Hostname: page-component-76fb5796d-x4r87 Total loading time: 0 Render date: 2024-04-26T01:08:09.941Z Has data issue: false hasContentIssue false

A current review on the regulation of dormancy in vegetative buds

Published online by Cambridge University Press:  20 January 2017

Wun S. Chao
Affiliation:
U.S. Department of Agriculture, Agricultural Research Service, Plant Science Research, 1605 Albrecht Boulevard, P.O. Box 5674, State University Station, Fargo, ND 58105
David P. Horvath
Affiliation:
U.S. Department of Agriculture, Agricultural Research Service, Plant Science Research, 1605 Albrecht Boulevard, P.O. Box 5674, State University Station, Fargo, ND 58105

Abstract

In this review, we examine current techniques and recent advances directed toward understanding cellular mechanisms involved in controlling dormancy in vegetative propagules. Vegetative propagules (including stems, rhizomes, tubers, bulbs, stolons, creeping roots, etc.) contain axillary and adventitious buds capable of producing new stems/branches under permissive environments. Axillary and adventitious buds are distinct in that axillary buds are formed in the axil of leaves and are responsible for production of lateral shoots (branches). Adventitious buds refer to buds that arise on the plant at places (stems, roots, or leaves) other than leaf axils. Both axillary and adventitious buds generally undergo periods of dormancy. Dormancy has been described as a temporary suspension of visible growth of any plant structure containing a meristem (Lang et al. 1987). Dormancy can be subdivided into three categories: (1) ecodormancy-arrest is under the control of external environmental factors; (2) paradormancy-arrest is under the control of external physiological factors within the plant; and (3) endodormancy-arrest is under the control of internal physiological factors. One common feature in all of these processes is prevention of growth under conditions where growth should otherwise continue. There is growing evidence that lack of growth is due to blockage of cell division resulting from interactions between the signaling pathways controlling dormancy and those controlling the cell cycle.

Type
Review Article
Copyright
Copyright © Weed Science Society of America 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Literature Cited

Aharoni, A., Keizer, L.C.P., Bouwmeester, H. J., et al. 2000. Identification of the SAAT gene involved in strawberry flavor biogenesis by use of DNA microarrays. Plant Cell 12:647661.CrossRefGoogle ScholarPubMed
Anderson, J. V. and Horvath, D. P. 2000. A genomics approach to investigating bud dormancy and developmental regulation in underground adventitious buds. Page 135 in Plant Biology 2000 Final Program and Abstract Supplement, Rockville, MD: American Society of Plant Physiologists.Google Scholar
Anderson, J. V. and Horvath, D. P. 2001. Random sequencing of cDNAs and identification of mRNAs. Weed Sci. 49:590597.CrossRefGoogle Scholar
Beveridge, C. A., Symons, G. M., and Turnbull, C.G.N. 2000. Auxin inhibition of decapitation-induced branching is dependent on grafttransmissible signals regulated by genes Rms1 and Rms2. Plant Physiol. 123:689698.Google Scholar
Borchert, R. 1991. Growth periodicity and dormancy. Pages 221245 In Raghavendra, A. S., ed. Physiology of Trees. New York: Wiley.Google Scholar
Burton, R. A., Gibeaut, D., Bacic, A., Findlay, K., Roberts, K., Hamilton, A., Baulcombe, D. C., and Fincher, G. B. 2000. Virus-induced silencing of a plant cellulose synthase gene. Plant Cell 12:691705.CrossRefGoogle ScholarPubMed
Campbell, M. A., Suttle, J. C., and Sell, T. W. 1996. Changes in cell cycle status and expression of p34cdc2 kinase during potato tuber meristem dormancy. Physiol. Plant 98:743752.CrossRefGoogle Scholar
Cary, A. J., Liu, W., and Howell, S. H. 1995. Cytokinin action is coupled to ethylene in its effects on the inhibition of root and hypocotyl elongation in Arabidopsis thaliana seedlings. Plant Physiol. 107:10751082.Google Scholar
Chaloupkova, K. and Smart, C. C. 1994. The abscisic acid induction of a novel peroxidase is antagonized by cytokinin in Spirodela polyrrhiza L. Plant Physiol. 105:497507.Google Scholar
Chao, W. S., Anderson, J. V., and Horvath, D. P. 2000. Sugar plays a role in inhibition of underground adventitious bud growth in leafy spurge (Euphorbia esula L.). Page 46 in Plant Biology 2000 Final Program and Abstract Supplement, Rockville, MD: American Society of Plant Physiologists.Google Scholar
Chen, H. H., Li, P. H., and Brenner, M. L. 1983. Involvement of abscisic acid in potato cold acclimation. Plant Physiol. 71:362365.Google Scholar
Chiariello, M., Gomez, E., and Gutkind, J. S. 2000. Regulation of cyclin-dependent kinase (Cdk2) Thr-160 phosphorylation and activity by mitogen-activated kinase in late G1 phase. Biochem. J. 349:869876.CrossRefGoogle ScholarPubMed
Chien, M., Rinker-Schaeffer, C., and Stadler, W. M. 2000. A G2/M growth arrest response to low-dose intermittent H2O2 in normal uroepithelial cells. Int. J. Oncol. 17:425432.Google Scholar
Cline, M. G. 1991. Apical dominance. Bot. Rev. 57:318358.CrossRefGoogle Scholar
Cook, S. J., Balmanno, K., Garner, A., Millar, T., Taverner, C., and Todd, D. 2000. Regulation of cell cycle re-entry by growth survival and stress signaling pathways. Biochem. Soc. Trans. 28:233240.Google Scholar
Coupland, R. T., Selleck, G. W., and Alex, J. F. 1955. Distribution of vegetative buds on the underground parts of leafy spurge (Euphorbia esula L.). Can. J. Agric. Sci. 35:161167.Google Scholar
Crabbé, J. and Barnola, P. 1996. A new conceptual approach to bud dormancy in woody plants. Pages 83113 In Lang, G. A., ed. Plant Dormancy: Physiology, Biochemistry and Molecular Biology. Wallingford, UK: CAB International.Google Scholar
Dutt, M. J. and Lee, K. H. 2000. Proteomic analysis. Curr. Opin. Biotechnol. 11:176179.Google Scholar
Eagles, C. F. and Wareing, P. F. 1963. Experimental induction of dormancy in Betula pubescens . Nature 199:874875.Google Scholar
Eagles, C. F. and Wareing, P. F. 1964. The role of growth substances in the regulation of bud dormancy. Plant Physiol. 17:697709.Google Scholar
Eisen, M. B., Spellman, P. T., Brown, P. O., and Botstein, D. 1998. Cluster analysis and display of genome-wide expression patterns. Proc. Natl. Acad. Sci. USA 95:1486314868.Google Scholar
Epstein, C. B. and Butow, R. A. 2000. Microarray technology—enhanced versatility, persistent challenge. Curr. Opin. Biotechnol. 11:3641.Google Scholar
Faiss, M., Zalubìlová, J., Strnad, M., and Schmülling, T. 1997. Conditional transgenic expression of the ipt gene indicates a function for cytokinins in paracrine signaling in whole tobacco plants. Plant J. 12:401415.Google Scholar
Frewen, B. E., Chen, T.H.H., Howe, G. T., Davis, J., Rohde, A., Boerjan, W., and Bradshaw, H. D. Jr. 2000. Quantitative trait loci and candidate gene mapping of bud set and bud flush in Populus. Genetics 154:837845.Google Scholar
Frugis, G., Giannino, D., Mele, G., et al. 1999. Are homeobox knotted-like genes and cytokinins the leaf architects? Plant Physiol. 119:371374.Google Scholar
Galitz, D. S. 1994. The biology of leafy spurge. Pages 5762 In Schmidt, C. H., ed. Proceedings: Leafy Spurge Strategic Planning Workshop, March 29–30, Dickinson, ND: USDI, NPS.Google Scholar
Gaudin, V., Lunness, P. A., Fobert, P. R., Towers, M., Riou-Khamlichi, C., Murray, J.A.H., Coen, E., and Doonan, J. H. 2000. The expression of D-Cyclin genes defines distinct developmental zones in snapdragon apical meristems and is locally regulated by the cycloidea gene. Plant Physiol. 122:11371148.Google Scholar
Gendreau, E., Hofte, H., Orbovic, V., and Traas, J. 1999. Gibberellin and ethylene control endoreduplication levels in the Arabidopsis thaliana hypocotyl. Planta 209:513516.Google Scholar
Gutierrez, C. 1998. The retinoblastoma pathway in plant cell cycle and development. Curr. Opin. Plant Biol. 1:492497.Google Scholar
Horvath, D. P. 1998. The role of specific plant organs and polar auxin transport in correlative inhibition of leafy spurge (Euphorbia esula) root buds. Can. J. Bot. 76:11271231.Google Scholar
Horvath, D. P. 1999. Role of mature leaves in inhibition of root bud growth in Euphorbia esula . Weed Sci. 47:544550.Google Scholar
Horvath, D. P. and Anderson, J. V. 2000. The effect of photosynthesis on underground adventitious shoot bud dormancy/quiescence in leafy spurge (Euphorbia esula L.). Pages 3034 In Viemont, J.-D. and Crabbe, J. J., eds. 2nd International Symposium on Plant Dormancy: Short Communications, Angers, France: Presses de L’Universite d’Angers and CAB International.Google Scholar
Horvath, D. P. and Olson, P. A. 1998. Cloning and characterization of cold-regulated glycine-rich RNA-binding protein genes from leafy spurge (Euphorbia esula L.) and comparison to heterologous genomic clones. Plant Mol. Biol. 38:531538.CrossRefGoogle ScholarPubMed
Huijser, C., Kortstee, A., Pego, J., Weisbeek, P., Wisman, E., and Smeekens, S. 2000. The Arabidopsis sucrose uncoupled-6 gene is identical to absisic acid insensitive-4: involvement of absisic acid in sugar responses. Plant J. 23:577585.Google Scholar
Huntley, R. P. and Murray, A. H. 1999. The plant cell cycle. Curr. Opin. Plant Biol. 2:440446.Google Scholar
Ikeda, Y., Koizumi, N., Kusano, T., and Sano, H. 1999. Sucrose and cytokinin modulation of WPK4, a gene encoding a SNF1-related protein kinase from wheat. Plant Physiol. 121:813820.Google Scholar
Jackson, P.E.J., Carrera, E., Prat, S., and Thomas, B. 2000. Regulation of transcript levels of a potato gibberellin 20-oxidase gene by light and phytochrome B. Plant Physiol. 124:423430.Google Scholar
Johnson, D. G. and Walker, C. L. 1999. Cyclins and cell cycle checkpoints. Annu. Rev. Pharmacol. Toxicol. 39:295312.Google Scholar
Kaelin, W. G. Jr. 1999. Functions of the retinoblastoma protein. Bioessay 21:950958.Google Scholar
Katsuta, J. and Shibaoka, H. 1992. Inhibition by kinase inhibitors of the development and the disappearance of the preprophase band and microtubules in tobacco BY-2 cells. Cell Sci, J. 103:397405.Google Scholar
Kjemtrup, S., Sampson, K. S., Peele, C. G., Nguyen, L. V., Conkling, M. A., Thompson, W. F., and Robertson, D. 1998. Gene silencing from plant DNA carried by a geminivirus. Plant J. 14:91100.CrossRefGoogle ScholarPubMed
Kumagai, M. H., Donson, J., Della-Cioppa, G., Harvey, D., Hanley, K., and Grill, L. K. 1995. Cytoplasmic inhibition of carotenoid biosynthesis with virus-derived RNA. Proc. Natl. Acad. Sci. USA 92:16791683.Google Scholar
Koepp, D. M., Harper, J. W., and Elledge, S. J. 1999. How the cyclin became a cyclin: regulated proteolysis in the cell cycle. Cell 97:431434.Google Scholar
Korableva, N. P. and Ladyzhenskaya, E. P. 1995. Mechanism of hormonal regulation of potato (Solanum tuberosum L.) tuber dormancy. Biochemistry (Moscow) 60:3338.Google Scholar
Krysan, P. J., Young, J. C., and Sussman, M. R. 1999. T-DNA as an insertional mutagen in Arabidopsis. Plant Cell 11:22832290.Google Scholar
Kumagai, M. H., Donson, J., Della-Cioppa, G., Harvey, D., Hanley, K., and Grill, L. K. 1995. Cytoplasmic inhibition of carotenoid biosynthesis with virus-derived RNA. Proc. Natl. Acad. Sci. USA 92:16791683.Google Scholar
Kurata, S.-I. 2000. Selective activation of p38 MAPK cascade and mitotic arrest caused by low level oxidative stress. J. Biol. Chem. 275 (31): 2341323416.Google Scholar
Laby, R. J., Kincaid, M. S., Kim, D., and Gibson, S. I. 2000. The Arabidopsis sugar-insensitive mutants sis4 and sis5 are defective in absisic acid synthesis and response. Plant J. 23:587596.Google Scholar
Lang, G. A., Early, J. D., Martin, G. C., and Darnell, R. L. 1987. Endo-, para-, and ecodormancy: physiological terminology and classification for dormancy research. Hortic. Sci. 22:371377.Google Scholar
Leyser, H.M.O., Timpte, L.C.A., Lammer, D., Turner, J., and Estelle, M. 1993. Arabidopsis auxin-resistance gene AXR1 encodes a protein related to ubiquitin-activating enzyme E1. Nature 364:161164.Google Scholar
Liang, P. and Pardee, A. B. 1992. Differential display of eukaryotic messenger RNA by means of the polymerase chain reaction. Science 257:967971.Google Scholar
Lopez-Juez, E., Kobayashi, M., Sakurai, A., Kamiya, Y., and Kendrick, R. E. 1995. Phytochrome, gibberellins and hypocotyl growth. Plant Physiol. 107:131140.Google Scholar
Meskiene, I. and Hirt, H. 2000. MAP kinase pathways: molecular plug-and-play chips for the cell. Plant Mol. Biol. 42:791806.CrossRefGoogle ScholarPubMed
Mironov, V., De Veylder, L., Van Montagu, M., and Inze, D. 1999. Cyclin-dependent kinases and cell division in plants—the nexus. Plant Cell 11:509521.Google Scholar
Nakagami, H., Sekine, M., Murakami, H., and Shinmyo, A. 1999. Tobacco retinoblastoma-related protein phosphorylated by a distinct cyclin-dependent kinase complex with Cdc2/cyclin D in vitro. Plant J. 18:243252.CrossRefGoogle ScholarPubMed
Napoli, C. A., Beveridge, C. A., and Snowden, K. C. 1999. Reevaluating concepts of apical dominance and the control of axillary bud outgrowth. Curr. Topics Dev. Biol. 44:127169.Google Scholar
Neff, M. M., Fankhauser, C., and Chory, J. 2000. Light: an indicator of time and place. Genes Dev. 14:257271.Google Scholar
Nissen, S. J. and Foley, M. E. 1987. Correlative inhibition in root buds of leafy spurge. Weed Sci. 35:155159.Google Scholar
Nooden, L. D. and Weber, J. A. 1978. Environmental and hormonal control of dormancy in terminal buds of plants. Pages 221268 In Cutter, M. E., ed. Dormancy and Developmental Arrest. New York: Academic Press.Google Scholar
Nurse, P. and Bissett, Y. 1981. Gene required in G1 for commitment to cell cycle and in G2 for control of mitosis in fission yeast. Nature 292:558560.Google Scholar
Paterson, A. H., Schertz, K. F., Lin, Y.-R., Liu, S.-C., and Chang, Y.- L. 1995. The weediness of wild plants: molecular analysis of genes influencing dispersal and persistence of johnsongrass, Sorghum halepense (L.). Proc. Natl. Acad. Sci. USA 92:61276131.Google Scholar
Pauley, S. S. and Perry, T. O. 1954. Ecotype variation in the photoperiodic response in Populus. J. Arnold Arbor. Harv. Univ. 35:167188.Google Scholar
Perata, P., Matsukura, C., Vernieri, P., and Yamaguchi, J. 1997. Sugar repression of a gibberellin-dependent signaling pathway in barley embyos. Plant Cell 9:21972208.Google Scholar
Perry, T. O. 1971. Dormancy of trees in winter. Science 171:2936.CrossRefGoogle ScholarPubMed
Pétel, G. and Gendraud, M. 1996. Processes at the plasma membrane and plasmolemma ATPase during dormancy. Pages 233243 In Lang, G. A., ed. Plant Dormancy: Physiology, Biochemistry and Molecular Biology. Wallingford, UK: CAB International.Google Scholar
Pines, J. 1999. Four-dimensional control of the cell cycle. Nat. Cell Biol. 1:E73E79.Google Scholar
Reed, J. W., Foster, K. R., Morgan, P. W., and Chory, J. 1996. Phytochrom B affects responsiveness to gibberellin in Arabidopsis . Plant Physiol. 112:337342.Google Scholar
Reymond, P., Weber, H., Damond, M., and Farmer, E. E. 2000. Differential gene expression in response to mechanical wounding and insect feeding in Arabidopsis . Plant Cell 12:707719.Google Scholar
Roberts, C. J., Nelson, B., Maron, M. J., et al. 2000. Signaling and circuitry of multiple MAPK pathways revealed by matrix of global gene expression profiles. Science 287:873880.Google Scholar
Ruiz, M. T., Voinnet, O., and Baulcombe, D. C. 1998. Initiation and maintenance of virus-induced gene silencing. Plant Cell 10:937946.Google Scholar
Sauter, M., Mekhedov, S. L., and Kende, H. 1995. Gibberellin promotes histone H1 kinase activity and the expression of cdc2 and cyclin genes during the induction of rapid growth in deepwater rice internodes. Plant J. 7:623632.Google Scholar
Schimming, W. K. and Messersmith, C. G. 1988. Freezing resistance of over-wintering buds of four perennial weeds. Weed Sci. 36:568573.Google Scholar
Sekine, M., Ito, M., Uemukai, K., Maeda, Y., Nakagami, H., and Shinmyo, A. 1999. Isolation and characterization of the E2F-like gene in plants. FEBS Lett. 460:117122.Google Scholar
Soni, R., Carmichael, J. P., Shah, Z. H., and Murray, J.A.H. 1995. A family of cyclin D homologs from plants differentially controlled by growth regulators and containing the conserved retinoblastoma protein interaction motif. Plant Cell 7:85103.Google Scholar
Spollen, W. G., LeNoble, M. E., Samuels, T. D., Bernstein, N., and Sharp, R. E. 2000. Abscisic acid accumulation maintains maize primary root elongation at low water potentials by restricting ethylene production. Plant Physiol. 122:967976.Google Scholar
Springer, P. S. 2000. Gene traps: tools for plant development and genomics. Plant Cell 12:10071020.Google Scholar
Stafstrom, J. P. 1995. Developmental potential of shoot buds. Pages 257279 In Gartner, B. L., ed. Plant Stems: Physiology and Functional Morphology. San Diego: Academic Press.Google Scholar
Suttle, J. C. 1998. Involvement of ethylene in potato microtuber dormancy. Plant Physiol. 118:843848.Google Scholar
Suttle, J. C. 2000. The role of endogenous hormones in potato tuber dormancy. Pages 211226 In Viemont, J.-D. and Crabbe, J., eds. Dormancy in Plants. New York: CAB Publishing.Google Scholar
Suttle, J. C. and Hultstrand, J. F. 1994. Role of endogenous abscisic acid in potato microtuber dormancy. Plant Physiol. 105:891896.Google Scholar
Toyomasu, T., Kawaide, H., Mitsubashi, W., Inoue, Y., and Kamiya, Y. 1998. Phytochrome regulates gibberellin biosynthesis during germination of photoblastic lettuce seeds. Plant Physiol. 118:15171523.Google Scholar
Umeda, M., Bhalerao, R. P., Schell, J., Uchimiya, H., and Koncz, C. 1998. A distinct cyclin-dependent kinase-activating kinase of Arabidopsis thaliana . Proc. Natl. Acad. Sci. USA 95:50215026.Google Scholar
van Hal, N.L.W., Vorst, O., van Houwelingen, A.M.M.L., Kok, E. J., Peijnenburg, A., Aharoni, A., van Tunen, A. J., and Keijer, J. 2000. The application of DNA microarrays in gene expression analysis. J. Biotechnol. 78:271280.Google Scholar
Velculescu, V. E., Zhang, L., Vogelstein, B., and Kinzler, K. W. 1995. Serial analysis of gene expression. Science 270:484487.Google Scholar
Wang, H., Qi, Q., Schorr, P., Cutler, A. J., Crosby, W. L., and Fowke, L. C. 1998. ICK1, a cyclin-dependent protein kinase inhibitor from Arabidopsis thaliana interacts with both Cdc2a and CycD3, and its expression is induced by abscisic acid. Plant J. 15 (4): 501510.Google Scholar
Weller, J. L., Ross, J. J., and Reid, J. B. 1994. Gibberellins and phytochrome regulation of stem elongation in pea. Planta 192:489496.Google Scholar
Wen, X., Fuhrman, S., Michaels, G. S., Carr, D. B., Smith, S., Barker, J. L., and Somogyi, R. 1998. Large-scale temporal gene expression mapping of central nervous system development. Proc. Natl. Acad. Sci. USA 95:334339.Google Scholar
Whitmarsh, A. J. and Davis, R. J. 2000. Regulation of transcription factor function by phosphorylation. Cell. Mol. Life Sci. 57:11721183.Google Scholar
Wingler, A., von Schaewen, A., Leegood, R. C., Lea, P. J., and Paul Quick, W. 1998. Regulation of leaf senescence by cytokinin, sugars, and light: effects on NADH-dependent hydroxypyruvate reductase. Plant Physiol. 116:329335.Google Scholar
Xin, X., van Lammeren, A.A.M., Vermeer, E., and Vreugdenhil, D. 1998. The role of gibberellin, abscisic acid, and sucrose in the regulation of potato tuber formation in vitro. Plant Physiol. 117:575584.Google Scholar
Yamaguchi, M., Umeda, M., and Uchimiya, H. 1998a. A rice homolog of CDK7/MO15 phosphorylates both cyclin-dependent protein kinases and the carboxy-terminal domain of RNA polymerase II. Plant J. 15:613619.Google Scholar
Yamaguchi, S., Smith, M. W., Brown, R.G.S., Kamiya, Y., and Sun, T.- P. 1998b. Phytochrome regulation and differential expression of gibberellin 3β-hydroxylase genes in germinating Arabidopsis seeds. Plant Cell 10:21152126.Google Scholar
Zieslin, N. and Halevy, A. H. 1976. Components of axillary bud inhibition in rose plants. The effects of different plant parts (correlative inhibition). Bot. Gaz. 137:291296.Google Scholar