Search
2022 Volume 2
Article Contents
REVIEW   Open Access    

Research progress on biological functions of lncRNAs in major vegetable crops

  • # These authors contributed equally: Nan Li, Yujie Wang

More Information
  • With the advances in genomics and bioinformatics, particularly the extensive application of high-throughput sequencing technology, a large number of non-coding RNAs (ncRNAs) have been discovered, of which long ncRNAs (lncRNAs) refer to a class of transcripts that are more than 200 nucleotides in length. Accumulating evidence demonstrates that lncRNAs play significant roles in a wide range of biological processes, including regulating plant growth and development as well as modulating biotic and abiotic stress responses. Although the study of lncRNAs has been a hotspot of biological research in recent years, the functional characteristics of plant lncRNAs are still in their initial phase and face great challenges. Here, we summarize the characteristics and screening methods of lncRNAs and highlight their biological functions in major vegetable crops, including tomato, Brassica genus crops, cucumber, pepper, carrot, radish, potato, and spinach, which are implicated in the interaction of lncRNAs and miRNAs. This review enhances the understanding of lncRNAs' roles and can guide crop improvement programs in the future.
  • 加载中
  • Supplemental Table S1 Published papers on lncRNAs in major vegetable crops.
  • [1]

    Wilhelm BT, Marguerat S, Watt S, Schubert F, Wood V, et al. 2008. Dynamic repertoire of a eukaryotic transcriptome surveyed at single-nucleotide resolution. Nature 453:1239−43

    doi: 10.1038/nature07002

    CrossRef   Google Scholar

    [2]

    Wang H, Niu QW, Wu HW, Liu J, Ye J, et al. 2015. Analysis of non-coding transcriptome in rice and maize uncovers roles of conserved lncRNAs associated with agriculture traits. The Plant Journal 84:404−16

    doi: 10.1111/tpj.13018

    CrossRef   Google Scholar

    [3]

    Zhang P, Wu W, Chen Q, Chen M. 2019. Non-Coding RNAs and their Integrated Networks. Journal of Integrative Bioinformatics 16:20190027

    doi: 10.1515/jib-2019-0027

    CrossRef   Google Scholar

    [4]

    Mercer TR, Dinger ME, Mattick JS. 2009. Long non-coding RNAs: insights into functions. Nature Reviews Genetics 10:155−59

    doi: 10.1038/nrg2521

    CrossRef   Google Scholar

    [5]

    Jha UC, Nayyar H, Jha R, Khurshid M, Zhou M, et al. 2020. Long non-coding RNAs: emerging players regulating plant abiotic stress response and adaptation. BMC Plant Biology 20:466

    doi: 10.1186/s12870-020-02595-x

    CrossRef   Google Scholar

    [6]

    Wierzbicki AT, Haag JR, Pikaard CS. 2008. Noncoding transcription by RNA polymerase Pol IVb/Pol V mediates transcriptional silencing of overlapping and adjacent genes. Cell 135:635−48

    doi: 10.1016/j.cell.2008.09.035

    CrossRef   Google Scholar

    [7]

    Gil N, Ulitsky I. 2020. Regulation of gene expression by cis-acting long non-coding RNAs. Nature Reviews Genetics 21:102−17

    doi: 10.1038/s41576-019-0184-5

    CrossRef   Google Scholar

    [8]

    Ponting CP, Oliver PL, Reik W. 2009. Evolution and functions of long noncoding RNAs. Cell 136:629−41

    doi: 10.1016/j.cell.2009.02.006

    CrossRef   Google Scholar

    [9]

    Guil S, Esteller M. 2012. Cis-acting noncoding RNAs: friends and foes. Nature Structural & Molecular Biology 19:1068−75

    doi: 10.1038/nsmb.2428

    CrossRef   Google Scholar

    [10]

    Fatica A, Bozzoni I. 2014. Long non-coding RNAs: new players in cell differentiation and development. Nature Reviews Genetics 15:7−21

    doi: 10.1038/nrg3606

    CrossRef   Google Scholar

    [11]

    Zhang H, Guo H, Hu W, Ji W. 2020. The emerging role of long non-coding RNAs in plant defense against fungal stress. International Journal of Molecular Sciences 21:2659

    doi: 10.3390/ijms21082659

    CrossRef   Google Scholar

    [12]

    Zhao X, Li J, Lian B, Gu H, Li Y, et al. 2018. Global identification of Arabidopsis lncRNAs reveals the regulation of MAF4 by a natural antisense RNA. Nature Communications 9:5056

    doi: 10.1038/s41467-018-07500-7

    CrossRef   Google Scholar

    [13]

    Wang Y, Luo X, Sun F, Hu J, Zha X, et al. 2018. Overexpressing lncRNA LAIR increases grain yield and regulates neighbouring gene cluster expression in rice. Nature Communications 9:3516

    doi: 10.1038/s41467-018-05829-7

    CrossRef   Google Scholar

    [14]

    Zhou Y, Zhang Y, Sun Y, Yu Y, Lei M, et al. 2021. The parent-of-origin lncRNA MISSEN regulates rice endosperm development. Nature Communications 12:6525

    doi: 10.1038/s41467-021-26795-7

    CrossRef   Google Scholar

    [15]

    Song X, Hu J, Wu T, Yang Q, Feng X, et al. 2021. Comparative analysis of long noncoding RNAs in angiosperms and characterization of long noncoding RNAs in response to heat stress in Chinese cabbage. Horticulture Research 8:48

    doi: 10.1038/s41438-021-00484-4

    CrossRef   Google Scholar

    [16]

    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

    [17]

    Liu X, Hao L, Li D, Zhu L, Hu S. 2015. Long non-coding RNAs and their biological roles in plants. Genomics, Proteomics & Bioinformatics 13:137−47

    doi: 10.1016/j.gpb.2015.02.003

    CrossRef   Google Scholar

    [18]

    Fabbri M, Girnita L, Varani G, Calin GA. 2019. Decrypting noncoding RNA interactions, structures, and functional networks. Genome Research 29:1377−88

    doi: 10.1101/gr.247239.118

    CrossRef   Google Scholar

    [19]

    Carlevaro-Fita J, Johnson R. 2019. Global Positioning System: Understanding long noncoding RNAs through subcellular localization. Molecular Cell 73:869−83

    doi: 10.1016/j.molcel.2019.02.008

    CrossRef   Google Scholar

    [20]

    Kopp F, Mendell JT. 2018. Functional classification and experimental dissection of long noncoding RNAs. Cell 172:393−407

    doi: 10.1016/j.cell.2018.01.011

    CrossRef   Google Scholar

    [21]

    Wu J, Liu C, Liu Z, Li S, Li D, et al. 2019. Pol III-dependent cabbage BoNR8 long ncRNA affects seed germination and growth in Arabidopsis. Plant & Cell Physiology 60:421−35

    doi: 10.1093/pcp/pcy220

    CrossRef   Google Scholar

    [22]

    Wang M, Zhao W, Gao L, Zhao L. 2018. Genome-wide profiling of long non-coding RNAs from tomato and a comparison with mRNAs associated with the regulation of fruit ripening. BMC Plant Biology 18:75

    doi: 10.1186/s12870-018-1300-y

    CrossRef   Google Scholar

    [23]

    Chen L, Zhu QH, Kaufmann K. 2020. Long non-coding RNAs in plants: emerging modulators of gene activity in development and stress responses. Planta 252:92

    doi: 10.1007/s00425-020-03480-5

    CrossRef   Google Scholar

    [24]

    Mehraj H, Shea DJ, Takahashi S, Miyaji N, Akter A, et al. 2021. Genome-wide analysis of long noncoding RNAs, 24-nt siRNAs, DNA methylation and H3K27me3 marks in Brassica rapa. PLoS One 16:e0242530

    doi: 10.1371/journal.pone.0242530

    CrossRef   Google Scholar

    [25]

    Deng P, Liu S, Nie X, Song W, Wu L. 2018. Conservation analysis of long non-coding RNAs in plants. Science China Life Sciences 61:190−98

    doi: 10.1007/s11427-017-9174-9

    CrossRef   Google Scholar

    [26]

    van Dijk EL, Auger H, Jaszczyszyn Y, Thermes C. 2014. Ten years of next-generation sequencing technology. Trends in Genetics 30:418−26

    doi: 10.1016/j.tig.2014.07.001

    CrossRef   Google Scholar

    [27]

    Ilott NE, Ponting CP. 2013. Predicting long non-coding RNAs using RNA sequencing. Methods 63:50−59

    doi: 10.1016/j.ymeth.2013.03.019

    CrossRef   Google Scholar

    [28]

    Steijger T, Abril JF, Engström PG, Kokocinski F, Hubbard TJ, et al. 2013. Assessment of transcript reconstruction methods for RNA-seq. Nature Methods 10:1177−84

    doi: 10.1038/nmeth.2714

    CrossRef   Google Scholar

    [29]

    Roberts RJ, Carneiro MO, Schatz MC. 2013. The advantages of SMRT sequencing. Genome Biology 14:405

    doi: 10.1186/gb-2013-14-6-405

    CrossRef   Google Scholar

    [30]

    van Dijk EL, Jaszczyszyn Y, Naquin D, Thermes C. 2018. The third revolution in sequencing technology. Trends in Genetics 34:666−81

    doi: 10.1016/j.tig.2018.05.008

    CrossRef   Google Scholar

    [31]

    Wang B, Tseng E, Regulski M, Clark TA, Hon T, et al. 2016. Unveiling the complexity of the maize transcriptome by single-molecule long-read sequencing. Nature Communications 7:11708

    doi: 10.1038/ncomms11708

    CrossRef   Google Scholar

    [32]

    Dong L, Liu H, Zhang J, Yang S, Kong G, et al. 2015. Single-molecule real-time transcript sequencing facilitates common wheat genome annotation and grain transcriptome research. BMC Genomics 16:1039

    doi: 10.1186/s12864-015-2257-y

    CrossRef   Google Scholar

    [33]

    Zhu FY, Chen MX, Ye NH, Shi L, Ma KL, et al. 2017. Proteogenomic analysis reveals alternative splicing and translation as part of the abscisic acid response in Arabidopsis seedlings. The Plant Journal 91:518−33

    doi: 10.1111/tpj.13571

    CrossRef   Google Scholar

    [34]

    Tan C, Liu H, Ren J, Ye X, Feng H, et al. 2019. Single-molecule real-time sequencing facilitates the analysis of transcripts and splice isoforms of anthers in Chinese cabbage (Brassica rapa L. ssp. pekinensis). BMC Plant Biol 19:517

    doi: 10.1186/s12870-019-2133-z

    CrossRef   Google Scholar

    [35]

    Clavijo BJ, Venturini L, Schudoma C, Accinelli GG, Kaithakottil G, et al. 2017. An improved assembly and annotation of the allohexaploid wheat genome identifies complete families of agronomic genes and provides genomic evidence for chromosomal translocations. Genome Research 27:885−96

    doi: 10.1101/gr.217117.116

    CrossRef   Google Scholar

    [36]

    Zhang L, Hu J, Han X, Li J, Gao Y, et al. 2019. A high-quality apple genome assembly reveals the association of a retrotransposon and red fruit colour. Nature Communications 10:1494

    doi: 10.1038/s41467-019-09518-x

    CrossRef   Google Scholar

    [37]

    Li N, Meng Z, Tao M, Wang Y, Zhang Y, et al. 2020. Comparative transcriptome analysis of male and female flowers in Spinacia oleracea L. BMC Genomics 21:850

    doi: 10.1186/s12864-020-07277-4

    CrossRef   Google Scholar

    [38]

    Yu T, Tzeng DTW, Li R, Chen J, Zhong S, et al. 2019. Genome-wide identification of long non-coding RNA targets of the tomato MADS box transcription factor RIN and function analysis. Annals of Botany 123:469−82

    doi: 10.1093/aob/mcy178

    CrossRef   Google Scholar

    [39]

    Song J, Cao J, Yu X, Xiang X. 2007. BcMF11, a putative pollen-specific non-coding RNA from Brassica campestris ssp chinensis. Journal of Plant Physiology 164:1097−100

    doi: 10.1016/j.jplph.2006.10.002

    CrossRef   Google Scholar

    [40]

    Song JH, Cao JS, Wang CG. 2013. BcMF11, a novel non-coding RNA gene from Brassica campestris, is required for pollen development and male fertility. Plant Cell Reports 32:21−30

    doi: 10.1007/s00299-012-1337-6

    CrossRef   Google Scholar

    [41]

    Zuo J, Grierson D, Courtney LT, Wang Y, Gao L, et al. 2020. Relationships between genome methylation, levels of non-coding RNAs, mRNAs and metabolites in ripening tomato fruit. The Plant Journal 103:980−94

    doi: 10.1111/tpj.14778

    CrossRef   Google Scholar

    [42]

    Liu T, Wu P, Wang Q, Wang W, Zhang C, et al. 2018. Comparative transcriptome discovery and elucidation of the mechanism of long noncoding RNAs during vernalization in Brassica rapa. Plant Growth Regulation 85:27−39

    doi: 10.1007/s10725-018-0371-y

    CrossRef   Google Scholar

    [43]

    Huang L, Dong H, Zhou D, Li M, Liu Y, et al. 2018. Systematic identification of long non-coding RNAs during pollen development and fertilization in Brassica rapa. The Plant Journal 96:203−22

    doi: 10.1111/tpj.14016

    CrossRef   Google Scholar

    [44]

    Zhou Y, Mumtaz MA, Zhang Y, Yang Z, Hao Y, et al. 2022. Response of anthocyanin biosynthesis to light by strand-specific transcriptome and miRNA analysis in Capsicum annuum. BMC Plant Biology 22:79

    doi: 10.1186/s12870-021-03423-6

    CrossRef   Google Scholar

    [45]

    Eom SH, Lee HJ, Lee JH, Wi SH, Kim SK, et al. 2019. Identification and functional prediction of drought-responsive long non-coding RNA in tomato. Agronomy 9:629

    doi: 10.3390/agronomy9100629

    CrossRef   Google Scholar

    [46]

    Wang A, Hu J, Gao C, Chen G, Wang B, et al. 2019. Genome-wide analysis of long non-coding RNAs unveils the regulatory roles in the heat tolerance of Chinese cabbage (Brassica rapa ssp. chinensis). Scientific Reports 9:5002

    doi: 10.1038/s41598-019-41428-2

    CrossRef   Google Scholar

    [47]

    Wang Y, Gao L, Zhu B, Zhu H, Luo Y, et al. 2018. Integrative analysis of long non-coding RNA acting as ceRNAs involved in chilling injury in tomato fruit. Gene 667:25−33

    doi: 10.1016/j.gene.2018.05.030

    CrossRef   Google Scholar

    [48]

    Li N, Wang Z, Wang B, Wang J, Xu R, et al. 2022. Identification and characterization of long non-coding RNA in tomato roots under salt stress. Frontiers in Plant Science 13:834027

    doi: 10.3389/fpls.2022.834027

    CrossRef   Google Scholar

    [49]

    Cao W, Gan L, Wang C, Zhao X, Zhang M, et al. 2021. Genome-wide identification and characterization of potato long non-coding RNAs associated with Phytophthora infestans resistance. Frontiers in Plant Science 12:619062

    doi: 10.3389/fpls.2021.619062

    CrossRef   Google Scholar

    [50]

    Wang J, Yu W, Yang Y, Li X, Chen T, et al. 2015. Genome-wide analysis of tomato long non-coding RNAs and identification as endogenous target mimic for microRNA in response to TYLCV infection. Scientific Reports 5:16946

    doi: 10.1038/srep16946

    CrossRef   Google Scholar

    [51]

    Zhou C, Zhu J, Qian N, Guo J, Yan C. 2020. Bacillus subtilis SL18r Induces Tomato Resistance Against Botrytis cinerea, Involving Activation of Long Non-coding RNA, MSTRG18363, to Decoy miR1918. Frontiers in Plant Science 11:634819

    doi: 10.3389/fpls.2020.634819

    CrossRef   Google Scholar

    [52]

    Rosli HG, Sirvent E, Bekier FN, Ramos RN, Pombo MA. 2021. Genome-wide analysis uncovers tomato leaf lncRNAs transcriptionally active upon Pseudomonas syringae pv. tomato challenge. Scientific Reports 11:24523

    doi: 10.1038/s41598-021-04005-0

    CrossRef   Google Scholar

    [53]

    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

    [54]

    Meng X, Li A, Yu B, Li S. 2021. Interplay between miRNAs and lncRNAs: Mode of action and biological roles in plant development and stress adaptation. Computational and Structural Biotechnology Journal 19:2567−74

    doi: 10.1016/j.csbj.2021.04.062

    CrossRef   Google Scholar

    [55]

    Ameres SL, Horwich MD, Hung JH, Xu J, Ghildiyal M, et al. 2010. Target RNA-directed trimming and tailing of small silencing RNAs. Science 328:1534−39

    doi: 10.1126/science.1187058

    CrossRef   Google Scholar

    [56]

    Baccarini A, Chauhan H, Gardner TJ, Jayaprakash AD, Sachidanandam R, Brown BD. 2011. Kinetic analysis reveals the fate of a microRNA following target regulation in mammalian cells. Current Biology 21:369−76

    doi: 10.1016/j.cub.2011.01.067

    CrossRef   Google Scholar

    [57]

    Xie J, Ameres SL, Friedline R, Hung JH, Zhang Y, et al. 2012. Long-term, efficient inhibition of microRNA function in mice using rAAV vectors. Nature Methods 9:403−9

    doi: 10.1038/nmeth.1903

    CrossRef   Google Scholar

    [58]

    Yan J, Gu Y, Jia X, Kang W, Pan S, et al. 2012. Effective small RNA destruction by the expression of a short tandem target mimic in Arabidopsis. The Plant Cell 24:415−27

    doi: 10.1105/tpc.111.094144

    CrossRef   Google Scholar

    [59]

    Mei J, Jiang N, Ren G. 2019. The F-box protein HAWAIIAN SKIRT is required for mimicry target-induced microRNA degradation in Arabidopsis. Journal of Integrative Plant Biology 61:1121−27

    doi: 10.1111/jipb.12761

    CrossRef   Google Scholar

    [60]

    Wang J, Feng Y, Ding X, Huo J, Nie W. 2021. Identification of Long Non-Coding RNAs Associated with Tomato Fruit Expansion and Ripening by Strand-Specific Paired-End RNA Sequencing. Horticulturae 7:522

    doi: 10.3390/horticulturae7120522

    CrossRef   Google Scholar

    [61]

    Zhu B, Yang Y, Li R, Fu D, Wen L, et al. 2015. RNA sequencing and functional analysis implicate the regulatory role of long non-coding RNAs in tomato fruit ripening. Journal of Experimental Botany 66:4483−95

    doi: 10.1093/jxb/erv203

    CrossRef   Google Scholar

    [62]

    Li R, Fu D, Zhu B, Luo Y, Zhu H. 2018. CRISPR/Cas9-mediated mutagenesis of lncRNA1459 alters tomato fruit ripening. The Plant Journal 94:513−24

    doi: 10.1111/tpj.13872

    CrossRef   Google Scholar

    [63]

    Xiao Y, Kang B, Li M, Xiao L, Xiao H, et al. 2020. Transcription of lncRNA ACoS-AS1 is essential to trans-splicing between SlPsy1 and ACoS-AS1 that causes yellow fruit in tomato. RNA Biology 17:596−607

    doi: 10.1080/15476286.2020.1721095

    CrossRef   Google Scholar

    [64]

    Yang S, Yang T, Tang Y, Aisimutuola P, Zhang G, et al. 2020. Transcriptomic profile analysis of non-coding RNAs involved in Capsicum chinense Jacq. fruit ripening. Scientia Horticulturae 264:109158

    doi: 10.1016/j.scienta.2019.109158

    CrossRef   Google Scholar

    [65]

    Zuo J, Wang Y, Zhu B, Luo Y, Wang Q, et al. 2019. Network analysis of noncoding RNAs in pepper provides insights into fruit ripening control. Scientific Reports 9:8734

    doi: 10.1038/s41598-019-45427-1

    CrossRef   Google Scholar

    [66]

    Ou L, Liu Z, Zhang Z, Wei G, Zhang Y, et al. 2017. Noncoding and coding transcriptome analysis reveals the regulation roles of long noncoding RNAs in fruit development of hot pepper (Capsicum annuum L.). Plant Growth Regulation 83:141−56

    doi: 10.1007/s10725-017-0290-3

    CrossRef   Google Scholar

    [67]

    Bouché F, Woods DP, Amasino RM. 2017. Winter memory throughout the plant kingdom: different paths to flowering. Plant Physiology 173:27−35

    doi: 10.1104/pp.16.01322

    CrossRef   Google Scholar

    [68]

    Shea DJ, Nishida N, Takada S, Itabashi E, Takahashi S, et al. 2019. Long noncoding RNAs in Brassica rapa L. following vernalization. Scientific Reports 9:9302

    doi: 10.1038/s41598-019-45650-w

    CrossRef   Google Scholar

    [69]

    Shi F, Xu H, Liu C, Tan C, Ren J, et al. 2021. Whole-transcriptome sequencing reveals a vernalization-related ceRNA regulatory network in chinese cabbage (Brassica campestris L. ssp. pekinensis). BMC Genomics 22:819

    doi: 10.1186/s12864-021-08110-2

    CrossRef   Google Scholar

    [70]

    Mayr E. 1986. Joseph Gottlieb Kolreuter's contributions to biology. Osiris 2:135−76

    doi: 10.1086/368655

    CrossRef   Google Scholar

    [71]

    Chen L, Liu YG. 2014. Male sterility and fertility restoration in crops. Annual Review of Plant Biology 65:579−606

    doi: 10.1146/annurev-arplant-050213-040119

    CrossRef   Google Scholar

    [72]

    Shi F, Pang Z, Liu C, Zhou L, Tan C, et al. 2022. Whole-transcriptome analysis and construction of an anther development-related ceRNA network in Chinese cabbage (Brassica campestris L. ssp. pekinensis). Scientific Reports 12:2667

    doi: 10.1038/s41598-022-06556-2

    CrossRef   Google Scholar

    [73]

    Zhou D, Chen C, Jin Z, Chen J, Lin S, et al. 2022. Transcript profiling analysis and ncRNAs' identification of male-sterile systems of Brassica campestris reveal new insights into the mechanism underlying anther and pollen development. Frontiers in Plant Science 13:806865

    doi: 10.3389/fpls.2022.806865

    CrossRef   Google Scholar

    [74]

    Li P, Zhang D, Su T, Wang W, Yu Y, et al. 2020. Genome-wide analysis of mRNA and lncRNA expression and mitochondrial genome sequencing provide insights into the mechanisms underlying a novel cytoplasmic male sterility system, BVRC-CMS96, in Brassicarapa. Theoretical and Applied Genetics 133:2157−70

    doi: 10.1007/s00122-020-03587-z

    CrossRef   Google Scholar

    [75]

    Lv J, Liu Z, Yang B, Deng M, Wang J, et al. 2020. Systematic identification and characterization of long non-coding RNAs involved in cytoplasmic male sterility in pepper (Capsicum annuum L.). Plant Growth Regulation 91:277−88

    doi: 10.1007/s10725-020-00605-4

    CrossRef   Google Scholar

    [76]

    Liu L, Lu Y, Wei L, Yu H, Cao Y, et al. 2018. Transcriptomics analyses reveal the molecular roadmap and long non-coding RNA landscape of sperm cell lineage development. The Plant Journal 96:421−37

    doi: 10.1111/tpj.14041

    CrossRef   Google Scholar

    [77]

    He J, Giusti MM. 2010. Anthocyanins: natural colorants with health-promoting properties. Annual Review of Food Science and Technology 1:163−87

    doi: 10.1146/annurev.food.080708.100754

    CrossRef   Google Scholar

    [78]

    Mattioli R, Francioso A, Mosca L, Silva P. 2020. Anthocyanins: A comprehensive review of their chemical properties and health effects on cardiovascular and neurodegenerative diseases. Molecules 25:3809

    doi: 10.3390/molecules25173809

    CrossRef   Google Scholar

    [79]

    Jaakola L. 2013. New insights into the regulation of anthocyanin biosynthesis in fruits. Trends in Plant Science 18:477−83

    doi: 10.1016/j.tplants.2013.06.003

    CrossRef   Google Scholar

    [80]

    Tang R, Dong H, Wu W, Zhao C, Jia X, et al. 2021. A comparative transcriptome analysis of purple and yellow fleshed potato tubers reveals long non-coding RNAs and their targets functioned in anthocyanin biosynthesis. Preprint

    doi: 10.21203/rs.3.rs-515121/v1

    CrossRef   Google Scholar

    [81]

    Bao Y, Nie T, Wang D, Chen Q. 2022. Anthocyanin regulatory networks in Solanum tuberosum L. leaves elucidated via integrated metabolomics, transcriptomics, and StAN1 overexpression. BMC Plant Biology 22:1

    doi: 10.1186/s12870-021-03391-x

    CrossRef   Google Scholar

    [82]

    Chialva C, Blein T, Crespi M, Lijavetzky D. 2021. Insights into long non-coding RNA regulation of anthocyanin carrot root pigmentation. Scientific Reports 11:4093

    doi: 10.1038/s41598-021-83514-4

    CrossRef   Google Scholar

    [83]

    Liao X, Wang J, Zhu S, Xie Q, Wang L, et al. 2020. Transcriptomic and functional analyses uncover the regulatory role of lncRNA000170 in tomato multicellular trichome formation. The Plant Journal 104:18−29

    doi: 10.1111/tpj.14902

    CrossRef   Google Scholar

    [84]

    Sonnewald S, Sonnewald U. 2014. Regulation of potato tuber sprouting. Planta 239:27−38

    doi: 10.1007/s00425-013-1968-z

    CrossRef   Google Scholar

    [85]

    Kolachevskaya OO, Lomin SN, Arkhipov DV, Romanov GA. 2019. Auxins in potato: molecular aspects and emerging roles in tuber formation and stress resistance. Plant Cell Reports 38:681−98

    doi: 10.1007/s00299-019-02395-0

    CrossRef   Google Scholar

    [86]

    Li L, Deng M, Lyu C, Zhang J, Peng J, et al. 2020. Quantitative phosphoproteomics analysis reveals that protein modification and sugar metabolism contribute to sprouting in potato after BR treatment. Food Chemistry 325:126875

    doi: 10.1016/j.foodchem.2020.126875

    CrossRef   Google Scholar

    [87]

    Hou X, Du Y, Liu X, Zhang H, Liu Y, et al. 2017. Genome-wide analysis of long non-coding RNAs in potato and their potential role in tuber sprouting process. International Journal of Molecular Sciences 19:101

    doi: 10.3390/ijms19010101

    CrossRef   Google Scholar

    [88]

    Ramírez Gonzales L, Shi L, Bergonzi SB, Oortwijn M, Franco-Zorrilla JM, et al. 2021. Potato CYCLING DOF FACTOR 1 and its lncRNA counterpart StFLORE link tuber development and drought response. The Plant Journal 105:855−69

    doi: 10.1111/tpj.15093

    CrossRef   Google Scholar

    [89]

    Ghorbani F, Abolghasemi R, Haghighi M, Etemadi N, Wang S, et al. 2021. Global identification of long non-coding RNAs involved in the induction of spinach flowering. BMC Genomics 22:704

    doi: 10.1186/s12864-021-07989-1

    CrossRef   Google Scholar

    [90]

    Shen E, Zhu X, Hua S, Chen H, Ye C, et al. 2018. Genome-wide identification of oil biosynthesis-related long non-coding RNAs in allopolyploid Brassica napus. BMC Genomics 19:745

    doi: 10.1186/s12864-018-5117-8

    CrossRef   Google Scholar

    [91]

    Zhu X, Tai X, Ren Y, Chen J, Bo T. 2019. Genome-wide analysis of coding and long non-coding RNAs involved in cuticular wax biosynthesis in cabbage (Brassica oleracea L. var. capitata). International Journal of Molecular Sciences 20:2820

    doi: 10.3390/ijms20112820

    CrossRef   Google Scholar

    [92]

    Zhu J. 2016. Abiotic Stress Signaling and Responses in Plants. Cell 167:313−24

    doi: 10.1016/j.cell.2016.08.029

    CrossRef   Google Scholar

    [93]

    Zandalinas SI, Mittler R. 2022. Plant responses to multifactorial stress combination. New Phytologist 234:1161−67

    doi: 10.1111/nph.18087

    CrossRef   Google Scholar

    [94]

    Markham KK, Greenham K. 2021. Abiotic stress through time. New Phytologist 231:40−46

    doi: 10.1111/nph.17367

    CrossRef   Google Scholar

    [95]

    Waititu JK, Zhang C, Liu J, Wang H. 2020. Plant non-coding RNAs: origin, biogenesis, mode of action and their roles in abiotic stress. International Journal of Molecular Sciences 21:8401

    doi: 10.3390/ijms21218401

    CrossRef   Google Scholar

    [96]

    Li Y, Li X, Yang J, He Y. 2020. Natural antisense transcripts of MIR398 genes suppress microR398 processing and attenuate plant thermotolerance. Nature Communications 11:5351

    doi: 10.1038/s41467-020-19186-x

    CrossRef   Google Scholar

    [97]

    Di C, Yuan J, Wu Y, Li J, Lin H, et al. 2014. Characterization of stress-responsive lncRNAs in Arabidopsis thaliana by integrating expression, epigenetic and structural features. The Plant Journal 80:848−61

    doi: 10.1111/tpj.12679

    CrossRef   Google Scholar

    [98]

    Kindgren P, Ard R, Ivanov M, Marquardt S. 2018. Transcriptional read-through of the long non-coding RNA SVALKA governs plant cold acclimation. Nature Communications 9:4561

    doi: 10.1038/s41467-018-07010-6

    CrossRef   Google Scholar

    [99]

    Qin T, Zhao H, Cui P, Albesher N, Xiong L. 2017. A nucleus-localized long non-coding RNA enhances drought and salt stress tolerance. Plant Physiology 175:1321−36

    doi: 10.1104/pp.17.00574

    CrossRef   Google Scholar

    [100]

    Ullah A, Sun H, Yang X, Zhang X. 2017. Drought coping strategies in cotton: increased crop per drop. Plant Biotechnology Journal 15:271−84

    doi: 10.1111/pbi.12688

    CrossRef   Google Scholar

    [101]

    Yang SJ, Vanderbeld B, Wan JX, Huang YF. 2010. Narrowing Down the Targets: Towards Successful Genetic Engineering of Drought-Tolerant Crops. Molecular Plant 3:469−90

    doi: 10.1093/mp/ssq016

    CrossRef   Google Scholar

    [102]

    McDowell N, Pockman WT, Allen CD, Breshears DD, Cobb N, et al. 2008. Mechanisms of plant survival and mortality during drought: why do some plants survive while others succumb to drought. New Phytologist 178:719−39

    doi: 10.1111/j.1469-8137.2008.02436.x

    CrossRef   Google Scholar

    [103]

    Lamin-Samu AT, Zhuo S, Ali M, Lu G. 2022. Long non-coding RNA transcriptome landscape of anthers at different developmental stages in response to drought stress in tomato. Genomics 114:110383

    doi: 10.1016/j.ygeno.2022.110383

    CrossRef   Google Scholar

    [104]

    Tan X, Li S, Hu L, Zhang C. 2020. Genome-wide analysis of long non-coding RNAs (lncRNAs) in two contrasting rapeseed (Brassica napus L.) genotypes subjected to drought stress and re-watering. BMC Plant Biol 20:81

    doi: 10.1186/s12870-020-2286-9

    CrossRef   Google Scholar

    [105]

    Jian H, Sun H, Liu R, Zhang W, Shang L, et al. 2022. Construction of drought stress regulation networks in potato based on SMRT and RNA sequencing data. Preprint

    doi: 10.21203/rs.3.rs-1456188/v1

    CrossRef   Google Scholar

    [106]

    Eom SH, Lee HJ, Wi SH, Kim SK, Hyun TK. 2021. Identification and functional prediction of long non-coding RNAs responsive to heat stress in heading type Chinese cabbage. Zemdirbyste-Agriculture 108:371−76

    doi: 10.13080/z-a.2021.108.047

    CrossRef   Google Scholar

    [107]

    Bhatia G, Singh A, Verma D, Sharma S, Singh K. 2020. Genome-wide investigation of regulatory roles of lncRNAs in response to heat and drought stress in Brassica juncea (Indian mustard). Environmental and Experimental Botany 171:103922

    doi: 10.1016/j.envexpbot.2019.103922

    CrossRef   Google Scholar

    [108]

    Song X, Liu G, Huang Z, Duan W, Tan H, et al. 2016. Temperature expression patterns of genes and their coexpression with LncRNAs revealed by RNA-Seq in non-heading Chinese cabbage. BMC Genomics 17:297

    doi: 10.1186/s12864-016-2625-2

    CrossRef   Google Scholar

    [109]

    He X, Guo S, Wang Y, Wang L, Shu S, et al. 2020. Systematic identification and analysis of heat-stress-responsive lncRNAs, circRNAs and miRNAs with associated co-expression and ceRNA networks in cucumber (Cucumis sativus L.). Physiologia Plantarum 168:736−54

    doi: 10.1111/ppl.12997

    CrossRef   Google Scholar

    [110]

    Yang Z, Li W, Su X, Ge P, Zhou Y, et al. 2019. Early Response of Radish to Heat Stress by Strand-Specific Transcriptome and miRNA Analysis. Int J Mol Sci 20:3321

    doi: 10.3390/ijms20133321

    CrossRef   Google Scholar

    [111]

    Zuo J, Wang Y, Zhu B, Luo Y, Wang Q, et al. 2018. Analysis of the coding and non-coding RNA transcriptomes in response to bell pepper chilling. International Journal of Molecular Sciences 19:2001

    doi: 10.3390/ijms19072001

    CrossRef   Google Scholar

    [112]

    Xue L, Sun M, Wu Z, Yu L, Yu Q, et al. 2020. LncRNA regulates tomato fruit cracking by coordinating gene expression via a hormone-redox-cell wall network. BMC Plant Biology 20:162

    doi: 10.1186/s12870-020-02373-9

    CrossRef   Google Scholar

    [113]

    Crawford RMM. 1992. Oxygen availability as an ecological limit to plant distribution. Advances in Ecological Research 23:93−185

    doi: 10.1016/S0065-2504(08)60147-6

    CrossRef   Google Scholar

    [114]

    Sairam RK, Kumutha D, Ezhilmathi K, Deshmukh PS, Srivastava GC. 2008. Physiology and biochemistry of waterlogging tolerance in plants. Biologia Plantarum 52:401−12

    doi: 10.1007/s10535-008-0084-6

    CrossRef   Google Scholar

    [115]

    Kęska K, Szcześniak MW, Adamus A, Czernicka M. 2021. Waterlogging-stress-responsive LncRNAs, their regulatory relationships with miRNAs and target genes in cucumber (Cucumis sativus L.). International Journal of Molecular Sciences 22:8197

    doi: 10.3390/ijms22158197

    CrossRef   Google Scholar

    [116]

    Meharg A. 2012. Marschner's Mineral Nutrition of Higher Plants. Ed. Marschner P. Third Edition. Amsterdam, Netherlands: Academic Press, Elsevier. 684 pp

    [117]

    Schroeder JI, Delhaize E, Frommer WB, Guerinot ML, Harrison MJ, et al. 2013. Using membrane transporters to improve crops for sustainable food production. Nature 497:60−66

    doi: 10.1038/nature11909

    CrossRef   Google Scholar

    [118]

    Zhang Z, Liao H, Lucas WJ. 2014. Molecular mechanisms underlying phosphate sensing, signaling, and adaptation in plants. Journal of Integrative Plant Biology 56:192−220

    doi: 10.1111/jipb.12163

    CrossRef   Google Scholar

    [119]

    Puga MI, Rojas-Triana M, de Lorenzo L, Leyva A, Rubio V, et al. 2017. Novel signals in the regulation of Pi starvation responses in plants: facts and promises. Current Opinion in Biotechnology 39:40−49

    doi: 10.1016/j.pbi.2017.05.007

    CrossRef   Google Scholar

    [120]

    Ham BK, Chen J, Yan Y, Lucas WJ. 2018. Insights into plant phosphate sensing and signaling. Current Opinion in Plant Biology 49:1−9

    doi: 10.1016/j.copbio.2017.07.005

    CrossRef   Google Scholar

    [121]

    Zhang Z, Zheng Y, Ham BK, Zhang S, Fei Z, et al. 2019. Plant lncRNAs are enriched in and move systemically through the phloem in response to phosphate deficiency. Journal of Integrative Plant Biology 61:492−508

    doi: 10.1111/jipb.12715

    CrossRef   Google Scholar

    [122]

    Genchi G, Sinicropi MS, Lauria G, Carocci A, Catalano A. 2020. The effects of cadmium toxicity. International Journal of Environmental Research and Public Health 17:3782

    doi: 10.3390/ijerph17113782

    CrossRef   Google Scholar

    [123]

    Tchounwou PB, Yedjou CG, Patlolla AK, Sutton DJ. 2012. Heavy metal toxicity and the environment. In: Molecular, Clinical and Environmental Toxicology. Experientia Supplementum, ed. Luch, A. Vol 101. Basel: Springer. pp 133–64. https://doi.org/10.1007/978-3-7643-8340-4_6

    [124]

    Sanità di Toppi L, Gabbrielli R. 1999. Response to cadmium in higher plants. Environmental and Experimental Botany 41:105−30

    doi: 10.1016/S0098-8472(98)00058-6

    CrossRef   Google Scholar

    [125]

    Huybrechts M, Cuypers A, Deckers J, Iven V, Vandionant S, et al. 2019. Cadmium and plant development: An agony from seed to seed. International Journal of Molecular Sciences 20:3971

    doi: 10.3390/ijms20163971

    CrossRef   Google Scholar

    [126]

    Feng S, Zhang X, Liu X, Tan S, Chu S, et al. 2016. Characterization of long non-coding RNAs involved in cadmium toxic response in Brassica napus. RSC Advances 6:82157−73

    doi: 10.1039/C6RA05459E

    CrossRef   Google Scholar

    [127]

    Chen L, Wang Q, Zhou L, Ren F, Li D, et al. 2013. Arabidopsis CBL-interacting protein kinase (CIPK6) is involved in plant response to salt/osmotic stress and ABA. Molecular Biology Reports 40:4759−67

    doi: 10.1007/s11033-013-2572-9

    CrossRef   Google Scholar

    [128]

    Takemoto D, Shibata Y, Ojika M, Mizuno Y, Imano S, et al. 2018. Resistance to Phytophthora infestans: exploring genes required for disease resistance in Solanaceae plants. Journal of General Plant Pathology 84:312−20

    doi: 10.1007/s10327-018-0801-8

    CrossRef   Google Scholar

    [129]

    Ivanov AA, Ukladov EO, Golubeva TS. 2021. Phytophthora infestans: An overview of methods and attempts to combat late blight. Journal of Fungi 7:1071

    doi: 10.3390/jof7121071

    CrossRef   Google Scholar

    [130]

    Cui J, Luan Y, Jiang N, Bao H, Meng J. 2017. Comparative transcriptome analysis between resistant and susceptible tomato allows the identification of lncRNA16397 conferring resistance to Phytophthora infestans by co-expressing glutaredoxin. The Plant Journal 89:577−89

    doi: 10.1111/tpj.13408

    CrossRef   Google Scholar

    [131]

    Cui J, Jiang N, Hou X, Wu S, Zhang Q, et al. 2020. Genome-wide identification of lncRNAs and analysis of ceRNA networks during tomato resistance to Phytophthora infestans. Phytopathology 110:456−64

    doi: 10.1094/phyto-04-19-0137-r

    CrossRef   Google Scholar

    [132]

    Cui J, Jiang N, Meng J, Yang G, Liu W, et al. 2019. LncRNA33732-respiratory burst oxidase module associated with WRKY1 in tomato-Phytophthora infestans interactions. The Plant Journal 97:933−46

    doi: 10.1111/tpj.14173

    CrossRef   Google Scholar

    [133]

    Jiang N, Cui J, Hou X, Yang G, Xiao Y, et al. 2020. Sl-lncRNA15492 interacts with Sl-miR482a and affects Solanum lycopersicum immunity against Phytophthora infestans. The Plant Journal 103:1561−74

    doi: 10.1111/tpj.14847

    CrossRef   Google Scholar

    [134]

    Liu W, Cui J, Luan Y. 2022. Overexpression of lncRNA08489 enhances tomato immunity against Phytophthora infestans by decoying miR482e-3p. Biochemical and Biophysical Research Communications 587:36−41

    doi: 10.1016/j.bbrc.2021.11.079

    CrossRef   Google Scholar

    [135]

    Jiang N, Cui J, Shi Y, Yang G, Zhou X, et al. 2019. Tomato lncRNA23468 functions as a competing endogenous RNA to modulate NBS-LRR genes by decoying miR482b in the tomato-Phytophthora infestans interaction. Horticulture Research 6:28

    doi: 10.1038/s41438-018-0096-0

    CrossRef   Google Scholar

    [136]

    Hou X, Cui J, Liu W, Jiang N, Zhou X, et al. 2020. LncRNA39026 enhances tomato resistance to Phytophthora infestans by decoying miR168a and inducing PR gene expression. Phytopathology 110:873−80

    doi: 10.1094/PHYTO-12-19-0445-R

    CrossRef   Google Scholar

    [137]

    Zhang Y, Hong Y, Liu Y, Cui J, Luan Y. 2021. Function identification of miR394 in tomato resistance to Phytophthora infestans. Plant Cell Reports 40:1831−44

    doi: 10.1007/s00299-021-02746-w

    CrossRef   Google Scholar

    [138]

    Wang J, Yang Y, Jin L, Ling X, Liu T, et al. 2018. Re-analysis of long non-coding RNAs and prediction of circRNAs reveal their novel roles in susceptible tomato following TYLCV infection. BMC Plant Biology 18:104

    doi: 10.1186/s12870-018-1332-3

    CrossRef   Google Scholar

    [139]

    Yang Y, Liu T, Shen D, Wang J, Ling X, et al. 2019. Tomato yellow leaf curl virus intergenic siRNAs target a host long noncoding RNA to modulate disease symptoms. PLoS Pathogens 15:e1007534

    doi: 10.1371/journal.ppat.1007534

    CrossRef   Google Scholar

    [140]

    Yang F, Zhao D, Fan H, Zhu X, Wang Y, et al. 2020. Functional analysis of long non-coding RNAs reveal their novel roles in biocontrol of bacteria-induced tomato resistance to Meloidogyne incognita. International Journal of Molecular Sciences 21:911

    doi: 10.3390/ijms21030911

    CrossRef   Google Scholar

    [141]

    Zheng Y, Wang Y, Ding B, Fei Z. 2017. Comprehensive transcriptome analyses reveal that potato spindle tuber viroid triggers genome-wide changes in alternative splicing, inducible trans-acting activity of phased secondary small interfering RNAs, and immune responses. Journal of Virology 91:e00247-17

    doi: 10.1128/jvi.00247-17

    CrossRef   Google Scholar

    [142]

    Joshi RK, Megha S, Basu U, Rahman MH, Kav NN. 2016. Genome wide identification and functional prediction of long non-coding RNAs responsive to Sclerotinia sclerotiorum infection in Brassica napus. PLoS One 11:e0158784

    doi: 10.1371/journal.pone.0158784

    CrossRef   Google Scholar

    [143]

    Zhu H, Li X, Xi D, Zhai W, Zhang Z, et al. 2019. Integrating long noncoding RNAs and mRNAs expression profiles of response to Plasmodiophora brassicae infection in Pakchoi (Brassica campestris ssp. chinensis Makino). PLoS One 14:e0224927

    doi: 10.1371/journal.pone.0224927

    CrossRef   Google Scholar

    [144]

    Summanwar A, Basu U, Rahman H, Kav N. 2019. Identification of lncRNAs responsive to infection by Plasmodiophora brassicae in clubroot-susceptible and -resistant Brassica napus lines carrying resistance introgressed from rutabaga. Molecular Plant-Microbe Interactions 32:1360−77

    doi: 10.1094/MPMI-12-18-0341-R

    CrossRef   Google Scholar

    [145]

    Zhang B, Su T, Li P, Xin X, Cao Y, et al. 2021. Identification of long noncoding RNAs involved in resistance to downy mildew in Chinese cabbage. Horticulture Research 8:44

    doi: 10.1038/s41438-021-00479-1

    CrossRef   Google Scholar

    [146]

    Akter MA, Mehraj H, Miyaji N, Takahashi S, Takasaki-Yasuda T, et al. 2022. Transcriptional association between mRNAs and their paired natural antisense transcripts following Fusarium oxysporum inoculation in Brassica rapa L. Horticulturae 8:17

    doi: 10.3390/horticulturae8010017

    CrossRef   Google Scholar

    [147]

    Yin J, Yan J, Hou L, Jiang L, Xian W, et al. 2021. Identification and functional deciphering suggested the regulatory roles of long intergenic ncRNAs (lincRNAs) in increasing grafting pepper resistance to Phytophthora capsici. BMC Genomics 22:868

    doi: 10.1186/s12864-021-08183-z

    CrossRef   Google Scholar

    [148]

    Kwenda S, Birch PRJ, Moleleki LN. 2016. Genome-wide identification of potato long intergenic noncoding RNAs responsive to Pectobacterium carotovorum subspecies brasiliense infection. BMC Genomics 17:614

    doi: 10.1186/s12864-016-2967-9

    CrossRef   Google Scholar

    [149]

    Glushkevich A, Spechenkova N, Fesenko I, Knyazev A, Samarskaya V, et al. 2022. Transcriptomic reprogramming, alternative splicing and RNA methylation in potato (Solanum tuberosum L.) plants in response to potato virus Y infection. Plants 11:635

    doi: 10.3390/plants11050635

    CrossRef   Google Scholar

    [150]

    Nie J, Wang H, Zhang W, Teng X, Yu C, et al. 2021. Characterization of lncRNAs and mRNAs involved in powdery mildew resistance in cucumber. Phytopathology 111:1613−24

    doi: 10.1094/PHYTO-11-20-0521-R

    CrossRef   Google Scholar

    [151]

    Wang Y, Gao L, Li J, Zhu B, Zhu H, et al. 2018. Analysis of long-non-coding RNAs associated with ethylene in tomato. Gene 674:151−60

    doi: 10.1016/j.gene.2018.06.089

    CrossRef   Google Scholar

    [152]

    Wang X, Ai G, Zhang C, Cui L, Wang J, et al. 2016. Expression and diversification analysis reveals transposable elements play important roles in the origin of Lycopersicon-specific lncRNAs in tomato. New Phytologist 209:1442−55

    doi: 10.1111/nph.13718

    CrossRef   Google Scholar

    [153]

    Zhang J, Wei L, Jiang J, Mason AS, Li H, et al. 2018. Genome-wide identification, putative functionality and interactions between lncRNAs and miRNAs in Brassica species. Scientific Reports 8:4960

    doi: 10.1038/s41598-018-23334-1

    CrossRef   Google Scholar

    [154]

    Shu HY, Zhou H, Mu HL, Wu SH, Jiang YL, et al. 2021. Integrated analysis of mRNA and non-coding RNA transcriptome in pepper (Capsicum chinense) hybrid at seedling and flowering stages. Frontiers in Genetics 12:685788

    doi: 10.3389/fgene.2021.685788

    CrossRef   Google Scholar

    [155]

    Wang P, Yu X, Zhu Z, Zhai Y, Zhao Q, et al. 2020. Global profiling of lncRNAs expression responsive to allopolyploidization in Cucumis. Genes 11:1500

    doi: 10.3390/genes11121500

    CrossRef   Google Scholar

    [156]

    Wang R, Zou J, Meng J, Wang J. 2018. Integrative analysis of genome-wide lncRNA and mRNA expression in newly synthesized Brassica hexaploids. Ecology and Evolution 8:6034−52

    doi: 10.1002/ece3.4152

    CrossRef   Google Scholar

    [157]

    Wang Z, Gerstein M, Snyder M. 2009. RNA-Seq: a revolutionary tool for transcriptomics. Nature Reviews Genetics 10:57−63

    doi: 10.1038/nrg2484

    CrossRef   Google Scholar

    [158]

    Jiang Q, Ma R, Wang J, Wu X, Jin S, et al. 2015. LncRNA2Function: a comprehensive resource for functional investigation of human lncRNAs based on RNA-seq data. BMC Genomics 16:S2

    doi: 10.1186/1471-2164-16-s3-s2

    CrossRef   Google Scholar

    [159]

    Liu J, Wang H, Chua NH. 2015. Long noncoding RNA transcriptome of plants. Plant Biotechnology Journal 13:319−28

    doi: 10.1111/pbi.12336

    CrossRef   Google Scholar

    [160]

    Bai Y, Dai X, Harrison AP, Chen M. 2015. RNA regulatory networks in animals and plants: a long noncoding RNA perspective. Briefings in Functional Genomics 14:91−101

    doi: 10.1093/bfgp/elu017

    CrossRef   Google Scholar

    [161]

    Waseem M, Liu Y, Xia R. 2020. Long non-coding RNAs, the dark matter: An emerging regulatory component in plants. International Journal of Molecular Sciences 22:86

    doi: 10.3390/ijms22010086

    CrossRef   Google Scholar

    [162]

    Rosenlund IA, Calin GA, Dragomir MP, Knutsen E. 2021. CRISPR/Cas9 to Silence Long Non-Coding RNAs. In Long Non-Coding RNAs in Cancer. Methods in Molecular Biology, ed. Navarro A. New York: Humana. pp. 175−87 https://doi.org/10.1007/978-1-0716-1581-2_12

    [163]

    Zhu S, Li W, Liu J, Chen CH, Liao Q, et al. 2016. Genome-scale deletion screening of human long non-coding RNAs using a paired-guide RNA CRISPR-Cas9 library. Nature Biotechnology 34:1279−86

    doi: 10.1038/nbt.3715

    CrossRef   Google Scholar

    [164]

    Huang W, Li H, Yu Q, Xiao W, Wang DO. 2022. LncRNA-mediated DNA methylation: an emerging mechanism in cancer and beyond. Journal of Experimental & Clinical Cancer Research 41:100

    doi: 10.1186/s13046-022-02319-z

    CrossRef   Google Scholar

    [165]

    Qian X, Zhao J, Yeung PY, Zhang QC, Kwok CK. 2019. Revealing lncRNA structures and interactions by sequencing-based approaches. Trends in Biochemical Sciences 44:33−52

    doi: 10.1016/j.tibs.2018.09.012

    CrossRef   Google Scholar

    [166]

    Ferrè F, Colantoni A, Helmer-Citterich M. 2016. Revealing protein–lncRNA interaction. Briefings in Bioinformatics 17:106−16

    doi: 10.1093/bib/bbv031

    CrossRef   Google Scholar

    [167]

    Xu W, Yang T, Wang B, Han B, Zhou H, et al. 2018. Differential expression networks and inheritance patterns of long non-coding RNAs in castor bean seeds. The Plant Journal 95:324−40

    doi: 10.1111/tpj.13953

    CrossRef   Google Scholar

    [168]

    Liu Y, Ke L, Wu G, Xu Y, Wu X, et al. 2017. miR3954 is a trigger of phasiRNAs that affects flowering time in citrus. The Plant Journal 92:263−75

    doi: 10.1111/tpj.13650

    CrossRef   Google Scholar

    [169]

    Sun Y, Hao P, Lv X, Tian J, Wang Y, et al. 2020. A long non-coding apple RNA, MSTRG. 85814.11, acts as a transcriptional enhancer of SAUR32 and contributes to the Fe-deficiency response. The Plant Journal 103:53−67

    doi: 10.1111/tpj.14706

    CrossRef   Google Scholar

    [170]

    Hu R, Sun X. 2016. lncRNATargets: A platform for lncRNA target prediction based on nucleic acid thermodynamics. Journal of Bioinformatics and Computational Biology 14:1650016

    doi: 10.1142/s0219720016500165

    CrossRef   Google Scholar

    [171]

    Furió-Tarí P, Tarazona S, Gabaldón T, Enright AJ, Conesa A. 2016. spongeScan: A web for detecting microRNA binding elements in lncRNA sequences. Nucleic Acids Research 44:W176−W180

    doi: 10.1093/nar/gkw443

    CrossRef   Google Scholar

    [172]

    Jiang Q, Wang J, Wang Y, Ma R, Wu X, Li Y. 2014. TF2LncRNA: identifying common transcription factors for a list of lncRNA genes from ChIP-Seq data. Biomed Research International 2014:317642

    doi: 10.1155/2014/317642

    CrossRef   Google Scholar

    [173]

    Huang HY, Chien CH, Jen KH, Huang HD. 2006. RegRNA: an integrated web server for identifying regulatory RNA motifs and elements. Nucleic Acids Research 34:W429−W434

    doi: 10.1093/nar/gkl333

    CrossRef   Google Scholar

    [174]

    Graf J, Kretz M. 2020. From structure to function: Route to understanding lncRNA mechanism. BioEssays 42:2000027

    doi: 10.1002/bies.202000027

    CrossRef   Google Scholar

    [175]

    Diederichs S. 2014. The four dimensions of noncoding RNA conservation. Trends in Genetics 30:121−23

    doi: 10.1016/j.tig.2014.01.004

    CrossRef   Google Scholar

  • Cite this article

    Li N, Wang Y, Zheng R, Song X. 2022. Research progress on biological functions of lncRNAs in major vegetable crops. Vegetable Research 2:14 doi: 10.48130/VR-2022-0014
    Li N, Wang Y, Zheng R, Song X. 2022. Research progress on biological functions of lncRNAs in major vegetable crops. Vegetable Research 2:14 doi: 10.48130/VR-2022-0014

Figures(2)  /  Tables(2)

Article Metrics

Article views(3080) PDF downloads(682)

Other Articles By Authors

REVIEW   Open Access    

Research progress on biological functions of lncRNAs in major vegetable crops

Vegetable Research  2 Article number: 14  (2022)  |  Cite this article

Abstract: With the advances in genomics and bioinformatics, particularly the extensive application of high-throughput sequencing technology, a large number of non-coding RNAs (ncRNAs) have been discovered, of which long ncRNAs (lncRNAs) refer to a class of transcripts that are more than 200 nucleotides in length. Accumulating evidence demonstrates that lncRNAs play significant roles in a wide range of biological processes, including regulating plant growth and development as well as modulating biotic and abiotic stress responses. Although the study of lncRNAs has been a hotspot of biological research in recent years, the functional characteristics of plant lncRNAs are still in their initial phase and face great challenges. Here, we summarize the characteristics and screening methods of lncRNAs and highlight their biological functions in major vegetable crops, including tomato, Brassica genus crops, cucumber, pepper, carrot, radish, potato, and spinach, which are implicated in the interaction of lncRNAs and miRNAs. This review enhances the understanding of lncRNAs' roles and can guide crop improvement programs in the future.

    • In higher eukaryotic genomes, approximately 90% of the genetic information can pervasively transfer to RNAs[1]. More than 75% of the transcripts do not have protein-coding potential and are classified as non-coding RNAs (ncRNAs)[2,3]. Long non-coding RNAs are a group of ncRNAs with a transcript length of more than 200 nt[4]. Compared with that of mRNAs, their transcript level is generally low and has strong tissue or condition expression specificity[4]. In addition, the sequence conservation of lncRNAs is very low across plant species, which may result from rapid sequence evolution[5]. Most lncRNAs have also been found to be transcribed by RNA pol II, while the others are produced by pol III, IV, and V[6,7]. Based on their location relative to adjacent protein-coding genes in the genome, the lncRNAs are classified into five types: sense lncRNA, antisense lncRNA, bidirectional lncRNA, intronic lncRNA (incRNA), and large intergenic lncRNA (lincRNA)[8]. Each lncRNA is produced by a specific mechanism and can act in cis or trans to regulate gene expression through diverse modes at chromatin, transcription, post-transcription, translation, and post-translation levels[9,10]. With the wide applications of high throughput RNA-sequencing technology, thousands of lncRNAs have been identified in diverse plant species. They act not only as regulators of basic cellular mechanisms but also participate in the regulation of developmental processes as well as biotic and abiotic stress responses[5,1114]. In recent years, the research on the function of lncRNAs in vegetable crops has gradually increased. This paper reviews the characteristics of lncRNAs and their biological functions in vegetables.

    • LncRNAs refer to ncRNAs longer than 200 nt, sometimes in a range of tens of kilo-nucleotides. By comprehensive comparative analysis of lncRNAs among 37 species, we found that the length of lncRNAs fluctuates greatly among different species, ranging from 550.83 nt of mean length in Brassica rapa to 12,053.52 nt in Manihot esculenta[15]. Most plant lncRNAs identified so far are polyadenylated and 5'-capped. However, there are some non-polyadenylated lncRNAs[4,16]. In comparison with those polyadenylated lncRNAs, the length of non-polyadenylated lncRNAs is shorter, the transcript abundance is lower, and the specificity in response to stresses is stronger[17]. Like most proteins, the structure of some lncRNAs is simple, while others appear to have a complex but poorly understood secondary/or tertiary structure, which is generally believed to be necessary for their function. There are two classes of functional elements in lncRNA: One is necessary for physical interactions with partner nucleic acids or proteins, and the other governs the secondary and/or tertiary structure, which further directs interaction partners' binding sites[18].

    • The transcript abundance of lncRNAs is generally low, only 1/30 to 1/60 of the average mRNA expression level[8]. However, there are also exceptions: In a previous study, we found that some lncRNAs had very high expression abundance after comprehensive analysis of the lncRNAs in 37 species[15]. Furthermore, there are significant differences in lncRNA expression patterns across species[15]. Most lncRNAs reside in the nucleus, while they can also export to the cytosol or other organelles, such as mitochondria, which was demonstrated by ribosome profiling and RNA FISH[19]. In the nucleus, lncRNA may perform its function in either cis or trans mode; it has been suggested that lncRNAs with low transcript abundance may work in cis, while those transcribed at a higher level are likely to act in trans[20].

      The expression of lncRNAs was highly specific in different tissues and developmental stages. For example, in cabbage, lncRNA BoNR8 was specifically expressed in the epidermal tissue of the elongation region of germinating seeds[21]. In tomato (Solanum lycopersicum), 4,079, 4,135, and 4,311 lncRNAs that were expressed in tomato fruits at the mature green, breaker, and breaker plus 7 days, respectively, were identified by integrating 134 datasets. Only 20 lncRNAs were expressed in all three developmental stages[22]. It was proposed that the apparent specificity was partly attributed to the generally low expression level of lncRNAs as well as limitations in detection by standard mRNA-sequencing protocols[23]. Most lncRNA sequences are weakly conserved. This shows that only a small part of lncRNA in Chinese cabbage has high homology with lncRNA in other Brassica crops[24]. Based on the analysis of lncRNA from five monocot and five dicot species, it was found that lncRNA had higher sequence conservativeness at the intra-species and sub-species levels but lower inter-species conservativeness[25].

    • With the rapid development of next-generation sequencing (NGS) technology, RNA-Seq has become the first choice for studying the whole transcriptome due to its advantages of high throughput, high accuracy, high sensitivity, and low cost, which has also greatly facilitated the development of lncRNA identification and prediction[26]. However, the construction and sequencing of general transcriptome libraries cannot separate the sense strand and the antisense strand, therefore, a strand-specific RNA-seq (ssRNA-seq) technique was developed to facilitate the identification of transcript orientations[27]. Although NGS techniques are effective, they still suffer from several drawbacks. One major disadvantage is short read lengths, and it is difficult to ensure the accuracy of reconstructed transcripts during assembly[28]. Single-molecule real-time sequencing technology (SMRT) is a third-generation sequencing method that can overcome these limitations and generate long reads without further assembly[29,30]. The third-generation sequencing technology (isoform sequencing, ISO-seq) based on the SMRT sequencing platform has recently been applied to analyze the full-length transcriptome and lncRNA prediction of various species[3134]. In addition, in order to solve the problem of high error rate of SMRT, the 'SMRT + NGS' sequencing joint analysis method, which uses high-quality, high-coverage NGS to correct SMRT data, has been more and more widely used[3537]. ChIP-seq technology, which combines chromatin immunoprecipitation (ChIP) and NGS, provides massive data for the identification of transcription factor binding sites, and it can also be used to identify lncRNA targets of specific transcription factors[38].

    • As an important new regulatory factor, in recent years, the function of lncRNA in vegetable crops has received attention. Here, we summarize the studies involving lncRNA research for some important vegetables, including tomato, Brassica crops, cucumber (Cucumis sativus L.), pepper (Capsicum annuum L.), carrot (Daucus carota L.), radish (Raphanus sativus L.), potato (Solanum tuberosum L.), and spinach (Spinacia oleracea L.), which were also the most studied among the various vegetable species. It was found that the first report about lncRNAs on these vegetables was the discovery of BcMF11 in 2007, which was predicted as an ncRNA associated with pollen development of Chinese cabbage[39]. Then in 2013, the function of BcMF11 was further explored[40]. Based on our statistics, there are fewer than 100 relevant studies in the literature to date (Supplemental Table S1). From 2017 to 2020, the number of published papers increased gradually, then decreased slightly in 2021 (Fig. 1, Supplemental Table S1). Among the studied species, studies on tomato were the most common (35 papers), followed by Brassica crops (32 papers), with relatively few reports on the other six vegetable crops: 8, 7, 6, 2, 2, and 1 for pepper, potato, cucumber, spinach, carrot, and radish, respectively (Fig. 1, Supplemental Table S1).

      Figure 1. 

      Statistics on the published number of lnRNA-related papers of major vegetables.

    • With the advances in genomic and bioinformatic techniques, lncRNAs in vegetable crops were suggested to be involved in various biological processes, and in our study, these processes were mainly categorized into three groups, including growth and development, abiotic stress, as well as biotic stress (Table 1). In different species, lncRNAs were found to be related to various developmental events, such as fruit ripening, vernalization, anther or pollen development, anthocyanin biosynthesis, and sex differentiation[37,4144]. Moreover, lncRNAs were implicated in a variety of abiotic stress responses, such as drought, heat, chilling, and salt stresses[4548]. In addition, lncRNAs may play an important role in plant immunity[4952] (Table 1). Even though a large number of lncRNAs were identified by high-throughput sequencing and suggested to be associated with different physiological processes, only a small portion of lncRNAs have been assessed by functional analysis using molecular biology approaches (Fig. 2, Table 2). The regulation modes of plant lncRNAs in different biological processes are complex and variable[14]. Among them, the interaction between lncRNAs and miRNAs was the most reported relationship in plants. First, lncRNAs can function as an endogenous target mimic (eTM) to sequester miRNAs via base pairing to complementary sites, therefore blocking the interaction of miRNAs and their potential targets[53]. These kinds of lncRNAs are also known as competitive endogenous RNAs (ceRNAs)[53,54]. Second, TMs with extensive complementarity to the 5' and 3' ends of endogenous miRNAs were recently found to trigger miRNA destruction in animals, a process known as target-directed miRNA degradation (TDMD)[5557]. Similarly, by expressing a short tandem target mimic (STTM) in plants, specific endogenous miRNAs can be disrupted. This technology was developed to investigate the function of specific miRNAs[58]. Furthermore, the F-box protein HAWAIIAN SKIRT (HWS) was found to be involved in the degradation pathway and may play a role in the clearance of RNA-induced silencing complexes (RISCs)[59]. Third, some lncRNAs were discovered as precursors of miRNAs, which positively regulate the maturation of miRNAs[54]. Lastly, some lncRNAs can bind and be cleaved by the sequence of complementary miRNAs, that are further processed into phased small-interfering RNAs (phasiRNAs) and guide RNA silencing[54].

      Table 1.  List of long non-coding RNAs (lncRNAs) identified in major vegetable crops.

      RolesSpeciesPathwaysApproachesLncRNAs
      number
      DE-LncRNAs numberRef.
      Growth and developmentSolanum lycopersicumFruit ripeningRNA-seq378[41]
      Solanum lycopersicumFruit ripeningssRNA-seq3,679677[61]
      Solanum lycopersicumFruit expansion and
      ripening
      ssRNA-seq17,674tissue- and stage-dependent[60]
      Solanum lycopersicumRIN target lncRNAs; fruit ripeningChIP-seq & RNA-seq187[38]
      Solanum lycopersicumFruit ripeningintegrate 134 data sets79,322tissue- and stage- specificity[22]
      Capsicum chinense Jacq.Fruit ripeningRNA-seq20,5631,1826[64]
      Capsicum annuumFruit ripeningRNA-seq11,999366[65]
      Capsicum annuumFruit developmentssRNA-seq2,5051,066[66]
      Brassica rapaVernalizationRNA-seq1,961254[42]
      Brassica rapa var. pekinensisVernalizationRNA-seq2,088549[68]
      Brassica
      campestris ssp. pekinensis
      VernalizationssRNA-seq1,858151[69]
      Brassica rapaPollen developmentRNA-seq12,05114[43]
      Brassica rapa ssp. pekinensisAnther developmentSMRT407[34]
      Brassica campestris ssp. pekinensisAnther developmentRNA-seq2,3841,344[72]
      Brassica rapa ssp. pekinensisCytoplasmic male sterilityRNA-seq3,312529[74]
      Brassica
      campestris
      Male sterileRNA-seq13,879361[73]
      Capsicum annuumCytoplasmic male sterilityeRNA-seq10,6551,137[75]
      Solanum lycopersicumSperm cell lineage
      development
      ssRNA-seq31,931cell/tissue-type specificity[76]
      Capsicum annuumAnthocyanin biosynthesisssRNA-seq172[44]
      Solanum tuberosumAnthocyanin BiosynthesisssRNA-seq4,3761,421[80]
      Solanum tuberosumAnthocyanin BiosynthesisRNA-seq1,0726[81]
      Daucus carotaAnthocyanin biosynthesisRNA-seq7,288639[82]
      Solanum lycopersicumTrichome formationssRNA-seq1,303196[83]
      Solanum tuberosumPotato tuber sproutingRNA-seq3,175723[87]
      Spinacia oleraceaFloweringRNA-seq1,141111[89]
      Spinacia oleraceaSex differentiationPacBio Iso-seq & RNA-seq50042[37]
      Growth and developmentBrassica napusOil
      biosynthesis
      ssRNA-seq & RNA-seq datasets8,90513[90]
      Brassica oleracea
      var. capitata
      Cuticular wax
      biosynthesis
      RNA-seq4,459148[91]
      Abiotic stressSolanum lycopersicumDrought responseRNA-seq521244[45]
      Solanum lycopersicumdrought-responseRNA-seq67,7703,053[103]
      Brassica napusDrought responseRNA-seq-477/706[104]
      Solanum tuberosumDrought responseNGS & SMRT3,445[105]
      Brassica rapaHeat responsessRNA-seq4,5941,686[46]
      Brassica junceaHeat and drought responseRNA-seq7,6131,614[107]
      Brassica rapaHeat responseRNA-seq18,2531,229[15]
      Brassica rapa ssp. pekinensisHeat responseRNA-seq27865[106]
      Brassica rapa ssp. chinensis
      (NHCC)
      Cold and heat responseRNA-seq10,0012,236[108]
      Cucumis sativus Heat responseRNA-seq2,085108[109]
      Raphanus sativusHeat responsessRNA-seq169[110]
      Solanum lycopersicumChilling injuryRNA-seq1,411239[47]
      Capsicum annuumChilling injuryRNA-seq9,848 380[111]
      Solanum lycopersicumFruit crackingRNA-seq2,508[112]
      Solanum pennellii and M82Salt responsessRNA-seq1,044154/137[48]
      Cucumis sativusWaterlogging responseRNA-seq3,738922/514/1,476/
      1,270
      [115]
      Cucumis sativusPhosphate-deficiency
      response
      ssRNA-seq14,27722[121]
      Brassica
      napus
      Cadmium toxic responsessRNA-seq5,038301[126]
      Biotic stress
      Solanum tuberosumPhytophthora infestans resistanceRNA-seq2,857133[49]
      Solanum lycopersicumPhytophthora
      infestans resistance
      RNA-seq28,256688[130]
      Solanum lycopersicum L3708Phytophthora
      infestans resistance
      RNA-seq9,011196[131]
      Solanum lycopersicumTYLCV resistancessRNA-seq1,565529[50]
      Solanum lycopersicumTYLCV resistancessRNA-seq2,056345[138]
      Biotic stress
      Solanum lycopersicumBacillus subtilis SL18r-induced tomato resistance against Botrytis cinereaRNA-seq55/34/15[51]
      Solanum lycopersicumPseudomonas putida Sneb821- induced tomato resistance against Meloidogyne incognitaRNA-seq3,37178[140]
      Solanum lycopersicumPst resistanceRNA-seq2,609Different in each comparison[52]
      Solanum lycopersicumPSTVd resistanceRNA-seq6,72644[141]
      Brassica campestris ssp.chinensis Makino Plasmodiophora brassicae resistanceRNA-seq1,492114[143]
      Brassica napusPlasmodiophora brassicae resistancessRNA-seq4,558530[144]
      Brassica rapa ssp. pekinensisDowny mildew resistanceRNA-seq3,711[145]
      Brassica napusSclerotinia sclerotiorum resistanceRNA-seq3,181931[142]
      Brassica rapaFusarium oxysporum
      resistance
      qPCR[146]
      Capsicum
      annuum
      Phytophthora
      capsica resistance
      RNA-seq2,388607[147]
      Solanum tuberosum Pectobacterium carotovorum resistancessRNA-seq1,113559[148]
      Solanum tuberosum Potato Virus Y resistance and heat stressRNA-seq4,007421[149]
      Cucumis sativusPowdery mildew resistancessRNA-seq12,903119[150]
      OthersSolanum lycopersicumEthylene signalingRNA-seq39712[151]
      Solanum
      pimpinellifolium LA1589,
      S. lycopersicum Heinz1706
      Lycopersicon specificityssRNA-seq413/70992/161[152]
      Brassica napus, B. oleracea and
      B. rapa
      Species divergenceRNA-seq1,885/1,910/
      1,299
      186 /157/161[153]
      Capsicum chinenseHeterosis effectssRNA-seq2,5251,932/ 593[154]
      Cucumis hytivusAllopolyploidizationRNA-seq2,2061,328[155]
      Brassica rapa, B. carinata, and
      B. hexaploid
      PolyploidizationRNA-seq2,725/1,672/
      2,810
      725[156]

      Figure 2. 

      The predictive model of regulatory mechanisms of lncRNAs with known functions under various developmental events or stress conditions in different vegetable crops. Full-line arrows represent positive regulatory interactions, blunt-ended bars represent negative regulation and dotted-line arrow indicates that the regulatory mechanism is unclear. White boxes with blue lines represent lncRNAs, light green boxes represent different developmental events or stresses. The letters in red denote positive regulators, while the letters in black denote negative regulators.

      Table 2.  Summary of functionally validated lncRNAs in major vegetables.

      SpeciesLncRNA nameBiological functionsInteraction targetsReferences
      Solanum lycopersicumlncRNA000170Trichome formationSolyc10g006360[83]
      lncRNA1459, lncRNA1840Fruit ripening[61, 62]
      lncRNA2155Fruit ripeningRIN[38]
      ACoS-AS1Trans-splicing; carotenoids biosynthesisSlPSY1[63]
      lncRNA33732Resistance to Phytophthora infestansRBOH[132]
      lncRNA16397Resistance to Phytophthora infestansSlGRX21, SlGRX22[130]
      lncRNA15492Resistance to Phytophthora infestansSl-miR482a[133]
      lncRNA08489Resistance to Phytophthora infestansmiR482e-3p[134]
      lncRNA23468Resistance to Phytophthora infestansmiR-482b[135]
      lncRNA39026Resistance to Phytophthora infestansmiR-168a[136]
      lncRNA40787Resistance to Phytophthora infestansmiR394[137]
      lncRNA42705, lncRNA08711Resistance to Phytophthora infestansmiR159[131]
      slylnc0049, slylnc0761Resistance to TYLCV[50]
      S-slylnc0957Resistance to TYLCV[138]
      SlLNR1Resistance to TYLCV[139]
      MSTRG18363Bacillus subtilis SL18r-induced tomato resistance against Botrytis cinereamiR1918[51]
      lncRNA44664,Pseudomonas putida Sneb821- induced tomato resistance against Meloidogyne incognitamiR396[140]
      lncRNA48734Pseudomonas putida Sneb821- induced tomato resistance against Meloidogyne incognitamiR156[140]
      Brassica oleraceaBoNR8Seed germination; root and silique growth[21]
      Brassica rapabra-eTM160-1, bra-eTM160-2Pollen developmentmiR160-5p[43]
      Brassica rapa ssp. pekinensisMSTRG.19915Resistance to downy mildewBrMAPK15[145]
      Brassica campestrisbra-miR5718HGPollen tube growthmiR5718[73]
      BcMF11Pollen development; male fertility[39, 40]
      Solanum tuberosumStFLORETuber development; drought responseStCDF1[88]
      StLNC0004Resistance to Phytophthora infestansNbEXT[49]
    • Based on previous studies, many lncRNAs were found to be involved in fruit development and the ripening process of vegetable crops. Tomato is a model plant to study flesh fruit development and ripening, and emerging evidence has shown that lncRNAs play crucial roles in this process[22,38,41,60,61],. It was found that lncRNAs may function as ceRNAs of miRNA, interfering with the expression of genes associated with ethylene and carotenoid pathways or directing the methylation of some critical genes involved in fruit ripening[41]. Silencing of either lncRNA1459 or lncRNA1840 resulted in a repressed tomato fruit ripening process[61]. The knockout mutant of lncRNA1459 was obtained by using the clustered regularly interspaced short palindromic repeats (CRISPR)-associated protein 9 (Cas9) system, which, in addition to severely delayed fruit ripening, significantly reduced ethylene biosynthesis and lycopene accumulation compared with wild-type, and meanwhile the expression of fruit-ripening-related genes and lncRNAs was also impaired[62]. RIPENING INHIBITOR (RIN) is one of the known core regulators of fruit ripening, in vivo and in vitro experiments have shown that lncRNA2155 could be targeted by RIN, lncRNA2155 knockout mutant exhibited delayed fruit ripening and the expression of ripening-related transcription factors, ethylene and carotenoids biosynthetic genes were also declined[38]. The ripening process of the tomato fruit is generally accompanied by the accumulation of carotenoids, and phytoene synthase (PSY) is the rate-limiting enzyme of carotenoid biosynthesis. Evidence showed that the trans-splicing between lncRNA ACoS-AS1 and its cognate sense transcript SlPSY1 may be responsible for the loss of function of SlPSY1, which further resulted in the yellow color of fruit in Solanum Lycopersicum var. cerasiforme accession PI 114490[63]. ACoS-AS1 was found to be an essential regulator of the trans-splicing event by generating ACoS-AS1 mutate, which gave rise to red fruit color in PI 114490[63]. Pepper is also an important vegetable worldwide and a model plant for studying the ripening process of non-climacteric flesh fruits. Yang et al. systematically identified 20,563 lncRNAs during three fruit development stages in C. chinense Jacq[64]. Among these, 11,826 were differentially expressed with 5,918 upregulated and 5,908 downregulated[64]. To investigate the regulatory roles of non-coding RNAs in bell pepper fruit ripening, Zuo et al. conducted RNA-seq to explore the expression pattern of lncRNAs in the bell pepper fruit ripening process, and 366 lncRNAs were discovered to exhibit distinct expression patterns in mature green and red ripe fruit[65]. LncRNAs were also involved in hot pepper fruit development, which was verified by comparative analysis of the lncRNA transcript abundance in successive fruit development stages[66].

    • Plants have evolved mechanisms to sense their environment and alter their growth and development for adaptation accordingly. Most varieties of Brassica vegetables must undergo low-temperature vernalization to realize the transition from vegetative growth to reproductive growth[67]. This process is crucial for floral organ formation as well as flowering time regulation. By conducting comparative transcriptome analysis, some lncRNAs were found to be differentially expressed before and after vernalization in Brassica crops[42,68,69]. Furthermore, some lncRNAs were identified as key lncRNAs involved in vernalization through bioinformatic analysis. For instance, in B. rapa, the antisense transcript of BrFLC and BrMAF, which act as repressors of flowering, may play a role in the transcriptional response to vernalization[68]. In Chinese cabbage, the vernalization-related lncRNAs, cirRNAs, miRNAs, and mRNAs were screened for ceRNA network construction, and several lncRNAs were identified as valuable candidates in the vernalization pathway based on this network[69].

    • Male plant sterility, broadly defined as the inability to produce dehiscent anthers, functional pollen, and viable male gametes, opens up new avenues for the utilization of heterosis. In 1763, male sterility was first observed by German botanist Joseph Gottlieb Kolreuter, and more than 610 plant species have been reported to be sterile[70,71]. At present, many Brassica crops have abundant male sterility variant materials, which have been widely utilized in production, but the molecular mechanism of male sterility is still elusive. Pollen abortion is a phenotypic feature of male sterility; therefore, a more in-depth exploration of the molecular regulation mechanism of pollen or anther development is an effective method to understand male sterility. LncRNAs were identified as participants in the process of pollen/anther development and male sterility in Brassica crops[34,43,7274]. For example, an RNA-seq experiment was performed to investigate the dynamic gene expression changes during successive pollen development stages of B. rapa. It is worth noting that 14 lncRNAs were revealed to be strongly co-expressed with 10 function-known coding genes which were related to pollen development. In particular, further exploration of these lncRNAs demonstrated that two lncRNAs, braeTM160-1 and bra-eTM160-2, were negatively involved in pollen formation and male fertility by acting as eTMs of miR160-5p, which further released the transcript of ARF genes[43]. Another study performed whole transcriptome sequencing to enclose the regulatory network of pollen development in different B. campestris sterile lines, of which bra-miR5718HG was demonstrated to reduce the expression of miR5718 and upregulate purple acid phosphorylase 10 (braPAP10), thus inhibiting the growth of pollen tubes and influencing seed set[73]. BcMF11 is a lncRNA that was strongly expressed in the floral organs, and it was confirmed to play an essential role in pollen development by conducting antisense RNA strategy-mediated downregulation of BcMF11 transcript, which leads to abnormal pollen development[39,40]. In addition to Brassica crops, lncRNAs are regarded as a critical regulator in the floral bud development process in pepper through performing RNA-seq and bioinformatic analysis of the transcript abundance in the cytoplasmic male sterility (CMS) line and maintainer line, which laid the foundation for further study of the molecular mechanisms underlying CMS[75]. Moreover, by conducting strand-specific RNA sequencing (ssRNA-seq), lncRNAs were found to be involved in sperm cell lineage development in tomato[76].

    • Anthocyanins are important pigments that are beneficial to health and have major contributions to the quality of fruit[77,78]. At present, the biosynthetic pathway of anthocyanins is well understood, and key regulatory genes have been identified in many species[79]. However, the role of lncRNAs in anthocyanin biosynthesis remains unclear. It is known that anthocyanins are accumulated under light exposure, and in pepper, 172 differentially expressed lncRNAs were identified on the light-exposed and shaded surface of pepper fruit[44]. In potato, Tang et al. found 1,421 differentially expressed lncRNAs between purple- and yellow-fleshed potato tubers. Furthermore, through constructing a lncRNA–mRNA interaction network, lncRNAs such as XLOC_060098 and XLOC_017372 were identified as positive regulators in anthocyanin biosynthesis by target anthocyanin-associated genes[80]. LncRNAs were also implicated in the anthocyanin biosynthesis of potato leaves[81]. Gene annotation suggested that lncRNAs could regulate the expression of PAL, F3H, and CHS, which are critical genes in the anthocyanin biosynthesis pathway and thus modulate the color of potato leaves[81]. Carrot is also an important vegetable that has been cultivated for thousands of years. Carrots were originally purple, and modern yellow varieties were domesticated from mutants lacking anthocyanins. By comparative analysis of the expression profile of lncRNAs in two carrot genotypes with a strong difference in anthocyanin accumulation in roots, Chialva et al. identified 639 lncRNAs with distinct expression patterns between these two genotypes, of which the natural antisense transcript of DcMYB7 was suggested to play an important role in anthocyanin pigmentation[82].

    • LncRNAs are also involved in other developmental events. In young tomato stems, 196 lncRNAs were discovered to be differentially expressed between woolly mutant LA3560 (Wo) and its non-woolly segregants (WT). Among them, lncRNA000170 and its cognate sense transcript, Solyc10g006360, exhibited a common expression trend, and overexpression of either of them could inhibit type I trichome formation[83]. Sprouting is the key factor leading to a quality deterioration of potato tubers and other huge storage losses[84]. Many studies have attempted to reveal the molecular mechanisms underlying potato sprouting[85,86]. Among them, the expression of 723 lncRNAs was distinct in potato tubers from dormancy to sprouting, and these lncRNAs may function by affecting the cellular components and cellular metabolic processes of potato apical buds[87]. Furthermore, a lncRNA named StFLORE together with its counterpart StCDF1 was found to be involved in tuber development and drought response by creating StFLORE knockout mutants and overexpression lines[88]. In spinach, several well-known flowering-related genes such as ELF, COL1, FLT, and FPF1 and also some important flowering transcription factor genes such as MYB, WRKY, GATA, and MADS-box were potential targets for lncRNAs[89]. Based on PacBio Iso-seq and Illumina RNA-seq data, Li et al. discovered 42 differentially expressed lncRNAs in male and female spinach flowers, suggesting the role of lncRNAs in sex determination[37]. In cabbage, Wu et al. identified a lncRNA homologous to Arabidopsis AtR8, BoNR8. Studies have shown that BoNR8 could respond to abiotic stress and negatively regulate seed germination and root and silique growth[21]. B. napus is a conventional oil crop with high economic value. Some lncRNAs were found to be important regulators in oil biosynthesis after comparative analysis of lncRNAs at multiple seed development stages and co-expression analysis[90]. Moreover, lncRNAs were implicated in cuticular wax biosynthesis in cabbage[91].

    • Plants are constantly affected by adverse environmental factors. To survive under various abiotic stresses, plants have to rapidly activate defense mechanisms and adapt to stressful environments[9294]. Among them, lncRNAs have been reported to be involved in multiple abiotic stress responses[5,9599].

      Drought is an important stress factor that affects the normal growth and development of plants. The research on the effects of drought stress on the growth and development of vegetables and crops has always been one of the hotspots in the field of stress research[100102]. In a previous study, a total of 244 lncRNAs were identified and characterized in drought-exposed tomato leaves[45]. Some of them may act as eTMs of miRNAs or through lncRNA–mRNA interactions to respond to drought stress[45]. According to strand-specific RNA-seq, 67,770 lncRNAs were discovered at different anther development stages of tomato, of which 3,053 were drought-responsive[103]. In drought-tolerant B. napus Q2 and drought-sensitive B. napus Qinyou8, 477 and 706 lncRNAs were differentially expressed between the two genotypes under drought stress and rehydration treatment, respectively[104]. Furthermore, a co-expression network of lncRNAs and mRNAs was constructed for functional prediction of these lncRNAs[104]. In potatoes, the role of lncRNAs under drought stress was also explored. A total of 3,445 lncRNAs were identified in different periods of drought stress, and function enrichment analysis indicated that they may be involved in drought response by modulating the 'ubiquitin-mediated proteolysis' pathway[105].

      In the 21st century, the frequent occurrence of extreme high-temperature events will bring a great threat to agricultural production. Growing evidence showed that lncRNAs may play an essential role in heat resistance in Brassica crops[15,46,106108]. In Chinese cabbage, lncRNAs could interact with mRNAs and miRNAs to form a network that affected plant hormone pathways and responded to heat stress[46]. In B. juncea, lncRNAs can also respond to heat and drought stress by functioning as putative targets of miRNAs, or through interaction with abiotic-stress-related transcription factors[107]. Furthermore, in our previous study, 1,229 differentially expressed lncRNAs were identified as being heat-responsive in Chinese cabbage; they can confer thermotolerance by affecting the 'protein processing in the endoplasmic reticulum' and 'plant hormone signaling' pathways, as well as the expression patterns of HSPs and ABA receptor PYL genes[15]. The role of lncRNAs in cucumber and radish under heat stress has also been explored. In cucumber, a total of 2,085 lncRNAs were found to be differentially expressed when exposed to heat stress, and some of them may have executive functions by acting as ceRNAs to compete for miRNA binding sites with mRNAs[109]. Radish is a semi-hardy vegetable, and high temperature is one of the greatest threats to its growth and development. Through performing ssRNA-seq, 169 lncRNAs were predicted to be heat-responsive and one lncRNA–miRNA–mRNA combination was constructed that provided valuable clues for further studies to elucidate their precise functions[110].

      Low-temperature storage is a common storage method for fruits and vegetables after harvest, but for cold-sensitive vegetables such as tomatoes and peppers, improper storage will often cause serious chilling damage. The regulatory relationship between lncRNA and fruit chilling stress has also been investigated in previous studies[47,111]. Combined with RNA-seq and bioinformatic analysis, 239 lncRNAs involved in chilling injury were identified in tomato, some of which may function by targeting chilling-injury-related genes[47]. In particular, a complex regulatory network composed of miRNAs, lncRNAs, and their regulatory targets was established to fully understand the molecular mechanism of lncRNAs in chilling stress response[47]. Likewise, 380 chilling-responsive lncRNA were identified in bell pepper, and their potential targets and relationship with miRNAs, circRNAs, and mRNAs were also assessed to uncover the influenced pathways and processes[111].

      LncRNAs also play important roles in other types of abiotic stresses. In tomatoes, fruit cracking occurs easily under abiotic stresses. Plants have evolved defense mechanisms and regulatory networks to combat this damage. Xue et al. investigated the expression profiles of mRNAs and lncRNAs at different stages of saturated irrigation-treated tomato fruits, and some lncRNAs (XLOC 16662, XLOC_033910, etc.) were identified as participants in regulating tomato fruit cracking via a lncRNA–mRNA (hormone–redox–cell wall) network[112]. By examining the differences in the expression of lncRNAs before and after salt treatment in wild and cultivated tomato materials, Li et al. screened some salt-induced LncRNAs, which may affect tomato salt tolerance by regulating the expression of hormone-pathway-related genes[48]. Cucumber is characterized by a shallow root system. Limited availability of oxygen often occurs during the cucumber cultivation period in unfavorable environmental conditions, one of which is excess water in the soil, which causes leaf wilting, chlorosis, and necrosis and decreased growth rates and yields due to the lack of available oxygen required to support aerobic respiration[113,114]. Through conducting high-throughput RNA-seq, 71 lncRNAs were predicted as members participating in acquiring hypoxia tolerance under long-term waterlogging stress in cucumber, and some of them may function by interacting with miRNAs[115]. In plants, phosphorus is a macronutrient essential for plant growth and yield and plays an important role in nucleic acid, phospholipid composition, energy transfer, and signal transduction[116]. Available forms of phosphorus (phosphate, Pi) are generally low in soil, and many plant species have evolved complex adaptive responses to maintain Pi homeostasis[117120]. LncRNAs were implicated in maintaining phosphate homeostasis in cucumber. Grafting studies combined with RNA-seq identified 22 lncRNAs that could serve as systemic signals during the early Pi deficiency response and can move a long distance from the source region into sink tissues[121]. Cadmium (Cd), a toxic heavy metal, is one of the main inorganic pollutants in the environment[122,123]. It can be freely absorbed and accumulated by plants, resulting in the disruption of nutrient homeostasis, the recurrence of toxicity symptoms, and interference with many physiological processes[124,125]. LncRNAs were also involved in mediating cadmium toxic response and detoxication in B. napus. Of the 5,038 lncRNAs identified, 301 were cadmium-responsive[126].

    • Vegetables often suffer from various biotic stresses during their growth and development, such as infection by fungi, bacteria, viruses, and nematodes[127]. Late blight is one of the most devastating diseases affecting Solanaceae crops and can cause a massive reduction in or even the extinction of potato and tomato production[128,129]. Phytophthora infestans is the causal agent of late blight; therefore, it is of great significance to study the resistance mechanism of tomato and potato to P. infestans. Based on the published RNA-sequencing data, Cao et al. discovered 133 lncRNAs involved in the resistance of P. infestans in potatoes and their regulatory mechanisms by constructing an interaction network[49]. It was remarkable that after transient transformation of StLNC0004 into tobacco, the expression of extensin (NbEXT) was activated, accompanied by the enhancement of resistance to P. infestans[49]. In tomato, the role of lncRNAs in P. infestans resistance has been widely explored[130,131]. Tomato lncRNA33732 activated by WRKY1 is positively involved in tomato resistance to P. infestans by inducing the expression of RESPIRATORY BURST OXIDASE (RBOH) and increasing H2O2 accumulation during early infecting stages[132]. lncRNA16397 could induce the expression of SlGRXs, resulting in a reduction in the accumulation of ROS and damage to the cell membrane, which in turn enhances tomato resistance[130]. Sl-lncRNA15492 acts against P. infestans infection via inhibiting the expression of mature Sl-miR482a, which could target Sl-NBS-LRR resistance genes[133]. Additionally, lncRNAs could function as ceRNA to modulate the expression of resistance-related genes by decoying miRNAs in the tomato-P. infestans interaction. Among them, lncRNA23468 and lncRNA08489 could decoy miR482b and miR482e-3p, respectively, to affect the expression of NBS-LRR genes[134,135]. lncRNA39026 can positively regulate Argonaute proteins 1(AGO1) by decoying miR168a and improve the transcript level of PR genes[136]. lncRNA40787 can suppress the expression of miR394, thereby improving the transcript abundance of LCR and changing the expression of JA-related genes[137]. Furthermore, some lncRNAs could modulate the expression of resistance-related transcription factors by decoying miRNAs, thus enhancing tomato resistance[131].

      Apart from the role of lncRNAs in P. infestans resistance, they were also implicated in yellow leaf curl virus (TYLCV) infection responses[50,138]. Wang et al. identified 529 lncRNAs that could respond to TYLCV infection in the resistant tomato breeding line CLN2777a, and several lncRNAs could serve as miRNA target mimics involved in disease resistance. Two of the lncRNAs, slylnc0049 and slylnc0761, that exhibited a substantial increase after TYLCV inoculation, were functionally characterized by virus-induced gene silencing (VIGS), and it was found that silenced tomato plants accumulated more virus than controls[50]. Furthermore, the role of lncRNAs in virus resistance in TYLCV-susceptible tomato line JS-CT-9210 was explored, and silencing of S-slylnc0957 resulted in improved resistance of tomato to TYLCV infection[102]. In addition, in tomato, the host lncRNA SlLNR1 in susceptible but not in resistant cultivars could interact with viral siRNA which was derived from intergenic region (IR) of TYLCV genome, thereby affecting virus accumulation and disease development during TYLCV infection[139].

      LncRNAs can also mediate Bacillus subtilis SL18r-induced tomato resistance to Botrytis cinerea, in which MSTRG18363 may modulate the expression of SlATL20 by decoying miR1918, thereby triggering the process of induced systemic resistance (ISR) against pathogens[51]. Yang et al. identified 78 lncRNAs that were implicated in Pseudomonas putida Sneb821-induced tomato resistance to Meloidogyne incognita, of which lncRNA44664 and lncRNA48734 could decoy miR396 and miR156, respectively, to competitively inhibit the expression of their target genes, thereby conferring resistance to M. incognita infection[140]. In addition, according to a comprehensive assessment of lncRNA expression profiles, lncRNAs were found to be involved in the immune response against Pseudomonas syringae pv. tomato (Pst) and potato spindle tuber viroid (PSTVd) in tomato[52,141].

      As with all crops, Brassica species are constantly threatened by biotic stresses during production, resulting in huge economic losses. The role of lncRNAs in Brassica crops in mediating responses to Plasmodiophora brassicae, Hyaloperonospora brassica, Sclerotinia sclerotiorum, and Fusarium oxysporum was explored[142146]. For instance, in Chinese cabbage, by comparing the lncRNA expression profiles before and after P. brassicae infection, 114 differentially expressed lncRNAs were identified, and 16 of them were predicted to interact with 15 defense-responsive genes based on the expression correlation between lncRNAs and mRNAs[143]. The role of lncRNAs in P. brassicae response was also explored in B. napus, of which 530 lncRNAs were found to exhibit distinct expression patterns in clubroot-susceptible and clubroot-resistant lines[144]. Downy mildew is an important oomycete disease threatening the production of Brassica vegetables worldwide. It was found that lncRNAs may participate in the disease defense response by regulating the expression of resistance-related genes[145]. Furthermore, silencing lncRNA MSTRG.19915 induced the expression of BrMAPK15 and improved resistance to downy mildew[145]. Additionally, 931 lncRNAs were involved in S. sclerotiorum infection response in B. napus[142]. Following F. oxysporum f. sp. conglutinans (Foc) inoculation, the expression of natural antisense lncRNAs was positively correlated with their cognate sense genes in B. rapa[146].

      Comprehensive analysis of the expression of lncRNAs in Phytophthora capsici-resistant grafted peppers and susceptible samples revealed a total of 607 differentially expressed lncRNAs[147]. These lncRNAs participate in disease resistance responses in part through a lincRNA–miRNA–mRNA interaction network that regulates the expression of disease defense-related genes[147]. LncRNAs were also involved in resistance to Pectobacterium carotovorum and potato virus Y (PVY) in potato[148,149]. Kwenda et al. identified 559 lncRNAs that are P. carotovorum-responsive, and 17 lncRNAs were highly correlated with 12 defense-related genes through co-expression analysis[148]. A systematic RNA-seq analysis explored a comprehensive landscape of 4,007 lncRNAs in tomato infected by PVY at normal and elevated temperature status, of which 12 lncRNAs participated in stress response regulation by recruiting complex mechanisms based on eTM[149]. Cucumber downy mildew (DM) is the most serious epidemic disease in the production of cucumbers in solar greenhouses. After the onset of the disease, most of the leaves of the cucumbers can be withered, and the cucumber fields will turn yellow. To reveal the resistance mechanism of this disease, Nie et al. have performed ssRNA-seq and miRNA-seq to explore the roles of lncRNAs, mRNAs, and miRNAs in DM resistance[150]. According to the expression profiles in resistant and susceptible cucumber lines, a total of 119 lncRNAs were identified to be associated with DM resistance since their expression changed after inoculation with DM. Furthermore, a lncRNA–miRNA–mRNA interaction network was set up to reveal the action mode of lncRNAs in DM response[150].

    • Apart from participating in growth and development, in response to abiotic and biotic stresses, lncRNAs are also involved in many other biological processes in vegetable crops, such as ethylene response and species divergence[151153]. Heterosis is a universal phenomenon in biology. The hybrid generation obtained by crossing different strains, varieties, and even different species often exhibits stronger growth rates and metabolic functions than its parents. Allopolyploid is a manifestation of hybrid vigor, which is obtained by doubling the chromosomes of hybrids produced by crossing different species. LncRNAs were found to be implicated in heterosis and displayed distinct expression patterns in allopolyploids and their parents[154,155]. In pepper, 1,932 lncRNAs were identified to be associated with heterosis, and a co-expression network was constructed to illustrate the functional modes of lncRNAs[154]. In Cucumis, the allotetraploid Cucumis hytivus was produced by chromosome doubling after crossing cultivated cucumber C. sativus with wild-type C. hystrix. Through systemic analysis of the transcriptome, 1,328 lncRNAs were found to be activated following hybridization. Some of their cis-regulatory targets were involved in the regulation of biological chloroplasts, and the others may be associated with epigenetic regulation of leaf verticillium and enhanced photosynthesis[155]. The function of lncRNAs in allopolyploidization was also explored in Brassica genus. Wang et al. discovered 725 differentially expressed lncRNAs between Brassica hexaploid and its parents, and the lncRNAs in the hexaploidy exhibited a significant paternal expression bias. The lncRNA–mRNA interaction network was constructed to visually display the relationship between lncRNAs and their potential target genes. Furthermore, the lncRNAs may perform their roles partially by functioning as ceRNAs or miRNA precursors[156].

    • The wide application of high-throughput RNA sequencing has provided revolutionary ways to discover novel lncRNAs[157,158]. In this review, we introduced the structures and expression features and highlighted the biological functions of lncRNAs in major vegetable crops. This work shows that lncRNAs could participate in a wide range of biological processes, including many development events, such as vernalizaion, fruit ripening, pollen or anther development, anthocyanin biosynthesis, flowering, and sex differentiation. LncRNAs were also confirmed to be involved in a serious of abiotic and biotic stress responses, such as drought, heat, cold, salt, chilling, P. infestans, TYLCV, Pst, and PSTVd (Table 1). However, in comparison with the research on humans and animals, the research involving plants is still in its infancy[159,160]. Although genome sequencing data have been reported for dozens of plants, annotations in most plant species lack information on lncRNAs, and studies of lncRNA functions are limited to only a few model angiosperms[161]. Among the species we explored in vegetable crops, the lncRNA research was mainly concentrated on tomato and Brassica crops, while few studies existed for other vegetable species. Furthermore, the function annotated lncRNAs are limited and only confined to a few cases (Fig. 2 & Table 2). Therefore, it is imperative to expand the research of lncRNA into other vegetables, and more efforts should be made towards a systematic analysis of the regulatory roles of non-coding RNAs in biological processes. Furthermore, the application of traditional reverse genetics based on highly efficient and stable plant genetic transformation systems, such as over-expression and RNAi as well as CRISPR/Cas9 gene-editing technology, would enrich our understanding of the precise function of lncRNAs in plants[162,163].

      LncRNA regulates the function of its target genes in a cis or trans manner through various mechanisms of interaction with DNA, RNA, or proteins[164166]. lncRNAs often work in highly intricate networks to regulate plant growth and development, as well as stress responses[167169]. Several tools have been developed to predict the function modes of lncRNA, for example, the lncRNATargets platform was conducted to predict the interaction of lncRNAs and mRNAs, SpongeScan for lncRNAs and miRNAs, TF2LncRNA for lncRNAs, and transcription factors and RegRNA for identification of functional sites of lncRNAs[170173]. Among them, the interaction between lncRNAs and miRNAs has received a lot of attention. In this paper, we outline many examples of the lncRNA functions as ceRNAs involved in fruit ripening, pollen development, resistance to P. infestans infection, salt stress tolerance, etc.[41,48,72,131]. Furthermore, the crosstalk network between lncRNAs and miRNAs was constructed by bioinformatics researchers for different vegetables under various experimental conditions, which expanded our knowledge of the function modes of lncRNAs. It is generally believed that the specific spatial structures of lncRNAs affect their interactions with other molecular elements, and the functional motif is necessary for physical interaction with various partners[174,175]. Therefore, it is of high importance to further explore the sequence motifs and secondary/tertiary structures, which is essential for fully elucidating the mechanisms of lncRNA regulation and developing new methods to predict lncRNA targets. These studies will provide a new perspective on the involvement of lncRNAs in the complex gene regulatory networks of plant growth and development and stress responses.

      • This work was supported by the National Natural Science Foundation of China (32172583), Natural Science Foundation of Hebei (C2021209005, C2021209019), and the China Postdoctoral Science Foundation (2020M673188, 2021T140097).

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

      • # These authors contributed equally: Nan Li, Yujie Wang

      • Copyright: © 2022 by the author(s). Published by Maximum Academic Press, Fayetteville, GA. This article is an open access article distributed under Creative Commons Attribution License (CC BY 4.0), visit https://creativecommons.org/licenses/by/4.0/.
    Figure (2)  Table (2) References (175)
  • About this article
    Cite this article
    Li N, Wang Y, Zheng R, Song X. 2022. Research progress on biological functions of lncRNAs in major vegetable crops. Vegetable Research 2:14 doi: 10.48130/VR-2022-0014
    Li N, Wang Y, Zheng R, Song X. 2022. Research progress on biological functions of lncRNAs in major vegetable crops. Vegetable Research 2:14 doi: 10.48130/VR-2022-0014

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return