Search
2023 Volume 3
Article Contents
REVIEW   Open Access    

Alternative splicing control of light and temperature stress responses and its prospects in vegetable crops

More Information
  • Vegetable crop quality and productivity are impaired by adverse light and temperature stresses, such as low light, cold, and heat stress. Alternative splicing is a powerful strategy for coping with these abiotic stresses by producing multiple mRNA splicing variants with different subcellular locations, translation efficiency, and coding sequences. Pre-mRNA is alternatively spliced by the spliceosome consisting of different splicing-related proteins termed splicing factors. Recent studies have shown that splicing factors and splicing variants have diverse effects on the development, flowering, and fertility of model plants under light and temperature stresses, but there is very little information available for vegetable crops. Therefore, we conducted a comprehensive literature review on the roles of alternative splicing in regulating plants' responses to light and temperature stress. Several suggestions for future studies and implementation in the field of vegetable crops are presented. We can manipulate alternative splicing using CRISPR/Cas9 technology to produce splicing variants with diverse roles for improving performance of vegetable crop under stress environments.
  • 加载中
  • Supplemental Table S1 Vegetable crops splicing-related proteins.
  • [1]

    Shipman EN, Yu J, Zhou J, Albornoz K, Beckles DM. 2021. Can gene editing reduce postharvest waste and loss of fruit, vegetables, and ornamentals? Horticulture Research 8:1

    doi: 10.1038/s41438-020-00428-4

    CrossRef   Google Scholar

    [2]

    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

    [3]

    Li X, Liang T, Liu H. 2021. How plants coordinate their development in response to light and temperature signals. The Plant Cell 34:955−66

    doi: 10.1093/plcell/koab302

    CrossRef   Google Scholar

    [4]

    Ling Y, Mahfouz MM, Zhou S. 2021. Pre-mRNA alternative splicing as a modulator for heat stress response in plants. Trends in Plant Science 26:1153−70

    doi: 10.1016/j.tplants.2021.07.008

    CrossRef   Google Scholar

    [5]

    Kathare PK, Huq E. 2021. Light-regulated pre-mRNA splicing in plants. Current Opinion in Plant Biology 63:102037

    doi: 10.1016/j.pbi.2021.102037

    CrossRef   Google Scholar

    [6]

    Godoy Herz MA, Kubaczka MG, Brzyżek G, Servi L, Krzyszton M, et al. 2019. Light regulates plant alternative splicing through the control of transcriptional elongation. Molecular Cell 73:1066−1074.e3

    doi: 10.1016/j.molcel.2018.12.005

    CrossRef   Google Scholar

    [7]

    Martín G, Márquez Y, Mantica F, Duque P, Irimia M. 2021. Alternative splicing landscapes in Arabidopsis thaliana across tissues and stress conditions highlight major functional differences with animals. Genome Biology 22:35

    doi: 10.1186/s13059-020-02258-y

    CrossRef   Google Scholar

    [8]

    Jiang J, Zhang C, Wang X. 2015. A recently evolved isoform of the transcription factor BES1 promotes brassinosteroid signaling and development in Arabidopsis thaliana. The Plant Cell 27:361−74

    doi: 10.1105/tpc.114.133678

    CrossRef   Google Scholar

    [9]

    Guo M, Zhang Y, Jia X, Wang X, Zhang Y, et al. 2022. Alternative splicing of REGULATOR OF LEAF INCLINATION 1 modulates phosphate starvation signaling and growth in plants. The Plant Cell 34:3319−38

    doi: 10.1093/plcell/koac161

    CrossRef   Google Scholar

    [10]

    Liu T, Zhang X, Zhang H, Cheng Z, Liu J, et al. 2022. Dwarf and high Tillering1 represses rice tillering through mediating the splicing of D14 pre-mRNA. The Plant Cell 34:3301−18

    doi: 10.1093/plcell/koac169

    CrossRef   Google Scholar

    [11]

    Wu Z, Liang J, Wang C, Ding L, Zhao X, et al. 2019. Alternative splicing provides a mechanism to regulate LlHSFA3 function in response to heat stress in lily. Plant Physiology 181:1651−67

    doi: 10.1104/pp.19.00839

    CrossRef   Google Scholar

    [12]

    Shikata H, Hanada K, Ushijima T, Nakashima M, Suzuki Y, et al. 2014. Phytochrome controls alternative splicing to mediate light responses in Arabidopsis. Proceedings of the National Academy of Sciences of the United States of America 111:18781−86

    doi: 10.1073/pnas.1407147112

    CrossRef   Google Scholar

    [13]

    Li C, Zheng L, Zhang J, Lv Y, Liu J, et al. 2017. Characterization and functional analysis of four HYH splicing variants in Arabidopsis hypocotyl elongation. Gene 619:44−49

    doi: 10.1016/j.gene.2017.04.001

    CrossRef   Google Scholar

    [14]

    Wang BB, Brendel V. 2004. The ASRG database: identification and survey of Arabidopsis thaliana genes involved in pre-mRNA splicing. Genome Biology 5:R102

    doi: 10.1186/gb-2004-5-12-r102

    CrossRef   Google Scholar

    [15]

    Wilkinson ME, Charenton C, Nagai K. 2020. RNA splicing by the spliceosome. Annual Review of Biochemistry 89:359−88

    doi: 10.1146/annurev-biochem-091719-064225

    CrossRef   Google Scholar

    [16]

    Laloum T, Martín G, Duque P. 2018. Alternative splicing control of abiotic stress responses. Trends in Plant Science 23:140−50

    doi: 10.1016/j.tplants.2017.09.019

    CrossRef   Google Scholar

    [17]

    Capovilla G, Delhomme N, Collani S, Shutava I, Bezrukov I, et al. 2018. PORCUPINE regulates development in response to temperature through alternative splicing. Nature Plants 4:534−39

    doi: 10.1038/s41477-018-0176-z

    CrossRef   Google Scholar

    [18]

    Nibau C, Gallemí M, Dadarou D, Doonan JH, Cavallari N. 2019. Thermo-sensitive alternative splicing of FLOWERING LOCUS M is modulated by cyclin-dependent kinase G2. Frontiers in Plant Science 10:1680

    doi: 10.3389/fpls.2019.01680

    CrossRef   Google Scholar

    [19]

    Jia J, Long Y, Zhang H, Li Z, Liu Z, et al. 2020. Post-transcriptional splicing of nascent RNA contributes to widespread intron retention in plants. Nature Plants 6:780−88

    doi: 10.1038/s41477-020-0688-1

    CrossRef   Google Scholar

    [20]

    Blencowe BJ. 2006. Alternative splicing: new insights from global analyses. Cell 126:37−47

    doi: 10.1016/j.cell.2006.06.023

    CrossRef   Google Scholar

    [21]

    Ule J, Blencowe BJ. 2019. Alternative splicing regulatory networks: functions, mechanisms, and evolution. Molecular Cell 76:329−45

    doi: 10.1016/j.molcel.2019.09.017

    CrossRef   Google Scholar

    [22]

    Wahl MC, Will CL, Lührmann R. 2009. The spliceosome: design principles of a dynamic RNP machine. Cell 136:701−18

    doi: 10.1016/j.cell.2009.02.009

    CrossRef   Google Scholar

    [23]

    Cabezas-Fuster A, Micol-Ponce R, Fontcuberta-Cervera S, Ponce MR. 2022. Missplicing suppressor alleles of Arabidopsis PRE-MRNA PROCESSING FACTOR 8 increase splicing fidelity by reducing the use of novel splice sites. Nucleic Acids Research 50:5513−27

    doi: 10.1093/nar/gkac338

    CrossRef   Google Scholar

    [24]

    Deng X, Lu T, Wang L, Gu L, Sun J, et al. 2016. Recruitment of the NineTeen Complex to the activated spliceosome requires AtPRMT5. Proceedings of the National Academy of Sciences of the United States of America 113:5447−52

    doi: 10.1073/pnas.1522458113

    CrossRef   Google Scholar

    [25]

    Reddy ASN, Marquez Y, Kalyna M, Barta A. 2013. Complexity of the alternative splicing landscape in plants. The Plant Cell 25:3657−83

    doi: 10.1105/tpc.113.117523

    CrossRef   Google Scholar

    [26]

    Marquez Y, Brown JWS, Simpson C, Barta A, Kalyna M. 2012. Transcriptome survey reveals increased complexity of the alternative splicing landscape in Arabidopsis. Genome Research 22:1184−95

    doi: 10.1101/gr.134106.111

    CrossRef   Google Scholar

    [27]

    James AB, Syed NH, Bordage S, Marshall J, Nimmo GA, et al. 2012. Alternative splicing mediates responses of the Arabidopsis circadian clock to temperature changes. The Plant Cell 24:961−81

    doi: 10.1105/tpc.111.093948

    CrossRef   Google Scholar

    [28]

    Gu Y, Zebell SG, Liang Z, Wang S, Kang BH, et al. 2016. Nuclear pore permeabilization is a convergent signaling event in effector-triggered immunity. Cell 166:1526−1538.e11

    doi: 10.1016/j.cell.2016.07.042

    CrossRef   Google Scholar

    [29]

    Williamson L, Saponaro M, Boeing S, East P, Mitter R, et al. 2017. UV irradiation induces a non-coding RNA that functionally opposes the protein encoded by the same gene. Cell 168:843−855.e13

    doi: 10.1016/j.cell.2017.01.019

    CrossRef   Google Scholar

    [30]

    Huang J, Lu X, Wu H, Xie Y, Peng Q, et al. 2020. Phytophthora effectors modulate genome-wide alternative splicing of host mRNAs to reprogram plant immunity. Molecular Plant 13:1470−84

    doi: 10.1016/j.molp.2020.07.007

    CrossRef   Google Scholar

    [31]

    Dever TE, Ivanov IP, Sachs MS. 2020. Conserved upstream open reading frame nascent peptides that control translation. Annual Review of Genetics 54:237−64

    doi: 10.1146/annurev-genet-112618-043822

    CrossRef   Google Scholar

    [32]

    Dong J, Chen H, Deng X, Irish VF, Wei N. 2020. Phytochrome B induces intron retention and translational inhibition of PHYTOCHROME-INTERACTING FACTOR3. Plant Physiology 182:159−66

    doi: 10.1104/pp.19.00835

    CrossRef   Google Scholar

    [33]

    Penfield S, Josse EM, Halliday KJ. 2010. A role for an alternative splice variant of PIF6 in the control of Arabidopsis primary seed dormancy. Plant Molecular Biology 73:89−95

    doi: 10.1007/s11103-009-9571-1

    CrossRef   Google Scholar

    [34]

    Tognacca RS, Servi L, Hernando CE, Saura-Sanchez M, Yanovsky MJ, et al. 2019. Alternative splicing regulation during light-induced germination of Arabidopsis thaliana seeds. Frontiers in Plant Science 10:1076

    doi: 10.3389/fpls.2019.01076

    CrossRef   Google Scholar

    [35]

    Xin R, Zhu L, Salomé PA, Mancini E, Marshall CM, et al. 2017. SPF45-related splicing factor for phytochrome signaling promotes photomorphogenesis by regulating pre-mRNA splicing in Arabidopsis. Proceedings of the National Academy of Sciences of the United States of America 114:E7018−E7027

    doi: 10.1073/pnas.1706379114

    CrossRef   Google Scholar

    [36]

    Charng YY, Liu HC, Liu NY, Chi WT, Wang CN et al. 2007. A heat-inducible transcription factor, HsfA2, is required for extension of acquired thermotolerance in Arabidopsis. Plant Physiology 143:251−62

    doi: 10.1104/pp.106.091322

    CrossRef   Google Scholar

    [37]

    Liu J, Sun N, Liu M, Liu J, Du B, et al. 2013. An autoregulatory loop controlling Arabidopsis HsfA2 expression: role of heat shock-induced alternative splicing. Plant Physiology 162:512−21

    doi: 10.1104/pp.112.205864

    CrossRef   Google Scholar

    [38]

    Moreno AA, Mukhtar MS, Blanco F, Boatwright JL, Moreno I, et al. 2012. IRE1/bZIP60-mediated unfolded protein response plays distinct roles in plant immunity and abiotic stress responses. PLoS ONE 7:e31944

    doi: 10.1371/journal.pone.0031944

    CrossRef   Google Scholar

    [39]

    Deng Y, Humbert S, Liu J, Srivastava R, Rothstein SJ, et al. 2011. Heat induces the splicing by IRE1 of a mRNA encoding a transcription factor involved in the unfolded protein response in Arabidopsis. Proceedings of the National Academy of Sciences of the United States of America 108:7247−52

    doi: 10.1073/pnas.1102117108

    CrossRef   Google Scholar

    [40]

    Li Z, Tang J, Srivastava R, Bassham DC, Howell SH. 2020. The transcription factor bZIP60 links the unfolded protein response to the heat stress response in maize. The Plant Cell 32:3559−75

    doi: 10.1105/tpc.20.00260

    CrossRef   Google Scholar

    [41]

    Deng Y, Srivastava R, Howell SH. 2013. Protein kinase and ribonuclease domains of IRE1 confer stress tolerance, vegetative growth, and reproductive development in Arabidopsis. Proceedings of the National Academy of Sciences of the United States of America 110:19633−38

    doi: 10.1073/pnas.1314749110

    CrossRef   Google Scholar

    [42]

    Deng Y, Srivastava R, Quilichini TD, Dong H, Bao Y, et al. 2016. IRE1, a component of the unfolded protein response signaling pathway, protects pollen development in Arabidopsis from heat stress. The Plant Journal 88:193−204

    doi: 10.1111/tpj.13239

    CrossRef   Google Scholar

    [43]

    Iwata Y, Fedoroff NV, Koizumi N. 2008. Arabidopsis bZIP60 is a proteolysis-activated transcription factor involved in the endoplasmic reticulum stress response. The Plant Cell 20:3107−21

    doi: 10.1105/tpc.108.061002

    CrossRef   Google Scholar

    [44]

    Posé D, Verhage L, Ott F, Yant L, Mathieu J, et al. 2013. Temperature-dependent regulation of flowering by antagonistic FLM variants. Nature 503:414−17

    doi: 10.1038/nature12633

    CrossRef   Google Scholar

    [45]

    Zhu J, Lou Y, Shi Q, Zhang S, Zhou W, et al. 2020. Slowing development restores the fertility of thermo-sensitive male-sterile plant lines. Nature Plants 6:360−67

    doi: 10.1038/s41477-020-0622-6

    CrossRef   Google Scholar

    [46]

    Huang X, Niu J, Sun M, Zhu J, Gao J, et al. 2013. CYCLIN-DEPENDENT KINASE G1 is associated with the spliceosome to regulate CALLOSE SYNTHASE5 splicing and pollen wall formation in Arabidopsis. The Plant Cell 25:637−48

    doi: 10.1105/tpc.112.107896

    CrossRef   Google Scholar

    [47]

    Riegler S, Servi L, Scarpin MR, Godoy Herz MA, Kubaczka MG, et al. 2021. Light regulates alternative splicing outcomes via the TOR kinase pathway. Cell Reports 36:109676

    doi: 10.1016/j.celrep.2021.109676

    CrossRef   Google Scholar

    [48]

    Jung JH, Domijan M, Klose C, Biswas S, Ezer D, et al. 2016. Phytochromes function as thermosensors in Arabidopsis. Science 354:886−89

    doi: 10.1126/science.aaf6005

    CrossRef   Google Scholar

    [49]

    Jiang B, Shi Y, Peng Y, Jia Y, Yan Y, et al. 2020. Cold-induced CBF-PIF3 interaction enhances freezing tolerance by stabilizing the phyB thermosensor in Arabidopsis. Molecular Plant 13:894−906

    doi: 10.1016/j.molp.2020.04.006

    CrossRef   Google Scholar

    [50]

    Xin R, Kathare PK, Huq E. 2019. Coordinated regulation of pre-mRNA splicing by the SFPS-RRC1 complex to promote photomorphogenesis. The Plant Cell 31:2052−69

    doi: 10.1105/tpc.18.00786

    CrossRef   Google Scholar

    [51]

    Shikata H, Shibata M, Ushijima T, Nakashima M, Kong SG, et al. 2012. The RS domain of Arabidopsis splicing factor RRC1 is required for phytochrome B signal transduction. The Plant Journal 70:727−38

    doi: 10.1111/j.1365-313X.2012.04937.x

    CrossRef   Google Scholar

    [52]

    Hartmann L, Drewe-Boß P, Wießner T, Wagner G, Geue S, et al. 2016. Alternative splicing substantially diversifies the transcriptome during early photomorphogenesis and correlates with the energy availability in Arabidopsis. The Plant Cell 28:2715−34

    doi: 10.1105/tpc.16.00508

    CrossRef   Google Scholar

    [53]

    Schwenk P, Sheerin DJ, Ponnu J, Staudt AM, Lesch KL, et al. 2021. Uncovering a novel function of the CCR4-NOT complex in phytochrome A-mediated light signalling in plants. eLife 10:e63697

    doi: 10.7554/eLife.63697

    CrossRef   Google Scholar

    [54]

    Li Y, Du Y, Huai J, Jing Y, Lin R. 2022. The RNA helicase UAP56 and the E3 ubiquitin ligase COP1 coordinately regulate alternative splicing to repress photomorphogenesis in Arabidopsis. The Plant Cell 34:4191−212

    doi: 10.1093/plcell/koac235

    CrossRef   Google Scholar

    [55]

    Vogel JL, Parsell DA, Lindquist S. 1995. Heat-shock proteins Hsp104 and Hsp70 reactivate mRNA splicing after heat inactivation. Current Biology 5:306−17

    doi: 10.1016/S0960-9822(95)00061-3

    CrossRef   Google Scholar

    [56]

    Song P, Jia Q, Xiao X, Tang Y, Liu C, et al. 2021. HSP70-3 interacts with phospholipase Dδ and participates in heat stress defense. Plant Physiology 185:1148−65

    doi: 10.1093/plphys/kiaa083

    CrossRef   Google Scholar

    [57]

    Wang TY, Wu JR, Duong NKT, Lu CA, Yeh CH, et al. 2021. HSP70-4 and farnesylated AtJ3 constitute a specific HSP70/HSP40-based chaperone machinery essential for prolonged heat stress tolerance in Arabidopsis. Journal of Plant Physiology 261:153430

    doi: 10.1016/j.jplph.2021.153430

    CrossRef   Google Scholar

    [58]

    Tiwari LD, Khungar L, Grover A. 2020. AtHsc70-1 negatively regulates the basal heat tolerance in Arabidopsis thaliana through affecting the activity of HsfAs and Hsp101. The Plant Journal 103:2069−83

    doi: 10.1111/tpj.14883

    CrossRef   Google Scholar

    [59]

    Guan Q, Wu J, Zhang Y, Jiang C, Liu R, et al. 2013. A DEAD box RNA helicase is critical for pre-mRNA splicing, cold-responsive gene regulation, and cold tolerance in Arabidopsis. The Plant Cell 25:342−56

    doi: 10.1105/tpc.112.108340

    CrossRef   Google Scholar

    [60]

    Guan Q, Wen C, Zeng H, Zhu J. 2013. A KH domain-containing putative RNA-binding protein is critical for heat stress-responsive gene regulation and thermotolerance in Arabidopsis. Molecular Plant 6:386−95

    doi: 10.1093/mp/sss119

    CrossRef   Google Scholar

    [61]

    Schlaen RG, Mancini E, Sanchez SE, Perez-Santángelo S, Rugnone ML, et al. 2015. The spliceosome assembly factor GEMIN2 attenuates the effects of temperature on alternative splicing and circadian rhythms. Proceedings of the National Academy of Sciences of the United States of America 112:9382−87

    doi: 10.1073/pnas.1504541112

    CrossRef   Google Scholar

    [62]

    Kim JS, Jung HJ, Lee HJ, Kim KA, Goh CH, et al. 2008. Glycine-rich RNA-binding protein7 affects abiotic stress responses by regulating stomata opening and closing in Arabidopsis thaliana. The Plant Journal 55:455−466

    doi: 10.1111/j.1365-313X.2008.03518.x

    CrossRef   Google Scholar

    [63]

    de Francisco Amorim M, Willing EM, Szabo EX, Francisco-Mangilet AG, Droste-Borel I, et al. 2018. The U1 snRNP subunit LUC7 modulates plant development and stress responses via regulation of alternative splicing. The Plant Cell 30:2838−54

    doi: 10.1105/tpc.18.00244

    CrossRef   Google Scholar

    [64]

    Calixto CPG, Guo W, James AB, Tzioutziou NA, Entizne JC, et al. 2018. Rapid and dynamic alternative splicing impacts the Arabidopsis cold response transcriptome. The Plant Cell 30:1424−44

    doi: 10.1105/tpc.18.00177

    CrossRef   Google Scholar

    [65]

    Lu CA, Huang CK, Huang WS, Huang TS, Liu HY, et al. 2020. DEAD-box RNA helicase 42 plays a critical role in pre-mRNA splicing under cold stress. Plant Physiology 182:255−71

    doi: 10.1104/pp.19.00832

    CrossRef   Google Scholar

    [66]

    Kim GD, Yoo SD, Cho YH. 2018. STABILIZED1 as a heat stress-specific splicing factor in Arabidopsis thaliana. Plant Signaling & Behavior 13:e1432955

    doi: 10.1080/15592324.2018.1432955

    CrossRef   Google Scholar

    [67]

    Lee BH, Kapoor A, Zhu J, Zhu JK. 2006. STABILIZED1, a stress-upregulated nuclear protein, is required for pre-mRNA splicing, mRNA turnover, and stress tolerance in Arabidopsis. The Plant Cell 18:1736−49

    doi: 10.1105/tpc.106.042184

    CrossRef   Google Scholar

    [68]

    Nibau C, Dadarou D, Kargios N, Mallioura A, Fernandez-Fuentes N, et al. 2020. A functional kinase is necessary for cyclin-dependent kinase G1 (CDKG1) to maintain fertility at high ambient temperature in Arabidopsis. Frontiers in Plant Science 11:586870

    doi: 10.3389/fpls.2020.586870

    CrossRef   Google Scholar

    [69]

    Zheng T, Nibau C, Phillips DW, Jenkins G, Armstrong SJ, et al. 2014. CDKG1 protein kinase is essential for synapsis and male meiosis at high ambient temperature in Arabidopsis thaliana. Proceedings of the National Academy of Sciences of the United States of America 111:2182−87

    doi: 10.1073/pnas.1318460111

    CrossRef   Google Scholar

    [70]

    Cavallari N, Nibau C, Fuchs A, Dadarou D, Barta A, et al. 2018. The cyclin-dependent kinase G group defines a thermo-sensitive alternative splicing circuit modulating the expression of Arabidopsis ATU2AF65A. The Plant Journal 94:1010−22

    doi: 10.1111/tpj.13914

    CrossRef   Google Scholar

    [71]

    Lee KC, Chung KS, Lee HT, Park JH, Lee JH, et al. 2020. Role of Arabidopsis splicing factor SF1 in temperature-responsive alternative splicing of FLM pre-mRNA. Frontiers in Plant Science 11:596354

    doi: 10.3389/fpls.2020.596354

    CrossRef   Google Scholar

    [72]

    Zhao N, Su X, Liu Z, Zhou J, Su Y, et al. 2022. The RNA recognition motif-containing protein UBA2c prevents early flowering by promoting transcription of the flowering repressor FLM in Arabidopsis. New Phytologist 233:751−65

    doi: 10.1111/nph.17836

    CrossRef   Google Scholar

    [73]

    Chang P, Hsieh HY, Tu SL. 2022. The U1 snRNP component RBP45d regulates temperature-responsive flowering in Arabidopsis. The Plant cell 34:834−51

    doi: 10.1093/plcell/koab273

    CrossRef   Google Scholar

    [74]

    Zhao Z, Dent C, Liang H, Lv J, Shang G, et al. 2022. CRY2 interacts with CIS1 to regulate thermosensory flowering via FLM alternative splicing. Nature Communications 13:7045

    doi: 10.1038/s41467-022-34886-2

    CrossRef   Google Scholar

    [75]

    Du L, Song J, Forney C, Palmer LC, Fillmore S, et al. 2016. Proteome changes in banana fruit peel tissue in response to ethylene and high-temperature treatments. Horticulture Research 3:16012

    doi: 10.1038/hortres.2016.12

    CrossRef   Google Scholar

    [76]

    Dai H, Zhu Z, Wang Z, Zhang Z, Kong W, et al. 2022. Galactinol synthase 1 improves cucumber performance under cold stress by enhancing assimilate translocation. Horticulture Research 9:uhab063

    doi: 10.1093/hr/uhab063

    CrossRef   Google Scholar

    [77]

    Li T, Yamane H, Tao R. 2021. Preharvest long-term exposure to UV-B radiation promotes fruit ripening and modifies stage-specific anthocyanin metabolism in highbush blueberry. Horticulture Research 8:67

    doi: 10.1038/s41438-021-00503-4

    CrossRef   Google Scholar

    [78]

    Alabd A, Ahmad M, Zhang X, Gao Y, Peng L, et al. 2022. Light-responsive transcription factor PpWRKY44 induces anthocyanin accumulation by regulating PpMYB10 expression in pear. Horticulture Research 9:uhac199

    doi: 10.1093/hr/uhac199

    CrossRef   Google Scholar

    [79]

    Liu Z, Qin J, Tian X, Xu S, Wang Y, et al. 2018. Global profiling of alternative splicing landscape responsive to drought, heat and their combination in wheat (Triticum aestivum L.). Plant Biotechnology Journal 16:714−26

    doi: 10.1111/pbi.12822

    CrossRef   Google Scholar

    [80]

    Liu B, Zhao S, Li P, Yin Y, Niu Q, et al. 2021. Plant buffering against the high-light stress-induced accumulation of CsGA2ox8 transcripts via alternative splicing to finely tune gibberellin levels and maintain hypocotyl elongation. Horticulture Research 8:2

    doi: 10.1038/s41438-020-00430-w

    CrossRef   Google Scholar

    [81]

    Hu Y, Mesihovic A, Jiménez-Gómez JM, Röth S, Gebhardt P, et al. 2020. Natural variation in HsfA2 pre-mRNA splicing is associated with changes in thermotolerance during tomato domestication. New Phytologist 225:1297−310

    doi: 10.1111/nph.16221

    CrossRef   Google Scholar

    [82]

    Schornack S, Ballvora A, Gürlebeck D, Peart J, Ganal M, et al. 2004. The tomato resistance protein Bs4 is a predicted non-nuclear TIR-NB-LRR protein that mediates defense responses to severely truncated derivatives of AvrBs4 and overexpressed AvrBs3. The Plant Journal 37:46−60

    doi: 10.1046/j.1365-313X.2003.01937.x

    CrossRef   Google Scholar

    [83]

    Piffanelli P, Zhou F, Casais C, Orme J, Jarosch B, et al. 2002. The barley MLO modulator of defense and cell death is responsive to biotic and abiotic stress stimuli. Plant Physiology 129:1076−85

    doi: 10.1104/pp.010954

    CrossRef   Google Scholar

    [84]

    Xu Y, Lei Y, Li R, Zhang L, Zhao Z, et al. 2017. XAP5 CIRCADIAN TIMEKEEPER positively regulates RESISTANCE TO POWDERY MILDEW8.1-mediated immunity in Arabidopsis. Frontiers in Plant Science 8:2044

    doi: 10.3389/fpls.2017.02044

    CrossRef   Google Scholar

    [85]

    Chehab EW, Kim S, Savchenko T, Kliebenstein D, Dehesh K, et al. 2011. Intronic T-DNA insertion renders Arabidopsis opr3 a conditional jasmonic acid-producing mutant. Plant Physiology 156:770−78

    doi: 10.1104/pp.111.174169

    CrossRef   Google Scholar

    [86]

    Zhang C, Huang L, Zhang H, Hao Q, Lyu B, et al. 2019. An ancestral NB-LRR with duplicated 3′UTRs confers stripe rust resistance in wheat and barley. Nature Communications 10:4023

    doi: 10.1038/s41467-019-11872-9

    CrossRef   Google Scholar

    [87]

    Wang L, Wang L, Yang T, Wang B, Lin Q, Zhu S, et al. 2020. RALF1-FERONIA complex affects splicing dynamics to modulate stress responses and growth in plants. Science Advances 6:eaaz1622

    doi: 10.1126/sciadv.aaz1622

    CrossRef   Google Scholar

    [88]

    Lin J, Shi J, Zhang Z, Zhong B, Zhu Z. 2022. Plant AFC2 kinase desensitizes thermomorphogenesis through modulation of alternative splicing. iScience 25:104051

    doi: 10.1016/j.isci.2022.104051

    CrossRef   Google Scholar

    [89]

    Dressano K, Weckwerth PR, Poretsky E, Takahashi Y, Villarreal C, et al. 2020. Dynamic regulation of Pep-induced immunity through post-translational control of defence transcript splicing. Nature Plants 6:1008−19

    doi: 10.1038/s41477-020-0724-1

    CrossRef   Google Scholar

    [90]

    Gao C. 2021. Genome engineering for crop improvement and future agriculture. Cell 184:1621−35

    doi: 10.1016/j.cell.2021.01.005

    CrossRef   Google Scholar

    [91]

    Yang T, Ali M, Lin L, Li P, He H, et al. 2023. Recoloring tomato fruit by CRISPR/Cas9-mediated multiplex gene editing. Horticulture Research 10:uhac214

    doi: 10.1093/hr/uhac214

    CrossRef   Google Scholar

    [92]

    Quint M, Delker C, Franklin KA, Wigge PA, Halliday KJ, et al. 2016. Molecular and genetic control of plant thermomorphogenesis. Nature Plants 2:15190

    doi: 10.1038/nplants.2015.190

    CrossRef   Google Scholar

    [93]

    Yu X, Liu H, Klejnot J, Lin C. 2010. The cryptochrome blue light receptors. The Arabidopsis Book 8:e0135

    doi: 10.1199/tab.0135

    CrossRef   Google Scholar

  • Cite this article

    Cao H, Wu T, Shi L, Li Y, Zhang C. 2023. Alternative splicing control of light and temperature stress responses and its prospects in vegetable crops. Vegetable Research 3:17 doi: 10.48130/VR-2023-0017
    Cao H, Wu T, Shi L, Li Y, Zhang C. 2023. Alternative splicing control of light and temperature stress responses and its prospects in vegetable crops. Vegetable Research 3:17 doi: 10.48130/VR-2023-0017

Figures(4)  /  Tables(1)

Article Metrics

Article views(2673) PDF downloads(265)

REVIEW   Open Access    

Alternative splicing control of light and temperature stress responses and its prospects in vegetable crops

Vegetable Research  3 Article number: 17  (2023)  |  Cite this article

Abstract: Vegetable crop quality and productivity are impaired by adverse light and temperature stresses, such as low light, cold, and heat stress. Alternative splicing is a powerful strategy for coping with these abiotic stresses by producing multiple mRNA splicing variants with different subcellular locations, translation efficiency, and coding sequences. Pre-mRNA is alternatively spliced by the spliceosome consisting of different splicing-related proteins termed splicing factors. Recent studies have shown that splicing factors and splicing variants have diverse effects on the development, flowering, and fertility of model plants under light and temperature stresses, but there is very little information available for vegetable crops. Therefore, we conducted a comprehensive literature review on the roles of alternative splicing in regulating plants' responses to light and temperature stress. Several suggestions for future studies and implementation in the field of vegetable crops are presented. We can manipulate alternative splicing using CRISPR/Cas9 technology to produce splicing variants with diverse roles for improving performance of vegetable crop under stress environments.

    • One of the most evident consequences of global warming is the escalation of extreme weather events, which enables abiotic stress to occur more frequently[1,2]. Vegetable crops are diverse in nature including cucumber, watermelon, tomato, pepper, and lettuce, etc. All have significant economic, nutritional, and aesthetic value. Adverse light and temperature are two major environmental factors that can trigger a variety of common abiotic stressors including low light, high light, cold, and heat stresses, etc. These stressors will not only damage vegetable crops, but will also have a substantial impact on their development, yield, and quality[3].

      In plants, alternative splicing (AS) is disproportionately used to produce critical specific mRNA splicing variants by the spliceosome for coping with fluctuating light and temperature environments. The spliceosome is a dynamical complex composed of hundreds of splicing factor (SF) proteins responsible for recognizing cis-elements, removing introns, and selectively ligating exons of pre-mRNA[46]. Almost 85% of genes are multiexon, and 70% of them are alternatively spliced in Arabidopsis[7]. In recent times, numerous studies have been carried out to determine the role of AS in various plant species. Splicing variants of SL (strigolactone) receptor gene D14 and REGULATOR OF LEAF INCLINATION 1 (RLI1) act as very critical roles in controlling key agronomic traits in rice[810]. Furthermore, the AS of heat-shock transcription factors (HSFs), hy5 homolog (HYH), and phytochrome interacting factors (PIFs) from Arabidopsis are regulated, and these induced splicing variants play significant roles under diverse light and temperature conditions[5,1113]. So, AS is an important and widely used strategy for plants to cope with light and temperature stresses.

      SFs, including glycine-rich RNA-binding proteins (GRPs), serine/arginine-rich (SR) proteins, and small nuclear ribonucleoprotein particles (snRNPs), are essential components of the spliceosome that govern AS[14,15]. The majority of these SFs have been recognized as crucial for stress tolerance, and mutants of these SFs are usually defective in growth and development, with poor tolerance against light and temperature stresses[1618]. Hence, studies on AS events and SFs involved in the cascade of light and temperature stress signals is emerging and crucial. In this review, we first summarize the current understanding of the core spliceosome and discuss the functions of different types of AS events and SFs in plants. Then, we briefly discuss the classical regulatory pathways in the control of AS under adverse light and temperature conditions. Moreover, knocking out the gene by CRISPR/Cas9 technique or overexpressing its major coding splicing variant usually will lead to undesired phenotype changes in plant growth and development. Thus, a possible strategy is proposed for discussion. Through CRISPR/Cas9 technology we can editing the splicing codes or splicing factors that can manipulate the AS to produce splicing variants with distinct function for understanding the mechanism of AS and developing genetic improvements in vegetable crops.

    • In eukaryotes, most genes are interrupted by introns and protein-coding sequences known as exons. gDNA is transcribed by RNA polymerase-II (Pol-II) to produce pre-mRNA with introns cotranscriptionally[19]. Pre-mRNAs contain critical cis-elements, including splice site (SS) consensus sequences (5'SS, 3'SS, branch point, and polypyrimidine tract), exonic splicing enhancers/silencers (ESEs/ESSs), and intronic splicing enhancers/silencers (ISEs/ISSs) (Fig.1). The cis-acting components listed above are referred to as splicing code, and they all have a substantial impact on splice-site selection and AS consequences[20,21]. First, the recognized efficiency of splicing factors is based on how intense the SS consensus sequences are[22]. If the SSs or splicing code sequences are altered, their relative strengths will change. This might end up causing the spliceosome to recognize new SSs and change how pre-mRNA has been spliced[23]. Secondly, flanking pre-mRNA regulatory sequences (ESEs/ESSs and ISEs/ISSs) served as the binding site for different splicing factors[20]. Among them, SR proteins are responsible for binding ESEs and ISEs whereas heterogeneous nuclear ribonucleoproteins (hnRNPs) mainly bind ESSs and ISSs. Binding the flanking pre-mRNA regulatory sequences by these SFs can help snRNPs choose the right SS. Meanwhile, these different types of SFs and cis-acting elements show antagonistic functions in selecting splice sites[22].

      Figure 1. 

      Pre-mRNA splicing process of plants. In plants, most coding sequences of genes are interrupted by introns. Transcription factors (TFs) recognize promoters and guide the expression of genes. The gDNA is transcribed by RNA polymerase-II (Pol-II) to produce pre-mRNA with both exons and introns. Pre-mRNA owns critical splicing codes, including 5'SS, 3'SS, branch point, polypyrimidine tract, ESEs/ESSs, and ISEs/ISSs. SFs are responsible for recognizing splicing codes. The active spliceosome removes introns, ligating exons of pre-mRNA selectively to produce splicing variants. Stresses have significant effects on AS through SFs. Different splicing variants will have multiple molecular consequences. Firstly, the nuclear pore complex (NPC) will prevent certain intron-containing mRNAs leakage into the cytoplasm and leading to nuclear retention events. Secondly, mRNA transcripts containing premature translation-termination codons (PTCs) with 3'-untranslated regions (3'UTRs) (≥ 350 nt) can be degraded by nonsense-mediated RNA decay (NMD). Finally, the remaining mRNA transcripts will be translated into different spliced proteins by ribosome with different translation efficiency due to the presence of (upstream open reading frames) uORFs.

      Numerous splicing factors recognize distinct splicing codes and can assemble a highly dynamic spliceosome complex composed of five core uridine-rich snRNPs (U1, U2, U4, U5, and U6), hundreds of small nuclear RNA (snRNA) genes, and accessory proteins[22]. Sequence comparison and motif searches in Arabidopsis identified 395 splicing-related proteins encoding genes[14]. Splicing-related proteins are referred to as splicing factors in this context. Most of them are different types of RNA-binding proteins responsible for recognizing the splicing code on pre-mRNA[21]. For example, pre-mRNA-processing factor 8 (Prp8) was a scaffolding protein of the U5 snRNP that regulates 5'SS and 3'SS utilization. It shares a multidomain structure characterized by one RNA recognition motif (RRMs), which bind to the pre-mRNA targets, and two U6/5-snRNA binding sites involved in snRNA–protein interactions[24]. More and more splicing variants and SFs have been discovered as crucial for various stress responses in Arabidopsis and rice, but vegetable crops have achieved modest progress in this field.

      The modes of alternative splicing are complicated and diverse, including exon skipping (ES), intron retention (IR), etc. (Fig. 2a). Among them, IR is the most prevalent. More than 51% of genes producing AS transcripts show IR events in Arabidopsis, but in metazoans, ES is prevalent[25,26]. Earlier studies revealed that the majority of splicing variants induced by stress were ineffective[27] . Yet, increasing evidence suggests that mRNA splicing variants perform active functions in response to stressful conditions. At first, the nuclear pore complex (NPC) will prevent intron-containing splicing variants from leakage into the cytoplasm[28] . Usually, these splicing variants in the nucleus might act as the non-coding RNA rather than coding proteins. The activating signal cointegrator complex 3 (ASCC3) gene, for example, can be spliced into coding and non-coding transcript isoforms with antagonistic activities in regulating transcription recovery following UV-induced DNA damage[29]. Secondly, splicing variants containing premature translation-termination codons (PTCs) with 3'UTRs (≥ 350 nt) can be degraded by nonsense-mediated RNA decay (NMD)[30]. Then, the splicing variants in the cytoplasm will be translated into different spliced proteins (Fig. 2b). Furthermore, if AS occurs in 5'UTR, it will not affect the coding sequence but change the presence of uORF (Fig.1 and Fig. 2b). uORF is widely present in the 5'leader sequences of 30–50% of eukaryotic mRNAs. The translation of uORF can greatly repress the translation efficiency of the downstream major coding sequences[31].

      Figure 2. 

      (a) Major types of AS events in plants include constitutive splicing (CS), intron retention (IR), exon skipping (ES), alternative 5' splice sites (A5SS), alternative 3' splice sites (A3SS), mutually exclusive splicing (MES), and exitron splicing (EIS). Usually, an mRNA splicing variant is formed by one or more AS events. (b) Different AS event combinations lead to generate multiple splicing variants from one gene locus. The subcellular location, stability, translation efficiency, and coding sequence of splicing variants are different.

    • Light and temperature are the most important environmental factors regulating plant growth and development. Key factors in the light signaling pathway, including PIF3, PIF6, ELONGATED HYPOCOTYL5 (HY5), suppressor of phyA-105 3(SPA3), and cryptochromes 1 (CRY1), are proven to suffer AS[5]. PIF3 was extensively investigated among them, and red light greatly increased amounts of the intron retained in PIF3 5'UTR. A uORF contained within this intron caused an 80% reduction in PIF3 protein levels[32]. In addition to PIF3, AS pattern change of PIF6 contributes to seed dormancy response[33,34], and the selection of alternative 5’SS within SPA3 intron 4 leads to increased truncated nonfunctional SPA3 protein[12]. Interestingly, phytochromes (PHYB and PHYA) are responsible for light-dependent AS of SPA3, PIF3, and PIF6[12,35]. The roles of these phytochromes in AS will be discussed later.

      Heat-shock transcription factors (HSFs) are vital for plant heat stress tolerance. HsfA2, a key regulator in response to heat stress in Arabidopsis, is also heat stress-induced spliced (a cryptic splice site in the intron) to produce a small truncated HsfA2 isoform with important function under heat stress conditions[36,37]. Another type of transcription factor bZIP60 involved in heat stress tolerance by controlling unfolded protein response[3842] , and its mRNA can be alternatively spliced by inositol requiring enzyme1 (IRE1) to produce two transcripts encoding different functional spliced proteins[43]. Moreover, flower time and fertility could also be greatly affected by temperature variation. FLOWERING LOCUS M (FLM) splicing variants play key roles in flower time. The FLM-β is predominately formed at low temperatures and prevents precocious flowering. By contrast, the FLM-δ encodes a truncated protein with impaired DNA binding and acts as a dominant-negative activator of flowering at higher temperatures[44]. Temperature-sensitive genic male sterility (TGMS) line is important for the breeding of hybrid crops. The splicing of CALLOSE SYNTHASE5 (CalS5) pre-mRNA is involved in controlling TGMS[45,46]. So, light and temperature change significantly affect AS patterns, which are essential for plant stress adaption.

    • Light dramatically affects alternative splicing[6,47]. Critical genes of light signaling pathways in plants, including PHYB, PHYA, CRY2, and CONSTITUTIVELY PHOTOMORPHOGENIC 1 (COP1), act similarly to SFs in determining AS (Fig. 3 & Table 1). Among them, PHYB regulates photomorphogenesis, thermomorphogenesis, and freezing tolerance[48,49]. PHYB could also directly interact with SPLICING FACTOR FOR PHYTOCHROME SIGNALING (SFPS) and red-light responses in CRY1CRY2 background1 (RRC1) in nuclear speckles to synergistically regulate pre-mRNA splicing of a group of genes involved in light signaling and circadian clock pathways[12,35,5052]. Furthermore, PHYA regulates AS via interaction with the Arabidopsis homologue of human CNOT9 (NOT9B)[53]. During photomorphogenesis, COP1 can regulate AS of U2 small nuclear ribonucleoprotein auxiliary factor 65A (U2AF65A) and SR30 by interacting with U2AF65-associated protein (UAP56)[54]. Hence, we can conclude that these proteins in light signaling pathways can directly regulate SFs in order to modulate AS in response to changes in the light environment.

      Figure 3. 

      The major pathway for regulating pre-mRNA AS under light and temperature stresses. At first, some SFs could act as sensors to perceive the changes in light and temperatures. Secondly, those sensors will phosphorylate or interact with other SFs and change their location and activity. These modifications of SFs will affect AS by controlling spliceosome function. Finally, plant growth, development, flower, fertility, and tolerance are regulated under adverse light and temperature conditions.

      Table 1.  Splicing factors involved in light and temperature stress adaption.

      Annotation Name Species Gene ID Function Reference
      DEAD-box RNA heatlicases RH42 Rice Os08G0159900 Cold Tolerance Lu et al.[65]
      U1 snRNP Prp39a Arabidopsis AT1G04080 Temperature-induced flowering time Chang et al.[73]
      hnRNPs RBP45d Arabidopsis AT5G19350 Temperature-induced flowering time Chang et al.[73]
      Phytochrome CRY2 Arabidopsis AT1G04400 Blue-light regulation of thermosensory flowering Zhao et al.[74]
      mRNA binding SF1/BBP Arabidopsis AT5G51300 Temperature-induced flowering time Lee et al.[71]
      hnRNP A/B family UBA2c Arabidopsis AT3G15010 Light and temperature-induced flowering time Zhao et al.[72]
      Core proteins (Sm) SmE-b/PCP/SME1 Arabidopsis AT2G18740 Arrest of growth and male sterile at
      low temperature
      Capovilla et al.[17]
      U1 snRNP Luc7a Arabidopsis AT3G03340 Low temperature resulted in arrest
      of growth
      Amorim et al.[63]
      U1 snRNP Luc7b Arabidopsis AT5G17440 Low temperature resulted in arrest
      of growth
      Amorim et al.[63]
      U1 snRNP Luc7-rl Arabidopsis AT5G51410 Low temperature resulted in arrest
      of growth
      Amorim et al.[63]
      Glycine rich protein GRPs-7 Arabidopsis AT2G30560 Cold tolerance Kim et al.[62]
      DEAD box RNA helicase Prp5-1b/RCF1 Arabidopsis AT1G20920 Cold tolerance and cold acclimion Guan et al.[59]
      U5 snRNP U5-200-2b/Brr2b Arabidopsis AT2G42270 Cold tolerance and cold acclimion Guan et al.[59]
      DEAD-box RNA helicase U2B"a/U2B"-LIKE Arabidopsis AT1G06960 Cold tolerance and cold acclimion Calixto et al.[64]
      SR protein kinase AFC3/AME3 Arabidopsis AT4G32660 Cold tolerance and cold acclimion L. Savitch, unpublisheatd data
      17S U2 snRNP GEMIN2 Arabidopsis AT1G54380 Cold tolerance and cold acclimion Schlaen et al.[61]
      U5 snRNP U5-102KD/STA1 Arabidopsis AT4G03430 Cold tolerance and cold acclimion;
      heat tolerance and heat acclimion
      Kim et al.[66];
      Guan et al.[59]
      Inositol-requiring enzyme-1 IRE1b Arabidopsis AT5G24360 Heat stress responses Deng et al.[39]; Moreno et al.[38]
      Inositol-requiring enzyme-1 IRE1a Arabidopsis AT2G17520 Heat stress responses Deng et al.[39]; Moreno et al.[38]
      hnRNP HOS5/RCF3 Arabidopsis AT5G53060 Heat stress tolerance Guan et al.[60]
      HSP70-3 HSP70-3 Arabidopsis AT3G09440 Heat stress tolerance Song et al.[56]
      HSP70-4 HSP70-4 Arabidopsis AT3G12580 Heat stress tolerance Wang et al.[57]
      HSC70-1 HSC70-1 Arabidopsis AT5G02500 Heat stress tolerance Tiwari et al.[58]
      Cyclin-dependent kinase CDCL2p110-1 Arabidopsis AT1G67580 High temperature-induced male fertility defects Cavallari et al.[70]; Zheng et al.[69]; Nibau et al.[68]
      Cyclin-dependent kinase CDCL2p110-2/CDKG1 Arabidopsis AT5G63370 High temperature induce fertility defects Cavallari et al.[70]; Zheng et al.[69]; Nibau et al.[68]
      SR protein Kinase AFC2 Arabidopsis AT4G24740 Thermosensors and thermomorphogenesis Lin et al.[88]
      Phytochrome PHYB Arabidopsis AT2G18790 Thermosensors and thermomorphogenesis;
      photoreceptors and photomorphogenesis
      Jung et al.[48]; Shikata et al.[12,51]; Hartmann et al.[52]; Xin et al.[35,50]
      Phytochrome PHYA Arabidopsis AT1G09570 Photomorphogenesis Shikata et al.[12]; Schwenk et al.[53]
      E3 ubiquitin ligase COP1 Arabidopsis AT2G32950 Photomorphogenesis Li et al.[54]
      17S U2 associated proteins SR140-1/RRC1 Arabidopsis AT5G25060 Photomorphogenesis Shikata et al.[12,51]; Hartmann et al.[52]; Xin et al.[35,50]

      SFs are also discovered to play a critical role in regulating plant temperature stress tolerance. Heat shock proteins (HSP70s), as a spliceosome component, are involved in the repair of AS during heat stress[55]. In Arabidopsis, HSP70-3, HSP70-4, and heat shock cognate protein 70–1 (HSC70–1) are all required for heat stress tolerance[5658]. Apart from heat shock protein, a nuclear-localized SF with a K homology (KH) domain, REGULATOR OF CBF GENE EXPRESSION 3 (RCF3) serves as a negative regulator in heat stress tolerance[59,60]. For cold stress, SFs including RCF1, U5 small nuclear ribonucleoprotein helicase (Brr2b)[59],Gemin2[61], GRP7[62], the LAMMER kinase 3 (AME3), LETHAL UNLESS CBC 7 (LUC7)[63], and U2 small nuclear ribonucleoprotein particle (snRNP)-specific proteins (U2B”-LIKE)[64] play important roles in cold acclimation and tolerance in Arabidopsis. In rice, DEAD-box RNA Helicase42 (OsRH42) modulates pre-mRNA splicing at low temperatures by directly interacting with U2 small nuclear RNA. OsRH42-knockdown and OsRH42 overexpression dramatically affect the pre-mRNA splicing pattern, results in retarded plants growth and impaired cold tolerance[65]. Interestingly, the U5-snRNP-interacting protein homolog STABILIZED1 (STA1) is necessary for both cold and heat stress tolerance (Fig. 3 & Table 1). STA1 acts as a high temperature-induced SF for controlling AS of HSFA3[66]. Cold stress has also been shown to enhance STA1 gene expression, and the sta1-1 mutant is defective in splicing the cold-induce gene COR15[67]. Thus, the preceding studies indicate that the majority of SFs positively influence temperature stress tolerance.

      Under adverse temperature conditions, SFs can also have a significant impact on plant development, flowering, and fertility. Sm protein E1 (SME1), for instance, is a temperature-specific SF in Arabidopsis, and mis-splicing WUSCHEL and CLAVATA3 rendered sme1 mutants growth and male sterility[17]. Similarly, CDKG1 regulates fertility under temperature stress and can interact with splicing factor Arginine/serine-rich zinc knuckle-containing protein33 (RSZ33), U2AF65A, and RS-containing zinc finger protein22 (SRZ22) to facilitate the splicing of the sixth intron of CalS5, consequently modifying pollen wall development (Fig. 3)[68,69]. CDKG1 can also be spliced into two transcript variants (CDKG1L and CDKG1S), and the ratio between long and short forms is regulated by temperature and CDKG2[46,70]. Furthermore, SFs are primarily engaged in flowering time control. In response to temperature stress, AtSF1[71], UBIQUITIN-SPECIFIC PROTEASE 1-ASSOCIATED PROTEIN 2C (UBA2c)[72], RNA binding protein 45d (RBP45d)[73], CRY2[74], U2AF65A[74], CRY2 INTERACTING SPLICING FACTOR 1 (CIS1)[74], PRP39a, CYCLYN L1 (CYCL1), and CDKG2[68], all of them can govern the AS of FLM to fine-tune flowering time. Among them, CDKG2 together with its cognate CYCL1 controls the AS of FLM. CRY2 can promote CIS1 binding to pre-mRNA of FLM, then CIS1 interact with SF1 and U2AF65A to recruit U2-snRNP to control AS of FLM. Thus, the CRY2–CIS1–U2AF65A/SF1–FLM signaling pathway shows the specific roles of AS under light and temperature stresses (Fig. 3).

    • The development of essential vegetable traits including yield, shape, color, nutritional value, flower, and fruit, is usually dependent on strict temperature and light requirements. Adverse light and temperature will produce low-quality vegetable crops, drastically reducing their economic and nutritional value[75]. For instance, low light and high temperature stress will lead to the formation of slender seedlings with a long stem, smaller leaf, and short root. Thus, these seedlings typically have lower stress tolerance, are susceptible to lodging, and difficult to transplant (Fig. 4a). In order to ensure a stable supply of vegetables and fruits during the off-season, protected fruit trees and vegetables are generally cultivated in early spring and late autumn. But still, low temperatures and low light stress are prevalent during this season, resulting in significant reductions in agricultural yields and quality. Temperature stress, in particular for fruit development, can considerably alter the accumulation level of fruit sugar content, and the quality and intensity of light have a substantial impact on the color formation[7678].

      Figure 4. 

      Expression patterns analysis of the HSP70 gene family members in cucurbit crops. (a) Low light or high temperature stress causes slender cucumber seedlings in the breeding greenhouse. (b) The phylogenetic tree of the HSP70 gene family from cucumber, pumpkin, watermelon, tomato, melon, and Arabidopsis. (c) Heatmap of melon HSP70 gene family after 8 h and 24 h cold treatment. (d) Gene expression patterns of the pumpkin HSP70 gene family after 24 h cold treatment.

      AS has been shown to be an efficient post-transcriptional regulatory mechanism in Arabidopsis for coping with fluctuating light and temperature stressors[6,7,12,47,79]. However, the distinctive roles of SFs and AS in vegetable crops have not been thoroughly investigated under light and temperature constraints. Vegetable crops have only published a couple of research results associated to AS regulation during light and temperature stress in recent times. In cucumber, AS of CsGA2ox8 is affected by high light to control hypocotyl elongation[80]. In lily, heat stress leads LlHSFA3B intron intention to produce the splicing variant LlHSFA3B-III with PTCs, which could increase plant tolerance of heat stress by encoding a truncated protein[11]. In tomato, it was found that intronic polymorphisms of SlHsfA2 affected the splicing efficiency and heat stress tolerance[81]. The findings demonstrate that AS is also crucial in the response of vegetable crops to light and temperature stressors.

    • In Arabidopsis, 395 protein-coding genes involved in AS are identified and classified into five categories (91 small nuclear ribonucleoproteins, 109 splicing factors, 60 splicing regulation proteins, 84 novel spliceosome proteins, and 51 possible splicing related proteins)[14]. Many splicing factors have been shown to be essential for stress tolerance. However, they have not been identified and analyzed effectively in most horticultural crops. Low light or high temperature stress are usual in the greenhouse when seedlings are being raised. It can cause slender seedlings with long hypocotyl and lower lodging resistance, especially in cucurbit crops (Fig. 4a). To find SFs involved in seedlings, we take vegetable crops (cucumber, pumpkin, watermelon, tomato, melon) as an example to carry out blastp search. One thousand seven hundred and eight SF protein sequences were retrieved and classified, including cucumber 326, melon 304, pumpkin 429, watermelon 303, and tomato 346 (Supplemental Table S1). The following has established a solid platform for us to investigate the role of splicing factors under light and temperature stresses. We identified that temperature stress had a significant effect on the expression of HSP70s by examining the expression patterns of all the SFs genes under light and temperature stresses. Hence, we systematically identified the HSP70 gene family and constructed the phylogenetic tree and heatmap (Fig. 4b). Surprisingly, low temperature stress substantially induced the majority of HSP70s gene family members, with HSP70−4 the most distinctive among them (Fig. 4c & d). HSP70 is a spliceosome component that plays a vital role in the repair of AS during temperature stress[55]. In cucurbit crops, its members are doubled, including two homologs in melon (MELO3C021658 and MELO3C021640), four homologous genes in pumpkin (CmoCh04G027550, CmoCh04G027600, CmoCh15G004100, and CmoCh15G004130), and all of which are highly temperature-responsive (Fig. 4b−d). Unfortunately, their involvement in controlling AS and adaptation to the light and temperature stressors of vegetable crops are unclear and require extensive research.

    • Traditionally, most studies only examine at one major splicing variant, and it is also presumed that the splicing variant of interest can be translated into proteins. One gene can produce several or even dozens of splicing variants through AS. Yet, only some splicing variants can be translated into distinct proteins. Other splicing variants without coding capacities maybe act as active non-coding RNA (Fig. 1). It is not enough to explain the gene function by analyzing one or two splicing variants with coding capacities. In plants, splicing factors and splicing codes are two major intrinsic factors in the control of AS. Many studies have proven that DNA polymorphism in splicing code regions could lead to changes in AS and phenotype[8286]. SFs are another factor controlling AS, and there is a complex hierarchical regulatory relationship between SFs (Fig. 3). SFs including receptor-like kinase receptor FERONIA (FER) and LAMMER kinase 2(AFC2) could act as sensors to detect changes in suboptimal light and temperature and subsequently phosphorylate or interact with other SFs (GRP7 and RSZ21) to alter their location or activity in regulation of AS of Abscisic acid response element binding factor 1 (ABF1) and PIF4[8789]. Furthermore, the interaction between SFs and splicing codes cannot be neglected since SFs can also influence splicing fidelity by restricting the usage of novel splicing code[23]. Consequently, a thorough investigation of splicing codes, splicing factors, and their effects on AS under stress will allow us to better understand and exploit it in the future for vegetable crops.

      CRISPR-Cas9 is the most advanced genome editing approach, and it has been extensively used in precision breeding of important cereals and vegetable crops[90,91]. AS modulates many thermomorphogenesis and photomorphogenesis genes in Arabidopsis, including CRY1, PHYB, and EARLY FLOWERING 3 (ELF3). They are also important in modulating seedling elongation under low light or high temperature stresses[92]. Yet, knockout of these genes by CRISPR/Cas9 will make plants extremely susceptible to low light and high temperature. They will also cause undesirable alterations in plant growth and development, including flowering timing, fruit set, seed germination, and so on[93]. Whether we can edit their splicing code (intron or UTR region) or splicing factors to manipulate AS of these genes by CRISPR/Cas9 to produce critical splicing variants for inhibiting excessive elongation but without other undesired effects? Thus, we propose an application potential of this AS mechanism for vegetable crops in (Fig. 5) and hope that this research strategy will promote studies on AS regulation in combination with gene editing technology under light and temperature stresses, which can be used in the future for the genetic improvement of vegetable crops.

      • This research was supported by Guangdong Basic and Applied Basic Research Foundation (2021A1515110515), the Innovation Fund Projects of Guangdong Academy of Agricultural Sciences (202208 and 202149), the Science and Technology Project of Zhaoqing City (2021N004), and the Guidance Project for Young People of Guangdong Academy of Agricultural Sciences (R2020QD-064).

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

      • Copyright: © 2023 by the author(s). Published by Maximum Academic Press, Fayetteville, GA. This article is an open access article distributed under Creative Commons Attribution License (CC BY 4.0), visit https://creativecommons.org/licenses/by/4.0/.
    Figure (4)  Table (1) References (93)
  • About this article
    Cite this article
    Cao H, Wu T, Shi L, Li Y, Zhang C. 2023. Alternative splicing control of light and temperature stress responses and its prospects in vegetable crops. Vegetable Research 3:17 doi: 10.48130/VR-2023-0017
    Cao H, Wu T, Shi L, Li Y, Zhang C. 2023. Alternative splicing control of light and temperature stress responses and its prospects in vegetable crops. Vegetable Research 3:17 doi: 10.48130/VR-2023-0017

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return