Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-sh8wx Total loading time: 0 Render date: 2024-07-22T11:22:13.650Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 July 2011

Lionel G. Harrison
Affiliation:
University of British Columbia, Vancouver
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
The Shaping of Life
The Generation of Biological Pattern
, pp. 226 - 242
Publisher: Cambridge University Press
Print publication year: 2010

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

Aida, M., Ishida, T., Fukaki, H., Fujisawa, H. and Tasaka, M. (1997). Genes involved in organ separation in Arabidopsis: an analysis of cup-shaped cotyledon mutant. Plant Cell 9, 841–57.CrossRefGoogle ScholarPubMed
Akam, M. E. (1987). The molecular basis for metameric pattern in the Drosophila embryo. Development 101, 1–22.Google ScholarPubMed
Akam, M. E. (1989). Making stripes inelegantly. Nature 341, 282–3.CrossRefGoogle ScholarPubMed
Alexandrov, T., Golyandina, N. and Spirov, A. V. (2008). Singular spectrum analysis of gene expression profiles of the early Drosophila embryo: exponential-in-distance patterns. Res. Lett. in Signal Processing, 12, n.p.Google Scholar
Amtmann, A., Klieber, H. G. and Gradmann, D. (1992). Cytoplasmic free Ca2+ in the marine alga Acetabularia: measurement with Ca2+-selective microelectrodes and kinetic analysis. J. Exp. Bot. 43, 875–85.CrossRefGoogle Scholar
Angier, N. (2007). The Canon. Boston and New York: Houghton Mifflin.Google Scholar
Armstrong, J. B. (1989). A Turing model to explain heart development. Axolotl Newsletter 18, 23–5.Google Scholar
Armstrong, J. B. and Graveson, A. C. (1988). Progressive patterning precedes segmentation in the Mexican axolotl (Ambystoma mexicanum). Dev. Biol. 126, 1–6.CrossRefGoogle Scholar
Armstrong, J. B. and Malacinski, G. M. (1989). Developmental Biology of the Axolotl. New York: Oxford University Press.Google Scholar
Baker, R. E., Schnell, S. and Maini, P. K. (2009). Waves and patterning in developmental biology: vertebrate segmentation and feather bud formation as case studies. Int. J. Dev. Biol. 53, 783–94.CrossRefGoogle ScholarPubMed
Ball, P. (2009). Nature's Patterns: A Tapestry in Three Parts. Oxford: Oxford University Press.Google Scholar
Bar-Even, A., Paulsson, J., Maheshri, N., Carmi, M., O'Shea, E., Pilpel, Y. and Barkai, N. (2006). Noise in protein expression scales with natural protein abundance. Nat. Genetics 38, 636–43.CrossRefGoogle ScholarPubMed
Barlow, P. W. (1984). Positional controls in root development. In Positional Controls in Plant Development, ed. Barlow, P. W. and Carr, D. J., pp. 281–318. Cambridge: Cambridge University Press.Google Scholar
Battey, N. H. and Blackbourn, H. D. (1993). The control of exocytosis in plant cells. New Phytol. 125, 307–38.CrossRefGoogle Scholar
Belousov, B. P. (1959). A periodic reaction and its mechanism. In Collection of Short Papers on Radiation Medicine for 1958. Moscow: Med. Publ. Reprinted In Oscillations and Traveling Waves in Chemical Systems, eds R. J. Field and M. Burger. New York: Wiley, 1985.Google Scholar
Benková, E., Michniewicz, M., Sauer, M., Teichman, T., Seifertova, D., Jurgens, G. and Friml, J. (2003). Local, efflux-dependent auxin gradients as a common module for plant organ formation. Cell 115, 591–602.CrossRefGoogle ScholarPubMed
Berger, S. and Kaever, M. J. (1992). Dasycladales: An Illustrated Monograph of a Fascinating Algal Order. Stuttgart and New York: Thieme.Google Scholar
Bergmann, S., Sandler, O., Sberro, H., Schejter, E., Shilo, B.-Z. and Barkai, N. (2007). Pre steady-state decoding of the Bicoid morphogen gradient. PLoS Biology 5, e46.CrossRefGoogle ScholarPubMed
Bernfield, M., Banerjee, S. D., Koda, J. E. and Rapraeger, A. C. (1984). Remodelling of the basement membrane as a mechanism of morphogenetic tissue interaction. In The Role of Extracellular Matrix in Development, ed. Trelstad, R. L., pp. 545–71. New York: Alan R. Liss.Google Scholar
Blake, W. J., Kærn, M., Cantor, C. R. and Collins, J. J. (2003). Noise in eukaryotic gene expression. Nature 422, 633–7.CrossRefGoogle ScholarPubMed
Bohn, S., Andreotti, B., Douady, S., Munzinger, J. and Couder, Y. (2002). Constitutive property of the local organization of leaf venation networks. Phys. Rev. E65, 061914.Google Scholar
Bollenbach, T., Kruse, K., Pantazis, P., González-Gaitán, M. and Jüllicher, F. (2005). Robust formation of morphogen gradients. Phys. Rev. Lett. 94, 018103.CrossRefGoogle ScholarPubMed
Bollenbach, T., Pantazis, P., Kicheva, A., Bokel, C., González-Gaitán, M. and Jüllicher, F. (2008). Precision of the Dpp gradient. Development 135, 1137–46.CrossRefGoogle ScholarPubMed
Bonotto, S., and Berger, S. (1997). Proceeding/Symposium Ecology and Biology of Giant Unicellular Algae. Torino: Museo regionale di scienze naturali.Google Scholar
Borckmans, P., Dewel, G., Wit, A. and Walgraef, D. (1995). Turing bifurcations and pattern selection. In Chemical Waves and Patterns, eds. Kapral, R. and Showalter, K., pp. 323–63. Dordrecht: Kluwer.CrossRefGoogle Scholar
Bornholdt, S. (2005). Less is more in modeling large genetic networks. Science 310, 449–51.CrossRefGoogle ScholarPubMed
Braybrook, S. and Kuhlemeier, C. (2010). How a plant builds leaves. Plant Cell 22, (in press).CrossRefGoogle ScholarPubMed
Brière, C. and Goodwin, , B. C. (1988). Geometry and dynamics of tip morphogenesis in Acetabularia. J. Theor. Biol. 131, 461–75.CrossRefGoogle Scholar
Briscoe, J., Lawrence, P. A., and Vincent, J-P., eds. (2010). Generation and Interpretation of Morphogen Gradients. Cold Spring Harbor: Cold Spring Harbor Laboratory Press.Google Scholar
Byrne, G. and Cox, E. C. (1986). Spatial patterning in Polysphondylium: monoclonal antibodies specific for whorl prepatterns. Dev. Biol. 117, 442–55.CrossRefGoogle ScholarPubMed
Byrne, G. and Cox, E. C. (1987). Genesis of a spatial pattern in the cellular slime mold Polysphondylium pallidum. Proc. Nat. Acad. Sci. USA 84, 4140–4.CrossRefGoogle ScholarPubMed
Cahn, J. W. (1965). Phase separation by spinodal decomposition in isotropic systems. J. Chem. Phys. 42, 93–9.CrossRefGoogle Scholar
Cahn, J. W. and Hilliard, J. E. (1958). Free energy of a nonuniform system: I. Interfacial energy. J. Chem. Phys 28, 258–67.CrossRefGoogle Scholar
Castets, V., Dulos, E., Boissonade, J. and Kepper, P. (1990). Experimental evidence of a sustained standing Turing-type nonequilibrium chemical pattern. Phys. Rev. Lett. 64, 2953–6.CrossRefGoogle ScholarPubMed
Chadefaud, M. (1952). La leçon des algues: comment elles ont évolué; comment leur évolution peut éclairer celle des plantes supérieures. L'Année Biologique 28, C9–C25.Google Scholar
Chevaillier, P. (1970). Le noyau du spermatozöide et son évolution au cours de la spermiogenèse. In Comparative Spermatology, Vol. 5, ed. Baccetti, B., pp. 499–514. New York: Academic Press.Google Scholar
Church, A. H. (1895). The structure of the thallus of Neomeris dumetosa, Lamour. Ann. Bot. 9, 581–608.CrossRefGoogle Scholar
Church, A. H. (1919). Thalassiophyta and the subaerial transmigration. Botanical Memoirs 3, 1–95.Google Scholar
Claxton, J. H. (1964). The determination of patterns with special reference to that of the central primary skin follicles in sheep. J. Theor. Biol. 7, 302–17.CrossRefGoogle Scholar
Clyde, D., Corado, M., Wu, X., Pare, A., Papatsenko, D. and Small, S. (2003). A self-organizing system of repressor gradients establishes segmental complexity in Drosophila. Nature 426, 849–53.CrossRefGoogle ScholarPubMed
Colman-Lerner, A., Gordon, A., Serra, E., Chin, T., Resnekov, O., Endy, D., Pesce, C. G. and Brent, R. (2005). Regulated cell-to-cell variation in a cell-fate decision system. Nature 437, 699–706.CrossRefGoogle Scholar
Coppey, M., Berezhovskii, A. M., Kim, Y., Boettiger, A. N. and Shvartsman, S. Y. (2007). Modeling the bicoid gradient: diffusion and reversible nuclear trapping of a stable protein. Dev. Biol. 312, 623–30.CrossRefGoogle ScholarPubMed
Coppey, M., Boettiger, A. N., Berezhovskii, A. M. and Shvartsman, S. Y. (2008). Nuclear trapping shapes the terminal gradient in the Drosophila embryo. Curr. Biol. 18, 915–19.CrossRefGoogle ScholarPubMed
Couté, A. and Tell, G. (1981). Ultrastructure de la paroi cellulaire des Desmidiacees au microscope electronique a balayage. Vaduz: J. Cramer.Google Scholar
Crampin, E. J., Hackborn, W. W. and Maini, P. K. (2002). Pattern formation in reaction–diffusion models with nonuniform domain growth. Bull. Math. Biol. 64, 747–69.CrossRefGoogle ScholarPubMed
Crauk, O. and Dostatni, N. (2005). Bicoid determines sharp and precise target gene expression in the Drosophila embryo. Curr. Biol. 15, 1888–98.CrossRefGoogle ScholarPubMed
Crombie, A. C. (1959). Medieval and Early Modern Science, 2nd ed., 2 vols. Garden City, N Y: Doubleday.Google Scholar
Damen, W. G. M. (2007). Evolutionary conservation and divergence of the segmentation process in arthropods. Dev. Dyn. 236, 1379–91.CrossRefGoogle ScholarPubMed
Davis, G. K. and Patel, N. H. (1999). The origin and evolution of segmentation. Trends Genet. 15, M68–M72.CrossRefGoogle Scholar
Davis, G. K., Jaramillo, C. A. and Patel, N. H. (2001). Pax group III genes and the evolution of insect pair-rule patterning. Development 128, 3445–58.Google ScholarPubMed
Degn, H. (1972). Oscillating chemical reactions in homogeneous phase. J. Chem. Ed. 49, 302–7.CrossRefGoogle Scholar
Reuille, P. B., Bohn-Courseau, I., Ljung, K., et al. (2006). Computer simulations reveal properties of the cell–cell signalling network at the shoot apex in Arabidopsis. Proc. Nat. Acad. Sci. USA 103, 1627–32.CrossRefGoogle Scholar
Robertis, E. M. (2008). The molecular ancestry of segmentation mechanisms. Proc. Nat. Acad. Sci. USA 105, 16411–12.CrossRefGoogle ScholarPubMed
Digiuni, S., Schellmann, S., Geier, F., et al. (2008). A competitive complex formation mechanism underlies trichome patterning on Arabidopsis leaves. Mol. Sys. Biol. 4, 217.Google ScholarPubMed
Douady, S. and Couder, Y. (1996). Phyllotaxis as a dynamical self organizing process. I, II, III. J. Theor. Biol. 178, 255–312.CrossRefGoogle Scholar
Driever, W. and Nüsslein-Volhard, C. (1988). A gradient of bicoid protein in Drosophila embryos: the bicoid protein determines position in the Drosophila embryo in a concentration-dependent manner. Cell 54, 83–93, 95–104.CrossRefGoogle Scholar
Dumais, J. and Harrison, L. G. (2000). Whorl morphogenesis in the dasycladalean algae: the pattern formation viewpoint. Phil. Trans. R. Soc. Lond. B 355, 281–305.CrossRefGoogle ScholarPubMed
Dumais, J. and Steele, C. (2000). New evidence for the role of mechanical forces in the shoot apical meristem. J. Plant Growth Regulation 19, 7–18.CrossRefGoogle ScholarPubMed
Dumais, J., Long, S. R. and Shaw, S. L. (2004). The mechanics of surface expansion anisotropy in Medicago truncatula root hairs. Plant Physiol. 136, 3266–75.CrossRefGoogle ScholarPubMed
Dumais, J., Shaw, S. L., Steele, C. R., Long, S. R. and Ray, P. M. (2006). An anisotropic-viscoplastic model of plant cell morphogenesis by tip growth. Int. J. Dev. Biol. 50, 209–22.CrossRefGoogle ScholarPubMed
Easton, H. S., Armstrong, J. B. and Smith, S. C. (1994). Heart specification in the Mexican axolotl (Ambystoma mexicanum). Dev. Dyn. 200, 313–20.CrossRefGoogle Scholar
Edelstein-Keshet, L. (1988). Mathematical Models in Biology. New York: Random House.Google Scholar
Eldar, A., Rosin, D., Shilo, B.-Z. and Barkai, N. (2003). Self-enhanced ligand degradation underlies robustness of morphogen gradients. Dev. Cell 5, 635–46.CrossRefGoogle ScholarPubMed
Elowitz, M. B. and Leibler, S. (2000). A synthetic oscillatory network of transcriptional regulators. Nature 403, 335–8.CrossRefGoogle ScholarPubMed
Elowitz, M. B., Levine, A. J., Siggia, E. D. and Swain, P. S. (2002). Stochastic gene expression in a single cell. Science 297, 1183–6.CrossRefGoogle Scholar
Emberger, L. (1968). Les plantes fossiles dans leurs rapports avec les végétaux vivants. Paris: Masson.Google Scholar
Ermentrout, B. (1991). Stripes or spots? Nonlinear effects in bifurcation of reaction–diffusion equations on the square. Proc. R. Soc. Lond. A 434, 413–17.CrossRefGoogle Scholar
Field, R. J., Körös, E. and Noyes, R. M. (1972). Oscillations in chemical systems: II. Thorough analysis of temporal oscillations in the bromate–cerium–malonate system. J. Am. Chem. Soc. 94, 8649–64.CrossRefGoogle Scholar
Forgacs, G. and Newman, S. A. (2005). Biological Physics of the Developing Embryo. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Foty, R. A. and Steinberg, M. S. (2005). The differential adhesion hypothesis: a direct evaluation. Dev. Biol. 278, 255–63.CrossRefGoogle ScholarPubMed
Foty, R. A., Pfleger, C. M., Forgacs, G. and Steinberg, M. S. (1996). Surface tensions of embryonic cells predict their mutual envelopment behavior. Development 122, 1611–20.Google Scholar
Fowlkes, C. C., Luengo Hendriks, C. L., Keränen, S. V. E., et al. (2008). A quantitative spatio-temporal atlas of gene expression in the Drosophila blastoderm. Cell 133, 364–74.CrossRefGoogle Scholar
Frank, F. C. (1953). On spontaneous asymmetric synthesis. Biochim. Biophys. Acta 11, 459–63.CrossRefGoogle ScholarPubMed
Frankel, J. (1989). Pattern Formation: Ciliate Studies and Models. Oxford: Oxford University Press.Google Scholar
Frasch, M., Hoey, T., Rushlow, C., Doyle, H. and Levine, M. (1987). Characterization and localization of the even-skipped protein of Drosophila. EMBO J. 6, 749–59.Google ScholarPubMed
Fraser, H. B., Hirsh, A. E., Giaever, G., Kumm, J. and Eisen, M. B. (2004). Noise minimization in eukaryotic gene expression. PLoS Biology 2, 834–838.CrossRefGoogle ScholarPubMed
French, V., Bryant, P. J. and Bryant, S. V. (1976). Pattern regulation in epimorphic fields. Science 193, 969–81.CrossRefGoogle ScholarPubMed
Gardner, M. (1982). The Ambidextrous Universe, 2nd ed. Harmondsworth: Penguin Books.Google Scholar
Gibor, A. (1966). Acetabularia: a useful giant cell. Sci. Am. 215, 118–24.CrossRefGoogle ScholarPubMed
Giddings, T. H., Brower, D. L. and Staehelin, L. A. (1980). Visualization of particle complexes in the plasma membrane of Micrasterias denticulata associated with the formation of cellulose fibrils in primary and secondary cell walls. J. Cell Biol. 84, 327–39.CrossRefGoogle ScholarPubMed
Gierer, A. and Meinhardt, , H. (1972). A theory of biological pattern formation. Kybernetik 12, 30–9.CrossRefGoogle ScholarPubMed
Gilbert, S. F. (2006). Developmental Biology, 8th ed. Sunderland, MA: Sinauer Associates.Google Scholar
Gilbert, S. F. and Sarkar, S. (2000). Embracing complexity: organicism for the twenty-first century. Dev. Dyn. 219, 1–9.3.0.CO;2-A>CrossRefGoogle Scholar
Glansdorff, P. and Prigogine, , I. (1971). Thermodynamic Theory of Structure, Stability, and Fluctuations. New York: Wiley-Interscience.Google Scholar
Glazier, J. A. and Graner, F. (1993). Simulation of the differential adhesion driven rearrangement of biological cells. Phys. Rev. E 47, 2128–54.CrossRefGoogle ScholarPubMed
Goel, N. S. and Rogers, G. (1978). Computer simulations of engulfment and other movements of embryonic tissues. J. Theor. Biol. 71, 103–40.CrossRefGoogle ScholarPubMed
Goodwin, B. C. and Trainor, L. E. H. (1985). Tip and whorl morphogenesis in Acetabularia by calcium-regulated strain fields. J. Theor. Biol. 117, 79–106.CrossRefGoogle Scholar
Goriely, A. and Tabor, M. (2003). Biomechanical models of hyphal growth in actinomycetes. J. Theor. Biol. 222, 211–18.CrossRefGoogle ScholarPubMed
Graner, F. and Glazier, J. A. (1992). Simulation of biological cell sorting using a two-dimensional extended Potts model. Phys. Rev. Lett. 69, 2013–16.CrossRefGoogle ScholarPubMed
Gratzl, M. (1980). Transport of membranes and vesicle contents during exocytosis. In Biological Chemistry of Organelle Formation, eds Bücher, T., Seebald, W., and Weiss, H., pp. 165–74. Berlin, Heidelberg and New York: Springer.CrossRefGoogle Scholar
Green, P. B. (1999). Expression of pattern in plants: combining molecular and calculus-based biophysical paradigms. Am. J. Bot. 86, 1059–76.CrossRefGoogle ScholarPubMed
Green, P. B. and King, A. (1966). A mechanism for the origin of specifically oriented textures in development with special reference to Nitella wall texture. Aust. J. Biol. Sci. 19, 421–37.CrossRefGoogle Scholar
Green, P. B. and Linstead, P. (1990). A procedure for SEM of complex shoot structures applied to the inflorescence of snapdragon (Antirrhinum). Protoplasma 158, 33–8.CrossRefGoogle Scholar
Green, P. B., Steele, C. R. and Rennich, S. C. (1996). Phyllotactic patterns: a biophysical mechanism for their origin. Ann. Bot. 77, 515–27.CrossRefGoogle Scholar
Gregor, T., Bialek, W., Steveninck, R. R. R., Tank, D. W. and Wieschaus, E. F. (2005). Diffusion and scaling during early embryonic pattern formation. Proc. Nat. Acad. Sci. USA 102, 18403–7.CrossRefGoogle ScholarPubMed
Gregor, T., McGregor, A. P. and Wieschaus, E. W. (2008). Shape and function of the bicoid morphogen gradient in dipteran species with different sized embryos. Dev. Biol. 316, 350–8.CrossRefGoogle ScholarPubMed
Gregor, T., Tank, D. W., Wieschaus, E. F. and Bialek, W. (2007a). Probing the limits to positional information. Cell 130, 153–64.CrossRefGoogle ScholarPubMed
Gregor, T., Wieschaus, E. F., McGregor, A. P., Bialek, W. and Tank, D. W. (2007b). Stability and nuclear dynamics of the bicoid morphogen gradient. Cell 130, 141–52.CrossRefGoogle ScholarPubMed
Grieneisen, V. A., Xu, J., Maree, A. F. M., Hogeweg, P. and Scheres, B. (2007). Auxin transport is sufficient to generate a maximum and gradient guiding root growth. Nature 449, 1008–13.CrossRefGoogle ScholarPubMed
Gross, J. D., Peacey, M. J. and Strandmann, R. P. (1988). Plasma membrane proton pump inhibition and stalk cell differentiation in Dictyostelium discoideum. Differentiation 38, 91–8.CrossRefGoogle ScholarPubMed
Guidi, G. M. and Goldbeter, A. (2000). Oscillations and bistability predicted by a model for a cyclical bienzymatic system involving the regulated isocitrate dehydrogenase reaction. Biophys. Chem. 83, 153–70.CrossRefGoogle ScholarPubMed
Gunning, B. E. S. (1981). Microtubules and cytomorphogenesis in a developing organ: the root primordium of Azolla pinnata. In Cytomorphogenesis in Plants, ed. Kiermayer, O., pp. 301–26. Vienna: Springer-Verlag.CrossRefGoogle Scholar
Gunning, B. E. S. (1982). The root of the water fern Azolla: cellular basis of development and multiple roles for cortical microtubules. In Developmental Order: Its Origin and Regulation, ed. Subtelny, S. and Green, P. B., pp. 379–421. New York: Alan R. Liss.Google Scholar
Gursky, V. V., Kozlov, K. N., Samsonov, A. M. and Reinitz, J. (2006). Cell divisions as a mechanism for selection in stable steady states of multi-stationary gene circuits. Physica D 218, 70–6.CrossRefGoogle Scholar
Hafen, E., Kuroiwa, A. and Gehring, W. J. (1984). Spatial distribution of transcripts from the segmentation gene fushi tarazu during Drosophila embryonic development. Cell 37, 833–41.CrossRefGoogle ScholarPubMed
Hagemann, W. (1992). The relationship of anatomy to morphology in plants: a new theoretical perspective. Int. J. Plant Sci. 153, S38–S48.CrossRefGoogle Scholar
Hamant, O., Heisler, M. G., Jönsson, H., et al. (2008). Developmental patterning by mechanical signals in Arabidopsis. Science 322, 1650–5.CrossRefGoogle ScholarPubMed
Hardway, H., Mukjopadhyay, B., Burke, T., Hichman, T. J. and Forman, R. (2008). Modeling the precision and robustness of the Hunchback boundary during Drosophila embryonic development. J. Theor. Biol. 254, 390–9.CrossRefGoogle ScholarPubMed
Harrison, L. G. (1973). Evolution of biochemical systems with specific chirality: a model involving territorial behaviour. J. Theor. Biol. 39, 333–41.CrossRefGoogle Scholar
Harrison, L. G. (1974). The possibility of spontaneous resolution of enantiomers on a catalyst surface. J. Mol. Evol. 4, 99–111.CrossRefGoogle ScholarPubMed
Harrison, L. G. (1979). Molecular asymmetry and morphology: big hands from little hands. In Origins of Optical Activity in Nature, ed. Walker, D. C., pp. 125–40. Amsterdam: Elsevier.Google Scholar
Harrison, L. G. (1981). Physical chemistry of biological morphogenesis. Chem. Soc. Rev. 10, 491–528.CrossRefGoogle Scholar
Harrison, L. G. (1982). An overview of kinetic theory in developmental modelling. In Developmental Order: Its Origin and Regulation, ed. Subtelny, S. and Green, P. B., pp. 3–33. New York: Alan R. Liss.Google Scholar
Harrison, L. G. (1992). Reaction–diffusion theory and intracellular differentiation. Int. J. Plant Sci.153, S76–S85.CrossRefGoogle Scholar
Harrison, L. G. (1993). Kinetic Theory of Living Pattern. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Harrison, L. G. (1994). Kinetic theory of living pattern. Endeavour 18, 130–6.CrossRefGoogle ScholarPubMed
Harrison, L. G. and Hillier, N. A. (1985). Quantitative control of Acetabularia morphogenesis by extracellular calcium: a test of kinetic theory. J. Theor. Biol. 114, 177–92.CrossRefGoogle Scholar
Harrison, L. G. and Kolář, M. (1988). Coupling between reaction–diffusion prepattern and expressed morphogenesis. J. Theor. Biol. 130, 493–515.CrossRefGoogle Scholar
Harrison, L. G. and Lacalli, T. C. (1978). Hyperchirality: a mathematically convenient and biochemically possible model for the kinetics of morphogenesis. Proc. R. Soc. Lond. B202, 361–97.CrossRefGoogle Scholar
Harrison, L. G. and Tan, K. Y. (1988). Where may reaction–diffusion mechanisms be operating in metameric patterning of Drosophila embryos?BioEssays 8, 118–24.CrossRefGoogle ScholarPubMed
Harrison, L. G. and Aderkas, P. (2004). Spatially quantitative control of the number of cotyledons in a clonal population of somatic embryo of hybrid larch Larix X leptoeuropaea. Ann. Bot. 93, 423–34.CrossRefGoogle Scholar
Harrison, L. G., Graham, K. T. and Lakowski, B. C. (1988). Calcium localization during Acetabularia whorl formation: evidence supporting a two-stage hierarchical mechanism. Development 104, 255–62.Google Scholar
Harrison, L. G., Holloway, D. M. and Orchard, J. J. (1996). Hearts and somites: problems of pattern formation in vertebrate embryology modelled by reaction–diffusion. Dev. Biol. 175, 73.Google Scholar
Harrison, L. G., Wehner, S. and Holloway, D. M. (2001). Complex morphogenesis of surfaces: theory and experiment on coupling of reaction–diffusion to growth. Faraday Disc. 120, 277–94.CrossRefGoogle Scholar
Harrison, L. G., Kasinsky, H. E., Ribes, E. and Chiva, M. (2005). Possible mechanisms for early and intermediate stages of sperm chromatin condensation patterning involving phase separation dynamics. J. Exp. Zool. 303A, 76–92.CrossRefGoogle Scholar
Harrison, L. G., Snell, J., Verdi, R., Vogt, D. E., Zeiss, G. D. and Green, , B. R. (1981). Hair morphogenesis in Acetabularia mediterranea: temperature-dependent spacing and models of morphogen waves. Protoplasma 106, 211–21.CrossRefGoogle Scholar
Harrison, L. G., Donaldson, G., Lau, W., et al. (1997). CaEGTA uncompetitively inhibits calcium activation of whorl morphogenesis in Acetabularia. Protoplasma 196, 190–6.CrossRefGoogle Scholar
Hartmann, C., Taubert, H., Jäckle, H. and Pankratz, M. J. (1994). A 2-step mode of stripe formation in the Drosophila blastoderm requires interactions among primary pair-rule genes. Mech. of Dev. 45, 3–13.CrossRefGoogle Scholar
He, F., Wen, Y., Lin, X., et al. (2008). Probing intrinsic properties of a robust morphogen gradient in Drosophila. Dev. Cell 15, 558–67.CrossRefGoogle ScholarPubMed
Herschkowitz-Kaufman, M. (1975). Bifurcation analysis of nonlinear reaction–diffusion equations: II. Steady-state solutions and comparison with numerical simulations. Bull. Math. Biol. 37, 589–636.CrossRefGoogle Scholar
Holloway, D. M. (1995). Reaction–Diffusion Theory of Localized Structures with Application to Vertebrate Organogenesis. PhD thesis, University of British Columbia, Canada.Google Scholar
Holloway, D. M. (2010). The role of chemical dynamics in plant morphogenesis. Bioch. Soc. Trans. 38, 645–50.CrossRefGoogle ScholarPubMed
Holloway, D. M. and Harrison, L. G. (1995). Order and localization in reaction–diffusion pattern. Physica A 222, 210–33.CrossRefGoogle Scholar
Holloway, D. M. and Harrison, L. G. (1999a). Algal morphogenesis: modelling interspecific variation in Micrasterias with reaction–diffusion patterned catalysis of cell surface growth. Phil. Trans. R. Soc. Lond. B 354, 417–33.CrossRefGoogle Scholar
Holloway, D. M. and Harrison, L. G. (1999b). Suppression of positional errors in biological development. Math. Biosc. 156, 271–90.CrossRefGoogle ScholarPubMed
Holloway, D. M. and Harrison, L. G. (2008). Pattern selection in plants: coupling chemical dynamics to surface growth in three dimensions. Ann. Bot. 101, 361–74.CrossRefGoogle ScholarPubMed
Holloway, D. M., Harrison, L. G. and Armstrong, J. B. (1994). Computations of post-inductive dynamics in axolotl heart formation. Dev. Dyn. 200, 242–56.CrossRefGoogle ScholarPubMed
Holloway, D. M., Harrison, L. G. and Spirov, A. V. (2003). Noise in the segmentation gene network of Drosophila, with implications for mechanisms of body axis specification. Proc. SPIE 5110, 180–91.CrossRefGoogle Scholar
Holloway, D. M., Harrison, L. G., Kosman, D., Vanario-Alonso, C. E. and Spirov, A. V. (2006). Analysis of pattern precision shows that Drosophila segmentation develops substantial independence from gradients of maternal gene products. Dev. Dyn. 235, 2949–60.CrossRefGoogle ScholarPubMed
Holzinger, A., Callaham, D. A., Hepler, P. K. and Meindl, U. (1995). Free calcium in Micrasterias: local gradients are not detected in growing lobes. Eur. J. Cell Biol. 67, 363–71.Google Scholar
Houchmandzadeh, B., Wieschaus, E. and Leibler, S. (2002). Establishment of developmental precision and proportions in the early Drosophila embryo. Nature 415, 798–802.CrossRefGoogle ScholarPubMed
Houchmandzadeh, B., Wieschaus, E. and Leibler, S. (2005). Precise domain specification in the developing Drosophila embryo. Phys. Rev. E 72, 061920.CrossRefGoogle ScholarPubMed
Howard, M. and Kruse, K. (2005). Cellular organization by self-organization: mechanisms and models for Min protein dynamics. J. Cell Biol. 168, 533–6.CrossRefGoogle ScholarPubMed
Howard, M. and Rein ten Wolde, P. (2005). Finding the center reliably: robust patterns of developmental gene expression. Phys. Rev. Lett. 95, 208103.CrossRefGoogle ScholarPubMed
Howard, M., Rutenberg, A. D. and Vet, S. (2001). Dynamic compartmentalization of bacteria: accurate division in E. coli. Phys. Rev. Lett. 87, 278102.CrossRefGoogle ScholarPubMed
Humphrey, R. R. (1972). Genetic and experimental studies on a mutant (c) determining absence of heart action in embryos of the Mexican axolotl (Ambystoma mexicanum). Dev. Biol. 27, 365–75.CrossRefGoogle Scholar
Hunding, A. (1989). Turing patterns of the second kind simulated on supercomputers in three curvilinear coordinates and time. In Cell to Cell Signalling: From Experiments to Theoretical Models, ed. Goldbeter, A., pp. 229–36. London: Academic Press.CrossRefGoogle Scholar
Hunding, A. (1993). Supercomputer simulation of Turing structures in Drosophila morphogenesis. In Experimental and Theoretical Advances in Biological Pattern Formation, eds. Othmer, H. G., Maini, P. K., and Murray, J. D., pp. 149–59. New York: Plenum Press.CrossRefGoogle Scholar
Hunding, A. (2004). Microtubule dynamics may embody a stationary bipolarity forming mechanism related to the prokaryotic division site mechanism (pole-to-pole oscillations). J. Biol. Phys. 30, 325–44.CrossRefGoogle Scholar
Ingram, G. C., Goodrich, J., Wilkinson, M. D., Simon, R., Haughn, G. W. and Coen, E. S. (1995). Parallels between UNUSUAL FLORAL ORGANS and FIMBRIATA, genes controlling flower development in Arabidopsis and Antirrhinum. Plant Cell 7, 1501–10.CrossRefGoogle ScholarPubMed
Irish, V., Lehmann, R. and Akam, M. (1989). The Drosophila posterior-group gene nanos functions by repressing hunchback activity. Nature 338, 646–8.CrossRefGoogle ScholarPubMed
Isaacs, F. J, Blake, W. J. and Collins, J. J. (2005). Signal processing in single cells. Science 307, 1886–88.CrossRefGoogle ScholarPubMed
Isalan, M. (2009). Gene networks and liar paradoxes. BioEssays 31, 1110–15.CrossRefGoogle ScholarPubMed
Jacobson, A. G. (1960). Influences of ectoderm and endoderm on heart differentiation in the newt. Dev. Biol. 2, 138–54.CrossRefGoogle ScholarPubMed
Jacobson, A. G. and Duncan, J. T. (1968). Heart induction in salamanders. J. Exp. Zool. 167, 79–103.CrossRefGoogle ScholarPubMed
Jaeger, J. (2009). Modelling the Drosophila embryo. Mol. BioSyst. 5, 1549–68.CrossRefGoogle ScholarPubMed
Jaeger, J. and Goodwin, , B. C. (2001). A cellular oscillator model for periodic pattern formation. J. Theor. Biol. 213, 171–81.CrossRefGoogle ScholarPubMed
Jaeger, J., Blagov, M., Kosman, D., et al. (2004a). Dynamical analysis of regulatory interactions in the gap gene system of Drosophila melanogaster. Genetics 167, 1721–37.CrossRefGoogle ScholarPubMed
Jaeger, J., Surkova, S., Blagov, M., et al. (2004b). Dynamic control of positional information in the early Drosophila embryo. Nature 430, 368–71.CrossRefGoogle ScholarPubMed
Jiang, T. X., Jung, H. S., Widelitz, R. B. and Chuong, C. M. (1999). Self-organization of periodic patterns by dissociated feather mesenchymal cells and the regulation of size, number and spacing of primordia. Development 126, 4997–5009.Google ScholarPubMed
Jönsson, H., Heisler, M. G., Shapiro, B. E., Mjolsness, E. and Meyerowitz, E. M. (2006). An auxin-driven polarized transport model for phyllotaxis. Proc. Nat. Acad. Sci. USA 103, 1633–8.CrossRefGoogle ScholarPubMed
Jung, H. S., Francis-West, P. H., Widelitz, R. B., et al. (1998). Local inhibitory action of BMPs and their relationships with activators in feather formation: implications for periodic patterning. Dev. Biol. 196, 11–23.CrossRefGoogle ScholarPubMed
Kaandorp, J. A. (1994). Fractal Modelling. Berlin: Springer.CrossRefGoogle Scholar
Kaern, M., Menzinger, M. and Hunding, A. (2000). Segmentation and somitogenesis derived from phase dynamics in growing oscillatory media. J. Theor. Biol. 207, 473–93.CrossRefGoogle ScholarPubMed
Kaern, M., Elston, T. C., Blake, W. J. and Collins, J. J. (2005). Stochasticity in gene expression: from theories to phenotypes. Nature Rev. Genet. 6, 451–64.CrossRefGoogle ScholarPubMed
Kaern, M., Menzinger, M., Satnoianu, R. and Hunding, A. (2001). Chemical waves in open flows of active media: their relevance to axial segmentation in biology. Faraday Disc. 120, 295–312.CrossRefGoogle Scholar
Käfer, J., Hogeweg, P. and Marée, A. F. M. (2006). Moving forward moving backward: directional sorting of chemotactic cells due to size and adhesion differences. PLoS Comput. Biol. 2, e56.CrossRefGoogle ScholarPubMed
Kaplan, D. R. (1992). The relationship of cells to organisms in plants: problem and implications of an organismal perspective. Int. J. Plant Sci. 153, S28–S37.CrossRefGoogle Scholar
Kauffman, S. A., Shymko, R. M. and Trabert, K. (1978). Control of sequential compartment formation in Drosophila. Science 199, 259–70.CrossRefGoogle ScholarPubMed
Kauzmann, W. (1957). Quantum Chemistry. New York: Academic Press.Google Scholar
Kent, L. J. and Knight, E. C., eds. (1969). Selected Writings of E. T. A. Hoffmann, 2 vols. Chicago and London: University of Chicago Press.Google Scholar
Kerszberg, M. and Wolpert, L. (2007). Specifying positional information in the embryo: looking beyond morphogens. Cell 130, 205–9.CrossRefGoogle ScholarPubMed
Kiermayer, O. (1964). Untersuchungen über die Morphogenese und Zellwandbildung bei Micrasterias denticulata Bréb. Protoplasma 59, 382–420.CrossRefGoogle Scholar
Kiermayer, O. (1967). Das Septum-Initialmuster von Micrasterias denticulata und seine Bildung. Protoplasma 64, 481–4.CrossRefGoogle Scholar
Kiermayer, O. (1968). Hemmung der Kern- und Chloroplasten-migration von Micrasterias durch Colchicin. Naturwissenschaften 55, 299–300.CrossRefGoogle Scholar
Kiermayer, O. (1970). Causal aspects of cytomorphogenesis in Micrasterias. Ann. NY Acad. Sci. 175, 686–701.CrossRefGoogle Scholar
Kiermayer, O. (1981). Cytoplasmic basis of morphogenesis in Micrasterias. In Cytomorphogenesis in Plants, ed. Kiermayer, O.. Wien and New York: Springer.CrossRefGoogle Scholar
Knight, J. (2003). Scientific literacy: clear as mud. Nature 423, 376–8.CrossRefGoogle Scholar
Ko, M. S. H. (1991). A stochastic model for gene induction. J. Theor. Biol. 153, 181.CrossRefGoogle ScholarPubMed
Ko, M. S. H. (1992). Induction mechanism of a single gene molecule: stochastic or deterministic?BioEssays 14, 341–6.CrossRefGoogle ScholarPubMed
Kondo, S. and Asai, R. (1995). A reaction–diffusion wave on the skin of the marine angelfish Pomocanthus. Nature 376, 765–8.CrossRefGoogle Scholar
Kondratiev, V. N. (1969). The Theory of Kinetics. New York: Elsevier.Google Scholar
Kosman, D., Reinitz, J. and Sharp, D. H. (1997). Automated assay of gene expression at cellular resolution. In Proceedings of the 1998 Pacific Symposium on Biocomputing, eds Altman, R., Dunker, K., Hunter, L. and Klein, T., pp. 6–17. Singapore: World Scientific Press.Google Scholar
Kosman, D., Small, S. and Reinitz, J. (1998). Rapid preparation of a panel of polyclonal antibodies to Drosophila segmentation proteins. Dev. Genes Evol. 208, 290–4.CrossRefGoogle ScholarPubMed
Kuhn, T. S. (1962). The Structure of Scientific Revolutions. Chicago: University of Chicago Press.Google Scholar
Kuijt, J. (1967). On the structure and origin of the seedling of Psittacanthus schiedeanus (Loranthaceae). Can. J. Bot. 45, 1497–506.CrossRefGoogle Scholar
Kwiatkowska, D. and Dumais, J. (2003). Growth and morphogenesis at the vegetative shoot apex of Anagallis arvensis L. J. Exp. Bot. 54, 1585–95.CrossRefGoogle ScholarPubMed
Lacalli, T. C. (1973a). Morphogenesis in Micrasterias. PhD thesis, University of British Columbia, Canada.Google Scholar
Lacalli, T. C. (1973b). Cytokinesis in Micrasterias rotata. Protoplasma 78, 433–42.CrossRefGoogle Scholar
Lacalli, T. C. (1975). Morphogenesis in Micrasterias. I: Tip growth. II: Patterns of morphogenesis. J. Embryol. Exp. Morphol. 33, 95–116, 117–27.Google ScholarPubMed
Lacalli, T. C. (1981). Dissipative structures and morphogenetic pattern in unicellular algae. Phil. Trans. R. Soc. Lond. B 294, 547–88.CrossRefGoogle Scholar
Lacalli, T. C. (1990). Modeling the Drosophila pair-rule pattern by reaction–diffusion: gap input and pattern control in a 4-morphogen system. J. Theor. Biol. 144, 171–94.CrossRefGoogle Scholar
Lacalli, T. C. and Harrison, L. G. (1978a). The regulatory capacity of Turing's model for morphogenesis, with application to slime moulds. J. Theor. Biol. 70, 273–95.CrossRefGoogle ScholarPubMed
Lacalli, T. C. and Harrison, L. G. (1978b). Development of ordered arrays of cell wall pores in Desmids: a nucleation model. J. Theor. Biol. 74, 109–38.CrossRefGoogle ScholarPubMed
Lacalli, T. C. and Harrison, L. G. (1991). From gradients to segments: models for pattern formation in early Drosophila embryogenesis. Sem. Dev. Biol. 2, 107–17.Google Scholar
Lacalli, T. C., Wilkinson, D. A. and Harrison, L. G. (1988). Theoretical aspects of stripe formation in relation to Drosophila segmentation. Development 104, 105–13.Google ScholarPubMed
Lander, A. D. (2007). Morpheus unbound: reimagining the morphogen gradient. Cell 128, 245–56.CrossRefGoogle ScholarPubMed
Lander, A. D., Nie, Q. and Wan, F. Y. (2002). Do morphogen gradients arise by diffusion?Dev. Cell 2, 785–96.CrossRefGoogle ScholarPubMed
Lander, A. D., Lo, W.-C., Nie, Q. and Wan, F. Y. M. (2009). The measure of success: constraints, objectives, and tradeoffs in morphogen-mediated patterning. Cold Spring Harb. Perspect. Biol. 1, a002022.CrossRefGoogle ScholarPubMed
Laskowski, M., Grieneisen, V. A., Hofhuis, H., et al. (2008). Root system architecture from coupling cell shape to auxin transport. PLoS Biol. 6, e307.CrossRefGoogle ScholarPubMed
Lengyel, I. and Epstein, I. R. (1991). Modeling of Turing structures in the chlorite–iodide–malonic acid–starch reaction system. Science 251, 650–2.CrossRefGoogle ScholarPubMed
Li, R. and Bowerman, B., eds. (2010). Symmetry Breaking in Biology. Cold Spring Harbor: Cold Spring Harbor Laboratory Press.Google ScholarPubMed
Lisman, J. E. (1985). A mechanism for memory storage insensitive to molecular turnover: a bistable autophosphorylating kinase. Proc. Nat. Acad. Sci. USA 82, 3055–7.CrossRefGoogle ScholarPubMed
Lopes, F. J. P., Vieira, F. M. C., Holloway, D. M., Bisch, P. M. and Spirov, A. V. (2008). Spatial bistability generates hunchback expression sharpness in the Drosophila embryo. PLoS Comp. Biol. 4, e1000184.CrossRefGoogle ScholarPubMed
Lott, S. E., Kreitman, M., Palsson, A., Alekseeva, E. and Ludwig, M. Z. (2007). Canalization of segmentation and its evolution in Drosophila. Proc. Nat. Acad. Sci. USA 104, 10926–31.CrossRefGoogle ScholarPubMed
Lucchetta, E. M., Vincent, M. E. and Ismagilov, R. F. (2008). A precise Bicoid gradient is nonessential during cycles 11–13 for precise patterning in the Drosophila blastoderm. PLoS ONE 3, e3651.CrossRefGoogle ScholarPubMed
Lucchetta, E. M., Lee, J. H., Fu, L. A., Patel, N. H. and Ismagilov, R. F. (2005). Dynamics of Drosophila embryonic patterning network perturbed in space and time using microfluidics. Nature 434, 1134–8.CrossRefGoogle ScholarPubMed
Lyons, M. J. and Harrison, L. G. (1991). A class of reaction–diffusion mechanisms which preferentially select striped patterns. Chem. Phys. Lett. 183, 158–64.CrossRefGoogle Scholar
Lyons, M. J. and Harrison, L. G. (1992a). Stripe selection: an intrinsic property of some pattern-forming models with nonlinear dynamics. Dev. Dyn. 195, 201–15.CrossRefGoogle ScholarPubMed
Lyons, M. J. and Harrison, L. G. (1992b). Non-linear analysis of models for biological pattern formation: application to ocular dominance stripes. In Analysis and Modeling of Neural Systems II, ed. Eeckman, F. H.. Norwell, MD: Kluwer Academic.Google Scholar
Lyons, M. J., Harrison, L. G., Lakowski, B. C. and Lacalli, T. C. (1990). Reaction–diffusion modelling of biological pattern formation: application to the embryogenesis of Drosophila melanogaster. Can. J. Phys. 68, 772–7.CrossRefGoogle Scholar
Maini, P. K., Baker, R. E. and Chuong, C. M. (2006). The Turing model comes of molecular age. Science 314, 1397–8.CrossRefGoogle ScholarPubMed
Mandoli, D. F. (1998). Elaboration of body plan and phase change during development of Acetabularia: how is the complex architecture of a giant unicell built?A. Rev. Plant Physiol. Plant Mol. Biol. 49, 173–98.CrossRefGoogle Scholar
Manu, Surkova S., Spirov, A. V., Gursky, V. V., et al. (2009a). Canalization of gene expression in the Drosophila blastoderm by gap gene cross regulation. PLoS Biology 7, e1000049.CrossRefGoogle ScholarPubMed
Manu, Surkova S., Spirov, A. V., Gursky, V. V., et al. (2009b). Canalization of gene expression and domain shifts in the Drosophila blastoderm by dynamical attractors,PLoS Comp. Biol. 5, e1000303.CrossRefGoogle ScholarPubMed
Marée, A. F. M. and Hogeweg, P. (2001). How amoeboids self-organize into a fruiting body: multicellular coordination in Dictyostelium discoideum. Proc. Nat. Acad. Sci. USA. 98, 3879–83.CrossRefGoogle ScholarPubMed
Marée, A. F. M. and Hogeweg, P. (2002). Modelling Dictyostelium discoideum morphogenesis: the culmination. Bull. Math. Biol. 64, 327–53.CrossRefGoogle ScholarPubMed
Marée, S. (2000). From Pattern Formation to Morphogenesis: Multicellular Coordination in Dictyostelium discoideum. PhD thesis. University of Utrecht, Netherlands.Google Scholar
Martens, G., Humphrey, E. C., Harrison, L. G., et al. (2009). High-pressure freezing of spermiogenic nuclei supports a dynamic chromatin model for the histone-to-protamine transition. J. Cell. Biochem. 108, 1399–409.CrossRefGoogle ScholarPubMed
Matela, R. J. and Fletterick, R. J. (1980). Computer simulation of cellular self-sorting: a topological exchange model. J. Theor. Biol. 84, 673–90.CrossRefGoogle ScholarPubMed
McNally, J. G. and Cox, E. C. (1989). Spots and stripes: the patterning spectrum in the cellular slime mould Polysphondylium pallidum. Development 105, 323–33.Google Scholar
Meindl, U. (1982). Local accumulations of membrane associated calcium according to cell pattern formation in Micrasterias denticulata, visualized by chlorotetracycline fluorescence. Protoplasma 110, 143–6.CrossRefGoogle Scholar
Meinhardt, H. (1982). Models of Biological Pattern Formation. London: Academic Press.Google Scholar
Meinhardt, H. (1984). Models of pattern formation and their application to plant development. In Positional Controls in Plant Development, ed. Barlow, P. W. and Carr, D. J., pp. 1–32. Cambridge: Cambridge University Press.Google Scholar
Meinhardt, H. (1995). The Algorithmic Beauty of Seashells. 1st edn, Berlin, Heidelberg and New York:Springer-Verlag. (2nd edn 1998; 3rd edn, 2003).CrossRefGoogle Scholar
Meinhardt, H. (2001). Organizer and axes formation as a self-organizing process. Int. J. Dev. Biol. 45, 177–88.Google ScholarPubMed
Meinhardt, H. and Boer, P. A. J. (2001). Pattern formation in Esherichia coli: a model for pole-to-pole oscillations of min proteins and the localization of the division site. Proc. Nat. Acad. Sci. USA 98, 14202–7.CrossRefGoogle Scholar
Mills, W. H. (1932). Stereochemistry and catalysis. Chem. Ind., 750–9.CrossRefGoogle Scholar
Milo, R., Shen-Orr, S., Itzkovitz, S., Kashtan, N., Chklovskii, D. and Alon, U. (2002). Network motifs: simple building blocks of complex networks. Science 298, 824–7.CrossRefGoogle ScholarPubMed
Mitchison, G. J. (1980). A model for vein formation in higher plants. Proc. R. Soc. Lond. B 207, 79–109.CrossRefGoogle Scholar
Mitchison, G. J. (1981). The polar transport of auxin and vein patterns in plants. Phil. Trans. R. Soc. Lond. B 295, 461–71.CrossRefGoogle Scholar
Mizutani, C. M., Nie, Q., Wan, F. Y. M., et al. (2005). Formation of the BMP activity gradient in the Drosophila embryo. Dev. Cell 8, 915–24.CrossRefGoogle ScholarPubMed
Moore, K. L. (1974). The Developing Human: Clinically Oriented Embryology. Philadelphia, London and Toronto: W. B. Saunders Co.Google Scholar
Moore, W. J. (1972). Physical Chemistry, 4th edn. Englewood Cliffs, N J: Prentice-Hall.Google Scholar
Muratov, C. B. and Shvartsman, S. Y. (2003). An asymptotic study of the inductive pattern formation mechanism in Drosophila egg development. Physica D 186, 93–108.CrossRefGoogle Scholar
Murray, J. D. (1981a). A pre-pattern formation mechanism for animal coat markings. J. Theor. Biol. 88, 161–99.CrossRefGoogle Scholar
Murray, J. D. (1981b). On pattern formation mechanisms for lepidopteran wing patterns and mammalian coat markings. Phil. Trans. R. Soc. Lond. B 295, 473–96.CrossRefGoogle ScholarPubMed
Murray, J. D. (1988). How the leopard gets its spots. Sci. Am. 258, 80–7.CrossRefGoogle Scholar
Murray, J. D. (1989). Mathematical Biology. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Nagata, W., Harrison, L. G. and Wehner, S. (2003). Reaction–diffusion models of growing plant tips: bifurcations on hemispheres. Bull. Math. Biol. 65, 571–607.CrossRefGoogle ScholarPubMed
Nägeli, C. (1847). Die neuern Algensysteme. Zürich, Switzerland: Schulthess.Google Scholar
Nagorcka, B. N. (1988). A pattern formation mechanism to control spatial organization in the embryo of Drosophila melanogaster. J. Theor. Biol. 132, 277–306.CrossRefGoogle ScholarPubMed
Nagorcka, B. N. and Mooney, J. R. (1982). The role of a reaction–diffusion system in the formation of hair fibres. J. Theor. Biol. 98, 575–607.CrossRefGoogle ScholarPubMed
Nagorcka, B. N. and Mooney, J. R. (1985). The role of a reaction–diffusion system in the initiation of primary hair follicles. J. Theor. Biol. 114, 243–72.CrossRefGoogle ScholarPubMed
Neville, A. A., Matthews, P. C. and Byrne, H. M. (2006). Interactions between pattern formation and domain growth. Bull. Math. Biol. 68, 1975–2003.CrossRefGoogle ScholarPubMed
Newman, S. A. (1988). Lineage and pattern in the developing vertebrate limb. Trends Genet. 4, 329–32.CrossRefGoogle ScholarPubMed
Newman, S. A. (1996). Sticky fingers: Hox genes and cell adhesion in vertebrate limb development. BioEssays 18, 171–4.CrossRefGoogle ScholarPubMed
Newman, S. A. and Frisch, H. L. (1979). Dynamics of skeletal pattern formation in developing chick limb. Science 205, 662–8.CrossRefGoogle ScholarPubMed
Nicolis, G. and Prigogine, , I. (1977). Self-Organization in Non-Equilibrium Systems. New York: Wiley.Google Scholar
Nieuwkoop, P. D. (1969). The formation of the mesoderm in urodele amphibians: I. Induction by the endoderm. Wilhelm Roux Arch. Entwicklungsmech. Org. 162, 341–73.CrossRefGoogle Scholar
Nieuwkoop, P. D. (1973). The ‘organisation center’ of the amphibian embryo: its origin, spatial organisation and morphogenetic action. Adv. Morphogenet. 10, 1–310.CrossRefGoogle Scholar
Nieuwkoop, P. D. (1977). Origin and establishment of embryonic polar axes in amphibian development. Curr. Top. Dev. Biol. 11, 115–32.CrossRefGoogle ScholarPubMed
Nieuwkoop, P. D., Johnen, A. G. and Albers, B. (1985). The Epigenetic Nature of Early Chordate Development. Cambridge: Cambridge University Press.Google Scholar
Nishimura, N. J. and Mandoli, D. F. (1992). Vegetative growth of Acetabularia acetabulum (Chlorophyta): structural evidence for juvenile and adult phases in development. J. Phycol. 28, 669–77.CrossRefGoogle Scholar
Okabe-Oho, Y., Murakami, H., Oho, S. and Sasai, M. (2009). Stable, precise, and reproducible patterning of Bicoid and Hunchback molecules in the early Drosophila embryo. PLoS Comp. Biol. 5: e1000486.CrossRefGoogle ScholarPubMed
Orchard, J. J. (1996). Reaction–diffusion Modelling of Somite Formation: Computed Dynamics and Bifurcation Analysis. MSc thesis, University of British Columbia, Canada.Google Scholar
Oster, G. F., Murray, J. D. and Harris, A. K. (1983). Mechanical aspects of mesenchymal morphogenesis. J. Embryol. Exp. Morphol. 78, 83–125.Google ScholarPubMed
Othmer, H. and Pate, E. (1980). Scale invariance in reaction–diffusion models of spatial pattern formation. Proc. Nat. Acad. Sci. USA 77, 4180–4.CrossRefGoogle ScholarPubMed
Ouyang, Q. and Swinney, H. L. (1991). Transition from a uniform state to hexagonal and striped Turing patterns. Nature 352, 610–11.CrossRefGoogle Scholar
Page, K. M., Maini, P. K., Monk, N. A. M. and Stern, C. D. (2001). A model of primitive streak initiation in the chick embryo, J. Theor. Biol. 208, 419–38.CrossRefGoogle ScholarPubMed
Palmeirim, I., Henrique, D., Ish-Horowicz, D. and Pourquié, O. (1997). Avian hairy gene expression identifies a molecular clock linked to vertebrate segmentation and somitogenesis. Cell 91, 639–48.CrossRefGoogle ScholarPubMed
Palsson, E. (2007). Modeling wave propagation, chemotaxis, cell adhesion and cell sorting: examples with Dictyostelium, using a 3-D cell-based model. In Modeling Biology, Structures, Behaviors, Evolution, Vienna Series in Theoretical Biology, ed. Laubichler, M. D. and Muller, G. D., pp. 165–94. Cambridge, MA: MIT Press.Google Scholar
Palsson, E. and Cox, E. C. (1997). Selection for spiral waves in the social amoebae Dictyostelium.Proc. Nat. Acad. Sci. USA 94, 13719–23.CrossRefGoogle ScholarPubMed
Pankratz, M. J., Gaul, U., Hoch, M., et al. (1990). Overlapping gene activities generate pair-rule stripes and delimit the expression domains of homeotic genes along the longitudinal axis of the Drosophila blastoderm embryo. In Genetics of Pattern Formation and Growth Control, ed. Mahowald, A. P., pp. 17–29. New York: Wiley-Liss.Google Scholar
Papatsenko, D. (2009). Stripe formation in the early fly embryo: principles, models, and networks. BioEssays 31, 1172–80.CrossRefGoogle Scholar
Peck, A. L. (trans.) (1942). Aristotle: Generation of Animals. Cambridge, MA: Harvard University Press.Google Scholar
Pedraza, J. M. and Oudenaarden, A. (2005). Noise propagation in gene networks. Science 307, 1965–9.CrossRefGoogle ScholarPubMed
Peel, A. (2004). The evolution of arthropod segmentation mechanisms. Bioessays 26: 1108–16.CrossRefGoogle ScholarPubMed
Peel, A., Chipman, A. D. and Akam, M. (2005). Arthropod segmentation: beyond the Drosophila paradigm. Nat. Rev. Genet. 6, 905–16.CrossRefGoogle ScholarPubMed
Phillips, H. M. and Steinberg, M. S. (1969). Equilibrium measurements of embryonic chick cell adhesiveness: I. Shape equilibrium in centrifugal fields. Proc. Nat. Acad. Sci. USA 64, 121–7.CrossRefGoogle ScholarPubMed
Pick, L. (1998). Segmentation: painting stripes from flies to vertebrates. Dev. Genetics 23, 1–10.3.0.CO;2-A>CrossRefGoogle ScholarPubMed
Potts, R. B. (1952). Some generalized order–disorder transformations. Proc. Camb. Phil. Soc. 48, 106−9.CrossRefGoogle Scholar
Pourquié, O. (2003). The segmentation clock: converting embryonic time into spatial pattern. Science 301, 328–30.CrossRefGoogle ScholarPubMed
Poustelnikova, E., Pisarev, A., Blagov, M., Samsonova, M. and Reinitz, J. (2004). A database for management of gene expression data in situ. Bioinformatics 20, 2212–21.CrossRefGoogle ScholarPubMed
Prescott, G. W., Croasdale, H. T. and Vinyard, W. C. (1977). A Synopsis of North American Desmids: II. Desmidiaceae: Placodermae, section 2. Lincoln and London: University of Nebraska Press.Google Scholar
Prigogine, I. (1967). Dissipative structures in chemical systems. In Fast Reactions and Primary Processes in Chemical Kinetics, ed. Claesson, S., pp. 371–82. New York: Interscience.Google Scholar
Prigogine, I. and Lefever, R. (1968). Symmetry-breaking instabilities in dissipative systems: II. J. Chem. Phys. 48, 1695–700.CrossRefGoogle Scholar
Pueyo, J. I., Lanfear, R. and Couso, J. P. (2008). Ancestral Notch-mediated segmentation revealed in the cockroach Periplaneta americana. Proc. Nat. Acad. Sci. USA 105, 16614–19.CrossRefGoogle ScholarPubMed
Puiseux-Dao, S. (1962). Recherches biologiques et physiologiques sur quelques Dasycladacées, en particulier, le Batophora oerstedii et l'Acetabularia mediterranea Lam. Rev. Gen. Bot. 69, 409–503.Google Scholar
Rao, C. V. and Arkin, A. P. (2001). Control motifs for intracellular regulatory networks. Annu. Rev. Biomed. Eng. 3, 391–419.CrossRefGoogle ScholarPubMed
Rao, C. V., Wolf, D. M. and Arkin, A. P. (2002). Control, exploitation and tolerance of intracellular noise. Nature 420, 231–7.CrossRefGoogle ScholarPubMed
Rashevsky, N. (1940). An approach to the mathematical biophysics of biological self-regulation and of cell polarity. Bull. Math. Biophys. 2, 15–25.CrossRefGoogle Scholar
Reddy, V. G. and Meyerowitz, E. M. (2005). Stem-cell homeostasis and growth dynamics can be uncoupled in the Arabidopsis shoot apex. Science 310, 663–7.CrossRefGoogle ScholarPubMed
Reeves, G. T., Muratov, C. B., Schüpbach, T. and Shvartsman, S. Y. (2007). Quantitative models of developmental pattern formation. Dev. Cell 11, 289–300.CrossRefGoogle Scholar
Reinhardt, D., Pesce, E. -R., Stieger, P., et al. (2003). Regulation of phyllotaxis by polar auxin transport. Nature 426, 255–60.CrossRefGoogle ScholarPubMed
Reinitz, J. and Sharp, D. H. (1995). Mechanism of formation of eve stripes. Mech. Dev. 49, 133–58.CrossRefGoogle Scholar
Reinitz, J., Mjolsness, E. and Sharp, D. H. (1995). Model for cooperative control of positional information in Drosophila by bicoid and hunchback. J. Exp. Zool. 271, 47–56.CrossRefGoogle ScholarPubMed
Rennich, S. C. and Green, , P. B. (1997). The mathematics of plate bending. In The Dynamics of Cell and Tissue Motion, eds Alt, W., Deutsch, A. and Dunn, G., pp. 251–3. Basel: Birkhauser Verlag.CrossRefGoogle Scholar
Rida, P. C., Minh, N. and Jiang, Y. J. (2004). A Notch feeling of somite segmentation and beyond. Dev. Biol. 265, 2–22.CrossRefGoogle ScholarPubMed
Rolland-Lagan, A.-G., Bangham, J. A. and Coen, E. (2003). Growth dynamics underlying petal shape and asymmetry. Nature 422, 161–3.CrossRefGoogle ScholarPubMed
Rosenfeld, N., Young, J. W., Alon, U., Swain, P. S. and Elowitz, M. B. (2005). Gene regulation at the single-cell level. Science 307, 1962–5.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I. and Jernvall, J. (2004). How different types of pattern formation mechanisms affect the evolution of form and development. Evol. Dev. 6, 6–16.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I., Jernvall, J. and Newman, S. A. (2003). Mechanisms of pattern formation in development and evolution. Development 130, 2027–37.CrossRefGoogle ScholarPubMed
Sater, A. K. and Jacobson, A. G. (1990). The restriction of the heart morphogenetic field inXenopus laevis. Dev. Biol. 140, 328–36.CrossRefGoogle ScholarPubMed
Saunders, T. and Howard, M. (2009a). Morphogen profiles can be optimized to buffer against noise. Phys. Rev. E 80, 041902.CrossRefGoogle ScholarPubMed
Saunders, T. and Howard, M. (2009b). When it pays to rush: interpreting morphogen gradients prior to steady-state. Phys. Biol. 6, 046020.CrossRefGoogle ScholarPubMed
Serrano, N. and O'Farrell, P. H. (1997). Limb morphogenesis: connections between patterning and growth. Curr. Biol. 7, R186–R195.CrossRefGoogle Scholar
Shipman, P. D. and Newell, A. C. (2005). Polygonal planforms and phyllotaxis on plants. J. Theor. Biol. 236, 154–97.CrossRefGoogle ScholarPubMed
Sick, S., Reinker, S., Timmer, J. and Schlake, T. (2006). WNT and DKK determine hair follicle spacing through a reaction–diffusion mechanism. Science 314, 1447–50.CrossRefGoogle ScholarPubMed
Simon, R., Carpenter, R., Doyle, S., Coen, E. (1994). FIMBRIATA controls flower development by mediating between meristem and organ identity genes. Cell 78, 99–107.CrossRefGoogle ScholarPubMed
Slack, J. M. W. (1983). From Egg to Embryo. Cambridge: Cambridge University Press.Google Scholar
Smith, R. S., Guyomarc'h, S., Mandel, T., et al. (2006). A plausible model of phyllotaxis. Proc. Nat. Acad. Sci, USA 103, 1301–6.CrossRefGoogle ScholarPubMed
Smith, S. C. and Armstrong, J. B. (1990). Heart induction in wild-type and cardiac mutant axolotls (Ambystoma mexicanum). J. Exp. Zool. 254, 48–54.CrossRefGoogle Scholar
Smith, S. C. and Armstrong, J. B. (1991). Heart development in normal and cardiac-lethal mutant axolotls: a model for the control of vertebrate cardiogenesis. Differentiation 47, 129–34.CrossRefGoogle ScholarPubMed
Smith, S. C. and Armstrong, J. B. (1993). Reaction–diffusion control of heart development: evidence for activation and inhibition in precardiac mesoderm. Dev. Biol. 160, 535–42.CrossRefGoogle ScholarPubMed
Solms-Laubach, , zu, H. Graf. (1895). Monograph of the Acetabularieae. Trans. Linn. Soc. Lond. 2 (Botany) 5, 1–39.Google Scholar
Solursh, M., Jensen, K. L., Zanetti, N. C., Linsenmayer, T. F. and Reiter, R. S. (1984). Extracellular matrix mediates epithelial effects on chondrogenesis in vitro. Dev. Biol. 105, 451–7.CrossRefGoogle ScholarPubMed
Spirov, A. V. and Holloway, D. M. (2003). Making the body plan: precision in the genetic hierarchy of Drosophila embryo segmentation. In. Silico. Biol. 3, 89–100.Google ScholarPubMed
Spirov, A. V., Lopes, F. J. P. and Holloway, D. M. (2008). Molecular fluctuations and interpreting spatial gradients, applied to Hunchback pattern formation. Dev. Biol. 319, 589.CrossRefGoogle Scholar
Spirov, A. V., Fahmy, K., Schneider, M., et al. (2009). Formation of the bicoid morphogen gradient: an mRNA gradient dictates the protein gradient. Development 136, 605–14.CrossRefGoogle ScholarPubMed
Staehelin, A. and Giddings, T. H. (1982). Membrane-mediated control of cell wall microfibrillar order. In Developmental Order: Its Origin and Regulation, ed. Subtelny, S. and Green, P. B., pp. 133–47. New York: Alan R. Liss.Google Scholar
Steer, M. W. (1988). The role of calcium in exocytosis and endocytosis in plant cells. Physiologia Pl. 72, 213–20.CrossRefGoogle Scholar
Steinberg, M. S. (1970). Does differential adhesion govern self-assembly processes in histogenesis? Equilibrium configurations and the emergence of a hierarchy among populations of embryonic cells. J. Exp. Zool. 173, 395–434.CrossRefGoogle ScholarPubMed
Streit, A., Berliner, A. J., Papanayotou, C., Sirulnik, A. and Stern, C. D. (2000). Initiation of neural induction by FGF signalling before gastrulation. Nature 406, 74–8.CrossRefGoogle ScholarPubMed
Surkova, S., Kosman, D., Kozlov, K., et al. (2008). Characterization of the Drosophila segment determination morphome. Dev. Biol. 313, 844–62.CrossRefGoogle ScholarPubMed
Swindale, N. V. (1980). A model for the formation of ocular dominance stripes. Proc. R. Soc. Lond. B 208, 243–64.CrossRefGoogle ScholarPubMed
Tardent, P. (1978). Coelenterata, Cnidaria. In Morphogenese der Tiere, pp. 199–302. Stuttgart: Gustav Fischer Verlag.Google Scholar
Thaller, C. and Eichele, G. (1987). Identification and spatial distribution of retinoids in the developing chick limb bud. Nature 327, 625–8.CrossRefGoogle ScholarPubMed
Thompson, D. W. (1917). On Growth and Form, 1st edn (2nd edn, 1942; abridged edn, ed. Bonner, J. T., 1961). Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Tickle, C., Summerbell, D. and Wolpert, L. (1975). Positional signalling and specification of digits in chick limb morphogenesis. Nature 254, 199–202.CrossRefGoogle ScholarPubMed
Tippit, D. H. and Pickett-Heaps, J. D. (1974). Experimental investigations into morphogenesis in Micrasterias. Protoplasma 81, 271–96.CrossRefGoogle Scholar
Tkacik, G., Gregor, T. and Bialek, W. (2008). The role of input noise in transcriptional regulation. PLoS One 3, e2774.CrossRefGoogle ScholarPubMed
Tostevin, F., ten Wolde, P. R. and Howard, M. (2007). Fundamental limits to position determination by concentration gradients. PLoS Comp. Biol. 3, e78.CrossRefGoogle ScholarPubMed
Townes, P. S. and Holtfreter, J. (1955). Directed movements and selective adhesion of embryonic amphibian cells. J. Exp. Zool. 128, 53–120.CrossRefGoogle Scholar
Turing, A. M. (1952). The chemical basis of morphogenesis. Phil. Trans. R. Soc. Lond. B 237, 37–72.CrossRefGoogle Scholar
Tyson, J. J. and Kauffman, S. A. (1975). Control of mitosis by a continuous biochemical oscillation. J. Math. Biol. 1, 289–310.CrossRefGoogle Scholar
Tyson, J. J. and Light, J. C. (1973). Properties of two-component bimolecular and trimolecular chemical reaction systems. J. Chem. Phys. 59, 4164–73.CrossRefGoogle Scholar
Ueda, K. and Noguchi, T. (1988). Microfilament bundles of F-actin and cytomorphogenesis in the green alga Micrasterias crux-melitensis. Eur. J. Cell Biol. 46, 61–7.Google Scholar
Ueda, K. and Yoshioka, S. (1976). Cell wall development of Micrasterias americana, especially in isotonic and hypertonic solutions. J. Cell Sci. 21, 617–31.Google ScholarPubMed
Umulis, D. M., Serpe, M., O'Connor, M. B. and Othmer, H. G. (2006). Robust, bistable patterning of the dorsal surface of the Drosophila embryo. Proc. Nat. Acad. Sci. USA 103, 11613–18.CrossRefGoogle ScholarPubMed
Umulis, D. M., Shimmi, O., O'Connor, M. B. and Othmer, H. (2010). Organism-scale modeling of early Drosophila pattering via bone morphogenetic proteins. Dev. Cell 7, 1–15.Google Scholar
Virag, A. (1999). Discovering Genes Involved in Branching Decisions in Neurospora crassa. PhD thesis, University of British Columbia, Canada.Google Scholar
Virag, A. and Griffiths, A. J. F. (2004). A mutation in the Neurospora crassa actin gene results in multiple defects in tip growth and branching. Fungal genet. biol. 41, 213–25.CrossRefGoogle Scholar
Vöchting, H. (1877). Ueber Theilbarkeit im Pflanzenreich und die Wirkung innerer un äusserer Krafte auf Organbildung an Pflanzentheilen. Pflüger's Arch. 15, 153–90.CrossRefGoogle Scholar
Vöchting, H. (1878). Ueber Organbildung im Pflanzenreich, Vol. 1. Bonn, Germany: Max Cohen & Sohn.Google Scholar
Aderkas, P. (2002). In vitro phenotypic variation in larch cotyledon number. Int. J. Plant Sci. 163, 301–7.CrossRefGoogle Scholar
Waddington, C. H. (1956). Principles of Embryology. London: Allen & Unwin.Google Scholar
Wald, G. (1957). The origin of optical activity in nature. Ann. N. Y. Acad. Sci. 69, 352–67.CrossRefGoogle Scholar
Walker, D. C., ed. (1979). Origins of Optical Activity in Nature. Amsterdam: Elsevier.Google Scholar
Waller, M. D. (1961). Chladni Figures: A Study in Symmetry. London: G. Bell & Sons.Google Scholar
Wang, Y. C. and Ferguson, E. L. (2005). Spatial bistability of Dpp-receptor interactions during Drosophila dorsal–ventral patterning. Nature 434, 229–34.CrossRefGoogle ScholarPubMed
Wässle, H. and Riemann, H. J. (1978). The mosaic of nerve cells in the mammalian retina. Proc. R. Soc. Lond. B 200, 441–61.CrossRefGoogle ScholarPubMed
Weil, T. T., Forrest, K. M. and Gavis, E. R. (2006). Localization of bicoid mRNA in late oocytes is maintained by continual active transport. Dev. Cell 11, 251–62.CrossRefGoogle ScholarPubMed
Weil, T. T., Parton, R., Davis, I. and Gavis, E. R. (2008). Changes in bicoid mRNA anchoring highlight conserved mechanisms during the oocyte-to-embryo transition. Curr. Biol. 18, 1055–61.CrossRefGoogle ScholarPubMed
Werz, G. (1965). Determination and realization of morphogenesis in Acetabularia. Brookhaven Symp. Biol. 18, 185–203.Google Scholar
West, W. and West, G. S. (1905). A Monograph of the British Desmidiaceae, Vol. 2. London: Adlard & Son, for the Ray Society.Google Scholar
Wigglesworth, V. B. (1940). Local and general factors in the development of ‘pattern’ in Rhodnius prolixus (Hemiptera). J. Exp. Biol. 17, 180–200.Google Scholar
Wolff, C., Sommer, R., Schroeder, R., Glaser, G. and Tautz, D. (1995). Conserved and divergent expression aspects of the Drosophila segmentation gene hunchback in the short germ band embryo of the flour beetleTribolium. Development 121, 4227–36.Google ScholarPubMed
Wolpert, L. (1970). Positional information and pattern formation. In Towards a Theoretical Biology, ed. Waddington, C. H., Vol. 3, pp. 198–230. Edinburgh: Edinburgh University Press.Google Scholar
Wolpert, L. (1971). Positional information and pattern formation. In Current Topics in Developmental Biology, Vol. 6, pp. 183–224. New York: Academic Press.Google Scholar
Wolpert, L. (1975). Control processes in development: pattern formation in chick limb morphogenesis. Ann. Biomed. Eng. 3, 401–5.CrossRefGoogle ScholarPubMed
Wolpert, L. (1981). Positional information and pattern formation. Phil. Trans. R. Soc. Lond. B 295, 441–50.CrossRefGoogle ScholarPubMed
Wolpert, L. (1990). The embryonic position. The New Biologist 2, 1075–8.Google ScholarPubMed
Wolpert, L. (1992). Gastrulation and the evolution of development. Development 116, 7–13.Google Scholar
Wolpert, L. (2002). Principles of development. 2nd edn. London: Oxford University Press.Google Scholar
Wortis, M., Seifert, U., Berndl, K., et al. (1993). Curvature-controlled shapes of lipid-bilayer vesicles: budding, vesiculation and other phase transitions. In Dynamical Phenomena at Interfaces, Surfaces and Membranes, ed. Beysens, D., Boccara, N. and Forgacs, G., pp. 221–36. Commack, N Y: Nova Science Publishers.Google Scholar
Wu, Y. F., Myasnikova, E. and Reinitz, J. (2007). Master equation simulation analysis of immunostained Bicoid morphogen gradient. BMC Sys. Biol. 1, 52.CrossRefGoogle ScholarPubMed
Zhabotinskii, A. M. (1964). Periodical oxidation of malonic acid in solution (a study of the Belousov reaction kinetics). Biofizika 9, 306–11.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
  • Lionel G. Harrison, University of British Columbia, Vancouver
  • Book: The Shaping of Life
  • Online publication: 05 July 2011
  • Chapter DOI: https://doi.org/10.1017/CBO9780511973970.016
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
  • Lionel G. Harrison, University of British Columbia, Vancouver
  • Book: The Shaping of Life
  • Online publication: 05 July 2011
  • Chapter DOI: https://doi.org/10.1017/CBO9780511973970.016
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
  • Lionel G. Harrison, University of British Columbia, Vancouver
  • Book: The Shaping of Life
  • Online publication: 05 July 2011
  • Chapter DOI: https://doi.org/10.1017/CBO9780511973970.016
Available formats
×