobromae. PhytoFrontiers™ 1:364−67

    doi: 10.1094/phytofr-03-21-0020-a

    CrossRef   Google Scholar

    [86]

    Zhang Y, Li P, She G, Xu Y, Peng A, et al. 2021. Molecular basis of the distinct metabolic features in shoot tips and roots of tea plants (Camellia sinensis): characterization of MYB regulator for root theanine synthesis. Journal of Agricultural and Food Chemistry 69:3415−29

    doi: 10.1021/acs.jafc.0c07572

    CrossRef   Google Scholar

    [87]

    Wang Y, Yang Y, Jiang X, Yang Y, Jiang S, et al. 2022. The sequence and integrated analysis of competing endogenous RNAs originating from tea leaves infected by the pathogen of tea leaf spot, Didymella segeticola. Plant Disease 106:1286−90

    doi: 10.1094/PDIS-06-21-1324-A

    CrossRef   Google Scholar

    [88]

    Guo D, Xia Z, Jiang X, Huang H, Yang Y, et al. 2022. Sequencing and functional annotation of competing endogenous RNAs and microRNAs in tea leaves during infection by Lasiodiplodia theobromae. PhytoFrontiers™ 2:307−12

    doi: 10.1094/phytofr-11-21-0075-a

    CrossRef   Google Scholar

    [89]

    Yang Y, Yin Q, Qiu C, Xia Z, Huang H, et al. 2022. Analysis of competing endogenous RNAs and microRNAs in tea (Camellia sinensis) leaves during infection by the leaf spot pathogen, Pestalotiopsis trachicarpicola. Molecular Plant-Microbe Interactions 35:432−38

    doi: 10.1094/mpmi-10-21-0262-a

    CrossRef   Google Scholar

    [90]

    Jiang M, Chen H, Du Q, Wang L, Liu X, et al. 2021. Genome-wide identification of circular RNAs potentially involved in the biosynthesis of secondary metabolites inSalvia miltiorrhiza. Frontiers in Genetics 12:645115

    doi: 10.3389/fgene.2021.645115

    CrossRef   Google Scholar

    [91]

    Tong W, Yu J, Hou Y, Li F, Zhou Q, et al. 2018. Circular RNA architecture and differentiation during leaf bud to young leaf development in tea (Camellia sinensis). Planta 248:1417−29

    doi: 10.1007/s00425-018-2983-x

    CrossRef   Google Scholar

    [92]

    Chen C, Zeng Z, Liu Z, Xia R. 2018. Small RNAs, emerging regulators critical for the development of horticultural traits. Horticulture Research 5:63

    doi: 10.1038/s41438-018-0072-8

    CrossRef   Google Scholar

    [93]

    Ulitsky I, Bartel DP. 2013. lincRNAs: genomics, evolution, and mechanisms. Cell 154:26−46

    doi: 10.1016/j.cell.2013.06.020

    CrossRef   Google Scholar

    [94]

    Zhou Y, Deng R, Xu X, Yang Z. 2021. Isolation of mesophyll protoplasts from tea (Camellia sinensis) and localization analysis of enzymes involved in the biosynthesis of specialized metabolites. Beverage Plant Research 1:2

    doi: 10.48130/bpr-2021-0002

    CrossRef   Google Scholar

    [95]

    Deng F, Zeng F, Shen Q, Abbas A, Cheng J, et al. 2022. Molecular evolution and functional modification of plant miRNAs with CRISPR. Trends in Plant Science 27:890−907

    doi: 10.1016/j.tplants.2022.01.009

    CrossRef   Google Scholar

    [96]

    Zhou C, Chang X, Zhu C, Cheng C, Chen Y, et al. 2022. Establishment of an efficientin planta transformation method for Camellia sinensis. Biotechnology Bulletin 38:263−68

    doi: 10.13560/j.cnki.biotech.bull.1985.2021-0635

    CrossRef   Google Scholar

    [97]

    Zhang Q, Su L, Zhang S, Xu X, Chen X, et al. 2020. Analyses of microRNA166 gene structure, expression, and function during the early stage of somatic embryogenesis in Dimocarpus longan Lour. Plant Physiology and Biochemistry 147:205−14

    doi: 10.1016/j.plaphy.2019.12.014

    CrossRef   Google Scholar

    [98]

    Ormancey M, Guillotin B, San Clemente H, Thuleau P, Plaza S, et al. 2021. Use of microRNA-encoded peptides to improve agronomic traits. Plant Biotechnology Journal 19:1687−89

    doi: 10.1111/pbi.13654

    CrossRef   Google Scholar

    [99]

    Isemura M. 2019. Catechin in human health and disease. Molecules 24:528

    doi: 10.3390/molecules24030528

    CrossRef   Google Scholar

  • Cite this article

    Zhou C, Tian C, Zhu C, Lai Z, Lin Y, et al. 2022. Hidden players in the regulation of secondary metabolism in tea plant: focus on non-coding RNAs. Beverage Plant Research 2:19 doi: 10.48130/BPR-2022-0019
    Zhou C, Tian C, Zhu C, Lai Z, Lin Y, et al. 2022. Hidden players in the regulation of secondary metabolism in tea plant: focus on non-coding RNAs. Beverage Plant Research 2:19 doi: 10.48130/BPR-2022-0019

Figures(3)  /  Tables(1)

Article Metrics

Article views(3973) PDF downloads(782)

REVIEW   Open Access    

Hidden players in the regulation of secondary metabolism in tea plant: focus on non-coding RNAs

Beverage Plant Research  2 Article number: 19  (2022)  |  Cite this article

Abstract: Non-coding RNAs (ncRNAs) are functional transcripts with minimal or no protein-coding capacity that comprise a large portion of the plant transcriptome. Among them, the microRNAs (miRNAs), linear long ncRNAs (lncRNAs), and circular long ncRNAs (circRNAs) have been widely proven to play essential regulatory roles in the biosynthesis of secondary metabolites (SMs) by modulating the expression of key synthesis-related genes in plants. Tea boasts numerous characteristic SMs, such as catechins, theanine, caffeine, volatile compounds, etc., which have distinguished health properties and largely determine the pleasant flavor quality. Thus, understanding how the tea plant produces these specialized metabolites is of great research interest. With the innovation and progress of biotechnologies in recent years, significant progress has been made in research on the regulation mechanism of SMs in tea plants at the DNA, mRNA, protein and metabolite levels. The release of the genome sequences of tea plants paves a path for precisely exploring ncRNAs and their functions in tea, and their huge potential for the biosynthesis regulation of SMs has gradually received attention. We herein summarize recent progress on miRNAs, lncRNAs, and circRNAs in tea plants and discuss their regulatory roles in the accumulation of SMs to enlighten the development of novel agronomic tools to enhance the quality of tea.

    • Tea, processed from the leaves of the tea plant [Camellia sinensis (L.) O. Kuntze], is one of the world's most popular non-alcoholic beverages[1]. The appeal of tea stems from the presence of various distinctive secondary metabolites (SMs) such as catechins, theanine, caffeine, volatile chemicals, and so on, all of which have great health benefits and play a significant role in determining the pleasant flavour quality[2]. Thus, understanding how tea produces these special metabolites has become a hot topic. Unlike primary metabolites, SMs are not essential for plant growth and development, while they are related to plant reproduction and stress resistance, and are crucial in managing the interaction between plants and their ecological environment[3]. To date, more than 1,400 specialized metabolites have been isolated and identified in tea[4]. Although most of these are widespread in the plant kingdom with similar functions, in terms of quality and quantity, catechins, caffeine, theanine, and diverse volatile compounds form the unique characteristics of tea[2]. For example, catechins are the source of the bitter and astringent taste of tea; caffeine gives tea the bitter taste; theanine is recognized as an important umami-enhancing compound in tea; and volatile compounds endow tea with abundant and diverse aroma quality[5,6]. Moreover, these distinctive SMs have substantial health benefits, making their metabolism a hot research issue and the primary focus of this paper. As the most abundant ingredients of tea, catechins account for 12%–24% of dry tea[7]. They consist of four major nongalloylated catechins and four galloylated catechins[8]. All of them are derived from the flavonoid pathway[2,9]. Caffeine is the most well-known purine alkaloid and accounts for 2%–5% of dry tea. The key enzymes involved in caffeine biosynthesis [N-methyltransferase (NMT), N-methyl nucleosidase (N-MeNase), theobromine synthase (MXMT), and tea caffeine synthase (TCS)] have been identified[10]. Theanine is a kind of non-protein amino acid, accounting for 1%–2% of dry tea. Recent studies have revealed that alanine decarboxylase (AlaDC) and theanine synthase (TSI) collaborate to regulate the biosynthesis of theanine in tea[11]. Volatile compounds have low content (accounting for less than 0.03% of dry tea) but are fundamental for tea aroma. According to tea volatile biosynthetic pathways, they can be divided into fatty acid derived volatiles (FADVs), volatile terpenes (VTs), and volatile phenylpropanoids/benzenoids (VPBs)[12]. Although the biosynthesis of these SMs follows the specific machinery, significant changes in both number and composition can be detected in varied tea varieties, seasons, geographic locations, or final tea products[13,14]. Many studies also indicated that tea SMs are synthesized to assist tea plants in adapting to their unstable environment and normal growth[15,16]. Taking advantage of this property, many man-made stresses are applied during tea growth or manufacturing processes to 'modify' tea metabolites[6,17,18]. The biosynthesis and accumulation behavior of tea SMs are tightly regulated by their biosynthetic machinery[2], and the changes in target metabolite content arise from variable expression of key synthesis-related genes[1921]. Therefore, understanding the regulatory mechanisms of gene expression related to SMs biosynthesis is crucial for improving tea quality. With the innovation and development of biotechnology in recent years, great progress has been made in understanding the regulation of secondary metabolism of tea plants at the levels of DNA, mRNA, protein and metabolites[22]. However, regarding the regulation of the transmission of information flow in DNA–RNA–protein, a class of hidden players, non-coding RNAs (ncRNAs), has been ignored for a long time.

      The ncRNAs were initially regarded as transcriptional 'noise' as they do not encode proteins or have low protein-coding potential. The latest development of high-throughput sequencing technology and experimental verification shows that ncRNAs are widely involved in the regulation of the expression of protein-encoded genes[23]. The regulatory ncRNAs mainly belong to small RNAs (18–30 nt) and long ncRNAs (> 200 nt)[24]. Specifically, microRNAs (miRNAs) are 21–24 nt in length, constituting a large portion of small RNAs in plants, which function as negative regulators to targets at the post-transcriptional or translation level or mediate DNA methylation[25]. The biogenesis process of miRNAs is widely conserved in plants, including transcription, precursor processing, methylation, and assembly of miRNA-induced silencing complex (miRISC) (as described in detail earlier by Song et al.[26] and He et al.[27]). In addition, long ncRNAs are classed into linear long ncRNAs (commonly known as lncRNAs) and circular long ncRNAs (circRNAs) and are essential modulators of protein-coding gene expression[28]. Each long ncRNA type is synthesized by a specific mechanism and has distinct regulatory properties in cis or trans form (as described in detail earlier by Yu et al.[23] and Xu et al.[29]).

      Many studies have found that ncRNA- 'protein-coding gene' interactions play important regulatory roles in plants, such as growth, development, and stress responses[23]. Emerging evidence also suggests that ncRNA-'protein-coding gene' interactions are vital in the modulation of secondary metabolism in horticultural plants[30]. However, there is relatively little research on ncRNAs, especially lncRNAs and circRNAs, concerning the tea plant. We herein summarize recent progress on tea plant miRNAs, lncRNAs, and circRNAs and discuss their regulatory roles in the production of SMs to enlighten the development of novel agronomic tools to improve the quality of tea.

    • miRNAs are the ultimate regulators of gene expression in plants, including tea plants, via direct target mRNA cleavage or translational repression[31,32], which play a key role in the regulation of secondary metabolism[33,34]. In addition, according to the competing endogenous RNA (ceRNA) hypothesis, lncRNAs, circRNAs, and mRNAs can operate as ceRNAs to competitively bind miRNA response elements (MREs) that modulate a variety of life activities[35]. Therefore, miRNAs play a role in the center of secondary metabolic regulatory networks.

    • Twenty-eight articles were retrieved from the Web of Science database (core collection) from 2010 to 2022. The studies on miRNAs in tea can be traced back to 2010. Das & Mondal[36], and Prabu & Mandal[37] identified tea conserved miRNAs and their targets from expressed sequence tags (ESTs) by computational methods. In the same way, Zhu & Luo[38] identified that 14 new miRNAs may target 51 mRNAs in tea, which took part in 13 metabolic networks. In 2012, Mohanpuria & Yadav[39] created a tea small RNA library from a mixture of two leaves and a bud and cloned six novel tea-specific miRNAs. Small RNA sequencing combined with bioinformatics prediction has been employed in several studies to identify novel miRNAs and explore their biological activities since the emergence of second-generation sequencing technology[4044]. For example, Zhang et al.[40] used Solexa sequencing technology to discover and analyze cold-responsive miRNAs and their targets derived from cold-treated tea leaves. Guo et al.[43] sequenced four small RNA libraries obtained from tea leaves exposed to four different levels of drought treatments. Also, degradome sequencing was frequently used to identify miRNA pairing information with degraded segments of their targets[4042,4547]. Due to, however, most miRNAs arising from the intergenic region, studies on the miRNAs identification and functional verification of tea plants progressed slowly and with low precision before the release of the C. sinensis genome. The draft genome of C. sinensis var. assamica 'Yunkang 10' released in 2017 paves the way for more detailed exploration of tea plant miRNAs and their functions[48]. Benefit from the revolution of third-generation sequencing technology, the draft genome of C. sinensis var. sinensis 'Shuchazao' with better assembly continuity was released the following year[49]. The genomes of tea plants were assembled to the chromosomal level[5054], even to haplotype-resolved levels[55,56], thanks to the availability of high-throughput chromatin conformation capture (Hi-C) technology. This means that raw data yielded from small RNA sequencing can be mapped with high-quality reference genomes, further increasing the accuracy of tea plant miRNA identification and providing a good basis for experimental verification of miRNA functions[5765]. Because many prediction methods for plant miRNA targets produce a substantial number of false positives in non-modal plants[66], validated miRNA-target interactions can better reflect the actual actions of miRNAs before the potential of miRNAs to regulate SMs biosynthesis can be further considered. As the biogenesis and actions of miRNA have been revealed, many experimental methods such as 5' RLM-RACE, transient co-transformation, northern blot, miRNA-agomir and miRNA-antagomir treatments, etc., have been used to verify miRNA-target interactions[30]. We briefly summarize the currently validated miRNA-target pairs in the tea plant from 18 published articles (Table 1). There were 75 miRNA-target interactions which have been validated at least using the 5'RLM-RACE method. Similar to other plants, most of the miRNAs' targets are TF genes in the tea plant[30], indicating miRNAs have a wide range of regulatory functions. In addition, many miRNA-target pairs specific to tea plants were found to be involved in the regulation of the biosynthesis of SMs. Next, it should be clarified under what conditions miRNAs can regulate the biosynthesis of SMs and we summarize the relevant content below.

      Table 1.  Summary of validated miRNA-target pairs in tea plant.

      miRNA famliymiRNA nameTargetActionVerification methodsBiology functionsReference
      miR156miR156fSPL-3 (Squamosa promoter-binding-like protein)Cleavage5’RLM-RACE; qRT-PCRColletotrichum gloeosporioides immune response[59]
      miR156iSBP (S-RNase binding protein)Cleavage5’RLM-RACE; qRT-PCRInfection of Pestalotiopsis-like species response[60]
      miR156SBP3Cleavage5’RLM-RACE; qRT-PCR; northern blot/co-transformation in tobacco leavesRegulation of the biosynthesis of catechins[62]
      miR156f-3pSBP; AP2/ERF (Ethylene-
      responsive transcription factors)
      Cleavage5’RLM-RACE; qRT-PCR; northern blotRegulation of the biosynthesis of terpenoids[65]
      miR156SPLCleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of catechins[67]
      miR156g-3pF3H (flavanone 3-hydroxylase)Cleavage5’RLM-RACERegulation of the biosynthesis of catechins[68]
      miR159miR159aGAMYBCleavage5’RLM-RACE; qRT-PCRThe transformation of plant development stages[34]
      miR160miR160a-5pARF17 (Auxin responsive factor17)Cleavage5’RLM-RACE; qRT-PCRAbiotic stresses response[45]
      miR160cARFCleavage5’RLM-RACE; qRT-PCRC. gloeosporioides immune response[47]
      miR160kDirigentCleavage5’RLM-RACE; qRT-PCRInfection of Pestalotiopsis-like species response[60]
      miR164miR164aNAC100 (NAC domain transcription factors100)Cleavage5’RLM-RACE; qRT-PCRRegulation of leaf development[45]
      miR164aNACCleavage5’RLM-RACE; qRT-PCRC. gloeosporioides immune response[47]
      miR164aNAC-17Cleavage5’RLM-RACE; qRT-PCRC. gloeosporioides immune response[59]
      miR164aNACCleavage5’RLM-RACE; qRT-PCR; northern blotRegulation of the biosynthesis of catechins[62]
      miR166miR166HD-ZIP III (Homeodomain-
      leucine zipperIII)
      Cleavage5’RLM-RACE; qRT-PCRDrought stress response[69]
      miR166aHD-ZIP4Cleavage5’RLM-RACE; qRT-PCR; northern blotRegulation of the biosynthesis of catechins[62]
      miR166d-5p_1ABCC1-2 (ABC transctipters)Cleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of flavonoid[61]
      miR166d-5p_1ABCG2Cleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of terpenoids[61]
      miR167miR167aCHL (Chalcone isomerase)Cleavage5’RLM-RACE; qRT-PCRRegulation of flavonoid biosynthesis[44]
      miR167aARF6Cleavage5’RLM-RACERegulation of the biosynthesis of catechins[44]
      miR167aARF8Cleavage5’RLM-RACERegulation of the biosynthesis of catechins[44]
      miR167ARFCleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of catechins[34]
      miR167d_1ARFCleavage5’RLM-RACE; qRT-PCR; miRNA-agomir/antagomirRegulation of the biosynthesis of flavonoid and terpenoids[61]
      miR167d_1GH3 (Flavanone 3-hydroxylase)CleavageqRT-PCR; miRNA-agomir/antagomirRegulation of the biosynthesis of flavonoid and terpenoids[61]
      miR167dARFCleavage5’RLM-RACE; qRT-PCR; northern blotRegulation of the biosynthesis of catechins[62]
      miR169miR169eNFY (Nuclear transcription
      factor Y)
      Cleavage5’RLM-RACE; qRT-PCRC. gloeosporioides immune response[47]
      miR169d-5p_1ACX (acyl-CoA oxidase)Cleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of JA/MeJA and terpenoids[70]
      miR169aNF-YA (Nuclear
      factor Y)
      Cleavage5’RLM-RACE; qRT-PCR; northern blotRegulation of the biosynthesis of theanine and caffeine[62]
      miR171miR171b-3pCRE1 (cytokinin receptor 1)Cleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of flavonoid and terpenoids[61]
      miR171b-3p_2DELLA1 (DELLA protein1)Cleavage5’RLM-RACE; qRT-PCR; miRNA-agomir/antagomirRegulation of the biosynthesis of flavonoid and terpenoids[61]
      miR172miR172kERF_RAP2-7Cleavage5’RLM-RACE; qRT-PCRC. gloeosporioides immune response[59]
      miR172g-3pMYC2Cleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of terpenoids[61]
      miR319miR319cTPC (TCP family transcription factor)Cleavage5’RLM-RACEC. gloeosporioides immune response[59]
      miR395miR395eAPS (Aspartic proteases)Cleavage5’RLM-RACE; qRT-PCRInfection of Pestalotiopsis-like species response[60]
      miR396miR396b-5pGRF-1Cleavage5’RLM-RACE; qRT-PCRC. gloeosporioides immune response[59]
      miR396b-5PF3HCleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of flavonoid[61]
      miR396GRFCleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of catechins[34]
      miR396dGRF1Cleavage5’RLM-RACE; qRT-PCR; northern blotRegulation of the biosynthesis of catechins[62]
      miR396dGRF4Cleavage5’RLM-RACE; qRT-PCR; northern blot; co-transformationRegulation of the biosynthesis of catechins[62]
      miR396dGRF13Cleavage5’RLM-RACE; qRT-PCR; northern blotRegulation of the biosynthesis of catechins[62]
      miR397miR397LAC17 (Laccase17)Cleavage5’RLM-RACE; qRT-PCRInfection of Pestalotiopsis-like species response[60]
      miR398miR398a-3p-1CSD4 (Cu/Zn-Superoxide dismutases4)Cleavage5’RLM-RACE; qRT-PCRCold stress response[66]
      miR408miR408-3p_2ss18GT19GTSTK1 (Ser/Thr-protein Kinse)Cleavage5’RLM-RACE; qRT-PCRC. gloeosporioides immune response[47]
      miR447miR447g-p5PAL (Phenylalanine ammonia-lyase)Cleavage5’RLM-RACE; qRT-PCRC. gloeosporioides immune response[59]
      miR529miR529dCHI (chalcone isomerase)Cleavage5’RLM-RACE;Regulation of the biosynthesis of catechins[68]
      miR530miR530bERF96Cleavage5’RLM-RACE; qRT-PCR; northern blot/co-transformationInfection of Pestalotiopsis-like species response[60]
      miR530aDHBPCleavage5’RLM-RACE; qRT-PCRInfection of Pestalotiopsis-like species response[60]
      miR828miR828WERCleavage5’RLM-RACE; qRT-PCRRegulation of plant development[45]
      miR828aMYB (MYB domain protein)Cleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of terpenoids[65]
      miR828aMYB75Cleavage5’RLM-RACE; qRT-PCRC. gloeosporioides immune response[47]
      miR845miR845ABCC1-3Cleavage5’RLM-RACE; qRT-PCR/miRNA-agomir/antagomirRegulation of the biosynthesis of flavonoid and terpenoids[61]
      miR845ABCC2CleavageqRT-PCR; miRNA-agomir/antagomirInhibite flavonoid biosynthesis but enhance terpenoid biosynthesis[61]
      miR858miR858aMYB12Cleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of polyphenols[45]
      miR858b_R-3MYBCleavage5’RLM-RACE; qRT-PCR; northern blotRegulation of the biosynthesis of terpenoids[65]
      miR858bMYBCleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of terpenoids[65]
      miR858aR2R3-MYBCleavage5’RLM-RACE; qRT-PCRC. gloeosporioides immune response[59]
      858b_R-3MYB114Cleavage5’RLM-RACE; qRT-PCRC. gloeosporioides immune response[47]
      miR2593miR2593eANR (Anthocyanidinreductase)Cleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of catechins[44]
      miR2868miR2868LAR (leucoanthocyanidin reductase)Cleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of catechins[68]
      miR3444miR3444bDFR (Dihydroflavonol 4-reductase)Cleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of catechins[44]
      miR3630miR3630-3p_L-3pMYCCleavageqRT-PCR; northern blotRegulation of the biosynthesis of terpenoids[65]
      miR4380miR4380aDFRCleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of catechins[62]
      miR5240miR5240DFRCleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of catechins[68]
      miR5251miR5251C4HCleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of catechins[62]
      miR5559miR5559-5pANR1; ANR2Cleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of catechins[68]
      miR5564miR5564ANR2Cleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of catechins[68]
      miR7777miR7777-5pC4HCleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of catechins[62]
      miR7814miR7814CHS1 (chalcone synthase1)Cleavage5’RLM-RACERegulation of the biosynthesis of catechins[68]
      miR7814CHS3Cleavage5’RLM-RACERegulation of the biosynthesis of catechins[68]
      miR8757miR8757fPPO (Polyphenol oxidase)Cleavage5’RLM-RACE; qRT-PCRInfection of Pestalotiopsis-like species response[60]
      novel miRNAmiRn211TLP (thaumatin-like protein)Cleavage5’RLM-RACE; qRT-PCR; northern blot; co-transformationInfection of Pestalotiopsis-like species response[60]
      PC-5p-80764_22WRKY (WRKY transcription
      factor)
      Cleavage5’RLM-RACE; qRT-PCRC. gloeosporioides and abiotic stresses response[47]
      PC-3p-81-33418AP2/ERFCleavageqRT-PCR; northern blotRegulation of the biosynthesis of terpenoids[65]
      novel_miR44LOX (Lipoxygenase)Cleavage5’RLM-RACE; qRT-PCRRegulation of the biosynthesis of JA/MeJA and terpenoids[70]
    • miRNAs can negatively modulate the content of related SMs in plants by directly regulating the expression of target genes in SM biosynthesis pathways. However, the opposite may be true if the two metabolites belong to different synthetic branches of a metabolic pathway and have a common synthetic precursor[39]. For instance, Sharma et al.[71] identified that miR858 can positively modulate the content of flavonoids, because inhibiting the expression of miR858 can increase the expression level of its targets MYBs, which results in the metabolic flow redirecting to the flavonoid synthesis at the expense of lignin synthesis. MiRNAs can regulate the amount of a class of chemicals belonging to the same metabolic pathway because the majority of their target genes are TFs[33]. While the roles of structural genes, in general, is not as extensive as that of TFs, miRNA-structural gene interactions are more targeted for regulating specific SM content. For instance, the nta-miRX27-QPT2 interaction specific to Nicotiana tabacum has been confirmed to have a regulatory function in nicotine accumulation[72]. In tea plants, miRNA-target pairs modulate the biosynthesis of SMs show lineage-specific target genes.

    • An illustration of the general flavonoid pathway leading to the biosynthesis of catechins has been described (Fig. 1). Their accumulation is affected by developmental cues and environmental stimuli[73]. During which miRNAs are confirmed to be involved in the modulation of catechins (Fig. 1).

      Figure 1. 

      Schematic representation of the biosynthesis of catechins regulated by ncRNAs in tea plants. The miRNA, circRNA, transcription factor gene, structural gene, and phytohormone are marked in the orange triangles, yellow ovals, light blue ovals, light green boxes, and dark blue ovals, respectively. The solid line with the arrow indicates direct regulation; the dashed line with the arrow is supported by less experimental evidence or prediction results; the black and red T-shaped lines represent negative and positive regulation, respectively; the black dashed T-shaped line is supported by less experimental evidence or prediction results. The miRNA-target pairs shown in the figure have been verified to have regulatory relationships.

      The content of catechins varies between tissues of the tea plant, which can be mediated by miRNAs. In general, the total amount of catechins collected in young leaves and buds was greater than that in mature leaves[34,44,62]. Sun et al.[44] investigated the functions of miRNAs in modulating the accumulation of catechins in tea buds, first to fifth leaves, and mature leaves. It was found that miRNA167a-CHI, miR2593e-ANR, miR4380a-DFR, miR3444b-DFR, csn-miR5251-C4H, and miR7777-5p.1-C4H pairs may be involved in the regulation of catechin accumulation in tea leaves of different maturity by integrated 5'RLM-RACE and qRT-PCR analyses[44]. Among these, miR3444b and DFR showed a similar expression trend instead of an opposite expression pattern. Considering the mode of action of miRNAs negatively regulating targets, the effect of miR3444b-DFR on the regulation of catechins needs to be further validated. At the same time, miR529d and miR156g-3p were shown to be involved in the regulation of catechin accumulation in the first, third, and oldest leaves of the tea plant, respectively, via targeting CHI and flavanone 3-hydroxylase (F3H)[68]. Although the miR7814-CHS1, miR7814-CHS3, miR5240-DFR, miR2868-LAR, miR5559-5p-ANR1, miR5559-5p-ANR2, and miR5264-ANR2 interactions that exist in tea plants have also been validated, they were not related to the regulation of differential accumulation of catechins in different leaf positions[68]. This requires further elucidation of the regulatory role of these miRNAs on catechins synthesis under what conditions, such as abiotic stress and biotic stress. In addition to structural genes, miRNA-TF gene modules have been implicated in the control of catechin accumulation. Zhao et al.[62] conducted a thorough investigation to explore the regulatory mechanisms that miRNA-mediated the biosynthesis of taste compounds in the different tea leaf positions. It was found that miR156, miR164a, miR166a, miR167d, and miR396d may negatively regulate the the accumulation of catechins by targeting S-RNase binding protein3 (SBP3), NAC domain transcription factor (NAC), homeodomain-leucine zipper4 (HD-ZIP4), auxin responsive factor (ARF), and growth-regulating factor1, 4, and 13 (GRF1, 4, and 13) cleavage, respectively. These modules and phytohormones, such as indole-3-acetic acid (IAA), jasmonic acid (JA), abscisic acid (ABA), zeatin (ZA), and salicylic acid (SA), might synergistically regulate the biosynthesis of catechins although there is no direct evidence to support this claim[62]. Li et al.[34] studied the regulatory link between essential taste components and miRNAs in different tissues of the tea plant. According to integrated 5'RLM-RACE and qRT-PCR validation results, three miRNA-target pairs (miR159-GAMYB, miR167-ARF, and miR396-GRF) may cause the distinct accumulation of catechins, yet this has not been verified by experiments.

      The accumulation of catechins in tea leaves differs depending on the nitrogen form, which can be regulated by miRNAs. miR156 regulated the accumulation of catechins in tea shoots by inhibiting the expression of the target SPL, a TF gene that promote the expression of DFR[67]. In the presence of $\text{NO}^-_3 $, the expression of miR156 would be inhibited thereby promoting the accumulation of catechins compared to the presence of $ \text{NH}^+_4$.

      miRNAs can mediate the biosynthetic modulation of catechins during the manufacturing processes of postharvest tea leaves. The accumulation of SM in tea leaves during the preharvest stage serves as the foundation for developing finished tea quality, while the manufacturing processes of postharvest tea leaves determine the final flavor quality of the finished tea. Especially in withering and turning-over processes, tea leaves remain alive under multi-stresses[74]. During this time, miRNAs have been discovered to control catechin metabolism via miRNA-TF gene modules. Solar-withering is a crucial process in tea production that can improve the palatability of tea by reducing catechin content appropriately[70]. Tea leaves are subjected to UV irradiation, heat, and dryness during the withering process, which can cause stress responses that are mediated by miRNAs to have an impact on the accumulation of related SMs[61]. The accumulation of catechins was linked to the miR845-ABCC1-3/ABCC2, miR319c_3-PIF-ARF, miR166d-5p_1-ABCC1-2, and, miR167d_1-ARF-GH3 modules[61]. These suggested that when the tea plant is under abiotic stress, miRNAs may indirectly regulate the accumulation of catechins by regulating genes related to phytohormone synthesis or metabolite transport.

    • Caffeine biosynthesis in tea plants is relatively clear[2] (Fig. 2). The regulatory mechanism of protein-coding genes affecting caffeine synthesis has been widely explored[20,75], but there are relatively few studies on the regulatory mechanism of miRNA affecting caffeine synthesis. Similarly, the study of theanine metabolism in tea plants has achieved a breakthrough[11,76] (Fig. 2), but the regulatory miRNAs involved in theanine biosynthesis remain largely unknown. To date, only one case of validated miRNA-target pair being involved in regulating caffeine and theanine synthesis has been reported[62]. The miR169 was proved to target NF-YA cleavage by 5'RLM-RACE. The decrease of miR169 expression from the tea bud to the fourth leaf will enhance the expression of its target NF-YA, indicating that the expression level of miR169 positively linked with the content of caffeine and theanine. According to weighted gene co-expression network analysis (WGCNA), it was found that both the content of IAA, JA, ZA, and ABA and the content of caffeine and theanine had significant correlations with the expression of miR169. It suggested that miRNA-mediated modulation of phytohormone crosstalk during specific developmental processes may modulate the caffeine and theanine biosynthesis. Of course, interesting findings require more experiments for verification.

      Figure 2. 

      Schematic representation of the biosynthesis of caffeine and theanine regulated by ncRNAs in tea plants. The miRNA, transcription factor gene, structural gene, and phytohormone are marked by the orange triangle, light blue oval, light green boxes, and dark blue ovals, respectively. The solid line with the arrow indicates direct regulation; the dashed line with the arrow is supported by less experimental evidence or prediction results; and the black T-shaped line represents negative regulation. The miRNA-target pairs shown in the figure have been verified to have regulatory relationships.

    • VTs mainly include monoterpenes and sesquiterpenes, which can be synthesized via the plasmatic methylerythritol phosphate (MEP) pathway and the cytosolic mevalonic acid (MVA) pathway (Fig. 3). They are the primary components of floral and fruity aromas in tea[12]. The content of VTs in preharvest tea leaves are relatively low. Both the environmental stimuli and the abiotic stresses induced by the postharvest manufacturing processes can lead to tea leaves accumulating high levels of VTs. It has recently been shown that miRNAs regulate to the biosynthesis of VTs in tea leaves.

      Figure 3. 

      Schematic representation of the biosynthesis of (a) JA/MeJA and (b) volatile terpenoids in tea plants. The miRNA, circRNA, lncRNA, transcription factor gene, structural gene, and phytohormone are marked in the orange triangles, yellow ovals, green ovals, light blue ovals, light green boxes, and dark blue ovals, respectively. The solid line with the arrow indicates direct regulation; the dashed line with the arrow is supported by less experimental evidence or prediction results; the black T-shaped line indicates negative regulation; the black dashed T-shaped line is supported by less experimental evidence or prediction results. The miRNA-target pairs shown in the figure have been verified to have regulatory relationships.

      Tea plant miRNAs can mediate the regulation of differential synthesis of VTs in different growing stages. Zhao et al.[65] investigated the relationship between miRNAs and VTs and found that four validated miRNA-target interactions, including miR156-SPL, miR3630-MYC, miR858b_R-3-MYB, and PC-3p-81_33418-ERF, may be involved in VTs accumulation at different growing time points. However, due to the lack of experimental evidence of direct interaction between TFs and structural genes related to VT synthesis, the above results are only speculative based on WGCNA.

      Tea plant miRNAs can intervene in the biosynthetic regulation of VTs during withering. Zhu et al.[70] validated that novel_miR44 and miR169d-5p_1 can negatively regulate JA biosynthesis by cleavage of the upstream structural genes of JA synthesis LOX (lipoxygenase) and ACX (acyl-CoA oxidase), respectively (Fig. 3a)[70]. Many studies have shown that JA promotes the biosynthesis of VTs in plants[77], which may be crucial for increasing the content of VTs. Moreover, miR166d-5p_1-ABCG2-MYC2 and miR171b-3p_2-DELLA-MYC2 modules were involved in VTs biosynthesis indirectly[61] since the MYC2 acts as a major regulator of JA biosynthesis[78]. Due to the fact that miRNAs have a wide range of regulatory functions when their targets are TFs, both miRNAs (miR166d-5p_1-ABCG2-MYC2 and miR171b-3p_2-DELLA-MYC2 modules), acting as positive regulators, have boosted VT accumulation in sunlight-withered leaves (Fig. 3b). Undoubtedly, the indirect regulatory functions of miRNAs on SMs need further research for verification.

    • Tea leaves are subjected to several abiotic stresses during postharvest processing. Because the leaf cells are alive, transcriptional alterations would have a significant impact on the accumulation of SMs[79]. Many plants have been reported to have lncRNAs and circRNAs that are involved in the abiotic stress response[80]. The large potential of lncRNAs and circRNAs in the modulation of SMs biosynthesis is also gradually being taken seriously by tea scientists.

    • Many studies have shown that lncRNAs play important roles in SMs biosynthesis by regulating related gene expression in cis or trans[72,81,82]. Because a large part of lncRNAs arises from non-coding regions, there was no research on tea plant lncRNAs until 2019, i.e. after the publication of the tea plant genome sequences[83]. So far, to the best of our knowledge, there have been eight articles published concerning tea plant lncRNAs[70,8389], which showed that lncRNAs were involved in response to pathogenic microorganisms attack[85,8789], salt stress[84], and regulation of secondary metabolism[70,83,86]. We focus on lncRNAs that relate in this section. Varshney et al.[83] first employed the bioinformatics prediction method that identified 33,400 lncRNAs of tea plants based on 170 public RNA-seq data from 11 harvestable tissues. That same year, Zhu et al.[70] performed a systemic analysis of lncRNAs during tea production (Fig. 3b). It was identified that 32,036 lncRNAs from fresh leaves, solar-withered leaves, and indoor-withered leaves of tea plant. An analysis of the differentially expressed lncRNAs and their target genes were identified as related to flavonoid metabolism as well as terpenoid accumulation. In addition, by integrated bioinformatics prediction and qRT-PCR analysis, lncRNAs LTCONS_00026271 and LTCONS_00020084 were preliminarily proven that can serve as the endogenous target mimics (eTMs) of novel_miR44 and miR169d-5p_1, respectively, that promote the biosynthesis of VTs in solar-withered leaves (Fig. 3). In another study, a total of 1,679 lncRNAs were discovered from the full-length transcriptome sequencing data of shoot tips and roots of tea plants. The bioinformatics prediction showed that lincRNAs are closely related to the regulation of characteristic secondary metabolites, such as catechins, theanine, and caffeine by influencing the expression of related genes[86]. These results suggested that lncRNAs are ubiquitous regulators in tea plant SMs biosynthesis.

    • The potential involvement of circRNAs in the modulation of biosynthesis of SMs in plants has been explored in Salvia miltiorrhiza[90]. In 2018, Tong et al.[91] first pursued an analysis of circRNAs in the tea plant and identified 342 high-confidence circRNAs. The expression levels of circRNAs were positively correlated with the expression levels of their parental transcripts in different mature leaves. Moreover, the gene sets encoding circRNAs revealed their candidate roles in metabolite biosynthesis base on the KEGG enrichment analysis. Subsequently, the circRNAs functioning as miRNA sponges during the tea withering process was predicted in another study. Combined with qRT-PCR analysis, circRNA498, circRNA1806, circRNA212/circRNA533, and circRNA174 were preliminarily proved that can mediate flavor-related metabolites by sponging corresponding miRNAs related to phytohormone signaling and ATP binding cassette (ABC) transporters[61] (Fig. 1 & Fig. 3b). Given the limited number of studies on tea circRNAs, the above-described circRNA functions may not reflect their general rules in the regulation of SMs biosynthesis in tea. In the future, functional studies will be useful for uncovering their regulatory functions.

    • We summarized recent research progress in the biosynthetic regulation of tea SMs exerted by ncRNAs. In general, the mechanisms by which tea plant miRNAs regulate the biosynthesis of SMs have been demonstrated by some experimental evidence. It is also encouraging to note that the potential of lncRNAs and circRNAs in regulating SM biosynthesis in tea have also received increasing attention from scholars. The rapid development of second-generation sequencing technology and the increasingly powerful bioinformatics tools accelerating the application of large-scale sequencing efforts to functional studies of tea plant ncRNAs. As data accumulates, it will be exciting to witness further developments in our understanding of the functional roles of ncRNAs in the biosynthesis modulation of SMs in tea. Future research on tea plant ncRNAs can be conducted mainly on the following aspects:

      (1) Establishment of a tea plant ncRNA database is needed. To facilitate the use and mining of plant ncRNA data, many published ncRNA databases have been established. However, the tea plant has also evolved a variety of specific ncRNA-mediated regulatory mechanisms on unique SMs[44,68]. To date, it is inconvenient to obtain detailed information such as the sequences, chromosome locations, functions, and annotations when researchers look at particular ncRNA-target interactions involved in the biosynthesis modulation of SMs in tea plants. Hence, a public database for tea plant ncRNA sequences and annotation are urgently required.

      (2) More experimental methods need to be applied to validate the functions of ncRNAs-mediated biosynthetic regulation of SMs in tea. Generally, plant miRNAs are classified into conserved miRNAs, less conserved miRNAs, and species-specific miRNAs that are present in angiosperms, a lineage or group of plants, and a single species, respectively[92]. While lncRNAs have an even lower degree of conservation among plant kingdoms[93]. To date, the specific functions of many ncRNAs have been clarified in model plants such as A. thaliana, Oryza sativa, N. tabacum, etc., which have stable genetic transformation systems[23,26]. Compared with ncRNA studies in model plants, progress in tea plants is relatively slow, especially because of the lack of a tea plant genetic transformation system for functional verification of ncRNAs. As described above, most regulatory mechanisms of ncRNA-mediated accumulation of tea SMs were indirectly demonstrated by qRT-PCR combined with bioinformatics analysis, or other in vitro experiments[61,70]. Several in vivo functional identification of miRNAs have used heterologous expression systems or tea transient transformation systems[57,60,61]. To make ncRNAs promising regulatory tools for tea quality improvement, the specific regulatory roles of tea plant ncRNAs on secondary metabolism in vivo need to be analyzed in detail. Recently, an efficient mesophyll protoplast isolation method for tea has been established[94], which provided a foundation for chitosan-complexed carbon nanotube carriers and nanoparticles with polyethylene glycol-mediated protoplast transformation in tea plants in the future[95]. In addition, an in planta transformation approach for gene functional verification of tea plants has been reported recently[96]. In the future, it is believed that vector construction and genetic transformation based on the in planta transformation system opens an intriguing research area to verify the specific ncRNA-target functions in tea plants.

      (3) The miRNAs have the potential as important agronomic tools to improve tea quality. If the expected quality of tea is to be improved via conventional breeding methods, long-term efforts are needed to identify and introduce beneficial traits into the tea plants. Moreover, the lack of an effective genetic transformation system for tea plants makes the advantage of transgenics over traditional breeding in terms of time disappear. Recently, it has been found that the first ORF after the transcription start site of plant primary miRNA (pri-miRNAs) corresponded to a translated ORF coding regulatory peptides called miRNA-encoded peptides (miPEPs), able to increase the expression of pri-miRNAs[97]. Ormancey et al. proved that watering plants with miPEPs can lead to an overall change in plant development by elevating the expression of corresponding miRNAs and, correlatively, decreasing the expression of the respective miRNA target genes, thus being a suitable alternative to the use of chemicals in agronomy[98]. While we are working on the molecular mechanisms of tea quality regulation, are there any methods that have the potential to be quickly and cost-effectively applied to the quality improvement of tea? Let's make a hypothesis. Catechins undoubtedly have a wide range of health benefits for the human body[99]. In addition, catechins are responsible for bitter and astringent tastes in the tea infusion[8]. Thus, moderately reducing the content of catechins may improve the palatability of tea. Many miRNAs have been found to negatively regulate the accumulation of catechins[44,68]. Suppose that their nascent pri-miRNAs have the potential to encode miPEPs. In that case, we can imagine using a cocktail of several miPEPs to increase the expression of their nascent pri-miRNAs and decrease the expression of the respective miRNA targets correlatively, thereby achieving the fine-tuning of flavor quality of tea.

      • This work was supported by the 6.18 Tea Industry Technology Branch of Collaborative Innovation Institute (K1520001A), the Construction Project for Technological Innovation and Service System of Tea Industry Chain of Fujian Agriculture and Forestry University (K1520005A01), the Rural Revitalization Tea Industry Technical Service Project of Fujian Agriculture and Forestry University (11899170145), the 'Double First-class' Scientific and Technological Innovation Capacity and Enhancement Cultivation Plan of Fujian Agriculture and Forestry University (KSYLP004), the Construction of Plateau Discipline of Fujian Province (102/71201801101), and the Tea Industry Branch of Collaborative Innovation Institute of Fujian Agriculture and Forestry University (K1521015A). The authors thank Xin Huang, Master of Translation, from the School of Foreign Languages of Fuzhou University for her linguistic assistance.

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

      • Copyright: © 2023 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 (3)  Table (1) References (99)
  • About this article
    Cite this article
    Zhou C, Tian C, Zhu C, Lai Z, Lin Y, et al. 2022. Hidden players in the regulation of secondary metabolism in tea plant: focus on non-coding RNAs. Beverage Plant Research 2:19 doi: 10.48130/BPR-2022-0019
    Zhou C, Tian C, Zhu C, Lai Z, Lin Y, et al. 2022. Hidden players in the regulation of secondary metabolism in tea plant: focus on non-coding RNAs. Beverage Plant Research 2:19 doi: 10.48130/BPR-2022-0019

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return