Search
2022 Volume 2
Article Contents
REVIEW   Open Access    

The long road to bloom in conifers

More Information
  • More than 600 species of conifers (phylum Pinophyta) serve as the backbone of the Earth’s terrestrial plant community and play key roles in global carbon and water cycles. Although coniferous forests account for a large fraction of global wood production, their productivity relies largely on the use of genetically improved seeds. However, acquisition of such seeds requires recurrent selection and testing of genetically superior parent trees, eventually followed by the establishment of a seed orchard to produce the improved seeds. The breeding cycle for obtaining the next generation of genetically improved seeds can be significantly lengthened when a target species has a long juvenile period. Therefore, development of methods for diminishing the juvenile phase is a cost-effective strategy for shortening breeding cycle in conifers. The molecular regulatory programs associated with the reproductive transition and annual reproductive cycle of conifers are modulated by environmental cues and endogenous developmental signals. Mounting evidence indicates that an increase in global average temperature seriously threatens plant productivity, but how conifers respond to the ever-changing natural environment has yet to be fully characterized. With the breakthrough of assembling and annotating the giant genome of conifers, identification of key components in the regulatory cascades that control the vegetative to reproductive transition is imminent. However, comparison of the signaling pathways that control the reproductive transition in conifers and the floral transition in Arabidopsis has revealed many differences. Therefore, a more complete understanding of the underlying regulatory mechanisms that control the conifer reproductive transition is of paramount importance. Here, we review our current understanding of the molecular basis for reproductive regulation, highlight recent discoveries, and review new approaches for molecular research on conifers.
  • 加载中
  • [1]

    Wang X, Ran J. 2014. Evolution and biogeography of gymnosperms. Molecular Phylogenetics and Evolution 75:24−40

    doi: 10.1016/j.ympev.2014.02.005

    CrossRef   Google Scholar

    [2]

    Florin R. 1951. Evolution in cordaites and conifers. Acta Horticulturae Bergiani 15:285−388

    Google Scholar

    [3]

    Carlsbecker A, Sundström J, Tandre K, Englund M, Kvarnheden A, et al. 2003. The DAL10 gene from Norway spruce (Picea abies) belongs to a potentially gymnosperm-specific subclass of MADS-box genes and is specifically active in seed cones and pollen cones. Evolution and Development 5:551−61

    doi: 10.1046/j.1525-142X.2003.03060.x

    CrossRef   Google Scholar

    [4]

    Becker A, Saedler H, Theissen G. 2003. Distinct MADS-box gene expression patterns in the reproductive cones of the gymnosperm Gnetum gnemon. Development Genes and Evolution 213:567−72

    doi: 10.1007/s00427-003-0358-0

    CrossRef   Google Scholar

    [5]

    Cooke JEK, Eriksson M, 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

    [6]

    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

    [7]

    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

    [8]

    Niu S, Li J, Bo W, Yang W, Zuccolo A, et al. 2021. The Chinese pine genome and methylome unveil key features of conifer evolution. Cell 185:204−217.e14

    doi: 10.1016/j.cell.2021.12.006

    CrossRef   Google Scholar

    [9]

    Liu Y, Yang K, Wei X, Wang X. 2017. Revisiting the phosphatidylethanolamine-binding protein (PEBP) gene family reveals cryptic FLOWERING LOCUS T gene homologs in gymnosperms and sheds new light on functional evolution. New Phytologist 212:730−44

    doi: 10.1111/nph.14066

    CrossRef   Google Scholar

    [10]

    Holefors A, Opseth L, Ree Rosnes AK, Ripel L, Snipen L, et al. 2009. Identification of PaCOL1 and PaCOL2, two CONSTANS-like genes showing decreased transcript levels preceding short day induced growth cessation in Norway spruce. Plant Physiology and Biochemistry 47:105−15

    doi: 10.1016/j.plaphy.2008.11.003

    CrossRef   Google Scholar

    [11]

    Zimin AV, Stevens KA, Crepeau MW, Puiu D, Wegrzyn JL, et al. 2017. An improved assembly of the loblolly pine mega-genome using long-read single-molecule sequencing. Gigascience 6:giw016

    doi: 10.1093/gigascience/giw016

    CrossRef   Google Scholar

    [12]

    Stevens KA, Wegrzyn JL, Zimin A, Puiu D, Crepeau M, et al. 2016. Sequence of the sugar pine megagenome. Genetics 204:1613−26

    doi: 10.1534/genetics.116.193227

    CrossRef   Google Scholar

    [13]

    Warren RL, Keeling CI, Yuen MM, Raymond A, Taylor GA, et al. 2015. Improved white spruce (Picea glauca) genome assemblies and annotation of large gene families of conifer terpenoid and phenolic defense metabolism. The Plant Journal 83:189−212

    doi: 10.1111/tpj.12886

    CrossRef   Google Scholar

    [14]

    Nystedt B, Street NR, Wetterbom A, Zuccolo A, Lin Y, et al. 2013. The Norway spruce genome sequence and conifer genome evolution. Nature 497:579−84

    doi: 10.1038/nature12211

    CrossRef   Google Scholar

    [15]

    Kosiński G, Giertych M. 1982. Light conditions inside developing buds affect floral induction. Planta 155:93−94

    doi: 10.1007/BF00402938

    CrossRef   Google Scholar

    [16]

    Mishra P, Panigrahi KC. 2015. GIGANTEA - an emerging story. Frontiers in Plant Science 6:8

    doi: 10.3389/fpls.2015.00008

    CrossRef   Google Scholar

    [17]

    Shim JS, Kubota A, Imaizumi T. 2017. Circadian clock and photoperiodic flowering in Arabidopsis: CONSTANS is a hub for signal integration. Plant Physiology 173:5−15

    doi: 10.1104/pp.16.01327

    CrossRef   Google Scholar

    [18]

    Li D, Zhang H, Mou M, Chen Y, Xiang S, et al. 2019. Arabidopsis class II TCP transcription factors integrate with the FT-FD module to control flowering. Plant physiology 181:97−111

    doi: 10.1104/pp.19.00252

    CrossRef   Google Scholar

    [19]

    Bendix C, Mendoza JM, Stanley DN, Meeley R, Harmon FG. 2013. The circadian clock-associated gene gigantea1 affects maize developmental transitions. Plant, Cell & Environment 36:1379−90

    doi: 10.1111/pce.12067

    CrossRef   Google Scholar

    [20]

    Kubota A, Kita S, Ishizaki K, Nishihama R, Yamato KT, et al. 2014. Co-option of a photoperiodic growth-phase transition system during land plant evolution. Nature Communications 5:3668

    doi: 10.1038/ncomms4668

    CrossRef   Google Scholar

    [21]

    Karlgren A, Gyllenstrand N, Källman T, Lagercrantz U. 2013. Conserved Function of Core Clock Proteins in the Gymnosperm Norway Spruce (Picea abies). PLoS One 8:e60110

    doi: 10.1371/journal.pone.0060110

    CrossRef   Google Scholar

    [22]

    Jun C, Yoshiaki T, Michael S, Thomas K, Nannan X, et al. 2014. Clinal variation at phenology-related genes in spruce: parallel evolution in FTL2 and Gigantea? Genetics 197:1025−38

    doi: 10.1534/genetics.114.163063

    CrossRef   Google Scholar

    [23]

    Olsen JE. 2010. Light and temperature sensing and signaling in induction of bud dormancy in woody plants. Plant Molecular Biology 73:37−47

    doi: 10.1007/s11103-010-9620-9

    CrossRef   Google Scholar

    [24]

    Eriksson ME, Millar AJ. 2003. The circadian clock. A plant's best friend in a spinning world. Plant Physiology 132:732−38

    doi: 10.1104/pp.103.022343

    CrossRef   Google Scholar

    [25]

    Søgaard G, Johnsen Ø, Nilsen J, Junttila O. 2008. Climatic control of bud burst in young seedlings of nine provenances of Norway spruce. Tree Physiology 28:311−20

    doi: 10.1093/treephys/28.2.311

    CrossRef   Google Scholar

    [26]

    Nose M, Kurita M, Tamura M, Matsushita M, Hiraoka Y, et al. 2020. Effects of day length- and temperature-regulated genes on annual transcriptome dynamics in Japanese cedar (Cryptomeria japonica D. Don), a gymnosperm indeterminate species. PLoS One 15:e0229843

    doi: 10.1371/journal.pone.0229843

    CrossRef   Google Scholar

    [27]

    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

    [28]

    Bowe LM, Coat G, dePamphilis CW. 2000. Phylogeny of seed plants based on all three genomic compartments: Extant gymnosperms are monophyletic and Gnetales' closest relatives are conifers. PNAS 97:4092−97

    doi: 10.1073/pnas.97.8.4092

    CrossRef   Google Scholar

    [29]

    Lagercrantz U. 2009. At the end of the day: a common molecular mechanism for photoperiod responses in plants. Journal of Experimental Botany 60:2501−15

    doi: 10.1093/jxb/erp139

    CrossRef   Google Scholar

    [30]

    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

    [31]

    Asante DKA, Yakovlev IA, Fossdal CG, Holefors A, Opseth L, et al. 2011. Gene expression changes during short day induced terminal bud formation in Norway spruce. Plant, Cell & Environment 34:332−46

    doi: 10.1111/j.1365-3040.2010.02247.x

    CrossRef   Google Scholar

    [32]

    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

    [33]

    Helliwell CA, Wood CC, Robertson M, James Peacock W, Dennis ES. 2006. The Arabidopsis FLC protein interacts directly in vivo with SOC1 and FT chromatin and is part of a high-molecular-weight protein complex. The Plant Journal 46:183−92

    doi: 10.1111/j.1365-313X.2006.02686.x

    CrossRef   Google Scholar

    [34]

    Searle I, He Y, Turck F, Vincent C, Fornara F, et al. 2006. The transcription factor FLC confers a flowering response to vernalization by repressing meristem competence and systemic signaling in Arabidopsis. Genes & Development 20:898−912

    doi: 10.1101/gad.373506

    CrossRef   Google Scholar

    [35]

    Hepworth SR, Valverde F, Ravenscroft D, Mouradov A, Coupland G. 2002. Antagonistic regulation of flowering-time gene SOC1 by CONSTANS and FLC via separate promoter motifs. The EMBO Journal 21:4327−37

    doi: 10.1093/emboj/cdf432

    CrossRef   Google Scholar

    [36]

    Wan T, Liu Z, Li L, Leitch AR, Leitch IJ, et al. 2018. A genome for gnetophytes and early evolution of seed plants. Nature Plants 4:82−89

    doi: 10.1038/s41477-017-0097-2

    CrossRef   Google Scholar

    [37]

    Zhang H, Cui X, Guo Y, Luo C, Zhang L. 2018. Picea wilsonii transcription factor NAC2 enhanced plant tolerance to abiotic stress and participated in RFCP1-regulated flowering time. Plant Molecular Biology 98:471−93

    doi: 10.1007/s11103-018-0792-z

    CrossRef   Google Scholar

    [38]

    Fu D, Dunbar M, Dubcovsky J. 2007. Wheat VIN3-like PHD finger genes are up-regulated by vernalization. Molecular Genetics and Genomics 277:301−13

    doi: 10.1007/s00438-006-0189-6

    CrossRef   Google Scholar

    [39]

    André D, Marcon A, Lee KC, Goretti D, Zhang B E. et al. 2022. FLOWERING LOCUS T paralogs control the annual growth cycle in Populus trees. Current Biology 32:2988−96

    doi: 10.1016/j.cub.2022.05.023

    CrossRef   Google Scholar

    [40]

    Chiang CM, Viejo M, Aas OT, Hobrak KT, Stromme CB, et al. 2021. Interactive effects of light quality during day extension and temperature on bud set, bud burst and PaFTL2, PaCOL1-2 and PaSOC1 expression in Norway Spruce (Picea abies (L.) Karst.). Forests 12:337

    doi: 10.3390/f12030337

    CrossRef   Google Scholar

    [41]

    Heide OM. 1974. Growth and dormancy in Norway Spruce ecotypes (Picea abies) Interaction of photoperiod and temperature. Physiology Plantarum 30:1−12

    doi: 10.1111/j.1399-3054.1974.tb04983.x

    CrossRef   Google Scholar

    [42]

    Hamilton JA, El Kayal W, Hart AT, Runcie DE, Arango-Velez A, et al. 2016. The joint influence of photoperiod and temperature during growth cessation and development of dormancy in white spruce (Picea glauca). Tree Physiology 36:1432−48

    doi: 10.1093/treephys/tpw061

    CrossRef   Google Scholar

    [43]

    Michaels SD, Amasino RM. 2001. Loss of FLOWERING LOCUS C activity eliminates the late-flowering phenotype of FRIGIDA and autonomous pathway mutations but not responsiveness to vernalization. The Plant Cell 13:935−941

    doi: 10.1105/tpc.13.4.935

    CrossRef   Google Scholar

    [44]

    Feng W, Jacob Y, Veley KM, Ding L, Yu X, et al. 2011. Hypomorphic alleles reveal FCA-independent roles for FY in the regulation of FLOWERING LOCUS C. Plant Physiology 155:1425−34

    doi: 10.1104/pp.110.167817

    CrossRef   Google Scholar

    [45]

    Hornyik C, Terzi LC, Simpson GG. 2010. The spen family protein FPA controls alternative cleavage and polyadenylation of RNA. Developmental Cell 18:203−13

    doi: 10.1016/j.devcel.2009.12.009

    CrossRef   Google Scholar

    [46]

    Pascual MB, Canovas FM and Avila C. 2015. The NAC transcription factor family in maritime pine (Pinus Pinaster): molecular regulation of two genes involved in stress responses. BMC Plant Biology 15:254

    doi: 10.1186/s12870-015-0640-0

    CrossRef   Google Scholar

    [47]

    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

    [48]

    Avia K, Kärkkäinen K, Lagercrantz U, Savolainen O. 2014. Association of FLOWERING LOCUS T/TERMINAL FLOWER 1-like gene FTL2 expression with growth rhythm in Scots pine (Pinus sylvestris). New Phytologist 204:159−70

    doi: 10.1111/nph.12901

    CrossRef   Google Scholar

    [49]

    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

    [50]

    Moyroud E, Kusters E, Monniaux M, Koes R, Parcy F. 2010. LEAFY blossoms. Trends in Plant Science 15:346−52

    doi: 10.1016/j.tplants.2010.03.007

    CrossRef   Google Scholar

    [51]

    Vázquez-Lobo A, Carlsbecker A, Vergara-Silva F, Alvarez-Buylla ER, Piñero D, et al. 2007. Characterization of the expression patterns of LEAFY/FLORICAULA and NEEDLY orthologs in female and male cones of the conifer genera Picea, Podocarpus, and Taxus: implications for current evo-devo hypotheses for gymnosperms. Evolution & Development 9:446−59

    doi: 10.1111/j.1525-142x.2007.00182.x

    CrossRef   Google Scholar

    [52]

    Frohlich MW, Parker DS. 2000. The mostly male theory of flower evolutionary origins: from genes to fossils. Systematic Botany 25:155−70

    doi: 10.2307/2666635

    CrossRef   Google Scholar

    [53]

    Mellerowicz EJ, Horgan K, Walden A, Coker A, Walter C. 1998. PRFLL - A Pinus radiata homologue of FLORICAULA and LEAFY is expressed in buds containing vegetative shoot and undi erentiated male cone primordia. Planta 206:619−29

    doi: 10.1007/s004250050440

    CrossRef   Google Scholar

    [54]

    Mouradov A, Glassick T, Hamdorf B, Murphy L, Fowler B, et al. 1998. NEEDLY, a Pinus radiata ortholog of FLORICAULA/LEAFY genes, expressed in both reproductive and vegetative meristems. PNAS 95:6537−42

    doi: 10.1073/pnas.95.11.6537

    CrossRef   Google Scholar

    [55]

    Shiokawa T, Yamada S, Futamura N, Osanai K, Murasugi D, et al. 2008. Isolation and functional analysis of the CjNdly gene, a homolog in Cryptomeria japonica of FLORICAULA/LEAFY genes. Tree Physiology 28:21−28

    doi: 10.1093/treephys/28.1.21

    CrossRef   Google Scholar

    [56]

    Dornelas MC, Rodriguez APM. 2005. A FLORICAULA/LEAFY gene homolog is preferentially expressed in developing female cones of the tropical pine Pinus caribaea var. caribaea. Genetics and Molecular Biology 2:299−307

    doi: 10.1590/s1415-47572005000200021

    CrossRef   Google Scholar

    [57]

    Carlsbecker A, Tandre K, Johanson U, Englund M, Engström P. 2004. The MADS-box gene DAL1 is a potential mediator of the juvenile-to-adult transition in Norway spruce (Picea abies). The Plant Journal 40:546−557

    doi: 10.1111/j.1365-313X.2004.02226.x

    CrossRef   Google Scholar

    [58]

    Carlsbecker A, Sundström JF, Englund M, Uddenberg D, Izquierdo L, et al. 2013. Molecular control of normal and acrocona mutant seed cone development in Norway spruce (Picea abies) and the evolution of conifer ovule-bearing organs. New Phytologist 200:261−75

    doi: 10.1111/nph.12360

    CrossRef   Google Scholar

    [59]

    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

    [60]

    Zhang P, Tan HTW, Pwee HK, Kumar PP. 2004. Conservation of class C function of floral organ development during 300 million years of evolution from gymnosperms to angiosperms. The Plant Journal 37:566−77

    doi: 10.1046/j.1365-313X.2003.01983.x

    CrossRef   Google Scholar

    [61]

    Winter KU, Saedler H, Theißen G. 2002. On the origin of class B floral homeotic genes: functional substitution and dominant inhibition in Arabidopsis by expression of an orthologue from the gymnosperm Gnetum. The Plant Journal 4:457−75

    doi: 10.1046/j.1365-313x.2002.01375.x

    CrossRef   Google Scholar

    [62]

    Moyroud E, Monniaux M, Thévenon E, Dumas R, Scutt CP, et al. 2017. A link between LEAFY and B-gene homologues in Welwitschia mirabilis sheds light on ancestral mechanisms prefiguring floral development. New Phytologist 216:469−81

    doi: 10.1111/nph.14483

    CrossRef   Google Scholar

    [63]

    Albani MC, Coupland G. 2010. Comparative analysis of flowering in annual and perennial plants. In Current Topics in Developmental Biology, ed. Timmermans MCP. Vol. 91. UK: Academic Press, Elsevier. pp. 323−48. https://doi.org/10.1016/S0070-2153(10)91011-9

    [64]

    Bergonzi S, Albani MC, Ver Loren van Themaat E, Nordström KJ, Wang R, et al. 2013. Mechanisms of age-dependent response to winter temperature in perennial flowering of Arabis alpina. Science 340:1094−97

    doi: 10.1126/science.1234116

    CrossRef   Google Scholar

    [65]

    Nodine MD, Bartel DP. 2010. MicroRNAs prevent precocious gene expression and enable pattern formation during plant embryogenesis. Genes & Development 24:2678−92

    doi: 10.1101/gad.1986710

    CrossRef   Google Scholar

    [66]

    Wötzel S, Andrello M, Albani MC, Koch MA, Coupland G, et al. 2022. Arabis alpina: A perennial model plant for ecological genomics and life-history evolution. Molecular Ecology Resources 22:468−86

    doi: 10.1111/1755-0998.13490

    CrossRef   Google Scholar

    [67]

    Albani MC, Castaings L, Wötzel S, Mateos JL, Wunder J, et al. 2012. PEP1 of Arabis alpina is encoded by two overlapping genes that contribute to natural genetic variation in perennial flowering. PLoS Genetics 8:e1003130

    doi: 10.1371/journal.pgen.1003130

    CrossRef   Google Scholar

    [68]

    Wang R, Farrona S, Vincent C, Joecker A, Schoof H, et al. 2009. PEP1 regulates perennial flowering in Arabis alpina. Nature 459:423−27

    doi: 10.1038/nature07988

    CrossRef   Google Scholar

    [69]

    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

    [70]

    Zhang Q, Li J, Sang Y, Xing S, Wu Q, et al. 2015. Identification and characterization of MicroRNAs in Ginkgo biloba var. epiphylla Mak. PLoS One 10:e0127184

    doi: 10.1371/journal.pone.0127184

    CrossRef   Google Scholar

    [71]

    Niu S, Liu C, Yuan HW, Li P, Li Y, et al. 2015. Identification and expression profiles of sRNAs and their biogenesis and action-related genes in male and female cones of Pinus tabuliformis. BMC Genomics 1:693

    doi: 10.1186/s12864-015-1885-6

    CrossRef   Google Scholar

    [72]

    Akhter S, Westrin KJ, Zivi N, Nordal V, Kretzschmar WW, et al. 2022. Cone-setting in spruce is regulated by conserved elements of the age-dependent flowering pathway. New Phytologist 236:1951−63

    doi: 10.1111/nph.18449

    CrossRef   Google Scholar

    [73]

    Zhang M, Chen Y, Jin X, Cai Y, Yuan Y, et al. 2019. New different origins and evolutionary processes of AP2/EREBP transcription factors in Taxus chinensis. BMC Plant Biology 19:413

    doi: 10.1186/s12870-019-2044-z

    CrossRef   Google Scholar

    [74]

    Nilsson L, Carlsbecker A, Sundås-Larsson A, Vahala T. 2007. APETALA2-like genes from Picea abies show functional similarities to their Arabidopsis homologues. Planta 225:589−602

    doi: 10.1007/s00425-006-0374-1

    CrossRef   Google Scholar

    [75]

    Shigyo M, Hasebe M, Ito M. 2006. Molecular evolution of the AP2 subfamily. Gene 366:256−65

    doi: 10.1016/j.gene.2005.08.009

    CrossRef   Google Scholar

    [76]

    Mouradov A, Glassick TV, Hamdorf BA, Murphy LC, Marla SS, et al. 1998. Family of MADS-Box genes expressed early in male and female reproductive structures of monterey pine. Plant Physiology 117:55−62

    doi: 10.1104/pp.117.1.55

    CrossRef   Google Scholar

    [77]

    Xiang W, Li W, Zhang S, Qi L. 2019. Transcriptome-wide analysis to dissect the transcription factors orchestrating the phase change from vegetative to reproductive development in Larix kaempferi. Tree Genetics & Genomes 15:681−89

    doi: 10.1007/s11295-019-1376-z

    CrossRef   Google Scholar

    [78]

    Chen F, Zhang X, Liu X, Zhang L. 2017. Evolutionary analysis of MIKCc-Type MADS-box genes in gymnosperms and angiosperms. Frontiers in Plant Science 8:895

    doi: 10.3389/fpls.2017.00895

    CrossRef   Google Scholar

    [79]

    Akhter S, Kretzschmar W, Nordal V, Delhomme N, Street NR, et al. 2018. Integrative analysis of three RNA sequencing methods identifies mutually exclusive exons of MADS-box isoforms during early bud development in Picea abies. Frontiers in Plant Science 9:1625

    doi: 10.3389/fpls.2018.01625

    CrossRef   Google Scholar

    [80]

    Uddenberg D, Reimegård J, Clapham D, Almqvist C, von Arnold S, et al. 2013. Early cone setting in Picea abies acrocona is associated with increased transcriptional activity of a MADS-box transcription factor. Plant Physiology 161:813−23

    doi: 10.1104/pp.112.207746

    CrossRef   Google Scholar

    [81]

    Chen X, Zhu Q, NieY, Han F, Li Y, et al. 2021. Determination of conifer age biomarker DAL1 interactome using Y2H-seq. Forestry Research 1:12

    doi: 10.48130/fr-2021-0012

    CrossRef   Google Scholar

    [82]

    Bao S, Hua C, Shen L, Yu H. 2020. New insights into gibberellin signaling in regulating flowering in Arabidopsis. Journal of Integrative Plant Biology 62:118−131

    doi: 10.1111/jipb.12892

    CrossRef   Google Scholar

    [83]

    Srikanth A, Schmid M. 2011. Regulation of flowering time: all roads lead to Rome. Cellular and Molecular Life Sciences 68:2013−2037

    doi: 10.1007/s00018-011-0673-y

    CrossRef   Google Scholar

    [84]

    Achard P, Herr A, Baulcombe DC, Harberd NP. 2004. Modulation of floral development by a gibberellin-regulated microRNA. Development 131:3357−65

    doi: 10.1242/dev.01206

    CrossRef   Google Scholar

    [85]

    Yu S, Galvão VC, Zhang Y, Horrer D, Zhang T, et al. 2012. Gibberellin regulates the Arabidopsis floral transition through miR156-targeted SQUAMOSA PROMOTER BINDING-LIKE Transcription factors. The Plant Cell 24:3320−32

    doi: 10.1105/tpc.112.101014

    CrossRef   Google Scholar

    [86]

    Blázquez 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

    [87]

    Schwechheimer C, Willige BC. 2009. Shedding light on gibberellic acid signalling. Current Opinion in Plant Biology 12:57−62

    doi: 10.1016/j.pbi.2008.09.004

    CrossRef   Google Scholar

    [88]

    de Lucas M, Davière JM, Rodríguez-Falcón M, Pontin M, Iglesias-Pedraz JM, et al. 2008. A molecular framework for light and Gibberellin control of cell elongation. Nature 451:480−84

    doi: 10.1038/nature06520

    CrossRef   Google Scholar

    [89]

    Millar AA, Lohe A, Wong G. 2019. Biology and Function of miR159 in Plants. Plants 8:255

    doi: 10.3390/plants8080255

    CrossRef   Google Scholar

    [90]

    Rosenberg O, Almqvist C, Weslien J. 2012. Systemic insecticide and Gibberellin reduced cone damage and increased flowering in a spruce seed orchard. Journal of Economic Entomology 105:916−22

    doi: 10.1603/EC11388

    CrossRef   Google Scholar

    [91]

    Cecich RA, Kang H, Chalupka W. 1994. Regulation of early flowering in Pinus banksiana. Tree Physiology 14:275−84

    doi: 10.1093/treephys/14.3.275

    CrossRef   Google Scholar

    [92]

    Li Y, Li X, Zhao MH, Pang ZY, Wei JT, et al. 2021. An overview of the practices and management methods for enhancing seed production in conifer plantations for commercial use. Horticulturae 7:252

    doi: 10.3390/horticulturae7080252

    CrossRef   Google Scholar

    [93]

    Kong L, Von Aderkas P, Irina Zaharia L. 2016. Effects of exogenously applied Gibberellins and Thidiazuron on phytohormone profiles of long-shoot buds and cone gender determination in lodgepole pine. Journal of Plant Growth Regulation 35:172−82

    doi: 10.1007/s00344-015-9517-6

    CrossRef   Google Scholar

    [94]

    Niu S, Yuan L, Zhang Y, Chen X, Li W. 2014. Isolation and expression profiles of Gibberellin metabolism genes in developing male and female cones of Pinus tabuliformis. Functional & Integrative Genomics 14:697−705

    doi: 10.1007/s10142-014-0387-y

    CrossRef   Google Scholar

    [95]

    Shearer RC, Stoehr MU, Webber JE, Ross SD. 1999. Seed cone production enhanced by injecting 38-year-old Larix occidentalis Nutt. with GA4/7. New Forests 18:289−300

    doi: 10.1023/A:1006612506340

    CrossRef   Google Scholar

    [96]

    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

    [97]

    Kong L, Aderkas P, Zaharia I, Abrams SR, Lee T, et al. 2012. Analysis of phytohormone profiles during male and female cone initiation and early differentiation in long-shoot buds of lodgepole pine. Journal of Plant Growth Regulation 31:478−89

    doi: 10.1007/s00344-011-9257-1

    CrossRef   Google Scholar

    [98]

    Du R, Niu S, Liu Y, Sun X, Porth I, et al. 2017. The gibberellin GID1-DELLA signalling module exists in evolutionarily ancient conifers. Scientific Reports 7:16637

    doi: 10.1038/s41598-017-11859-w

    CrossRef   Google Scholar

    [99]

    Katahata S, Futamura N, Igasaki T, Shinohara K. 2014. Functional analysis of SOC1-like and AGL6-likeMADS-box genes of the gymnosperm Cryptomeria japonica. Tree Genetics & Genomes 10:317−27

    doi: 10.1007/s11295-013-0686-9

    CrossRef   Google Scholar

    [100]

    Li W, Liu S, Ma J, Liu H, Han F, et al. 2020. Gibberellin signaling is required for far-red light-induced shoot elongation in Pinus tabuliformis seedlings. Plant Physiology 182:658−68

    doi: 10.1104/pp.19.00954

    CrossRef   Google Scholar

    [101]

    Street NR. 2019. Genomics of forest trees. In Advances in Botanical Research, ed. Cánovas FM. Vol. 89. UK: Academic Press. pp. 1−37. https://doi.org/10.1016/bs.abr.2018.11.001

    [102]

    Poovaiah C, Phillips L, Geddes B, Reeves C, Sorieul M, et al. 2021. Genome editing with CRISPR/Cas9 in Pinus radiata (D. Don). BMC Plant Biology 21:363

    doi: 10.1186/s12870-021-03143-x

    CrossRef   Google Scholar

    [103]

    Cui Y, Zhao J, Gao Y, Zhao R, Zhang J, et al. 2021. Efficient multi-sites genome editing and plant regeneration via somatic embryogenesis in Picea glauca. Frontiers in Plant Science 12:751891

    doi: 10.3389/fpls.2021.751891

    CrossRef   Google Scholar

    [104]

    Elfstrand M, Baison J, Lundén K, Zhou L, Vos I, et al. 2020. Association genetics identifies a specifically regulated Norway spruce laccase gene, PaLAC5, linked to Heterobasidion parviporum resistance. Plant, Cell & Environment 43:1779−91

    doi: 10.1111/pce.13768

    CrossRef   Google Scholar

    [105]

    Mukrimin M, Kovalchuk A, Neves LG, Jaber EHA, Haapanen M, et al. 2018. Genome-wide exon-capture approach identifies genetic variants of Norway Spruce genes associated with susceptibility to Heterobasidion parviporum infection. Frontiers in Plant Science 9:793

    doi: 10.3389/fpls.2018.00793

    CrossRef   Google Scholar

    [106]

    Di Pierro EA, Mosca E, Rocchini D, Binelli G, Neale DB, et al. 2016. Climate-related adaptive genetic variation and population structure in natural stands of Norway spruce in the South-Eastern Alps. Tree Genetics & Genomes 12:16

    doi: 10.1007/s11295-016-0972-4

    CrossRef   Google Scholar

    [107]

    Sena JS, Lachance D, Duval I, Nguyen TTA, Stewart D, et al. 2019. Functional analysis of the PgCesA3 white spruce cellulose synthase gene promoter in secondary xylem. Frontiers in Plant Science 10:626

    doi: 10.3389/fpls.2019.00626

    CrossRef   Google Scholar

    [108]

    Lamara M, Parent GJ, Giguère I, Beaulieu J, Bousquet J, et al. 2018. Association genetics of acetophenone defence against spruce budworm in mature white spruce. BMC Plant Biology 18:231

    doi: 10.1186/s12870-018-1434-y

    CrossRef   Google Scholar

    [109]

    Calleja-Rodriguez A, Li Z, Hallingbäck HR, Sillanpää MJ, Wu HX, et al. 2019. Analysis of phenotypic- and Estimated Breeding Values (EBV) to dissect the genetic architecture of complex traits in a Scots pine three-generation pedigree design. Journal of Theoretical Biology 462:283−92

    doi: 10.1016/j.jtbi.2018.11.007

    CrossRef   Google Scholar

    [110]

    Li Z, Hallingbäck HR, Abrahamsson S, Fries A, Gull BA, et al. 2014. Functional multi-locus QTL mapping of temporal trends in Scots pine wood traits. G3 Genes|Genomes|Genetics 4:2365−79

    doi: 10.1534/g3.114.014068

    CrossRef   Google Scholar

    [111]

    Plomion C, Chancerel E, Endelman J, Lamy JB, Mandrou E, et al. 2014. Genome-wide distribution of genetic diversity and linkage disequilibrium in a mass-selected population of maritime pine. BMC Genomics 15:171

    doi: 10.1186/1471-2164-15-171

    CrossRef   Google Scholar

    [112]

    Lepoittevin C, Harvengt L, Plomion C, Garnier-Géré P. 2012. Association mapping for growth, straightness and wood chemistry traits in the Pinus pinaster Aquitaine breeding population. Tree Genetics & Genomes 8:113−126

    doi: 10.1007/s11295-011-0426-y

    CrossRef   Google Scholar

    [113]

    Telfer E, Graham N, Macdonald L, Li Y, Klápště J, et al. 2019. A high-density exome capture genotype-by-sequencing panel for forestry breeding in Pinus radiata. PLoS One 14:e0222640

    doi: 10.1371/journal.pone.0222640

    CrossRef   Google Scholar

    [114]

    Liu JJ, Schoettle AW, Sniezko RA, Yao F, Zamany A, et al. 2019. Limber pine (Pinus flexilis James) genetic map constructed by exome-seq provides insight into the evolution of disease resistance and a genomic resource for genomics-based breeding. The Plant Journal 98:745−58

    doi: 10.1111/tpj.14270

    CrossRef   Google Scholar

    [115]

    Han X, Chen Q, Yang Q, Zeng Q, Lan T, et al. 2019. Genome-wide analysis of superoxide dismutase genes inLarix kaempferi. Gene 686:29−36

    doi: 10.1016/j.gene.2018.10.089

    CrossRef   Google Scholar

    [116]

    Niu S, Liu S, Ma J, Han F, Li Y, et al. 2019. The transcriptional activity of a temperature-sensitive transcription factor module is associated with pollen shedding time in pine. Tree Physiology 39:1173−1186

    doi: 10.1093/treephys/tpz023

    CrossRef   Google Scholar

    [117]

    Ma J, Liu S, Han F, Li W, Li Y, et al. 2020. Comparative transcriptome analyses reveal two distinct transcriptional modules associated with pollen shedding time in pine. BMC Genomics 21:504

    doi: 10.1186/s12864-020-06880-9

    CrossRef   Google Scholar

    [118]

    Langfelder P, Horvath S. 2008. WGCNA: an R package for weighted correlation network analysis. BMC Bioinformatics 9:559

    doi: 10.1186/1471-2105-9-559

    CrossRef   Google Scholar

    [119]

    Hakman IC, Arnold S. 1983. Isolation and growth of protoplasts from cell suspensions of Pinus contorta dougl. ex loud. Plant Cell Reports 2:92−94

    doi: 10.1007/BF00270174

    CrossRef   Google Scholar

    [120]

    Bekkaoui F, Datla RS, Pilon M, Tautorus TE, Crosby WL, et al. 1990. The effects of promoter on transient expression in conifer cell lines. Theoretical and Applied Genetics 79:353−59

    doi: 10.1007/BF01186079

    CrossRef   Google Scholar

    [121]

    Géomez-Maldonado J, Crespillo R, ÉAvila C, Céanovas FM. 2001. Efficient preparation of maritime pine (Pinus pinaster) protoplasts suitable for transgene expression analysis. Plant Molecular Biology Reporter 19:361−66

    doi: 10.1007/BF02772834

    CrossRef   Google Scholar

    [122]

    Guo Y, Song X, Zhao S, Lv J, Lu M. 2015. A transient gene expression system in Populus euphratica Oliv. protoplasts prepared from suspension cultured cells. Acta Physiologiae Plantarum 37:160

    doi: 10.1007/s11738-015-1906-8

    CrossRef   Google Scholar

    [123]

    Liang Z, Zong Y, Gao C. 2016. An efficient targeted mutagenesis system using CRISPR/Cas in monocotyledons. Current Protocols in Plant Biology 1:329−44

    doi: 10.1002/cppb.20021

    CrossRef   Google Scholar

    [124]

    Liu Z, Wu Y, Yang F, Zhang Y, Chen S, et al. 2013. BIK1 interacts with PEPRs to mediate ethylene-induced immunity. PNAS 110:6205−10

    doi: 10.1073/pnas.1215543110

    CrossRef   Google Scholar

    [125]

    Denyer T, Timmermans MCP. 2022. Crafting a blueprint for single-cell RNA sequencing. Trends in Plant Science 27:92−103

    doi: 10.1016/j.tplants.2021.08.016

    CrossRef   Google Scholar

    [126]

    Denyer T, Timmermans MCP. 2022. High-throughput single-cell RNA sequencing. Trends in Plant Science 27:104−105

    doi: 10.1016/j.tplants.2021.09.003

    CrossRef   Google Scholar

    [127]

    Liu S, Ma J, Liu H, Guo Y, Li W, et al. 2020. An efficient system for Agrobacterium-mediated transient transformation in Pinus tabuliformis. Plant Methods 16:52

    doi: 10.1186/s13007-020-00594-5

    CrossRef   Google Scholar

    [128]

    Borthakur D, Busov V, Cao XH, 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

    [129]

    Lin YH, Li W, Sun YH, Kumari S, Wei H, et al. 2013. SND1 transcription factor-directed quantitative functional hierarchical genetic regulatory network in wood formation in Populus trichocarpa. The Plant Cell 25:4324−41

    doi: 10.1105/tpc.113.117697

    CrossRef   Google Scholar

    [130]

    Wei H. 2019. Construction of a hierarchical gene regulatory network centered around a transcription factor. Briefings in Bioinformatics 20:1021−31

    doi: 10.1093/bib/bbx152

    CrossRef   Google Scholar

    [131]

    Meng D, Yang Q, Dong B, Song Z, Niu L, Wang L, et al. 2019. Development of an efficient root transgenic system for pigeon pea and its application to other important economically plants. Plant Biotechnology Journal 17:1804−13

    doi: 10.1111/pbi.13101

    CrossRef   Google Scholar

    [132]

    Bao W, Wang J, Wang Q, O'Hare D, Wan Y. 2016. Layered double hydroxide Nanotransporter for molecule delivery to intact plant cells. Science Reports 6:26738

    doi: 10.1038/srep26738

    CrossRef   Google Scholar

    [133]

    Hasanzadeh A, Radmanesh F, Hosseini ES, Hashemzadeh I, Kiani, J, et al. 2021. Highly photoluminescent Nitrogen- and Zinc-doped carbon dots for efficient delivery of CRISPR/Cas9 and mRNA. Bioconjugate Chemistry 32:1875−87

    doi: 10.1021/acs.bioconjchem.1c00309

    CrossRef   Google Scholar

  • Cite this article

    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
    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

Figures(2)

Article Metrics

Article views(4786) PDF downloads(700)

REVIEW   Open Access    

The long road to bloom in conifers

Forestry Research  2 Article number: 16  (2022)  |  Cite this article

Abstract: More than 600 species of conifers (phylum Pinophyta) serve as the backbone of the Earth’s terrestrial plant community and play key roles in global carbon and water cycles. Although coniferous forests account for a large fraction of global wood production, their productivity relies largely on the use of genetically improved seeds. However, acquisition of such seeds requires recurrent selection and testing of genetically superior parent trees, eventually followed by the establishment of a seed orchard to produce the improved seeds. The breeding cycle for obtaining the next generation of genetically improved seeds can be significantly lengthened when a target species has a long juvenile period. Therefore, development of methods for diminishing the juvenile phase is a cost-effective strategy for shortening breeding cycle in conifers. The molecular regulatory programs associated with the reproductive transition and annual reproductive cycle of conifers are modulated by environmental cues and endogenous developmental signals. Mounting evidence indicates that an increase in global average temperature seriously threatens plant productivity, but how conifers respond to the ever-changing natural environment has yet to be fully characterized. With the breakthrough of assembling and annotating the giant genome of conifers, identification of key components in the regulatory cascades that control the vegetative to reproductive transition is imminent. However, comparison of the signaling pathways that control the reproductive transition in conifers and the floral transition in Arabidopsis has revealed many differences. Therefore, a more complete understanding of the underlying regulatory mechanisms that control the conifer reproductive transition is of paramount importance. Here, we review our current understanding of the molecular basis for reproductive regulation, highlight recent discoveries, and review new approaches for molecular research on conifers.

    • Living gymnosperms comprise four of the five main lineages of seed plants: cycads, ginkgos, gnetophytes, and conifers[1]. In contrast to annual plants like Arabidopsis, conifers are perennials that undergo a long juvenile phase and repeated cycles of vegetative growth, dormancy, and reproductive growth controlled by distinct, complex reproductive regulatory mechanisms. Conifer cones are reproductive shoots that are more similar to inflorescences than to individual flowers[2], and reproductive organ identity and development in conifers differ markedly from those in angiosperms[3, 4]. In addition, most conifer reproductive cycle spans at least two years. Coniferous meristems and perennating organs therefore endure tremendous environmental changes and rely, to a great extent, on specific reproductive strategies. Environmental cues (photoperiod, temperature) and endogenous factors (age, developmental stage, plant hormone levels) influence the timing of the developmental transition from vegetative to reproductive growth, which is critical for reproductive success. Conifers in boreal and temperate regions survive climatic extremes by integrating endogenous developmental signals with environmental cues to initiate reproductive growth at an opportune time[58]. In recent years, a number of crucial molecular regulators that control conifer reproduction have been identified[9, 10], largely as a result of large-scale genomic sequencing in a variety of species, such as Pinus tabuliformis (Chinese pine)[8], Pinus taeda (loblolly pine)[11], Pinus lambertiana (sugar pine)[12], Picea glauca (white spruce)[13], and Picea abies (Norway spruce)[14]. In this review, we summarize our current understanding of the cellular and molecular mechanisms involved in reproductive induction and highlight future prospects for conifer molecular biology research.

    • Day length is a major environmental factor that controls photoperiodism and influences flowering, bud break, and dormancy in angiosperm plants[5, 15]. GIGANTEA (GI), which promotes the transcription of CONSTANS (CO), performs central functions in the transmission of light signals in the photoperiodic pathway of Arabidopsis[16]. The steady, continuous accumulation of CO protein directly induces expression of the downstream target gene FLOWERING LOCUS T (FT) in leaves, and FT protein is then transported to the apical meristem through the phloem[17]. FT forms protein complexes with the bZIP transcription factor FLOWERING LOCUS D (FD) in the apical meristem to activate SUPPRESSOR OF OVEREXPRESSION OF CO 1 (SOC1), APETALA1 (AP1) and FRUITFULL (FUL) to participate in flower induction[18].

      The function of GI is thought to have been conserved during plant evolutionary history, not only in angiosperms[19] but also in early land plants such as Selaginella tamariscina[20] and conifers such as P. abies[21]. Overexpression of GI from P. abies in Arabidopsis produced no obvious phenotype but partially rescued the late-flowering phenotype of gi-2 mutants[21]. The expression of GI in P. abies and Picea obovata confirmed its key roles in the control of seasonal growth cessation in spruce species[22]. Moreover, endogenous silencing of GI or FLAVIN-BINDING KELCH REPEAT F-BOX1 (KFK) homologs in S. tamariscina completely eliminated its reproductive phase transition, which relied on day length, and ectopic expression of GI and KFK promoted the floral transition under short days in Arabidopsis[20]. The mechanism by which the GI-FKF1 system regulates reproductive growth upstream of the photoperiodic pathway may thus be conserved throughout vascular plants.

      Plants have adapted to the day/night cycle by evolving a circadian clock system that is closely related to the photoperiodic pathway and drives matching rhythms in many aspects of metabolism and physiology[23, 24]. Nucleotide diversity data from P. abies indicate that PSEUDO RESPONSE REGULATOR 3 (PRR3) and ZEITLUPE (ZTL) harbor multiple non-synonymous variants and appear to be excellent candidate genes for control of the photoperiod response[21]. CIRCADIAN CLOCK ASSOCIATED 1 (CCA1), GI, ZTL, and PRR1, which are major components of the circadian clock loops, show functional conservation between P. abies and Arabidopsis, although they displayed different expression patterns and their expression was rapidly dampened under constant light conditions[21]. In short, the biological circadian clock network has an important role in the photoperiodic control of reproductive development, and it appears to have been largely present before the divergence of conifers and angiosperms. The GI gene in conifers may have both conserved and specific roles in the regulation of annual rhythms upstream of the photoperiodic pathway, together with other circadian clock genes (Fig. 1).

      Figure 1. 

      Current understanding of flowering pathways in conifers. Five interdependent pathways control the reproductive transition in conifers: the vernalization, photoperiod, autonomous, gibberellin, and aging pathways. Arrows indicate promotion, blunt-ended lines indicate genetic inhibition, and curves indicate protein–protein interactions. Solid lines denote interactions that are supported by experimental evidence, whereas dashed lines denote proposed interactions. Genes that act as major regulators in different pathways are written in green blocks. Environmental factors are represented by pink ellipses, and hormones involved in reproductive growth are represented by red hexagons.

      Bud break in P. abies is promoted under long-day conditions[25], and the expression of two CO homologs increased after transfer from dark to light conditions[10]. Analysis of the annual transcriptome dynamics of Cryptomeria japonica also revealed conserved expression patterns of CO homologs in angiosperms and conifers[26], suggesting that conifer CO genes may be candidate inducers of reproductive growth initiation in response to photoperiod (Fig. 1).

      The functional conservation of the CO-FT regulatory module in the photoperiod response has been confirmed in perennial woody trees like poplar[27]. However, angiosperms and conifers diverged about 300 million years ago[28], and functional differentiation of FT-like genes has occurred. In a phylogenetic study, the FT/TFL1-like genes of P. abies were clustered at the base of the branch node that separates FT and TERMINAL FLOWER LIKE-1 (TFL1) genes, and key amino acid sites for FT function were preserved[29]. The conifer FTL1 and FTL2 genes arose from a duplication event in a common ancestor of gymnosperms and play roles in the pathways that control growth rhythm and reproductive development[9]. Expression of conifer FTL2 declined rapidly during spring bud break and increased before bud set and the onset of dormancy in late summer and autumn, suggesting that it has an important role in the annual growth rhythm[9, 21, 22, 30]. FTL1 displayed the opposite pattern of photoperiodic expression and controlled bud set and temperature-mediated bud break[31, 32]. Interestingly, overexpression of both PaFTL1 and PaFTL2 in transgenic Arabidopsis lines inhibited flowering, and a similar result was also observed in P. tabuliformis[4], suggesting that conifer FTL proteins are more functionally similar to TFL1 than to FT of angiosperms[32]. FT-like genes may have undergone functional divergence over the course of subsequent evolution in seed plants, including FT genes, which encode growth activators, and FT/TFL1-like genes, which encode growth inhibitors and are more functionally similar to TFL1. Further studies are needed on the photoperiod-related functions and regulatory mechanisms of FT/TFL1-like genes in conifers.

    • In addition to light, plants also respond strongly to other environmental stimuli like temperature. In Arabidopsis, FLOWERING LOCUS C (FLC) functions as a key temperature monitor that integrates floral signals in the vernalization pathway and releases the inhibition of FT and SOC1 genes[3335]. In conifers, autumn dormancy in response to low temperatures and reproductive induction occur during the same growth period[23, 31]. Transcriptome and genome sequencing data suggest that FLC-like homologs arose after the divergence of angiosperms and conifers[14, 36], and gene(s) with a homologous function in the regulation of conifer vernalization pathways have not yet been found[32]. However, some potential key genes that may be involved in conifer vernalization have been identified (Fig. 1). NAM/ATAF/CUC2 (NAC2) from Picea wilsonii enhanced drought and salt stress tolerance via several signaling pathways and promoted flowering in transgenic Arabidopsis through its interaction with the Resemble-FCA-contain-PAT1 domain (RFCP1) protein[37]. Cold stimulation in winter did not lead to an increase in VERNALIZATION INSENSITIVE 3 (VIN3) expression in P. abies, whereas VIN3 transcription was promoted by vernalization in wheat[38]; these contrasting results may reflect differences in the physiological functions of conifer needles and spring wheat apical meristems.

      In Populus trees (poplars, aspens and cottonwoods), the CO/FT2 regulatory module regulates the short-day–induced growth cessation in fall[27]. While, FT1 is hyper-induced by chilling and functions on the release of winter dormancy in Populus trees[39]. The photoperiod pathway and vernalization pathway may thus also share common components in conifers. The conifer FT/TFL1-like genes, which are regulated by low temperature and short-day conditions in the autumn, play important roles in growth cessation and endogenous dormancy in response to chilling stress[21, 30]. FTL2 was also reported to function as a key integrator of the photoperiod pathway during growth rhythm control in P. abies[40]. Long-day conditions with high temperatures during the day and low temperatures at night could bypass the typical rhythm cycle and bring about growth cessation[41]. Light and temperature are important environmental signals for the seasonal acclimation process in conifers[42]. Although the specific mechanisms remain unclear and require further study, it appears that the FT/TFL1-like genes may function as key regulators of both the photoperiod and vernalization pathways in conifers (Fig. 1).

    • At a specific stage of their life cycle, plants may undergo reproductive development independent of day length under the control of endogenous signals via the so-called autonomous pathway, enabling their survival even under unsuitable external environmental conditions. FLC serves as the key node of the gene network that controls this autonomous pathway in angiosperms[33, 43]. In the upstream pathway, the RNA-binding protein FCA controls the expression of alternatively polyadenylated antisense RNAs at the FLC locus[44], and the RNA-binding protein FPA prevents the accumulation of FLC mRNA in order to induce flowering[43, 45]. However, current evidence suggests that homologs of angiosperm FLC genes do not exist in the conifer lineage[14, 36]. Identification of transcription factors that functionally substitute for FLC will provide further insight into the control of reproduction via the autonomous pathway in conifers.

      Researches on Pinus pinaster and P. wilsonii revealed the potential functions of the NAC transcription factors not only on stress responses but also related to reproductive regulation[37,46]. Ectopic expression of the RFCP1 transcription factor from P. wilsonii in Arabidopsis significantly accelerated flowering by negatively regulating FLC expression[37]. Moreover, promotion of hypocotyl growth by PwRFCP1 in Arabidopsis was independent of light, suggesting that RFCP1 may modulate reproductive growth by the autonomous pathway, i.e. independently of photoperiod, in conifers[37]. Specifically, RFCP1 could function as a key component in the conifer autonomous pathway by negatively regulating reproductive inhibitors (Fig. 1). The upstream components of the autonomous pathway are more likely to be conserved in angiosperms and conifers, whereas the downstream mechanisms may differ in conifers because they lack an FLC ortholog.

      The FTL2 gene, which acts as a reproduction suppressor in conifers, displays expression patterns similar to those of angiosperm FLC, with high accumulation in bud crowns[32, 47]. In addition, expression of FTL2 increases before the formation of reproductive buds in P. abies and Pinus sylvestris but decreases when the reproductive buds open[30, 48, 49]. Although conifer FTL2 belongs to a different gene family than angiosperm FLC and its homolog functions differently in angiosperms, current studies indicate that the expression patterns and functions of FTL2 in reproductive growth inhibition of conifers are similar to those of FLC in angiosperms[30, 47]. The inhibitory effects of FTL2 on reproductive growth, which are biochemically more similar to those of angiosperm TFL1-like, are conserved in conifers such as Picea sitchensis, P. glauca, Picea engelmannii × glauca, Pinus tabuliformis, and Pinus contorta[32, 49]. Moreover, conifer FTL2 prevented flowering and rescued the phenotypes of tf1-14 mutants when ectopically expressed in Arabidopsis[49]. FTL2 may therefore function as a key component of the autonomous pathway to regulate the reproductive transition in conifers, and the mechanism by which it controls transcriptional activity requires further exploration (Fig. 1).

      LFY/FLO, the downstream target of FT/TFL1 in the floral repression pathway, regulates B- and C-class floral organ identity genes to control floral meristem development in angiosperms[50]. Two similar paralogs of LFY-like genes are present in all major extant conifer groups[51, 52]; they were first isolated from Pinus radiata and named PrLFY and PrNLY[53, 54]. Phylogenetic analysis revealed that NEEDLY (NLY) was lost in flowering plants before the expansion and subsequent evolution of extant angiosperm lineages[52]. In all conifers studied to date, LFY was highly accumulated during reproductive organ development, revealing its functional conservation in the initiation of reproductive development in both angiosperms and gymnosperms[51, 5357]. The acrocona mutant in P. abies bear female cones on the vegetative branches, and LFY expression is upregulated in the transformed reproductive structures, supporting a vital role for LFY in the female organ formation of P. abies[58]. Seed and pollen cones are separate reproductive shoots that may be regulated by different mechanisms, and B-type genes act as dominant activators of male cone identity[59]. Ectopic expression of gymnosperm B-class gene APETALA3/PISTILLATA-LIKE (AP3/PI-like) from Gnetum and C-class gene AGAM0US-LIKE (AG-like) from Cycas edentata rescued the phenotypes of ap3-1, pi-1 and ag-2 Arabidopsis mutants, respectively, suggesting the biochemical conservation of B- and C-class floral genes in seed plants[60, 61]. The two LFY-like paralogs of Welwitschia mirabilis, LFY, and NLY, displayed significantly different DNA binding specificities, and only LFY effectively bound to the AP3/PI-like genes promoter genes in Welwitschia[62]. Therefore, LFY-like genes in gymnosperms appear to have undergone functional differentiation over the course of evolution, such that conifer LFY shares with its angiosperm ortholog the capacity to regulate reproductive growth by binding directly to B-gene promoters (Fig. 1).

    • The age-related pathway in perennial Arabis alpina is similar to that in annual Arabidopsis, which is regulated by the sequential action of two microRNAs, miR156 and miR172[63]. Typically, miR156 levels decline as A. alpina and Arabidopsis age increases, whereas miR172 shows the opposite expression pattern[64, 65]. PERPETUAL FLOWERING 2 (PEP2), an APETALA2 transcription factor, is a target of miR172 and prevents flowering before vernalization in A. alpina[66]. Reduced levels of miR156 cause increased production of SQUAMOSA PROMOTER BINDING PROTEIN LIKESPL (SPL) transcription factors to promote the transition from vegetative growth to reproduction in both A. alpina and Arabidopsis[67,68]. The A. alpina gene PERPETUAL FLOWERING 1 (PEP1), the ortholog of Arabidopsis FLC, mechanistically links polycarpy with seasonal flowering[68], and continuous flowering forms have arisen multiple times through PEP1-1 mutations[67]. Although homologs of angiosperm FLCs are not present in the conifer lineage[14, 36], identification of transcription factors that functionally substitute for PEP1 may provide further insight into the ageing pathway in conifers.

      miR156 and miR172 post-transcriptional regulatory modules and their target genes have been identified in conifer species[6971]. SBP-box genes contain highly conserved miR156 target sites in conifers such as P. taeda and P. glauca[71], and miR156 and miR172 specifically cut the target mRNAs SPL1,2,3 and AP2L1,2,3 in P. tabuliformis[70]. SPL1 of P. abies harbors conserved binding sites for miR156 and miR529, and the SPL-miR156/miR529 regulatory module in the age-dependent pathway appears to be highly conserved[72]. miR172 also has highly conserved AP2 homolog target sites in conifers[7375]. In general, miR156 and miR172 target genes appear to be conserved in seed plants, although miR156 and miR172 levels may uncoupled in perennial plants[64,65]. Further study is needed to assess the regulatory relationships between miR156 and miR172 and their functions in the vegetative growth phase transition of conifers.

      A study that specifically screened MADS-box genes from a cDNA library of P. abies seedlings identified three DEFICIENS AGAMOUS-LIKE (DAL) genes (DAL1DAL3), as homologs of Arabidopsis AGL6[76]. DAL1 expression increased with development and could serve as an age-related marker in P. abies and Larix kaempferi, whose physiological and morphological characteristics were consistent with the age-related pattern of reproductive growth[57, 77]. Constitutive expression of conifer DAL1 in transgenic Arabidopsis plants dramatically accelerated flowering, suggesting a regulatory role for DAL1 in the transformation from vegetative to reproductive growth in conifers[7, 57]. Moreover, DAL1 physically interacted with MADS11 (SOC1-like), and the MADS11–DAL1 module appeared to function as a regulatory component of the juvenile-adult phase transition in P. tabuliformis[7]. The number of genes in the SOC1-like clade is greatly expanded in conifers compared with angiosperms[78], resulting not only from expansion of the gene family through gene duplication events but also from the production of numerous splice variants[79]. Members of this subclade also express distinct splice variants in different bud types. The SOC1-like gene DEFICIENS AGAMOUS LIKE 19 (DAL19) is specifically upregulated in cone-setting shoots, and its two mutually exclusive exons play key roles in the vegetative-to-reproductive phase change in P. abies[79, 80]. Interestingly, the DAL1 was found to have widely physical interaction with many transcription factors including DAL19 in P. tabuliformis[8, 81]. Taken together, these results suggest that DAL1 has a conserved age-related expression pattern and clearly affects the phase transition process through interaction with SOC1-like proteins. It may function as a key regulator in the conifer maturation pathway and is therefore deserving of continued research attention (Fig. 1).

    • Gibberellin (GA), an essential plant hormone, is involved in regulating many events during the plant life cycle, and its role in floral development has been widely studied[8284]. In Arabidopsis, GA promotes flowering by activating LEAFY and eliminating the inhibition of SPL transcription factors by DELLA protein, thereby activating FUL and SOC1 genes to promote flowering[82, 85, 86]. DELLA also mediates FT expression to control flowering time by directly regulating the PIF gene[87, 88]. In addition to DELLA protein, miR159 also has important functions upstream in the GA pathway[84, 89].

      Various biotechnological approaches have been used to shorten the breeding cycles of conifers, and exogenous GA in particular has been highly effective and is widely applied[9092]. GA3 is most commonly used to promote reproductive growth and increase yields in Cupressaceae and Taxodiaceae species[92, 93], whereas non-polar GA4/7 is more efficient for application to Pinaceae; the latter has been shown to stimulate reproduction and increase production in at least 12 pine species, as well as Larix lepepis and L. occidentalis[9496]. Combined exogenous application of GA4/7 and the cytokinin analog thidiazuron (TDZ) to long-shoot buds increases female strobili formation in P. contorta, highlighting the potential function of GA in conifer sex determination[93, 97].

      Despite ongoing efforts to elucidate the mechanisms by which GA promotes reproductive growth and sexual reversion at an early developmental stage in conifers, the specific genes that respond to GA signals and the downstream regulatory mechanisms of the GA pathway remain unclear. A DELLA homolog in P. tabuliformis interacts with PtGID1, which functions as a GA receptor, suggesting the conservation of the GA–GID1–DELLA signaling module in conifers[98]. Expression of an SOC1-like (MADS15) gene was significantly upregulated after exogenous GA3 application in C. japonica, indicating that the conifer SOC1-like gene may be a downstream target of GA signaling in conifers (Fig. 1)[99]. A study in P. tabuliformis revealed that the regulatory targets for GA biosynthesis differ between conifers and angiosperms[100]. In short, angiosperms and conifers share similar regulatory mechanisms in the GA signaling pathway, but the metabolic pathway of GA signaling appears to be different. Identification of more genes that respond to GA signaling and construction of the associated gene regulatory network should be directions for future research.

    • Studies of growth rhythms and developmental regulation have been lagging behind in conifers owing to their very large genomes and highly heterozygous genetic backgrounds. Moreover, because of the large genetic distance between angiosperms and gymnosperms, technical systems that are widely used in Arabidopsis and crops are difficult to apply directly to conifers, greatly impeding progress in conifer molecular research. Conifers have great ecological and economic value as well as a significant impact on forest carbon sinks, and studying their reproductive patterns is crucial for advancing our understanding of seed plant evolution. Next, we summarize techniques that are currently used in conifer molecular biology research and propose three research strategies for future investigations of genetic regulatory mechanisms in conifers (Fig. 2).

      Figure 2. 

      Strategies for identification and characterization of conifer genes and their regulatory relationships. Omics technologies combine GWAS and WGCNA based on RNA-seq data to identify key genes that determine important traits. Transient transformation systems overcome restrictions on genetic transformation, enabling integration of target plasmids into conifer chromosomes to produce functional proteins. A genetic regulatory network can then be constructed from RNA-seq data. The aim of genome editing is to precisely modify target genes or regulatory elements in conifers. Tissue culture–free delivery systems include delivery via plant germline or meristematic cells and nanotechnology-based delivery systems.

    • Because the large genomes of conifers contain 70%–80% repetitive sequences and numerous redundant genes[101], traditional mutation techniques (EMS, radiation, UV) that do not rely on genetic transformation are inefficient for obtaining functional mutations in conifers. To date, there have been few reports on genome editing in conifers, mainly because it is difficult to transform and integrate exogenous genes. Recently, CRISPR/Cas9-mediated targeted mutagenesis has been reported in P. radiata[102] and P. glauca[103]. Thus, reverse genetics can be used in conifer research, obtaining the sequence of a gene of interest before mutating the gene to verify its function. Genome-wide association analysis (GWAS) has been used in a variety of conifer species such as P. abies[104106], P. glauca[107, 108], P. sylvestris[109, 110], P. pinaster[111, 112], P. radiata[113], Pinus flexilis[114], and L. kaempferi[115], to identify candidate genes associated with reproductive development, and research on conifer molecular mechanisms has thus entered the genomics era. Transcriptomic sequencing combined with gene co-expression network analysis in P. tabuliformis successfully identified gene modules that control pollen shedding time in response to temperature[116]. Taking full advantage of existing transcriptome data and sophisticated analytical methods such as weighted gene co-expression network analysis (WGCNA) can therefore overcome the current impasse in conifer molecular investigation[117, 118].

    • Transient transformation is another potential strategy for investigating molecular mechanisms in conifers. In conifer biotechnology, protoplast extraction was first performed in P. contorta, laying the foundation for establishment of a conifer transient transformation system[119]. Protoplasts from suspension cultures of P. glauca somatic embryos have been electroporated with plasmids[120], and a technique for isolating shoot protoplasts and driving transient gene expression via electroporation has been reported in P. pinaster[121]. In related tree biotechnology research, a transient gene expression protocol was developed for the simultaneous co-transformation of two proteins in the same protoplasts of Populus euphratica[122]. The protoplast transient expression system has also been widely used for CRISPR/Cas-based genome modification as a powerful tool for in-depth investigation of gene function[123, 124]. Protoplast transient transformation technology is thus very valuable for the rapid assessment of gene functions and physical interactions (Fig. 2), and it will be particularly useful for systems studies of conifers in which stable transgenic plants and mutants are unavailable. Stable and efficient protoplast transformation may enable the use of high-throughput, droplet-based single-cell RNA sequencing (scRNA-Seq) in conifers, allowing researchers to examine cell-cell heterogeneity in tissues and organs with an unprecedented degree of resolution[125]. At present, the large size of conifer protoplasts (~70 nm diameter) limits this approach: oil droplets can only wrap cells less than 40 nm in diameter owing to surface tension[126]. Improvements in the capacity of oil droplets to wrap larger cells will thus promote the application of scRNA-Seq to conifers.

      In addition to protoplast transformation, Agrobacterium-mediated transient transformation of callus and hypocotyls in P. tabuliformis has been reported; combined with transcriptome analysis, this approach could efficiently confirm gene regulatory relationships in conifers[127]. However, because of the tissue-specificity of plant gene expression, genes related to reproductive development are typically silenced in callus and hypocotyl tissues. The use of transient callus or hypocotyl transformation to study molecular mechanisms of reproduction and development will thus require further improvements.

      No matter which transient expression system is employed, computational methods have been developed for inferring the direct target genes or the impacted genes of a transformed gene[128]. For example, Top-down GGM Algorithm[129,130] is especially suitable for using transient expression data to identify the direct target genes or the influenced genes of an overexpressed/suppressed gene. This is because the gene delivered into a transient expression system is generally perturbed, allowing the target genes or impacted genes to be recognized.

    • A simple, fast, and efficient technique for generating stable transgenic roots in living plants by Agrobacterium rhizogenes-mediated transformation has recently been reported[131]. Positively charged nanosheets have also been used to facilitate the transport of biologically active materials across the plasma membrane into plant cells via non-endocytic pathways, a strategy that might also be applied to conifers[132]. Naturally occurring carbon dots have been used as rapid vehicles for carrying plasmids into mature plant cells, resulting in transient transformation[133]. All these approaches can be used without a regeneration system and therefore show great promise for conifer transformation (Fig. 2). In future research, genetic modification using nanomaterials will broaden the horizons of plant molecular research, especially for conifers, which lack systems for regeneration and stable genetic transformation.

    • Molecular genetic approaches have provided insights into the mechanisms involved in the reproductive transition of conifers. Positive and negative regulators integrate signals from different regulatory pathways to modulate the timing of the reproductive process. However, there are no direct homologs of FLC in conifers, and all of their FT/TFL1-like genes appear to function more like TFL1, acting as reproductive repressors[14, 32]. TFL2 may function as an integrator of the photoperiod and vernalization pathways in conifers; its expression patterns respond to SD conditions and display annual rhythms, suggesting that the reproductive and developmental regulatory pathways of conifers may reflect more ancient evolutionary mechanisms[21, 22]. The complexity of the conifer genetic background and the lack of a reproductive transformation system significantly impede the research progress on conifer regulatory mechanisms. Identifying key genes in conifer regulatory networks and establishing regeneration-free techniques for gene functional characterization are therefore important scientific challenges.

      • This work was supported by the Scientific Research Development Fund Project of Zhejiang Agriculture and Forestry University (2021LFR051), the National Natural Science Foundation of China (No. 31870651), and the Fundamental Research Funds for the Central Universities (No. 2015ZCQ-SW-02).

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

      • 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)  References (133)
  • About this article
    Cite this article
    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
    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

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return