Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-26T13:42:04.127Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  09 February 2018

Alessandro Minelli
Affiliation:
Università degli Studi di Padova, Italy
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Plant Evolutionary Developmental Biology
The Evolvability of the Phenotype
, pp. 340 - 435
Publisher: Cambridge University Press
Print publication year: 2018

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

Aagaard, J. E., Olmstead, R. G., Willis, J. H. & Phillips, P. C. (2005). Duplication of floral regulatory genes in Lamiales. American Journal of Botany, 92: 12841293.Google Scholar
Aida, M., Ishida, T., Fukaki, H. et al. (1997). Genes involved in organ separation in Arabidopsis: an analysis of the cup-shaped cotyledon mutant. Plant Cell, 9: 841857.Google Scholar
Aida, M., Ishida, T. & Tasaka, M. (1999). Shoot apical meristem and cotyledon formation during Arabidopsis embryogenesis: interaction among the CUP-SHAPED COTYLEDON and SHOOT MERISTEMLESS genes. Development, 126: 15631570.CrossRefGoogle ScholarPubMed
Aida, M. & Tasaka, M. (2006). Morphogenesis and patterning at the organ boundaries in the higher plant shoot apex. Plant Molecular Biology, 60: 915928.Google Scholar
Ainsworth, C., Crossley, S., Buchanan-Wollaston, V., Thangavelu, M. & Parker, J. (1995). Male and female flowers of the dioecious plant sorrel show different patterns of MADS box gene expression. Plant Cell, 7: 15831598.Google Scholar
Airoldi, C. A., Bergonzi, S. & Davies, B. (2010). Single amino acid change alters the ability to specify male or female organ identity. Proceedings of the National Academy of Sciences of the United States of America, 107, 1889818902.Google Scholar
Airoldi, C. A. & Davies, B. (2012). Gene duplication and the evolution of plant MADS-box transcription factors. Journal of Genetics and Genomics, 39: 157165.Google Scholar
Al-Shehbaz, I. A., Beilstein, M. A. & Kellogg, E. A. (2006). Systematics and phylogeny of the Brassicaceae (Cruciferae): an overview. Plant Systematics and Evolution, 259: 89120.Google Scholar
Alados, C. L., Escos, J., Emlen, J. M. & Freeman, D. C. (1999). Characterization of branch complexity by fractal analyses. International Journal of Plant Sciences, 160: S147S155.Google Scholar
Alapetite, E., Baker, W. J. & Nadot, S. (2014). Evolution of stamen number in Ptychospermatinae (Arecaceae): insights from a new molecular phylogeny of the subtribe. Molecular Phylogenetics and Evolution, 76: 227240.Google Scholar
Alberch, P. (1991). From genes to phenotype: dynamical systems and evolvability. Genetica, 84: 511.Google Scholar
Alberch, P., Gould, S. J., Oster, G. F. & Wake, D. B. (1979). Size and shape in ontogeny and phylogeny. Paleobiology, 5: 296317.CrossRefGoogle Scholar
Albert, V. A., Gustafsson, M. H. G. & Di Laurenzio, L. (1998). Ontogenetic systematics, molecular developmental genetics, and the angiosperm petal. In Molecular Systematics of Plants II: DNA Sequencing, eds. Soltis, D. E., Soltis, P. S. & Doyle, J. A.. Boston, MA: Kluwer, pp. 349374.Google Scholar
Albert, V. A., Oppenheimer, D. & Lindqvist, C. (2002). Pleiotropy, redundancy and the evolution of flowers. Trends in Plant Science, 7: 297301.Google Scholar
Almeida, A. M. R., Yockteng, R., Otoni, W. C. & Specht, C. D. (2015a). Positive selection on the K domain of the AGAMOUS protein in the Zingiberales suggests a mechanism for the evolution of androecial morphology. EvoDevo, 6: 7.Google Scholar
Almeida, A. M. R., Yockteng, R. & Specht, C. D. (2015b). Evolution of petaloidy in the Zingiberales: an assessment of the relationship between ultrastructure and gene expression patterns. Developmental Dynamics, 244: 11211132.Google Scholar
Alonso-Cantabrana, H., Ripoll, J. J., Ochando, I. et al. (2007). Common regulatory networks in leaf and fruit patterning revealed by mutations in the Arabidopsis ASYMMETRIC LEAVES1 gene. Development, 134: 26632671.Google Scholar
Alvarez, J. & Smyth, D.R. (1999). CRABS CLAW and SPATULA, two Arabidopsis genes that control carpel development in parallel with AGAMOUS. Development, 126: 23772386.Google Scholar
Alvarez-Buylla, E. R., Ambrose, B. A., Flores-Sandoval, E. et al. (2010). B-function expression in the flower center underlies the homeotic phenotype of Lacandonia schismatica (Triuridaceae). Plant Cell, 22: 35433559.Google Scholar
Amasino, R. (2010). Seasonal and developmental timing of flowering. The Plant Journal, 61: 10011013.Google Scholar
Ambros, V. & Moss, E. G. (1994). Heterochronic genes and the temporal control of C. elegans development. Trends in Genetics, 10: 123127.Google Scholar
Ambrose, B. A., Espinosa-Matías, S., Vázquez-Santana, S. et al. (2006). Comparative floral developmental series of the Mexican triurids support a euanthial interpretation for the unusual floral structures of Lacandonia schismatica (Lacandoniaceae). American Journal of Botany, 93: 1535.Google Scholar
Ambrose, B. A. & Ferrándiz, C. (2013). Development and the evolution of plant form. Annual Plant Reviews, 45: 277320.Google Scholar
Ambrose, B. A., Lerner, D. R., Ciceri, P. et al. (2000). Molecular and genetic analyses of the silky1 gene reveal conservation in floral organ specification between eudicots and monocots. Molecular Cell, 5: 569579.Google Scholar
Andreasen, K. & Baldwin, B. G. (2001). Unequal evolutionary rates between annual and perennial lineages of checker mallows (Sidalcea, Malvaceae): evidence from 18S–26S rDNA internal and external transcribed spacers. Molecular Biology and Evolution, 18: 936944.Google Scholar
Andrés, F. & Coupland, G. (2012). The genetic basis of flowering responses to seasonal cues. Nature Reviews Genetics, 13: 627639.Google Scholar
Andriankaja, M., Dhondt, S., De Bodt, S. et al. (2012). Exit from proliferation during leaf development in Arabidopsis thaliana: a not-so-gradual process. Developmental Cell, 22: 6478.Google Scholar
Angenent, G. C. & Colombo, L. (1996). Molecular control of ovule development. Trends in Plant Science, 1: 228232.Google Scholar
Antonelli, A. (2009). Have giant lobelias evolved several times independently? Life form shifts and historical biogeography of the cosmopolitan and highly diverse subfamily Lobelioideae (Campanulaceae). BMC Biology, 7: 82.Google Scholar
Antonius, K. & Ahokas, H. (1996). Flow cytometric determination of polyploidy level in spontaneous clones of strawberries. Hereditas, 124: 285.Google Scholar
APG IV (2016). An update of the Angiosperm Phylogeny Group classification for the orders and families of flowering plants: APG IV. Botanical Journal of the Linnean Society, 181: 120.Google Scholar
Appel, O. & Al-Shehbaz, I. A. (2003). Cruciferae. In The Families and Genera of Vascular Plants, Vol. 5, eds. Kubitzki, K. & Bayer, C.. Berlin: Springer, pp. 75174.Google Scholar
Arabidopsis Genome Initiative (2000). Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature, 408: 796815.Google Scholar
Arber, A. (1918). The phyllode theory of the monocotyledonous leaf, with special reference to anatomical evidence. Annals of Botany, 32: 465501.Google Scholar
Arber, A. (1920). Water Plants: A Study of Aquatic Angiosperms. Cambridge: Cambridge University Press.Google Scholar
Arber, A. (1921). The leaf structure of the Iridaceae, considered in relation to the phyllode theory. Annals of Botany, 35: 301336.Google Scholar
Arber, A. (1928). Studies in the Gramineae. V. 1). On Luziola and Dactylis. 2). On Lygeum and Nardus. Annals of Botany, 42: 391407.Google Scholar
Arber, A. (1941). The interpretation of leaf and root in the angiosperms. Biological Review, 16: 81105.Google Scholar
Arber, A. (1950). The Natural Philosophy of Plant Form. Cambridge: Cambridge University Press.Google Scholar
Armbruster, W. S., Debevec, E. M. & Willson, M. F. (2002). Evolution of syncarpy in angiosperms: theoretical and phylogenetic analyses of the effects of carpel fusion on offspring quantity and quality. Journal of Evolutionary Biology, 15: 657672.Google Scholar
Armstrong, J. & Douglas, A. W. (1989). The ontogenetic basis for corolla aestivation in Scrophulariaceae. Bulletin of the Torrey Botanical Club, 116: 378389.Google Scholar
Arnaud, N. & Laufs, P. (2013). Plant development: brassinosteroids go out of bounds. Current Biology, 23, 152154.Google Scholar
Ashton, P. S. (2003). Dipterocarpaceae. In The Families and Genera of Vascular Plants, Vol. 5, eds. Kubitzki, K. & Bayer, C.. Berlin: Springer, pp. 182197.Google Scholar
Ashton, P. S., Givnish, T. J. & Appanah, S. (1988). Staggered flowering in the Dipterocarpaceae: new insights into floral induction and the evolution of mast fruiting in the aseasonal tropics. American Naturalist, 132: 4466.Google Scholar
Autran, D., Jonak, C., Belcram, K. et al. (2002). Cell numbers and leaf development in Arabidopsis: a functional analysis of the STRUWWELPETER gene. EMBO Journal, 21: 60366049.Google Scholar
Bachmann, K. & Gailing, O. (2003). The genetic dissection of the stepwise evolution of morphological characters. In Deep Morphology: Toward a Renaissance of Morphology in Plant Systematics, eds. Stuessy, T. F., Mayer, V. & Hörandl, E.. Königstein: Koeltz, pp. 3562.Google Scholar
Bahadur, B., Reddy, N. P., Rao, M. M. & Farooqui, S. M. (1984). Corolla handedness in Oxalidaceae, Linaceae and Plumbaginaceae. Journal of the Indian Botanical Society, 63: 408411.Google Scholar
Bailey, C. D., Koch, M. A., Mayer, M. et al. (2006). Toward a global phylogeny of the Brassicaceae. Molecular Biology and Evolution, 23: 21422160.Google Scholar
Bainbridge, K., Guyomarc’h, S., Bayer, E. et al. (2008). Auxin influx carriers stabilize phyllotactic patterning. Genes and Development, 22: 810823.Google Scholar
Baker, C. C., Sieber, P., Wellmer, F. & Meyerowitz, E. M. (2005). The early extra petals1 mutant uncovers a role for microRNA miR164c in regulating petal number in Arabidopsis. Current Biology, 15: 303315.Google Scholar
Balazadeh, S., Parlitz, S., Mueller-Roeber, B. & Meyer, R. C. (2008). Natural developmental variation in leaf and plant senescence in Arabidopsis thaliana. Plant Biology, 10: 136147.Google Scholar
Baldwin, J. M. (1896). A new factor in evolution. American Naturalist, 30: 441451, 536–553.Google Scholar
Banks, J. A., Nishiyama, T., Hasebe, M. et al. (2011). The Selaginella genome identifies genetic changes associated with the evolution of vascular plants. Science, 332: 960963.Google Scholar
Barclay, I. R. (1975). High frequencies of haploid production in wheat (Triticum aestivum) by chromosome elimination. Nature, 256: 410411.Google Scholar
Barker, M. S., Kane, N. C., Matvienko, M. et al. (2008). Multiple paleopolyploidizations during the evolution of the Compositae reveal parallel patterns of duplicate gene retention after millions of years. Molecular Biology and Evolution, 25: 24452455.Google Scholar
Barkoulas, M., Galinha, C., Grigg, S. P. & Tsiantis, M. (2007). From genes to shape: regulatory interactions in leaf development. Current Opinion in Plant Biology, 10: 660666.Google Scholar
Barkoulas, M., Hay, A., Kougioumoutzi, E. et al. (2008). A developmental framework for dissected leaf formation in the Arabidopsis relative Cardamine hirsuta. Nature Genetics, 40: 11361141.Google Scholar
Barow, M. & Meister, A. (2003). Endopolyploidy in seed plants is differently correlated to systematics, organ, life strategy and genome size. Plant Cell and Environment, 26: 571584.Google Scholar
Barrett, R. D. H. & Schluter, D. (2008). Adaptation from standing genetic variation. Trends in Ecology and Evolution, 23: 3844.Google Scholar
Barrett, S. C. H. (ed.) (1992). Evolution and Function of Heterostyly. Berlin: Springer.Google Scholar
Barrett, S. C. H. (2002). The evolution of plant sexual diversity. Nature Reviews Genetics, 3: 274283.Google Scholar
Barrett, S. C. H. (2010). Darwin’s legacy: The forms, function and sexual diversity of flowers. Philosophical Transactions of the Royal Society B, 365: 351368.Google Scholar
Barrett, S. C. H., Jesson, L. K. & Baker, A. M. (2000). The evolution of stylar polymorphisms in plants. Annals of Botany, 85 (Suppl. A): 253265.Google Scholar
Barth, S., Geier, T., Eimert, K. et al. (2009). KNOX overexpression in transgenic Kohleria (Gesneriaceae) prolongs the activity of proximal leaf blastozones and drastically alters segment fate. Planta, 230: 10811091.Google Scholar
Barthélémy, D. & Caraglio, Y. (2007). Plant architecture: a dynamic, multilevel and comprehensive approach to plant form, structure and ontogeny. Annals of Botany, 99: 375407.Google Scholar
Bartholmes, C., Hidalgo, O., Gleissberg, S. (2012). Evolution of the YABBY gene family with emphasis on the basal eudicot Eschscholzia californica (Papaveraceae). Plant Biology, 14: 1123.Google Scholar
Bartlett, M. E. & Specht, C. D. (2010). Evidence for the involvement of GLOBOSA-like gene duplications and expression divergence in the evolution of floral morphology in the Zingiberales. New Phytologist, 187: 521541.Google Scholar
Bartlett, M. E. & Specht, C. D. (2011). Changes in expression pattern of the TEOSINTE BRANCHED1-like genes in the Zingiberales provide a mechanism for evolutionary shifts in symmetry across the order. American Journal of Botany, 98: 227243.Google Scholar
Barton, M. K. (2010). Twenty years on: the inner workings of the shoot apical meristem, a developmental dynamo. Developmental Biology, 341: 95113.Google Scholar
Bateman, R. M. & DiMichele, W. A. (1994). Saltational evolution of form in vascular plants: a neoGoldschmidtian synthesis. In Shape and Form in Plants and Fungi, eds. Ingram, D. S. & Hudson, A.. London: Academic Press, pp. 63102.Google Scholar
Bateman, R. M. & DiMichele, W. A. (2002). Generating and-filtering-major phenotypic novelties: NeoGoldschmidtian saltation revisited. In Developmental Genetics and Plant Evolution, eds. Cronk, Q. C. B., Bateman, R. M. & Hawkins, J. A.. London: Taylor & Francis, pp. 109159.Google Scholar
Bateman, R. M., Hilton, J. & Rudall, P. J. (2006). Morphological and molecular phylogenetic context of the angiosperms: contrasting the ‘top-down’ and ‘bottom-up’ approaches to inferring the likely characteristics of the first flowers. Journal of Experimental Botany, 57: 34713503.Google Scholar
Bateman, R. M., Hilton, J. & Rudall, P. J. (2011). Spatial separation and developmental divergence of male and female reproductive units in gymnosperms, and their relevance to the origin of the angiosperm flower. In Flowers on the Tree of Life, eds. Wanntorp, L. & De Craene, L. P. Ronse. Cambridge: Cambridge University Press, pp. 848.Google Scholar
Bateson, W. & Bateson, A. (1891). On variations in the floral symmetry of certain plants having irregular corollas. Journal of the Linnean Society, Botany, 28: 386424.Google Scholar
Baum, D. A. (1998). The evolution of plant development. Current Opinion in Plant Biology, 1: 7986.Google Scholar
Baum, D. A. & Donoghue, M. J. (2002). Transference of function, heterotopy and the evolution of plant development. In Developmental Genetics and Plant Evolution, eds. Cronk, Q. C. B., Bateman, R. M. & Hawkins, J. A.. London: Taylor & Francis, pp. 5269.Google Scholar
Baum, D. A. & Hileman, L. C. (2006). A developmental genetic model for the origin of the flower. In Flowering and its Manipulation, ed. Ainsworth, C.. Sheffield: Blackwell, pp. 327.Google Scholar
Beaulieu, J. M. & Donoghue, M. J. (2013). Fruit evolution and diversification in campanulid angiosperms. Evolution, 67: 31323144.Google Scholar
Beilstein, M. A., Al-Shehbaz, I. A. & Kellogg, E. A. (2006). Brassicaceae phylogeny and trichome evolution. American Journal of Botany, 93: 607619.Google Scholar
Beilstein, M. A., Nagalingum, N. S., Clements, M. D. et al. (2010). Dated molecular phylogenies indicate a Miocene origin for Arabidopsis thaliana. Proceedings of the National Academy of Sciences of the United States of America, 107: 1872418728.Google Scholar
Bell, A. (2008). Plant Form: An Illustrated Guide to Flowering Plant Morphology, new edition. Portland, OR; London: Timber Press.Google Scholar
Bell, E. M., Lin, W., Husbands, A. Y. et al. (2012). Arabidopsis LATERAL ORGAN BOUNDARIES negatively regulates brassinosteroid accumulation to limit growth in organ boundaries. Proceedings of the National Academy of Sciences of the United States of America, 109: 2114621151.Google Scholar
Bellini, C., Pacurar, D. I. & Perrone, I. (2014). Adventitious roots and lateral roots: similarities and differences? Annual Review of Plant Biology, 65: 639666.Google Scholar
Bello, M. A., Álvarez, I., Torices, R. & Fuertes-Aguilar, J. (2013). Floral development and evolution of capitulum structure in Anacyclus (Anthemideae, Asteraceae). Annals of Botany, 112: 15971612.Google Scholar
Bello, M. A., Bruneau, A., Forest, F. & Hawkins, J. A. (2009). Elusive relationships within order Fabales: phylogenetic analyses using matK and rbcL sequence data. Systematic Botany, 34: 102114.Google Scholar
Bello, M. A., Rudall, P. J. & Hawkins, J. A. (2012). Combined phylogenetic analyses reveal interfamilial relationships and patterns of floral evolution in the eudicot order Fabales. Cladistics, 28: 393421.Google Scholar
Benková, E., Michniewicz, M., Sauer, E. et al. (2003). Local, efflux-dependent auxin gradients as a common module for plant organ formation. Cell, 115: 591602.Google Scholar
Benlloch, R., Berbel, A., Serrano-Mislata, A. & Madueno, F. (2007). Floral initiation and inflorescence architecture: a comparative view. Annals of Botany, 100: 659676.Google Scholar
Bennett, M. D. & Leitch, I. J. (2012). Plant DNA C-values Database (release 6.0, December 2012). http://data.kew.org/cvalues (accessed September 2017).Google Scholar
Berbel, A., Navarro, C. & Ferrándiz, C. et al. (2001). Analysis of PEAM4, the pea AP1 functional homologue, supports a model for AP1-like genes controlling both floral meristem and floral organ identity in different plant species. The Plant Journal, 25: 441451.Google Scholar
Berger, B. A., Thompson, V., Lim, A., Ricigliano, V. & Howarth, D.G. (2016). Elaboration of bilateral symmetry across Knautia macedonica capitula related to changes in ventral petal expression of CYCLOIDEA-like genes. EvoDevo, 7: 8.Google Scholar
Berger, Y., Harpaz-Saad, S., Brand, A. et al. (2009). The NAC-domain transcription factor GOBLET specifies leaflet boundaries in compound tomato leaves. Development, 136: 823832.Google Scholar
Bergthorsson, U., Adams, K. L., Thomason, B. & Palmer, J. D. (2003). Widespread horizontal transfer of mitochondrial genes in flowering plants. Nature, 424: 197201.Google Scholar
Bergthorsson, U., Richardson, A. O., Young, G. J., Goertzen, L. R. & Palmer, J. D. (2004). Massive horizontal transfer of mitochondrial genes from diverse land plant donors to the basal angiosperm Amborella. Proceedings of the National Academy of Sciences of the United States of America, 101: 1774717752.Google Scholar
Bertrand-Garcia, R. & Freeling, M. (1991). Hairy-sheath-frayed1-0: a systemic, heterochronic mutant of maize that specifies slow developmental stage transitions. American Journal of Botany, 78: 747765.Google Scholar
Bharathan, G., Goliber, T. E., Moore, C. et al. (2002). Homologies in leaf form inferred from KNOXI gene expression during development. Science, 296: 18581860.Google Scholar
Bissell, E. K. & Diggle, P. K. (2008). Floral morphology in Nicotiana: architectural and temporal effects on phenotypic integration. International Journal of Plant Sciences, 169: 225240.Google Scholar
Blaser, J. L. (1954). The morphology of the flower and inflorescence of Mitchella repens. American Journal of Botany, 41: 533539.Google Scholar
Blázquez, M. A., Ferrándiz, C., Madueno, F. & Parcy, F. (2006). How floral meristems are built. Plant Molecular Biology, 60: 855870.Google Scholar
Blázquez, M. A., Soowal, L. N., Lee, I. & Weigel, D. (1997). LEAFY expression and flower initiation in Arabidopsis. Development, 124: 38353844.Google Scholar
Blázquez, M. A. & Weigel, D. (2000). Integration of floral inductive signals in Arabidopsis. Nature, 404: 889892.Google Scholar
Blein, T., Hasson, A. & Laufs, P. (2010). Leaf development: what it needs to be complex. Current Opinion in Plant Biology, 13: 7582.Google Scholar
Blein, T., Pulido, A., Vialette-Guiraud, A. et al. (2008). A conserved molecular framework for compound leaf development. Science, 322: 18351839.Google Scholar
Bliss, B. J., Wanke, S., Barakat, A. et al. (2013). Characterization of the basal angiosperm Aristolochia fimbriata: a potential experimental system for genetic studies. BMC Plant Biology, 13: 13.Google Scholar
Bohs, L., Weese, T., Myers, N. et al. (2007). Zygomorphy and heteranthery in Solanum in a phylogenetic context. Acta Horticulturae, 745: 201224.Google Scholar
Borsch, T., Löhne, C. & Wiersema, J. (2008). Phylogeny and evolutionary patterns in Nymphaeales: integrating genes, genomes and morphology. Taxon, 57: 10521081.Google Scholar
Boss, P. K. & Thomas, M. R. (2002). Association of dwarfism and floral induction with a grape ‘green revolution’ mutation. Nature, 416: 847850.Google Scholar
Bouche, F., Lobet, G., Tocquin, P. & Perilleux, C. (2016). FLOR-ID: an interactive database of flowering-time gene networks in Arabidopsis thaliana. Nucleic Acids Research, 44: D1167D1171.Google Scholar
Boudaoud, A. (2010). An introduction to the mechanics of morphogenesis for plant biologists. Trends in Plant Science, 15: 353360.Google Scholar
Bowers, J. E., Chapman, B. A., Rong, J. & Paterson, A. H. (2003). Unravelling angiosperm genome evolution by phylogenetic analysis of chromosomal duplication events. Nature, 422: 433438.Google Scholar
Bowman, J. L. (1997). Evolutionary conservation of angiosperm flower development at the molecular and genetic levels. Journal of Biosciences, 22: 515527.Google Scholar
Bowman, J. L., Alvarez, J., Weigel, D., Meyerowitz, E. M. & Smyth, D. R. (1993). Control of flower development in Arabidopsis thaliana by APETALA1 and interacting genes. Development, 119: 721743.Google Scholar
Bowman, J. L., Bruggemann, H., Lee, J.-Y. & Mummenhoff, K. (1999). Evolutionary changes in floral structure within Lepidium L. (Brassicaceae). International Journal of Plant Sciences, 160: 917929.Google Scholar
Bowman, J. L., Smyth, D. R. & Meyerowitz, E. M. (1989). Genes directing flower development in Arabidopsis. Plant Cell, 1: 3752.Google Scholar
Bown, D. (2000). Aroids: Plants of the Arum Family, 2nd edn. Portland, OR: Timber Press.Google Scholar
Box, M. S., Bateman, R. M., Glover, B. J. & Rudall, P. J. (2008). Floral ontogenetic evidence of repeated speciation via paedomorphosis in subtribe Orchidinae (Orchidaceae). Botanical Journal of the Linnean Society, 157: 429454.Google Scholar
Box, M. S. & Glover, B. J. (2010). A plant developmentalist’s guide to paedomorphosis: reintroducing a classic concept to a new generation. Trends in Plant Science, 15: 241246.Google Scholar
Bradley, D., Carpenter, R., Copsey, L. et al. (1996). Control of inflorescence architecture in Antirrhinum. Nature, 379: 791797.Google Scholar
Bradley, D., Carpenter, R., Sommer, H., Hartley, N. & Coen, E. (1993). Complementary floral homeotic phenotypes result from opposite orientations of a transposon at the plena locus of Antirrhinum. Cell, 72: 8595.Google Scholar
Bradley, D., Ratcliffe, O., Vincent, C., Carpenter, R. & Coen, E. (1997). Inflorescence commitment and architecture in Arabidopsis. Science, 275: 8083.Google Scholar
Braybrook, S. A. & Kuhlemeier, C. (2010). How a plant builds leaves. Plant Cell, 22: 10061018.Google Scholar
Breeze, E., Harrison, E., Page, T. et al. (2008). Transcriptional regulation of plant senescence: from functional genomics to systems biology. Plant Biology, 10 Suppl 1: 99109.Google Scholar
Breuil-Broyer, S., Trehin, C., Morel, P. et al. (2016). Analysis of the Arabidopsis superman allelic series and the interactions with other genes demonstrate developmental robustness and joint specification of male–female boundary, flower meristem termination and carpel compartmentalization. Annals of Botany, 117: 905923.Google Scholar
Brigandt, I. & Love, A. C. (2010). Evolutionary novelty and the evo-devo synthesis: field notes. Evolutionary Biology, 37: 9399.Google Scholar
Brigandt, I. & Love, A. C. (2012). Conceptualizing evolutionary novelty: moving beyond definitional debates. Journal of Experimental Zoology Part B: Molecular and Developmental Evolution, 318: 417427.Google Scholar
Broadley, M. R., White, P. J., Hammond, J. P. et al. (2008). Evidence of neutral transcriptome evolution in plants. New Phytologist, 180: 587593.Google Scholar
Brockington, S. F., Roolse, A., Randall, J. et al. (2009). Phylogeny of the Caryophyllales sensu lato: revisiting hypotheses on pollination biology and perianth differentiation in the core Caryophyllales. International Journal of Plant Sciences, 170: 627643.Google Scholar
Brockington, S. F., Rudall, P. J., Frohlich, M. W. et al. (2011). ‘Living stones’ reveal alternative petal identity programs within the core eudicots. The Plant Journal, 69: 193203.Google Scholar
Brody, A. & Morita, S. I. (2000). A positive association between oviposition and fruit set: female choice or manipulation? Oecologia, 124: 418425.Google Scholar
Broholm, S. K., Tähtiharju, S., Laitinen, R. A. et al. (2008). A TCP domain transcription factor controls flower type specification along the radial axis of the Gerbera (Asteraceae) inflorescence. Proceedings of the National Academy of Sciences of the United States of America, 105: 91179122.Google Scholar
Broholm, S. K., Teeri, T. H. & Elomaa, P. (2014). Molecular control of inflorescence development in Asteraceae. Advances in Botanical Research, 27: 297334.Google Scholar
Brookfield, J. F. Y. (2009). Evolution and evolvability: Celebrating Darwin 200. Biology Letters, 5: 4446.Google Scholar
Brunetti, R., Gissi, C., Pennati, R. et al. (2015). Morphological evidence that the molecularly determined Ciona intestinalis type A and type B are different species: Ciona robusta and Ciona intestinalis. Journal of Zoological Systematics and Evolutionary Research, 53: 186193.Google Scholar
Buchanan-Wollaston, V. (2007). Senescence in plants. In Encyclopedia of Life Sciences. Chichester: Wiley.Google Scholar
Buchholz, J. T. (1946). Volumetric studies of seeds, endosperm, and embryos in Pinus ponderosa during embryonic differentiation. Botanical Gazette, 108: 232244.Google Scholar
Budd, G. E. (1999). Does evolution in body patterning genes drive morphological change or vice versa? BioEssays, 21: 326332.Google Scholar
Burgeff, C., Liljegren, S. J., Tapia-Lopez, R., Yanosky, M. F. & Alvarez-Buylla, E. R. (2002). MADS-box gene expression in lateral primordia, meristems and differentiated tissues of Arabidopsis thaliana roots. Planta, 214: 365372.Google Scholar
Bürglin, T. R. (2005). Homeodomain proteins. In Encyclopedia of Molecular Cell Biology and Molecular Medicine, ed. Meyers, R. A.. Weinheim: Wiley-VCH, pp. 179222.Google Scholar
Busch, A., Horn, S., Mühlhausen, A., Mummenhoff, K. & Zachgo, S. (2012). Corolla monosymmetry: evolution of a morphological novelty in the Brassicaceae family. Molecular Biology and Evolution, 29: 12411254.Google Scholar
Busch, A. & Zachgo, S. (2007). Control of corolla monosymmetry in the Brassicaceae Iberis amara. Proceedings of the National Academy of Sciences of the United States of America, 104: 1671416719.Google Scholar
Busch, A. & Zachgo, S. (2009). Flower symmetry evolution: towards understanding the abominable mystery of angiosperm radiation. BioEssays, 31: 11811190.Google Scholar
Buzgo, M. & Endress, P. K. (2000). Floral structure and development of Acoraceae and its systematic relationships with basal angiosperms. International Journal of Plant Sciences, 161: 2341.Google Scholar
Buzgo, M., Soltis, P. S. & Soltis, D. S. (2004). Floral developmental morphology of Amborella trichopoda (Amborellaceae). International Journal of Plant Sciences, 165: 925947.Google Scholar
Byng, J. W. (2014). The Flowering Plants Handbook: A Practical Guide to Families and Genera of the World. Hertford: Plant Gateway.Google Scholar
Byrne, M. (2012). Making leaves. Current Opinion in Plant Biology, 15: 2430.Google Scholar
Byrne, M. E. (2006). Shoot meristem function and leaf polarity: the role of class III HD-ZIP genes. PLoS Genetics 2 (6): e89.Google Scholar
Byrne, M. E., Barley, R., Curtis, M. et al. (2000). ASYMMETRIC LEAVES1 mediates leaf patterning and stem cell function in Arabidopsis. Nature, 408: 967971.Google Scholar
Caddick, L. R., Rudall, P. J. & Wilkin, P. (2000). Floral morphology and development in Dioscoreales. Feddes Repertorium, 111: 189230.Google Scholar
Cai, H., Liu, X., Vanneste, K. et al. (2014). The genome sequence of the orchid Phalaenopsis equestris. Nature Genetics, 47: 6572.Google Scholar
Callos, J. D. & Medford, J. I. (1994). Organ positions and pattern formation in the shoot apex. The Plant Journal, 6: 17.Google Scholar
Calonje, M., Cubas, P., Martinez-Zapater, J. M. & Carmona, M. J. (2004). Floral meristem identity genes are expressed during tendril development in grapevine. Plant Physiology, 135: 14911501.Google Scholar
Cameron, R. J. (1970). Light intensity and growth of Eucalyptus seedlings. I. Ontogenetic variation in E. fastigiata. Australian Journal of Botany, 18: 2943.Google Scholar
Canales, C., Barkoulas, M., Galinha, C. & Tsiantis, M. (2010). Weeds of change: Cardamine hirsuta as a new model system for studying dissected leaf development. Journal of Plant Research, 123, 2533.Google Scholar
Cantino, P. D., Doyle, J. A., Graham, S. W. et al. (2007). Towards a phylogenetic nomenclature of Tracheophyta. Taxon, 56: 822846.Google Scholar
Carles, C. C., Choffnes-Inada, D., Reville, K., Lertpiriyapong, K. & Fletcher, J. C. (2005). ULTRAPETALA1 encodes a SAND domain putative transcriptional regulator that controls shoot and floral meristem activity in Arabidopsis. Development, 132: 897911.Google Scholar
Carles, C. C. & Fletcher, J. C. (2003). Shoot apical meristem maintenance: the art of a dynamic balance. Trends in Plant Science, 8: 394401.Google Scholar
Carlquist, S. (1969). Toward acceptable evolutionary interpretations of floral anatomy. Phytomorphology, 19: 332362.Google Scholar
Carlquist, S. (1980). Hawaii, a Natural History, 2nd edn. Lawai, Kauai, HI: Pacific Tropical Botanical Garden.Google Scholar
Carlsbecker, A., Tandre, K., Johanson, U., Englund, M. & Engstrom, 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: 546557.Google Scholar
Carlson, J. E., Leebens-Mack, J. H., Wall, P. K. et al. (2006). EST database for early flower development in California poppy (Eschscholzia californica Cham., Papaveraceae) tags over 6,000 genes from a basal eudicot. Plant Molecular Biology, 62: 351369.Google Scholar
Carlson, S. E., Howarth, D. G. & Donoghue, M. J. (2011). Diversification of CYCLOIDEA-like genes in Dipsacaceae (Dipsacales): implications for the evolution of capitulum inflorescences. BMC Evolutionary Biology, 11: 325.Google Scholar
Carpenter, R. & Coen, E. S. (1990). Floral homeotic mutations produced by transposon-mutagenesis in Antirrhinum majus. Genes and Development, 4: 14831493.Google Scholar
Carraro, N., Peaucelle, A., Laufs, P. & Traas, J. (2006). Cell differentiation and organ initiation at the shoot apical meristem. Plant Molecular Biology, 60: 811826.Google Scholar
Caruso, C., Rigato, E. & Minelli, A. (2012). Finalism and adaptationism in contemporary biological literature. Atti dell’Istituto Veneto di Scienze Lettere ed Arti, Classe di Scienze Fisiche, Matematiche e Naturali, 170: 6976.Google Scholar
Castel, R., Kusters, E. & Koes, R. (2010). Inflorescence development in petunia: through the maze of botanical terminology. Journal of Experimental Botany, 61: 22352246.Google Scholar
Causier, B., Castillo, R., Xue, Y., Schwarz-Sommer, Z. & Davies, B. (2010b). Tracing the evolution of the floral homeotic B- and C-function genes through genome synteny. Molecular Biology and Evolution, 27: 26512664.Google Scholar
Causier, B., Castillo, R., Zhou, J. et al. (2005). Evolution in action: following function in duplicated floral homeotic genes. Current Biology, 15: 15081512.Google Scholar
Causier, B., Schwarz-Sommer, Z. & Davies, B. (2010a). Floral organ identity: 20 years of ABCs. Seminars in Cell and Developmental Biology, 21: 7379.Google Scholar
Cavalier-Smith, T., Chao, E. E., Snell, E. A. et al. (2014). Multigene eukaryote phylogeny reveals the likely protozoan ancestors of opisthokonts (animals, fungi, choanozoans) and Amoebozoa. Molecular Phylogenetics and Evolution, 81: 7185.Google Scholar
Cevik, V., Ryder, C. D., Popovich, A. et al. (2010). A FRUITFULL-like gene is associated with genetic variation for fruit flesh firmness in apple (Malus domestica Borkh.). Tree Genetics and Genomes, 6: 271279.Google Scholar
Chae, E., Tan, Q. K. G., Hill, T. A. & Irish, V. F. (2008). An Arabidopsis F-box protein acts as a transcriptional co-factor to regulate floral development. Development, 135: 12351245.Google Scholar
Champagne, C. & Sinha, N. (2004). Compound leaves: equal to the sum of their parts? Development, 131: 44014412.Google Scholar
Champagne, C. E., Goliber, T. E., Wojchiechowski, M. F. et al. (2007). Compound leaf development and evolution in the legumes. Plant Cell, 19: 33693378.Google Scholar
Chanderbali, A. S., Albert, V. A., Leebens-Mack, J. et al. (2009). Transcriptional signatures of ancient floral developmental genetics in avocado (Persea americana; Lauraceae). Proceedings of the National Academy of Sciences of the United States of America, 106: 89298934.Google Scholar
Chanderbali, A. S., Berger, B. A., Howarth, D. G., Soltis, D. E. & Soltis, P. S. (2017). Evolution of floral diversity: genomics, genes and gamma. Philosophical Transactions of the Royal Society B, 372: 20150509.Google Scholar
Chandler, J., Nardmann, J. & Werr, W. (2008). Plant development revolves around axes. Trends in Plant Science, 13: 7884.Google Scholar
Chapman, M. A., Tang, S., Draeger, D. et al. (2012). Genetic analysis of floral symmetry in Van Gogh’s sunflowers reveals independent recruitment of CYCLOIDEA genes in the Asteraceae. PLoS Genetics, 8: e1002628.Google Scholar
Charlesworth, B. (1980). Evolution in Age-structured Populations. Cambridge: Cambridge University Press.Google Scholar
Charlesworth, D. & Guttman, D. S. (1999). The evolution of dioecy and plant sex chromosome systems. In Sex Determination in Plants, ed. Ainsworth, C. C.. Oxford: Bios Scientific, pp. 2549.Google Scholar
Chen, C., Xu, Y., Zeng, M. & Huang, H. (2001). Genetic control by Arabidopsis genes LEUNIG and FILAMENTOUS FLOWER in gynoecium fusion. Journal of Plant Research, 114: 465469.Google Scholar
Chen, C. B., Wang, S. P. & Huang, H. (2000). LEUNIG has multiple functions in gynoecium development in Arabidopsis. Genesis, 26: 4254.Google Scholar
Chen, J. J., Janssen, B. J., Williams, A. & Sinha, N. (1997). A gene fusion at a homeobox locus: alterations in leaf shape and implications for morphological evolution. Plant Cell, 9: 12891304.Google Scholar
Chen, T. C., Zhang, D. X., Larsen, K. & Larsen, S. S. (2010). Bauhinia Linnaeus. In Flora of China, 10, eds. Wu, Z. Y., Raven, P. H. & Hong, D. Y.. Beijing: Science Press; St. Louis, MO: Missouri Botanical Garden Press, pp. 621.Google Scholar
Chen, X. (2012). Small RNAs in development: insights from plants. Current Opinion in Genetics and Development, 22: 361367.Google Scholar
Chitwood, D. H., Headland, L. R., Ranjan, A. et al. (2012). Leaf asymmetry as a developmental constraint imposed by auxin-dependent phyllotactic patterning. Plant Cell, 24: 110.Google Scholar
Cho, E. & Zambryski, P. C. (2011). ORGAN BOUNDARY1 defines a gene expressed at the junction between the shoot apical meristem and lateral organs. Proceedings of the National Academy of Sciences of the United States of America, 108: 21542159.Google Scholar
Cho, J. W., Park, S. C., Shin, E. A. et al. (2004). Cyclin D1 and p22ack1 play opposite roles in plant growth and development. Biochemical and Biophysical Research Communications, 324: 5257.Google Scholar
Chuang, C. F., Running, M. P., Williams, R. W. & Meyerowitz, E. M. (1999). The PERIANTHIA gene encodes a bZIP protein involved in the determination of floral organ number in Arabidopsis thaliana. Genes and Development, 13: 334344.Google Scholar
Chuck, G., Meeley, R. &, Hake, S. (2008). Floral meristem initiation and meristem cell fate are regulated by the maize AP2 genes ids1 and sid1. Development, 135: 30133019.Google Scholar
Chung, Y. Y., Kim, S. R., Kang, H. G. et al. (1995). Characterization of two rice MADS box genes homologous to GLOBOSA. Plant Science, 109: 4556.Google Scholar
Citerne, H., Jabbour, F., Nadot, S. & Damerval, C. (2010). The evolution of floral symmetry. Advances in Botanical Research, 54: 85137.Google Scholar
Citerne, H. L., Möller, M. & Cronk, Q. C. B. (2000). Diversity of cycloidea-like genes in Gesneriaceae in relation to floral symmetry. Annals of Botany, 86: 167176.Google Scholar
Citerne, H. L., Pennington, R. T. & Cronk, Q. C. (2006). An apparent reversal in floral symmetry in the legume Cadia is a homeotic transformation. Proceedings of the National Academy of Sciences of the United States of America, 103: 1201712020.Google Scholar
Citerne, H. L., Reyes, E., Le Guilloux, M. et al. (2017). Characterization of CYCLOIDEA-like genes in Proteaceae, a basal eudicot family with multiple shifts in floral symmetry. Annals of Botany, 119: 367378.Google Scholar
Clark, S. E. (2001). Meristems: start your signaling. Current Opinion in Plant Biology, 4: 2832.Google Scholar
Clarke, E. (2011). Plant individuality: a solution to the demographer’s dilemma, Biology and Philosophy, 27: 321361.Google Scholar
Classen-Bockhoff, R. (1992). Florale Differenzierung in komplex organisierten Asteraceenköpfen. Flora, 186: 122.Google Scholar
Classen-Bockhoff, R. (1996). Functional units beyond the level of the capitulum and cypsela in Compositae. In Compositae: Biology and Utilization, ed. Hind, D. J. N.. Kew: Royal Botanic Gardens, pp. 129160.Google Scholar
Classen-Bockhoff, R. (2001). Plant morphology: the historic concepts of Wilhelm Troll, Walter Zimmermann and Agnes Arber. Annals of Botany, 88: 11531172.Google Scholar
Clausen, R. E. & Mann, M. C. (1924). Inheritance of Nicotiana tabacum. V. The occurrence of haploid plants in interspecific progenies. Proceedings of the National Academy of Sciences of the United States of America, 10: 121124.Google Scholar
Clausing, G. & Renner, S. S. (2001). Evolution of growth in epiphytic Dissochaeteae (Melastomataceae). Organisms Diversity and Evolution, 1: 4560.Google Scholar
Clay, K. & Ellstrand, N. (1981). Stylar polymorphism in Epigaea repens, a dioecious species. Bulletin of the Torrey Botanical Club, 108: 305310.Google Scholar
Clifford, H. T. (1987). Spikelet and floral morphology. In Grass Systematics and Evolution, eds. Soderstrom, T. R., Hilu, K. W., Campbell, C. S. & Barkworth, M. E.. Washington, DC: Smithsonian Institution Press, pp. 2130.Google Scholar
Clune, J., Mouret, J. B. & Lipson, H. (2013). The evolutionary origins of modularity. Proceedings of the Royal Society B, 280: 20122863.Google Scholar
Coen, E. (1999). The Art of Genes. How Organisms Make Themselves. Oxford: Oxford University Press.Google Scholar
Coen, E. S., Doyle, S., Romero, J. M. et al. (1991). Homeotic genes controlling flower development in Antirrhinum. Development, Supplement, 1: 149155.Google Scholar
Coen, E. S. & Meyerowitz, E. M. (1991). The war of the whorls: genetic interactions controlling flower development. Nature, 353: 3137.Google Scholar
Coen, E. S. & Nugent, J. M. (1994). The evolution of flowers and inflorescences. Development, 1994 (supplement): 107116.Google Scholar
Coen, E., Rolland-Lagan, A.-G., Matthews, M., Bangham, J.A. & Prusinkiewicz, P. (2004). The genetics of geometry. Proceedings of the National Academy of Sciences of the United States of America, 101: 47284735.Google Scholar
Coen, E. S., Romero, J. M., Doyle, S. et al. (1990). Floricaula: a homeotic gene required for flower development in Antirrhinum majus. Cell, 63: 13111322.Google Scholar
Cohen, J. I. (2014). A phylogenetic analysis of morphological and molecular characters of Boraginaceae: evolutionary relationships, taxonomy, and patterns of character evolution. Cladistics, 30: 139169.Google Scholar
Cole, M., Nolte, C. & Werr, W. (2006). Nuclear import of the transcription factor SHOOT MERISTEMLESS depends on heterodimerization with BLH proteins expressed in discrete sub-domains of the shoot apical meristem of Arabidopsis thaliana. Nucleic Acids Research, 34: 12811292.Google Scholar
Colombo, L., Battaglia, R. & Kater, M. M. (2008). Arabidopsis ovule development and its evolutionary conservation. Trends in Plant Science, 13: 444450.Google Scholar
Colombo, L., Franken, J., Koetje, E. et al. (1995). The petunia MADS box gene FBP11 determines ovule identity. Plant Cell, 7: 18591868.Google Scholar
Conner, J. K. (2012). Quantitative genetic approaches to evolutionary constraint: how useful? Evolution, 66: 33133320.Google Scholar
Conti, L. & Bradley, D. (2007). TERMINAL FLOWER1 is a mobile signal controlling Arabidopsis architecture. Plant Cell, 19: 767778.Google Scholar
Cook, C. D. K. & Rutishauser, R. (2007). Podostemaceae. In The Families and Genera of Vascular Plants, Vol. 9, ed. Kubitzki, K.. Berlin: Springer, pp. 304344.Google Scholar
Cooke, T. J., Poli, D. & Cohen, J. D. (2003). Did auxin play a crucial role in the evolution of novel body plans during the Late Silurian-Early Devonian radiation of land plants? In The Evolution of Plant Physiology: From Whole Plants to Ecosystems, eds. Hemsley, A. R. & Poole, I.. London: Linnean Society of London, pp. 85107.Google Scholar
Corbesier, L., Vincent, C., Jang, S. H. et al. (2007). FT protein movement contributes to long-distance signaling in floral induction of Arabidopsis. Science, 316: 10301033.Google Scholar
Corley, S. B., Carpenter, R., Copsey, L. & Coen, E. (2005). Floral asymmetry involves an interplay between TCP and MYB transcription factors in Antirrhinum. Proceedings of the National Academy of Sciences of the United States of America, 102: 50685073.Google Scholar
Coudert, Y., Périn, C., Courtois, B., Khong, N. G. & Gantet, P. (2010). Genetic control of root development in rice, the model cereal. Trends in Plant Science, 15: 219226.Google Scholar
Couvreur, T. L. P., Franzke, A., Al-Shehbaz, I. A. et al. (2010). Molecular phylogenetics, temporal diversification and principles of evolution in the mustard family (Brassicaceae). Molecular Biology and Evolution, 27: 5571.Google Scholar
Crane, P. R. (1985). Phylogenetic analysis of seed plants and the origin of angiosperms. Annals of the Missouri Botanical Garden, 72: 716793.Google Scholar
Crane, P. R., Friis, E. M. & Pedersen, K. R. (1995). The origin and early diversification of angiosperms. Nature, 374: 2733.Google Scholar
Crane, P. R. & Kenrick, P. (1997). Diverted development of reproductive organs: a source of morphological innovation in land plants. Plant Systematics and Evolution, 206: 161174.Google Scholar
Cremer, F., Lönnig, W. E., Saedler, H. & Huijser, P. (2001). The delayed terminal flower phenotype is caused by a conditional mutation in the CENTRORADIALIS gene of snapdragon. Plant Physiology, 126: 10311041.Google Scholar
Cridge, A. G., Dearden, P. K. & Brownfield, L. R. (2016). Convergent occurrence of the developmental hourglass in plant and animal embryogenesis? Annals of Botany, 117: 833843.Google Scholar
Cronk, Q. C. B. (2001). Plant evolution and development in a post-genomic context. Nature Reviews Genetics, 2: 607619.Google Scholar
Cronk, Q. C. B. (2009). The Molecular Organography of Plants. Oxford: Oxford University Press.Google Scholar
Cronquist, A. (1968). The Evolution and Classification of Flowering Plants. Boston, MA: Houghton Mifflin.Google Scholar
Cronquist, A. (1981). An Integrated System of Classification of Flowering Plants. New York, NY: Columbia University Press.Google Scholar
Cubas, P. (2004). Floral zygomorphy, the recurring evolution of a successful trait. Bioessays, 26: 11751184.Google Scholar
Cubas, P., Vincent, C. & Coen, E. (1999). An epigenetic mutation responsible for natural variation in floral symmetry. Nature, 401: 157161.Google Scholar
Cui, R., Han, J., Zhao, S. et al. (2010). Functional conservation and diversification of class E floral homeotic genes in rice (Oryza sativa). The Plant Journal, 61: 767781.Google Scholar
Czapek, A. (1898). Die inverse Orientierung der Blätter von Alstroemeria. Flora, 85: 418430.Google Scholar
Dahmann, C., Oates, A. C. & Brand, M. (2011). Boundary formation and maintenance in tissue development. Nature Reviews Genetics, 12: 4355.Google Scholar
Dai, M., Zhao, Y., Ma, O. et al. (2007). The rice YABBY1 gene is involved in the feedback regulation of gibberellin metabolism. Plant Physiology, 144: 121133.Google Scholar
Damerval, C., Citerne, H., Le Guilloux, M. et al. (2013). Asymmetric morphogenetic cues along the transverse plane: shift from disymmetry to zygomorphy in the flower of Fumarioideae. American Journal of Botany, 100: 391402.Google Scholar
Damerval, C., Le Guilloux, M., Jager, M. & Charon, C. (2007). Diversity and evolution of CYCLOIDEA-like TCP genes in relation to flower development in Papaveraceae. Plant Physiology, 143: 759772.Google Scholar
Damerval, C. & Nadot, S. (2007). Evolution of perianth and stamen characteristics with respect to floral symmetry in Ranunculales. Annals of Botany, 100: 631640.Google Scholar
Daniell, H., Lin, C. S., Yu, M. & Chang, W. J. (2016). Chloroplast genomes: diversity, evolution, and applications in genetic engineering. BMC Genome Biology, 17(1): 129.Google Scholar
Danyluk, J., Kane, N. A., Breton, G. et al. (2003). TaVRT-1, a putative transcription factor associated with vegetative to reproductive transition in cereals. Plant Physiology, 132: 18491860.Google Scholar
D’Arcy, W. (1991). The Solanaceae since 1976, with a review of its biogeography. In Solanaceae III: Taxonomy, Chemistry, Evolution, eds. Hawkes, J. G., Lester, R. W., Nee, M. & Estrada, R. N.. Kew: Royal Botanic Gardens; London: Linnean Society of London, pp. 75137.Google Scholar
Darwin, C. (1859). On the Origin of Species by Natural Selection. London: J. Murray.Google Scholar
Darwin, C. (1877). The Different Forms of Flowers on Plants of the Same Species. London: Murray.Google Scholar
Datson, P. M., Murray, B. G. & Steiner, K. E. (2008). Climate and the evolution of annual/perennial life-histories in Nemesia (Scrophulariaceae). Plant Systematics and Evolution, 270: 3957.Google Scholar
Davies, B., Motte, P., Keck, E. et al. (1999). PLENA and FARINELLI: redundancy and regulatory interactions between two Antirrhinum MADS-box factors controlling flower development. EMBO Journal, 18: 40234034.Google Scholar
Davis, C. & Wurdack, K. (2004). Host-to-parasite gene transfer in flowering plants: phylogenetic evidence from Malpighiales. Science, 305: 676678.Google Scholar
Davis, C. C. & Anderson, W. R. (2010). A complete generic phylogeny of Malpighiaceae inferred from nucleotide sequence data and morphology. American Journal of Botany, 97: 20312048.Google Scholar
Davis, C. C. & Xi, Z. (2015). Horizontal gene transfer in parasitic plants. Current Opinion in Plant Biology, 26: 1419.Google Scholar
Dawkins, R. (1976). The Selfish Gene. Oxford: Oxford UniversityPress.Google Scholar
de Beer, G. R. (1930). Embryology and Evolution. Oxford: Clarendon Press.Google Scholar
de Beer, G. R. (1940). Embryos and Ancestors. Oxford: Clarendon Press.Google Scholar
de Bruijn, S., Angenent, G. C. & Kaufmann, K. (2012). Plant ‘evo-devo’ goes genomic: from candidate genes to regulatory networks. Trends in Plant Science, 17: 441447.Google Scholar
de Lange, P. J., Heenan, P. B., Houliston, G. J., Rolfe, J. R. & Mitchell, A. D. (2013). New Lepidium (Brassicaceae) from New Zealand. PhytoKeys, 24: 1147.Google Scholar
de Martino, G., Pan, I., Emmanuel, E., Levy, A. & Irish, V. F. (2006). Functional analyses of two tomato APETALA3 genes demonstrate diversification in their roles in regulating floral development. Plant Cell, 18: 18331845.Google Scholar
De Smet, I., Lau, S., Mayer, U. & Jürgens, G. (2010). Embryogenesis: the humble beginnings of plant life. The Plant Journal, 61: 959970.Google Scholar
de Vries, H. (1904). Species and Varieties: Their Origin by Mutation. Chicago, IL: Open Court.Google Scholar
Della Pina, S., Souer, E. & Koes, R. (2014). Arguments in the evo-devo debate: say it with flowers! Journal of Experimental Botany, 65: 22312242.Google Scholar
DeMason, D. A. & Schmidt, R. J. (2001). Roles of the Uni gene in shoot and leaf development of pea (Pisum sativum): phenotypic characterization and leaf development in the uni and uni-tac mutants. International Journal of Plant Sciences, 162: 10331051.Google Scholar
Dengler, N. G. (1999). Anisophylly and dorsiventral shoot symmetry. International Journal of Plant Sciences, 160: S67S80.Google Scholar
Dengler, N. G., Dengler, R. E. & Kaplan, D. R. (1982). The mechanism of plication in palm leaves: histogenic observations on the pinnate leaf of Chrysalidocarpus lutescens. Canadian Journal of Botany, 60: 29762980.Google Scholar
Derelle, R., Lopez, P., Le Guyader, H. & Manuel, M. (2007). Homeodomain proteins belong to the ancestral molecular toolkit of eukaryotes. Evolution and Development, 9: 212219.Google Scholar
Deroin, T. (2007). Floral vascular pattern of the endemic Malagasy genus Fenerivia Diels (Annonaceae). Adansonia, 29: 712.Google Scholar
Dewitte, W., Scofield, S., Alcasabas, A. A. et al. (2007). Arabidopsis CYCD3 D-type cyclins link cell proliferation and endocycles and are rate-limiting for cytokinin responses. Proceedings of the National Academy of Sciences of the United States of America, 104: 1453714542.Google Scholar
Di Giacomo, E., Sestili, F., Iannelli, M. A. et al. (2008). Characterization of KNOX genes in Medicago truncatula. Plant Molecular Biology, 67: 135150.Google Scholar
Dickinson, T. A. (1978). Epiphylly in angiosperms. Botanical Review, 44: 181232.Google Scholar
Diggle, P. K. (1991). Labile sex expression in andromonoecious Solanum hirtum: sources of variation in mature floral structure. Canadian Journal of Botany, 69: 20332043.Google Scholar
Diggle, P. K. (1993). Developmental plasticity, genetic variation and the evolution of andromonoecy. American Journal of Botany, 80: 967973.Google Scholar
Diggle, P. K. (1994). The expression of andromonoecy in Solanum hirtum: phenotypic plasticity and ontogenetic contingency. American Journal of Botany, 81: 13541365.Google Scholar
Diggle, P. K. (1995). Architectural effects and the interpretation of patterns of fruit and seed development. Annual Review of Ecology and Systematics, 26: 531552.Google Scholar
Diggle, P. K. (1997). Extreme preformation in an alpine Polygonum viviparum: an architectural and developmental analysis. American Journal of Botany, 84: 154169.Google Scholar
Diggle, P. K. (1999). Heteroblasty and the evolution of flowering phenologies. International Journal of Plant Sciences, 160: S123S134.Google Scholar
Diggle, P. K. (2003). Architectural effects on floral form and function: a review. In Deep Morphology: Toward a Renaissance of Morphology in Plant Systematics, eds. Stuessy, T., Hörandl, E. & Mayer, V.. Königstein: Koeltz, pp. 6380.Google Scholar
Diggle, P. K. (2014). Modularity and intra-floral integration in metameric organisms: plants are more than the sum of their parts. Philosophical Transactions of the Royal Society B, 369: 20130253.Google Scholar
Diggle, P. K., Di Stilio, V. S., Gschwend, A. R. et al. (2011). Multiple developmental processes underlie sex differentiation in angiosperms. Trends in Genetics, 27: 368376.Google Scholar
Dilcher, D. L. & Crane, P. R. (1984). Archaeanthus: an early angiosperm from the Cenomanian of the western interior of North America. Annals of the Missouri Botanical Garden, 71: 351383.Google Scholar
Dinneny, J. R., Weigel, D. & Yanofsky, M. F. (2005). A genetic framework for fruit patterning in Arabidopsis thaliana. Development, 132: 46874696.Google Scholar
Dinneny, J. R., Yadegari, R., Fischer, R. L., Yanofsky, M. F. & Weigel, D. (2004). The role of JAGGED in shaping lateral organs. Development, 131: 11011110.Google Scholar
Dinneny, J. R. & Yanofsky, M. F. (2005). Drawing lines and borders: how the dehiscent fruit of Arabidopsis is patterned. BioEssays, 27: 4249.Google Scholar
Ditta, G., Pinyopich, A., Robles, P., Pelaz, S. & Yanofsky, M. F. (2004). The SEP4 gene of Arabidopsis thaliana functions in floral organ and meristem identity. Current Biology, 14: 19351940.Google Scholar
Dodds, P. N. & Rathjen, J. P. (2010). Plant immunity: towards an integrated view of plant-pathogen interactions. Nature Reviews Genetics, 11: 539548.Google Scholar
Doebley, J., Stec, A. & Hubbard, L. (1997). The evolution of apical dominance in maize. Nature, 386: 485488.Google Scholar
Domazet-Lošo, T., Brajković, J. & Tautz, D. (2007). A phylostratigraphy approach to uncover the genomic history of major adaptations in metazoan lineages. Trends in Genetics, 23, 533539.Google Scholar
Domoney, C., Duc, G., Ellis, T. H. N. et al. (2006). Genetic and genomic analysis of legume flowers and seeds. Current Opinion in Plant Biology, 9: 133141.Google Scholar
Donnelly, P. M., Bonetta, D., Tsukaya, H., Dengler, R. & Dengler, N. G. (1999). Cell cycling and cell enlargement in developing leaves of Arabidopsis. Developmental Biology, 215: 407419.Google Scholar
Donoghue, M. J. (1992). Homology. In Keywords in Evolutionary Biology, eds. Keller, E. F. & Lloyd, E. A.. Cambridge, MA: Harvard University Press, pp. 170179.Google Scholar
Donoghue, M. J. & Ree, R. H. (2000). Homoplasy and developmental constraint: a model and an example from plants. American Zoologist, 40: 759769.Google Scholar
Donoghue, M. J., Ree, R. H. & Baum, D. A. (1998). Phylogeny and the evolution of flower symmetry in the Asteridae. Trends in Plant Science, 3: 311317.Google Scholar
Doust, A. N. (2001). The developmental basis of floral variation in Drimys winteri (Winteraceae). International Journal of Plant Sciences, 162: 697717.Google Scholar
Doust, A. N., Mauro-Herrera, M., Francis, A. D. & Shand, L. C. (2014). Morphological diversity and genetic regulation of inflorescence abscission zones in grasses. American Journal of Botany, 101: 17591769.Google Scholar
Doyle, J. A. (2008). Integrating molecular phylogenetic evidence and paleobotanical evidence on the origin of the flower. International Journal of Plant Sciences, 167: 816843.Google Scholar
Doyle, J. A. & Endress, P. K. (2000). Morphological phylogenetic analysis of basal angiosperms: comparison and combination with molecular data. International Journal of Plant Sciences, 161: S121S153.Google Scholar
Doyle, J. A. & Endress, P. K. (2011). Tracing the evolutionary diversification of the flower in basal angiosperms. In Flowers on the Tree of Life, eds. Wanntorp, L. & De Craene, L. P. Ronse. Cambridge: Cambridge University Press, pp. 88119.Google Scholar
Draghi, J. & Wagner, G. P. (2008). Evolution of evolvability in a developmental model. Evolution, 62: 301315.Google Scholar
Dransfield, J. & Uhl, N.W. (1998). Palmae. In The Families and Genera of Vascular Plants. Vol. 4, ed. Kubitzki, K.. Berlin: Springer, pp. 306388.Google Scholar
Drea, S., Hileman, L. C., de Martino, G. & Irish, V. F. (2007). Functional analyses of genetic pathways controlling petal specification in poppy. Development, 134: 41574166.Google Scholar
Dreni, L., Jacchia, S., Fornara, F. et al. (2007). The D-lineage MADS-box gene OsMADS13 controls ovule identity in rice. The Plant Journal, 52: 690699.Google Scholar
Drinnan, A. N., Crane, P. R. & Hoot, S. B. (1994). Patterns of floral evolution in the early diversification of nonmagnoliid dicotyledons (eudicots). Plant Systematics and Evolution, Supplement, 8: 93122.Google Scholar
Drost, H. G., Gabel, A., Grosse, I. & Quint, M. (2015). Evidence for active maintenance of phylotranscriptomic hourglass patterns in animal and plant embryogenesis. Molecular Biology and Evolution, 32: 12211231.Google Scholar
Du, X., Xiao, Q., Zhao, R. et al. (2008). TrMADS3, a new MADS-box gene, from a perennial species Taihangia rupestris (Rosaceae) is upregulated by cold and experiences seasonal fluctuation in expression level. Development Genes and Evolution, 218: 281292.Google Scholar
Du, Z.-Y. & Wang, Y.-Z. (2008). Significance of RT-PCR expression patterns of CYC-like genes in Oreocharis benthamii (Gesneriaceae). Journal of Systematics and Evolution, 46: 2331.Google Scholar
Duboule, D. (1994). Temporal colinearity and the phylotypic progression: a basis for the stability of a vertebrate Bauplan and the evolution of morphologies through heterochrony. Development, Supplement, 1994: 135142.Google Scholar
Dubrovsky, J. G., Gambetta, G. A., Hernández-Barrera, A., Shishkova, S. & González, I. (2006). Lateral root initiation in Arabidopsis: developmental window, spatial patterning, density and predictability. Annals of Botany, 97: 903915.Google Scholar
Dunwell, J. M. (2010). Haploids in flowering plants: origins and exploitation. Plant Biotechnology Journal, 8: 377424.Google Scholar
Duthion, C., Ney, B. & Munier-Jolain, N. M. (1994). Development and growth of white lupin: implications for crop management. Agronomy Journal, 86: 10391045.Google Scholar
Eames, A. J. (1961). Morphology of the Angiosperms. New York, NY: McGraw-Hill.Google Scholar
Efroni, I., Blum, E., Goldshmidt, A. & Eshed, Y. (2008). A protracted and dynamic maturation schedule underlies Arabidopsis leaf development. Plant Cell, 20: 22932306.Google Scholar
Efroni, I., Eshed, Y. & Lifschitz, E. (2010). Morphogenesis of simple and compound leaves: a critical review. Plant Cell, 22: 10191032.Google Scholar
Egea-Cortines, M. & Weiss, J. (2013). Control of plant organ size. In Encyclopedia of Life Sciences. Chichester: Wiley.Google Scholar
Ehrenreich, I. M. & Pfennig, D. W. (2016). Genetic assimilation: a review of its potential proximate causes and evolutionary consequences. Annals of Botany, 117: 769779.Google Scholar
El Ottra, J. H. L., Pirani, J. R. & Endress, P. K. (2013). Fusion within and between whorls of floral organs in Galipeinae (Rutaceae): structural features and evolutionary implications. Annals of Botany, 111: 821837.Google Scholar
Elo, A., Lemmetyinen, J., Novak, A. et al. (2007). BpMADS4 has a central role in inflorescence initiation in silver birch (Betula pendula). Physiologia Plantarum, 131: 149158.Google Scholar
Elsner, J., Michalski, M. & Kwiatkowska, D. (2012). Spatiotemporal variation of leaf epidermal cell growth: a quantitative analysis of Arabidopsis thaliana wildtype and triple cyclinD3 mutant plants. Annals of Botany, 109: 897910.Google Scholar
Emery, J. F., Floyd, S. K., Alvarez, J. et al. (2003). Radial patterning of Arabidopsis shoots by class III HD-ZIP and KANADI genes. Current Biology, 13: 17681774.Google Scholar
Endress, M. E. (2001). Apocynaceae and Asclepiadaceae: united they stand. Haseltonia, 8: 29.Google Scholar
Endress, P. K. (1970). Die Infloreszenzen der apetalen Hamamelidaceen, ihre grundsätzliche morphologische und systematische Bedeutung. Botanische Jahrbücher für Systematik, Pflanzengeschichte und Pflanzengeographie, 90: 154.Google Scholar
Endress, P. K. (1976). Die Androeciumanlage bei polyandrischen Hamamelidaceen und ihre systematische Bedeutung. Botanische Jahrbücher für Systematik, Pflanzengeschichte und Pflanzengeographie, 97: 436457.Google Scholar
Endress, P. K. (1978). Blütenontogenese, Blütenabgrenzung und systematische Stellung der perianthlosen Hamamelidoideae. Botanische Jahrbücher für Systematik, Pflanzengeschichte und Pflanzengeographie, 100: 249317.Google Scholar
Endress, P. K. (1982). Syncarpy and alternative modes of escaping disadvantages of apocarpy in primitive angiosperms. Taxon, 31: 4852.Google Scholar
Endress, P. K. (1984). The flowering process in the Eupomatiaceae (Magnoliales). Botanische Jahrbücher für Systematik, Pflanzengeschichte und Pflanzengeographie, 104: 297319.Google Scholar
Endress, P. K. (1987a). The Chloranthaceae: reproductive structures and phylogenetic position. Botanische Jahrbücher für Systematik, Pflanzengeschichte und Pflanzengeographie, 109: 153226.Google Scholar
Endress, P. K. (1987b). Floral phyllotaxis and floral evolution. Botanische Jahrbücher für Systematik, Pflanzengeschichte und Pflanzengeographie, 108: 417438.Google Scholar
Endress, P. K. (1989). Chaotic floral phyllotaxis and reduced perianth in Achlys (Berberidaceae). Botanica Acta, 102: 159163.Google Scholar
Endress, P. K. (1990a). Patterns of floral construction in ontogeny and phylogeny. Biological Journal of the Linnean Society, 39: 153175.Google Scholar
Endress, P. K. (1990b). Evolution of reproductive structures and functions in primitive angiosperms (Magnoliidae). Memoirs of the New York Botanical Garden, 55: 534.Google Scholar
Endress, P. K. (1992). Evolution and floral diversity: the phylogenetic surroundings of Arabidopsis and Antirrhinum. International Journal of Plant Sciences, 153: S106S122.Google Scholar
Endress, P. K. (1994a). Diversity and Evolutionary Biology of Tropical Flowers. Cambridge: Cambridge University Press.Google Scholar
Endress, P. K. (1994b). Floral structure and evolution of primitive angiosperms: recent advances. Plant Systematics and Evolution, 192: 7997.Google Scholar
Endress, P. K. (1994c). Evolutionary aspects of the floral structure in Ceratophyllum. Plant Systematics and Evolution, Supplement, 8: 175183.Google Scholar
Endress, P. K. (1995). Floral structure and evolution in Ranunculanae. Plant Systematics and Evolution, Supplement, 9: 4761.Google Scholar
Endress, P. K. (1996). Homoplasy in angiosperm flowers. In Homoplasy: The Recurrence of Similarity in Evolution, eds. Sanderson, M. J. & Hufford, L.. San Diego, CA: Academic Press, pp. 303325.Google Scholar
Endress, P. K. (1997). Evolutionary biology of flowers: prospects for the next century. In Evolution and Diversification of Land Plants, ed. Iwatsuki, K. & Raven, P. H.. Tokyo: Springer-Verlag, pp. 99119.Google Scholar
Endress, P. K. (1998). Antirrhinum and Asteridae: evolutionary changes of floral symmetry. Symposia of the Society for Experimental Biology, 51: 133140.Google Scholar
Endress, P. K. (1999). Symmetry in flowers: diversity and evolution. International Journal of Plant Sciences, 160: S3S23.Google Scholar
Endress, P. K. (2001a). Evolution of floral symmetry. Current Opinion in Plant Biology, 4: 8691.Google Scholar
Endress, P. K. (2001b). Origins of flower morphology. Journal of Experimental Zoology (Molecular and Developmental Evolution), 291: 105115.Google Scholar
Endress, P. K. (2001c). The flowers in extant basal angiosperms and inferences on ancestral flowers. International Journal of Plant Sciences, 162: 11111140.Google Scholar
Endress, P. K. (2004). Structure and relationships of basal relictual angiosperms. Australian Systematic Botany, 17: 343366.Google Scholar
Endress, P. K. (2006). Angiosperm floral evolution: morphological and developmental framework. Advances in Botanical Research, 44: 161.Google Scholar
Endress, P. K. (2008a). Perianth biology in the basal grade of extant angiosperms. International Journal of Plant Sciences, 169: 844862.Google Scholar
Endress, P. K. (2008b). The whole and the parts: relationships between floral architecture and floral organ shape, and their repercussions on the interpretation of fragmentary floral fossils. Annals of the Missouri Botanical Garden, 95: 101120.Google Scholar
Endress, P. K. (2010a). Flower structure and trends of evolution in eudicots and their major subclades. Annals of the Missouri Botanical Garden, 97: 541583.Google Scholar
Endress, P. K. (2010b). Disentangling confusions in inflorescence morphology: patterns and diversity of reproductive shoot ramification in angiosperms. Journal of Systematics and Evolution, 48: 225239.Google Scholar
Endress, P. K. (2010c). The evolution of floral biology in basal angiosperms. Philosophical Transactions of the Royal Society B, 365: 411421.Google Scholar
Endress, P. K. (2010d). Synorganisation without organ fusion in the flowers of Geranium robertianum (Geraniaceae) and its not so trivial obdiplostemony. Annals of Botany, 106: 687695.Google Scholar
Endress, P. K. (2011). Evolutionary diversification of the flowers in angiosperms. American Journal of Botany, 98: 370396.Google Scholar
Endress, P. K. (2012). The immense diversity of floral monosymmetry and asymmetry across angiosperms. Botanical Review, 78: 345397.Google Scholar
Endress, P. K. (2014). Multicarpellate gynoecia in angiosperms - occurrence, development, organization and architectural constraints. Botanical Journal of the Linnean Society, 174: 143.Google Scholar
Endress, P. K. (2015). Patterns of angiospermy development before carpel sealing across living angiosperms: diversity, and morphological and systematic aspects. Botanical Journal of the Linnean Society, 178: 556591.Google Scholar
Endress, P. K. & Doyle, J. A. (2007). Floral phyllotaxis in basal angiosperms: development and evolution. Current Opinion in Plant Biology, 10: 5257.Google Scholar
Endress, P. K. & Doyle, J. A. (2015) Ancestral traits and specializations in the flowers of the basal grade of living angiosperms. Taxon, 64: 10931116.Google Scholar
Endress, P. K. & Igersheim, A. (2000a) Gynoecium structure and evolution in basal angiosperms. International Journal of Plant Sciences, 161: S211S223.Google Scholar
Endress, P. K. & Igersheim, A. (2000b). The reproductive structures of the basal angiosperm Amborella trichopoda (Amborellaceae). International Journal of Plant Sciences, 161: S237S248.Google Scholar
Endress, P. K., Igersheim, A., Sampson, F. B. & Schatz, G. E. (2000). Floral structure of Takhtajania and its systematic position in Winteraceae. Annals of the Missouri Botanical Garden, 87: 347365.Google Scholar
Endress, P. K. & Lorence, D. H. (2004). Heterodichogamy of a novel type in Hernandia (Hernandiaceae) and its structural basis. International Journal of Plant Sciences, 165: 753763.Google Scholar
Endress, P. K. & Matthews, M. L. (2006a). Elaborate petals and staminodes in eudicots: diversity, function, and evolution. Organisms, Diversity and Evolution, 6: 257293.Google Scholar
Endress, P. K. & Matthews, M. L. (2006b). First steps towards a floral structural characterization of the major rosid subclades. Plant Systematics and Evolution, 260: 223251.Google Scholar
Erbar, C. & Leins, P. (1996). Distribution of the character states ‘early’ and ‘late sympetaly’ within the ‘Sympetalae Tetracyclicae’ and presumably related groups. Botanica Acta, 109: 427440.Google Scholar
Erbar, C. & Leins, P. (1997). Different patterns of floral development in whorled flowers, exemplified by Apiaceae and Brassicaceae. International Journal of Plant Sciences, 158: S49S64.Google Scholar
Erwin, D. H., Laflamme, M., Tweedt, S. M. et al. (2011). The Cambrian conundrum: early divergence and later ecological success in the early history of animals. Science, 334: 10911097.Google Scholar
Eshed, Y., Baum, S. F. & Bowman, J. L. (1999). Distinct mechanisms promote polarity establishment in carpels of Arabidopsis. Cell, 99: 199209.Google Scholar
Eshed, Y., Baum, S. F., Perea, J. V. & Bowman, J. L. (2001). Establishment of polarity in lateral organs of plants. Current Biology, 11: 12511260.Google Scholar
Eshed, Y., Izhaki, A., Baum, S. F., Floyd, S. K. & Bowman, J. L. (2004). Asymmetric leaf development and blade expansion in Arabidopsis are mediated by KANADI and YABBY activities. Development, 131: 29973006.Google Scholar
Espadaler, X. & Gómez, C. (2001). Female performance in Euphorbia characias: effect of flower position on seed quantity and quality. Seed Science Research, 11: 163172.Google Scholar
Evered, D. & Marsh, J. (1989). The Cellular Basis of Morphogenesis. Chichester: Wiley.Google Scholar
Ewers, F. W. & Schmid, R. (1981). Longevity of needle fascicles of Pinus longaeva (bristlecone pine) and other North American pines. Oecologia, 51: 107115.Google Scholar
Fahlgren, N., Jogdeo, S., Kasschau, K. D. et al. (2010). MicroRNA gene evolution in Arabidopsis lyrata and Arabidopsis thaliana. Plant Cell, 22: 10741089.Google Scholar
Fambrini, M., Salvini, M. & Pugliesi, C. (2011). A transposon-mediate inactivation of a CYCLOIDEA-like gene originates polysymmetric and androgynous ray flowers in Helianthus annuus. Genetica, 139: 15211529.Google Scholar
Fauron, C., Allen, J. O., Clifton, S. & Newton, K. J. (2004). Plant mitochondrial genomes. In Molecular Biology and Biotechnology of Plant Organelles, eds. Daniell, H. & Chase, C.. Dordrecht: Kluwer, pp. 155171.Google Scholar
Feng, G., Qin, Z., Yan, J., Zhang, X. & Hu, Y. (2011). Arabidopsis ORGAN SIZE RELATED1 regulates organ growth and final organ size in orchestration with ARGOS and ARL. New Phytologist, 191: 635646.Google Scholar
Feng, M. & Lu, A.-M. (1998). Floral organogenesis and its systematic significance of the genus Nandina (Berberidaceae). Acta Botanica Sinica, 40: 102108.Google Scholar
Feng, X., Zhao, Z., Tian, Z. et al. (2006). Control of petal shape and floral zygomorphy in Lotus japonicus. Proceedings of the National Academy of Sciences of the United States of America, 103: 49704975.Google Scholar
Ferjani, A., Horiguchi, G., Yano, S. & Tsukaya, H. (2007). Analysis of leaf development in fugu mutants of Arabidopsis reveals three compensation modes that modulate cell expansion in determinate organs. Plant Physiology, 144: 988999.Google Scholar
Ferrándiz, C., Liljegren, S. J. & Yanofsky, M. F. (2000). Negative regulation of the SHATTERPROOF genes by FRUITFULL during Arabidopsis fruit development. Science, 289: 436438.Google Scholar
Ferreira, B. G. & Isaias, R. M. S. (2014). Floral-like destiny induced by a galling Cecidomyiidae on the axillary buds of Marcetia taxifolia (Melastomataceae). Flora, 209: 391400.Google Scholar
Ferrero, V., Arroyo, J., Vargas, P., Thompson, J. D. & Navarro, L. (2009). Evolutionary transitions of style polymorphisms in Lithodora (Boraginaceae). Perspectives in Plant Ecology, Evolution and Systematics, 11: 111125.Google Scholar
Ferrero, V., Rojas, D., Vale, A. & Navarro, L. (2012). Delving into the loss of heterostyly in Rubiaceae: is there a similar trend in tropical and non-tropical zones? Perspectives in Plant Ecology, Evolution and Systematics, 14: 161167.Google Scholar
Fink, W. L. (1982). The conceptual relationship between ontogeny and phylogeny. Paleobiology, 8: 254264.Google Scholar
Fink, W. L. (1988). Phylogenetic analysis and the detection of ontogenetic patterns. In Heterochrony in Evolution: A Multidisciplinary Approach, ed. McKinney, M. L.. New York, NY: Plenum Press, pp. 7191.Google Scholar
Fischer, E. (2015). Magnoliopsida (Angiosperms) p.p.: Subclass Magnoliidae [Amborellanae to Magnolianae, Lilianae p.p. (Acorales to Asparagales)]. In Syllabus of Plant Families, 13th edn, Part 4, ed. Frey, W.. Stuttgart: Borntraeger, pp. 111495.Google Scholar
Fishbein, M. (2001). Evolutionary innovation and diversification in the flowers of Asclepiadaceae. Annals of the Missouri Botanical Garden, 88: 603623.Google Scholar
Fisher, J. B. (2002). Indeterminate leaves of Chisocheton (Meliaceae): survey of structure and development. Botanical Journal of the Linnean Society, 139: 207221.Google Scholar
Fisher, J. B. & Rutishauser, R. (1990). Leaves and epiphyllous shoots in Chisocheton (Meliaceae): a continuum of woody leaf and stem-axes. Canadian Journal of Botany, 68: 23162328.Google Scholar
Fisher, R. A. (1930). The Genetical Theory of Natural Selection. Oxford: Clarendon Press.Google Scholar
Fitting, H. (1921). Das Verblühen der Blüten. Die Naturwissenschaften, 9: 19.Google Scholar
Floyd, S. K. & Bowman, J. L. (2007). The ancestral developmental toolkit of land plants. International Journal of Plant Sciences, 168: 135.Google Scholar
Floyd, S. K. & Bowman, J. L. (2010). Gene expression patterns in seed plant shoot meristems and leaves: homoplasy or homology? Journal of Plant Research, 123: 4355.Google Scholar
Forest, F., Chase, M. W., Persson, C., Crane, P. R. & Hawkins, J. A. (2007). The role of biotic and abiotic factors in the evolution of ant-dispersal in the milkwort family (Polygalaceae). Evolution, 61: 16751694.Google Scholar
Foucher, F., Morin, J., Courtiade, J. et al. (2003). DETERMINATE and LATE FLOWERING are two TERMINAL FLOWER1/CENTRORADIALIS homologs that control two distinct phases of flowering initiation and development in pea. Plant Cell, 15: 27422754.Google Scholar
Frajman, B. & Schönswetter, P. (2011). Giants and dwarfs: molecular phylogenies reveal multiple origins of annual spurges within Euphorbia subg. Esula. Molecular Phylogenetics and Evolution, 61: 413424.Google Scholar
Francis, D. (2008). Apical meristems. In Encyclopedia of Life Sciences. Chichester: Wiley.Google Scholar
Franzke, A., Lysak, M. A., Al-Shehbaz, I. A., Koch, M. A. & Mummenhoff, K. (2011). Cabbage family affairs: the evolutionary history of Brassicaceae. Trends in Plant Science, 16: 108116.Google Scholar
Friedman, W. E., Moore, R. C. & Purugganan, M. D. (2004). The evolution of plant development. American Journal of Botany, 91: 17261741.Google Scholar
Friis, E. M., Doyle, J. A., Endress, P. K. & Leng, Q. (2003). Archaefructus: angiosperm precursor or specialized early angiosperm? Trends in Plant Science, 8: 369373.Google Scholar
Friis, E. M., Eklund, H., Pedersen, K. R. & Crane, P. R. (1994). Virginianthus calycanthoides gen. et sp. nov.: a calycanthaceous flower from the Potomac group (Early Cretaceous) of eastern North America. International Journal of Plant Sciences, 155: 772785.Google Scholar
Friis, E. M., Pedersen, K. R. & Crane, P. R. (2000). Reproductive structure and organization of basal angiosperms from the early Cretaceous (Barremian or Aptian) of western Portugal. International Journal of Plant Sciences, 161: 51695182.Google Scholar
Frohlich, M. W. & Parker, D. S. (2000). The mostly male theory of flower evolutionary origins: from genes to fossils. Systematic Botany, 25: 155170.Google Scholar
Fruciano, C., Franchini, P. & Meyer, A. (2013). Resampling based approaches to study variation in morphological modularity. PLoS ONE, 8: e69376.Google Scholar
Fujikura, U., Horiguchi, G., Ponce, M. R., Micol, J. L. & Tsukaya, H. (2009). Coordination of cell proliferation and cell expansion mediated by ribosomerelated processes in the leaves of Arabidopsis thaliana. The Plant Journal, 59: 499508.Google Scholar
Fujinami, R., Ghogue, J. P. & Imaichi, R. (2013). Developmental morphology of the controversial ramulus organ of Tristicha trifaria (subfamily Tristichoideae, Podostemaceae): implications for evolution of a unique body plan in Podostemaceae. International Journal of Plant Sciences, 174: 609618.Google Scholar
Fujinami, R. & Imaichi, R. (2015). Developmental morphology of flattened shoots in Dalzellia ubonensis and Indodalzellia gracilis with implications for the evolution of diversified shoot morphologies in the subfamily Tristichoideae (Podostemaceae). American Journal of Botany, 102: 848859.Google Scholar
Fukuda, H. (2000). Programmed cell death of tracheary elements as a paradigm in plants. Plant Molecular Biology, 44: 245253.Google Scholar
Fukuda, T., Yokoyama, J. & Maki, M. (2003). Molecular evolution of cycloidea-like genes in Fabaceae. Journal of Molecular Evolution, 57: 588597.Google Scholar
Fukuda, Y. (1988). Phyllotaxis in two species of Rubia, R. akane and R. sikkimensis. Botanical Magazine (Tokyo), 101: 2538.Google Scholar
Fukuhara, T., Nagmasu, H. & Okada, H. (2003). Floral vasculature, sporogenesis and gametophyte development in Pentastemona egregia (Stemonaceae). Systematics and Geography of Plants, 73: 8390.Google Scholar
Fukushima, K., Fujita, H., Yamaguchi, T. et al. (2015). Oriented cell division shapes carnivorous pitcher leaves of Sarracenia purpurea. Nature Communications, 6: 6450.Google Scholar
Fukushima, K. & Hasebe, M. (2014). Adaxial–abaxial polarity: the developmental basis of leaf shape diversity. Genesis, 52: 118.Google Scholar
Furumizu, C., Alvarez, J. P., Sakakibara, K. & Bowman, J. L. (2015). Antagonistic roles for KNOX1 and KNOX2 genes in patterning the land plant body plan following an ancient gene duplication. PLoS Genetics, 11: e1004980.Google Scholar
Furutani, I., Watanabe, Y., Prieto, R. et al. (2000). The SPIRAL genes are required for directional control of cell elongation in Arabidopsis thaliana. Development, 127: 44434453.Google Scholar
Fusco, G. & Minelli, A. (2010). Phenotypic plasticity in development and evolution. Philosophical Transactions of the Royal Society B, 365: 547556.Google Scholar
Galego, L. & Almeida, J. (2002). Role of DIVARICATA in the control of dorsoventral asymmetry in Antirrhinum flowers. Genes and Development, 16: 880891.Google Scholar
Galimba, K. D. & Di Stilio, V. S. (2015). Sub-functionalization to ovule development following duplication of a floral organ identity gene. Developmental Biology, 405: 158172.Google Scholar
Galis, F. (1999). Why do almost all mammals have seven cervical vertebrae? Developmental constraints, Hox genes, and cancer. Journal of Experimental Zoology (Molecular and Developmental Evolution), 285: 1926.Google Scholar
Galis, F., Van Dooren, T. J., Feuth, J. D. et al. (2006). Extreme selection in humans against homeotic transformations of cervical vertebrae. Evolution, 60: 26432654.Google Scholar
Gallois, J. L., Woodward, C., Reddy, G. V. & Sablowski, R. (2002). Combined SHOOT MERISTEMLESS and WUSCHEL trigger ectopic organogenesis in Arabidopsis. Development, 129: 32073217.Google Scholar
Gao, J. Y., Ren, P. Y., Yang, Z. H. & Li, Q. J. (2006). The pollination ecology of Paraboea rufescens (Gesneriaceae): a buzz-pollinated tropical herb with mirror-image flowers. Annals of Botany, 97: 371376.Google Scholar
Gao, Q., Tao, J.-H., Wang, Y.-Z. & Li, Z.-H. (2008). Expression differentiation of CYC-like floral symmetry genes correlated with their protein sequence divergence in Chirita heterotricha (Gesneriaceae). Development, Genes and Evolution, 218: 341351.Google Scholar
Gao, X., Liang, W., Yin, C. et al. (2010). The SEPALLATA-like gene OsMADS34 is required for rice inflorescence and spikelet development. Plant Physiology, 153: 728740.Google Scholar
García, M. B. & Antor, R. J. (1995a). Age and size structure in populations of a long-lived dioecious geophyte: Borderea pyrenaica (Dioscoreaceae). International Journal of Plant Sciences, 156: 236243.Google Scholar
García, M. B. & Antor, R. J. (1995b). Sex-ratio and sexual dimorphism in the dioecious Borderea pyrenaica (Dioscoreaceae). Oecologia, 101: 5967.Google Scholar
García, M. B., Dahlgren, J. P. & Ehrlén, J. (2011). No evidence of senescence in a 300-year-old mountain herb. Journal of Ecology, 99: 14241430.Google Scholar
Garnock-Jones, P. J. & Johnson, P. N. (1987). Iti lacustris (Brassicaceae), a new genus and species from southern New Zealand. New Zealand Journal of Botany, 25: 603610.Google Scholar
Gautheret, R. J. (1934). Culture du tissus cambial. Comptes rendus hebdomadaires des séances de l’Académie des sciences, 198: 21952196.Google Scholar
Geeta, R., Davalos, L. M., Levy, A. et al. (2012). Keeping it simple: flowering plants tend to retain, and revert to, simple leaves. New Phytologist, 193: 481493.Google Scholar
Gehring, W. J., Affolter, M. & Bürglin, T. R. (1994). Homeodomain proteins. Annual Review of Biochemistry, 63: 487526.Google Scholar
Gendron, J. M., Liu, J.-S., Fan, M. et al. (2012). Brassinosteroids regulate organ boundary formation in the shoot apical meristem of Arabidopsis. Proceedings of the National Academy of Sciences of the United States of America, 109: 2115221157.Google Scholar
Gensel, P. G., Kotyk, M. E. & Basinger, J. E. (2001). Morphology of above- and below-ground structures in early Devonian (Pragian-Emsian) plants. In Plants Invade the Land: Evolutionary and Environmental Perspectives, eds. Gensel, P. G. & Edwards, D.. New York, NY: Columbia University Press, pp. 83102.Google Scholar
Gerats, T. & Vandenbussche, M. (2005). A model system for comparative research: Petunia. Trends in Plant Science, 10: 251256.Google Scholar
Gerrath, J. M. & Lacroix, C. R. (1997). Heteroblastic sequence and leaf development in Leea guineensis. International Journal of Plant Sciences, 158: 747756.Google Scholar
Gerrath, J. M., Posluszny, U. & Dengler, N. G. (2001). Primary vascular patterns in the Vitaceae. International Journal of Plant Sciences, 162: 729745.Google Scholar
Geuten, K., Becker, A., Kaufmann, K. et al. (2006). Petaloidy and petal identity MADS-box genes in the balsaminoid genera Impatiens and Marcgravia. The Plant Journal, 47: 501518.Google Scholar
Geuten, K. & Irish, V. (2010). Hidden variability of floral homeotic B genes in Solanaceae provides a molecular basis for the evolution of novel functions. Plant Cell, 22: 25622578.Google Scholar
Gibson, G. & Dworkin, I. (2004). Uncovering cryptic genetic variation. Nature Reviews Genetics, 5: 11991212.Google Scholar
Gilbert, S. F. & Bolker, J. A. (2001). Homologies of process and modular elements of embryonic construction. Journal of Experimental Zoology (Molecular and Developmental Evolution), 291: 112.Google Scholar
Gill, N., Buti, M., Kane, N. et al. (2014). Sequence-based analysis of structural organization and composition of the cultivated sunflower (Helianthus annuus L.) genome. Biology, 3: 295319.Google Scholar
Gillies, A. C. M., Cubas, P., Coen, E. S. & Abbott, R. J. (2002). Making rays in the Asteraceae: genetics and evolution of variation for radiate versus discoid flower heads. In Developmental Genetics and Plant Evolution, eds. Cronk, Q. C. B., Bateman, R. M. & Hawkins, J. A.. London: Taylor & Francis, pp. 237246.Google Scholar
Giménez, E., Pineda, B., Capel, J. et al. (2010). Functional analysis of the Arlequin mutant corroborates the essential role of the Arlequin/TAGL1 gene during reproductive development of tomato. PLoS ONE, 5: e14427.Google Scholar
Gissi, C., Hastings, K. E. M., Gasparini, F. et al. (2017). An unprecedented taxonomic revision of a model organism: the paradigmatic case of Ciona robusta and Ciona intestinalis. Zoologica Scripta, 46: 521522.Google Scholar
Givnish, T. J. (2010). Giant lobelias exemplify convergent evolution. BMC Biology, 8: 3.Google Scholar
Gleissberg, S. (1998a). Comparative analysis of leaf shape development in Papaveraceae-Papaveroideae. Flora, 193: 269301.Google Scholar
Gleissberg, S. (1998b). Comparative analysis of leaf shape development in Papaveraceae–Chelidonioideae. Flora, 193: 387409.Google Scholar
Gleissberg, S., Groot, E. P., Schmalz, M. et al. (2005). Developmental events leading to peltate leaf structure in Tropaeolum majus (Tropaeolaceae) are associated with expression domain changes of a YABBY gene. Development Genes and Evolution, 215: 313319.Google Scholar
Gleissberg, S. & Kadereit, J. W. (1999). Evolution of leaf morphogenesis: evidence from developmental and phylogenetic data in Papaveraceae. International Journal of Plant Sciences, 160: 787794.Google Scholar
Godfrey-Smith, P. (2009). Darwinian Populations and Natural Selection. New York, NY: Oxford University Press.Google Scholar
Goebel, K. (1900–1905). Organography of Plants. English translation by Balfour, I.. Parts 1 and 2. Oxford: Clarendon Press.Google Scholar
Goebel, K. (1920). Die Entfaltungsbewegungen der Pflanzen und deren teleologische Deutung. Jena: Fischer.Google Scholar
Goebel, K. (1924). Die Entfaltungsbewegungen der Pflanzen und deren teleologische Deutung, 2nd edn. Jena: Fischer.Google Scholar
Goebel, K. (1928). Organographie der Pflanzen. 1. Allgemeine Organographie, 3rd edn. Jena: Fischer.Google Scholar
Goebel, K. (1930). Blütenbildung und Sprossgestaltung (Anthokladien und Infloreszenzen). Jena: G. Fischer.Google Scholar
Goebel, K. (1933). Organographie der Pflanzen III. Samenpflanzen. Jena: G. Fischer.Google Scholar
Goethe, J. W. (1790). Versuch die Metamorphose der Pflanzen zu erklären. Gotha: Ettinger.Google Scholar
Goff, S. A., Ricke, D., Lan, T. H. et al. (2002). A draft sequence of the rice genome (Oryza sativa L. ssp. japonica). Science, 296: 92100.Google Scholar
Goldschmidt, R. (1940). The Material Basis of Evolution. New Haven, CT: Yale University Press.Google Scholar
Golz, J. F., Roccaro, M., Kuzoff, R. & Hudson, A. (2004). GRAMINIFOLIA promotes growth and polarity of Antirrhinum leaves. Development, 131: 36613670.Google Scholar
González, F. & Bello, M. A. (2009). Intra-individual variation of flowers in Gunnera subgenus Panke (Gunneraceae) and proposed apomorphies for Gunnerales. Botanical Journal of the Linnean Society, 160: 262283.Google Scholar
Gonzalez, N., Vanhaeren, H. & Inzé, D. (2012). Leaf size control: complex coordination of cell division and expansion. Trends in Plant Science, 17: 332340.Google Scholar
Gooh, K., Ueda, M., Aruga, K. et al. (2015). Live-cell imaging and optical manipulation of Arabidopsis early embryogenesis. Developmental Cell, 34: 242251.Google Scholar
Goto, K. & Meyerowitz, E. M. (1994). Function and regulation of the Arabidopsis floral homeotic gene PISTILLATA. Genes and Development, 8: 15481560.Google Scholar
Gould, K. S. (1993). Leaf heteroblasty in Pseudopanax crassifolius: functional significance of leaf morphology and anatomy. Annals of Botany, 71: 6170.Google Scholar
Gould, S. J. (1977). Ontogeny and Phylogeny. Cambridge, MA: Harvard University Press.Google Scholar
Gould, S. J. (1988). The uses of heterochrony. In Heterochrony in Evolution: A Multidisciplinary Approach, ed. McKinney, M. L.. New York, NY: Plenum Press, pp. 113.Google Scholar
Gourlay, C. W., Hofer, J. M. I. & Ellis, T. H. N. (2000). Pea compound leaf architecture is regulated by interactions among the genes UNIFOLIATA, COCHLEATA, AFILA, and TENDRIL-LESS. Plant Cell, 12: 12791294.Google Scholar
Govindan, B., Johnson, A. J., Nair, S. N. A. et al. (2016). Nutritional properties of the largest bamboo fruit Melocanna baccifera and its ecological significance. Scientific Reports, 6: 26135.Google Scholar
Graham, L. E., Cook, M. E. & Busse, J. S. (2000). The origin of plants: body plan changes contributing to a major evolutionary radiation. Proceedings of the National Academy of Sciences of the United States of America, 97: 45354540.Google Scholar
Granado-Yela, C., Balaguer, L., Cayuela, L. & Méndez, M. (2017). Unusual positional effects on flower sex in an andromonoecious tree: resource competition, architectural constraints, or inhibition by the apical flower? American Journal of Botany, 104: 608615.Google Scholar
Greb, T., Clarenz, O., Schäfer, E. et al. (2003). Molecular analysis of the LATERAL SUPPRESSOR gene in Arabidopsis reveals a conserved control mechanism for axillary meristem formation. Genes and Development, 17: 11751187.Google Scholar
Greb, T. & Lohmann, J. U. (2016). Plant stem cells. Current Biology, 26: R816R821.Google Scholar
Greenwood, M. S. (1995). Juvenility and maturation in conifers: current concepts. Tree Physiology, 15: 433438.Google Scholar
Greilhuber, J., Borsch, T., Müller, K. et al. (2006). Smallest angiosperm genomes found in Lentibulariaceae with chromosomes of bacterial size. Plant Biology, 8: 770777.Google Scholar
Grew, N. (1682). The Anatomy of Plants. London: Rawlings.Google Scholar
Griesemer, J. (2014). Reproduction and scaffolded developmental processes: an integrated evolutionary perspective. In Towards a Theory of Development, eds. Minelli, A. & Pradeu, T.. Oxford: Oxford University Press, pp. 183202.Google Scholar
Griffith, M. E., da Silva Conceição, A. & Smyth, D. R. (1999). PETAL LOSS gene regulates initiation and orientation of second whorl organs in the Arabidopsis flower. Development, 126: 56355644.Google Scholar
Grimes, J. (1999). Inflorescence morphology, heterochrony, and phylogeny in the Mimosoid tribes Ingeae and Acacieae (Leguminosae: Mimosoideae). Botanical Review, 65: 317347.Google Scholar
Grob, V., Moline, P., Pfeifer, E., Novelo, A. R. & Rutishauser, R. (2006). Developmental morphology of branching flowers in Nymphaea prolifera. Journal of Plant Research, 119: 561570.Google Scholar
Groover, A., DeWitt, N., Heidel, A. & Jones, A. (1997). Programmed cell death of tracheary elements differentiating in vitro. Protoplasma, 196: 197211.Google Scholar
Gu, Q., Ferrándiz, C., Yanofsky, M. F. & Martienssen, R. (1998). The FRUITFULL MADS-box gene mediates cell differentiation during Arabidopsis fruit development. Development, 125: 15091517.Google Scholar
Guédès, M. (1979). Morphology of Seed-Plants. Vaduz: Cramer.Google Scholar
Guerrant, E. O. (1982). Neotenic evolution of Delphinium nudicaule (Ranunculaceae): a hummingbird-pollinated larkspur. Evolution, 36: 699712.Google Scholar
Guha, S. & Maheshwari, S. C. (1964). In vitro production of embryos from anthers of Datura. Nature, 204: 497.Google Scholar
Gunawardena, A. H. L. A. N. & Dengler, N. (2006). Alternative modes of leaf dissection in monocotyledons. Botanical Journal of the Linnean Society, 150: 2544.Google Scholar
Gunawardena, A. H. L. A. N., Greenwood, J. S. & Dengler, N. G. (2004). Programmed cell death remodels lace plant leaf shape during leaf development. Plant Cell, 16: 6073.Google Scholar
Guo, H. S., Xie, Q., Fei, J. F. & Chua, N. H. (2005). MicroRNA directs mRNA cleavage of the transcription factor NAC1 to downregulate auxin signals for Arabidopsis lateral root development. Plant Cell, 17: 13761386.Google Scholar
Haeckel, E. (1866). Generelle Morphologie der Organismen. Allgemeine Grundzüge der organischen Formen-Wissenschaft, mechanisch begründet durch die von Charles Darwin reformirte Descendenz-Theorie, vol. 1: Allgemeine Anatomie der Organismen. Berlin: Reimer.Google Scholar
Hagemann, W. & Gleissberg, W. (1996). Organogenetic capacity of leaves: the significance of marginal blastozones in angiosperms. Plant Systematics and Evolution, 199: 121152.Google Scholar
Hake, S., Smith, H. M. S., Holtan, H. et al. (2004). The role of KNOX genes in plant development. Annual Review of Cell and Developmental Biology, 20: 125151.Google Scholar
Hall, B. K. (1992). Evolutionary Developmental Biology. London: Chapman & Hall.Google Scholar
Hall, B. K. (1998). Evolutionary Developmental Biology, 2nd edn. London: Chapman and Hall.Google Scholar
Hall, B. K. (2005). Consideration of the neural crest and its skeletal derivatives in the context of novelty/innovation. Journal of Experimental Zoology (Molecular and Developmental Evolution), 304B: 548557.Google Scholar
Hall, J. C., Tisdale, T. E., Donohue, K. & Kramer, E. M. (2006). Developmental basis of an anatomical novelty: heteroarthrocarpy in Cakile lanceolata and Erucaria erucarioides (Brassicaceae). International Journal of Plant Sciences, 167: 771789.Google Scholar
Hallgrímsson, B., Jamniczky, H., Young, N. M. et al. (2009). Deciphering the palimpsest: studying the relationship between morphological integration and phenotypic covariation. Evolutionary Biology, 36: 355376.Google Scholar
Hallgrímsson, B., Jamniczky, H., Young, N. M. et al. (2012). The generation of variation and the developmental basis for evolutionary novelty. Journal of Experimental Zoology (Molecular and Developmental Evolution), 318B: 501517.Google Scholar
Hamada, S., Onouchi, H., Tanaka, H. et al. (2000). Mutations in the WUSCHEL gene of Arabidopsis thaliana result in the development of shoots without juvenile leaves. The Plant Journal, 24: 91101.Google Scholar
Hamant, O., Heisler, M. G., Jönsson, H. et al. (2008). Developmental patterning by mechanical signals in Arabidopsis. Science, 322: 16501655.Google Scholar
Han, P., García-Ponce, B., Fonseca-Salazar, G., Alvarez-Buylla, E. R. & Yu, H. (2008). AGAMOUS-LIKE 17, a novel flowering promoter, acts in a FT independent photoperiod pathway. The Plant Journal, 55: 253265.Google Scholar
Hannan, G. L. (1980). Heteromericarpy and dual seed germination modes in Platystemon californicus (Papaveraceae). Madroño, 27: 164170.Google Scholar
Hansen, T. F. (2003). Is modularity necessary for evolvability? Remarks on the relationship between pleiotropy and evolvability. BioSystems, 69: 8394.Google Scholar
Hansen, T. F. (2006). The evolution of genetic architecture. Annual Review of Ecology and Systematics, 37: 123157.Google Scholar
Hareven, D., Gutfinger, T., Parnis, A. et al. (1996). The making of a compound leaf: genetic manipulation of leaf architecture in tomato. Cell, 84: 735744.Google Scholar
Harling, G., Wilder, G. J. & Eriksson, R. (1998). Cyclanthaceae. In The Families and Genera of Vascular Plants, Vol. 3, ed. Kubitzki, K.. Berlin: Springer, pp. 202215.Google Scholar
Harper, J. L. & White, J. (1974). The demography of plants. Annual Review of Ecology and Systematics, 5: 419463.Google Scholar
Harris, E. M. (1991). Floral initiation and early development in Erigeron philadelphicus (Asteraceae). American Journal of Botany, 78: 108121.Google Scholar
Harris, E. M. (1995). Inflorescence and floral ontogeny in Asteraceae: a synthesis of historical and current concepts. Botanical Review, 61: 3278.Google Scholar
Harris, M. A., Lock, A., Bühler, J., Oliver, S. G. & Wood, V. (2013). FYPO: the fission yeast phenotype ontology. Bioinforma, 29: 16711678.Google Scholar
Harrison, J., Möller, M., Langdale, J., Cronk, Q. C. B. & Hudson, A. (2005). The role of KNOX genes in the evolution of morphological novelty in Streptocarpus. Plant Cell, 17: 430443.Google Scholar
Harrison, J. C. (2017). Development and genetics in the evolution of land plant body plans. Philosophical Transactions of the Royal Society B, 372: 20150490.Google Scholar
Hartmann, U., Höhmann, S., Nettesheim, K. et al. (2000). Molecular cloning of SVP: a negative regulator of the floral transition in Arabidopsis. The Plant Journal, 21: 351360.Google Scholar
Hashimoto, T. (2002). Molecular genetic analysis of left–right handedness in plants. Philosophical Transactions of the Royal Society B, 357: 799808.Google Scholar
Hasson, A., Blein, T. & Laufs, P. (2010). Leaving the meristem behind: the genetic and molecular control of leaf patterning and morphogenesis. Comptes Rendus Biologies, 333: 350360.Google Scholar
Hawkins, J. A. (2002). Evolutionary developmental biology: impact on systematic theory and practice, and the contribution of systematics. In Developmental Genetics and Plant Evolution, eds. Cronk, Q. C. B., Bateman, R. M. & Hawkins, J. A.. London: Taylor & Francis, pp. 3251.Google Scholar
Hay, A., Barkoulas, M. & Tsiantis, M. (2006). ASYMMETRIC LEAVES1 and auxin activities converge to repress BREVIPEDICELLUS expression and promote leaf development in Arabidopsis. Development, 133, 39553961.Google Scholar
Hay, A., Jackson, D., Ori, N. & Hake, S. (2003). Analysis of the competence to respond to KNOTTED1 activity in Arabidopsis leaves using a steroid induction system. Plant Physiology, 131: 16711680.Google Scholar
Hay, A. & Tsiantis, M. (2006) The genetic basis for differences in leaf form between Arabidopsis thaliana and its wild relative Cardamine hirsuta. Nature Genetics, 38: 942947.Google Scholar
Hay, A. & Tsiantis, M. (2010). KNOX genes: versatile regulators of plant development and diversity. Development, 137: 31533165.Google Scholar
Hay, A. S., Pieper, B., Cooke, E. et al. (2014). Cardamine hirsuta: a versatile genetic system for comparative studies. The Plant Journal, 78: 115.Google Scholar
He, C. Y. & Saedler, H. (2005). Heterotopic expression of MPF2 is the key to the evolution of the Chinese lantern of Physalis, a morphological novelty in Solanaceae. Proceedings of the National Academy of Sciences of the United States of America, 102: 57795784.Google Scholar
He, Y., Doyle, M. R. & Amasino, R. M. (2004). PAF1-complex-mediated histone methylation of FLOWERING LOCUS C chromatin is required for the vernalization-responsive, winter-annual habit in Arabidopsis. Genes and Development, 18 : 27742784.Google Scholar
Hecht, V., Foucher, F., Ferrándiz, C. et al. (2005). Conservation of Arabidopsis flowering genes in model legumes. Plant Physiology, 137: 14201434.Google Scholar
Heenan, P. B. (2002). Cardamine lacustris, a new combination for Iti lacustris (Brassicaceae). New Zealand Journal of Botany, 40: 563569.Google Scholar
Heenan, P. B., Molloy, P. B. J. & Smissen, R. D. (2013). Cardamine cubita (Brassicaceae), a new species from New Zealand with a remarkable reduction in floral parts. Phytotaxa, 140: 4350.Google Scholar
Heide-Jorgensen, H. S. (2013). Introduction: the parasitic syndrome in higher plants. In Parasitic Orobanchaceae, eds. Joel, D. M., Gressel, J. & Musselman, L. J.. Heidelberg: Springer, pp. 118.Google Scholar
Held, L. I. Jr. (2009). Quirks of Human Anatomy: An Evo-Devo Look at the Human Body. Cambridge: Cambridge University Press.Google Scholar
Hemerly, A., Engler, J. de A., Bergounioux, C. et al. (1995). Dominant negative mutants of the Cdc2 kinase uncouple cell division from iterative plant development. EMBO Journal, 14: 39253936.Google Scholar
Hempel, F. D., Weigel, D., Mandel, M. A. et al. (1997). Floral determination and expression of floral regulatory genes in Arabidopsis. Development, 124: 38453853.Google Scholar
Hendrikse, J. L., Parsons, T. E. & Hallgrímsson, B. (2007). Evolvability as the proper focus of evolutionary developmental biology. Evolution and Development, 9: 393401.Google Scholar
Hennig, W. (1966). Phylogenetic Systematics. Urbana, IL: University of Illinois Press.Google Scholar
Henschel, K., Kofuji, R., Hasebe, M. et al. (2002). Two ancient classes of MIKC-type MADS-box genes are present in the moss Physcomitrella patens. Molecular Biology and Evolution, 19: 801814.Google Scholar
Heyn, C. C., Dagan, O. & Nachman, B. (1974). The annual Calendula species, taxonomy and relationships. Israel Journal of Botany, 23: 169201.Google Scholar
Hibara, K., Karim, M. R., Takada, S. et al. (2006). Arabidopsis CUP-SHAPED COTYLEDON3 regulates postembryonic shoot meristem and organ boundary formation. Plant Cell, 18: 29462957.Google Scholar
Hileman, L. C. (2012). Flowers. In Encyclopedia of Life Sciences. Chichester: Wiley.Google Scholar
Hileman, L. C. (2014a). Trends in flower symmetry evolution revealed through phylogenetic and developmental genetic advances. Philosophical Transactions of the Royal Society B, 369: 20130348.Google Scholar
Hileman, L. C. (2014b). Bilateral flower symmetry: how, when and why? Current Opinion in Plant Biology, 17: 146152.Google Scholar
Hileman, L. C. & Baum, D. A. (2003). Why do paralogs persist? Molecular evolution of CYCLOIDEA and related floral symmetry genes in Antirrhineae (Veronicaceae). Molecular Biology and Evolution, 20: 591600.Google Scholar
Hileman, L. C. & Irish, V. F. (2009). More is better: the uses of developmental genetic data to reconstruct perianth evolution. American Journal of Botany, 96: 8395.Google Scholar
Hileman, L. C., Kramer, E. M. & Baum, D. A. (2003). Differential regulation of symmetry genes and the evolution of floral morphologies. Proceedings of the National Academy of Sciences of the United States of America, 100: 1281412819.Google Scholar
Hileman, L. C., Sundstrom, J. F., Litt, A. et al. (2006). Molecular and phylogenetic analyses of the MADS-box gene family in tomato. Molecular Biology and Evolution, 23: 22452258.Google Scholar
Hill, J. P. & Lord, E. M. (1989). Floral development in Arabidopsis thaliana: comparison of the wild type and the homeotic pistillata mutant. Canadian Journal of Botany, 67: 29222936.Google Scholar
Hillson, C. J. (1979). Leaf development in Senecio rowleyanus (Compositae). American Journal of Botany, 66: 5963.Google Scholar
Hintz, M., Bartholmes, C., Nutt, P. et al. (2006). Catching a ‘hopeful monster’: shepherd’s purse (Capsella bursa-pastoris) as a model system to study the evolution of flower development. Journal of Experimental Botany, 57: 35313542.Google Scholar
Hnatiuk, R. J. (1977). Population structure of Livistona eastonii Gardn., Mitchell Plateau, Western Australia. Australian Journal of Ecology, 2: 461466.Google Scholar
Hódar, J. A. (2002). Leaf fluctuating asymmetry of holm oak in response to drought under contrasting climatic conditions. Journal of Arid Environments, 52: 233243.Google Scholar
Hodges, S. A. (1997a). Rapid radiation due to a key innovation in columbines (Ranunculaceae: Aquilegia). In Molecular Evolution and Adaptive Radiation, eds. Givnish, T. J. & Sytsma, K. J.. Cambridge: Cambridge University Press, pp. 391405.Google Scholar
Hodges, S. A. (1997b). Floral nectar spurs and diversification. International Journal of Plant Sciences, 158: S81S88.Google Scholar
Hoehndorf, R., Ngonga Ngomo, A.-C. & Kelso, J. (2010). Applying the functional abnormality ontology pattern to anatomical functions. Journal of Biomedical Semantics, 1: 4.Google Scholar
Hofer, J., Gourlay, C., Michael, A. & Ellis, T. H. N. (2001a). Expression of a class-1 knottedl-like homeobox gene is downregulated in pea compound-leaf primordia. Plant Molecular Biology, 45: 387398.Google Scholar
Hofer, J., Turner, L., Hellens, R. et al. (1997). UNIFOLIATA regulates leaf and flower morphogenesis in pea. Current Biology, 7: 581587.Google Scholar
Hofer, J., Turner, L., Moreau, C. et al. (2009). Tendril-less regulates tendril formation in pea leaves. Plant Cell, 21: 420428.Google Scholar
Hofer, J. M. I., Gourlay, C. W. & Ellis, T. H. N. (2001b). Genetic control of leaf morphology: a partial view. Annals of Botany, 88: 11291139.Google Scholar
Hofmeister, W. (1868). Allgemeine Morphologie der Gewächse. In Handbuch der physiologischen Botanik, vol. 1(2), ed. Hofmeister, W.. Leipzig: Engelmann, pp. 405664.Google Scholar
Hohmann, N., Wolf, E. M., Lysak, M. A. & Koch, M. A. (2015). A time-calibrated road map of Brassicaceae species radiation and evolutionary history. Plant Cell, 27: 27702784.Google Scholar
Hong, L., Dumond, M., Tsugawa, S. et al. (2016). Variable cell growth yields reproducible organ development through spatiotemporal averaging. Developmental Cell, 38: 1532.Google Scholar
Honma, T. & Goto, K. (2001). Complexes of MADS-box proteins are sufficient to convert leaves into floral organs. Nature, 409: 525529.Google Scholar
Horst, N. A., Katz, A., Pereman, I. et al. (2016). A single homeobox gene triggers phase transition, embryogenesis and asexual reproduction. Nature Plants, 2: 15209.Google Scholar
Horst, N. H. & Reski, R. (2016). Alternation of generations: unravelling the underlying molecular mechanism of a 165-year-old botanical observation. Plant Biology, 18: 549551.Google Scholar
Howarth, D. G. & Donoghue, M. J. (2006). Phylogenetic analysis of the ‘ECE’ (CYC/TB1) clade reveals duplications predating the core eudicots. Proceedings of the National Academy of Sciences of the United States of America, 103: 91019106.Google Scholar
Howarth, D. G., Martins, T., Chimney, E. & Donoghue, M. J. (2011). Diversification of CYCLOIDEA expression in the evolution of bilateral flower symmetry in Caprifoliaceae and Lonicera (Dipsacales). Annals of Botany, 107: 15211532.Google Scholar
Hu, T. T., Pattyn, P., Bakker, E. G. et al. (2011). The Arabidopsis lyrata genome sequence and the basis of rapid genome size change. Nature Genetics, 43: 476481.Google Scholar
Hu, Y., Xie, Q. & Chua, N. H. (2003). The Arabidopsis auxin-inducible gene ARGOS controls lateral organ size. Plant Cell, 15: 19511961.Google Scholar
Huang, L.-J., Wang, X.-W. & Wang, X.-F. (2014). The structure and development of incompletely closed carpels in an apocarpous species, Sagittaria trifolia (Alismataceae). American Journal of Botany, 101: 12291234.Google Scholar
Huang, T. & Irish, V. F. (2016). Gene networks controlling petal organogenesis. Journal of Experimental Botany, 67: 6168.Google Scholar
Hubbard, L., McSteen, P., Doebley, J. & Hake, S. (2002). Expression patterns and mutant phenotype of teosinte branched1 correlate with growth suppression in maize and teosinte. Genetics, 162: 19271935.Google Scholar
Huber, H. (1998). Dioscoreaceae. In The Families and Genera of Vascular Plants, Vol. 3, ed. Kubitzki, K.. Berlin: Springer, pp. 216235.Google Scholar
Hudson, A. & Jeffree, C. (2001). Leaf and internode. In Encyclopedia of Life Sciences. Chichester: Wiley.Google Scholar
Hudson, C. J., Freeman, J. S., Jones, R. C. et al. (2014). Genetic control of heterochrony in Eucalyptus globulus. G3 (Bethesda), 4: 12351245.Google Scholar
Huether, C. A. (1968). Exposure of natural genetic variability underlying the pentamerous corolla constancy in Linanthus androsaceus ssp. androsaceus. Genetics, 60: 123146.Google Scholar
Huether, C. A. (1969). Constancy of the pentamerous corolla phenotype in natural populations of Linanthus. Evolution, 23: 572588.Google Scholar
Huijser, P., Klein, J., Lönnig, W. E. et al. (1992). Bracteomania, an inflorescence anomaly, is caused by the loss of function of the MADS-box gene Squamosa in Antirrhinum majus. EMBO Journal, 11: 12391249.Google Scholar
Huxley, J. S. (1942). Evolution: The Modern Synthesis. London: Allen and Unwin.Google Scholar
Iannelli, F., Pesole, G., Sordino, P. & Gissi, C. (2007). Mitogenomics reveals two cryptic species in Ciona intestinalis. Trends in Genetics, 23: 419422.Google Scholar
Ichihashi, Y., Kawade, K., Usami, T. et al. (2011). Key proliferative activity in the junction between the leaf blade and leaf petiole of Arabidopsis. Plant Physiology, 157: 11511162.Google Scholar
Igersheim, A., Buzgo, M. & Endress, P. K. (2001). Gynoecium diversity and systematics in basal monocots. Botanical Journal of the Linnean Society, 136: 165.Google Scholar
Igersheim, A. & Endress, P. K. (1998). Gynoecium diversity and systematics of the paleoherbs. Botanical Journal of the Linnean Society, 127: 289370.Google Scholar
Ikeuchi, M., Ogawa, Y., Iwase, A. & Sugimoto, K. (2016). Plant regeneration: cellular origins and molecular mechanisms. Development, 143: 14421451.Google Scholar
Ikezaki, M., Kojima, M., Sakakibara, H. et al. (2010). Genetic networks regulated by ASYMMETRIC LEAVES1 (AS1) and AS2 in leaf development in Arabidopsis thaliana: KNOX genes control five morphological events. The Plant Journal, 61: 7082.Google Scholar
Imaichi, R., Hiyama, Y. & Kato, M. (2004). Developmental morphology of foliose shoots and seedlings of Dalzellia zeylanica (Podostemaceae) with special reference to their meristems. Botanical Journal of the Linnean Society, 144: 289302.Google Scholar
Imaichi, R., Hiyama, Y. & Kato, M. (2005). Leaf development in the absence of a shoot apical meristem in Zeylanidium subulatum (Podostemaceae). Annals of Botany, 96: 5158.Google Scholar
Imaichi, R., Nagumo, S. & Kato, M. (2000). Ontogenic anatomy of Streptocarpus grandis (Gesneriaceae) with implications for evolution of monophyly. Annals of Botany, 86: 3746.Google Scholar
Imbert, E. 2002. Ecological consequences and ontogeny of seed heteromorphism. Perspectives in Plant Ecology, Evolution and Systematics, 5: 1336.Google Scholar
Immink, R. G. H., Ferrario, S., Busscher-Lange, J. et al. (2003). Analysis of the petunia MADS-box transcription factor family. Molecular Genetics and Genomics, 268: 598606.Google Scholar
Immink, R. G. H., Hannapel, D. J., Ferrario, S. et al. (1999). A petunia MADS box gene involved in the transition from vegetative to reproductive development. Development, 126: 51175126.Google Scholar
Immink, R. G. H., Kaufmann, K. & Angenent, G. C. (2010). The ‘ABC’ of MADS domain protein behaviour and interactions. Seminars in Cell and Developmental Biology, 21: 8793.Google Scholar
Immink, R. G. H., Tonaco, I. A. N., de Folter, S. et al. (2009). SEPALLATA3: the ‘glue’ for MADS box transcription factor complex formation. Genome Biology, 10: R24.Google Scholar
Ingram, G. C., Goodrich, J., Wilkinson, M. D. et al. (1995). Parallels between UNUSUAL FLORAL ORGANS and FIMBRIATA, genes controlling flower development in Arabidopsis and Antirrhinum. Plant Cell, 7: 15011510.Google Scholar
Inzé, D. (2003). Why should we study the plant cell cycle? Journal of Experimental Botany, 54: 11251126.Google Scholar
Irish, V. F. (2006). Duplication, diversification, and comparative genetics of angiosperm MADS-box genes. Advances in Botanical Research, 44: 129161.Google Scholar
Irish, V. F. (2009). Evolution of petal identity. Journal of Experimental Botany, 60: 25172527.Google Scholar
Irish, V. F. & Sussex, I. M. (1990). Function of the apetala-1 gene during Arabidopsis floral development. Plant Cell, 2: 741753.Google Scholar
Ishida, T., Aida, M., Takada, S. & Tasaka, M. (2000). Involvement of CUP-SHAPED COTYLEDON genes in gynoecium and ovule development in Arabidopsis thaliana. Plant and Cell Physiology, 41: 6067.Google Scholar
Ishii, H. S. & Harder, L. D. (2012). Phenological associations of within- and among-plant variation in gender with floral morphology and integration in protandrous Delphinium glaucum. Journal of Ecology, 100: 10291038.Google Scholar
Itkin, M., Seybold, H., Breitel, D. et al. (2009). TOMATO AGAMOUS-LIKE 1 is a component of the fruit ripening regulatory network. The Plant Journal, 60: 10811095.Google Scholar
Itoh, J.-I., Hasegawa, A., Kitano, H. & Nagato, Y. (1998). A recessive heterochronic mutation, plastochron1, shortens the plastochron and elongates the vegetative phase in rice. Plant Cell, 10: 15111521.Google Scholar
Iwamoto, A., Shimuzu, A. & Ohba, H. (2003). Floral development and phyllotactic variation in Ceratophyllum demersum (Ceratophyllaceae). American Journal of Botany, 90: 11241130.Google Scholar
Jaakola, L., Poole, M., Jones, M. O. et al. (2010). A SQUAMOSA MADS box gene involved in the regulation of anthocyanin accumulation in bilberry fruits. Plant Physiology, 153: 16191629.Google Scholar
Jabbour, F., Cossard, G., Le Guilloux, M. et al. (2014). Specific duplication and dorsoventrally asymmetric expression patterns of cycloidea-like genes in zygomorphic species of Ranunculaceae. PLoS ONE, 9: e95727.Google Scholar
Jabbour, F., Ronse De Craene, L. P., Nadot, S. & Damerval, C. (2009). Establishment of zygomorphy on an ontogenetic spiral and evolution of perianth in the tribe Delphinieae (Ranunculaceae). Annals of Botany, 104: 809822.Google Scholar
Jabbour, F., Udron, M., Le Guilloux, M. et al. (2015). Flower development schedule and AGAMOUS-like gene expression patterns in two morphs of Nigella damascena (Ranunculaceae) differing in floral architecture. Botanical Journal of the Linnean Society, 178: 608619.Google Scholar
Jablonka, E. (2006). Genes as followers in evolution: a post-synthesis synthesis? Biology and Philosophy, 21: 143154.Google Scholar
Jack, T. (2004). Molecular and genetic mechanisms of floral control. Plant Cell, 16: S1S17.Google Scholar
Jack, T., Brockman, L. L. & Meyerowitz, E. M. (1992). The homeotic gene APETALA3 of Arabidopsis thaliana encodes a MADS box and is expressed in petals and stamens. Cell, 68: 683697.Google Scholar
Jack, T., Fox, G. L. & Meyerowitz, E. M. (1994). Arabidopsis homeotic gene APETALA3 ectopic expression: transcriptional and posttranscriptional regulation determine floral organ identity. Cell, 76: 703716.Google Scholar
Jäger-Zürn, I. (2007). The shoot apex of Podostemaceae: de novo structure or reduction of the conventional type? Flora, 202: 383394.Google Scholar
Jaillon, O., Aury, J. M., Noel, B. et al. (2007). The grapevine genome sequence suggests ancestral hexaploidization in major angiosperm phyla. Nature, 449: 463467.Google Scholar
James, P. J. (2009). ‘Tree and leaf’: a different angle. The Linnean, 25 (1): 1319.Google Scholar
Janssen, B.-J., Lund, L. & Sinha, N. (1998). Overexpression of a homeobox gene, LeT6, reveals indeterminate features in the tomato compound leaf. Plant Physiology, 117: 771786.Google Scholar
Janzen, D. H. (1976). Why bamboos wait so long to flower. Annual Review of Ecology and Systematics, 7: 347391.Google Scholar
Jaramillo, M. A. & Kramer, E. M. (2004). APETALA3 and PISTILLATA homologs exhibit novel expression patterns in the unique perianth of Aristolochia (Aristolochiaceae). Evolution and Development, 6: 449458.Google Scholar
Jaramillo, M. A. & Kramer, E. M. (2007). The role of developmental genetics in understanding homology and morphological evolution in plants. International Journal of Plant Sciences, 168: 6172.Google Scholar
Jasinski, S., Vialette-Guiraud, A. C. & Scutt, C. P. (2010). The evolutionary-developmental analysis of plant microRNAs. Philosophical Transactions of the Royal Society B, 365: 469476.Google Scholar
Jaya, E., Kubien, D. S., Jameson, P. E. & Clemens, J. (2010). Vegetative phase change and photosynthesis in Eucalyptus occidentalis: architectural simplification prolongs juvenile traits. Tree Physiology, 30: 393403.Google Scholar
Jesson, L. K. & Barrett, S. C. H. (2002a). Enantiostyly in Wachendorfia (Haemodoraceae); the influence of reproductive systems on the maintenance of the polymorphism. American Journal of Botany, 89: 253263.Google Scholar
Jesson, L. K. & Barrett, S. C. H. (2002b). The genetics of mirror-image flowers. Proceedings of the Royal Society B, 269: 18351839.Google Scholar
Jesson, L. K. & Barrett, S. C. H. (2003). The comparative biology of mirror-image flowers. International Journal of Plant Sciences, 164: S237S249.Google Scholar
Jeune, B. & Sattler, R. (1992). Multivariate analysis in process morphology of plants. Journal of Theoretical Biology, 156: 147167.Google Scholar
Jeune, B. & Sattler, R. (1996). Quelques aspects d’une morphologie continuiste et dynamique. Canadian Journal of Botany, 74: 10231039.Google Scholar
Jiao, Y., Leebens-Mack, J., Ayyampalayam, S. et al. (2012). A genome triplication associated with early diversification of the core eudicots. Genome Biology, 13: R3.Google Scholar
Jiao, Y. N., Wickett, N. J., Ayyampalayam, S. et al. (2011). Ancestral polyploidy in seed plants and angiosperms. Nature, 473, 97100.Google Scholar
Jibran, R., Hunter, D. A. & Dijkwel, P. P. (2013). Hormonal regulation of leaf senescence through integration of developmental and stress signals. Plant Molecular Biology, 82: 547561.Google Scholar
Jiménez, S., Lawton-Rauh, A. L., Reighard, G. L., Abbott, A. G. & Bielenberg, D. G. (2009). Phylogenetic analysis and molecular evolution of the dormancy associated MADS-box genes from peach. BMC Plant Biology, 9: 81.Google Scholar
Jones, C. S. (1992). Comparative ontogeny of a wild cucurbit and its derived cultivar. Evolution, 46: 18271847.Google Scholar
Jong, K. (1973). Streptocarpus (Gesneriaceae) and the phyllomorph concept. Acta Botanica Neerlandica, 22: 243255.Google Scholar
Jong, K. & Burtt, B. L. (1975). The evolution of morphological novelty exemplified in the growth patterns of some Gesneriaceae. New Phytologist, 75: 297311.Google Scholar
Juarez, M. T., Twigg, R. W. & Timmermans, M. C. (2004). Specification of adaxial cell fate during maize leaf development. Development, 131: 45334544.Google Scholar
Juncosa, A. M. (1988). Floral development and character evolution in Rhizophoraceae. In Aspects of Floral Development, eds. Leins, P., Tucker, S. C. & Endress, P. K.. Berlin: Cramer, pp. 83101.Google Scholar
Juntheikki-Palovaara, I., Tähtiharju, S., Lan, T. et al. (2014). Functional diversification of duplicated CYC2 clade genes in regulation of inflorescence development in Gerbera hybrida (Asteraceae). The Plant Journal, 79: 783796.Google Scholar
Kagale, S., Robinson, S. J., Nixon, J. et al. (2014). Polyploid evolution of the Brassicaceae during the Cenozoic era. Plant Cell, 26: 27772791.Google Scholar
Kalinka, A. T., Varga, K. M., Gerrard, D. T. et al. (2010). Gene expression divergence recapitulates the developmental hourglass model. Nature, 468: 811814.Google Scholar
Kampny, C. M., Dickinson, T. A. & Dengler, N. G. (1993). Quantitative comparison of floral development in Veronica chamaedrys and Veronicastrum virginicum (Scrophulariaceae). American Journal of Botany, 80: 449460.Google Scholar
Kampny, C. M., Dickinson, T. A. & Dengler, N. G. (1994). Quantitative floral development in Pseudolysimachion (Scrophulariaceae): intraspecific variation and comparison with Veronica and Veronicastrum. American Journal of Botany, 81: 13431353.Google Scholar
Kang, H. G., Jeon, J. S., Lee, S. & An, G. H. (1998). Identification of class B and class C floral organ identity genes from rice plants. Plant Molecular Biology, 38: 10211029.Google Scholar
Kanno, A., Saeki, H., Kameya, T., Saedler, H. & Theissen, G. (2003). Heterotopic expression of class B floral homeotic genes supports a modified ABC model for tulip (Tulipa gesneriana). Plant Molecular Biology, 52: 831841.Google Scholar
Kaplan, D. (1975). Comparative developmental evaluation of the morphology of unifacial leaves in the monocotyledons. Botanische Jahrbücher für Systematik, Pflanzengeschichte und Pflanzengeographie, 95: 1105.Google Scholar
Kaplan, D. R. (1970). Comparative foliar histogenesis in Acorus calamus and its bearing on the phyllode theory of monocotyledonous leaves. American Journal of Botany, 57: 331361.Google Scholar
Kaplan, D. R. (1973). The monocotyledons: their evolution and comparative biology. VII. The problem of leaf morphology and evolution in the monocotyledons. Quarterly Review of Biology, 48: 437457.Google Scholar
Kaplan, D. R. (1984). Alternative modes of organogenesis in higher plants. In Contemporary Problems in Plant Anatomy, eds. White, R. A. & Dickison, W. C.. New York, NY: Academic Press, pp. 261300.Google Scholar
Kaplan, D. R. (1992). The relationship of cells to organisms in plants: problem and implications of an organismal perspective. International Journal of Plant Sciences, 153: S28S37.Google Scholar
Kaplan, D. R. (2001). Fundamental concepts of leaf morphology and morphogenesis: a contribution to the interpretation of molecular genetic mutants. International Journal of Plant Sciences, 162: 465474.Google Scholar
Kaplan, D. R., Dengler, N. G. & Dengler, R. E. (1982). The mechanisms of plication inception in palm leaves: histogenetic observations of the palmate leaves of Rhapis excelsa. Canadian Journal of Botany, 60: 29993016.Google Scholar
Kaplan, D. R. & Hagemann, W. (1991). The relationship of cell and organism in vascular plants. Bioscience, 41: 693703.Google Scholar
Karoly, K. & Conner, J. K. (2000). Heritable variation in a family-diagnostic trait. Evolution, 54: 14331438.Google Scholar
Kasha, K. J. & Kao, K. N. (1970). High frequency haploid production in barley (Hordeum vulgare L.). Nature, 225: 874876.Google Scholar
Katayama, N., Kato, M., Nishiuchi, T. & Yamada, T. (2011). Comparative anatomy of embryogenesis in three species of Podostemaceae and evolution of the loss of embryonic shoot and root meristems. Evolution and Development, 13: 333342.Google Scholar
Katayama, N., Kato, M. & Yamada, T. (2013). Origin and development of the cryptic shoot meristem in Zeylanidium lichenoides. American Journal of Botany, 100: 635646.Google Scholar
Katayama, N., Koi, S. & Kato, M. (2010). Expression of SHOOT MERISTEMLESS, WUSCHEL, and ASYMMETRIC LEAVES1 homologs in the shoots of Podostemaceae: implications for the evolution of novel shoot organogenesis. Plant Cell, 22: 21312140.Google Scholar
Kaufmann, K., Pajoro, A. & Angenent, G. C. (2010). Regulation of transcription in plants: mechanisms controlling developmental switches. Nature Reviews Genetics, 11: 830842.Google Scholar
Kawade, K., Horiguchi, G. & Tsukaya, H. (2010). Non-cell-autonomously coordinated organ size regulation in leaf development. Development, 137: 42214227.Google Scholar
Kawamura, E., Horiguchi, G. & Tsukaya, H. (2010). Mechanisms of leaf tooth formation in Arabidopsis. The Plant Journal, 62: 429441.Google Scholar
Kazama, Y., Fujiwara, M. T., Koizumi, A. et al. (2009). A SUPERMAN-like gene is exclusively expressed in female flowers of the dioecious plant Silene latifolia. Plant Cell Physiology, 50: 11271141.Google Scholar
Kellogg, E. A. (2000). A model of inflorescence development. In Monocots: Systematics and Evolution, eds. Wilson, K. L. & Morrison, D. A.. Melbourne: CSIRO, pp. 8488.Google Scholar
Kellogg, E. A. (2001). Evolutionary history of the grasses. Plant Physiology, 125: 11981205.Google Scholar
Kellogg, E. A., Camara, P. E. A. S., Rudall, P. J. et al. (2013). Early inflorescence development in the grasses (Poaceae). Frontiers in Plant Science, 4: 250.Google Scholar
Kerr, A. D. (1972). Ephemeral means ‘don’t turn your head’. American Orchid Society Bulletin, 41: 208211.Google Scholar
Kerstetter, R. A., Bollman, K., Taylor, R. A., Bomblies, K. & Poethig, R. S. (2001). KANADI regulates organ polarity in Arabidopsis. Nature, 411: 706709.Google Scholar
Kerstetter, R. A. & Poethig, R. S. (1998). The specification of leaf identity during shoot development. Annual Review of Cell and Developmental Biology, 14: 373398.Google Scholar
Kessler, S. & Sinha, N. (2004). Shaping up: the genetic control of leaf shape. Current Opinion in Plant Biology, 7: 6572.Google Scholar
Khan, M. R., Hu, J. Y., Riss, S., He, C. & Saedler, H. (2009). MPF2-like-A MADS-box genes control the inflated calyx syndrome in Withania (Solanaceae): roles of Darwinian selection. Molecular Biology and Evolution, 26: 24632473.Google Scholar
Kidner, C. A. & Timmermans, M. C. P. (2010). Signaling sides: adaxial–abaxial patterning in leaves. Current Topics in Developmental Biology, 91: 141168.Google Scholar
Kierzkowski, D., Nakayama, N., Routier-Kierzkowska, A.-L. et al. (2012). Elastic domains regulate growth and organogenesis in the plant shoot apical meristem. Science, 335: 10961099.Google Scholar
Kim, G., LeBlanc, M. L., Wafula, E. K., de Pamphilis, C. W. & Westwood, J. H. (2014). Genomic-scale exchange of mRNA between a parasitic plant and its hosts. Science, 345: 808811.Google Scholar
Kim, K.-J., Choi, K.-S. & Jansen, R. K. (2005a). Two chloroplast DNA inversions originated simultaneously during the early evolution of the sunflower family (Asteraceae). Molecular Biology and Evolution, 22: 17831792.Google Scholar
Kim, M., Cui, M., Cubas, P. et al. (2008). Regulatory genes control a key morphological and ecological trait transferred between species. Science, 322: 11161119.Google Scholar
Kim, M., McCormick, S., Timmermans, M & Sinha, N. (2003). The expression domain of PHANTASTICA determines leaflet placement in compound leaves. Nature, 424: 438443.Google Scholar
Kim, S., Koh, J., Ma, H. et al. (2005b). Sequence and expression studies of A-, B-, and E-class MADS-box homologues in Eupomatia (Eupomatiaceae): support for the bracteate origin of the calyptra. International Journal of Plant Sciences, 166: 185198.Google Scholar
Kim, S., Koh, J., Yoo, M. J. et al. (2005c). Expression of floral MADS-box genes in basal angiosperms: implications for the evolution of floral regulators. The Plant Journal, 43: 724744.Google Scholar
Kim, S., Yoo, M. J., Albert, V. A. et al. (2004). Phylogeny and diversification of B-function MADS-box genes in angiosperms: evolutionary and functional implications of a 260-million-year-old duplication. American Journal of Botany, 91: 21022118.Google Scholar
Kirchoff, B. K. (2000). Hofmeister’s rule and primordium shape: influences on organ position in Hedychium coronarium (Zingiberaceae). In Monocots: Systematics and Evolution, eds. Wilson, K. L. & Morrison, D. A.. Melbourne: CSIRO, pp. 7583.Google Scholar
Kirchoff, B. K. (2001). Character description in phylogenetic analysis: insights from Agnes Arber’s concept of the plant. Annals of Botany, 88: 12031214.Google Scholar
Kirchoff, B. K. (2017). Inflorescence and flower development in Musa velutina H. Wendl. & Drude (Musaceae), with a consideration of developmental variability, restricted phyllotactic direction and hand initiation. International Journal of Plant Sciences, 178: 259272.Google Scholar
Kirkpatrick, M. (2009). Patterns of quantitative genetic variation in multiple dimensions. Genetica, 136: 271284.Google Scholar
Kirschner, M. & Gerhart, J. (1998). Evolvability. Proceedings of the National Academy of Sciences of the United States of America, 95: 84208427.Google Scholar
Kitomi, Y., Ogawa, A., Kitano, H. & Inukai, Y. (2008). CRL4 regulates crown root formation through auxin transport in rice. Plant Root, 2: 1928.Google Scholar
Kliber, A. & Eckert, C. (2004). Sequential decline in allocation among flowers within inflorescences: proximate mechanisms and adaptive significance. Ecology, 85: 16751687.Google Scholar
Klingenberg, C. P. (1998). Heterochrony and allometry: the analysis of evolutionary change in ontogeny. Biological Reviews, 73: 79123.Google Scholar
Klingenberg, C. P. (2005). Developmental constraints, modules and evolvability. In Variation: A Central Concept in Biology, eds. Hallgrímsson, B. & Hall, B. K.. Burlington, MA: Elsevier, pp. 219247.Google Scholar
Klingenberg, C. P. (2008). Morphological integration and developmental modularity. Annual Review of Ecology and Systematics, 39: 115132.Google Scholar
Knapp, S. (2010). On ‘various contrivances’: pollination, phylogeny and flower form in the Solanaceae. Philosophical Transactions of the Royal Society B, 365: 449460.Google Scholar
Knight, C., Perroud, P. F. & Cove, D. (2009). The Moss Physcomitrella patens. London: Wiley-Blackwell.Google Scholar
Kobayashi, K., Maekawa, M., Miyao, A., Hirochika, H. & Kyozuka, J. (2010). PANICLE PHYTOMER2 (PAP2), encoding a SEPALLATA subfamily MADS-box protein, positively controls spikelet meristem identity in rice. Plant Cell Physiology, 51: 4757.Google Scholar
Koch, M. & Al-Shehbaz, I. A. (2002). Molecular data indicate complex intra- and intercontinental differentiation of American Draba (Brassicaceae). Annals of the Missouri Botanical Garden, 89: 88109.Google Scholar
Koch, M., Al-Shehbaz, I. A. & Mummenhoff, K. (2003). Molecular systematics, evolution, and population biology in the mustard family (Brassicaceae). Annals of the Missouri Botanical Garden, 90: 151171.Google Scholar
Koch, M., Haubold, B. & Mitchell-Olds, T. (2001). Molecular systematics of the Brassicaceae: evidence from coding plastidic matK and nuclear Chs sequences. American Journal of Botany, 88: 534544.Google Scholar
Koenig, D., Bayer, E., Kang, J., Kuhlemeier, C. & Sinha, N. (2009). Auxin patterns Solanum lycopersicum leaf morphogenesis. Development, 136: 29973006.Google Scholar
Koenig, D. & Weigel, D. (2015). Beyond the thale: comparative genomics and genetics of Arabidopsis relatives. Nature Reviews Genetics, 16: 285298.Google Scholar
Koes, R. (2008). Evolution and development of virtual inflorescences. Trends in Plant Science, 13: 13.Google Scholar
Kofuji, R., Sumikawa, N., Yamasaki, M. et al. (2003). Evolution and divergence of the MADS-box gene family based on genome-wide expression analyses. Molecular Biology and Evolution, 20: 19631977.Google Scholar
Kohn, J. R., Graham, S. W., Morton, B., Doyle, J. J. & Barrett, S. C. H. (1996). Reconstruction of the evolution of reproductive characters in Pontederiaceae using phylogenetic evidence from chloroplast DNA restriction-site variation. Evolution, 50: 14541469.Google Scholar
Koi, S., Imaichi, R. & Kato, M. (2005). Endogenous leaf initiation in the apical-meristemless shoot of Cladopus queenslandicus (Podostemaceae) and implications for evolution of shoot morphology. International Journal of Plant Sciences, 166: 199206.Google Scholar
Kölsch, A. & Gleissberg, S. (2006). Diversification of CYCLOIDEA-like TCP genes in the basal eudicot families Fumariaceae end Papaveraceae s.str. Plant Biology, 8: 680687.Google Scholar
Konishi, S., Izawa, T., Lin, S.Y. et al. (2006). An SNP caused loss of seed shattering during rice domestication. Science, 312: 13921396.Google Scholar
Kosuge, K. (1994). Petal evolution in Ranunculaceae. Plant Systematics and Evolution, Supplement, 8: 185191.Google Scholar
Kozlov, M. V., Wilsey, B. J., Koricheva, J. & Haukioja, E. (1996). Fluctuating asymmetry of birch leaves increases under pollution impact. Journal of Applied Ecology, 33: 14891495.Google Scholar
Kozo Poljanski, B. (1936). On some ‘third’ conceptions in floral morphology. New Phytologist, 35: 479492.Google Scholar
Kramer, E. M. (2009a). Aquilegia: a new model for plant development, ecology, and evolution. Annual Review of Plant Biology, 60: 261277.Google Scholar
Kramer, E. M. (2009b). New model systems for the study of developmental evolution in plants. Current Topics in Developmental Biology, 86: 65107.Google Scholar
Kramer, E. M., Di Stilio, V. S. & Schluter, P. M. (2003). Complex patterns of gene duplication in the Apetala3 and Pistillata lineages of the Ranunculaceae. International Journal of Plant Sciences, 164: 111.Google Scholar
Kramer, E. M., Dorit, R. L. & Irish, V. F. (1998). Molecular evolution of genes controlling petal and stamen development: duplication and divergence within the APETALA3 and PISTILLATA MADS-box gene lineages. Genetics, 149: 765783.Google Scholar
Kramer, E. M. & Hodges, S. A. (2010). Aquilegia as a model system for the evolution and ecology of petals. Philosophical Transactions of the Royal Society B, 365: 477490.Google Scholar
Kramer, E. M., Holappa, L., Gould, B. et al. (2007). Elaboration of B gene function to include the identity of novel floral organs in the lower eudicot Aquilegia. Plant Cell, 19: 750766.Google Scholar
Kramer, E. M. & Irish, V. F. (1999). Evolution of genetic mechanisms controlling petal development. Nature, 399: 144148.Google Scholar
Kramer, E. M. & Irish, V. F. (2000). Evolution of the petal and stamen developmental programs: evidence from comparative studies of the lower eudicots and basal angiosperms. International Journal of Plant Sciences, 161: S29S40.Google Scholar
Kramer, E. M. & Jaramillo, M. A. (2005). Genetic basis for innovations in floral organ identity. Journal of Experimental Zoology (Molecular and Developmental Evolution), 304B: 526535.Google Scholar
Kramer, E. M., Jaramillo, M. A. & Di Stilio, V. S. (2004). Patterns of gene duplication and functional evolution during the diversification of the AGAMOUS subfamily of MADS box genes in angiosperms. Genetics, 166: 10111023.Google Scholar
Kramer, E. M., Su, H. J., Wu, C. C. & Hu, J. M. (2006). A simplified explanation for the frameshift mutation that created a novel C-terminal motif in the APETALA3 gene lineage. BMC Evolutionary Biology, 6: 30.Google Scholar
Krizek, B. A. & Meyerowitz, E. M. (1996). The Arabidopsis homeotic genes APETALA3 and PISTILLATA are sufficient to provide the B class organ identity function. Development, 122: 1122.Google Scholar
Krolikowski, K. A., Victor, J. L., Wagler, T. N., Lolle, S. J. & Pruitt, R. J. (2003). Isolation and characterization of the Arabidopsis organ fusion gene HOTHEAD. The Plant Journal, 35: 501511.Google Scholar
Kubitzki, K. (1998). Taccaceae. In The Families and Genera of Vascular Plants, Vol. 3, ed. Kubitzki, K.. Berlin: Springer, pp. 425428.Google Scholar
Kuhlemeier, C. & Reinhardt, D. (2001). Auxin and phyllotaxis. Trends in Plant Science, 6: 187189.Google Scholar
Kuittinen, H., de Haan, A. A., Vogl, C. et al. (2004). Comparing the linkage maps of the close relatives Arabidopsis lyrata and A. thaliana. Genetics, 168: 15751584.Google Scholar
Kümpers, B. M. C., Richardson, J. E., Anderberg, A. A., Wilkie, P. & Ronse De Craene, L. (2016). The significance of meristic changes in the flowers of Sapotaceae. Botanical Journal of the Linnean Society, 180: 161192.Google Scholar
Kutschera, U., Langguth, H., Kuo, D.-H., Weisblat, D. A. & Shankland, M. (2013). Description of a new leech species from North America, Helobdella austinensis n. sp. (Hirudinea: Glossiphoniidae), with observations on its feeding behaviour. Zoosystematics and Evolution, 89: 239246.Google Scholar
Kuwabara, A., Tsukaya, H. & Nagata, T. (2001). Identification of factors that cause heterophylly in Ludwigia arcuata Walt. (Onagraceae). Plant Biology, 3: 98105.Google Scholar
Kwiatkowska, D. (1995). Ontogenetic changes of phyllotaxis in Anagallis arvensis L. Acta Societatis Botanicorum Poloniae, 64: 319325.Google Scholar
Kwiatkowska, D. (1999). Formation of pseudowhorls in Peperomia verticillata (L.) A. Dietr. shoots exhibiting various phyllotactic patterns. Annals of Botany, 83: 675685.Google Scholar
Kyozuka, J., Kobayashi, T., Morita, M. & Shimamoto, K. (2000). Spatially and temporally regulated expression of rice MADS box genes with similarity to Arabidopsis class A, B and C genes. Plant Cell Physiology, 41: 710718.Google Scholar
Labonne, J. D. J., Tamari, F. & Shore, J. S. (2010). Characterization of X-ray-generated floral mutants carrying deletions at the S-locus of distylous Turnera subulata. Heredity, 105: 235243.Google Scholar
Lacroix, C., Jeune, B. & Purcell-Macdonald, S. (2003). Shoot and compound leaf comparisons in eudicots: dynamic morphology as an alternative approach. Botanical Journal of the Linnean Society, 143: 219230.Google Scholar
Lacroix, C. & Sattler, R. (1988). Phyllotaxis theories and tepal-stamen superposition in Basella rubra. American Journal of Botany, 75: 906917.Google Scholar
Lacroix, C. R. & Sattler, R. (1994). Expression of shoot features in early leaf development of Murraya paniculata (Rutaceae). Canadian Journal of Botany, 72: 678687.Google Scholar
Lahti, D. C., Johnson, N. A., Ajie, B. C. et al. (2009). Relaxed selection in the wild. Trends in Ecology and Evolution, 24: 487496.Google Scholar
Laland, K. N., Uller, T., Feldman, M. W. et al. (2015). The extended evolutionary synthesis: its structure, assumptions and predictions. Proceedings of the Royal Society B, 282: 20151019.Google Scholar
Lamont, B. B. (1980). Tissue longevity of the arborescent monocotyledon, Kingia australis (Xanthorrhoeaceae). American Journal of Botany, 67: 12621264.Google Scholar
Lamsdell, J. C. & Selden, P. A. (2013). Babes in the wood: a unique window into sea scorpion ontogeny. BMC Evolutionary Biology, 13, 98.Google Scholar
Landis, J. B., Barnett, L. L. & Hileman, L. C. (2012). Evolution of petaloid sepals independent of shifts in B-class MADS box gene expression. Development Genes and Evolution, 222: 1928.Google Scholar
Landrein, B., Refahi, Y., Besnard, F. et al. (2015). Meristem size contributes to the robustness of phyllotaxis in Arabidopsis. Journal of Experimental Botany, 66: 13171324.Google Scholar
Lang, D., van Gessel, N., Ullrich, K. K. & Reski, R. (2016). The genome of the model moss Physcomitrella patens. Advances in Botanical Research, 78: 97140.Google Scholar
Langdale, J. & Harrison, J. C. (2008). Developmental transitions during the evolution of plant form. In Evolving Pathways: Key Themes in Evolutionary Developmental Biology, eds. Minelli, A. & Fusco, G.. Cambridge: Cambridge University Press, pp. 299319.Google Scholar
Langer, R. H. & Wilson, D. (1965). Environmental control of cleistogamy in prairie grass (Bromus unioloides HBK). New Phytologist, 65: 8085.Google Scholar
Langham, R. J., Walsh, J., Dunn, M. et al. (2004). Genomic duplication, fractionation and the origin of regulatory novelty. Genetics, 166: 935945.Google Scholar
Lanner, R. M. & Connor, K. F. (2001). Does bristlecone pine senesce? Experimental Gerontology, 36: 675685.Google Scholar
Lau, S., Slane, D., Herud, O., Kong, J. & Jürgens, G. (2012). Early embryogenesis in flowering plants: setting up the basic body pattern. Annual Review of Plant Biology, 63: 483506.Google Scholar
Laufs, P., Grandjean, O., Jonak, C., Kieu, K. & Traas, J. (1998). Cellular parameters of the shoot apical meristem in Arabidopsis. Plant Cell, 10: 13751389.Google Scholar
Laux, T., Mayer, K. F. X., Berger, J. & Jürgens, G. (1996). The WUSCHELL gene is required for shoot and floral meristem integrity in Arabidopsis. Development, 122: 8796.Google Scholar
Lavin, M., Herendeen, P. S. & Wojciechowski, M. F. (2005). Evolutionary rates analysis of Leguminosae implicates a rapid diversification of lineages during the Tertiary. Systematic Biology, 54: 575594.Google Scholar
Lawrence, G. H. (1951). Taxonomy of Flowering Plants. New York, NY: Macmillan.Google Scholar
Layton, D. J. & Kellogg, E. A. (2014). Morphological, phylogenetic, and ecological diversity of the new model species Setaria viridis (Poaceae: Paniceae) and its close relatives. American Journal of Botany, 101: 539557.Google Scholar
Le Rouzic, A. & Carlborg, O. (2008). Evolutionary potential of hidden genetic variation. Trends in Ecology and Evolution, 23: 3337.Google Scholar
Lee, J., Park, J.-J., Kim, S. L., Yim, J. & An, G. (2007). Mutations in the rice liguleless gene result in a complete loss of the auricle, ligule and laminar joint. Plant Molecular Biology, 65: 487499.Google Scholar
Lee, J.-H., Lin, H., Joo, S. & Goodenough, U. (2008). Early sexual origins of homeoprotein heterodimerization and evolution of the plant KNOX/BELL family. Cell, 133: 829840.Google Scholar
Lee, J.-Y., Baum, S. F., Oh, S. H. et al. (2005). Recruitment of CRABS CLAW to promote nectary development within the eudicot clade. Development, 132: 50215032.Google Scholar
Lee, J.-Y., Mummenhoff, K. & Bowman, J. L. (2002). Allopolyploidization and evolution of species with reduced floral structures in Lepidium L. (Brassicaceae). Proceedings of the National Academy of Sciences of The United States of America, 99: 1683516840.Google Scholar
Leinfellner, W. (1958). Über die peltaten Kronblätter der Sapindaceen. Österreichische botanische Zeitschrift, 105: 443514.Google Scholar
Leins, P. & Erbar, C. (1997). Floral developmental studies: some old and new questions. International Journal of Plant Sciences, 158: S3S12.Google Scholar
Leins, P. & Erbar, C. (2004). Floral organ sequences in Apiales (Apiaceae, Araliaceae, Pittosporaceae). South African Journal of Botany, 70: 468474.Google Scholar
Leins, P. & Erbar, C. (2008). Blüte und Frucht, 2nd edn. Stuttgart: Schweizerbart.Google Scholar
Lewis, D. & Jones, D. A. (1992). The genetics of heterostyly. In Evolution and Function of Heterostyly, ed. Barrett, S. C. H.. Berlin: Springer, pp. 129150.Google Scholar
Li, D., Liu, C., Shen, L. et al. (2008). A repressor complex governs the integration of flowering signals in Arabidopsis. Developmental Cell, 15: 110120.Google Scholar
Li, G. S., Meng, Z., Kong, H. Z. et al. (2005). Characterization of candidate class A, B and E floral homeotic genes from the perianthless basal angiosperm Chloranthus spicatus (Chloranthaceae). Development Genes and Evolution, 215: 437449.Google Scholar
Li, H., Liang, W., Hu, Y. et al. (2011). Rice MADS6 interacts with the floral homeotic genes SUPERWOMAN1, MADS3, MADS58, MADS13, and DROOPING LEAF in specifying floral organ identities and meristem fate. Plant Cell, 23: 25362552.Google Scholar
Li, J., Webster, M. A., Furuya, M. & Gilmartin, P. M. (2007). Identification and characterization of pin and thrum alleles of two genes that co-segregate with the Primula S locus. The Plant Journal, 51: 1831.Google Scholar
Li, P. & Johnston, M. O. (2000). Heterochrony in plant evolutionary studies through the twentieth century. Botanical Review, 66: 5788.Google Scholar
Li, P. & Johnston, M. O. (2010). Flower development and the evolution of self-fertilization in Amsinckia: the role of heterochrony. Evolutionary Biology, 37: 143168.Google Scholar
Li, Z., Reighard, G. L., Abbott, A. G. & Bielenberg, D. G. (2009). Dormancy-associated MADS genes from the EVG locus of peach [Prunus persica (L.) Batsch] have distinct seasonal and photoperiodic expression patterns. Journal of Experimental Botany, 60: 35213530.Google Scholar
Liljegren, S. J., Ditta, G. S., Eshed, H. Y. et al. (2000). SHATTERPROOF MADS-box genes control seed dispersal in Arabidopsis. Nature, 404: 766770.Google Scholar
Linnaeus, C. (1754). Genera plantarum: eorumque characteres naturales secundum numerum, figuram, situm, et proportionem omnium fructificationis partium. Editio Quinta. Holmiæ: Laurentius Salvius.Google Scholar
Litt, A. (2007). An evaluation of A-function: evidence from the APETALA1 and APETALA2 gene lineages. International Journal of Plant Sciences, 168: 7391.Google Scholar
Litt, A. (2013). Comparative evolutionary genomics of land plants. Annual Plant Reviews, 45: 227276.Google Scholar
Litt, A. & Irish, V. F. (2003). Duplication and diversification in the APETALA1/ FRUITFULL floral homeotic gene lineage: implications for the evolution of floral development. Genetics, 165: 821833.Google Scholar
Litt, A. & Kramer, E. M. (2010). The ABC model and the diversification of floral organ identity. Seminars in Cell and Developmental Biology, 21: 129137.Google Scholar
Liu, C., Thong, Z. & Yu, H. (2009a). Coming into bloom: the specification of floral meristems. Development, 136: 33793391.Google Scholar
Liu, S., Wang, J., Wang, L. et al. (2009b) Adventitious root formation in rice requires OsGNOM1 and is mediated by the OsPINs family. Cell Research, 19: 11101119.Google Scholar
Liu, Z. C., Franks, R. G. & Klink, V. P. (2000). Regulation of gynoecium marginal tissue formation by LEUNIG and AINTEGUMENTA. Plant Cell, 12: 18791891.Google Scholar
Liu, Z.-J. & Wang, X. (2016). A perfect flower from the Jurassic of China. Historical Biology, 28: 707719.Google Scholar
Lloyd, D. G. & Webb, C. J. (1977). Secondary sex characters in plants. Botanical Review, 43: 177216.Google Scholar
Lloyd, D. G. & Webb, C. J. (1986). The avoidance of interference between the presentation of pollen and stigmas in angiosperms. I. Dichogamy. New Zealand Journal of Botany, 24: 135162.Google Scholar
Loiseau, J.-E. (1969). La Phyllotaxie. Paris: Masson.Google Scholar
Lolle, S. J., Cheung, A. Y. & Sussex, I. M. (1992). Fiddlehead: an Arabidopsis mutant constitutively expressing an organ fusion program that involves interactions between epidermal cells. Developmental Biology, 152: 383392.Google Scholar
Long, J. & Barton, M. K. (2000). Initiation of axillary and floral meristems in Arabidopsis. Developmental Biology, 218: 341353.Google Scholar
Long, J. A., Moan, E. I., Medford, J. I. et al. (1996). A member of the KNOTTED class of homeodomain proteins encoded by the STM gene of Arabidopsis. Nature, 379: 6669.Google Scholar
Lord, E. M. (1979). The development of cleistogamous and chasmogamous flowers in Lamium amplexicaule (Labiatae): an example of heteroblastic inflorescence development. Botanical Gazette, 140: 3950.Google Scholar
Lord, E. M. (1981). Cleistogamy: a tool for the study of floral morphogenesis, function and evolution. Botanical Review, 47: 421449.Google Scholar
Lord, E. M. (1982). Floral morphogenesis in L. amplexicaule L. (Labiateae) with a model for the evolution of the cleistogamous flower. Botanical Gazette, 143: 6372.Google Scholar
Lord, E. M. (2001). Heterochrony in plants. In Encyclopedia of Life Sciences. Chichester: Wiley.Google Scholar
Lord, E. M. & Hill, J. P. (1987). Evidence for heterochrony in the evolution of plant form. In Development as an Evolutionary Process, eds. Raff, R. A. & Raff, E. C.. New York, NY: Alan R. Liss, pp. 4770.Google Scholar
Love, A. C. (2010). Idealization in evolutionary developmental investigation: a tension between phenotypic plasticity and normal stages. Philosophical Transactions of the Royal Society B, 365: 679690.Google Scholar
Luo, D., Carpenter, R., Copsey, L. et al. (1999). Control of organ asymmetry in flowers of Antirrhinum. Cell, 99: 367376.Google Scholar
Luo, D., Carpenter, R., Vincent, C., Copsey, L. & Coen, E. (1996). Origin of floral asymmetry in Antirrhinum. Nature, 383: 794799.Google Scholar
Lynch, M. & Conery, J. S. (2000). The evolutionary fate and consequences of duplicate genes. Science, 290: 11511155.Google Scholar
Lyndon, R. F. (1998). The Shoot Apical Meristem: Its Growth and Development. Cambridge: Cambridge University Press.Google Scholar
Lysak, M. A., Berr, A., Pecinka, A. et al. (2006). Mechanisms of chromosome number reduction in Arabidopsis thaliana and related Brassicaceae species. Proceedings of the National Academy of Sciences of the United States of America, 103: 52245229.Google Scholar
Ma, H. (1998). To be, or not to be, a flower: control of floral meristem identity. Trends in Genetics, 14: 2632.Google Scholar
Maizel, A. (2016). Plant organ growth: stopping under stress. Current Biology, 26: R417R419.Google Scholar
Malcomber, S. T. & Kellogg, E. A. (2004). Heterogeneous expression patterns and separate roles of the SEPALLATA gene LEAFY HULL STERILE1 in grasses. Plant Cell, 16: 16921706.Google Scholar
Malcomber, S. T. & Kellogg, E. A. (2005). SEPALLATA gene diversification: brave new whorls. Trends in Plant Science, 10: 427435.Google Scholar
Malcomber, S. T. & Kellogg, E. A. (2006). Evolution of unisexual flowers in grasses (Poaceae) and the putative sex-determination gene, TASSELSEED2 (TS2). New Phytologist, 170: 885899.Google Scholar
Malcomber, S. T., Preston, J. C., Reinheimer, R., Kossuth, J. & Kellogg, E. A. (2006). Developmental gene evolution and the origin of grass inflorescence diversity. Advances in Botanical Research, 44: 425481.Google Scholar
Manger, H. L. (1783). Vollstaendige Anleitung zu einer Systematischen Pomologie wodurch die Genauigste Kentniß von der Natur, Beschaffenheit und den Unterschiedenen Merkmalen aller Obstarten Enthalten Werden Kann. Zweyter Theil von den Birnen. Leipzig: J. F. Junius.Google Scholar
Manos, P. S. & Stone, D. E. (2001). Evolution, phylogeny, and systematics of the Juglandaceae. Annals of the Missouri Botanical Garden, 88: 231262.Google Scholar
Mantegazza, R., Möller, M., Harrison, C. J. et al. (2007). Anisocotyly and meristem initiation in an unorthodox plant, Streptocarpus rexii (Gesneriaceae). Planta, 225: 653663.Google Scholar
Mantegazza, R., Tononi, P., Möller, M. & Spada, A. (2009). WUS and STM homologues are linked to the expression of lateral dominance in the acaulescent Streptocarpus rexii (Gesneriaceae). Planta, 230: 529542.Google Scholar
Marazzi, B. & Endress, P. K. (2008). Patterns and development of floral asymmetry in Senna (Leguminosae, Cassiinae). American Journal of Botany, 95: 2240.Google Scholar
Marazzi, B., Endress, P. K., Paganucci de Queiroz, L. & Conti, E. (2006). Phylogenetic relationships within Senna (Leguminosae, Cassiinae) based on three chloroplast DNA regions: patterns in the evolution of floral symmetry and extrafloral nectaries. American Journal of Botany, 93: 288303.Google Scholar
Martín-Trillo, M. & Cubas, P. (2010). TCP genes: a family snapshot ten years later. Trends in Plant Science, 15: 3139.Google Scholar
Martinez, C. C., Chitwood, D. H., Smith, R. S. & Sinha, N. R. (2016). Left–right leaf asymmetry in decussate and distichous phyllotactic systems. Philosophical Transactions of the Royal Society B, 371: 20150412.Google Scholar
Martínez-Laborda, A. & Vera, A. (2009). Arabidopsis fruit development. Annual Plant Reviews, 38: 172203.Google Scholar
Masel, J. & Siegal, M. L. (2009). Robustness: mechanisms and consequences. Trends in Genetics, 25: 395403.Google Scholar
Masel, J. & Trotter, M. V. (2010). Robustness and evolvability. Trends in Genetics, 26: 406414.Google Scholar
Masiero, S., Li, M. A., Will, I. et al. (2004). INCOMPOSITA: a MADS-box gene controlling prophyll development and floral meristem identity in Antirrhinum. Development, 131: 59815990.Google Scholar
Matsuhashi, S., Sakai, S. & Kudoh, H. (2012). Temperature-dependent fluctuation of stamen number in Cardamine hirsuta (Brassicaceae). International Journal of Plant Sciences, 173: 391398.Google Scholar
Matsui, K., Nishio, T. & Tetsuka, T. (2004). Genes outside the S supergene suppress S functions in buckwheat (Fagopyrum esculentum). Annals of Botany, 94: 805809.Google Scholar
Matthews, M. L. & Endress, P. K. (2008). Comparative floral structure and systematics in Chrysobalanaceae s.l. (Chrysobalanaceae, Dichapetalaceae, Euphroniaceae, and Trigoniaceae; Malpighiales). Botanical Journal of the Linnean Society, 157: 249309.Google Scholar
Mauseth, J. D. (1991). Botany. Orlando, FL: Holt Rhinehart & Winston.Google Scholar
Mayer, V., Möller, M., Perret, M. & Weber, A. (2003). Phylogenetic position and generic differentiation of Epithemateae (Gesneriaceae) inferred from plastid DNA sequence data. American Journal of Botany, 90: 321329.Google Scholar
Mayers, A. M. & Lord, E. M. (1983). Comparative flower development in the cleistogamous species Viola odorata. I. A growth rate study. American Journal of Botany, 70: 15481555.Google Scholar
Mayers, A. M. & Lord, E. M. (1984). Comparative floral development in the cleistogamous species Viola odorata. III. A histological study. Botanical Gazette, 145: 8391.Google Scholar
Mayo, S. J., Bogner, J. & Boyce, P. C. (1997). The Genera of Araceae. Kew: Royal Botanical Gardens.Google Scholar
Mayo, S. J., Bogner, J. & Boyce, P. C. (1998). Araceae. In The Families and Genera of Vascular Plants, Vol. 4, ed. Kubitzki, K.. Berlin: Springer, pp. 2673.Google Scholar
Mayr, E. (1982). The Growth of Biological Thought: Diversity, Evolution, and Inheritance. Cambridge, MA: Harvard University Press.Google Scholar
McConnell, J. R. & Barton, M. K. (1998). Leaf polarity and meristem formation in Arabidopsis. Development, 125: 29352942.Google Scholar
McConnell, J. R., Emery, J., Eshed, Y. et al. (2001). Role of PHABULOSA and PHAVOLUTA in determining radial patterning in shoots. Nature, 411: 709713.Google Scholar
McCouch, S. R. (2008). Gene nomenclature system for rice. Rice, 1: 7284.Google Scholar
McDill, J., Repplinger, M., Simpson, B. B. & Kadereit, J. W. (2009). The phylogeny of Linum and Linaceae subfamily Linoideae, with implications for their systematics, biogeography, and evolution of heterostyly. Systematic Botany, 34: 386405.Google Scholar
McHale, N. A. & Koning, R. E. (2004). PHANTASTICA regulates development of the adaxial mesophyll in Nicotiana leaves. Plant Cell, 16: 12511262.Google Scholar
McIntosh, R. A. (1988). A catalogue of gene symbols for wheat. In Proceedings of the 7th International Wheat Genetics Symposium, eds. Miller, T. E. & Koebner, R. M. D.. Cambridge: IPSR, pp. 12251324.Google Scholar
McIntyre, G. I. & Best, K. F. (1975). Studies on the flowering of Thlaspi arvense L. II. A comparative study of early- and late-flowering strains. Botanical Gazette, 136: 151158.Google Scholar
McIntyre, G. I. & Best, K. F. (1978). Studies on the flowering of Thlaspi arvense L. IV. Genetic and ecological differences between early- and late-flowering strains. Botanical Gazette, 139: 190195.Google Scholar
McKone, M. J. & Tonkyn, D. W. (1986). Intrapopulation gender variation in common ragweed (Asteracae: Ambrosia artemisiifolia L.), a monecious, annual herb. Oecologia, 70: 6367.Google Scholar
McNamara, K. J. (1986). A guide to the nomenclature of heterochrony. Journal of Paleontology, 60: 413.Google Scholar
McNeill, J., Barrie, F. R., Buck, W. R. et al. (eds.) (2012). International Code of Nomenclature for Algae, Fungi, and Plants (Melbourne Code) Adopted by the Eighteenth International Botanical Congress Melbourne, Australia, July 2011. Ruggell: Gantner.Google Scholar
M’Cosh, J. (1851). Some remarks on the plant morphologically considered. Transactions of the Botanical Society, 4: 127132.Google Scholar
M’Cosh, J. & Dickie, G. (1856). Typical Forms and Special Ends in Creation. New York, NY: Carter.Google Scholar
Meeuse, A. D. J. (1972). Sixty-five years of theories of the multiaxial flower. Acta Biotheoretica, 21: 167202.Google Scholar
Meijer, M. & Murray, J. A. (2001). Cell cycle controls and the development of plant form. Current Opinion in Plant Biology, 4: 4449.Google Scholar
Meinke, D. W., Cherry, J. M., Dean, C., Rounsley, S. D. & Koornneef, M. (1998). Arabidopsis thaliana: a model plant for genome analysis. Science, 282: 662682.Google Scholar
Melville, R. (1960). A new theory of the angiosperm flower. Nature, 118: 1418.Google Scholar
Melzer, R. & Theißen, G. (2016). The significance of developmental robustness for species diversity. Annals of Botany, 117: 725732.Google Scholar
Melzer, R., Verelst, W. & Theißen, G. (2009). The class E floral homeotic protein SEPALLATA3 is sufficient to loop DNA in ‘floral quartet’-like complexes in vitro. Nucleic Acids Research, 37: 144157.Google Scholar
Melzer, R., Wang, Y. Q. & Theißen, G. (2010). The naked and the dead: the ABCs of gymnosperm reproduction and the origin of the angiosperm flower. Seminars in Cell and Developmental Biology, 21: 118128.Google Scholar
Mena, M., Ambrose, B. A., Meeley, R. B. et al. (1996). Diversification of C-function activity in maize flower development. Science, 274: 15371540.Google Scholar
Meng, A., Zhang, Z., Li, J., Ronse De Craene, L. & Wang, H. (2012). Floral development of Stephania (Menispermaceae): impact of organ reduction on symmetry. International Journal of Plant Sciences, 173: 861874.Google Scholar
Merckx, V. (2013). Mycoheterotrophy: The Biology of Plants Living on Fungi. Berlin: Springer.Google Scholar
Merrill, E. K. (1979). Comparison of ontogeny of three types of leaf architecture in Sorbus L. (Rosaceae). Botanical Gazette, 140: 328337.Google Scholar
Mestek Boukhibar, L. & Barkoulas, M. (2016). The developmental genetics of biological robustness. Annals of Botany, 117: 699707.Google Scholar
Metzger, R. J. & Krasnow, M. A. (1999). Genetic control of branching morphogenesis. Science, 284: 16351639.Google Scholar
Meyen, S. V. (1973). Plant morphology in its nomothetical aspects. Botanical Review, 39: 205260.Google Scholar
Meyerowitz, E. M. (1994). The genetics of flower development. Scientific American, 271: 4047.Google Scholar
Meyerowitz, E. M. (1997). Control of cell division patterns in developing shoots and flowers of Arabidopsis thaliana. Cold Spring Harbor Symposia on Quantitative Biology, 62: 369375.Google Scholar
Michaels, S. D. & Amasino, R. M. (1999). FLOWERING LOCUS C encodes a novel MADS domain protein that acts as a repressor of flowering. Plant Cell, 11: 949956.Google Scholar
Miller, A. P. (1995). Leaf-mining insects and fluctuating asymmetry in elm Ulmus glabra leaves. Journal of Animal Ecology, 64: 697707.Google Scholar
Minelli, A. (2000). Limbs and tail as evolutionarily diverging duplicates of the main body axis. Evolution and Development, 2: 157165.Google Scholar
Minelli, A. (2003). The Development of Animal Form: Ontogeny, Morphology, and Evolution. Cambridge: Cambridge University Press.Google Scholar
Minelli, A. (2009a). Forms of Becoming: The Evolutionary Biology of Development. Princeton, NJ: Princeton University Press.Google Scholar
Minelli, A. (2009b). Perspectives in Animal Phylogeny and Evolution. Oxford: Oxford University Press.Google Scholar
Minelli, A. (2011). A principle of developmental inertia. In Epigenetics: Linking Genotype and Phenotype in Development and Evolution, eds. Hallgrímsson, B. & Hall, B. K.. San Francisco, CA: University of California Press, pp. 116133.Google Scholar
Minelli, A. (2014). Developmental disparity. In Towards a Theory of Development, eds. Minelli, A. & Pradeu, T.. Oxford: Oxford University Press, pp. 227245.Google Scholar
Minelli, A. (2015a). Biological systematics in the evo-devo era. European Journal of Taxonomy, 125: 123.Google Scholar
Minelli, A. (2015b). Evo devo and its significance for animal evolution and phylogeny. In Evolutionary Developmental Biology of Invertebrates. 1. Introduction, Non-Bilateria, Acoelomorpha, Xenoturbellida, Chaetognatha, ed. Wanninger, A.. Wien: Springer, pp. 123.Google Scholar
Minelli, A. (2015c). Grand challenges in evolutionary developmental biology. Frontiers in Ecology and Evolution, 2: 85.Google Scholar
Minelli, A. (2015d). Constraints on animal (and plant) form in nature and art. Art and Perception, 3: 265281.Google Scholar
Minelli, A. (2016a). Scaffolded biology. Theory in Biosciences, 135: 163173.Google Scholar
Minelli, A. (2016b). Tracing homologies in an ever-changing world. Rivista di estetica, n.s., 56: 4055.Google Scholar
Minelli, A. (2016c). Species diversity vs. morphological disparity in the light of evolutionary developmental biology. Annals of Botany, 117: 781794.Google Scholar
Minelli, A. (2017). Evolvability and its evolvability. In Challenging the Modern Synthesis: Adaptation, Development, and Inheritance, eds. Huneman, P. & Walsh, D.. Oxford: Oxford University Press, pp. 211238.Google Scholar
Minelli, A. & Fusco, G. (2005). Conserved vs. innovative features in animal body organization. Journal of Experimental Zoology (Molecular and Developmental Evolution), 304B: 520525.Google Scholar
Minelli, A. & Fusco, G. (2012). On the evolutionary developmental biology of speciation. Evolutionary Biology, 39: 242254.Google Scholar
Minelli, A. & Fusco, G. (2013). Homology. In The Philosophy of Biology: A Companion for Educators, History, Philosophy and Theory of the Life Sciences, ed. Kampourakis, K.. Dordrecht: Springer, pp. 289322.Google Scholar
Minelli, A., Negrisolo, E. & Fusco, G. (2006). Reconstructing animal phylogeny in the light of evolutionary developmental biology. In Reconstructing the Tree of Life: Taxonomy and Systematics of Species Rich Taxa, eds. Hodkinson, T. R., Parnell, J. A. N. & Waldren, S.. Boca Raton, FL: Taylor & Francis/CRC Press, pp. 177190.Google Scholar
Minelli, A. & Pradeu, T. (eds.) (2014). Towards a Theory of Development. Oxford: Oxford University Press.Google Scholar
Minter, T. C. & Lord, E. M. (1983). A comparison of cleistogamous and chasmogamous floral development in Collomia grandiflora Dougl. Ex Lindl. (Polemoniaceae). American Journal of Botany, 70: 14991508.Google Scholar
Mirabet, V., Das, P., Boudaoud, A. & Hamant, O. (2011). The role of mechanical forces in plant morphogenesis. Annual Review of Plant Biology, 62: 365385.Google Scholar
Mitchell, C. H. & Diggle, P. K. (2005). The evolution of unisexual flowers: morphological and functional convergence results from diverse developmental transitions. American Journal of Botany, 92: 10681076.Google Scholar
Mitchell-Olds, T. (2001). Arabidopsis thaliana and its wild relatives: a model system for ecology and evolution. Trends in Ecology and Evolution, 16: 693700.Google Scholar
Mitchell-Olds, T., Al-Shehbaz, I. A., Koch, M. & Sharbel, T. F. (2005). Crucifer evolution in the post-genomic era. In Plant Diversity and Evolution: Genotypic and Phenotypic Variation in Higher Plants, ed. Henry, R. J.. Cambridge, MA: CAB International, pp. 119137.Google Scholar
Mitsuda, N. & Ohme-Takagi, M. (2009). Functional analysis of transcription factors in Arabidopsis. Plant Cell Physiology, 50: 12321248.Google Scholar
Miwa, H., Kinoshita, A., Fukuda, H. & Sawa, S. (2009). Plant meristems: CLAVATA3/ESR-related signaling in the shoot apical meristem and the root apical meristem. Journal of Plant Research, 122: 3139.Google Scholar
Mizukami, Y. & Fischer, R. L. (2000). Plant organ size control: AINTEGUMENTA regulates growth and cell numbers during organogenesis. Proceedings of the National Academy of Sciences of the United States of America, 97: 942947.Google Scholar
Moczek, A. P. (2008). On the origins of novelty in development and evolution. BioEssays, 30: 432447.Google Scholar
Moczek, A. P. (2010). Phenotypic plasticity and diversity in insects. Philosophical Transactions of the Royal Society B, 365: 593603.Google Scholar
Molinero-Rosales, N., Jamilena, M., Zurita, S. et al. (1999). FALSIFLORA, the tomato orthologue of FLORICAULA and LEAFY, controls flowering time and floral meristem identity. The Plant Journal, 20: 685693.Google Scholar
Möller, M., Pfosser, M., Jang, C. G. et al. (2009). A preliminary phylogeny of the ‘didymocarpoid Gesneriaceae’ based on three molecular data sets: incongruence with available tribal classifications. American Journal of Botany, 96: 9891010.Google Scholar
Mondragón-Palomino, M., Hiese, L., Harter, A., Koch, M. A. & Theißen, G. (2009). Positive selection and ancient duplications in the evolution of class B floral homeotic genes of orchids and grasses. BMC Evolutionary Biology, 9: 81.Google Scholar
Mondragón-Palomino, M. & Theißen, G. (2008). MADS about the evolution of orchid flowers. Trends in Plant Science, 13: 5159.Google Scholar
Mondragón-Palomino, M. & Theißen, G. (2009). Why are orchid flowers so diverse? Reduction of evolutionary constraints by paralogues of class B floral homeotic genes. Annals of Botany, 104: 583594.Google Scholar
Mondragón-Palomino, M. & Theißen, G. (2011). Conserved differential expression of paralogous DEFICIENS- and GLOBOSA-like MADS-box genes in the flowers of Orchidaceae: refining the ‘orchid code’. The Plant Journal, 66: 10081019.Google Scholar
Mondragón-Palomino, M. & Trontin, C. (2011). High time for a roll call: gene duplication and phylogenetic relationships of TCP-like genes in monocots. Annals of Botany, 107: 15331544.Google Scholar
Monniaux, M., Pieper, B. & Hay, A. (2016). Stochastic variation in Cardamine hirsuta petal number. Annals of Botany, 117: 881887.Google Scholar
Moody, A., Diggle, P. K. & Steingraeber, D. A. (1999). Developmental analysis of the evolutionary origin of vegetative propagules in Mimulus gemmiparus (Scrophulariaceae). American Journal of Botany, 86: 15121522.Google Scholar
Mordhorst, A. P., Toonen, M. A. J., de Vries, S. C. & Meinke, D. (1997). Plant embryogenesis. Critical Reviews in Plant Sciences, 16: 535576.Google Scholar
Moreno-Risueno, M. A., Busch, W. & Benfey, P. N. (2010). Omics meet networks: using systems approaches to-infer regulatory networks in plants. Current Opinion in Plant Biology, 13: 126131.Google Scholar
Moreno-Risueno, M. A., Van Norman, J. M. & Benfey, P. N. (2012). Transcriptional switches direct plant organ formation and patterning. Current Topics in Developmental Biology, 98: 229257.Google Scholar
Mouradov, A., Glassick, T., Hamdorf, B. et al. (1998). NEEDLY, a Pinus radiata ortholog of FLORICAULA/LEAFY genes, expressed in both reproductive and vegetative meristems. Proceedings of the National Academy of Sciences of the United States of America, 95: 65376542.Google Scholar
Mower, J. P., Stefanovic, S., Young, G. J. & Palmer, J. D. (2004). Gene transfer from parasitic to host plants. Nature, 432: 165166.Google Scholar
Moyroud, E., Kusters, E., Monniaux, M., Koes, R. & Parcy, F. (2010). LEAFY blossoms. Trends in Plant Science, 15: 346352.Google Scholar
Mueller, A.L., Solow, T. H., Taylor, N. et al. (2005). The SOL Genomics Network (SGN): a comparative resource for solanaceous biology and beyond. Plant Physiology, 138: 13101317.Google Scholar
Mukherjee, K. & Brocchieri, L. (2010). Evolution of plant homeobox genes. In Encyclopedia of Life Sciences. Chichester: Wiley.Google Scholar
Mukherjee, K., Brocchieri, L. & Burglin, T. R. (2009). A comprehensive classification and evolutionary analysis of plant homeobox genes. Molecular Biology and Evolution, 26: 27752794.Google Scholar
Müller, M. & Cronk, Q. C. B. (2001). Evolution of morphological novelty: a phylogenetic analysis of growth patterns in Streptocarpus (Gesneriaceae). Evolution, 55: 918929.Google Scholar
Mummenhoff, K., Al-Shehbaz, I. A., Bakker, F. T., Linder, H. P. & Mühlhausen, A. (2005). Phylogeny, morphological evolution, and speciation of endemic Brassicaceae genera in the Cape flora of southern Africa. Annals of the Missouri Botanical Garden, 92: 400424.Google Scholar
Mummenhoff, K., Brüggemann, H. & Bowman, J. L. (2001). Chloroplast DNA phylogeny and biogeography of Lepidium (Brassicaceae). American Journal of Botany, 88: 20512063.Google Scholar
Mummenhoff, K., Polster, A., Mühlhausen, A. & Theißen, G. (2009). Lepidium as a model system for studying the evolution of fruit development in Brassicaceae. Journal of Experimental Botany, 60: 15031513.Google Scholar
Mungall, C., Gkoutos, G. V., Smith, C., Haendel, M. & Ashburner, M. (2010). Integrating phenotype ontologies across multiple species. Genome Biology, 11: R2.Google Scholar
Munné-Bosch, S. (2008). Do perennials really senesce? Trends in Plant Science, 13: 216220.Google Scholar
Münster, T., Pahnke, J., Di Rosa, A et al. (1997). Floral homeotic genes were recruited from homologous MADS-box genes preexisting in the common ancestor of ferns and seed plants. Proceedings of the National Academy of Sciences of the United States of America, 94: 24152420.Google Scholar
Münster, T., Wingen, L. U., Faigl, W. et al. (2001). Characterization of three GLOBOSA-like MADS-box genes from maize: evidence for ancient paralogy in one class of floral homeotic B-function genes of grasses. Gene, 262: 113.Google Scholar
Müntzing, A. (1936). The evolutionary significance of autopolyploidy. Hereditas, 21: 263378.Google Scholar
Murai, K., Miyamae, M., Kato, H., Takumi, S. & Ogihara, Y. (2003). WAP1, a wheat APETALA1 homolog, plays a central role in the phase transition from vegetative to reproductive growth. Plant and Cell Physiology, 44: 12551265.Google Scholar
Murray, N. A. & Johnson, D. M. (1987). Synchronous dichogamy in a Mexican anonillo Rollinia jimenezi var. nelsonii. Contributions from the University of Michigan Herbarium, 16: 173178.Google Scholar
Nagasawa, N., Miyoshi, M., Sano, Y. et al. (2003). SUPERWOMAN1 and DROOPING LEAF genes control floral organ identity in rice. Development, 130: 705718.Google Scholar
Nah, G. & Chen, J. (2010). Tandem duplication of the FLC locus and the origin of a new gene in Arabidopsis related species and their functional implications in allopolyploids. New Phytologist, 186: 228238.Google Scholar
Naiki, A. (2012). Heterostyly and the possibility of its breakdown by polyploidization. Plant Species Biology, 27: 329.Google Scholar
Nakada, M., Komatsu, M., Ochiai, T. et al. (2006). Isolation of MaDEF from Muscari armeniacum and analysis of its expression using laser microdissection. Plant Science, 170: 143150.Google Scholar
Nakamura, T., Fukuda, T., Nakano, M. et al. (2005). The modified ABC model explains the development of the petaloid perianth of Agapanthus praecox ssp. orientalis (Agapanthaceae) flowers. Plant Molecular Biology, 58: 435445.Google Scholar
Nakano, T., Kimbara, J., Fujisawa, M., Kitagawa, M. et al. (2012). MACROCALYX and JOINTLESS interact in the transcriptional regulation of tomato fruit abscission zone development. Plant Physiology, 158: 439450.Google Scholar
Nakata, M., Matsumoto, N., Tsugeki, R. et al. (2012). Roles of the middle domain-specific WUSCHEL-RELATED HOMEOBOX genes in early development of leaves in Arabidopsis. Plant Cell, 24: 519535.Google Scholar
Nakayama, H., Nakayama, N., Seiki, S. et al. (2014). Regulation of the KNOX-GA gene module induces heterophyllic alteration in North American lake cress. Plant Cell, 26: 47334748.Google Scholar
Nakayama, H., Yamaguchi, T. & Tsukaya, H. (2012). Acquisition and diversification of cladodes: leaf-like organs in the genus Asparagus. Plant Cell, 24: 929940.Google Scholar
Nardmann, J. & Werr, W. (2006). The shoot stem cell niche in angiosperms: expression patterns of WUS orthologues in rice and maize imply major modifications in the course of mono- and dicot evolution. Molecular Biology and Evolution, 23: 24922504.Google Scholar
Nath, U., Crawford, B. C., Carpenter, R. & Coen, E. (2003). Genetic control of surface curvature. Science, 299: 14041407.Google Scholar
Nickrent, D. L., Blarer, A., Qiu, Y. L., Vidal-Russell, R. & Anderson, F. E. (2004). Phylogenetic inference in Rafflesiales: the influence of rate heterogeneity and horizontal gene transfer. BMC Evolutionary Biology, 4: 40.Google Scholar
Nicolas, M. & Cubas, P. (2016). The role of TCP transcription factors in shaping flower structure, leaf morphology, and plant architecture. In Plant Transcription Factors: Evolutionary, Structural and Functional Aspects, ed. Gonzalez, D. H.. Amsterdam: Elsevier, pp. 249267.Google Scholar
Niklas, K. J. (2016). Plant Evolution: An Introduction to the History of Life. Chicago, IL: University of Chicago Press.Google Scholar
Nikolov, L. A., Staedler, Y. M., Manickam, S. et al. (2014). Floral structure and development in Rafflesiaceae with emphasis on their exceptional gynoecia. American Journal of Botany, 101: 225243.Google Scholar
Nikovics, K., Blein, T., Peaucelle, A. et al. (2006). The balance between the MIR164A and CUC2 genes controls leaf margin serration in Arabidopsis. Plant Cell, 18: 29292945.Google Scholar
Nishiyama, T., Fujita, T., Shin-I, T. et al. (2003). Comparative genomics of Physcomitrella patens gametophytic transcriptome and Arabidopsis thaliana: implication for land plant evolution. Proceedings of the National Academy of Sciences of the United States of America, 100: 80078012.Google Scholar
Nodine, M. D. & Bartel, D. P. (2012). Maternal and paternal genomes contribute equally to the transcriptome of early plant embryos. Nature, 482: 9497.Google Scholar
Nowak, M. D., Russo, G., Schlapbach, R. et al. (2015). The draft genome of Primula veris yields insights into the molecular basis of heterostyly. Genome Biology, 16: 12.Google Scholar
Nuraliev, M. S., Degtajareva, G. V., Sokoloff, D. D. et al. (2014). Flower morphology and relationships of Schefflera subintegra (Araliaceae, Apiales): an evolutionary step towards extreme floral polymery. Botanical Journal of the Linnean Society, 175: 553597.Google Scholar
Nürnberger, T. & Brunner, F. (2002). Innate immunity in plants and animals: emerging parallels between the recognition of general elicitors and pathogen-associated molecular patterns. Current Opinion in Plant Biology, 5: 318324.Google Scholar
Nutt, P., Ziermann, J., Hintz, M., Neuffer, B. & Theißen, G. (2006). Capsella as a model system to study the evolutionary relevance of floral homeotic mutants. Plant Systematics and Evolution, 259: 217235.Google Scholar
Ocarez, N. & Mejía, N. (2016). Suppression of the D-class MADS-box AGL11 gene triggers seedlessness in fleshy fruits. Plant Cell Reports, 35: 239254.Google Scholar
Ochando, I., Jover-Gil, S., Ripoll, J. J. et al. (2006). Mutations in the microRNA complementarity site of the INCURVATA4 gene perturb meristem function and adaxialize lateral organs in Arabidopsis. Plant Physiology, 141: 607619.Google Scholar
Ochiai, T., Nakamura, T., Mashiko, Y. et al. (2004). The differentiation of sepal and petal morphologies in Commelinaceae. Gene, 343: 253262.Google Scholar
Ogura, T. & Busch, W. (2016). Genotypes, networks, phenotypes: moving toward plant systems genetics. Annual Review of Cell and Developmental Biology, 32: 103126.Google Scholar
Ohmori, S., Kimizu, M., Sugita, M. et al. (2009). MOSAIC FLORAL ORGANS1, an AGL6-like MADS box gene, regulates floral organ identity and meristem fate in rice. Plant Cell, 21: 30083025.Google Scholar
Okada, K., Komaki, M. K. & Shimura, Y. (1989). Mutational analysis of pistil structure and development of Arabidopsis thaliana. Cell Differentiation and Development, 28: 2738.Google Scholar
Olmstead, R. G., Bohs, L., Migid, H. A. et al. (2008). A molecular phylogeny of the Solanaceae. Taxon, 57: 11591181.Google Scholar
Olmstead, R. G., Michaels, H. J., Scott, K. M. & Palmer, J. D. (1992). Monophyly of the Asteridae sensu lato and identification of their major lineages inferred from DNA sequences of rbcL. Annals of the Missouri Botanical Garden, 79: 249265.Google Scholar
Olmstead, R. G. & Palmer, J. D. (1992). A chloroplast DNA phylogeny of the Solanaceae: subfamilial relationships and character evolution. Annals of the Missouri Botanical Garden, 79: 346360.Google Scholar
Olmstead, R. G. & Reeves, P. A. (1995). Evidence for the polyphyly of the Scrophulariaceae based on chloroplast rbcL and ndhF sequences. Annals of the Missouri Botanical Garden, 82: 176193.Google Scholar
Olson, M. E. (2003). Ontogenetic origins of floral bilateral symmetry in Moringaceae (Brassicales). American Journal of Botany, 90: 4971.Google Scholar
Orkwiszewski, J. A. & Poethig, R. S. (2000). Phase identity of the maize leaf is determined after leaf initiation. Proceedings of the National Academy of Sciences of the United States of America, 97: 1063110636.Google Scholar
Otsuga, D., DeGuzman, B., Prigge, M. J., Drews, G. N. & Clark, S. E. (2001). REVOLUTA regulates meristem initiation at lateral positions. The Plant Journal, 25: 223236.Google Scholar
Owen, R. (1843). Lectures on the Comparative Anatomy and Physiology of the Invertebrate Animals, Delivered at the Royal College of Surgeons. London: Longman, Brown, Green and Longmans.Google Scholar
Oyama, S. (2000). The Ontogeny of Information, 2nd edn. Durham, NC: Duke University Press.Google Scholar
Ozerova, L. V. & Timonin, A. C. (2009). On the evidence of subunifacial and unifacial leaves: developmental studies in leaf-succulent Senecio L. species (Asteraceae). Wulfenia, 16: 6177.Google Scholar
Pabón-Mora, N. & González, F. (2012). Leaf development, metamorphic heteroblasty and heterophylly in Berberis s.l. (Berberidaceae). Botanical Review, 74: 463489.Google Scholar
Palauqui, J. C. & Laufs, P. (2011). Phyllotaxis: in search of the golden angle. Current Biology,: 21: R502-R504.Google Scholar
Pallakies, H. & Simon, R. (2010). Positional information in plant development. In Encyclopedia of Life Sciences. Chichester: Wiley.Google Scholar
Panero, J. L. & Funk, V. A. (2008). The value of sampling anomalous taxa in phylogenetic studies: major clade of the Asteraceae revealed. Molecular Phylogenetics and Evolution, 47: 757782.Google Scholar
Pang, H.-B., Sun, Q.-W., He, S.-Z. & Wang, Y.-Z. (2010). Expression pattern of CYC-like genes relating to a dorsalized actinomorphic flower in Tengia (Gesneriaceae). Journal of Systematics and Evolution, 48: 309317.Google Scholar
Parcy, F. (2005). Flowering: a time for integration. International Journal of Developmental Biology, 49: 585593.Google Scholar
Parcy, F., Nilsson, O., Busch, M. A., Lee, I. & Weigel, D. (1998). A genetic framework for floral patterning. Nature, 395: 561566.Google Scholar
Park, J.-H., Ishikawa, Y., Ochiai, T., Kanno, A. & Kameya, T. (2004). Two GLOBOSA-like genes are expressed in second and third whorls of homochlamydeous flowers in Asparagus officinalis L. Plant Cell Physiology, 45: 325332.Google Scholar
Park, J.-H., Ishikawa, Y., Yoshida, R., Kanno, A. & Kameya, T. (2003). Expression of AODEF, a B-functional MADS-box gene, in stamens and inner tepals of the dioecious species Asparagus officinalis L. Plant Molecular Biology, 51: 867875.Google Scholar
Paterson, A. H., Bowers, J. E. & Chapman, B. A. (2004). Ancient polyploidization predating divergence of the cereals, and its consequences for comparative genomics. Proceedings of the National Academy of Sciences of the United States of America, 101: 99039908.Google Scholar
Pauw, A. (2005). Inversostyly: a new stylar polymorphism in an oil-secreting plant, Hemimeris racemosa (Scrophulariaceae). American Journal of Botany, 92: 18781886.Google Scholar
Pavlicev, M. & Hansen, T. H. (2011). Genotype-phenotype maps maximizing evolvability: modularity revisited. Evolutionary Biology, 38: 371389.Google Scholar
Peaucelle, A., Louvet, R., Johansen, J. N. et al. (2008). Arabidopsis phyllotaxis is controlled by the methylesterification status of cell-wall pectins. Current Biology, 18: 19431948.Google Scholar
Pelaz, S., Ditta, G. S., Baumann, E., Wisman, E. & Yanofsky, M. F. (2000). B and C floral organ identity functions require SEPALLATA MADS-box genes. Nature, 405: 200203.Google Scholar
Pelaz, S., Gustafson-Brown, C., Kohalmi, S. E., Crosby, W. L. & Yanofsky, M. F. (2001a). APETALA1 and SEPALLATA3 interact to promote flower development. The Plant Journal, 26: 385394.Google Scholar
Pelaz, S., Tapia-Lopez, R., Alvarez-Buylla, E. R. & Yanofsky, M. F. (2001b). Conversion of leaves into petals in Arabidopsis. Current Biology, 11: 182184.Google Scholar
Pellicer, J., Fay, M. F. & Leitch, I. J. (2010). The largest eukaryotic genome of them all? Botanical Journal of the Linnean Society, 164: 1015.Google Scholar
Pennington, R. T., Klitgaard, B. B., Ireland, H. & Lavin, M. (2000). New insights into floral evolution of basal Papilionoideae from molecular phylogenies. In Advances in Legume Systematics, 9, eds. Herendeen, P. S. & Bruneau, A.. Kew: Royal Botanic Gardens, pp. 233248.Google Scholar
Perez-Rodriguez, P., Riano-Pachon, D. M., Correa, L. G. G. et al. (2010). PInTFDB: updated content and new features of the plant transcription factor database. Nucleic Acids Research, 38: D822D827.Google Scholar
Peris, C. I., Rademacher, E. H. & Weijers, D. (2010). Green beginnings: pattern formation in the early plant embryo. Current Topics in Developmental Biology, 91: 127.Google Scholar
Peterson, R. L. (1992). Adaptations of root structure in relation to biotic and abiotic factors. Canadian Journal of Botany, 70: 661675.Google Scholar
Petricka, J. J., Winter, C. M. & Benfey, P. N. (2012). Control of Arabidopsis root development. Annual Review of Plant Biology, 63: 563590.Google Scholar
Pham, T. & Sinha, N. (2003). Role of KNOX genes in shoot development of Welwitschia mirabilis. International Journal of Plant Sciences, 164: 333343.Google Scholar
Piazza, P., Bailey, C. D., Cartolano, M. et al. (2010). Arabidopsis thaliana leaf form evolved via loss of KNOX expression in leaves in association with a selective sweep. Current Biology, 20: 22232228.Google Scholar
Pigliucci, M. (2001). Phenotypic Plasticity: Beyond Nature and Nurture. Baltimore, MD: Johns Hopkins University Press.Google Scholar
Pigliucci, M. (2008). Is evolvability evolvable? Nature Reviews Genetics, 9: 7582.Google Scholar
Pigliucci, M. & Müller, G. (eds.) (2010). Evolution: The Extended Synthesis. Cambridge, MA: MIT Press.Google Scholar
Pigliucci, M. & Murren, C. (2003). Genetic assimilation and a possible evolutionary paradox: can macroevolution sometimes be so fast as to pass us by? Evolution, 57: 14551464.Google Scholar
Pigliucci, M., Murren, C. J. & Schlichting, C. D. (2006). Phenotypic plasticity and evolution by genetic assimilation. Journal of Experimental Biology, 209: 23622367.Google Scholar
Piñeyro-Nelson, A., Almeida, A. M. R., Sass, C., Iles, W. J. D. & Specht, C. D. (2017). Change of fate and staminodial laminarity as potential agents of floral diversification in the Zingiberales. Journal of Experimental Zoology (Molecular and Developmental Evolution), 328: 4154.Google Scholar
Pires, N. D. & Dolan, L. (2012). Morphological evolution in land plants: new designs with old genes. Philosophical Transactions of the Royal Society B, 367: 508518.Google Scholar
Poethig, R. S. (1988). Heterochronic mutations affecting shoot development in maize. Genetics, 119: 959973.Google Scholar
Poethig, R. S. (2009). Small RNAs and developmental timing in plants. Current Opinion in Genetics and Development, 19: 374378.Google Scholar
Poethig, R. S. & Sussex, I. M. (1985). The developmental morphology and growth dynamics of the tobacco leaf. Planta, 165: 158169.Google Scholar
Porras, R. & Muñoz, J. M. (2000). Cleistogamous capitulum in Centaurea melitensis (Asteraceae): heterochronic origin. American Journal of Botany, 87: 925933.Google Scholar
Posé, D., Yant, L. & Schmid, M. (2012). The end of innocence: flowering networks explode in complexity. Current Opinion in Plant Biology, 15: 4550.Google Scholar
Povilus, R. A., Losada, J. M. & Friedman, W. E. (2015). Floral biology and ovule and seed ontogeny of Nymphaea thermarum, a water lily at the brink of extinction with potential as a model system for basal angiosperms. Annals of Botany, 115: 211226.Google Scholar
Powell, A. E. & Lenhard, M. (2012). Control of organ size in plants. Current Biology, 22: R360R367.Google Scholar
Powell, R. (2007). Is convergence more than an analogy? Homoplasy and its implications for macroevolutionary predictability. Biology and Philosophy, 22: 565578.Google Scholar
Pradeu, T. (2012). The Limits of the Self: Immunology and Biological Identity. Oxford: Oxford University Press.Google Scholar
Pradeu, T. (2016). Organisms or biological individuals? Combining physiological and evolutionary individuality. Biology and Philosophy, 31: 797817.Google Scholar
Pradeu, T., Laplane, L., Prévot, K. et al. (2016). Defining ‘development’. Current Topics in Developmental Biology, 117: 171183.Google Scholar
Prenner, G. (2004). Floral development in Polygala myrtifolia (Polygalaceae) and its similarities with Leguminosae. Plant Systematics and Evolution, 249: 6776.Google Scholar
Prenner, G. (2014). Floral ontogeny in Passiflora lobata (Malpighiales, Passifloraceae) reveals a rare pattern in petal formation and provides new evidence for interpretation of the tendril and corona. Plant Systematics and Evolution, 300: 12851297.Google Scholar
Prenner, G. & Klitgaard, B. B. (2008). Towards unlocking the deep nodes of Leguminosae: floral development and morphology of the enigmatic Duparquetia orchidacea (Leguminosae, Caesalpinioideae). American Journal of Botany, 95: 13491365.Google Scholar
Prenner, G. & Rudall, P. J. (2007). Comparative ontogeny of the cyathium in Euphorbia and its allies: exploring the organ-flower-inflorescence boundaries. American Journal of Botany, 94: 16121629.Google Scholar
Preston, J. C. & Hileman, L. C. (2009). Developmental genetics of floral symmetry evolution. Trends in Plant Science, 14: 147154.Google Scholar
Preston, J. C. & Hileman, L. C. (2010). SQUAMOSA-PROMOTER BINDING PROTEIN 1 initiates flowering in Antirrhinum majus through the activation of meristem identity genes. The Plant Journal, 62: 704712.Google Scholar
Preston, J. C. & Hileman, L. C. (2012). Parallel evolution of TCP and B-class genes in Commelinaceae flower bilateral symmetry. Evo Devo, 3: 6.Google Scholar
Preston, J. C., Hileman, L. C. & Cubas, P. (2011a). Reduce, reuse, and recycle: developmental evolution of trait diversification. American Journal of Botany, 98: 397403.Google Scholar
Preston, J. C. & Kellogg, E. A. (2006). Reconstructing the evolutionary history of paralogous APETALA1/FRUITFULL-like genes in grasses (Poaceae). Genetics, 174: 421437.Google Scholar
Preston, J. C. & Kellogg, E. A. (2007). Conservation and divergence of APETALA1/FRUITFULL-like gene function in grasses: evidence from gene expression analyses. The Plant Journal, 52: 6981.Google Scholar
Preston, J. C. & Kellogg, E. A. (2008). Discrete developmental roles for temperate cereal grass VERNALIZATION1/FRUITFULL-like genes in flowering competency and the transition to flowering. Plant Physiology, 146: 265276.Google Scholar
Preston, J. C., Kost, M. A. & Hileman, L. C. (2009). Conservation and diversification of the symmetry developmental program among close relatives of snapdragon with divergent floral morphologies. New Phytologist, 182: 751762.Google Scholar
Preston, J. C., Martinez, C. C. & Hileman, L. C. (2011b). Gradual disintegration of the floral symmetry gene network is implicated in the evolution of a wind-pollination syndrome. Proceedings of the National Academy of Sciences of the United States of America, 108: 23432348.Google Scholar
Prigge, M. J., Otsuga, D., Alonso, J. M. et al. (2005). Class III homeodomain-leucine zipper gene family members have overlapping, antagonistic, and distinct roles in Arabidopsis development. Plant Cell, 17: 6176.Google Scholar
Primack, R. B. (1985). Longevity of individual flowers. Annual Reviews of Ecology and Systematics, 16: 1537.Google Scholar
Pruitt, R. E., Vielle-Calzada, J. P., Ploense, S. E., Grossniklaus, U. & Lolle, S. J. (2000). FIDDLEHEAD, a gene required to suppress epidermal cell interactions in Arabidopsis, encodes a putative lipid biosynthetic enzyme. Proceedings of the National Academy of Sciences of the United States of America, 97: 13111316.Google Scholar
Prusinkiewicz, P. (2004). Modeling plant growth and development. Current Opinion in Plant Biology, 7: 7984.Google Scholar
Prusinkiewicz, P., Erasmus, Y., Lane, B., Harder, J. D. & Coen, E. (2007). Evolution and development of inflorescence architectures. Science, 316: 14521456.Google Scholar
Prusinkiewicz, P. & Lindenmayer, A. (1990). The Algorithmic Beauty of Plants. Berlin: Springer.Google Scholar
Quint, M., Drost, H.-G., Gabel, A. et al. (2012). A transcriptomic hourglass in plant embryogenesis. Nature, 490: 98101.Google Scholar
Quodt, V., Faigl, W., Saedler, H. & Münster, T. (2007). The MADS-domain protein PPM2 preferentially occurs in gametangia and sporophytes of the moss Physcomitrella patens. Gene, 400: 2534.Google Scholar
Raff, R. A. & Kaufman, T. C. (1983). Embryos, Genes, and Evolution. New York, NY: Macmillan.Google Scholar
Rao, N. N., Prasad, K., Kumar, P. R. & Vijayraghavan, U. (2008). Distinct regulatory role for RFL, the rice LFY homolog, in determining flowering time and plant architecture. Proceedings of the National Academy of Sciences of the United States of America, 105: 36463651.Google Scholar
Rasmussen, D. A., Kramer, E. M. & Zimmer, E. A. (2009). One size fits all? Molecular evidence for a commonly inherited petal identity program in Ranunculales. American Journal of Botany, 96: 96109.Google Scholar
Rast, M. I. & Simon, R. (2008). The meristem-to-organ boundary: more than an extremity of anything. Current Opinion in Genetics and Development, 18: 287294.Google Scholar
Raven, J. A. & Edwards, D. (2001). Roots: evolutionary origins and biogeochemical significance. Journal of Experimental Botany, 52: 381401.Google Scholar
Ravi, M. & Chan, S. W. (2010). Haploid plants produced by centromere-mediated genome elimination. Nature, 464: 615618.Google Scholar
Reardon, W., Fitzpatrick, D. A., Fares, M. A. & Nugent, J. M. (2009). Evolution of flower shape in Plantago lanceolata. Plant Molecular Biology, 71: 241250.Google Scholar
Rebocho, A. B., Bliek, M., Kusters, E. et al. (2008). Role of EVERGREEN in the development of the cymose petunia inflorescence. Developmental Cell, 15: 437447.Google Scholar
Ree, R. H. & Donoghue, M. J. (1999). Inferring rates of change in flower symmetry in asterid angiosperms. Systematic Botany, 48: 633641.Google Scholar
Reeves, P. A., He, Y., Schmitz, R. J. et al. (2007). Evolutionary conservation of the FLOWERING LOCUS C mediated vernalization response: evidence from the sugar beet (Beta vulgaris). Genetics, 176: 295307.Google Scholar
Reeves, P. A. & Olmstead, R. G. (1998). Evolution of novel morphological and reproductive traits in a clade containing Antirrhinum majus (Scrophulariaceae). American Journal of Botany, 85: 10471056.Google Scholar
Reilly, S. M. (1997). An integrative approach to heterochrony: the distinction between interspecific and intraspecific phenomena. Biological Journal of the Linnean Society, 60: 119143.Google Scholar
Reinhardt, D., Mandel, T. & Kuhlemeier, C. (2000). Auxin regulates the initiation and radial position of plant lateral organs. Plant Cell, 12: 507518.Google Scholar
Reinhardt, D., Pesce, E. R., Stieger, P. et al. (2003). Regulation of phyllotaxis by polar auxin transport. Nature, 426: 255260.Google Scholar
Reinhart, B. J., Weinstein, E. G., Rhoades, M. W., Bartel, B. & Bartel, D. P. (2002). MicroRNAs in plants. Genes and Development, 16: 16161626.Google Scholar
Reiser, L., Sanchez-Baracaldo, P. & Kake, S. (2000). Knots in the family tree: evolutionary relationships and functions of knox homeobox genes. Plant Molecular Biology, 42: 151166.Google Scholar
Remizowa, M. V., Sokoloff, D. D. & Rudall, P. J. (2010). Evolutionary history of the monocot flower. Annals of the Missouri Botanical Garden, 97: 617645.Google Scholar
Ren, J.-B. & Guo, Y.-P. (2015). Behind the diversity: ontogenies of radiate, disciform, and discoid capitula of Chrysanthemum and its allies. Journal of Systematics and Evolution, 53: 520528.Google Scholar
Ren, Y., Chang, H.-L. & Endress, P. K. (2010). Floral development in Anemoneae (Ranunculaceae). Botanical Journal of the Linnean Society, 162: 77100.Google Scholar
Reski, R. (2003). Physcomitrella patens as a novel tool for plant functional genomics. In Plant Biotechnology 2002 and Beyond, ed. Vasil, I. K.. Dodrecht: Kluwer, pp. 205209.Google Scholar
Reski, R. & Cove, D. J. (2004). Quick guide: Physcomitrella patens. Current Biology, 14: R261R262.Google Scholar
Reut, M. S. & Fineran, B. A. (2000). Ecology and vegetative morphology of the carnivorous plant Utricularia dichotoma (Lentibulariaceae) in New Zealand. New Zealand Journal of Botany, 38: 433450.Google Scholar
Rhoades, M. W., Reinhart, B. J., Lim, L. P. et al. (2002). Prediction of plant microRNA targets. Cell, 110: 513520.Google Scholar
Rice, D. W., Alverson, A. J., Richardson, A. O. et al. (2013). Horizontal transfer of entire genomes via mitochondrial fusion in the angiosperm Amborella. Science, 342: 14681473.Google Scholar
Richards, A. J. (1997). Plant Breeding Systems, 2nd edn. London: Chapman & Hall.Google Scholar
Richardson, A. O. & Palmer, J. D. (2007). Horizontal gene transfer in plants. Journal of Experimental Botany, 58: 19.Google Scholar
Rigato, E. & Minelli, A. (2013). The great chain of being is still here. Evolution: Education and Outreach, 6: 18.Google Scholar
Rijpkema, A. S., Royaert, S., Zethof, J. et al. (2006). Analysis of the Petunia TM6 MADS box gene reveals functional divergence within the DEF/AP3 lineage. Plant Cell, 18: 18191832.Google Scholar
Rijpkema, A. S., Vandenbussche, M., Koes, R., Heijmans, K. & Gerats, T. (2010). Variations on a theme: changes in the floral ABCs in angiosperms. Seminars in Cell and Developmental Biology, 21: 100107.Google Scholar
Rijpkema, A. S., Zethof, J., Gerats, T. & Vandenbussche, M. (2009). The petunia AGL6 gene has a SEPALLATA-like function in floral patterning. The Plant Journal, 60: 19.Google Scholar
Roberts, J. A., Elliott, K. A. & González-Carranza, Z. H. (2002). Abscission, dehiscence, and other cell separation processes. Annual Review of Plant Biology, 53: 131158.Google Scholar
Roberts, J. A. & González-Carranza, Z. H. (2013). Abscission. In Encyclopedia of Life Sciences. Chichester: Wiley.Google Scholar
Robinson, B. W. & Dukas, R. (1999). The influence of phenotypic modifications on evolution: the Baldwin effect and modern perspectives. Oikos, 85: 528589.Google Scholar
Roeder, A. H. K. (2010). Sepals. In Encyclopedia of Life Sciences. Chichester: Wiley.Google Scholar
Roeder, A. H. K., Ferrándiz, C. & Yanofsky, M. F. (2003). The role of the REPLUMLESS homeodomain protein in patterning the Arabidopsis fruit. Current Biology, 13: 16301635.Google Scholar
Roeder, A. H. K. & Yanofsky, M. F. (2006). Fruit development in Arabidopsis. In The Arabidopsis Book. Rockville, MD: American Society of Plant Biologists, 4: e0075.Google Scholar
Rohde, A. & Bhalerao, R. P. (2007). Plant dormancy in the perennial context. Trends in Plant Science, 12: 217223.Google Scholar
Rolland-Lagan, A.-G., Bangham, J. A. & Coen, E. (2003). Growth dynamics underlying petal shape and asymmetry. Nature, 422: 161163.Google Scholar
Ronse De Craene, L. P. (2003). The evolutionary significance of homeosis in flowers: a morphological perspective. International Journal of Plant Sciences, 164: S225S235.Google Scholar
Ronse De Craene, L. P. (2004). Floral development of Berberidopsis corallina: a crucial link in the evolution of flowers in the core eudicots. Annals of Botany, 94: 111.Google Scholar
Ronse De Craene, L. P. (2007). Are petals sterile stamens or bracts? The origin and evolution of petals in the core eudicots. Annals of Botany, 100: 621630.Google Scholar
Ronse De Craene, L. P. (2008). Homology and evolution of petals in the core eudicots. Systematic Botany, 33: 301325.Google Scholar
Ronse De Craene, L. P. (2010). Floral Diagrams: An Aid to Understanding Flower Morphology and Evolution. Cambridge: Cambridge University Press.Google Scholar
Ronse De Craene, L. P. (2011) Floral development of Napoleonaea (Lecythidaceae), a deceptively complex flower. In Flowers on the Tree of Life, ed. Wanntorp, L. & Ronse De Craene, L. P.. Cambridge: Cambridge University Press, pp. 279295.Google Scholar
Ronse De Craene, L. P. (2013). Reevaluation of the perianth and androecium in Caryophyllales: implications for flower evolution. Plant Systematics and Evolution, 299: 15991636.Google Scholar
Ronse De Craene, L. P. (2016). Meristic changes in flowering plants: how flowers play with numbers. Flora, 221: 2237.Google Scholar
Ronse De Craene, L. P. (2017). Floral development of Berberidopsis beckleri – can an additional species of the Berberidopsidaceae add evidence to floral evolution in the core eudicots? Annals of Botany, 119: 599610.Google Scholar
Ronse De Craene, L. P. & Brockington, S. F. (2013). Origin and evolution of petals in angiosperms. Plant Ecology and Evolution, 146: 525.Google Scholar
Ronse De Craene, L. P., Linder, H. P. & Smets, E. F. (2002). Ontogeny and evolution of the flower of South African Restionaceae with special emphasis on the gynoecium. Plant Systematics and Evolution, 231: 225258.Google Scholar
Ronse De Craene, L. P. & Smets, E. F. (1990). The floral development of Popowia whitei (Annonaceae). Nordic Journal of Botany, 10: 411420. (Correction: Nordic Journal of Botany, 11: 420 (1991)).Google Scholar
Ronse De Craene, L. P. & Smets, E. F. (1994). Merosity in flowers: definition, origin, and taxonomic significance. Plant Systematics and Evolution, 191: 83104.Google Scholar
Ronse De Craene, L. P. & Smets, E. F. (1995). Evolution of the androecium in the Ranunculiflorae. Plant Systematics and Evolution, Supplement, 9: 6370.Google Scholar
Ronse De Craene, L. P. & Smets, E. F. (1998). Meristic changes in gynoecium morphology, exemplified by floral ontogeny and anatomy. In Reproductive Biology in Systematics, Conservation and Economic Botany, eds. Owens, S. J. & Rudall, P. J.. Kew: Royal Botanic Gardens, pp. 85112.Google Scholar
Ronse De Craene, L. P. & Smets, E. F. (2000). Floral development of Galopina tomentosa with a discussion of sympetaly and placentation in the Rubiaceae. Systematics and Geography of Plants, 70: 155170.Google Scholar
Ronse De Craene, L. P. & Smets, E. F. (2001). Staminodes: their morphological and evolutionary significance. Botanical Review, 67: 351402.Google Scholar
Ronse De Craene, L. P., Soltis, P. S. & Soltis, D. E. (2003). Evolution of floral structure in basal angiosperms. International Journal of Plant Sciences, 164: S329S363.Google Scholar
Ronse De Craene, L. P. & Wanntorp, L. (2008). Morphology and anatomy of the flower of Meliosma (Sabiaceae): implications for pollination biology. Plant Systematics and Evolution, 271: 7991.Google Scholar
Roquet, C., Coissac, E., Cruaud, C. et al. (2016). Understanding the evolution of holoparasitic plants: the complete plastid genome of the holoparasite Cytinus hypocistis (Cytinaceae). Annals of Botany, 118: 885896.Google Scholar
Rosin, F. M. & Kramer, E. M. (2009). Old dogs, new tricks: regulatory evolution in conserved genetic modules leads to novel morphologies in plants. Developmental Biology, 332: 2535.Google Scholar
Rudall, P. J. (2008). Fascicles and filamentous structures: comparative ontogeny of morphological novelties in the mycoheterotrophic family Triuridaceae. International Journal of Plant Sciences, 169: 10231037.Google Scholar
Rudall, P. J. (2010). All in a spin: centrifugal organ formation and floral patterning. Current Opinion in Plant Biology, 13: 108114.Google Scholar
Rudall, P. J. (2013). Identifying key features in the origin and early diversification of angiosperms. Annual Plant Reviews, 45: 163188.Google Scholar
Rudall, P. J., Alves, M. & das Graças Sajo, M. (2016). Inside-out flowers of Lacandonia brasiliana (Triuridaceae) provide new insights into fundamental aspects of floral patterning. PeerJ, 4: e1653.Google Scholar
Rudall, P. J. & Bateman, R. M. (2002). Roles of synorganisation, zygomorphy and heterotopy in floral evolution: the gynostemium and labellum of orchids and other lilioid monocots. Biology Reviews, 77: 403441.Google Scholar
Rudall, P. J. & Bateman, R. M. (2003). Evolutionary change in flowers and inflorescences: evidence from naturally occurring terata. Trends in Plant Science, 8: 7682.Google Scholar
Rudall, P. J. & Bateman, R. M. (2004). Evolution of zygomorphy in monocot flowers: iterative patterns and developmental constraints. New Phytologist, 162: 2544.Google Scholar
Rudall, P. J. & Bateman, R. M. (2006). Morphological phylogenetic analysis of Pandanales: testing contrasting hypotheses of floral evolution. Systematic Botany, 31: 223238.Google Scholar
Rudall, P. J. & Bateman, R. M. (2010). Defining the limits of flowers: the challenge of distinguishing between the evolutionary products of simple versus compound strobili. Philosophical Transactions of the Royal Society B, 365: 397409.Google Scholar
Rudall, P. J. & Buzgo, M. (2002). Evolutionary history of the monocot leaf. In Developmental Genetics and Plant Evolution, eds. Cronk, Q. C. B., Bateman, R. M. & Hawkins, J. A.. London: Taylor & Francis, pp. 431458.Google Scholar
Rudall, P. J., Cunniff, J., Wilkin, P. & Caddick, L. R. (2005a). Evolution of dimery, pentamery and the monocarpellary condition in the monocot family Stemonaceae (Pandanales). Taxon, 54: 701711.Google Scholar
Rudall, P. J., Remizowa, M. V., Prenner, G. et al. (2009). Non-flowers near the base of extant angiosperms? Spatiotemporal arrangement of organs in reproductive units of Hydatellaceae, and its bearing on the origin of the flower. American Journal of Botany, 96: 6782.Google Scholar
Rudall, P. J., Sokoloff, D. D., Remizowa, M. V. et al. (2007). Morphology of Hydatellaceae, an anomalous aquatic family recently recognized as an early-divergent angiosperm lineage. American Journal of Botany, 94: 10731092.Google Scholar
Rudall, P. J., Stuppy, W., Cunniff, J., Kellogg, E. A. & Briggs, B. G. (2005b). Evolution of reproductive structures in grasses (Poaceae) inferred by sister-group comparison with their putative closest living relatives, Ecdeiocoleaceae. American Journal of Botany, 92: 14321443.Google Scholar
Ruiz de Clavijo, E. (1994). Heterocarpy and seed polymorphism in Ceratocapnos heterocarpa (Fumariaceae). International Journal of Plant Sciences, 155: 196202.Google Scholar
Ruiz-Sanchez, E. & Sosa, V. (2015). Origin and evolution of fleshy fruit in woody bamboos. Molecular Phylogenetics and Evolution, 91: 123134.Google Scholar
Ruskin, J. (1900). Modern Painters. New York, NY: The Kelmscott Society.Google Scholar
Rutishauser, R. (1981). Blattstellung und Sprossentwicklung bei Blütenpflanzen unter besonderer Berücksichtigung der Nelkengewächse (Caryophyllaceen s.l.). Dissertationes Botanicae, 62: 1165.Google Scholar
Rutishauser, R. (1984). Blattquirle, Stipeln und Kolleteren bei den Rubieae (Rubiaceae) im Vergleich mit anderen Angiospermen. Beiträge zur Biologie der Pflanzen, 59: 375424.Google Scholar
Rutishauser, R. (1995). Developmental patterns of leaves in Podostemonaceae as compared to more typical flowering plants: saltational evolution and fuzzy morphology. Canadian Journal of Botany, 73: 13051317.Google Scholar
Rutishauser, R. (1997). Structural and developmental diversity in Podostemaceae (river-weeds). Aquatic Botany, 57: 2970.Google Scholar
Rutishauser, R. (1998). Plastochrone ratio and leaf arc as parameters of a quantitative phyllotaxis analysis in vascular plants. In Symmetry in Plants, eds. Jean, R.V. & Barabé, D.. Singapore: World Scientific, pp. 171212.Google Scholar
Rutishauser, R. (1999). Polymerous leaf whorls in vascular plants: developmental morphology and fuzziness of organ identity. International Journal of Plant Sciences, 160: S81S103.Google Scholar
Rutishauser, R. (2016a). Evolution of unusual morphologies in Lentibulariaceae (bladderworts and allies) and Podostemaceae (river-weeds): a pictorial report at the interface of developmental biology and morphological diversification. Annals of Botany, 117: 811832.Google Scholar
Rutishauser, R. (2016b). Acacia (wattle) and Cananga (ylang-ylang): from spiral to whorled and irregular (chaotic) phyllotactic patterns – a pictorial report. Acta Societatis Botanicorum Poloniae, 85 (4): 3531.Google Scholar
Rutishauser, R., Grob, V. & Pfeifer, E. (2008). Plants are used to having identity crises. In Evolving Pathways. Key Themes in Evolutionary Developmental Biology, eds. Minelli, A. & Fusco, G.. Cambridge: Cambridge University Press, pp. 194213.Google Scholar
Rutishauser, R. & Grubert, M. (1999). The architecture of Mourera fluviatilis (Podostemaceae). Developmental morphology of inflorescences, flowers, and seedlings. American Journal of Botany, 86: 907922.Google Scholar
Rutishauser, R. & Huber, K. A. (1991). The developmental morphology of Indotristicha ramosissima (Podostemaceae, Tristichoideae). Plant Systematics and Evolution, 178: 195223.Google Scholar
Rutishauser, R. & Isler, B. (2001). Developmental genetics and morphological evolution of flowering plants, especially bladderworts (Utricularia): fuzzy Arberian morphology complements classical morphology. Annals of Botany, 88: 11731202.Google Scholar
Rutishauser, R. & Moline, P. (2005). Evo-devo and the search for homology (‘sameness’) in biological systems. Theory in Biosciences, 124: 213241.Google Scholar
Rutishauser, R., Pfeifer, E., Moline, P. & Philbrick, C. T. (2003). Developmental morphology of roots and shoots of Podostemum ceratophyllum (Podostemaceae-Podostemoideae). Rhodora, 105: 337353.Google Scholar
Rutishauser, R., Ronse De Craene, L. P., Smets, E. & Mendoza-Heuer, I. (1998). Theligonum cynocrambe: developmental morphology of a peculiar rubaceous herb. Plant Systematics and Evolution, 210: 124.Google Scholar
Rutishauser, R. & Sattler, R. (1986). Architecture and development of the phyllode–stipule whorls in Acacia longipedunculata: controversial interpretations and continuum approach. Canadian Journal of Botany, 64: 19872019.Google Scholar
Rutishauser, R. & Sattler, R. (1997). Expression of shoot processes in leaf development of Polemonium caeruleum as compared to other dicotyledons. Botanische Jahrbücher für Systematik, Pflanzengeschichte und Pflanzengeographie, 119: 563582.Google Scholar
Saddic, L. A., Huvermann, B., Bezhani, S. et al. (2006). The LEAFY target LMI1 is a meristem identity regulator and acts together with LEAFY to regulate expression of CAULIFLOWER. Development, 133: 16731682.Google Scholar
Sajo, M. G., Mello-Silva, R. & Rudall, P. J. (2010). Homologies of floral structures in Velloziaceae, with particular reference to the corona. International Journal of Plant Sciences, 171: 595606.Google Scholar
Sakai, H., Medrano, L. J. & Meyerowitz, E. M. (1995). Role of SUPERMAN in maintaining Arabidopsis floral whorl boundaries. Nature, 378: 199203.Google Scholar
Sakakibara, K., Ando, S., Yip, H. K. et al. (2013). KNOX2 genes regulate the haploid-to-diploid morphological transition in land plants. Science, 339: 10671070.Google Scholar
Salomé, P. A., Bomblies, K., Laitinen, R. A. et al. (2011). Genetic architecture of flowering-time variation in Arabidopsis thaliana. Genetics, 188: 421433.Google Scholar
Sander, K. (1983). The evolution of patterning mechanisms: gleanings from insect embryogenesis and spermatogenesis. In Development and Evolution: the Sixth Symposium of the British Society for Developmental Biology, eds. Goodwin, B. C., Holder, N. & Wylie, C. C.. Cambridge: Cambridge University Press, pp. 137160.Google Scholar
Sarojam, R., Sappl, P. G., Goldshmidt, A. et al. (2010). Differentiating Arabidopsis shoots from leaves by combined YABBY activities. Plant Cell, 22: 21132130.Google Scholar
Sato, S., Nakamura, Y., Kaneko, T. et al. (2008). Genome structure of the legume, Lotus japonicus. DNA Research, 15: 227239.Google Scholar
Sattler, R. (1972). Centrifugal primordial inception in floral development. Advances in Plant Morphology, 1972: 170178.Google Scholar
Sattler, R. (1992). Process morphology: structural dynamics in development and evolution. Canadian Journal of Botany, 70: 708714.Google Scholar
Sattler, R. (1994). Homology, homeosis, and process morphology in plants. In Homology: The Hierarchical Basis of Comparative Biology, ed. Hall, B. K.. London: Academic Press, pp. 423475.Google Scholar
Sattler, R. (1996). Classical morphology and continuum morphology: opposition and continuum. Annals of Botany, 78: 577581.Google Scholar
Sattler, R. & Jeune, B. (1992). Multivariate analysis confirms the continuum view of plant form. Annals of Botany, 69: 249262.Google Scholar
Sattler, R. & Rutishauser, R. (1990). Structural and dynamic descriptions of the development of Utricularia foliosa and U. australis. Canadian Journal of Botany, 68: 19892003.Google Scholar
Sattler, R. & Rutishauser, R. (1992). Partial homology of pinnate leaves and shoots: orientation of leaflet inception. Botanische Jahrbücher für Systematik, Pflanzengeschichte und Pflanzengeographie, 114: 6179.Google Scholar
Sattler, R. & Rutishauser, R. (1997). The fundamental relevance of morphology and morphogenesis to plant research. Annals of Botany, 80: 571582.Google Scholar
Saunders, R. M. K. (2010). Floral evolution in the Annonaceae: hypotheses of homeotic mutations and functional convergence. Biological Reviews, 85: 571591.Google Scholar
Savage, A. J. P. & Ashton, P. S. (1983). The population structure of the double coconut and some other Seychelles palms. Biotropica, 15: 1525.Google Scholar
Sawa, S., Ito, T., Shimura, Y. & Okada, K. (1999b). FILAMENTOUS FLOWER controls the formation and development of Arabidopsis inflorescences and floral meristems. Plant Cell, 11: 6986.Google Scholar
Sawa, S., Watanabe, K., Goto, K. et al. (1999a). FILAMENTOUS FLOWER, a meristem and organ identity gene of Arabidopsis, encodes a protein with a zinc finger and HMG-related domains. Genes and Development, 13: 10791088.Google Scholar
Scanlon, M. J. (2000). Developmental complexities of simple leaves. Current Opinion in Plant Biology, 3: 3136.Google Scholar
Scarpella, E., Barkoulas, M. & Tsiantis, M. (2010). Control of leaf and vein development by auxin. Cold Spring Harbor Perspectives in Biology, 2: a001511.Google Scholar
Scarpella, E., Marcos, D., Friml, J. & Berleth, T. (2006). Control of leaf vascular patterning by polar auxin transport. Genes and Development, 20: 10151027.Google Scholar
Schäferhoff, B., Fleischmann, A., Fischer, E. et al. (2010). Towards resolving Lamiales relationships: insights from rapidly evolving chloroplast sequences. BMC Evolutionary Biology, 10: 352.Google Scholar
Scheres, B., Wolkenfelt, H., Willemsen, V. et al. (1994). Embryonic origin of the Arabidopsis primary root and root meristem initials. Development, 120: 24752487.Google Scholar
Schlichting, C. D. & Murren, C. J. (2004). Evolvability and the raw materials for adaptation. In Plant Adaptation: Molecular Genetics and Ecology, eds. Cronk, Q. C. B., Whitton, J., Ree, R. H. & Taylor, I. E. P.. Ottawa: NRC Research Press, pp. 1829.Google Scholar
Schlichting, C. D. & Pigliucci, M. (1998). Phenotypic Evolution: A Reaction Norm Perspective. Sunderland, MA: Sinauer Associates.Google Scholar
Schlichting, C. D. & Wund, M. A. (2014). Phenotypic plasticity and epigenetic marking: an assessment of evidence for genetic accommodation. Evolution, 68: 656672.Google Scholar
Schmid, M., Davison, T. S., Henz, S. R. et al. (2005). A gene expression map of Arabidopsis thaliana development. Nature Genetics, 37: 501506.Google Scholar
Schmuths, H., Meister, A., Horres, R. & Bachmann, K. (2004). Genome size variation among accessions of Arabidopsis thaliana. Annals of Botany, 93: 317321.Google Scholar
Schneider, H., Pryer, K. M., Cranfill, R., Smith, A. R. & Wolf, P. G. (2002). Evolution of vascular plant body plans: a phylogenetic perspective. In Developmental Genetics and Plant Evolution, eds. Cronk, Q. C. B., Bateman, R. M. & Hawkins, J. A.. London: Taylor & Francis, pp. 114.Google Scholar
Schneitz, K. & Balasubramanian, S. (2009). Floral meristems. In Encyclopedia of Life Sciences. Chichester: Wiley.Google Scholar
Schoen, D. J., Johnston, M. O., L’Heureux, A. M. & Marsolais, J. V. (1997). Evolutionary history of the mating system in Amsinckia (Boraginaceae). Evolution, 51: 10901099.Google Scholar
Schönenberger, J., Anderberg, A. A. & Sytsma, K. J. (2005). Molecular phylogenetics and patterns of floral evolution in the Ericales. International Journal of Plant Sciences, 166: 265288.Google Scholar
Schoof, H., Lenhard, M., Haeker, A. et al. (2000). The stem cell population of Arabidopsis shoot meristems is maintained by a regulatory loop between the CLAVATA and WUSCHEL genes. Cell, 100: 635644.Google Scholar
Schranz, M. E., Quijada, P., Sung, S. B. et al. (2002). Characterization and effects of the replicated flowering time gene FLC in Brassica rapa. Genetics, 162: 14571468.Google Scholar
Schultz, E. A. & Haughn, G. W. (1993). Genetic analysis of the floral initiation process (FLIP) in Arabidopsis. Development, 119: 745765.Google Scholar
Schumacher, K., Schmitt, T., Rossberg, M., Schmitz, C. & Theres, K. (1999). The lateral suppressor (ls) gene of tomato encodes a new member of the vhiid protein family. Proceedings of the National Academy of Sciences of the United States of America, 96: 290295.Google Scholar
Schwander, T. & Leimar, O. (2011). Genes as leaders and followers in evolution. Trends in Ecology and Evolution, 26: 143151.Google Scholar
Schwarz, S., Grande, A. V., Bujdoso, N., Saedler, H. & Huijser, P. (2008). The microRNA regulated SBP-box genes SPL9 and SPL15 control shoot maturation in Arabidopsis. Plant Molecular Biology, 67: 183195.Google Scholar
Schwarz-Sommer, Z., Davies, B. & Hudson, A., (2003). An everlasting pioneer: the story of Antirrhinum research. Nature Reviews Genetics, 4: 657666.Google Scholar
Schwarz-Sommer, Z., Huijser, P., Nacken, W., Saedler, H. & Sommer, H. (1990). Genetic control of flower development by homeotic genes in Antirrhinum majus. Science, 250: 931936.Google Scholar
Scofield, S. & Murray, J. A. (2006). KNOX gene function in plant stem cell niches. Plant Molecular Biology, 60: 929946.Google Scholar
Scribailo, R. W. & Tomlinson, P. B. (1992). Shoot and floral development in Calla palustris (Araceae-Calloideae). International Journal of Plant Sciences, 153: 113.Google Scholar
Searle, I., He, Y., Turck, F. et al. (2006). The transcription factor FLC confers a flowering response to vernalization by repressing meristem competence and systemic signaling in Arabidopsis. Genes and Development, 20: 898912.Google Scholar
Seifriz, W. (1950). Gregarious flowering of Chusquea. Nature, 165: 635636.Google Scholar
Semiarti, E., Ueno, Y., Tsukaya, H. et al. (2001). The ASYMMETRIC LEAVES2 gene of Arabidopsis thaliana regulates formation of a symmetric lamina, establishment of venation and repression of meristem-related homeobox genes in leaves. Development, 128: 17711783.Google Scholar
Sentoku, N., Sato, Y., Kurata, N. et al. (1999). Regional expression of the rice KN1-type homeobox gene family during embryo, shoot, and flower development. Plant Cell, 11: 16511663.Google Scholar
Seymour, G. B., Østergaard, L., Chapman, N. H., Knapp, S. & Martin, C. (2013). Fruit development and ripening. Annual Review of Plant Biology, 64: 219241.Google Scholar
Seymour, G. B., Ryder, C. D., Cevik, V. et al. (2011). A SEPALLATA gene is involved in the development and ripening of strawberry (Fragaria × ananassa Duch.) fruit, a non-climacteric tissue. Journal of Experimental Botany, 62: 11791188.Google Scholar
Shah, J. J. & Dave, Y. S. (1970). Tendrils of Passiflora foetida: histogenesis and morphology. American Journal of Botany, 57: 786793.Google Scholar
Shan, H., Su, K., Lu, W. et al. (2006). Conservation and divergence of candidate class B genes in Akebia trifoliata (Lardizabalaceae). Development Genes and Evolution, 216: 785795.Google Scholar
Shannon, S. & Meeks-Wagner, D. R. (1991). A mutation in the Arabidopsis Tfl1 gene affects inflorescence meristem development. Plant Cell, 3: 877892.Google Scholar
Sharma, P. P., Clouse, R. M. & Wheeler, W. C. (2017). Hennig’s semaphoront concept and the use of ontogenetic stages in phylogenetic reconstruction. Cladistics, 33: 93108.Google Scholar
Sheffield, E. & Bell, P. R. (1987). Current studies of the pteridophyte life cycle. Botanical Reviews, 53: 442490.Google Scholar
Sheldon, C. C., Burn, J. E., Perez, P. P. et al. (1999). The FLF MADS box gene: a repressor of flowering in Arabidopsis regulated by vernalization and methylation. Plant Cell, 11: 445458.Google Scholar
Shepard, K. A. & Purugganan, M. D. (2002). The genetics of plant morphological evolution. Current Opinion in Plant Biology, 5: 4955.Google Scholar
Shitsukawa, N., Ikari, C., Shimada, S. et al. (2007). The einkorn wheat (Triticum monococcum) mutant, maintained vegetative phase, is caused by a deletion in the VRN1 gene. Genes and Genetic Systems, 82: 167170.Google Scholar
Siddall, M. E. & Borda, E. (2003). Phylogeny and revision of the leech genus Helobdella (Glossiphoniidae) based on mitochondrial gene sequences and morphological data and a special consideration of the triserialis complex. Zoologica Scripta, 32: 2333.Google Scholar
Sieber, P., Wellmer, F., Gheyselinck, J., Riechmann, J. L. & Meyerowitz, E. M. (2007). Redundancy and specialization among plant microRNAs: role of the MIR164 family in developmental robustness. Development, 134: 10511060.Google Scholar
Siegfried, K. R., Eshed, Y., Baum, S. F. et al. (1999). Members of the YABBY gene family specify abaxial cell fate in Arabidopsis. Development, 126: 41174128.Google Scholar
Simon, R., Carpenter, R., Doyle, S. & Coen, E. (1994). Fimbriata controls flower development by mediating between meristem and organ identity genes. Cell, 78: 99107.Google Scholar
Singer, S. D., Krogan, N. T. & Ashton, N. W. (2007). Clues about the ancestral roles of plant MADS-box genes from a functional analysis of moss homologues. Plant Cell Reports, 26: 11551169.Google Scholar
Sinha, N. (1999). Leaf development in angiosperms. Annual Review of Plant Physiology and Plant Molecular Biology, 50: 419446.Google Scholar
Skippington, E., Barkman, T. J., Rice, D. W. & Palmer, J. D. (2015). Miniaturized mitogenome of the parasitic plant Viscum scurruloideum is extremely divergent and dynamic and has lost all nad genes. Proceedings of the National Academy of Sciences of the United States of America, 112: E3515E3524.Google Scholar
Slotte, T., Hazzouri, K. M., Agren, J. A. et al. (2013). The Capsella rubella genome and the genomic consequences of rapid mating system evolution. Nature Genetics, 45: 831835.Google Scholar
Smaczniak, C., Immink, R. G. H., Angenent, G. C. & Kaufmann, K. (2012). Developmental and evolutionary diversity of plant MADS-domain factors: insights from recent studies. Development, 139: 30813098.Google Scholar
Smith, J. F., Brown, K. D., Carroll, C. L. & Denton, D. S. (1997). Familial placement of Cyrtandromoea, Titanotrichum and Sanango, three problematic genera of the Lamiales. Taxon, 46: 6574.Google Scholar
Smith, J. F., Hileman, L. C., Powell, M. P. & Baum, D. A. (2004). Evolution of GCYC, a Gesneriaceae homolog of CYCLOIDEA, within Gesnerioideae (Gesneriaceae). Molecular Phylogenetics and Evolution, 31: 765779.Google Scholar
Smith, K. K. (2001). Heterochrony revisited: the evolution of developmental sequences. Biological Journal of the Linnean Society, 73: 169186.Google Scholar
Smith, R.S., Guyomarc’h, S., Mandel, T. et al. (2006). A plausible model of phyllotaxis. Proceedings of the National Academy of Sciences of the United States of America, 103: 13011306.Google Scholar
Smith, Z. R. & Long, J. A. (2010). Control of Arabidopsis apical–basal embryo polarity by antagonistic transcription factors. Nature, 464: 423426.Google Scholar
Smýkal, P., Aubert, G., Burstin, J. et al. (2012). Pea (Pisum sativum L.) in the genomic era. Agronomy, 2: 74115.Google Scholar
Smyth, D. R., Bowman, J. L. & Meyerowitz, E. M. (1990). Early flower development in Arabidopsis. Plant Cell, 2: 755768.Google Scholar
Sokoloff, D. D., Rudall, P. J. & Remizowa, M. (2006). Flower-like terminal structures in racemose inflorescences: a tool in morphogenetic and evolutionary research. Journal of Experimental Botany, 57: 35173530.Google Scholar
Sokoloff, D. D., Sokolski, A. A., Remizowa, M. V. & Nuraliev, M. S. (2007). Flower structure and development in Tupidanthus calyptratus (Araliaceae): an extreme case of polymery among asterids. Plant Systematics and Evolution, 268: 209234.Google Scholar
Soltis, D. E. (2007). Saxifragaceae. In The Families and Genera of Vascular Plants, Vol. 9, ed. Kubitzki, K.. Berlin: Springer, pp. 418435.Google Scholar
Soltis, D. E., Albert, V. A., Leebens-Mack, J. et al. (2009). Polyploidy and angiosperm diversification. American Journal of Botany, 96: 336348.Google Scholar
Soltis, D. E., Ma, H., Frohlich, M. W. et al. (2007). The floral genome: an evolutionary history of gene duplication and shifting patterns of gene expression. Trends in Plant Science, 12: 358367.Google Scholar
Soltis, D. E., Senters, A. E., Zanis, M. J. et al. (2003). Gunnerales are sister to other core eudicots: implications for the evolution of pentamery. American Journal of Botany, 90: 461470.Google Scholar
Soltis, D. E., Soltis, P. S., Endress, P. K. & Chase, M. W. (2005). Phylogeny and Evolution of Angiosperms. Sunderland, MA: Sinauer.Google Scholar
Soltis, P. S., Soltis, D. E., Kim, S., Chanderbali, A. & Buzgo, M. (2006). Expression of floral regulators in basal angiosperms and the origin and evolution of ABC function. In Developmental Genetics of the Flower, eds. Soltis, D. E., Leebens-Mack, J. H. & Soltis, P. S.. San Diego, CA: Elsevier, pp. 483506.Google Scholar
Somerville, C. R. & Meyerowitz, E. M. (eds.) (2002–) The Arabidopsis Book. Rockville, MD: American Society of Plant Biologists.Google Scholar
Song, Y. H., Ito, S. & Imaizumi, T. (2013). Flowering time regulation: photoperiod- and temperature-sensing in leaves. Trends in Plant Science, 18: 575583.Google Scholar
Souer, E., Rebocho, A. B., Bliek, M. et al. (2008). Patterning of inflorescences and flowers by the F-box protein DOUBLE TOP and the LEAFY homolog ABERRANT LEAF AND FLOWER in Petunia. Plant Cell, 20: 20332048.Google Scholar
Souer, E., van der Krol, A., Kloos, D. et al. (1998). Genetic control of branching pattern and floral identity during Petunia inflorescence development. Development, 125: 733742.Google Scholar
Sousa-Baena, M. S., Lohmann, L. G., Rossi, M. & Sinha, N. R. (2014a). Acquisition and diversification of tendrilled leaves in Bignonieae (Bignoniaceae) involved changes in expression patterns of SHOOTMERISTEMLESS (STM), LEAFY/FLORICAULA (LFY/FLO), and PHANTASTICA (PHAN). New Phytologist, 201: 9931008.Google Scholar
Sousa-Baena, M. S., Sinha, N. R. & Lohmann, L. G. (2014b). Evolution and development of tendrils in Bignonieae (Lamiales, Bignoniaceae). Annals of the Missouri Botanical Garden, 99: 323347.Google Scholar
Specht, C. D. & Bartlett, M. E. (2009). Flower evolution: the origin and subsequent diversification of the angiosperm flower. Annual Reviews in Ecology, Evolution and Systematics, 40: 217243.Google Scholar
Specht, C. D., Yockteng, R., Almeida, A. M., Kirchoff, B. K. & Kress, W. J. (2012). Homoplasy, pollination, and emerging complexity during the evolution of floral development in the tropical gingers (Zingiberales). Botanical Review, 78: 440462.Google Scholar
Stahle, M. I., Kuehlich, J., Staron, L., von Arnim, A. G. & Golz, J. F. (2009). YABBYs and the transcriptional corepressors LEUNIG and LEUNIG_HOMOLOG maintain leaf polarity and meristem activity in Arabidopsis. Plant Cell, 21: 31053118.Google Scholar
Stebbins, G. L. (1974). Flowering Plants. Evolution Above the Species Level. Cambridge, MA: Belknap Press.Google Scholar
Steeves, T. A. & Sussex, I. M. (1989). Patterns in Plant Development, 2nd edn. Cambridge: Cambridge University Press.Google Scholar
Steingraeber, D. A. & Fisher, J. B. (1986). Indeterminate growth of leaves in Guarea (Meliaceae): a twig analogue. American Journal of Botany, 73: 852862.Google Scholar
Stevens, P. F. (1975). Review of Chisocheton (Meliaceae) in Papuasia. Contributions from Herbarium Australiense, 11: 155.Google Scholar
Strable, J. & Scanlon, M. J. (2009). Maize (Zea mays): a model organism for basic and applied research in plant biology. Cold Spring Harbor Protocols, 2009 (10): pdb.emo132.Google Scholar
Stuessy, T. F. & Urtubey, E. (2006). Phylogenetic implications of corolla morphology in subfamily Barnadesioideae (Asteraceae). Flora, 201: 340352.Google Scholar
Suárez-Baron, H., Pérez-Mesa, P., Ambrose, B. A., González, F. & Pabón-Mora, N. (2017). Deep into the aristolochia flower: expression of C, D, and E-class genes in Aristolochia fimbriata (Aristolochiaceae). Journal of Experimental Zoology (Molecular and Developmental Evolution), 328B: 5571.Google Scholar
Sulman, J. D., Drew, B. T., Drummond, C., Hayasaka, E. & Sytsma, K. J. (2013). Systematics, biogeography, and character evolution of Sparganium (Typhaceae): diversification of a widespread, aquatic lineage. American Journal of Botany, 100: 20232039.Google Scholar
Sun, G., Dilcher, D. L., Zheng, S. & Zhou, Z. (1998). In search of the first flower: a Jurassic angiosperm, Archaefructus, from northeast China. Science, 282: 16921695.Google Scholar
Sun, G., Ji, Q., Dilcher, D. L. et al. (2002). Archaefructaceae, a new basal angiosperm family. Science, 296: 899904.Google Scholar
Szymkowiak, E. J. & Sussex, I. M. (1996). What chimeras can tell us about plant development. Annual Review of Plant Physiology and Plant Molecular Biology, 47: 351376.Google Scholar
Tadege, M., Sheldon, C. C., Helliwell, C. A. et al. (2001). Control of flowering time by FLC orthologues in Brassica napus. The Plant Journal, 28: 545553.Google Scholar
Takada, S., Hibara, K., Ishida, T. & Tasaka, M. (2001). The CUP-SHAPED COTYLEDON1 gene of Arabidopsis regulates shoot apical meristem formation. Development, 128: 11271135.Google Scholar
Takhtajan, A. (1969). Flowering Plants: Origin and Dispersal. Edinburgh: Oliver & Boyd.Google Scholar
Takhtajan, A. (1991). Evolutionary Trends in Flowering Plants. New York, NY: Columbia University Press.Google Scholar
Tamura, M. (1995). Ranunculaceae. In Die natürlichen Pflanzenfamilien, 17a, Part IV, eds. Engler, A. & Prantl, K.. Berlin: Ducker & Humblot.Google Scholar
Tanabe, Y., Hasebe, M., Sekimoto, H. et al. (2005). Characterization of MADS-box genes in charophycean green algae and its implication for the evolution of MADS-box genes. Proceedings of the National Academy of Sciences of the United States of America, 102: 24362441.Google Scholar
Tang, G., Reinhart, B. J., Bartel, D. P. & Zamore, P. D. (2003). A biochemical framework for RNA silencing in plants. Genes and Development, 17: 4963.Google Scholar
Tank, D. C. & Olmstead, R. G. (2008). From annuals to perennials: phylogeny of subtribe Castillejinae (Orobanchaceae). American Journal of Botany, 95: 608625.Google Scholar
Tattersall, A. D., Turner, L., Knox, M. R. et al. (2005). The mutant crispa reveals multiple roles for PHANTASTICA in pea compound leaf development. Plant Cell, 17: 10461060.Google Scholar
Taylor, D. W. & Hickey, L. J. (1996). Evidence for and implications of an herbaceous origin of angiosperms. In Flowering Plant Origin, Evolution and Phylogeny, eds. Taylor, D. W. & Hickey, L. J.. New York, NY: Chapman & Hall, pp. 232266.Google Scholar
Taylor, P. (1989). The Genus Utricularia: A Taxonomic Monograph. London: HMSO.Google Scholar
Taylor, S., Hofer, J. & Murfet, I. (2001). Stamina pistilloida, the pea ortholog of Fim and UFO, is required for normal development of flowers, inflorescences, and leaves. Plant Cell, 13: 3146.Google Scholar
Taylor, S. A., Hofer, J. M., Murfet, I. C. et al. (2002). PROLIFERATING INFLORESCENCE MERISTEM, a MADS-box gene that regulates floral meristem identity in pea. Plant Physiology, 129: 11501159.Google Scholar
Teeri, T. H., Elomaa, P., Kotilainen, M. & Albert, V. A. (2006a). Mining plant diversity: Gerbera as a model system for plant developmental and biosynthetic research. BioEssays, 28: 756767.Google Scholar
Teeri, T. H., Uimari, A., Kotilainen, M. et al. (2006b). Reproductive meristem fates in Gerbera. Journal of Experimental Botany, 57: 34453455.Google Scholar
Telfer, A. & Poethig, R. S. (1998) HASTY, a gene that regulates the timing of shoot maturation in Arabidopsis thaliana. Development, 125: 18891898.Google Scholar
Terpstra, I. & Heidstra, R. (2009). Stem cells: the root of all cells. Seminars in Cell and Developmental Biology, 20: 10891096.Google Scholar
Theißen, G. (2000). Evolutionary developmental genetics of floral symmetry: the revealing power of Linnaeus’ monstrous flower. Bioessays, 22: 209213.Google Scholar
Theißen, G. (2006). The proper place of hopeful monsters in evolutionary biology. Theory in Biosciences, 124: 349369.Google Scholar
Theißen, G. (2009). Saltational evolution: hopeful monsters are here to stay. Theory in Biosciences, 128: 4351.Google Scholar
Theißen, G., Becker, A., Winter, K.-U. et al. (2002). How the land plants learned their floral ABCs: the role of MADS box genes in the evolutionary origin of flowers. In Developmental Genetics and Plant Evolution, eds. Cronk, Q. C. B., Bateman, R. M. & Hawkins, J. A.. London: Taylor & Francis, pp. 173206.Google Scholar
Theißen, G., Kim, J. T. & Saedler, H. (1996). Classification and phylogeny of the MADS-box multigene family suggest defined roles of MADS-box gene subfamilies in the morphological evolution of eukaryotes. Journal of Molecular Evolution, 43: 484516.Google Scholar
Theissen, G. & Melzer, R. (2007). Molecular mechanisms underlying origin and diversification of the angiosperm flower. Annals of Botany, 100: 603619.Google Scholar
Theißen, G. & Melzer, R. (2016). Robust views on plasticity and biodiversity. Annals of Botany, 117: 693697.Google Scholar
Theißen, G. & Saedler, H. (2001). Floral quartets. Nature, 409: 469471.Google Scholar
Thien, L. B. (1980). Patterns of pollination in the primitive angiosperms. Biotropica, 12: 113.Google Scholar
Thomas, H. (2003). Do green plants age, and if so, how? Topics in Current Genetics, 3: 145171.Google Scholar
Thomas, M. M., Rudall, P. J., Ellis, A. G., Savolainen, V. & Glover, B. J. (2009). Development of a complex floral trait: the pollinator-attracting petal spots of the beetle daisy, Gorteria diffusa (Asteraceae). American Journal of Botany, 96: 21842196.Google Scholar
Thompson, B. E., Bartling, L., Whipple, C. et al. (2009). bearded-ear encodes a MADS box transcription factor critical for maize floral development. Plant Cell, 21: 25782590.Google Scholar
Thorpe, T. A. (2007). History of plant tissue culture. Molecular Biotechnology, 37: 169180.Google Scholar
Timmermans, M. C. P., Hudson, A., Becraft, P. W. & Nelson, T. (1999). Rough sheath2: a Myb protein that represses knox homeobox genes in maize lateral organ primordia. Science, 284: 151153.Google Scholar
Tomescu, A. M. F. (2009). Megaphylls, microphylls and the evolution of leaf development. Trends in Plant Science, 14: 512.Google Scholar
Tomlinson, P. B. (1990). The Structural Biology of Palms. Oxford: Clarendon Press.Google Scholar
Tomlinson, P. B. & Huggett, B. A. (2012). Cell longevity and sustained primary growth in palm stems. American Journal of Botany, 99: 18911902.Google Scholar
Tooke, F., Ordidge, M., Chiurugwi, T. & Battey, N. (2005). Mechanisms and function of flower and inflorescence reversion. Journal of Experimental Botany, 56: 25872599.Google Scholar
Toriba, T., Harada, K., Takamura, A. et al. (2007). Molecular characterization the YABBY gene family in Oryza sativa and expression analysis of OsYABBY1. Molecular Genetics and Genomics, 277: 457468.Google Scholar
Townsley, B. T. & Sinha, N. R. (2012). A new development: evolving concepts in leaf ontogeny. Annual Review of Plant Biology, 63: 535562.Google Scholar
Trevaskis, B., Bagnall, D. J., Ellis, M. H., Peacock, W. J. & Dennis, E. S. (2003). MADS box genes control vernalization-induced flowering in cereals. Proceedings of the National Academy of Sciences of the United States of America, 100: 1309913104.Google Scholar
Tröbner, W., Ramirez, L., Motte, P. et al. (1992). GLOBOSA: a homeotic gene which interacts with DEFICIENS in the control of Antirrhinum floral organogenesis. EMBO Journal, 11: 46934704.Google Scholar
Troll, W. (1937, 1939, 1943). Vergleichende Morphologie der höheren Pflanzen. Berlin: Borntraeger.Google Scholar
Troll, W. (1964, 1969). Die Infloreszenzen. Jena: Fischer.Google Scholar
True, J. R. & Haag, E. S. (2001). Developmental system drift and flexibility in evolutionary trajectories. Evolution and Development, 3: 109119.Google Scholar
Tsai, W. C., Chuang, M. H., Kuoh, C. S., Chen, W. H. & Chen, H. H. (2004). Four DEF-like MADS box genes displayed distinct floral morphogenetic roles in Phalaenopsis orchid. Plant and Cell Physiology, 45: 831844.Google Scholar
Tsai, W. C., Pan, Z. J., Hsiao, Y. Y., Chen, L. J. & Liu, Z. J. (2014). Evolution and function of MADS-box genes involved in orchid floral development. Journal of Systematics and Evolution, 52: 397410.Google Scholar
Tsiantis, M., Brown, M. I. N., Skibinski, G. & Langdale, J. A. (1999). Disruption of auxin transport is associated with aberrant leaf development in maize. Plant Physiology, 121: 11631168.Google Scholar
Tsuda, K. & Katagiri, F. (2010). Comparing signaling mechanisms engaged in pattern-triggered and effector-triggered immunity. Current Opinion in Plant Biology, 13: 459465.Google Scholar
Tsuda, K., Sato, M., Stoddard, T., Glazebrook, J. & Katagiri, F. (2009). Network properties of robust immunity in plants. PLoS Genetics, 5: e1000772.Google Scholar
Tsukaya, H. (1995). Developmental genetics of leaf morphogenesis in dicotyledonous plants. Journal of Plant Research, 108: 407416.Google Scholar
Tsukaya, H. (1997). Determination of the unequal fate of cotyledons of a one-leaf plant, Monophyllaea. Development, 124: 12751280.Google Scholar
Tsukaya, H. (2000). The role of meristematic activities in the formation of leaf blades. Journal of Plant Research, 113: 119126.Google Scholar
Tsukaya, H. (2002). Interpretation of mutants in leaf morphology: genetic evidence for a compensatory system in leaf morphogenesis that provides a new link between cell and organismal theory. International Review of Cytology, 217: 139Google Scholar
Tsukaya, H. (2003). Organ shape and size: a lesson from studies of leaf morphogenesis. Current Opinion in Plant Biology, 6: 5762.Google Scholar
Tsukaya, H. (2006). Mechanism of leaf-shape determination. Annual Review of Plant Biology, 57: 477496.Google Scholar
Tsukaya, H. (2008). Controlling size in multicellular organs: focus on the leaf. PLoS Biology, 6: 13731376.Google Scholar
Tsukaya, H. (2013). Leaf development. In The Arabidopsis Book. Rockville, MD: American Society of Plant Biologists, 11: e0163.Google Scholar
Tsukaya, H. (2014). Comparative leaf development in angiosperms. Current Opinion in Plant Biology, 17: 103109.Google Scholar
Tsukaya, H., Inaba-Higano, K. & Komeda, Y. (1995). Phenotypic and molecular mapping of an acaulis2 mutant of Arabidopsis thaliana with flower stalks of much reduced length. Plant Cell Physiology, 36: 239246.Google Scholar
Tucker, S. C. (1984a). Origin of symmetry in flowers. In Contemporary Problems in Plant Anatomy, eds. White, R. A. & Dickison, W. C.. New York, NY: Academic Press, pp. 351395.Google Scholar
Tucker, S. C. (1984b). Unidirectional organ initiation in leguminous flowers. American Journal of Botany, 71: 11391148.Google Scholar
Tucker, S. C. (1987). Floral initiation and development in legumes. In Advances in Legume Systematics, 3, ed. Stirton, C. H.. Kew: Royal Botanic Gardens, pp. 183239.Google Scholar
Tucker, S. C. (1991). Helical floral organogenesis in Gleditsia, a primitive caesalpinioid legume. American Journal of Botany, 78: 11301149.Google Scholar
Tucker, S. C. (1996). Trends in evolution of floral ontogeny in Cassia sensu stricto, Senna. and Chamaecrista (Leguminosae: Caesalpinoideae: Cassieae: Cassiinae): a study in convergence. American Journal of Botany, 83: 687711.Google Scholar
Tucker, S. C. (1999). Evolutionary lability of symmetry in early floral development. International Journal of Plant Sciences, 160: S25S39.Google Scholar
Tucker, S. C. (2000). Floral development in tribe Detarieae (Leguminosae: Caesalpinioideae): Amherstia, Brownea, and Tamarindus. American Journal of Botany, 87: 13851407.Google Scholar
Tucker, S. C. (2001). Floral development in Schotia and Cynometra (Leguminosae: Caesalpinioideae: Detarieae). American Journal of Botany, 88: 11641180.Google Scholar
Tucker, S. C. (2002). Floral ontogeny in Sophoreae (Leguminosae: Papilionoideae). III. Cadia purpurea with radial symmetry and random petal aestivation. American Journal of Botany, 89: 748757.Google Scholar
Tucker, S. C. (2003). Floral development in legumes. Plant Physiology, 131: 911926.Google Scholar
Tucker, S. C. & Douglas, A. W. (1996). Floral structure, development, and relationships of paleoherbs: Saruma, Cabomba, Lactoris, and selected Piperales. In Flowering Plant Origin, Evolution, and Phylogeny, eds. Taylor, D. W. & Hickey, L. J.. New York, NY: Chapman & Hall, pp. 141175.Google Scholar
Tuskan, G. A., Difazio, S., Jansson, S. et al. (2006). The genome of black cottonwood, Populus trichocarpa (Torr. & Gray). Science, 313: 15961604.Google Scholar
Tzeng, T.-Y., Chen, H.-Y. & Yang, C.-H. (2002). Ectopic expression of carpel-specific MADS box genes from lily and Lisianthus causes similar homeotic conversion of sepal and petal in Arabidopsis. Plant Physiology, 130: 18271836.Google Scholar
Tzeng, T.-Y. & Yang, C.-H. (2001). A MADS box gene from lily (Lilium longiflorum) is sufficient to generate dominant negative mutation by interacting with PISTILLATA (PI) in Arabidopsis thaliana. Plant Cell Physiology, 42: 11561168.Google Scholar
Überlacker, B., Klinge, B. & Werr, W. (1996). Ectopic expression of the maize homeobox genes ZmHox1a or ZmHox1b causes pleiotropic alterations in the vegetative and floral development of transgenic tobacco. Plant Cell, 8: 349362.Google Scholar
Uhl, N. W. & Dransfield, J. (1987). Genera Palmarum. Lawrence, KS: Allen Press.Google Scholar
Uimari, A., Kotilainen, M., Elomaa, P. et al. (2004). Integration of reproductive meristem fates by a SEPALLATA-like MADS box gene. Proceedings of the National Academy of Sciences of the United States of America, 101: 1581715822.Google Scholar
Uittien, H. (1928). Uber den Zusammenhang zwischen Blattnervatur und Sprossverzweigung. Recueil des Travaux Botaniques Neerlandais, 25: 390412.Google Scholar
Uller, T. & Helanterä, H. (2011). When are genes ‘leaders’ or ‘followers’ in evolution? Trends in Ecology and Evolution, 26: 435436.Google Scholar
Usami, T., Horiguchi, G., Yano, S. & Tsukaya, H. (2009). The more and smaller cells mutants of Arabidopsis thaliana identify novel roles for SQUAMOSA PROMOTER BINDING PROTEIN-LIKE genes in the control of heteroblasty. Development, 136: 955964.Google Scholar
Uyttewaal, M., Burian, A., Alim, K. et al. (2012). Mechanical stress acts via katanin to amplify differences in growth rate between adjacent cells in Arabidopsis. Cell, 149: 439451.Google Scholar
Vallejo-Marín, M., Manson, J. S., Thomson, J. D. & Barrett, S. C. H. (2009). Division of labour within flowers: heteranthery, a floral strategy to reconcile contrasting pollen fates. Journal of Evolutionary Biology, 22: 828839.Google Scholar
Vallius, E. (2000). Position-dependent reproductive success of flowers in Dactylorhiza maculata (Orchidaceae). Functional Ecology, 14: 573579.Google Scholar
Van de Peer, Y., Fawcett, J. A., Proost, S., Sterck, L. & Vandepoele, K. (2009). The flowering world: a tale of duplications. Trends in Plant Science, 14: 680688.Google Scholar
van der Graaff, E., Dulk-Ras, A. D., Hooykaas, P. J. J. & Keller, B. (2000). Activation tagging of the LEAFY PETIOLE gene affects leaf petiole development in Arabidopsis thaliana. Development, 127: 49714980.Google Scholar
van der Maesen, L. J. G. (1970). Primitiae Africanae VIII. A revision of the genus Cadia Forskål (Caes.) and some remarks regarding Dicraeopetalum Harms (Pap.) and Platycelyphium (Harms) (Pap.). Acta Botanica Neerlandica, 19: 227248.Google Scholar
van Doorn, W. G. & Stead, A. D. (1997). Abscission of flowers and floral parts. Journal of Experimental Botany, 48: 821837.Google Scholar
Van Dyken, J. & Wade, M. J. (2010). The genetic signature of conditional expression. Genetics, 84: 557570.Google Scholar
Vandenbussche, M., Horstman, A., Zethof, J. et al. (2009). Differential recruitment of WOX transcription factors for lateral development and organ fusion in Petunia and Arabidopsis. Plant Cell, 21: 22692283.Google Scholar
Vandenbussche, M., Zethof, J., Royaert, S., Weterings, K. & Gerats, T. (2004). The duplicated B-class heterodimer model: whorl-specific effects and complex genetic interactions in Petunia hybrida flower development. Plant Cell, 16: 741754.Google Scholar
Vandenbussche, M., Zethof, J., Souer, E. et al. (2003). Toward the analysis of the petunia MADS box gene family by reverse and forward transposon insertion mutagenesis approaches: B, C, and D floral organ identity functions require SEPALLATA-like MADS box genes in petunia. Plant Cell, 15: 26802693.Google Scholar
Vazquez, F. (2009). Small RNAs in plants. In Encyclopedia of Life Sciences. Chichester: Wiley.Google Scholar
Vázquez-Lobo, A., Carlsbecker, A., Vergara-Silva, F. 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 and Development, 9: 446459.Google Scholar
Vekemans, D., Proost, S., Vanneste, K. et al. (2012a). Gamma paleohexaploidy in the stem lineage of core eudicots: significance for MADS-box gene and species diversification. Molecular Biology and Evolution, 29: 37933806.Google Scholar
Vekemans, D., Viaene, T., Caris, P. & Geuten, K. (2012b). Transference of function shapes organ identity in the dove tree inflorescence. New Phytologist, 193: 216228.Google Scholar
Velhagen, W. A. (1997). Analyzing developmental sequences using sequence units. Systematic Biology, 46: 204210.Google Scholar
Viaene, T., Vekemans, D., Irish, V. F. et al. (2009). Pistillata: duplications as a mode for floral diversification in (basal) asterids. Molecular Biology and Evolution, 26: 26272645.Google Scholar
Vialette-Guiraud, A. C. M., Adam, H., Finet, C. et al. (2011). Insights from ANA-grade angiosperms into the early evolution of CUP-SHAPED COTYLEDON genes. Annals of Botany, 107: 15111519.Google Scholar
Vialette-Guiraud, A. C. M. & Scutt, C. P. (2009). Carpel evolution. Annual Plant Reviews, 38: 134.Google Scholar
Vijayraghavan, U., Prasad, K. & Meyerowitz, E. (2005). Specification and maintenance of the floral meristem: interactions between positively acting promoters of flowering and negative regulators. Current Science, 89: 18351843.Google Scholar
Vlad, D., Kierzkowski, D., Rast, M. I. et al. (2014). Leaf shape evolution through duplication, regulatory diversification, and loss of a homeobox gene. Science, 343: 780783.Google Scholar
Vollbrecht, E., Veit, B., Sinha, N. & Hake, S. (1991). The developmental gene Knotted-1 is a member of a maize homeobox gene family. Nature, 350: 241243.Google Scholar
von Baer, K. E. (1828). Über Entwicklungsgeschichte der Thiere: Beobachtung und Reflexion, Vol. 1. Königsberg: Bornträger.Google Scholar
von Balthazar, M. & Endress, P. K. (2002). Development of inflorescences and flowers in Buxaceae and the problem of perianth interpretation. International Journal of Plant Sciences, 163: 847876.Google Scholar
von Hagen, K. B. & Kadereit, J. W. (2002). Phylogeny and flower evolution of the Swertiinae (Gentianaceae-Gentianeae): homoplasy and the principle of variable proportions. Systematic Botany, 27: 548572.Google Scholar
von Wangenheim, D., Fangerau, J., Schmitz, A. et al. (2016). Rules and self-organizing properties of post-embryonic plant organ cell division patterns. Current Biology, 26: 439449.Google Scholar
Vrebalov, J., Pan, I. L., Arroyo, A. J. et al. (2009). Fleshy fruit expansion and ripening are regulated by the tomato SHATTERPROOF gene TAGL1. Plant Cell, 21: 30413062.Google Scholar
Vrebalov, J., Ruezinsky, D., Padmanabhan, V. et al. (2002). A MADS-box gene necessary for fruit ripening at the tomato ripeninginhibitor (rin) locus. Science, 296: 343346.Google Scholar
Vroemen, C. W., Mordhorst, A. P., Albrecht, C., Kwaaitaal, M. A. & de Vries, S. C. (2003). The CUP-SHAPED COTYLEDON3 gene is required for boundary and shoot meristem formation in Arabidopsis. Plant Cell, 15: 15631577.Google Scholar
Waddington, C. H. (1953). Genetic assimilation of an acquired character. Evolution, 7: 118126.Google Scholar
Wagner, A. (2005). Robustness and Evolvability in Living Systems. Princeton, NJ: Princeton University Press.Google Scholar
Wagner, A. (2011). The Origins of Evolutionary Innovations: A Theory of Transformative Change in Living Systems. Oxford: Oxford University Press.Google Scholar
Wagner, A. (2014). Arrival of the Fittest: Solving Evolution’s Greatest Puzzle. New York, NY: Penguin.Google Scholar
Wagner, G. P. (1996). Homologues, natural kinds and the evolution of modularity. American Zoologist, 36: 3643.Google Scholar
Wagner, G. P. & Altenberg, L. (1996). Complex adaptations and evolution of evolvability. Evolution, 50: 967976.Google Scholar
Wagner, G. P., Pavlicev, M. & Cheverud, J. M. (2007). The road to modularity. Nature Reviews Genetics, 8: 921931.Google Scholar
Waites, R. & Hudson, A. (1995). phantastica: a gene required for dorsoventrality of leaves in Antirrhinum majus. Development, 121: 21432154.Google Scholar
Waites, R. & Hudson, A. (2001). The Handlebars gene is required with Phantastica for dorsoventral asymmetry of organs and for stem cell activity in Antirrhinum. Development, 128: 19231931.Google Scholar
Waites, R., Selvadurai, H. R. N., Oliver, I. R. & Hudson, A. (1998). The Phantastica gene encodes a MYB transcription factor involved in growth and dorsoventrality of lateral organs in Antirrhinum. Cell, 93: 779789.Google Scholar
Wake, D. (2003). Homology and homoplasy. In Keywords and Concepts in Evolutionary Developmental Biology, eds. Hall, K. & Olson, W. M.. Cambridge, MA: Harvard University Press, pp. 191201.Google Scholar
Walbot, V. (1996). Sources and consequences of phenotypic and genotypic plasticity in flowering plants. Trends in Plant Science, 1: 2732.Google Scholar
Walker, J. W. & Walker, A. G. (1984). Ultrastructure of lower Cretaceous angiosperm pollen and the origin and early evolution of flowering plants. Annals of the Missouri Botanical Garden, 71: 464521.Google Scholar
Walker-Larsen, J. & Harder, L. D. (2000). The evolution of staminodes in angiosperms: patterns of stamen reduction, loss, and functional reinvention. American Journal of Botany, 87: 13671384.Google Scholar
Walsh, B. & Blows, M. W. (2009). Abundant genetic variation + strong selection = multivariate genetic constraints: a geometric view of adaptation. Annual Review of Ecology, Evolution and Systematics, 40: 4159.Google Scholar
Wang, J. W., Czech, B. & Weigel, D. (2009a). miR156-regulated SPL transcription factors define an endogenous flowering pathway in Arabidopsis thaliana. Cell, 138: 738749.Google Scholar
Wang, J. W., Park, M. Y., Wang, L. J. et al. (2011). MiRNA control of vegetative phase change in trees. PLoS Genetics, 7: e1002012.Google Scholar
Wang, R., Farrona, S., Vincent, C. et al. (2009b). PEP1 regulates perennial flowering in Arabis alpina. Nature, 459: 423428.Google Scholar
Wang, R. L., Stec, A., Hey, J., Lukens, L. & Doebley, J. (1999). The limits of selection during maize domestication. Nature, 398: 236239.Google Scholar
Wang, Z., Luo, Y., Li, X. et al. (2008). Genetic control of floral zygomorphy in pea (Pisum sativum L.). Proceedings of the National Academy of Sciences of the United States of America, 105: 1041410419.Google Scholar
Wanntorp, L. & Ronse De Craene, L. P. (2005). The Gunnera flower: key to eudicot diversification or response to pollination mode? International Journal of Plant Sciences, 166: 945953.Google Scholar
Wardlaw, C. W. (1955). Embryogenesis in Plants. London: Methuen.Google Scholar
Warner, K. A., Rudall, P. J. & Frohlich, M. W. (2008). Differentiation of perianth organs in Nymphaeales. Taxon, 57: 10961109.Google Scholar
Warner, K. A., Rudall, P. J. & Frohlich, M. W. (2009). Environmental control of sepalness and petalness in perianth organs of waterlilies: a new Mosaic Theory on the evolutionary origin of a differentiated perianth. Journal of Experimental Botany, 60: 35593574.Google Scholar
Washburn, J. D., Bird, K. A., Conant, G. C. & Pires, J. C. (2016). Convergent evolution and the origin of complex phenotypes in the age of systems biology. International Journal of Plant Sciences, 177: 305318.Google Scholar
Watanabe, K. & Okada, K. (2003). Two discrete cis elements control the abaxial side-specific expression of the FILAMENTOUS FLOWER gene in Arabidopsis. Plant Cell, 15: 25922602.Google Scholar
Weber, A. (2003). What is morphology and why is it time for its renaissance in plant systematics? In Deep Morphology: Towards a Renaissance of Morphology in Plant Systematics, eds. Stuessy, T. F., Mayer, V. & Hörandl, E.. Ruggell: Gantner, pp. 332.Google Scholar
Weber, A., Clark, J. L. & Möller, M. (2013). A new formal classification of Gesneriaceae. Selbyana, 31: 6894.Google Scholar
Weberling, F. (1989). Morphology of Flowers and Inflorescences. Cambridge: Cambridge University Press.Google Scholar
Webster, M. A. & Gilmartin, P. M. (2006). Analysis of late stage flower development in Primula vulgaris reveals novel differences in cell morphology and temporal aspects of floral heteromorphy. New Phytologist, 171: 591603.Google Scholar
Wei, L., Wang, Y.-Z. & Li, Z.-Y. (2011). Floral ontogeny of Ruteae (Rutaceae) and its systematic implications. Plant Biology, 14: 190197.Google Scholar
Weigel, D. (2012). Natural variation in Arabidopsis: from molecular genetics to ecological genomics. Plant Physiology, 158: 222.Google Scholar
Weigel, D., Alvarez, J., Smyth, D. R., Yanofsky, M. F. & Meyerowitz, E. M. (1992). LEAFY controls floral meristem identity in Arabidopsis. Cell, 69: 843859.Google Scholar
Weiss, M. R. (1995). Floral color change: a widespread functional convergence. American Journal of Botany, 82: 167185.Google Scholar
Wendel, J. F. (2015). The wondrous cycles of polyploidy in plants. American Journal of Botany, 102: 17531756.Google Scholar
West-Eberhard, M. J. (2003). Developmental Plasticity and Evolution. New York, NY: Oxford University Press.Google Scholar
Westerkamp, C. & Weber, A. (1999). Keel flowers of the Polygalaceae and Fabaceae: a functional comparison. Botanical Journal of the Linnean Society, 129: 207221.Google Scholar
Weston, P. H. (2000). Process morphology from a cladistic perspective. In Homology and Systematics, eds. Scotland, R. & Pennington, R. T.. London: Taylor & Francis, pp. 124144.Google Scholar
Westwood, J. H., Yoder, J. I., Timko, M. P. & dePamphilis, C. W. (2010). The evolution of parasitism in plants. Trends in Plant Science, 15: 227235.Google Scholar
Whipple, C. J., Ciceri, P., Padilla, C. M. et al. (2004). Conservation of B-class floral homeotic gene function between maize and Arabidopsis. Development, 131: 60836091.Google Scholar
Whipple, C. J., Zanis, M. J., Kellogg, E. A. & Schmidt, R. J. (2007). Conservation of B-class gene expression in the second whorl of a basal grass and outgroups links the origin of lodicules and petals. Proceedings of the National Academy of Sciences of the United States of America, 104: 10811086.Google Scholar
White, D. W. (2006). PEAPOD regulates lamina size and curvature in Arabidopsis. Proceedings of the National Academy of Sciences of the United States of America, 103: 1323813243.Google Scholar
Wigge, P. A., Kim, M. C., Jaeger, K. E. et al. (2005). Integration of spatial and temporal information during floral induction in Arabidopsis. Science, 309: 10561059.Google Scholar
Wiley, E. O. (1981). Phylogenetics: The Theory and Practice of Phylogenetic Systematics. New York, NY: Wiley.Google Scholar
Wilkinson, M., de Andrade Silva, E., Zachgo, S., Saedler, H. & Schwarz-Sommer, Z. (2000). CHORIPETALA and DESPENTEADO: general regulators during plant development and potential floral targets of FIMBRIATA-mediated degradation. Development, 127: 37253734.Google Scholar
Williams, D. M. & Ebach, M. C. (2007). Heterology: the shadows of a shade. Cladistics, 23: 6483.Google Scholar
Williston, S.W. (1914). Water Reptiles of the Past and Present. Chicago, IL: University of Chicago Press.Google Scholar
Willmann, M. R. & Poethig, R. S. (2007). Conservation and evolution of miRNA regulatory programs in plant development. Current Opinion in Plant Biology, 10: 503511.Google Scholar
Wiltshire, R. J. E., Murfet, I. C. & Reid, J. B. (1994). The genetic control of heterochrony: evidence from developmental mutants of Pisum sativum L. Journal of Evolutionary Biology, 7: 447465.Google Scholar
Wiltshire, R. J. E., Potts, B. M. & Reid, J. B. (1998). Genetic control of reproductive and vegetative phase change in the Eucalyptus risdonii–E. tenuiramis complex. Australian Journal of Botany, 46: 4563.Google Scholar
Winter, K.-U., Becker, A., Münster, T. et al. (1999). MADS-box genes reveal that gnetophytes are much more closely related to conifers than to flowering plants. Proceedings of the National Academy of Sciences of the United States of America, 96: 73427347.Google Scholar
Winther, R. G. (2015). Evo-devo as a trading zone. In Conceptual Change in Biology: Scientific and Philosophical Perspectives on Evolution and Development, ed. Love, A. C.. Dordrecht: Springer, pp. 459482.Google Scholar
Wojciechowski, M. F., Lavin, M. & Sanderson, M. J. (2004). A phylogeny of legumes (Leguminosae) based on analysis of the plastid matK gene resolves many well-supported subclades within the family. American Journal of Botany, 91: 18461862.Google Scholar
Wolfe, J. M. & Hegna, T. A. (2014). Testing the phylogenetic position of Cambrian pancrustacean larval fossils by coding ontogenetic stages. Cladistics, 30: 366390.Google Scholar
Woloszynska, M., Bocer, T., Mackiewicz, P. & Janska, H. (2004). A fragment of chloroplast DNA was transferred horizontally, probably from non-eudicots, to mitochondrial genome of Phaseolus. Plant Molecular Biology, 56: 811820.Google Scholar
Woodrick, R., Martin, P. R., Birman, I. & Pickett, F. B. (2000). The Arabidopsis embryonic shoot fate map. Development, 127: 813820.Google Scholar
Worley, A., Baker, A., Thompson, J. & Barrett, S. C. H. (2000). Floral display in Narcissus: variation in flower size and number at the species, population, and individual levels. International Journal of Plant Sciences, 161: 6979.Google Scholar
Wörz, A. (1996) Rubiaceae. Rötegewächse. In Die Farn und Blütenpflanzen Baden-Württembergs. Band 5, ed. Sebald, O., Seybold, S., Philippi, G. & Wörz, A.. Stuttgart: Ulmer, pp. 449484.Google Scholar
Wróblewska, M., Dołzbłasz, A. & Zagórska-Marek, B. (2016). The role of ABC genes in shaping perianth phenotype in the basal angiosperm Magnolia. Plant Biology, 18: 230238.Google Scholar
Wu, C. A., Lowry, D. B., Cooley, A. M. et al. (2008). Mimulus is an emerging model system for the integration of ecological and genomic studies. Heredity, 100: 220230.Google Scholar
Wu, G., Park, M. Y., Conway, S. R. et al. (2009). The sequential action of miR156 and miR172 regulates developmental timing in Arabidopsis. Cell, 138: 750759.Google Scholar
Wu, G. & Poethig, R. S. (2006). Temporal regulation of shoot development in Arabidopsis thaliana by miR156 and its target SPL3. Development, 133: 35393547.Google Scholar
Wund, M. A. (2012). Assessing the impacts of phenotypic plasticity on evolution. Integrative and Comparative Biology, 52: 515.Google Scholar
Wunderlin, R. P. (1983). Revision of the arborescent Bauhinias (Fabaceae: Caesalpinioideae: Cercideae) native to middle America. Annals of the Missouri Botanical Garden, 70: 95127.Google Scholar
Xi, Z., Bradley, R. K., Wurdack, K. J. et al. (2012). Horizontal transfer of expressed genes in a parasitic flowering plant. BMC Genomics, 13: 227.Google Scholar
Xi, Z., Wang, Y., Bradley, R. K. et al. (2013). Massive mitochondrial gene transfer in a parasitic flowering plant clade. PLoS Genetics, 9: e1003265.Google Scholar
Xiao, H., Wang, Y., Liu, D. et al. (2003). Functional analysis of the rice AP3 homologue OsMADS16 by RNA interference. Plant Molecular Biology, 52: 957966.Google Scholar
Xu, L., Xu, Y., Dong, A., Sun, Y. et al. (2003). Novel as1 and as2 defects in leaf adaxial–abaxial polarity reveal the requirement for ASYMMETRIC LEAVES1 and 2 and ERECTA functions in specifying leaf adaxial identity. Development, 130: 40974107.Google Scholar
Xu, Y., Sun, Y., Liang, W. Q. & Huang, H. (2002). The Arabidopsis AS2 gene encoding a predicted leucine-zipper protein is required for the leaf polarity formation. Acta Botanica Sinica, 44: 11941202.Google Scholar
Xu, Y., Teo, L. L., Zhou, J., Kumar, P. P. & Yu, H. (2006). Floral organ identity genes in the orchid Dendrobium crumenatum. The Plant Journal, 46: 5468.Google Scholar
Yamada, T., Yokota, S., Hirayama, Y. et al. (2011). Ancestral expression patterns and evolutionary diversification of YABBY genes in angiosperms. The Plant Journal, 67: 2636.Google Scholar
Yamaguchi, A., Wu, M. F., Yang, L. et al. (2009). The microRNA-regulated SBP-box transcription factor SPL3 is a direct upstream activator of LEAFY, FRUITFULL, and APETALA1. Developmental Cell, 17: 268278.Google Scholar
Yamaguchi, T., Lee, D. Y., Miyao, A. et al. (2006). Functional diversification of the two C-class MADS box genes OSMADS3 and OSMADS58 in Oryza sativa. Plant Cell, 18: 1528.Google Scholar
Yamaguchi, T., Yano, S. & Tsukaya, H. (2010). Genetic framework for flattened leaf blade formation in unifacial leaves of Juncus prismatocarpus. Plant Cell, 22: 21412155.Google Scholar
Yamaki, S., Nagato, Y., Kurata, N. & Nonomura, K.-I. (2011). Ovule is a lateral organ finally differentiated from the terminating floral meristem in rice. Developmental Biology, 351: 208216.Google Scholar
Yant, L., Mathieu, J. & Schmid, M. (2009). Just say no: floral repressors help Arabidopsis bide the time. Current Opinion in Plant Biology, 12: 580586.Google Scholar
Yephremov, A., Wisman, E., Huijser, P. et al. (1999). Characterization of the FIDDLEHEAD gene of Arabidopsis reveals a link between adhesion response and cell differentiation in the epidermis. Plant Cell, 11: 21872201.Google Scholar
Yockteng, R. B., Almeida, A. M. R., Morioka, K., Alvarez-Buylla, E. R. & Specht, C. D. (2013). Molecular evolution and patterns of duplications in the SEP/AGL6-like lineage of the Zingiberales: a proposed mechanism for floral diversification. Molecular Biology and Evolution, 30: 24012422.Google Scholar
Yogeeswaran, K., Frary, A., York, T. L. et al. (2005). Comparative genome analyses of Arabidopsis spp.: inferring chromosomal rearrangement events in the evolutionary history of A. thaliana. Genome Research, 15: 505515.Google Scholar
Yoo, M.-J., Bell, C. D., Soltis, P. S. & Soltis, D. E. (2005). Divergence times and historical biogeography of Nymphaeales. Systematic Botany, 30: 693704.Google Scholar
Yoo, M.-J., Soltis, P. S. & Soltis, D. E. (2010). Expression of floral MADS-box genes in two divergent water lilies: Nymphaeales and Nelumbo. International Journal of Plant Sciences, 171: 121146.Google Scholar
Yoshida, S., Barbier de Reuille, P., Lane, B. et al. (2014). Genetic control of plant development by overriding a geometric division rule. Developmental Cell, 29: 7587.Google Scholar
Yoshida, S., Cui, S., Ichihashi, Y. & Shirasu, K. (2016). The haustorium, a specialized invasive organ in parasitic plants. Annual Review of Plant Biology, 67: 643667.Google Scholar
Yu, J., Hu, S., Wang, J. et al. (2002). A draft sequence of the rice genome (Oryza sativa L. ssp. indica). Science, 296: 7992.Google Scholar
Yuan, Z., Gao, S., Xue, D. W. et al. (2009). RETARDED PALEA1 controls palea development and floral zygomorphy in rice. Plant Physiology, 149: 235244.Google Scholar
Žádníkova, P. & Simon, R. (2014). How boundaries control plant development. Current Opinion in Plant Biology, 17: 116125.Google Scholar
Zagotta, M. T., Shannon, S. & Jacobs, C. (1992). Early-flowering mutants of Arabidopsis thaliana. Australian Journal of Plant Physiology, 19: 411418.Google Scholar
Zahn, L. M., Kong, H., Leebens-Mack, J. H. et al. (2005). The evolution of the SEPALLATA subfamily of MADS-box genes: a preangiosperm origin with multiple duplications throughout angiosperm history. Genetics, 169: 22092223.Google Scholar
Zanis, M. J., Soltis, P. S., Qiu, Y.-L., Zimmer, E. & Soltis, D. E. (2003). Phylogenetic analyses and perianth evolution in basal angiosperms. Annals of the Missouri Botanical Garden, 90: 129150.Google Scholar
Zhang, J. Z., Li, Z. M., Mei, L., Yao, J. L. & Hu, C. G. (2009). PtFLC homolog from trifoliate orange (Poncirus trifoliata) is regulated by alternative splicing and experiences seasonal fluctuation in expression level. Planta, 229: 847859.Google Scholar
Zhang, N., Wen, J. & Zimmer, E. A. (2015). Expression patterns of AP1, FUL, FT and LEAFY orthologs in Vitaceae support the homology of tendrils and inflorescences throughout the grape family. Journal of Systematics and Evolution, 53: 469476.Google Scholar
Zhang, R., Guo, C. C., Zhang, W. G. et al. (2013a). Disruption of the petal identity gene APETALA3-3 is highly correlated with loss of petals within the buttercup family (Ranunculaceae). Proceedings of the National Academy of Sciences of the United States of America, 110: 50745079.Google Scholar
Zhang, W., Kramer, E. M. & Davis, C. C. (2010). Floral symmetry genes and the origin and maintenance of zygomorphy in a plant pollinator mutualism. Proceedings of the National Academy of Sciences of the United States of America, 107: 63886393.Google Scholar
Zhang, W., Kramer, E. M. & Davis, C. C. (2016). Differential expression of CYC2 genes and the elaboration of floral morphologies in Hiptage, an Old World genus of Malpighiaceae. International Journal of Plant Sciences, 177: 551558.Google Scholar
Zhang, W., Steinmann, V. W., Nikolov, L., Kramer, E. M. & Davies, C. C. (2013b). Divergent genetic mechanisms underlie reversals to radial floral symmetry from diverse zygomorphic flowered ancestors. Frontiers in Plant Science, 4: 302.Google Scholar
Zhao, D., Yu, Q., Chen, C. & Ma, H. (2001). Genetic control of reproductive meristems. In Meristematic Tissues in Plant Growth and Development, eds. McManus, M. T. & Veit, B.. Sheffield: Sheffield Academic Press, pp. 89142.Google Scholar
Zhong, J. & Kellogg, E. A. (2015). Stepwise evolution of corolla symmetry in CYCLOIDEA2-like and RADIALIS-like gene expression patterns in Lamiales. American Journal of Botany, 102: 12601267.Google Scholar
Zhong, J., Powell, S. & Preston, J. C. (2016). Organ boundary NAC-domain transcription factors are implicated in the evolution of petal fusion. Plant Biology, 18: 893902Google Scholar
Zhong, R. & Ye, Z. H. (2004). Molecular and biochemical characterization of three WD-repeat-domain-containing inositol polyphosphate 5-phosphatases in Arabidopsis thaliana. Plant Cell Physiology, 45: 17201728.Google Scholar
Zhou, Q., Wang, Y. & Xiaobai, J. (2002). Ontogeny of floral organs and morphology of floral apex in Phellodendron amurense (Rutaceae). Australian Journal of Botany, 50: 633644.Google Scholar
Zhou, X.-R., Wang, Y.-Z., Smith, J. F. & Chen, R. (2008). Altered expression patterns of TCP and MYB genes relating to the floral developmental transition from initial zygomorphy to actinomorphy in Bournea (Gesneriaceae). New Phytologist, 178: 532543.Google Scholar
Zimmerman, R. H., Hackett, W. P. & Pharis, R. P. (1985). Hormonal aspects of phase change and precocious flowering. Encyclopaedia of Plant Physiology, 11: 79115.Google Scholar
Zluvova, J., Nicolas, M., Berger, A. et al. (2006). Premature arrest of the male flower meristem precedes sexual dimorphism in the dioecious plant Silene latifolia. Proceedings of the National Academy of Sciences of the United States of America, 103: 1885418859.Google Scholar
Zobell, O., Faigl, W., Saedler, H. & Münster, T. (2010). MIKC MADS-box proteins: conserved regulators of the gametophytic generation of land plants. Molecular Biology and Evolution, 27: 12011211.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • Alessandro Minelli, Università degli Studi di Padova, Italy
  • Book: Plant Evolutionary Developmental Biology
  • Online publication: 09 February 2018
  • Chapter DOI: https://doi.org/10.1017/9781139542364.013
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • Alessandro Minelli, Università degli Studi di Padova, Italy
  • Book: Plant Evolutionary Developmental Biology
  • Online publication: 09 February 2018
  • Chapter DOI: https://doi.org/10.1017/9781139542364.013
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • Alessandro Minelli, Università degli Studi di Padova, Italy
  • Book: Plant Evolutionary Developmental Biology
  • Online publication: 09 February 2018
  • Chapter DOI: https://doi.org/10.1017/9781139542364.013
Available formats
×