Search
2023 Volume 3
Article Contents
REVIEW   Open Access    

Molecular mechanisms of flowering phenology in trees

More Information
  • Flower initiation is a phenological developmental process strictly regulated in all flowering plants. Studies in Arabidopsis thaliana, a model plant organism in plant biology and genetics, and major cereal crops have provided fundamental knowledge and understanding of the underlying molecular mechanisms and regulation in annuals. However, this flowering process and underly molecular mechanisms in perennials are much more complicated than those in annuals and remain poorly understood and documented. In recent years, the increasing availability of perennial plant genomes and advances in biotechnology have allowed the identification and characterization of flowering-associated gene orthologs in perennials. In this review, we compared and summarized the recent progress in regulation of flowering time in perennial trees, with an emphasis on the perennial-specific regulatory mechanisms. Pleiotropic effects on tree growth habits such as juvenility, seasonal activity–dormancy growth, and the applications of tree flowering phenology are discussed.
  • 加载中
  • [1]

    Gaudinier A, Blackman BK. 2020. Evolutionary processes from the perspective of flowering time diversity. New Phytologist 225:1883−98

    doi: 10.1111/nph.16205

    CrossRef   Google Scholar

    [2]

    Andrés F, Coupland G. 2012. The genetic basis of flowering responses to seasonal cues. Nature Reviews Genetics 13:627−39

    doi: 10.1038/nrg3291

    CrossRef   Google Scholar

    [3]

    Bouché F, Lobet G, Tocquin P, Périlleux C. 2016. FLOR-ID: an interactive database of flowering-time gene networks in Arabidopsis thaliana. Nucleic Acids Research 44:D1167−D1171

    doi: 10.1093/nar/gkv1054

    CrossRef   Google Scholar

    [4]

    Ó'Maoiléidigh DS, Graciet E, Wellmer F. 2014. Gene networks controlling Arabidopsis thaliana flower development. New Phytologist 201:16−30

    doi: 10.1111/nph.12444

    CrossRef   Google Scholar

    [5]

    Turck F, Fornara F, Coupland G. 2008. Regulation and identity of florigen: FLOWERING LOCUS T moves center stage. Annual Review of Plant Biology 59:573−94

    doi: 10.1146/annurev.arplant.59.032607.092755

    CrossRef   Google Scholar

    [6]

    Song YH, Shim JS, Kinmonth-Schultz HA, Imaizumi T. 2015. Photoperiodic flowering: time measurement mechanisms in leaves. Annual Review of Plant Biology 66:441−64

    doi: 10.1146/annurev-arplant-043014-115555

    CrossRef   Google Scholar

    [7]

    Cao S, Luo X, Xu D, Tian X, Song J, et al. 2021. Genetic architecture underlying light and temperature mediated flowering in Arabidopsis, rice, and temperate cereals. New Phytologist 230:1731−45

    doi: 10.1111/nph.17276

    CrossRef   Google Scholar

    [8]

    Hill CB, Li C. 2016. Genetic architecture of flowering phenology in cereals and opportunities for crop improvement. Frontiers in Plant Science 7:1906

    doi: 10.3389/fpls.2016.01906

    CrossRef   Google Scholar

    [9]

    Battey NH. 2000. Aspects of seasonality. Journal of Experimental Botany 51:1769−80

    doi: 10.1093/jexbot/51.352.1769

    CrossRef   Google Scholar

    [10]

    Chalupka W, Cecich RA. 1997. Control of the first flowering in forest trees. Scandinavian Journal of Forest Research 12:102−11

    doi: 10.1080/02827589709355390

    CrossRef   Google Scholar

    [11]

    Albani MC, Coupland G. 2010. Comparative analysis of flowering in annual and perennial plants. Current Topics in Developmental Biology 91:323−48

    doi: 10.1016/S0070-2153(10)91011-9

    CrossRef   Google Scholar

    [12]

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

    doi: 10.1093/jxb/eru246

    CrossRef   Google Scholar

    [13]

    Wang J, Czech B, Weigel D. 2009. miR156-regulated SPL transcription factors define an endogenous flowering pathway in Arabidopsis thaliana. Cell 138:738−49

    doi: 10.1016/j.cell.2009.06.014

    CrossRef   Google Scholar

    [14]

    Wu G, Park MY, Conway SR, Wang J, Weigel D, et al. 2009. The sequential action of miR156 and miR172 regulates developmental timing in Arabidopsis. Cell 138:750−59

    doi: 10.1016/j.cell.2009.06.031

    CrossRef   Google Scholar

    [15]

    Yant L, Mathieu J, Dinh TT, Ott F, Lanz C, et al. 2010. Orchestration of the floral transition and floral development in Arabidopsis by the bifunctional transcription factor APETALA2. Plant Cell 22:2156−70

    doi: 10.1105/tpc.110.075606

    CrossRef   Google Scholar

    [16]

    Mathieu J, Yant LJ, Mürdter F, Küttner F, Schmid M. 2009. Repression of flowering by the miR172 target SMZ. PLoS Biology 7:e1000148

    doi: 10.1371/journal.pbio.1000148

    CrossRef   Google Scholar

    [17]

    Yamaguchi A, Wu MF, Yang L, Wu G, Poethig RS, et al. 2009. The microRNA-regulated SBP-box transcription factor SPL3 is a direct upstream activator of LEAFY, FRUITFULL, and APETALA1. Developmental Cell 17:268−78

    doi: 10.1016/j.devcel.2009.06.007

    CrossRef   Google Scholar

    [18]

    Jung JH, Lee HJ, Ryu JY, Park CM. 2016. SPL3/4/5 Integrate developmental aging and photoperiodic signals into the FT-FD module in Arabidopsis flowering. Molecular Plant 9:1647−59

    doi: 10.1016/j.molp.2016.10.014

    CrossRef   Google Scholar

    [19]

    Wang H, Wang H. 2015. The miR156/SPL module, a regulatory hub and versatile toolbox, gears up crops for enhanced agronomic traits. Molecular Plant 8:677−88

    doi: 10.1016/j.molp.2015.01.008

    CrossRef   Google Scholar

    [20]

    Wang JW, Park MY, Wang LJ, Koo Y, Chen XY, et al. 2011. miRNA control of vegetative phase change in trees. PLoS Genetics 7:e1002012

    doi: 10.1371/journal.pgen.1002012

    CrossRef   Google Scholar

    [21]

    Li H, Luo Y, Ma B, Hu J, Lv Z, et al. 2021. Hierarchical Action of Mulberry miR156 in the Vegetative Phase Transition. International Journal of Molecular Sciences 22:5550

    doi: 10.3390/ijms22115550

    CrossRef   Google Scholar

    [22]

    Ahsan MU, Hayward A, Irihimovitch V, Fletcher S, Tanurdzic M, et al. 2019. Juvenility and Vegetative Phase Transition in Tropical/Subtropical Tree Crops. Frontiers in Plant Science 10:729

    doi: 10.3389/fpls.2019.00729

    CrossRef   Google Scholar

    [23]

    Xing L, Zhang D, Li Y, Zhao C, Zhang S, et al. 2014. Genome-wide identification of vegetative phase transition-associated microRNAs and target predictions using degradome sequencing in Malus hupehensis. BMC Genomics 15:1125

    doi: 10.1186/1471-2164-15-1125

    CrossRef   Google Scholar

    [24]

    Lawrence EH, Springer CJ, Helliker BR, Poethig RS. 2021. MicroRNA156-mediated changes in leaf composition lead to altered photosynthetic traits during vegetative phase change. New Phytologist 231:1008−22

    doi: 10.1111/nph.17007

    CrossRef   Google Scholar

    [25]

    Lawrence EH, Leichty AR, Doody EE, Ma C, Strauss SH, et al. 2021. Vegetative phase change in Populus tremula x alba. New Phytologist 231:351−64

    doi: 10.1111/nph.17316

    CrossRef   Google Scholar

    [26]

    Niu S, Yuan H, Sun X, Porth I, Li Y, et al. 2016. A transcriptomics investigation into pine reproductive organ development. New Phytologist 209:1278−89

    doi: 10.1111/nph.13680

    CrossRef   Google Scholar

    [27]

    Shalom L, Shlizerman L, Zur N, Doron-Faigenboim A, Blumwald E, et al. 2015. Molecular characterization of SQUAMOSA PROMOTER BINDING PROTEIN-LIKE (SPL) gene family from Citrus and the effect of fruit load on their expression. Frontiers in Plant Science 6:389

    doi: 10.3389/fpls.2015.00389

    CrossRef   Google Scholar

    [28]

    Jiang Y, Peng J, Wang M, Su W, Gan X, et al. 2020. The Role of EjSPL3, EjSPL4, EjSPL5, and EjSPL9 in Regulating Flowering in Loquat (Eriobotrya japonica Lindl.). International Journal of Molecular Sciences 21:248

    doi: 10.3390/ijms21010248

    CrossRef   Google Scholar

    [29]

    Zhou Y, Gan X, Viñegra de la Torre N, Neumann U, Albani MC. 2021. Beyond flowering time: diverse roles of an APETALA2-like transcription factor in shoot architecture and perennial traits. New Phytologist 229:444−59

    doi: 10.1111/nph.16839

    CrossRef   Google Scholar

    [30]

    Song Y, Ito S, Imaizumi T. 2013. Flowering time regulation: photoperiod- and temperature-sensing in leaves. Trends in Plant 18:575−83

    doi: 10.1016/j.tplants.2013.05.003

    CrossRef   Google Scholar

    [31]

    Kobayashi Y, Kaya H, Goto K, Iwabuchi M, Araki T. 1999. A pair of related genes with antagonistic roles in mediating flowering signals. Science 286:1960−2

    doi: 10.1126/science.286.5446.1960

    CrossRef   Google Scholar

    [32]

    Pin PA, Nilsson O. 2012. The multifaceted roles of FLOWERING LOCUS T in plant development. Plant, Cell & Environment 35:1742−55

    doi: 10.1111/j.1365-3040.2012.02558.x

    CrossRef   Google Scholar

    [33]

    Wickland DP, Hanzawa Y. 2015. The FLOWERING LOCUS T/TERMINAL FLOWER 1 gene family: Functional evolution and molecular mechanisms. Molecular Plant 8:983−97

    doi: 10.1016/j.molp.2015.01.007

    CrossRef   Google Scholar

    [34]

    Hsu CY, Liu Y, Luthe DS, Yuceer C. 2006. Poplar FT2 shortens the juvenile phase and promotes seasonal flowering. The Plant Cell 18:1846−61

    doi: 10.1105/tpc.106.041038

    CrossRef   Google Scholar

    [35]

    Endo T, Shimada T, Fujii H, Kobayashi Y, Araki T, et al. 2005. Ectopic expression of an FT homolog from citrus confers an early flowering phenotype on trifoliate orange (Poncirus trifoliata L. Raf.). Transgenic Research 14:703−12

    doi: 10.1007/s11248-005-6632-3

    CrossRef   Google Scholar

    [36]

    Li C, Luo L, Fu Q, Niu L, Xu ZF. 2014. Isolation and functional characterization of JcFT, a FLOWERING LOCUS T (FT) homologous gene from the biofuel Plant Jatropha curcas. BMC Plant Biology 14:125

    doi: 10.1186/1471-2229-14-125

    CrossRef   Google Scholar

    [37]

    Odipio J, Getu B, Chauhan RD, Alicai T, Bart R, et al. 2020. Transgenic overexpression of endogenous FLOWERING LOCUS T-like gene MeFT1 produces early flowering in cassava. PLoS One 15:e0227199

    doi: 10.1371/journal.pone.0227199

    CrossRef   Google Scholar

    [38]

    Sinn JP, Held JB, Vosburg C, Klee SM, Orbovic V, et al. 2021. Flowering Locus T chimeric protein induces floral precocity in edible citrus. Plant Biotechnology Journal 19:215−17

    doi: 10.1111/pbi.13463

    CrossRef   Google Scholar

    [39]

    Song G, Walworth A, Lin T, Chen Q, Han X, et al. 2019. VcFT-induced mobile florigenic signals in transgenic and transgrafted blueberries. Horticulture Research 6:105

    doi: 10.1038/s41438-019-0188-5

    CrossRef   Google Scholar

    [40]

    Srinivasan C, Dardick C, Callahan A, Scorza R. 2012. Plum (Prunus domestica) trees transformed with poplar FT1 result in altered architecture, dormancy requirement, and continuous flowering. PLoS One 7:e40715

    doi: 10.1371/journal.pone.0040715

    CrossRef   Google Scholar

    [41]

    Wenzel S, Flachowsky H, Hanke MV. 2013. The Fast-track breeding approach can be improved by heat-induced expression of the FLOWERING LOCUS T genes from poplar (Populus trichocarpa) in apple (Malus × domestica Borkh.). Plant Cell Tissue and Organ Culture 115:127−37

    doi: 10.1007/s11240-013-0346-7

    CrossRef   Google Scholar

    [42]

    Zhang H, Harry DE, Ma C, Yuceer C, Hsu CY, et al. 2010. Precocious flowering in trees: the FLOWERING LOCUS T gene as a research and breeding tool in Populus. Journal of Experimental Botany 61:2549−60

    doi: 10.1093/jxb/erq092

    CrossRef   Google Scholar

    [43]

    Tränkner C, Lehmann S, Hoenicka H, Hanke MV, Fladung M, et al. 2010. Over-expression of an FT-homologous gene of apple induces early flowering in annual and perennial plants. Planta 232:1309−24

    doi: 10.1007/s00425-010-1254-2

    CrossRef   Google Scholar

    [44]

    Böhlenius H, Huang T, Charbonnel-Campaa L, Brunner AM, Jansson S, et al. 2006. CO/FT regulatory module controls timing of flowering and seasonal growth cessation in trees. Science 312:1040−43

    doi: 10.1126/science.1126038

    CrossRef   Google Scholar

    [45]

    Kotoda N, Iwanami H, Takahashi S, Abe K. 2006. Antisense expression of MdTFL1, a TFL1-like gene, reduces the juvenile phase in apple. Journal of The American Society For Horticultural Science 131:74−81

    doi: 10.21273/JASHS.131.1.74

    CrossRef   Google Scholar

    [46]

    Mohamed R, Wang CT, Ma C, Shevchenko O, Dye SJ, et al. 2010. Populus CEN/TFL1 regulates first onset of flowering, axillary meristem identity and dormancy release in Populus. Plant Journal 62:674−88

    doi: 10.1111/j.1365-313X.2010.04185.x

    CrossRef   Google Scholar

    [47]

    Varkonyi-Gasic E, Wang T, Voogd C, Jeon S, Drummond RSM, et al. 2019. Mutagenesis of kiwifruit CENTRORADIALIS-like genes transforms a climbing woody perennial with long juvenility and axillary flowering into a compact plant with rapid terminal flowering. Plant Biotechnology Journal 17:869−80

    doi: 10.1111/pbi.13021

    CrossRef   Google Scholar

    [48]

    Lee J, Lee I. 2010. Regulation and function of SOC1, a flowering pathway integrator. Journal of Experimental Botany 61:2247−54

    doi: 10.1093/jxb/erq098

    CrossRef   Google Scholar

    [49]

    Ma J, Chen X, Song Y, Zhang G, Zhou X, et al. 2021. MADS-box transcription factors MADS11 and DAL1 interact to mediate the vegetative-to-reproductive transition in pine. Plant Physiology 187:247−62

    doi: 10.1093/plphys/kiab250

    CrossRef   Google Scholar

    [50]

    Wei J, Liu D, Liu G, Tang J, Chen Y. 2016. Molecular cloning, characterization, and expression of MiSOC1: A homolog of the flowering gene SUPPRESSOR OF OVEREXPRESSION OF CONSTANS1 from Mango (Mangifera indica L.). Frontiers in Plant Science 7:1758

    doi: 10.3389/fpls.2016.01758

    CrossRef   Google Scholar

    [51]

    Jiang Y, Peng J, Zhu Y, Su W, Zhang L, et al. 2019. The role of EjSOC1s in flower initiation in Eriobotrya japonica. Frontiers in Plant Science 10:253

    doi: 10.3389/fpls.2019.00253

    CrossRef   Google Scholar

    [52]

    Tan FC, Swain SM. 2007. Functional characterization of AP3, SOC1 and WUS homologues from citrus (Citrus sinensis). Physiologia Plantarum 131:481−95

    doi: 10.1111/j.1399-3054.2007.00971.x

    CrossRef   Google Scholar

    [53]

    Li G, Cao C, Yang H, Wang J, Wei W, et al. 2020. Molecular cloning and potential role of DiSOC1s in flowering regulation in Davidia involucrata Baill. Plant Physiology Biochemistry 157:453−59

    doi: 10.1016/j.plaphy.2020.11.003

    CrossRef   Google Scholar

    [54]

    Voogd C, Wang T, Varkonyi-Gasic E. 2015. Functional and expression analyses of kiwifruit SOC1-like genes suggest that they may not have a role in the transition to flowering but may affect the duration of dormancy. Journal of Experimental Botany 66:4699−710

    doi: 10.1093/jxb/erv234

    CrossRef   Google Scholar

    [55]

    Wang J, Gao Z, Li H, Jiu S, Qu Y, et al. 2020. Dormancy-associated MADS-box (DAM) genes influence chilling requirement of sweet cherries and co-regulate flower development with SOC1 gene. International Journal of Molecular Sciences 21:921

    doi: 10.3390/ijms21030921

    CrossRef   Google Scholar

    [56]

    Gómez-Soto D, Ramos-Sánchez JM, Alique D, Conde D, Triozzi PM, et al. 2021. Overexpression of a SOC1-related gene promotes bud break in ecodormant poplars. Frontiers in Plant Science 12:670497

    doi: 10.3389/fpls.2021.670497

    CrossRef   Google Scholar

    [57]

    Weigel D, Alvarez J, Smyth DR, Yanofsky MF, Meyerowitz EM. 1992. LEAFY controls floral meristem identity in Arabidopsis. Cell 69:843−59

    doi: 10.1016/0092-8674(92)90295-N

    CrossRef   Google Scholar

    [58]

    Blázquez MA, Soowal LN, Lee I, Weigel D. 1997. LEAFY expression and flower initiation in Arabidopsis. Development 124:3835−44

    doi: 10.1242/dev.124.19.3835

    CrossRef   Google Scholar

    [59]

    Rottmann WH, Meilan R, Sheppard LA, Brunner AM, Skinner JS, et al. 2000. Diverse effects of overexpression of LEAFY and PTLF, a poplar (Populus) homolog of LEAFY/FLORICAULA, in transgenic poplar and Arabidopsis. The Plant Journal 22:235−45

    doi: 10.1046/j.1365-313x.2000.00734.x

    CrossRef   Google Scholar

    [60]

    Weigel D, Nilsson O. 1995. A developmental switch sufficient for flower initiation in diverse plants. Nature 377:495−500

    doi: 10.1038/377495a0

    CrossRef   Google Scholar

    [61]

    Pena L, Martin-Trillo M, Juarez J, Pina JA, Navarro L, Martinez-Zapater JM. 2001. Constitutive expression of Arabidopsis LEAFY or APETALA1 genes in citrus reduces their generation time. Nat Biotechnology 19:263−67

    doi: 10.1038/85719

    CrossRef   Google Scholar

    [62]

    Wang Y, Yu H, He X, Lu T, Huang X, Luo C. 2022. Isolation and functional characterization of a LEAFY gene in mango (Mangifera indica L.). International Journal of Molecular Sciences 23:3974

    doi: 10.3390/ijms23073974

    CrossRef   Google Scholar

    [63]

    An L, Lei H, Shen X, Li T. 2012. Identification and Characterization of PpLFL, a Homolog of FLORICAULA/LEAFY in Peach (Prunus persica). Plant Molecular Biology Reporter 30:1488−95

    doi: 10.1007/s11105-012-0459-x

    CrossRef   Google Scholar

    [64]

    Ding F, Zhang S, Chen H, Peng H, Lu J, et al. 2018. Functional analysis of a homologue of the FLORICAULA/LEAFY gene in litchi (Litchi chinensis Sonn.) revealing its significance in early flowering process. Genes Genomics 40:1259−67

    doi: 10.1007/s13258-018-0739-4

    CrossRef   Google Scholar

    [65]

    Mandel MA, Yanofsky MF. 1995. A gene triggering flower formation in Arabidopsis. Nature 377:522−24

    doi: 10.1038/377522a0

    CrossRef   Google Scholar

    [66]

    Kaufmann K, Wellmer F, Muiño JM, Ferrier T, Wuest SE, et al. 2010. Orchestration of floral initiation by APETALA1. Science 328:85−89

    doi: 10.1126/science.1185244

    CrossRef   Google Scholar

    [67]

    Alejandra Mandel M, Gustafson-Brown C, Savidge B, Yanofsky MF. 1992. Molecular characterization of the Arabidopsis floral homeotic gene APETALA1. Nature 360:273−7

    doi: 10.1038/360273a0

    CrossRef   Google Scholar

    [68]

    Huang H, Wang S, Jiang J, Liu G, Li H, et al. 2014. Overexpression of BpAP1 induces early flowering and produces dwarfism in Betula platyphylla × Betula pendula. Physiologia Plantarum 151:495−506

    doi: 10.1111/ppl.12123

    CrossRef   Google Scholar

    [69]

    Moon J, Suh SS, Lee H, Choi KR, Hong CB, et al. 2003. The SOC1 MADS-box gene integrates vernalization and gibberellin signals for flowering in Arabidopsis. The Plant Journal 35:613−23

    doi: 10.1046/j.1365-313X.2003.01833.x

    CrossRef   Google Scholar

    [70]

    Blazquez MA, Green R, Nilsson O, Sussman MR, Weigel D. 1998. Gibberellins promote flowering of Arabidopsis by activating the LEAFY promoter. The Plant Cell 10:791−800

    doi: 10.1105/tpc.10.5.791

    CrossRef   Google Scholar

    [71]

    Yamaguchi N, Winter CM, Wu MF, Kanno Y, Yamaguchi A, et al. 2014. Gibberellin acts positively then negatively to control onset of flower formation in Arabidopsis. Science 344:638−41

    doi: 10.1126/science.1250498

    CrossRef   Google Scholar

    [72]

    Meilan R. 1997. Floral induction in woody angiosperms. New Forest 14:179−202

    doi: 10.1023/A:1006560603966

    CrossRef   Google Scholar

    [73]

    Bangerth KF. 2009. Floral induction in mature, perennial angiosperm fruit trees: Similarities and discrepancies with annual/biennial plants and the involvement of plant hormones. Scientia Horticulturae 122:153−63

    doi: 10.1016/j.scienta.2009.06.014

    CrossRef   Google Scholar

    [74]

    Williams DR, Ross JJ, Reid JB, Potts BM. 1999. Response of Eucalyptus nitens seedlings to gibberellin biosynthesis inhibitors. Plant Growth Regulation 27:125−29

    doi: 10.1023/A:1006176928861

    CrossRef   Google Scholar

    [75]

    Yuceer C, Kubiske ME, Harkess RL, Land SB. 2003. Effects of induction treatments on flowering in Populus deltoides. Tree Physiology 23:489−95

    doi: 10.1093/treephys/23.7.489

    CrossRef   Google Scholar

    [76]

    Hackett WP. 1985. Juvenility, Maturation, and Rejuvenation in Woody Plants. In Horticultural Reviews, ed. Janick J. 7: XI,569. US: John Wiley & Sons. pp. 109−55. https://doi.org/10.1002/9781118060735.ch3

    [77]

    Boss PK, Thomas MR. 2002. Association of dwarfism and floral induction with a grape 'green revolution' mutation. Nature 416:847−50

    doi: 10.1038/416847a

    CrossRef   Google Scholar

    [78]

    Pharis RP, Webber JE, Ross SD. 1987. The promotion of flowering in forest trees by gibberellin-A47 and Cultural treatments: A review of the possible mechanisms. Forest Ecology and Management 19:65−84

    doi: 10.1016/0378-1127(87)90012-0

    CrossRef   Google Scholar

    [79]

    Satake A, Nagahama A, Sasaki E. 2022. A cross-scale approach to unravel the molecular basis of plant phenology in temperate and tropical climates. New Phytologist 233:2340−53

    doi: 10.1111/nph.17897

    CrossRef   Google Scholar

    [80]

    Nagahama A, Yahara T. 2019. Quantitative comparison of flowering phenology traits among trees, perennial herbs, and annuals in a temperate plant community. American Journal of Botany 106:1545−1557

    doi: 10.1002/ajb2.1387

    CrossRef   Google Scholar

    [81]

    Singh KP, Kushwaha CP. 2006. Diversity of flowering and fruiting phenology of trees in a tropical deciduous forest in India. Annals of Botany 97:265−76

    doi: 10.1093/aob/mcj028

    CrossRef   Google Scholar

    [82]

    van Schaik CP, Terborgh JW, Wright SJ. 1993. The Phenology of Tropical Forests - Adaptive Significance and Consequences for Primary Consumers. Annual Review of Ecology and Systematics 24:353−77

    doi: 10.1146/annurev.es.24.110193.002033

    CrossRef   Google Scholar

    [83]

    Borchert R, Meyer SA, Felger RS, Porter-Bolland L. 2004. Environmental control of flowering periodicity in Costa Rican and Mexican tropical dry forests. Global Ecology and Biogeography 13:409−25

    doi: 10.1111/j.1466-822X.2004.00111.x

    CrossRef   Google Scholar

    [84]

    Butt N, Seabrook L, Maron M, Law BS, Dawson TP, et al. 2015. Cascading effects of climate extremes on vertebrate fauna through changes to low-latitude tree flowering and fruiting phenology. Global Change Biolology 21:3267−77

    doi: 10.1111/gcb.12869

    CrossRef   Google Scholar

    [85]

    Morellato LPC, Talora DC, Takahasi A, Bencke CC, Romera EC, et al. 2000. Phenology of Atlantic rain forest trees: A comparative study. Biotropica 32:811−23

    doi: 10.1111/j.1744-7429.2000.tb00620.x

    CrossRef   Google Scholar

    [86]

    Curran LM, Caniago I, Paoli GD, Astianti D, Kusneti M, et al. 1999. Impact of El Niño and logging on canopy tree recruitment in Borneo. Science 286:2184−8

    doi: 10.1126/science.286.5447.2184

    CrossRef   Google Scholar

    [87]

    Brearley FQ, Proctor J, Suriantata, Nagy L, Dalrymple G, et al. 2007. Reproductive phenology over a 10-year period in a lowland evergreen rain forest of central Borneo. Journal of Ecology 95:828−39

    doi: 10.1111/j.1365-2745.2007.01258.x

    CrossRef   Google Scholar

    [88]

    Grainger J. 1939. Studies upon the time of flowering of plants: Anatomical, floristic and phenological aspects of the problem. Annals of Applied Biology 26:684−704

    doi: 10.1111/j.1744-7348.1939.tb06994.x

    CrossRef   Google Scholar

    [89]

    Tan FC, Swain SM. 2006. Genetics of flower initiation and development in annual and perennial plants. Physiologia Plantarum 128:8−17

    doi: 10.1111/j.1399-3054.2006.00724.x

    CrossRef   Google Scholar

    [90]

    Tooke F, Battey NH. 2010. Temperate flowering phenology. Journal of Experimental Botany 61:2853−62

    doi: 10.1093/jxb/erq165

    CrossRef   Google Scholar

    [91]

    Brunner AM, Nilsson O. 2004. Revisiting tree maturation and floral initiation in the poplar functional genomics era. New Phytologist 164:43−51

    doi: 10.1111/j.1469-8137.2004.01165.x

    CrossRef   Google Scholar

    [92]

    Liang Q, Song K, Lu M, Dai T, Yang J, et al. 2022. Transcriptome and metabolome analyses reveal the involvement of multiple pathways in flowering intensity in mango. Frontiers in Plant Science 13:933923

    doi: 10.3389/fpls.2022.933923

    CrossRef   Google Scholar

    [93]

    Meng X, Li Y, Yuan Y, Zhang Y, Li H, et al. 2020. The regulatory pathways of distinct flowering characteristics in Chinese jujube. Horticulture Research 7:123

    doi: 10.1038/s41438-020-00344-7

    CrossRef   Google Scholar

    [94]

    Kudoh H. 2016. Molecular phenology in plants: in natura systems biology for the comprehensive understanding of seasonal responses under natural environments. New Phytologist 210:399−412

    doi: 10.1111/nph.13733

    CrossRef   Google Scholar

    [95]

    Chen Z, Rao P, Yang X, Su X, Zhao T, et al. 2018. A global view of transcriptome dynamics during male floral bud development in Populus tomentosa. Scientific Reports 8:722

    doi: 10.1038/s41598-017-18084-5

    CrossRef   Google Scholar

    [96]

    Fan Z, Li J, Li X, Wu B, Wang J, et al. 2015. Genome-wide transcriptome profiling provides insights into floral bud development of summer-flowering Camellia azalea. Scientific Reports 5:9729

    doi: 10.1038/srep09729

    CrossRef   Google Scholar

    [97]

    Hassankhah A, Rahemi M, Ramshini H, Sarikhani S, Vahdati K. 2020. Flowering in Persian walnut: patterns of gene expression during flower development. BMC Plant Biology 20:136

    doi: 10.1186/s12870-020-02372-w

    CrossRef   Google Scholar

    [98]

    Kobayashi MJ, Takeuchi Y, Kenta T, Kume T, Diway B, Shimizu KK. 2013. Mass flowering of the tropical tree Shorea beccariana was preceded by expression changes in flowering and drought-responsive genes. Molecular Ecology 22:4767−82

    doi: 10.1111/mec.12344

    CrossRef   Google Scholar

    [99]

    Liu K, Feng S, Pan Y, Zhong J, Chen Y, et al. 2016. Transcriptome analysis and identification of genes associated with floral transition and flower development in sugar apple (Annona squamosa L.). Frontiers in Plant Science 7:1695

    doi: 10.3389/fpls.2016.01695

    CrossRef   Google Scholar

    [100]

    Miyazaki Y, Maruyama Y, Chiba Y, Kobayashi MJ, Joseph B, et al. 2014. Nitrogen as a key regulator of flowering in Fagus crenata: understanding the physiological mechanism of masting by gene expression analysis. Ecology Letters 17:1299−309

    doi: 10.1111/ele.12338

    CrossRef   Google Scholar

    [101]

    Wang Z, Ma W, Zhu T, Lu N, Ouyang F, et al. 2020. Multi-omics sequencing provides insight into floral transition in Catalpa bungei. C.A. Mey. BMC Genomics 21:508

    doi: 10.1186/s12864-020-06918-y

    CrossRef   Google Scholar

    [102]

    Dai X, Lin Y, Zhou T, Li Y, Liao X, et al. 2022. Natural annual transcriptome dynamics of Eucalyptus reveal seasonal adaptation of tropical/sub-tropical trees. Tree Physiology Accepted:tpac136

    doi: 10.1093/treephys/tpac136

    CrossRef   Google Scholar

    [103]

    Hsu CY, Adams JP, Kim H, No K, Ma C, et al. 2011. FLOWERING LOCUS T duplication coordinates reproductive and vegetative growth in perennial poplar. PNAS 108:10756−61

    doi: 10.1073/pnas.1104713108

    CrossRef   Google Scholar

    [104]

    Cooke JE, Eriksson ME, Junttila O. 2012. The dynamic nature of bud dormancy in trees: environmental control and molecular mechanisms. Plant, Cell & Environment 35:1707−28

    doi: 10.1111/j.1365-3040.2012.02552.x

    CrossRef   Google Scholar

    [105]

    Maurya JP, Bhalerao RP. 2017. Photoperiod- and temperature-mediated control of growth cessation and dormancy in trees: a molecular perspective. Annals of Botany 120:351−60

    doi: 10.1093/aob/mcx061

    CrossRef   Google Scholar

    [106]

    Singh RK, Bhalerao RP, Eriksson ME. 2021. Growing in time: exploring the molecular mechanisms of tree growth. Tree Physiology 41:657−78

    doi: 10.1093/treephys/tpaa065

    CrossRef   Google Scholar

    [107]

    Jansson S, Douglas CJ. 2007. Populus: a model system for plant biology. Annual Review of Plant Biology 58:435−58

    doi: 10.1146/annurev.arplant.58.032806.103956

    CrossRef   Google Scholar

    [108]

    Ding J, Nilsson O. 2016. Molecular regulation of phenology in trees—because the seasons they are a-changin'. Current Opinion In Plant Biology 29:73−9

    doi: 10.1016/j.pbi.2015.11.007

    CrossRef   Google Scholar

    [109]

    Borthakur D, Busov V, Cao X, Du Q, Gailing O, et al. 2022. Current status and trends in forest genomics. Forestry Research 2:11

    doi: 10.48130/fr-2022-0011

    CrossRef   Google Scholar

    [110]

    Yordanov YS, Ma C, Strauss SH, Busov VB. 2014. EARLY BUD-BREAK 1 (EBB1) is a regulator of release from seasonal dormancy in poplar trees. PNAS 111:10001−6

    doi: 10.1073/pnas.1405621111

    CrossRef   Google Scholar

    [111]

    Azeez A, Zhao YC, Singh RK, Yordanov YS, Dash M, et al. 2021. EARLY BUD-BREAK 1 and EARLY BUD-BREAK 3 control resumption of poplar growth after winter dormancy. Nature Communications 12:1123

    doi: 10.1038/s41467-021-21449-0

    CrossRef   Google Scholar

    [112]

    Singh RK, Svystun T, AlDahmash B, Jönsson AM, Bhalerao RP. 2017. Photoperiod- and temperature-mediated control of phenology in trees — a molecular perspective. New Phytologist 213:511−24

    doi: 10.1111/nph.14346

    CrossRef   Google Scholar

    [113]

    Ding J, Böhlenius H, Rühl MG, Chen P, Sane S, et al. 2018. GIGANTEA-like genes control seasonal growth cessation in Populus. New Phytologist 218:1491−503

    doi: 10.1111/nph.15087

    CrossRef   Google Scholar

    [114]

    Ding J, Zhang B, Li Y, André D, Nilsson O. 2021. Phytochrome B and PHYTOCHROME INTERACTING FACTOR8 modulate seasonal growth in trees. New Phytologist 232:2339−52

    doi: 10.1111/nph.17350

    CrossRef   Google Scholar

    [115]

    Nilsson O. 2022. Winter dormancy in trees. Current Biology 32:R630−R634

    doi: 10.1016/j.cub.2022.04.011

    CrossRef   Google Scholar

    [116]

    Ramos-Sanchez JM, Triozzi PM, Alique D, Geng F, Gao M, et al. 2019. LHY2 integrates night-length information to determine timing of poplar photoperiodic growth. Current Biology 29:2402−2406.E4

    doi: 10.1016/j.cub.2019.06.003

    CrossRef   Google Scholar

    [117]

    Andre D, Marcon A, Lee KC, Goretti D, Zhang B, et al. 2022. FLOWERING LOCUS T paralogs control the annual growth cycle in Populus trees. Current Biology 32:2988−2996.E4

    doi: 10.1016/j.cub.2022.05.023

    CrossRef   Google Scholar

    [118]

    Karlgren A, Gyllenstrand N, Källman T, Sundström JF, Moore D, et al. 2011. Evolution of the PEBP gene family in plants: functional diversification in seed plant evolution. Plant Physiology 156:1967−77

    doi: 10.1104/pp.111.176206

    CrossRef   Google Scholar

    [119]

    Karlgren A, Gyllenstrand N, Clapham D, Lagercrantz U. 2013. FLOWERING LOCUS T/TERMINAL FLOWER1-Like Genes Affect Growth Rhythm and Bud Set in Norway Spruce. Plant Physiology 163:792−803

    doi: 10.1104/pp.113.224139

    CrossRef   Google Scholar

    [120]

    Gyllenstrand N, Clapham D, Källman T, Lagercrantz U. 2007. A Norway spruce FLOWERING LOCUS T homolog is implicated in control of growth rhythm in conifers. Plant Physiology 144:248−57

    doi: 10.1104/pp.107.095802

    CrossRef   Google Scholar

    [121]

    Chen J, Källman T, Ma X, Gyllenstrand N, Zaina G, et al. 2012. Disentangling the roles of history and local selection in shaping clinal variation of allele frequencies and gene expression in Norway spruce (Picea abies). Genetics 191:865−81

    doi: 10.1534/genetics.112.140749

    CrossRef   Google Scholar

    [122]

    Michaels SD, Amasino RM. 1999. FLOWERING LOCUS C encodes a novel MADS domain protein that acts as a repressor of flowering. The Plant Cell 11:949−56

    doi: 10.1105/tpc.11.5.949

    CrossRef   Google Scholar

    [123]

    Voogd C, Brian LA, Wu R, Wang T, Allan AC, et al. 2022. A MADS-box gene with similarity to FLC is induced by cold and correlated with epigenetic changes to control budbreak in kiwifruit. New Phytologist 233:2111−26

    doi: 10.1111/nph.17916

    CrossRef   Google Scholar

    [124]

    Díaz-Riquelme J, Lijavetzky D, Martínez-Zapater JM, Carmona MJ. 2009. Genome-wide analysis of MIKCC-type MADS box genes in grapevine. Plant Physiology 149:354−69

    doi: 10.1104/pp.108.131052

    CrossRef   Google Scholar

    [125]

    Leseberg CH, Li A, Kang H, Duvall M, Mao L. 2006. Genome-wide analysis of the MADS-box gene family in Populus trichocarpa. Gene 378:84−94

    doi: 10.1016/j.gene.2006.05.022

    CrossRef   Google Scholar

    [126]

    Kumar G, Arya P, Gupta K, Randhawa V, Acharya V, et al. 2016. Comparative phylogenetic analysis and transcriptional profiling of MADS-box gene family identified DAM and FLC-like genes in apple (Malusx domestica). Scientific Reports 6:20695

    doi: 10.1038/srep20695

    CrossRef   Google Scholar

    [127]

    Zhang J, Li Z, Mei L, Yao J, Hu C. 2009. PtFLC homolog from trifoliate orange (Poncirus trifoliata) is regulated by alternative splicing and experiences seasonal fluctuation in expression level. Planta 229:847−59

    doi: 10.1007/s00425-008-0885-z

    CrossRef   Google Scholar

    [128]

    Agustí M, Mesejo C, Muñoz-Fambuena N, Vera-Sirera F, de Lucas M, et al. 2020. Fruit-dependent epigenetic regulation of flowering in Citrus. New Phytologist 225:376−84

    doi: 10.1111/nph.16044

    CrossRef   Google Scholar

    [129]

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

    doi: 10.1007/s11295-007-0126-9

    CrossRef   Google Scholar

    [130]

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

    doi: 10.1101/gad.1518407

    CrossRef   Google Scholar

    [131]

    Yamane H, Wada M, Honda C, Matsuura T, Ikeda Y, et al. 2019. Overexpression of Prunus DAM6 inhibits growth, represses bud break competency of dormant buds and delays bud outgrowth in apple plants. PLoS One 14:e0214788

    doi: 10.1371/journal.pone.0214788

    CrossRef   Google Scholar

    [132]

    Wu R, Tomes S, Karunairetnam S, Tustin SD, Hellens RP, et al. 2017. SVP-like MADS Box Genes Control Dormancy and Budbreak in Apple. Frontiers in Plant Science 8:477

    doi: 10.3389/fpls.2017.00477

    CrossRef   Google Scholar

    [133]

    Leida C, Conesa A, Llácer G, Badenes ML, Ríos G. 2012. Histone modifications and expression of DAM6 gene in peach are modulated during bud dormancy release in a cultivar-dependent manner. New Phytologist 193:67−80

    doi: 10.1111/j.1469-8137.2011.03863.x

    CrossRef   Google Scholar

    [134]

    Saito T, Bai S, Imai T, Ito A, Nakajima I, et al. 2015. Histone modification and signalling cascade of the dormancy-associated MADS-box gene, PpMADS13-1, in Japanese pear (Pyrus pyrifolia) during endodormancy. Plant, Cell & Environment 38:1157−66

    doi: 10.1111/pce.12469

    CrossRef   Google Scholar

    [135]

    Wu R, Cooney J, Tomes S, Rebstock R, Karunairetnam S, et al. 2021. RNAi-mediated repression of dormancy-related genes results in evergrowing apple trees. Tree Physiology 41:1510−23

    doi: 10.1093/treephys/tpab007

    CrossRef   Google Scholar

    [136]

    Falavigna VDS, Guitton B, Costes E, Andrés F. 2018. I want to (Bud) break free: The potential role of DAM and SVP-like genes in regulating dormancy cycle in temperate fruit trees. Frontiers in Plant Science 9:1990

    doi: 10.3389/fpls.2018.01990

    CrossRef   Google Scholar

    [137]

    da Silveira Falavigna V, Severing E, Lai X, Estevan J, Farrera I, et al. 2021. Unraveling the role of MADS transcription factor complexes in apple tree dormancy. New Phytologist 232:2071−88

    doi: 10.1111/nph.17710

    CrossRef   Google Scholar

    [138]

    Moser M, Asquini E, Miolli GV, Weigl K, Hanke MV, et al. 2020. The MADS-box gene MdDAM1 controls growth cessation and bud dormancy in apple. Frontiers in Plant Science 11:1003

    doi: 10.3389/fpls.2020.01003

    CrossRef   Google Scholar

    [139]

    Singh RK, Maurya JP, Azeez A, Miskolczi P, Tylewicz S, et al. 2018. A genetic network mediating the control of bud break in hybrid aspen. Nature Communications 9:4173

    doi: 10.1038/s41467-018-06696-y

    CrossRef   Google Scholar

    [140]

    Singh RK, Miskolczi P, Maurya JP, Bhalerao RP. 2019. A Tree ortholog of SHORT VEGETATIVE PHASE Floral repressor mediates photoperiodic control of bud dormancy. Current Biology 29:128−133.E2

    doi: 10.1016/j.cub.2018.11.006

    CrossRef   Google Scholar

    [141]

    Yang Q, Gao Y, Wu X, Moriguchi T, Bai S, et al. 2021. Bud endodormancy in deciduous fruit trees: advances and prospects. Horticuture Research 8:139

    doi: 10.1038/s41438-021-00575-2

    CrossRef   Google Scholar

    [142]

    Wu R, Wang T, Warren BAW, Allan AC, Macknight RC, et al. 2017. Kiwifruit SVP2 gene prevents premature budbreak during dormancy. Journal of Experimental Botany 68:1071−82

    doi: 10.1093/jxb/erx014

    CrossRef   Google Scholar

    [143]

    André D, Zambrano JA, Zhang B, Lee KC, Rühl M, et al. 2022. Populus SVL acts in leaves to modulate the timing of growth cessation and bud set. Frontiers in Plant Science 13:823019

    doi: 10.3389/fpls.2022.823019

    CrossRef   Google Scholar

    [144]

    Goralogia GS, Howe GT, Brunner AM, Helliwell E, Nagle MF, et al. 2021. Overexpression of SHORT VEGETATIVE PHASE-LIKE (SVL) in Populus delays onset and reduces abundance of flowering in field-grown trees. Horticulture Research 8:167

    doi: 10.1038/s41438-021-00600-4

    CrossRef   Google Scholar

    [145]

    Moon J, Lee H, Kim M, Lee I. 2005. Analysis of flowering pathway integrators in Arabidopsis. Plant and Cell Physiology 46:292−99

    doi: 10.1093/pcp/pci024

    CrossRef   Google Scholar

    [146]

    Mouhu K, Kurokura T, Koskela EA, Albert VA, Elomaa P, et al. 2013. The Fragaria vesca homolog of suppressor of overexpression of constans1 represses flowering and promotes vegetative growth. The Plant Cell 25:3296−310

    doi: 10.1105/tpc.113.115055

    CrossRef   Google Scholar

    [147]

    Cseke LJ, Zheng J, Podila GK. 2003. Characterization of PTM5 in aspen trees: a MADS-box gene expressed during woody vascular development. Gene 318:55−67

    doi: 10.1016/S0378-1119(03)00765-0

    CrossRef   Google Scholar

    [148]

    Olukolu BA, Trainin T, Fan S, Kole C, Bielenberg DG, et al. 2009. Genetic linkage mapping for molecular dissection of chilling requirement and budbreak in apricot (Prunus armeniaca L.). Genome 52:819−28

    doi: 10.1139/G09-050

    CrossRef   Google Scholar

    [149]

    Manghwar H, Lindsey K, Zhang X, Jin S. 2019. CRISPR/Cas system: Recent advances and future prospects for genome editing. Trends In Plant Science 24:1102−25

    doi: 10.1016/j.tplants.2019.09.006

    CrossRef   Google Scholar

    [150]

    Yamagishi N, Li C, Yoshikawa N. 2016. Promotion of flowering by Apple latent spherical virus vector and virus elimination at high temperature allow accelerated breeding of apple and pear. Frontiers in Plant Science 7:171

    doi: 10.3389/fpls.2016.00171

    CrossRef   Google Scholar

    [151]

    Freiman A, Shlizerman L, Golobovitch S, Yablovitz Z, Korchinsky R, et al. 2012. Development of a transgenic early flowering pear (Pyrus communis L.) genotype by RNAi silencing of PcTFL1-1 and PcTFL1-2. Planta 235:1239−51

    doi: 10.1007/s00425-011-1571-0

    CrossRef   Google Scholar

    [152]

    Klocko AL, Ma C, Robertson S, Esfandiari E, Nilsson O, et al. 2016. FT overexpression induces precocious flowering and normal reproductive development in Eucalyptus. Plant Biotechnology Journal 14:808−19

    doi: 10.1111/pbi.12431

    CrossRef   Google Scholar

    [153]

    Putterill J, Varkonyi-Gasic E. 2016. FT and florigen long-distance flowering control in plants. Current Opinion In Plant Biology 33:77−82

    doi: 10.1016/j.pbi.2016.06.008

    CrossRef   Google Scholar

    [154]

    Fitter AH, Fitter RSR. 2002. Rapid changes in flowering time in British plants. Science 296:1689−91

    doi: 10.1126/science.1071617

    CrossRef   Google Scholar

    [155]

    Chen IC, Hill JK, Ohlemüller R, Roy DB, Thomas CD. 2011. Rapid Range Shifts of Species Associated with High Levels of Climate Warming. Science 333:1024−6

    doi: 10.1126/science.1206432

    CrossRef   Google Scholar

    [156]

    Satake A, Kawagoe T, Saburi Y, Chiba Y, Sakurai G, Kudoh H. 2013. Forecasting flowering phenology under climate warming by modelling the regulatory dynamics of flowering-time genes. Nature Communications 4:2303

    doi: 10.1038/ncomms3303

    CrossRef   Google Scholar

    [157]

    Diggle PK, Mulder CPH. 2019. Diverse Developmental Responses to Warming Temperatures Underlie Changes in Flowering Phenologies. Integrative And Comparative Biology 59:559−70

    doi: 10.1093/icb/icz076

    CrossRef   Google Scholar

    [158]

    Ma J, Chen X, Han F, Song Y, Zhou B, et al. 2022. The long road to bloom in conifers. Forestry Research 2:16

    doi: 10.48130/FR-2022-0016

    CrossRef   Google Scholar

    [159]

    Charrier A, Vergne E, Dousset N, Richer A, Petiteau A, Chevreau E. 2019. Efficient targeted mutagenesis in apple and first time edition of pear using the CRISPR-Cas9 system. Frontiers in Plant Science 10:40

    doi: 10.3389/fpls.2019.00040

    CrossRef   Google Scholar

    [160]

    Flachowsky H, Szankowski I, Waidmann S, Peil A, Trankner C, et al. 2012. The MdTFL1 gene of apple (Malus × domestica Borkh.) reduces vegetative growth and generation time. Tree Physiology 32:1288−301

    doi: 10.1093/treephys/tps080

    CrossRef   Google Scholar

    [161]

    Mimida N, Kotoda N, Ueda T, Igarashi M, Hatsuyama Y, et al. 2009. Four TFL1/CEN-like genes on distinct linkage groups show different expression patterns to regulate vegetative and reproductive development in apple (Malus × domestica Borkh.). Plant and Cell Physiology 50:394−412

    doi: 10.1093/pcp/pcp001

    CrossRef   Google Scholar

    [162]

    Flachowsky H, Hättasch C, Höfer M, Peil A, Hanke MV. 2010. Overexpression of LEAFY in apple leads to a columnar phenotype with shorter internodes. Planta 231:251−63

    doi: 10.1007/s00425-009-1041-0

    CrossRef   Google Scholar

    [163]

    Wada M, Cao Q, Kotoda N, Soejima J, Masuda T. 2002. Apple has two orthologues of FLORICAULA/LEAFY involved in flowering. Plant Molecular Biology 49:567−77

    doi: 10.1023/A:1015544207121

    CrossRef   Google Scholar

    [164]

    Kagaya H, Ito N, Shibuya T, Komori S, Kato K, Kanayama Y. 2020. Characterization of FLOWERING LOCUS C homologs in apple as a model for fruit trees. International Journal of Molecular Sciences 21:4562

    doi: 10.3390/ijms21124562

    CrossRef   Google Scholar

    [165]

    Ziv D, Zviran T, Zezak O, Samach A, Irihimovitch V. 2014. Expression profiling of FLOWERING LOCUS T-like gene in alternate bearing 'Hass' avocado trees suggests a role for PaFT in avocado flower induction. PLoS One 9:e110613

    doi: 10.1371/journal.pone.0110613

    CrossRef   Google Scholar

    [166]

    Song G, Walworth A, Zhao D, Jiang N, Hancock JF. 2013. The Vaccinium corymbosum FLOWERING LOCUS T-like gene (VcFT): a flowering activator reverses photoperiodic and chilling requirements in blueberry. Plant Cell Reports 32:1759−69

    doi: 10.1007/s00299-013-1489-z

    CrossRef   Google Scholar

    [167]

    Pillitteri LJ, Lovatt CJ, Walling LL. 2004. Isolation and characterization of a TERMINAL FLOWER homolog and its correlation with juvenility in citrus. Plant Physiology 135:1540−51

    doi: 10.1104/pp.103.036178

    CrossRef   Google Scholar

    [168]

    Orbović V, Ravanfar SA, Acanda Y, Narvaez J, Merritt BA, et al. 2021. Stress-inducible Arabidopsis thaliana RD29A promoter constitutively drives Citrus sinensis APETALA1 and LEAFY expression and precocious flowering in transgenic Citrus spp. Transgenic Research 30:687−99

    doi: 10.1007/s11248-021-00260-z

    CrossRef   Google Scholar

    [169]

    Liu X, Zhang J, Abuahmad A, Franks RG, Xie D, et al. 2016. Analysis of two TFL1 homologs of dogwood species (Cornus L.) indicates functional conservation in control of transition to flowering. Planta 243:1129−41

    doi: 10.1007/s00425-016-2466-x

    CrossRef   Google Scholar

    [170]

    Elorriaga E, Klocko AL, Ma C, du Plessis M, An X, et al. 2021. Genetic containment in vegetatively propagated forest trees: CRISPR disruption of LEAFY function in Eucalyptus gives sterile indeterminate inflorescences and normal juvenile development. Plant Biotechnology Journal 19:1743−55

    doi: 10.1111/pbi.13588

    CrossRef   Google Scholar

    [171]

    Brill EM, Watson JM. 2004. Ectopic expression of a Eucalyptus grandis SVP orthologue alters the flowering time of Arabidopsis thaliana. Functional Plant Biology 31:217

    doi: 10.1071/FP03180

    CrossRef   Google Scholar

    [172]

    Ikegami H, Nogata H, Inoue Y, Himeno S, Yakushiji H, et al. 2013. Expression of FcFT1, a FLOWERING LOCUS T-like gene, is regulated by light and associated with inflorescence differentiation in fig (Ficus carica L.). BMC Plant Biology 13:216

    doi: 10.1186/1471-2229-13-216

    CrossRef   Google Scholar

    [173]

    Li C, Chen L, Fan X, Qi W, Ma J, et al. 2020. MawuAP1 promotes flowering and fruit development in the basal angiosperm Magnolia wufengensis (Magnoliaceae). Tree Physiology 40:1247−59

    doi: 10.1093/treephys/tpaa057

    CrossRef   Google Scholar

    [174]

    Carmona MJ, Calonje M, Martínez-Zapater JM. 2007. The FT/TFL1 gene family in grapevine. Plant Molecular Biology 63:637−50

    doi: 10.1007/s11103-006-9113-z

    CrossRef   Google Scholar

    [175]

    Esumi T, Kitamura Y, Hagihara C, Yamane H, Tao R. 2010. Identification of a TFL1 ortholog in Japanese apricot (Prunus mume Sieb. et Zucc.). Scientia Horticulturae 125:608−16

    doi: 10.1016/j.scienta.2010.05.016

    CrossRef   Google Scholar

    [176]

    Tang M, Bai X, Wang J, Chen T, Meng X, et al. 2022. Efficiency of graft-transmitted JcFT for floral induction in woody perennial species of the Jatropha genus depends on transport distance. Tree Physiology 42:189−201

    doi: 10.1093/treephys/tpab116

    CrossRef   Google Scholar

    [177]

    Bai X, Ke J, Huang P, Fatima I, Cheng T, Tang M. 2022. Promotion of natural flowers by JcFT depends on JcLFY in the perennial woody species Jatropha curcas. Plant Science 318:111236

    doi: 10.1016/j.plantsci.2022.111236

    CrossRef   Google Scholar

    [178]

    Tang M, Tao Y, Fu Q, Song Y, Niu L, et al. 2016. An ortholog of LEAFY in Jatropha curcas regulates flowering time and floral organ development. Scientific Reports 6:37306

    doi: 10.1038/srep37306

    CrossRef   Google Scholar

    [179]

    Tang M, Tao Y, Xu Z. 2016. Ectopic expression of Jatropha curcas APETALA1 (JcAP1) caused early flowering in Arabidopsis, but not in Jatropha. PeerJ 4:e1969

    doi: 10.7717/peerj.1969

    CrossRef   Google Scholar

    [180]

    Li C, Fu Q, Niu L, Luo L, Chen J, Xu Z. 2017. Three TFL1 homologues regulate floral initiation in the biofuel plant Jatropha curcas. Scientific Reports 7:43090

    doi: 10.1038/srep43090

    CrossRef   Google Scholar

    [181]

    Varkonyi-Gasic E, Moss SMA, Voogd C, Wang T, Putterill J, et al. 2013. Homologs of FT, CEN and FD respond to developmental and environmental signals affecting growth and flowering in the perennial vine kiwifruit. New Phytologist 198:732−46

    doi: 10.1111/nph.12162

    CrossRef   Google Scholar

    [182]

    Voogd C, Brian LA, Wang T, Allan AC, Varkonyi-Gasic E. 2017. Three FT and multiple CEN and BFT genes regulate maturity, flowering, and vegetative phenology in kiwifruit. Journal of Experimental Botany 68:1539−53

    doi: 10.1093/jxb/erx044

    CrossRef   Google Scholar

    [183]

    Herath D, Voogd C, Mayo-Smith M, Yang B, Allan AC, et al. 2022. CRISPR-Cas9-mediated mutagenesis of kiwifruit BFT genes results in an evergrowing but not early flowering phenotype. Plant Biotechnology Journal 20:2064−76

    doi: 10.1111/pbi.13888

    CrossRef   Google Scholar

    [184]

    Wu R, Wang T, McGie T, Voogd C, Allan AC, et al. 2014. Overexpression of the kiwifruit SVP3 gene affects reproductive development and suppresses anthocyanin biosynthesis in petals, but has no effect on vegetative growth, dormancy, or flowering time. Journal of Experimental Botany 65:4985−95

    doi: 10.1093/jxb/eru264

    CrossRef   Google Scholar

    [185]

    Wu R, Walton EF, Richardson AC, Wood M, Hellens RP, et al. 2012. Conservation and divergence of four kiwifruit SVP-like MADS-box genes suggest distinct roles in kiwifruit bud dormancy and flowering. Journal of Experimental Botany 63:797−807

    doi: 10.1093/jxb/err304

    CrossRef   Google Scholar

    [186]

    Ding F, Zhang S, Chen H, Su Z, Zhang R, et al. 2015. Promoter difference of LcFT1 is a leading cause of natural variation of flowering timing in different litchi cultivars (Litchi chinensis Sonn.). Plant Science 241:128−37

    doi: 10.1016/j.plantsci.2015.10.004

    CrossRef   Google Scholar

    [187]

    Zhang J, Liu G, Guo C, He Y, Li Z, et al. 2011. The FLOWERING LOCUS T orthologous gene of Platanus acerifolia is expressed as alternatively spliced forms with distinct spatial and temporal patterns. Plant Biology 13:809−20

    doi: 10.1111/j.1438-8677.2010.00432.x

    CrossRef   Google Scholar

    [188]

    Winterhagen P, Tiyayon P, Samach A, Hegele M, Wunsche JN. 2013. Isolation and characterization of FLOWERING LOCUS T subforms and APETALA1 of the subtropical fruit tree Dimocarpus longan. Plant Physiology and Biochemistry 71:184−90

    doi: 10.1016/j.plaphy.2013.07.013

    CrossRef   Google Scholar

    [189]

    Jiang Y, Zhu Y, Zhang L, Su W, Peng J, et al. 2020. EjTFL1 genes promote growth but inhibit flower bud differentiation in loquat. Frontiers in Plant Science 11:576

    doi: 10.3389/fpls.2020.00576

    CrossRef   Google Scholar

    [190]

    Liu Y, Zhao Q, Meng N, Song H, Li C, et al. 2017. Over-expression of EjLFY-1 leads to an early flowering habit in strawberry (Fragaria × ananassa) and its asexual progeny. Frontiers in Plant Science 8:496

    doi: 10.3389/fpls.2017.00496

    CrossRef   Google Scholar

    [191]

    Gafni I, Rai AC, Halon E, Zviran T, Sisai I, et al. 2022. Expression profiling of four mango FT/TFL1-encoding genes under different fruit load conditions, and their involvement in flowering regulation. Plants 11:2409

    doi: 10.3390/plants11182409

    CrossRef   Google Scholar

    [192]

    Wang Y, He X, Yu H, Mo X, Fan Y, et al. 2021. Overexpression of four MiTFL1 genes from mango delays the flowering time in transgenic Arabidopsis. BMC Plant Biology 21:407

    doi: 10.1186/s12870-021-03199-9

    CrossRef   Google Scholar

    [193]

    Klintenäs M, Pin PA, Benlloch R, Ingvarsson PK, Nilsson O. 2012. Analysis of conifer FLOWERING LOCUS T/TERMINAL FLOWER1-like genes provides evidence for dramatic biochemical evolution in the angiosperm FT lineage. New Phytologist 196:1260−73

    doi: 10.1111/j.1469-8137.2012.04332.x

    CrossRef   Google Scholar

    [194]

    Haberman A, Bakhshian O, Cerezo-Medina S, Paltiel J, Adler C, et al. 2017. A possible role for flowering locus T-encoding genes in interpreting environmental and internal cues affecting olive (Olea europaea L.) flower induction. Plant, Cell & Environment 40:1263−80

    doi: 10.1111/pce.12922

    CrossRef   Google Scholar

    [195]

    Chen Y, Jiang P, Thammannagowda S, Liang H, Wilde HD. 2013. Characterization of peach TFL1 and comparison with FT/TFL1 gene families of the rosaceae. Journal of the American Society for Horticultural Science 138:12−7

    doi: 10.21273/JASHS.138.1.12

    CrossRef   Google Scholar

    [196]

    Cai Y, Wang L, Ogutu CO, Yang Q, Luo B, et al. 2021. The MADS-box gene PpPI is a key regulator of the double-flower trait in peach. Physiologia Plantarum 173:2119−29

    doi: 10.1111/ppl.13561

    CrossRef   Google Scholar

    [197]

    Zhang X, An L, Nguyen TH, Liang H, Wang R, et al. 2015. The cloning and functional characterization of peach CONSTANS and FLOWERING LOCUS T homologous genes PpCO and PpFT. PLoS One 10:e0124108

    doi: 10.1371/journal.pone.0124108

    CrossRef   Google Scholar

    [198]

    Freiman A, Golobovitch S, Yablovitz Z, Belausov E, Dahan Y, et al. 2015. Expression of flowering locus T2 transgene from Pyrus communis L. delays dormancy and leaf senescence in Malus × domestica Borkh, and causes early flowering in tobacco. Plant Science 241:164−76

    doi: 10.1016/j.plantsci.2015.09.012

    CrossRef   Google Scholar

    [199]

    Patil HB, Chaurasia AK, Azeez A, Krishna B, Subramaniam VR, et al. 2018. Characterization of two TERMINAL FLOWER1 homologs PgTFL1 and PgCENa from pomegranate (Punica granatum L.). Tree Physiol 38:772−84

    doi: 10.1093/treephys/tpx154

    CrossRef   Google Scholar

    [200]

    Azeez A, Miskolczi P, Tylewicz S, Bhalerao RP. 2014. A tree ortholog of APETALA1 mediates photoperiodic control of seasonal growth. Current Biology 24:717−24

    doi: 10.1016/j.cub.2014.02.037

    CrossRef   Google Scholar

    [201]

    Bi Z, Li X, Huang H, Hua Y. 2016. Identification, functional study, and promoter analysis of HbMFT1, a homolog of MFT from rubber tree (Hevea brasiliensis). International Journal of Molecular Sciences 17:247

    doi: 10.3390/ijms17030247

    CrossRef   Google Scholar

    [202]

    Yarur A, Soto E, León G, Almeida AM. 2016. The sweet cherry (Prunus avium) FLOWERING LOCUS T gene is expressed during floral bud determination and can promote flowering in a winter-annual Arabidopsis accession. Plant Reproduction 29:311−22

    doi: 10.1007/s00497-016-0296-4

    CrossRef   Google Scholar

    [203]

    Wang J, Jiu S, Xu Y, Sabir IA, Wang L, et al. 2021. SVP-like gene PavSVP potentially suppressing flowering with PavSEP, PavAP1, and PavJONITLESS in sweet cherries (Prunus avium L.). Plant Physiology and Biochemistry 159:277−84

    doi: 10.1016/j.plaphy.2020.12.013

    CrossRef   Google Scholar

    [204]

    Wang J, Zhang X, Yan G, Zhou Y, Zhang K. 2013. Over-expression of the PaAP1 gene from sweet cherry (Prunus avium L.) causes early flowering in Arabidopsis thaliana. Journal of Plant Physiology 170:315−20

    doi: 10.1016/j.jplph.2012.09.015

    CrossRef   Google Scholar

    [205]

    Lei H, Su S, Ma L, Wen Y, Wang X. 2017. Molecular cloning and functional characterization of CoFT1, a homolog of FLOWERING LOCUS T (FT) from Camellia oleifera. Gene 626:215−26

    doi: 10.1016/j.gene.2017.05.044

    CrossRef   Google Scholar

    [206]

    Velázquez K, Agüero J, Vives MC, Aleza P, Pina JA, et al. 2016. Precocious flowering of juvenile citrus induced by a viral vector based on Citrus leaf blotch virus: a new tool for genetics and breeding. Plant Biotechnology Journal 14:1976−85

    doi: 10.1111/pbi.12555

    CrossRef   Google Scholar

  • Cite this article

    Wang J, Ding J. 2023. Molecular mechanisms of flowering phenology in trees. Forestry Research 3:2 doi: 10.48130/FR-2023-0002
    Wang J, Ding J. 2023. Molecular mechanisms of flowering phenology in trees. Forestry Research 3:2 doi: 10.48130/FR-2023-0002

Figures(2)  /  Tables(1)

Article Metrics

Article views(6689) PDF downloads(1222)

Other Articles By Authors

REVIEW   Open Access    

Molecular mechanisms of flowering phenology in trees

Forestry Research  3 Article number: 2  (2023)  |  Cite this article

Abstract: Flower initiation is a phenological developmental process strictly regulated in all flowering plants. Studies in Arabidopsis thaliana, a model plant organism in plant biology and genetics, and major cereal crops have provided fundamental knowledge and understanding of the underlying molecular mechanisms and regulation in annuals. However, this flowering process and underly molecular mechanisms in perennials are much more complicated than those in annuals and remain poorly understood and documented. In recent years, the increasing availability of perennial plant genomes and advances in biotechnology have allowed the identification and characterization of flowering-associated gene orthologs in perennials. In this review, we compared and summarized the recent progress in regulation of flowering time in perennial trees, with an emphasis on the perennial-specific regulatory mechanisms. Pleiotropic effects on tree growth habits such as juvenility, seasonal activity–dormancy growth, and the applications of tree flowering phenology are discussed.

    • Flowering time is a complicated, environmentally responsive trait, which can impact the fitness and survival of all flowering plants[1]. The timing of flowering is determined by endogenous genetic factors, as well as various environmental signals, such as photoperiod, temperature, and stress[2]. Studies in Arabidopsis thaliana, a model plant, have provided a basis for understanding plant flowering regulation in annual plants, in which flowering time is precisely controlled by a gene regulatory network comprising more than 300 genes[3]. These genes are involved in complex signal pathways including the autonomous, age, circadian clock, and gibberellin (GA) pathways that respond to intracellular and intercellular signals, and vernalization, ambient temperature, and photoperiod pathways that react to environmental cues. Moreover, with some notable exceptions, genes with analogous functions and similar molecular mechanisms found in Arabidopsis have conserved functions in flowering regulation in annual crop species. Recent reviews have provided detailed descriptions of flowering genes and mechanisms in annual plants[48]. These provide a basis for the understanding of gene networks controlling the flowering phenology of trees discussed in this review.

      Most annual or biennial plants are monocarpic, flowering only once in their life cycle before death[2] (Fig. 1). Unlike annual plants, woody perennial species are typically polycarpic and undergo repeated vegetative and reproductive growth cycles[9]. Perennial trees take several years to undergo the juvenile to adult phase change to acquire reproductive capability[10]. Following first-time flowering, trees flower annually throughout their lifespan (Fig. 1). Therefore, flowering is split into two dimensions in perennial trees: the first onset of flowering after many years of juvenility, and seasonal flowering after reproductive maturity (Fig. 1). The very long lifespan and polycarpic growth habits require a more complex regulatory network to synchronize environmental cues and mediate the appropriate flowering time[11]. The availability of increasing genome assemblies for trees now allows the identification of flowering phenology-associated gene orthologs in perennial plants. Many flowering genes have been functionally characterized in perennial trees, and conserved as well as functionally divergent genes involved in flowering have been found. This review summarizes the current understanding of flowering time regulation in perennial trees. Moreover, we discuss the pleiotropic effects on tree growth habit such as juvenility, seasonal activity–dormancy growth, and the applications of tree flowering phenology.

      Figure 1. 

      Comparison of flowering phenology between annual and perennial woody plants. The life cycle of flowering plants can be considered as a succession of distinct growth phases: vegetative growth, followed by a reproductive phase and eventually seed set and senescence. Annuals are fast cyclers and only need several months from the stage of vegetative development to flowering, and complete their life cycle within one growing season (a). While perennial woody plants experience a prolonged vegetative phase with many years until the first onset of flowering. Following first-time flowering, trees undergo seasonal flowering throughout their lifespan (b). Tree’s seasonal flowering can mainly be divided into 'direct' and 'indirect' flowering types, based on whether the development from initiation to emergence is interrupted or includes a period of rest. The 'indirect' flowering is common among temperate/boreal trees. It displays extended periods between flower initiation and flower blooming, in which flowers initiate in the summer are dormant through the winter, and the trees do not blossom until the following spring. In comparison, 'direct' flowering is common among subtropical or tropical evergreen species. They finish their complete reproductive cycles during a single growing season without dormancy or a rest period. VPC, vegetative phase change; RPC, reproductive phase change.

    • The life cycle of flowering plants can be considered a succession of distinct growth phases: vegetative growth, followed by a reproductive development and seed set, and eventually senescence (Fig. 1). The length of these phases varies among species and is particularly extended in perennial plants. Annuals progress quickly from vegetative to reproductive stage to complete their life cycle in one growing season. While perennial woody plants undergo a prolonged vegetative phase varying from a few years to several decades until the first onset of flowering[10] (Fig. 1). The precise development phase transitions are essential for the success of plant adaptability, survival, and reproduction. Floral induction depends on the transition from the juvenile-to-adult vegetative phase (vegetative phase change, VPC), called the age pathway in flowering regulation[12,13]. In Arabidopsis, VPC and flowering transition are regulated by the sequential activity of two microRNAs, miR156 and miR172, and their respective target genes[13,14]. With the aging of the plant, a gradual decline in miR156 abundance occurs in accordance with a steady accumulation of SQUAMOSA PROMOTER-BINDING PROTEIN (SBP)-LIKE (SPL) transcription factors (TFs)[13,14]. miR156 reduction is also coupled with the gradual accumulation of miR172, which can repress APETALA2 (AP2)-like TFs[15,16]. SPL and AP2-like gene expression is regulated by diverse flowering signals and their products form the molecular output of a pathway that regulates VPC and flowering initiation[17,18]. The miR156/miR172 module is conserved and regulates VPC in several other crop species[19].

      Perennial woody plants experience a long period of vegetative growth before the first flower onset; thus, it is of more pragmatic value to study phase transitions of perennial woody plants. Studies of broad-leaved trees, such as Populus canadensis, Acacia confusa, A. colei, Hedera helix, Eucalyptus globulus, Quercus acutissima, Folium mori, Mangifera indica, Malus hupehensis, Persea americana, and Macadamia integrifolia, have shown that miR156 and miR172 have similar expression trends with age[2023], suggesting that miR156 and miR172 are common to almost all major plant taxa and their roles in the control of VPC appears conserved (Fig. 2). miR156 overexpression in both P. canadensis and P. tremula × alba prolongs the juvenile phase, providing a genetic support for its role in VPC in trees[20,24,25]. However, recent studies in a gymnosperm Pinus tabulaeformis showed that the expression pattern of miR156 and its target genes showed no correlation with age, suggesting diversity of VPC control in gymnosperm trees[26]. In Arabidopsis, the miR156/miR172 module showed strong connections between VPC and reproductive competence[12,13]. However, the relationship between VPC and floral induction in perennial plants is unclear. Morphological changes during VPC have been comprehensively characterized in P. tremula × alba using miR156 overexpression and knockdown transgenic plants, and the onset of adult traits already begins within three months of growth[25]. This phenomenon is interesting as it corrected our traditional understanding that trees have a long juvenile stage. If VPC is completed at the early stage, trees undergo a long period of the adult vegetative stage until floral induction. This raises a question on how the miR156/miR172 module coordinates these two processes. Although ectopic expression of SPL gene from Citrus clementina and Eriobotrya japonica could promote flowering in Arabidopsis[27,28]. Thus far, however, reports confirming that the miR156/miR172 module and related genes regulate floral initiation in trees are rare. To what extent miR156 and miR172 and their targets function in the first flowering of trees remains to be clarified (Fig. 2). Besides the age pathway, AP2-like genes contribute to polycarpy in Arabis alpina, which provides a valuable clue in understanding the molecular basis of the polycarpic growth habit of woody trees[29].

      Figure 2. 

      Molecular pathways of flowering phenology, and their shared mechanisms in seasonal vegetative growth regulation in Populus, the model tree for perennial plant phenology study. Flowering is split into two dimensions: one is the first onset of flowering after many years of juvenile and adult vegetative growth; another is seasonal flowering after reproductive maturity. Conserved to herbaceous plants, the juvenile to adult vegetative phase change is mainly regulated by two microRNAs, miR156 and miR172. The first onset of flowering is controlled by FT/TFL1 family genes and their downstream integrators such as AP1 and LFY. As trees will undergo a long period of adult vegetative stage until floral induction, how miR156/miR172 module regulate age-dependent flowering in trees remains an open question. Unlike in Arabidopsis, GA usually inhibits flowering in diverse woody angiosperms. Whether GA regulates flowering through SOC1-like genes, and do SOC1-like genes control reproductive competence in trees needs further investigation. For seasonal flowering, the expression of flowering integrator genes, such as FT (FT1 and FT2), CENL1, SVL, and LEAFY (LFY), are controlled by seasonal cues like photoperiod and temperature. The specific expression patterns of these genes ensures the tree undergoes floral initiation at a specific time of the year. Meanwhile, these flowering integrator genes also play key roles in the seasonal activity-dormancy vegetative growth, including photoperiod-induced growth cessation of shoot apex at the end of summer, dormancy induction in autumn, cold-induced release of dormancy in winter, and warm temperature-induced bud burst in spring. Thus, trees have evolved an ability to incorporate the environmental signal to different developing events. the diagram sketch of seasonal growth from the inside out represent seasons, the expression pattens of flowering integrator genes, the seasonal flowering events, the seasonal vegetative growth events and environment signals such as photoperiod (LD and SD, long day and short day) and temperature (high and low) respectively.

    • Most knowledge about the molecular mechanisms of flowering time comes from studies in the annual plant Arabidopsis, in which flowering initiation is induced by multiple pathways that converge to a few integrator genes, such as FLOWERING LOCUS T (FT) and SUPPRESSOR OF OVEREXPRESSION OF CONSTANS 1 (SOC1). These genes act as floral pathway integrators to activate downstream floral meristem identity genes, such as LEAFY (LFY) and APETALA1 (AP1), and cooperate to promote flowering[6,30]. Due to a lack of availability of molecular resources, the molecular mechanisms that regulate reproductive competence have not been widely explored in perennials. The most common way is to study the functional conservation of Arabidopsis genes regulating flowering by ectopic expression of these genes in transgenic trees (Table 1). The gene FT and its family member TERMINAL FLOWER 1 (TFL1) have contributed largely to our understanding of the molecular mechanisms that regulate reproductive competence in perennials. FT and TFL1 encode a pair of flowering regulators of the phosphatidylethanolamine-binding protein family[31]. FT promotes the reproductive transition and flowering, whereas TFL1 represses flowering[32]. The antagonistic roles of FT/TFL in mediating flowering signals have been documented in all angiosperm species examined[5,33]. In poplar trees, leaves in adult shoots have higher expression levels of FT than leaves in juvenile shoots[34]. Overexpression of FT orthologs can induce premature flowering in many perennial species[3444] (Table 1). In contrast, transgenic M. domestica, P. trichocarpa, and Actinidia chinensis plants with reduced TFL1 expression accelerated flowering and shortened the length of vegetative growth before first flowering[4547] (Table 1). Such an antagonistic function of FT and TFL1 observed in trees suggests their functional conservation in reproductive competence in perennial species.

      Table 1.  Functional orthologs of flowering integrator genes identified in perennial trees.

      SpeciesGeneConstructionFloweringOther effectsReferences
      Apple
      (Malus pumila Mill.)
      MdFT1OverexpressionA,NInduction[43]
      MdTFL1, MdTFL1.1
      MdCENa, MdCENb
      CRISPR/RNAiA,T,NRepression[45,159161]
      AFL1, AFL2OverexpressionA,NInduction[162,163]
      MdDAMa, MdDAMb, MdDAMc
      MdSVPa, MdSVPb
      RNAiNInductionRegulates bud dormancy[135]
      MdFLC1
      MdFLC3
      OverexpressionARepressionJuvenility regulation[164]
      Avocado
      (Persea americana)
      PaFTOverexpressionAInduction[165]
      Blueberry
      (Vaccinium corymbosum L.)
      VcFTOverexpressionT,NInduction[166]
      Birch (Betula)BpAP1OverexpressionNInduction[68]
      Citrus (Citrus sinensis)CsTFLOverexpressionARepression[167]
      Citrus
      (Citrus clementina)
      CsAP1Stress-inducible promoterNInduction[168]
      CsLFYStress-inducible promoterNInduction[168]
      CsSL1, CsSL2OverexpressionAInduction[52]
      Dogwood (Cornus L.)CorcanTFL1, CorfloTFL1OverexpressionARepression[169]
      Eucalyptus
      (Eucalyptus spp.)
      AtFTOverexpressionNInduction[152]
      PtFT1OverexpressionNInduction
      ELFYCRISPRAffects floral development[170]
      EgSVPOverexpressionARepressionAffects floral development[171]
      Fig (Ficus carica)FcFT1OverexpressionTInduction[172]
      MagnoliaceaeMawuAP1OverexpressionAInduction[173]
      Grapevine (Vitis spp.)VvTFL1AOverexpressionARepression[174]
      VvFTOverexpressionAInduction
      Japanese apricot
      (Prunus mume)
      PmFTOverexpressionAInduction[175]
      PmTFL1OverexpressionARepression
      Jatropha
      (Jatropha curcas L.)
      JcFTOverexpression /RNAiA,NInduction[36,176]
      JcLFYOverexpressionA,NInductionAffects floral fruit and seed development[177,178]
      JcAP1OverexpressionAInduction[179]
      JcTFL1a, JcTFL1b, JcTFL1cOverexpressionA,NRepression[180]
      JcTFL1RNAiNInduction
      Kiwifruit
      (Actinidia spp.)
      AcFT1, AcFT2OverexpressionA,NInduction[181,182]
      AcCEN1, AcCEN2,
      AcCEN3, AcCEN4
      Overexpression/CRISPRARepression[47,181,182]
      AcBFT1, AcBFT2, AcBFT3Overexpression/CRISPRARepressionAffects dormancy and bud break[182,183]
      SVP1-4OverexpressionA,T,NNormalAffects dormancy[142,184,185]
      AcSOC1e, AcSOC1f, AcSOC1iOverexpressionAInduction
      NNormal
      Affects dormancy[54]
      AcFLCLOverexpression/CRISPRRegulate bud break[123]
      Litchi
      (Litchi chinensis Sonn.)
      LcFT1, LcFT2OverexpressionA,TInduction[186]
      London plane
      (Platanus acerifolia)
      PaFTOverexpressionAInduction
      TInduction
      [187]
      Longan
      (Dimocarpus longan L.)
      DlFT1OverexpressionAInduction[188]
      DlFT2OverexpressionARepression
      Loquat
      (Eriobotrya japonica)
      EjTFL1-1, EjTFL1-2OverexpressionARepression[189]
      EjSOC1-1, EjSOC1-2OverexpressionAcenter[51]
      EjLFY-1OverexpressionSInduction[190]
      Mango
      (Mangiferaindica L.)
      MiFT1OverexpressionAInduction[191]
      MiFT2; MiTFL1-1, MiTFL1-2,
      MiTFL1-3, MiTFL1-4
      OverexpressionARepression[191,192]
      Norway spruce
      (Picea abies)
      PaFTL1, PaFTL2OverexpressionARepression[193]
      PaFTL2OverexpressionControl growth arrest[119]
      Olive (Olea europaea L.)OeFT1, OeFT2OverexpressionAInduction[194]
      Peach
      (Prunus persica L.)
      PpTFL1OverexpressionARepression[195]
      PpAP1OverexpressionAInduction[196]
      PpFTOverexpressionAInduction[197]
      Pear
      (Pyrus communis L.)
      PcTFL1-1, PcTFL1-2RNAiNInduction[151]
      PcFT2OverexpressionTInduction
      NNormal
      Regulate vegetative growth[198]
      PcTFL1.1CRISPRNInduction[159]
      Pomegranate
      (Punica granatum L.)
      PgTFL1, PgCENaOverexpressionARepression[199]
      Poplar
      (Populus spp.)
      FT1, FT2Overexpression/CRISPRNInductionFT1 regulates bud break; FT2 regulates growth cessation[34,44,
      103,117]
      LAP1Overexpression/RNAiAInductionRegulates growth cessation[200]
      PopCEN1, PopCEN2Overexpression/RNAiNRepressionRegulates bud break[46]
      SVLOverexpression/RNAiNRepressionRegulate growth cessation, dormancy and bud break[139,140,
      143,144]
      Rubber trees
      (Hevea brasiliensis)
      HbMFT1OverexpressionARepression[201]
      Sweet Cherry
      (Prunus avium L.)
      PavFTOverexpressionAInduction[202]
      PavSVPOverexpressionARepression[203]
      PavSOC1OverexpressionAInduction[55]
      PaAP1OverexpressionAInduction[204]
      Tea-oil tree
      (Camellia oleifera Abel.)
      CoFT1OverexpressionAInduction[205]
      Trifoliate orange
      (Poncirus trifoliate)
      CiFTOverexpressionA,NInduction[35,38,206]
      N, A, T and S represent function assessed in native plant, Arabidopsis, tobacco and strawberry respectively.

      SOC1 is another floral pathway integrator that integrates multiple flowering signals, including age-dependent signals in which SPL9 and miR156 are involved[48]. Thus, it is reasonable to speculate that SOC1-like genes would have roles in the cooperation of VPC and reproductive transition in trees. Recent studies in the perennial conifer P. tabulaeformis have identified 33 age-dependent TFs, among which 11 belong to the MADS-box family including SOC1-like genes[49]. Combined with transcriptome association analysis and genetic confirmation, the SOC1-like gene MADS11 was confirmed as a regulatory mediator of VPC in pine[49]. Thus far, SOC1-like genes have been widely studied in many angiosperm perennial trees. Many SOC1-like genes from different tree species have been confirmed to complement the late flowering of the soc1 mutant in Arabidopsis, suggesting their conserved roles in flowering induction[5054]. However, their roles in reproductive competence in native plants are less known. Ectopic expression of AcSOC1 in Actinidia chinensis failed to induce precocious flowering[54]. Instead, SOC1-like genes were associated with bud dormancy maintenace and dormancy released in many trees[5456], suggesting their functional diversification in woody plants. Whether and how SOC1-like genes regulate flowering induction in trees requires further investigation (Fig. 2).

      Besides FT and TFL1, overexpression of downstream floral meristem identity genes can overcome several years of the juvenile period in multiple woody species. For example, LFY plays an important role in both flowering initiation and floral meristem differentiation[57,58]. LFY homologs have been studied in many perennial trees, and their gene overexpression causes early flowering in hybrid aspen, citrus, litchi, and so on[5964]. AP1 is both a floral meristem identity and a floral organ morphology gene, and possibly regulates flowering[6567]. In perennial trees, though not all, overexpression of AP1 homologs can also induce flowering[61,68]. Therefore, orthologs of these floral initiation genes have conserved functions in regulating the first flowering after a long period of juvenility (Fig. 2).

      The phytohormone GA plays a major role in flowering regulation in Arabidopsis. It promotes flowering by inducing SOC1 and LFY expression under short-day conditions[69,70]. Increased LFY activity causes reduced GA levels by directly up-regulating the GA catabolic enzyme EUI-LIKE P450 A1 (ELA1) as well as GA2 oxidases, which in turn enables accumulation of DELLA proteins that complex with the SBP transcription factor SPL9 to activate AP1[71]. Thus, the GA plays dual opposite roles on flower formation onset in Arabidopsis. It promotes termination of vegetative development while inhibiting flower formation[71]. The role of GA in the floral initiation of woody perennials seems more complicated. It often inhibits flowering in diverse woody trees[72,73]. As perennial trees have two dimensions of flowering initiation: the first onset of flowering after many years of juvenility, and the seasonal flowering once reproductive maturity is reached. Most reports that showed GA inhibits floral induction refer to the seasonal flowering onset. This is also supported by the results of GA treatment of different physiological age of trees. Application of GA inhibitor only induce flowering in mature shoots in Eucalyptus nitens and Populus deltoides, but it appears to be inefficient on juvenile shoots[74,75]. In this case, the negative role of GA on seasonal flowering switch in woody angiosperms may be similar to the negative function of GA on flower formation in Arabidopsis. However, unlike its positive roles on vegetative termination in Arabidopsis, GA likely also inhibits the first onset of flowering in many woody angiosperms. Application of exogenous GAs in several perennial species can even cause a reverse from reproductive to vegetative development[76]. Previous studies have shown that juvenile shoot apices contained higher levels of endogenous GAs than adult shoot apices[76]. In grapevine, the GA inhibition of tree flowering is confirmed by an early flowering grapevine mutant that is defective in a grapevine homolog of the Arabidopsis gene GA INSENSITIVE (GAI), a key gene involved in GA signal transduction[77]. However, similar to Arabidopsis, GA appears as a flowering activator in conifers, and it is widely applied to stimulate flowering for breeding purposes[78]. Overall, the molecular mechanism of GA signals in the reproductive competence of trees is still an unsolved mystery (Fig. 2).

    • After perennials become capable of reproduction, they periodically flower with seasonal changes. Trees have evolved to time their flowering in appropriate seasons to adapt to geographically different environments. Thus, there is a rich diversity of flowering phenology from temperate to tropical climates[79]. In temperate regions, trees synchronize their flowering time to coincide with appropriate seasons by responding to seasonal environmental cues, particularly temperature and photoperiod, and flowering usually peaks from spring to early summer[80]. In tropical and subtropical forests where seasonal environmental cues are less available, there is a wide variation in flowering time patterns and its climatic drivers. Flowering can be seasonal or aseasonal with variation across years. For example, in tropical deciduous forests in India, five flowering types occur and the variation in flowering is relative to leaf flushing[81]. In seasonally dry tropical forests, flowering is driven by water availability, and flowering peaks usually occur at the end of the dry season or the beginning of the wet season[82,83]. Sometimes, the effect of climate on flowering phenology in subtropical forests is difficult to predict because of wide variations in rainfall seasonality. For instance, in Australia, flowering phenology varies among species, with both seasonally dynamic and spatially variable, driven by temperature, rainfall, and soil/substrate moisture[84]. Although there is no dry season in the Atlantic rainforest in Brazil, there is still clear seasonality in leafing and reproductive events that might be affected by slight changes in photoperiod and/or temperature[85]. Mass flowering occurs in some aseasonal Asian and South American tropical forests[86,87]. The differences in flowering phenological patterns observed among temperate forests, tropical dry forests, and tropical rainforests highlight differences in plant response to environmental cues. These environmental cues not only include the relatively stable seasonal environmental signals, such as temperature and photoperiod, but also biotic and abiotic stresses, such as drought, heat, and salinity. Such varied flowering phenology and their environmental drivers in trees support the idea that the altering flowering time is an evolutionary strategy for plant to maximize the chances of reproduction under diverse stress conditions[2]. We should keep in mind that the above observations of flowering phenology focus on flowering time rather than the time of floral initiation. Although flowering time is closely related to flower initiation, the environmental drivers can differ.

    • Flowering has been studied most extensively in Arabidopsis, in which temperature and photoperiod are two major environmental signals that regulate flowering initiation. However, knowledge of seasonal flowering initiation in perennial trees is scarce. On one hand, annual and perennial plants have different growth habits: most annual or biennial plants are monocarpic, whereas perennial species are typically polycarpic. These different growth habits are reflected by flowering patterns[88]. All meristems of annual plants transform to floral meristems, and the life cycle is completed within one year (Fig. 1). Perennial trees have asynchronous differentiation behavior of meristems, with some committing to reproductive development, whereas others retain vegetative growth[9,89]. Seasonal tree flowering can be mainly divided into 'direct' and 'indirect' flowering types, based on whether the development from initiation to emergence is interrupted or includes a period of rest[88]. 'Indirect' flowering is common among temperate trees (poplar, cherries, pears, plums, apples, etc.)[90,91]. It displays extended periods between flower initiation and blooming, in which flowering is initiated in summer, trees are dormant by winter, and trees do not blossom until the following spring (Fig. 1). In comparison, 'direct' flowering species (mango, jujuba, etc.) complete their reproductive cycles in a single growing season without dormancy or a rest period[92,93] (Fig. 1). On the other hand, seasonal flowering phenology is not mutually independent from other phenological events, such as seasonal growth cessation, dormancy, leaf flushing, and fruiting. Such a long flowering time accompanied by a complex natural environment and various phenological events, as well as asynchronous development of the axillary meristem, makes it more challenging to determine the time of flowering initiation and their environmental drivers.

      Although orthologs of floral pathway integrator genes, such as FT, LFY, and AP1, have been isolated from many trees, and their functions on reproductive competence are conserved among species (discussed in an earlier section of this review), much less is known about the molecular regulation of seasonal flowering. Applying advances in transcriptomics is an effective strategy to reveal the underlying mechanism of dynamic environmental responses in plants. Recently, a molecular phenology approach that monitors seasonal gene expression patterns in nature has been increasingly applied in a range of plants to explore plant responses to fluctuating natural environments[79,94]. These field-based seasonal transcriptomes provide ideal maps for associating genes with flowering phenology[95101]. This approach was also used to successfully identify environmental signals driving flowering in different tree species. For example, seasonal transcriptome studies in Fagus crenata showed that the expression levels of FT, LFY, and AP1 orthologs display clear between-year fluctuations[100]. These between-year fluctuations in gene expression coincided with the nitrogen change of current-year shoots. Plants fertilized with nitrogen can induce the expression of these three genes in F. crenata and flowered in two consecutive years. This result suggests that nitrogen is a key regulator of flowering initiation in this species[100]. Similarly, we recently characterized the annual transcriptome dynamics of the subtropical hardwood tree E. dunnii in natural field environments. Our transcriptome analysis, combined with geographical distribution, environmental cues, and heterologous transformation analyses, suggests that low temperature is one of the environmental triggers for its seasonal flowering[102]. Things are usually more complex because of gene duplication and sub-functionalization. For example, in Populus, two FT-like genes have been identified: PtFT1 and PtFT2. They have distinct seasonal expression patterns: PtFT2 peaks in late spring until early summer, and its expression is regulated by photoperiod, whereas PtFT1 is only induced by cold and peaks in late winter[103]. Both PtFT1 and PtFT2 can induce early flowering, suggesting their protein conservation in promoting flower initiation in Populus[34,44]. The expression pattern of PtFT2 is more similar to Arabidopsis FT, both of which are regulated by photoperiod[44]. However, PtFT1 is supposed to be a potential seasonal floral activator based on its seasonal expression pattern (Fig.2)[32,103]; however, more genetic and molecular evidence is needed to prove which one determines seasonal flowering initiation. In the future, such a molecular phenology strategy, combined with approaches in molecular biology, ecology, and mathematical modeling, will be useful to dissect the environmental factors regulating flowering traits in different climate zones.

    • Trees native to temperate and boreal regions have evolved an important adaptive trait in which they undergo a seasonal activity–dormancy growth cycle for survival and growth[104]. This activity–dormancy cycle includes cessation of apical growth, bud set, and dormancy induction in the fall; maintenance and release of dormancy in winter; and bud burst in spring (Fig. 2). Different developmental phases in the cycle have different responses to multiple environmental factors, adaptations that enable synchronization of these phases to the local climatic conditions. Temperature and photoperiod are two primary environmental cues, with the contribution of these cues varying among species. In the last two decades, the molecular mechanism of seasonal activity–dormancy growth has been intensively studied in trees, especially in the model species Populus[44,105111]. In Populus, the timing of growth cessation in the fall is primarily governed by photoperiod. The reason for this molecular mechanism breakthrough discovery is that PtFT2, besides promoting floral initiation, plays a key role in suppressing short-day induced growth cessation in the fall[44]. Since then, more genetic and molecular approaches have revealed a remarkable conservation of the photoperiod pathways in regulating growth cessation in Populus and controlling flowering time in Arabidopsis, originating from light perception by phytochromes (phyA and phyB), together with internal circadian clock genes such as LATE ELONGATED HYPOCOTYL, GIGANTEA, and CONSTANS, allowing plants to measure day length. Long photoperiods induce PtFT2 expression in the leaves. PtFT2 can move from the leaves to the shoot apex through the phloem where it interacts with TF FDL1 to induce the expression of LIKE APETALA1 and AINTEGUMENTA-LIKE1, which in turn activates the cell cycle genes and thus growth[108,112116]. After growth cessation, the continuation of short days induces bud dormancy. The dormant buds need a certain period of cold temperature to release dormancy for bud burst in spring. Recent studies have shown that another paralog of FT, PtFT1, plays a key role in cold-induced dormancy release. Plants with knockout PtFT1 showed inhibited dormancy release and delayed bud burst[117] (Fig. 2).

      Similar to the FT/TFL1 function in flowering regulation, RNAi downregulation of the poplar TFL1 homolog CENTRORADIALIS1 (CEN1) or CEN2 not only accelerated the first onset of flowering and increased the proportion of short shoots but also promoted dormancy release and advanced bud burst[46]. The antagonistically functioning paralogs FT and TFL1 likely arose after duplication in the angiosperm lineage, and the flowering-promoting function of FT evolved after the divergence of angiosperms from gymnosperms 300 million years ago[118]. Notably, an FT/TFL-based mechanism for seasonal growth has also been observed in conifers such as spruce[119,120]. Two FT/TFL1-like genes (PaFTL1 and PaFTL2) were identified in the conifer Norway spruce. Gene expression and population genetic studies have suggested that PaFTL1 and PaFTL2 act in concert to control perennial growth in Norway spruce. PaFTL1 expresses in the meristem and prevents meristematic cell proliferation during active extension growth in summer, whereas PaFTL2 attenuates extension growth in the fall[119,121]. Therefore, besides flowering time, FT/TFL1 genes have evolved roles in controlling seasonal growth before angiosperms and gymnosperms diverged.

      Many of the Arabidopsis MIKC MADS-box TFs are key regulators of reproductive development, including flowering time control, flower development, and inflorescence architecture. In Arabidopsis, the winter cold temperature response has a dominant effect on flowering time. Plants undergo vernalization to overcome prolonged cold, which suppresses flower initiation until cold acclimation is fulfilled by winter temperature under the control of a MADS-box protein, FLOWERING LOCUS C (FLC)[30,122]. Genes with sequence homology to FLC have been identified in many trees[123127]. However, FLC appeared specific to the Brassicaceae lineage as no sequences were similar enough to be regarded as individual orthologs to the FLC gene in Arabidopsis. Whether and how genes in the autonomous and vernalization pathways have evolved roles in flowering regulation in many other species is still less known. It has been reported that the citrus FLC-like genes CcMADS19 acts to regulate flowering by repressing the citrus FT[128]. Winter temperature response also exists in perennial trees from boreal and temperate climates, where plants undergo dormancy to overcome the harsh winter. Recent studies showed that AcFLCL, a kiwifruit FLC-like gene, is induced by cold and correlated with epigenitc changes to control budbreak in kiwifruit[123]. Besides FLC, another close homolog MADS-box TF, termed dormancy-associated MADS-box (DAM) plays a key role in this process. The DAM genes were discovered early in the nondormant evergrowing mutant of peach, which is incapable of going into dormancy[129]. DAM genes are orthologous to the floral repressor SHORT VEGETATIVE PHASE (SVP) of Arabidopsis, which is another MADS-box gene that plays a key role as a flowering repressor responding to ambient temperature[130]. DAM- and SVP-like genes have been characterized in many perennial species[131141], and the functions of these proteins in dormancy have been verified in transgenic plants of apple[131,132,137,138], poplar[139140], and kiwifruit[142]. For example, a Populus SVP ortholog termed SHORT VEGETATIVE PHASE-LIKE (SVL) has been reported to play extensive roles in seasonal growth. It is not only involved in photoperiod-regulated growth cessation and bud set, but also plays crucial roles in seasonal dormancy initiation and release[139,140,143]. Besides dormancy regulation, recent studies have shown that SVL overexpression poplars delay the onset of flowering by several years in field-grown conditions[144]. However, MdSVP overexpression in apple delayed bud burst in spring, but flower development and time to first flowering were normal[132]. In Arabidopsis, FT also promotes flowering by another MADS-box gene, SOC1, the first gene to be activated in the shoot apex[145]. Similar to SVP-like genes, SOC1-like genes have also been associated with seasonal dormancy in perennial trees[5456]. SOC1 controls seasonal vegetative and reproductive growth in strawberry[146]. In poplar, one SOC1 homolog PTM5 is implicated in seasonality and spring wood formation, and another SOC1 related gene promotes bud break[56, 147]. In apricot (Prunus armeniaca), a SOC1-like gene has been associated with chilling requirements during bud dormancy[148]. Functional studies of kiwifruit SOC1-like genes indicate that they affect the duration of dormancy but may not play a role in the transition to flowering[54].

      In brief, many of the genes involved in seasonal flowering are genes controlling seasonal activity–dormancy growth in trees. Thus, seasonal activity–dormancy vegetative growth and flowering may share a common mechanism in trees, although how these genes control tree flowering remains to be determined (Fig. 2).

    • Two decades ago, five central questions were raised about floral initiation in perennial trees[91]. To date, these questions have only been partially answered. However, some important underlying molecular mechanisms need to be further explored and understood. For example, 1) how flowering genes temporally and spatially regulate the first reproductive competency and seasonal initiation of flowering 2) How trees coordinate seasonal vegetative and reproduction growth, and to what extent do they share genetic pathways? 3) How trees adapt to local environments to precisely control flowering time in a season. The study on the molecular mechanisms of flowering phenology has been restricted due to a lack of genetic materials (e.g., a genome-wide mutant library), which take a long time to generate. Recent advances and applications of CRISPR/Cas9 technology have enabled the generation of knockout mutants of flowering-related genes[149]. Analysis of loss-of-function transgenics coupled with temporal–spatial expression analysis can provide important information to elucidate the functional roles of these genes during tree flowering. From the perspective of breeding, the prolonged juvenility of trees has greatly limited tree domestication. Thus far, many floral regulators can be used to regulate the switch from vegetative to reproductive growth in many tree species, such as FT, TFL1, LFY, and some MADS-box genes (Table 1). Gene manipulation has been used as a tool for accelerated breeding[40,47,150153]. In addition to regulating flowering time, flowering time genes also have pleiotropic effects on plant growth and development, including seasonal growth, pollen fertility, and wood development. Understanding the molecular mechanisms of how flowering genes finely control these traits can help optimize the breeding strategies and processes.

      With the progress of global warming, understanding how such a climate change impacts on the life cycle of organisms are critical for evaluating ecosystem vulnerability as the phenological shifts occur in the key life cycle of organisms[154,155]. Recently, the knowledge of the molecular basis of flowering genes in temperature response in Arabidopsis. halleri has been incorporated into a predictive model, which can be used to forecast flowering phenology under climate change[156]. With the increased knowledge of genetic architecture of flowering phenology in crops[8], this approach can be used to predict the adaptation of crops to the changing environment. The flowering phenology of perennial trees has become a major contributor to climate change metrics applied to understand the impact of global climate changes on plant ecosystems[154,155,157]. However, what we understand of the genetic base of flowering phenology is biased toward temperate regions[90,158]. Improved mechanistic understanding of environmental drivers of plant flowering phenology in other ecosystems is urgently needed[79]. Integration of molecular knowledge of flowering phenology, climate data, and ecological perspectives can help us to assess the vulnerability of the ecosystem and predict risks of climate change.

      • This work was supported by the National Natural Science Foundation of China (31971676) and the Fundamental Research Funds for Central Universities (2662019PY007).

      • Jihua Ding is the Editorial Board member of Journal Forestry Research. She is blinded from reviewing or making decisions on the manuscript. The article was subject to the journal's standard procedures, with peer-review handled independently of this Editorial Board member and her research groups.

      • 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 (1) References (206)
  • About this article
    Cite this article
    Wang J, Ding J. 2023. Molecular mechanisms of flowering phenology in trees. Forestry Research 3:2 doi: 10.48130/FR-2023-0002
    Wang J, Ding J. 2023. Molecular mechanisms of flowering phenology in trees. Forestry Research 3:2 doi: 10.48130/FR-2023-0002

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return