Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-26T23:13:36.810Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  24 May 2010

Gabor Forgacs
Affiliation:
University of Missouri, Columbia
Stuart A. Newman
Affiliation:
New York Medical College
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
Publisher: Cambridge University Press
Print publication year: 2005

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

Adams, C. L., Nelson, W. J., and Smith, S. J. (1996). Quantitative analysis of cadherin–catenin–actin reorganization during development of cell–cell adhesion. J. Cell Biol. 135, 1899–911.CrossRefGoogle ScholarPubMed
Adams, C. L., Chen, Y. T., Smith, S. J., and Nelson, W. J. (1998). Mechanisms of epithelial cell–cell adhesion and cell compaction revealed by high-resolution tracking of E-cadherin-green fluorescent protein. J. Cell Biol. 142, 1105–19.CrossRefGoogle ScholarPubMed
Afzelius, B. A. (1985). The immotile-cilia syndrome: a microtubule-associated defect. CRC Crit. Rev. Biochem. 19, 63–87.CrossRefGoogle ScholarPubMed
Agius, E., Oegeschlager, M., Wessely, O., Kemp, C., and Roberts, E. M. (2000). Endodermal Nodal-related signals and mesoderm induction in Xenopus. Development 127, 1173–83.Google ScholarPubMed
Agutter, P. S., and Wheatley, D. N. (2000). Random walk and cell size. BioEssays 22, 1018–23.3.0.CO;2-Y>CrossRefGoogle Scholar
Akam, M. (1989). Making stripes inelegantly. Nature 341, 282–3.CrossRefGoogle ScholarPubMed
Akatiya, T., and Bronner-Fraser, M. (1992). Expression of cell adhesion molecules during initiation and cessation of neural crest cell migration. Dev. Dyn. 194, 12–20.CrossRefGoogle Scholar
Alber, M. S., Kiskowski, M. A., Glazier, J. A., and Jiang, Y. (2003). On cellular automaton approaches to modeling biological cells. In Mathematical Systems Theory in Biology, Communication, and Finance (Rosenthal, J. and Gilliam, D. S., eds.), Vol. 134, pp. 1–40. New York, Springer-Verlag.CrossRefGoogle Scholar
Alber, M., Hentschel, H. G. E., Kazmierczak, B., and Newman, S. A. (2005). Existence of solutions to a new model of biological pattern formation. Journal of Mathematical Analysis and Application 308, 175–194.CrossRefGoogle Scholar
Alberts, B., Johnson, A., Lewis, J., Raff, M., Roberts, K., and Walter, P. (2002). Molecular Biology of the Cell. Garland Science, New York.Google Scholar
Allaerts, W. (1991). On the role of gravity and positional information in embryological axis formation and tissue compartmentalization. Acta Biotheor. 39, 47–62.CrossRefGoogle ScholarPubMed
Amonlirdviman, K., Khare, N. A., Tree, D. R., Chen, W. S., Axelrod, J. D., and Tomlin, C. J. (2005).Mathematical modeling of planar cell polarity to understand domineering nonautonomy. Science 307, 423–6.CrossRefGoogle ScholarPubMed
Anderson, A. R., and Chaplain, M. A. (1998). Continuous and discrete mathematical models of tumor-induced angiogenesis. Bull. Math. Biol. 60, 857–99.CrossRefGoogle ScholarPubMed
Anderson, K. V., and Ingham, P. W. (2003). The transformation of the model organism: a decade of developmental genetics. Nat. Genet. 33 Suppl., 285–93.CrossRefGoogle ScholarPubMed
Angres, B., Barth, A., and Nelson, W. J. (1996). Mechanism for transition from initial to stable cell–cell adhesion: kinetic analysis of E-cadherin-mediated adhesion using a quantitative adhesion assay. J. Cell Biol. 134, 549–57.CrossRefGoogle ScholarPubMed
Armstrong, P. B. (1989). Cell sorting out: the self-assembly of tissues in vitro. Crit. Rev. Biochem. and Mol. Biol. 24, 119–49.CrossRefGoogle ScholarPubMed
Artavanis-Tsakonas, S., Rand, M. D., and Lake, R. J. (1999). Notch signaling: cell fate control and signal integration in development. Science 284, 770–6.CrossRefGoogle ScholarPubMed
Arthur, W. (1997). The Origin of Animal Body Plans: A Study in Evolutionary Developmental Biology. Cambridge, New York, Cambridge University Press.CrossRefGoogle Scholar
Atherton-Fessler, S., Hannig, G., and Piwnica-Worms, H. (1993). Reversible tyrosine phosphorylation and cell cycle control. Semin. Cell Biol. 4, 433–42.CrossRefGoogle ScholarPubMed
Augustin, H. G. (2001). Tubes, branches, and pillars: the many ways of forming a new vasculature. Circ. Res. 89, 645–7.Google ScholarPubMed
Aulehla, A., Wehrle, C., Brand-Saberi, B., Kemler, R., Gossler, A., Kanzler, B., and Herrmann, B. G. (2003). Wnt3a plays a major role in the segmentation clock controlling somitogenesis. Dev. Cell 4, 395–406.CrossRefGoogle Scholar
Aylsworth, A. S. (2001). Clinical aspects of defects in the determination of laterality. Am. J. Med. Genet. 101, 345–55.CrossRefGoogle ScholarPubMed
Bachvarova, R. (1985). Gene expression during oogenesis and oocyte development in mammals. Dev. Biol. 1, 453–524.Google ScholarPubMed
Ball, W. D. (1974). Development of the rat salivary glands. 3. Mesenchymal specificity in the morphogenesis of the embryonic submaxillary and sublingual glands of the rat. J. Exp. Zool. 188, 277–88.CrossRefGoogle ScholarPubMed
Ballaro, B., and Reas, P. G. (2000). Chemical and mechanical waves on the cortex of fertilized egg cells: a bioexcitability effect. Rev. Biol. 93, 83–101.Google ScholarPubMed
Baoal, D. (2002). Mechanics of the Cell. Cambridge University Press, Cambridge.Google Scholar
Barabási, A.-L. (2002). Linked: The New Science of Networks. Perseus Publications, Cambridge, MA.Google Scholar
Bard, J. B. (1999). A bioinformatics approach to investigating developmental pathways in the kidney and other tissues. Int. J. Dev. Biol. 43, 397–403.Google ScholarPubMed
Barkai, N., and Leibler, S. (2000). Circadian clocks limited by noise. Nature 403, 267–8.CrossRefGoogle ScholarPubMed
Basu, S., Gerchmann, Y., Collins, C. H., Arnold, F. H., Weiss, R. (2005). A synthetic multicellular system for programmed pattern formation. Nature 434, 1130–4.CrossRefGoogle ScholarPubMed
Bateman, E. (1998). Autoregulation of eukaryotic transcription factors. Prog. Nucleic Acid Res. Mol. Biol. 60, 133–68.CrossRefGoogle ScholarPubMed
Bateson, W. (1894). Materials for the Study of Variation. Macmillan. London.Google Scholar
Baumgartner, W., Hinterdorfer, P., Ness, W., et al. (2000). Cadherin interaction probed by atomic force microscopy. Proc. Nat. Acad. Sci. USA 97, 4005–10.CrossRefGoogle ScholarPubMed
Bausch, A. R., Moller, W., and Sackmann, E. (1999). Measurement of local viscoelasticity and forces in living cells by magnetic tweezers. Biophys. J. 76, 573–9.CrossRefGoogle ScholarPubMed
Becker, M., Baumann, C., John, S., et al. (2002). Dynamic behavior of transcription factors on a natural promoter in living cells. EMBO Rep. 3, 1188–94.CrossRefGoogle ScholarPubMed
Beddington, R. S., and Robertson, E. J. (1999). Axis development and early asymmetry in mammals. Cell 96, 195–209.CrossRefGoogle ScholarPubMed
Bejsovec, A., and Wieschaus, E. (1993). Segment polarity gene interactions modulate epidermal patterning in Drosophila embryos. Development 119, 501–17.Google ScholarPubMed
Bell, G. I. (1978). Models for the specific adhesion of cells to cells. Science 200, 618–27.CrossRefGoogle ScholarPubMed
Beloussov, L. (1998). The Dynamic Architecture of a Developing Organism.Kluwer Academic Publishers, Dordrecht.CrossRefGoogle Scholar
Ben-Avraham, D., and Havlin, S. (2000). Diffusion and Reactions in Fractals and Disordered Systems. Cambridge University Press, Cambridge, New York.CrossRefGoogle Scholar
Benink, H. A., Mandato, C. A., and Bement, W. M. (2000). Analysis of cortical flow models in vivo. Mol. Biol. Cell 11, 2553–63.CrossRefGoogle ScholarPubMed
Berg, H. C. (1993). Random Walks in Biology. Princeton University Press, Princeton.Google Scholar
Berridge, M. J., Lipp, P., and Bootman, M. D. (2000). The versatility and universality of calcium signalling. Nat. Rev. Mol. Cell Biol. 1, 11–21.CrossRefGoogle ScholarPubMed
Berry, L. D., and Gould, K. L. (1996). Regulation of Cdc2 activity by phosphorylation at T14/Y15. Prog. Cell. Cycle Res. 2, 99–105.CrossRefGoogle ScholarPubMed
Bertrand, N., Castro, D. S., and Guillemot, F. (2002). Proneural genes and the specification of neural cell types. Nat. Rev. Neurosci. 3, 517–30.CrossRefGoogle ScholarPubMed
Bevilacqua, M., Butcher, E., Furie, B., et al. (1991). Selectins: a family of adhesion receptors. Cell 67, 233.CrossRefGoogle ScholarPubMed
Beysens, D. A., Forgacs, G., and Glazier, J. A. (2000). Cell sorting is analogous to phase ordering in fluids. Proc. Nat. Acad. Sci. USA 97, 9467–71.CrossRefGoogle ScholarPubMed
Bhalla, U. S., and Iyengar, R. (1999). Emergent properties of networks of biological signaling pathways. Science 283, 381–7.CrossRefGoogle ScholarPubMed
Bissell, M. J., and Barcellos-Hoff, M. H. (1987). The influence of extracellular matrix on gene expression: is structure the message?J. Cell Sci. Suppl. 8, 327–43.CrossRefGoogle ScholarPubMed
Blair, S. S. (2003). Developmental biology: boundary lines. Nature 424, 379–81.CrossRefGoogle ScholarPubMed
Boal, D. H. (2002). Mechanics of the Cell. Cambridge University Press, Cambridge, New York.Google Scholar
Boggon, T. J., Murray, J., Chappuis-Flament, S., et al. (2002). C-cadherin ectodomain structure and implications for cell adhesion mechanisms. Science 296, 1308–13.CrossRefGoogle ScholarPubMed
Boissonade, J., Dulos, E., and DeKepper, P. (1994). Turing patterns: from myth to reality. In Chemical Waves and Patterns (Kapral, R. and Showalter, K., eds.), pp. 221–68. Kluwer, Boston.Google Scholar
Bolouri, H., and Davidson, E. H. (2003). Transcriptional regulatory cascades in development: initial rates, not steady state, determine network kinetics. Proc. Nat. Acad. Sci. USA 100, 9371–6.CrossRefGoogle Scholar
Bonner, J. T. (1998). The origins of multicellularity. Integrative Biology 1, 27–36.3.0.CO;2-6>CrossRefGoogle Scholar
Boring, L. (1989). Cell–cell interactions determine the dorsoventral axis in embryos of an equally cleaving opisthobranch mollusc. Dev. Biol. 136, 239–53.CrossRefGoogle ScholarPubMed
Borisuk, M. T., and Tyson, J. J. (1998). Bifurcation analysis of a model of mitotic control in frog eggs. J. Theor. Biol. 195, 69–85.CrossRefGoogle ScholarPubMed
Borkhvardt, V. G. (2000). The growth and form development of the limb buds in vertebrate animals. Ontogenez 31, 192–200.Google ScholarPubMed
Bouligand, Y. (1972). Twisted fibrous arrangements in biological materials and cholesteric mesophases. Tissue Cell 4, 189–217.CrossRefGoogle ScholarPubMed
Braat, A. K., Zandbergen, T., Water, S., Goos, H. J., and Zivkovic, D. (1999). Characterization of zebrafish primordial germ cells: morphology and early distribution of vasa RNA. Dev. Dyn. 216, 153–67.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Braga, V. M. (2002). Cell–cell adhesion and signalling. Curr. Opin. Cell. Biol. 14, 546–56.CrossRefGoogle ScholarPubMed
Branford, W. W., and Yost, H. J. (2002). Lefty-dependent inhibition of Nodal- and Wnt-responsive organizer gene expression is essential for normal gastrulation. Curr. Biol. 12, 2136–41.CrossRefGoogle ScholarPubMed
Brasier, M., and Antcliffe, J. (2004). Paleobiology. Decoding the Ediacaran enigma. Science 305, 1115–7.CrossRefGoogle ScholarPubMed
Breckenridge, R. A., Mohun, T. J., and Amaya, E. (2001). A role for BMP signalling in heart looping morphogenesis in Xenopus. Dev. Biol. 232, 191–203.CrossRefGoogle ScholarPubMed
Brinker, C. J., and Scherer, G. W. (1990). Sol–Gel Science. Academic Press, New York.Google Scholar
Brock, H. W., and Fisher, C. L. (2005). Maintenance of gene expression patterns. Dev. Dyn. 232, 633–55.CrossRefGoogle ScholarPubMed
Bronner-Fraser, M. (1982). Distribution of latex beads and retinal pigment epithelial cells along the ventral neural crest pathways. Dev. Biol. 91, 50–63.CrossRefGoogle Scholar
Bronner-Fraser, M. (1984). Latex beads as probes of a neural crest pathway: effects of laminin, collagen, and surface charge on bead translocation. J. Cell Biol. 98, 1947–60.CrossRefGoogle ScholarPubMed
Bronner-Fraser, M. (1985). Effects of different fragments of the fibronectin molecule on latex bead translocation along neural crest migratory pathways. Dev. Biol. 108, 131–45.CrossRefGoogle ScholarPubMed
Bronner-Fraser, M., Wolf, J. J., and Murray, B. A. (1992). Effects of antibodies against N-cadherin and N-CAM on the cranial neural crest and neural tube. Dev. Biol. 153, 291–301.CrossRef
Brown, S. J., Hilgenfeld, R. B., and Denell, R. E. (1994a). The beetle Tribolium castaneum has a fushi tarazu homolog expressed in stripes during segmentation. Proc. Nat. Acad. Sci. USA 91, 12922–6.CrossRefGoogle Scholar
Brown, S. J., Patel, N. H., and Denell, R. E. (1994b). Embryonic expression of the single Tribolium engrailed homolog. Dev. Genet. 15, 7–18.CrossRefGoogle Scholar
Browne, E. N. (1909). The production of new hydranths in Hydra by insertion of small grafts. J. Exp. Zool. 7, 1–23.CrossRefGoogle Scholar
Brunet, J. F., and Ghysen, A. (1999). Deconstructing cell determination: proneural genes and neuronal identity. Bioessays 21, 313–8.3.0.CO;2-C>CrossRefGoogle ScholarPubMed
Bryant, P. J. (1999). Filopodia: fickle fingers of cell fate?Curr. Biol. 9, R655–7.CrossRefGoogle ScholarPubMed
Buckley, C. D., Rainger, G. E., Bradfield, P. F., Nash, G. B., and Simmons, D. L. (1998). Cell adhesion: more than just glue. Mol. Membr. Biol. 15, 167–76.CrossRefGoogle ScholarPubMed
Bugrim, A., Fontanilla, R., Eutenier, B. B., Keizer, J., and Nuccitelli, R. (2003). Sperm initiate a Ca2+ wave in frog eggs that is more similar to Ca2+ waves initiated by IP3 than by Ca2+. Biophys. J. 84, 1580–90.CrossRefGoogle ScholarPubMed
Bugrim, A. E., Zhabotinsky, A. M., and Epstein, I. R. (1997). Calcium waves in a model with a random spatially discrete distribution of Ca2+ release sites. Biophys. J. 73, 2897–906.CrossRefGoogle Scholar
Callamaras, N., Marchant, J. S., Sun, X. P., and Parker, I. (1998). Activation and co-ordination of InsP3-mediated elementary Ca2+ events during global Ca2+ signals in Xenopus oocytes. J. Physiol. 509, 81–91.CrossRefGoogle ScholarPubMed
Campochiaro, P. A. (2000). Retinal and choroidal neovascularization. J. Cell Physiol. 184, 301–10.3.0.CO;2-H>CrossRefGoogle ScholarPubMed
Canman, J. C., and Bement, W. M. (1997). Microtubules suppress actomyosin-based cortical flow in Xenopus oocytes. J. Cell Sci. 110, 1907–17.Google ScholarPubMed
Capco, D. G., and McGaughey, R. W. (1986). Cytoskeletal reorganization during early mammalian development: analysis using embedment-free sections. Dev. Biol. 115, 446–58.CrossRefGoogle ScholarPubMed
Carnac, G., and Gurdon, J. B. (1997). The community effect in Xenopus myogenesis is promoted by dorsalizing factors. Int. J. Dev. Biol. 41, 521–4.Google ScholarPubMed
Castets, V., Dulos, E., Boissonade, J., and DeKepper, P. (1990). Experimental evidence of a sustained standing Turing-type nonequilibrium chemical pattern. Phys. Rev. Lett. 64, 2953–6.CrossRefGoogle ScholarPubMed
Cayan, S., Conaghan, J., Schriock, E. D., Ryan, I. P., Black, L. D., and Turek, P. J. (2001). Birth after intracytoplasmic sperm injection with use of testicular sperm from men with Kartagener/immotile cilia syndrome. Fertil. Steril. 76, 612–4.CrossRefGoogle ScholarPubMed
Chambon, F., and Winter, H. H. (1987). Linear viscoelasticity at the gel point of a crosslinking PDMS with imbalanced stoichiometry. J. Rheol. 31, 683–97.CrossRefGoogle Scholar
Chan, A. P., and Etkin, L. D. (2001). Patterning and lineage specification in the amphibian embryo. Curr. Top. Dev. Biol. 51, 1–67.CrossRefGoogle ScholarPubMed
Chaturvedi, R., Huang, C., Kazmierczak, B., Schneider, T., Izaguirre, J. A., Glimm, T., Hentschel, H. G. E., Newman, S. A., Glazier, J. A., and Alber, M. (2005). On multiscale approaches to three-dimensional modeling of morphogenesis. J. Roy. Soc. London Interface 2, 237–53.CrossRefGoogle Scholar
Cheer, A., Vincent, J. P., Nuccitelli, R., and Oster, G. (1987). Cortical activity in vertebrate eggs. I: The activation waves. J. Theor. Biol. 124, 377–404.CrossRefGoogle ScholarPubMed
Chen, J. N., Eeden, F. J., Warren, K. S., et al. (1997). Left–right pattern of cardiac BMP4 may drive asymmetry of the heart in zebrafish. Development 124, 4373–82.Google ScholarPubMed
Chen, J.-Y., Bottjer, D. J., Oliveri, P., et al. (2004). Small bilaterian fossils from 40 to 55 million years before the Cambrian. Science 305, 218–22.CrossRefGoogle ScholarPubMed
Chen, Y., and Schier, A. F. (2002). Lefty proteins are long-range inhibitors of squint-mediated nodal signaling. Curr. Biol. 12, 2124–8.CrossRefGoogle ScholarPubMed
Cheng, A., Ross, K. E., Kaldis, P., and Solomon, M. J. (1999). Dephosphorylation of cyclin-dependent kinases by type 2C protein phosphatases. Genes Dev. 13, 2946–57.CrossRefGoogle ScholarPubMed
Christen, B., and Slack, J. (1999). Spatial response to fibroblast growth factor signalling in Xenopus embryos. Development 126, 119–25.Google ScholarPubMed
Chuong, C. M. (1993). The making of a feather: homeoproteins, retinoids and adhesion molecules. BioEssays 15, 513–21.CrossRefGoogle ScholarPubMed
Cickovski, T., Huang, C., Chaturvedi, R., et al. (2005). A framework for three-dimensional simulation of morphogenesis. IEEE/ACM Trans. Computat. Biol. Bioinformatics, in press.CrossRefGoogle ScholarPubMed
Cinquin, O., and Demongeof, J. (2005). High-dimensional switches and the modelling of cellular differentiation. J. Theor. Biol. 233, 391–411.CrossRefGoogle ScholarPubMed
Clements, D., Friday, R. V., and Woodland, H. R. (1999). Mode of action of VegT in mesoderm and endoderm formation. Development 126, 4903–11.Google ScholarPubMed
Clerk, J. P., Giraud, G., Laugier, J. M., and Luck, J. M. (1990). The AC electrical conductance of binary disordered systems, percolation custers, fractals and related models. Adv. Phys. 39, 191–309.CrossRefGoogle Scholar
Cline, C. A., Schatten, H., Balczon, R., and Schatten, G. (1983). Actin-mediated surface motility during sea urchin fertilization. Cell Motil. 3, 513–24.CrossRefGoogle ScholarPubMed
Clyde, D. E., Corado, M. S., 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
Colas, J. F., and Schoenwolf, G. C. (2001). Towards a cellular and molecular understanding of neurulation. Dev. Dyn. 221, 117–45.CrossRefGoogle ScholarPubMed
Collier, J. R., Monk, N. A., Maini, P. K., and Lewis, J. H. (1996). Pattern formation by lateral inhibition with feedback: a mathematical model of Delta–Notch intercellular signalling. J. Theor. Biol. 183, 429–46.CrossRefGoogle ScholarPubMed
Comper, W. D. (1996). Extracellular Matrix. Vol. I. Tissue Function; Vol. II. Molecular Components and Interactions. Harwood Academic Publishers, Amsterdam.Google Scholar
Comper, W. D., Pratt, L., Handley, C. J., and Harper, G. S. (1987). Cell transport in model extracellular matrices. Arch. Biochem. Biophys. 252, 60–70.CrossRefGoogle ScholarPubMed
Conway Morris, S. (2003). The Cambrian “explosion” of metazoans. In Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology (Müller, G. B. and Newman, S. A., eds.), pp. 13–32. MIT Press, Cambridge, MA.Google Scholar
Cooke, J., Nowak, M. A., Boerlijst, M., and Maynard-Smith, J. (1997). Evolutionary origins and maintenance of redundant gene expression during metazoan development. Trends Genet. 13, 360–4.CrossRefGoogle ScholarPubMed
Cooke, J., and Zeeman, E. C. (1976). A clock and wavefront model for control of the number of repeated structures during animal morphogenesis. J. Theor. Biol. 58, 455–76.CrossRefGoogle ScholarPubMed
Cormack, D. H. (1987). Ham's Histology, ninth edn. Lippincott, Philadelphia.Google Scholar
Cornish-Bowden, A. (1995). Fundamentals of Enzyme Kinetics. Ashgate Publ. Co., London.Google Scholar
Coulombe, J. N., and Bronner-Fraser, M. (1984). Translocation of latex beads after laser ablation of the avian neural crest. Dev. Biol. 106, 121–34.CrossRefGoogle ScholarPubMed
Crawford, K., and Stocum, D. L. (1988). Retinoic acid coordinately proximalizes regenerate pattern and blastema differential affinity in axolotl limbs. Development 102, 687–98.Google ScholarPubMed
Crick, F. H. C. (1970). Diffusion in embryogenesis. Nature 225, 420–2.CrossRefGoogle ScholarPubMed
Crick, F. H. C., and Lawrence, P. A. (1975). Compartments and polyclones in insect development. Science 189, 340–7.CrossRefGoogle ScholarPubMed
Cross, N. L., and Elinson, R. P. (1980). A fast block to polyspermy in frogs mediated by changes in the membrane potential. Dev. Biol. 75, 187–98.CrossRef
Cunliffe, V. T. (2003). Memory by modification: the influence of chromatin structure on gene expression during vertebrate development. Gene 305, 141–50.CrossRefGoogle ScholarPubMed
Czirok, A., Rupp, P. A., Rongish, B. J., and Little, C. D. (2002). Multi-field 3D scanning light microscopy of early embryogenesis. J. Microsc. 206, 209–17.CrossRefGoogle ScholarPubMed
Dale, J. K., Maroto, M., Dequeant, M. L., Malapert, P., McGrew, M., and Pourquie, O. (2003). Periodic notch inhibition by Lunatic fringe underlies the chick segmentation clock. Nature 421, 275–8.CrossRefGoogle ScholarPubMed
Danos, M. C., and Yost, H. J. (1996). Role of notochord in specification of cardiac left–right orientation in zebrafish and Xenopus. Dev. Biol. 177, 96–103.CrossRefGoogle ScholarPubMed
Dathe, V., Gamel, A., Manner, J., Brand-Saberi, B., and Christ, B. (2002). Morphological left–right asymmetry of Hensen's node precedes the asymmetric expression of Shh and Fgf8 in the chick embryo. Anat. Embryol. (Berlin) 205, 343–54.CrossRefGoogle ScholarPubMed
Davidson, E. H. (2001). Genomic Regulatory Systems: Development and Evolution. Academic Press, San Diego.Google Scholar
Davidson, E. H., Rast, J. P., Oliveri, P., et al. (2002). A genomic regulatory network for development. Science 295, 1669–78.CrossRefGoogle Scholar
Davidson, L. A., Koehl, M. A., Keller, R., and Oster, G. F. (1995). How do sea urchins invaginate? Using biomechanics to distinguish between mechanisms of primary invagination. Development 121, 2005–18.Google ScholarPubMed
Davidson, L. A., Oster, G. F., Keller, R. E., and Koehl, M. A. (1999). Measurements of mechanical properties of the blastula wall reveal which hypothesized mechanisms of primary invagination are physically plausible in the sea urchin Strongylocentrotus purpuratus. Dev. Biol. 209, 221–38.CrossRefGoogle ScholarPubMed
Dawes, R., Dawson, I., Falciani, F., Tear, G., and Akam, M. (1994). Dax, a locust Hox gene related to fushi-tarazu but showing no pair-rule expression. Development 120, 1561–72.Google ScholarPubMed
Dawson, S. P., Keizer, J., and Pearson, J. E. (1999). Fire–diffuse–fire model of dynamics of intracellular calcium waves. Proc. Nat. Acad. Sci. USA 96, 6060–3.CrossRefGoogle ScholarPubMed
Felici, M. (2000). Regulation of primordial germ cell development in the mouse. Int. J. Dev. Biol. 44, 575–80.Google ScholarPubMed
Gennes, P. G. (1976a). Critical dimensionality for a special percolation problem (relevant to the gelation in polymers). J. Physique (Paris) 30, 1049–54.Google Scholar
Gennes, P. G. (1976b). On the relation between percolation and the elasticity of gels. J. Physique (Paris) 37, L1–2.CrossRefGoogle Scholar
Gennes, P. G. (1992). Soft matter. Science 256, 495–7.CrossRefGoogle ScholarPubMed
Gennes, P. G., and Prost, J. (1993). The Physics of Liquid Crystals. Clarendon Press, Oxford.Google Scholar
Deguchi, R., Shirakawa, H., Oda, S., Mohri, T., and Miyazaki, S. (2000). Spatiotemporal analysis of Ca2+ waves in relation to the sperm entry site and animal–vegetal axis during Ca2+ oscillations in fertilized mouse eggs. Dev. Biol. 218, 299–313.CrossRefGoogle Scholar
DeMarais, A. A., and Moon, R. T. (1992). The armadillo homologs beta-catenin and plakoglobin are differentially expressed during early development of Xenopus laevis. Dev. Biol. 153, 337–46.CrossRefGoogle ScholarPubMed
Smedt, V., Poulhe, R., Cayla, X., et al. (2002). Thr-161 phosphorylation of monomeric Cdc2. Regulation by protein phosphatase 2C in Xenopus oocytes. J. Biol. Chem. 277, 28592–600.CrossRefGoogle ScholarPubMed
Dembo, M., Glushko, V., Aberlin, M. E., and Sonenberg, M. (1979). A method for measuring membrane microviscosity using pyrene excimer formation. Application to human erythrocyte ghosts. Biochim. Biophys. Acta 552, 201–11.CrossRefGoogle ScholarPubMed
Derganc, J., Bozic, B., Svetina, S., and Zeks, B. (2000). Stability analysis of micropipette aspiration of neutrophils. Biophys. J. 79, 153–62.CrossRefGoogle ScholarPubMed
Dewar, H., Tanaka, K., Nasmyth, K., and Tanaka, T. U. (2004). Tension between two kinetochores suffices for their bi-orientation on the mitotic spindle. Nature 428, 93–7.CrossRefGoogle ScholarPubMed
Dickinson, R. B., and Tranquillo, R. T. (1993). A stochastic model for adhesion-mediated cell random motility and haptotaxis. J. Math. Biol. 31, 563–600.CrossRefGoogle ScholarPubMed
Djabourov, M., Leblond, J., and Papon, P. (1988). Gelation of aqueous gelatin solutions. II. Rheology of the sol–gel transition. J. Phys. (Paris) 49, 333–343.CrossRefGoogle Scholar
Djabourov, M., Lechaire, J. P., and Gaill, F. (1993). Structure and rheology of gelatin and collagen gels. Biorheology 30, 191–205.CrossRefGoogle ScholarPubMed
Doedel, E. J., and Wang, X. J. (1995). AUTO94: Software for continuation and bifurcation problems in ordinary differential equations. Center for Research on Parallel Computing, California Institute of Technology, Pasadena, CA.Google Scholar
Dolmetsch, R. E., Xu, K., and Lewis, R. S. (1998). Calcium oscillations increase the efficiency and specificity of gene expression. Nature 392, 933–6.CrossRefGoogle ScholarPubMed
Dosch, R., Gawantka, V., Delius, H., Blumenstock, C., and Niehrs, C. (1997). Bmp-4 acts as a morphogen in dorsoventral mesoderm patterning in Xenopus. Development 124, 2325–34.Google ScholarPubMed
Downie, S. A., and Newman, S. A. (1994). Morphogenetic differences between fore and hind limb precartilage mesenchyme: relation to mechanisms of skeletal pattern formation. Dev. Biol. 162, 195–208.CrossRefGoogle ScholarPubMed
Downie, S. A., and Newman, S. A. (1995). Different roles for fibronectin in the generation of fore and hind limb precartilage condensations. Dev. Biol. 172, 519–30.CrossRefGoogle ScholarPubMed
Drasdo, D., and Forgacs, G. (2000). Modeling the interplay of generic and genetic mechanisms in cleavage, blastulation, and gastrulation. Dev. Dyn. 219, 182–91.3.3.CO;2-1>CrossRefGoogle ScholarPubMed
Drasdo, D., Kree, R., and McCaskill, J. S. (1995). Monte Carlo approach to tissue-cell populations. Phys. Rev. E Statistical Physics, Plasmas, Fluids, and Related Interdisciplinary Topics 52, 6635–57.Google ScholarPubMed
Duband, J. L., Monier, F., Delannet, M., and Newgreen, D. (1995). Epithelium–mesenchyme transition during neural crest development. Acta Anat. (Basel) 154, 63–78.CrossRefGoogle ScholarPubMed
Dubrulle, J., McGrew, M. J., and Pourquié, O. (2001). FGF signaling controls somite boundary position and regulates segmentation clock control of spatiotemporal Hox gene activation. Cell 106, 219–32.CrossRefGoogle ScholarPubMed
Ducibella, T., Huneau, D., Angelichio, E., Xu, Z., Schultz, R. M., Kopf, G. S., Fissore, R., Madoux, S., and Ozil, J. P. (2002). Egg-to-embryo transition is driven by differential responses to Ca2+ oscillation number. Dev. Biol. 250, 280–91.CrossRefGoogle Scholar
Duguay, D., Foty, R. A., and Steinberg, M. S. (2003). Cadherin-mediated cell adhesion and tissue segregation: qualitative and quantitative determinants. Dev. Biol. 253, 309–23.CrossRefGoogle ScholarPubMed
Dumollard, R., Carroll, J., Dupont, G., and Sardet, C. (2002). Calcium wave pacemakers in eggs. J. Cell Sci. 115, 3557–64.CrossRefGoogle ScholarPubMed
Dumont, J. N., and Brummett, A. R. (1985). Egg envelopes in vertebrates. Dev. Biol. 1, 235–88.Google ScholarPubMed
Durand, D., Delsanti, M., Adam, M., and Luck, J. M. (1987). Frequency dependence of viscoelastic properties of branched polymers near gelation threshold. Europhys. Lett. 3, 297–301.CrossRefGoogle Scholar
Edelman, G. M. (1992). Morphoregulation. Dev. Dyn. 193, 2–10.CrossRefGoogle ScholarPubMed
Eichmann, A., Pardanaud, L., Yuan, L., and Moyon, D. (2002). Vasculogenesis and the search for the hemangioblast. J. Hematother. Stem Cell Res. 11, 207–14.CrossRefGoogle ScholarPubMed
Eidne, K. A., Zabavnik, J., Allan, W. T., Trewavas, A. J., Read, N. D., and Anderson, L. (1994). Calcium waves and dynamics visualized by confocal microscopy in Xenopus oocytes expressing cloned TRH receptors. J. Neuroendocrinol. 6, 173–8.CrossRefGoogle ScholarPubMed
Ekblom, P. (1992). Renal Development. Raven Press, New York.Google Scholar
Elinson, R. P., and Rowning, B. (1988). A transient array of parallel microtubules in frog eggs: potential tracks for a cytoplasmic rotation that specifies the dorso-ventral axis. Dev. Biol. 128, 185–97.CrossRefGoogle ScholarPubMed
Ellis, R. J. (2001). Macromolecular crowding: an important but neglected aspect of the intracellular environment. Curr. Opin. Struct. Biol. 11, 114–9.CrossRefGoogle ScholarPubMed
Ellis, R. J., and Minton, A. P. (2003). Cell biology: join the crowd. Nature 425, 27–8.CrossRefGoogle Scholar
Elowitz, M. B., and Leibler, S. (2000). A synthetic oscillatory network of transcriptional regulators. Nature 403, 335–8.CrossRefGoogle ScholarPubMed
Entchev, E. V., Schwabedissen, A., and Gonzalez-Gaitan, M. (2000). Gradient formation of the TGF-β homolog Dpp. Cell 103, 981–91.CrossRefGoogle ScholarPubMed
Erickson, C. A. (1985). Control of neural crest cell dispersion in the trunk of the avian embryo. Dev. Biol. 111, 138–57.CrossRefGoogle ScholarPubMed
Erickson, C. A. (1988). Control of pathfinding by the avian trunk neural crest. Development 103, 63–80.Google ScholarPubMed
Erickson, C. A., and Isseroff, R. R. (1989). Plasminogen activator activity is associated with neural crest cell motility in tissue culture. J. Exp. Zool. 251, 123–33.CrossRefGoogle ScholarPubMed
Erickson, C. A., and Perris, R. (1993). The role of cell–cell and cell–matrix interactions in the morphogenesis of the neural crest. Dev. Biol. 159, 60–74.CrossRefGoogle ScholarPubMed
Eshkind, L., Tian, Q., Schmidt, A., et al. (2002). Loss of desmoglein 2 suggests essential functions for early embryonic development and proliferation of embryonal stem cells of mice and models: improved animal models for biomedical research. Synaptic vesicle alterations in rod photoreceptors of synaptophysin-deficient mice. Eur. J. Cell. Biol. 81, 592–8.CrossRefGoogle Scholar
Espeseth, A., Johnson, E., and Kintner, C. (1995). Xenopus F-cadherin, a novel member of the cadherin family of cell adhesion molecules, is expressed at boundaries in the neural tube. Mol. Cell. Neurosci. 6, 199–211.CrossRefGoogle ScholarPubMed
Essner, J. J., Vogan, K. J., Wagner, M. K., Tabin, C. J., Yost, H. J., and Brueckner, M. (2002). Conserved function for embryonic nodal cilia. Nature 418, 37–8.CrossRefGoogle ScholarPubMed
Ettensohn, C. A. (1999). Cell movements in the sea urchin embryo. Curr. Opin. Genet. Dev. 9, 461–5.CrossRefGoogle ScholarPubMed
Ettinger, L., and Doljanski, F. (1992). On the generation of form by the continuous interactions between cells and their extracellular matrix. Biol. Rev. Camb. Philos. Soc. 67, 459–89.CrossRefGoogle ScholarPubMed
Evans, E., and Yeoung, A. (1989). Apparent viscosity and cortical tension of blood granulocytes determined by micropipet aspiration. Biophys. J. 56, 151–60.CrossRefGoogle ScholarPubMed
Fagotto, F., Guger, K., and Gumbiner, B. M. (1997). Induction of the primary dorsalizing center in Xenopus by the Wnt/GSK/beta-catenin signaling pathway, but not by Vg1, Activin or Noggin. Development 124, 453–60.Google ScholarPubMed
Ferrell, J. E. Jr, Wu, M., Gerhart, J. C., and Martin, G. S. (1991). Cell cycle tyrosine phosphorylation of p34cdc2 and a microtubule-associated protein kinase homolog in Xenopus oocytes and eggs. Mol. Cell. Biol. 11, 1965–71.CrossRefGoogle Scholar
Fitch, J., Fini, M. E., Beebe, D. C., and Linsenmayer, T. F. (1998). Collagen type IX and developmentally regulated swelling of the avian primary corneal stroma. Dev. Dyn. 212, 27–37.3.0.CO;2-4>CrossRefGoogle ScholarPubMed
Fleming, T. P., and Goodall, H. (1986). Endocytic traffic in trophectoderm and polarised blastomeres of the mouse preimplantation embryo. Anat. Rec. 216, 490–503.CrossRefGoogle ScholarPubMed
Folkman, J. (2003). Angiogenesis and proteins of the hemostatic system. J. Thromb. Haemost. 1, 1681–2.CrossRefGoogle ScholarPubMed
Folkman, J., and Moscona, A. (1978). Role of cell shape in growth control. Nature 273, 345–9.CrossRefGoogle ScholarPubMed
Fontanilla, R. A., and Nuccitelli, R. (1998). Characterization of the sperm-induced calcium wave in Xenopus eggs using confocal microscopy. Biophys. J. 75, 2079–87.CrossRefGoogle ScholarPubMed
Forgacs, G. (1995). On the possible role of cytoskeletal filamentous networks in intracellular signalling: an approach based on percolation. J. Cell Sci. 108, 2131–2143.Google Scholar
Forgacs, G., and Foty, R. A. (2004). Biological implications of tissue viscoelasticity. In Function and Regulation of Cellular Systems: Experiments and Models (Deutsch, A., Falke, M., Howard, J., and Zimmerman, W., eds.), pp. 269–77. Biskhauser Basel.CrossRefGoogle Scholar
Forgacs, G., and Newman, S. A. (1994). Phase transitions, interfaces, and morphogenesis in a network of protein fibers. Int. Rev. Cytol. 150, 139–48.CrossRefGoogle Scholar
Forgacs, G., Jaikaria, N. S., Frisch, H. L., and Newman, S. A. (1989). Wetting, percolation and morphogenesis in a model tissue system. J. Theor. Biol. 140, 417–430.CrossRefGoogle Scholar
Forgacs, G., Newman, S. A., Obukhov, S. P., and Birk, D. E. (1991). Phase transition and morphogenesis in a model biological system. Phys. Rev. Lett. 67, 2399–402.CrossRefGoogle Scholar
Forgacs, G., Foty, R. A., Shafrir, Y., and Steinberg, M. S. (1998). Viscoelastic properties of living embryonic tissues: a quantitative study. Biophys. J. 74, 2227–34.CrossRefGoogle ScholarPubMed
Forgacs, G., Newman, S. A., Hinner, B., Maier, C. W., and Sackmann, E. (2003). Assembly of collagen matrices as a phase transition revealed by structural and rheologic studies. Biophys. J. 84, 1272–80.CrossRefGoogle ScholarPubMed
Forgacs, G., Yook, S. H., Janmey, P. A., Jeong, H., and Burd, C. G. (2004). Role of the cytoskeleton in signaling networks. J. Cell. Sci. 117, 2769–75.CrossRefGoogle ScholarPubMed
Foty, R. A., and Steinberg, M. S. (1997). Measurement of tumor cell cohesion and suppression of invasion by E- or P-cadherin. Cancer Res. 57, 5033–6.Google ScholarPubMed
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., Forgacs, G., Pfleger, C. M., and Steinberg, M. S. (1994). Liquid properties of embryonic tissues: measurement of interfacial tensions. Phys. Rev. Lett. 72, 2298–301.CrossRefGoogle ScholarPubMed
Foty, R. A., Pfleger, C. M., Forgacs, G., and Steinberg, M. S. (1996). Surface tensions of embryonic tissues predict their mutual envelopment behavior. Development 122, 1611–20.Google ScholarPubMed
Frasch, M., and Levine, M. (1987). Complementary patterns of even-skipped and fushi tarazu expression involve their differential regulation by a common set of segmentation genes in Drosophila. Genes Dev. 1, 981–95.CrossRefGoogle ScholarPubMed
Freeman, M. (2002). Morphogen gradients, in theory. Dev. Cell 2, 689–90.CrossRefGoogle ScholarPubMed
Frenz, D. A., Akiyama, S. K., Paulsen, D. F., and Newman, S. A. (1989a). Latex beads as probes of cell surface–extracellular matrix interactions during chondrogenesis: evidence for a role for amino-terminal heparin-binding domain of fibronectin. Dev. Biol. 136, 87–96.CrossRefGoogle Scholar
Frenz, D. A., Jaikaria, N. S., and Newman, S. A. (1989b). The mechanism of precartilage mesenchymal condensation: a major role for interaction of the cell surface with the amino-terminal heparin-binding domain of fibronectin. Dev. Biol. 136, 97–103.CrossRefGoogle Scholar
Freyman, T. M., Yannas, I. V., Yokoo, R., and Gibson, L. J. (2001). Fibroblast contraction of a collagen-GAG matrix. Biomaterials 22, 2883–91.CrossRefGoogle ScholarPubMed
Friedlander, D. R., Mege, R.-M., Cunningham, B. A., and Edelman, G. M. (1989). Cell sorting-out is modulated by both the specificity and amount of different cell adhesion molecules (CAMs) expressed on cell surfaces. Proc. Nat. Acad. Sci. USA 86, 7043–7047.CrossRefGoogle ScholarPubMed
Fristrom, D., and Chihara, C. (1978). The mechanism of evagination of imaginal discs of Drosophila melanogaster. V. Evagination of disc fragments. Dev. Biol. 66, 564–570.CrossRefGoogle ScholarPubMed
Fujiwara, T., Ritchie, K., Murakoshi, H., Jacobson, K., and Kusumi, A. (2002). Phospholipids undergo hop diffusion in compartmentalized cell membrane. J. Cell Biol. 157, 1071–81.CrossRefGoogle ScholarPubMed
Fung, Y. C. (1993). Biomechanics: Mechanical Properties of Living Tissues. Springer-Verlag, New York.CrossRefGoogle Scholar
Furusawa, C., and Kaneko, K. (1998). Emergence of rules in cell society: differentiation, hierarchy, and stability. Bull. Math. Biol. 60, 659–87.CrossRefGoogle ScholarPubMed
Furusawa, C., and Kaneko, K. (2001). Theory of robustness of irreversible differentiation in a stem cell system: chaos hypothesis. J. Theor. Biol. 209, 395–416.CrossRefGoogle Scholar
Gaill, F., Lechaire, J. P., and Denefle, J. P. (1991). Fibrillar pattern of self-assembled and cell-assembled collagen: resemblance and analogy. Biol. Cell 72, 149–58.CrossRefGoogle ScholarPubMed
Gamba, A., Ambrosi, D., Coniglio, A., et al. (2003). Percolation, morphogenesis, and Burgers dynamics in blood vessels formation. Phys. Rev. Lett. 90, 118 101.CrossRefGoogle ScholarPubMed
Garcia-Bellido, A. (1975). Genetic control of wing disc development in Drosophila. Ciba Found. Sym. 29, 169–78.Google Scholar
Garcia-Bellido, A., Ripoll, P., and Morata, G. (1976). Developmental compartmentalization in the dorsal mesothoracic disc of Drosophila. Dev. Biol. 48, 132–47.CrossRefGoogle ScholarPubMed
Garcia-Perez, A. I., Lopez-Beltran, E. A., Kluner, P., Luque, J., Ballesteros, P., and Cerdan, S. (1999). Molecular crowding and viscosity as determinants of translational diffusion of metabolites in subcellular organelles. Arch. Biochem. Biophys. 362, 329–38.CrossRefGoogle ScholarPubMed
Gardner, R. L. (2001). The initial phase of embryonic patterning in mammals. Int. Rev. Cytol. 203, 233–90.CrossRefGoogle ScholarPubMed
Gerhart, J. (2002). Changing the axis changes the perspective. Dev. Dyn. 225, 380–3.CrossRefGoogle Scholar
Gerhart, J., Ubbels, G., Black, S., Hara, K., and Kirschner, M. (1981). A reinvestigation of the role of the grey crescent in axis formation in Xenopus laevis. Nature 292, 511–6.CrossRefGoogle ScholarPubMed
Ghosh, S., and Comper, W. D. (1988). Oriented fibrillogenesis of collagen in vitro by ordered convection. Connect. Tissue. Res. 17, 33–41.CrossRefGoogle ScholarPubMed
Giancotti, F. G., and Ruoslahti, E. (1999). Integrin signaling. Science 285, 1028–32.CrossRefGoogle ScholarPubMed
Giansanti, M. G., Bonaccorsi, S., Bucciarelli, E., and Gatti, M. (2001). Drosophila male meiosis as a model system for the study of cytokinesis in animal cells. Cell Struct. Funct. 26, 609–17.CrossRefGoogle Scholar
Gierer, A. (1977). Physical aspects of tissue evagination and biological form. Quart. Rev. Biophys. 10, 529–93.CrossRefGoogle ScholarPubMed
Gilbert, S. F. (2003). Developmental Biology. Sinauer Associates, Sunderland, MA.Google ScholarPubMed
Gilkey, J. C., Jaffe, L. F., Ridgway, E. B., and Reynolds, G. T. (1978). A free calcium wave traverses the activating egg of the medaka, Oryzias latipes. J. Cell Biol. 76, 448–66.CrossRefGoogle ScholarPubMed
Gimlich, R. L. (1985). Cytoplasmic localization and chordamesoderm induction in the frog embryo. J. Embryol. Exp. Morphol. 89 Suppl., 89–111.Google ScholarPubMed
Gimlich, R. L. (1986). Acquisition of developmental autonomy in the equatorial region of the Xenopus embryo. Dev. Biol. 115, 340–52.CrossRefGoogle ScholarPubMed
Ginsburg, M., Snow, M. H., and McLaren, A. (1990). Primordial germ cells in the mouse embryo during gastrulation. Development 110, 521–8.Google ScholarPubMed
Giraud-Guille, M. M. (1996). Twisted liquid crystalline supramolecular arrangements in morphogenesis. Int. Rev. Cytol. 166, 59–101.CrossRefGoogle ScholarPubMed
Glahn, D., and Nuccitelli, R. (2003). Voltage-clamp study of the activation currents and fast block to polyspermy in the egg of Xenopus laevis. Dev. Growth Differ. 45, 187–97.CrossRefGoogle ScholarPubMed
Glazier, J. A., and Graner, F. (1993). A simulation of the differential adhesion driven rearrangement of biological cells. Phys. Rev. E 47, 2128–54.CrossRefGoogle ScholarPubMed
Godt, D., and Tepass, U. (1998). Drosophila oocyte localization is mediated by differential cadherin-based adhesion. Nature 395, 387–91.CrossRefGoogle ScholarPubMed
Goldbeter, A. (1996). Biochemical Oscillations and Cellular Rhythms: the Molecular Bases of Periodic and Chaotic Behaviour. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Gomperts, M., Wylie, C., and Heasman, J. (1994). Primordial germ cell migration. Ciba Found. Symp. 182, 121–34; discussion 134–9.Google ScholarPubMed
Gong, Y., Mo, C., and Fraser, S. E. (2004). Planar cell polarity signalling controls cell division orientation during zebrafish gastrulation. Nature 430, 689–93.CrossRefGoogle ScholarPubMed
Gonzalez-Reyes, A., and St Johnston, D. (1998). The Drosophila AP axis is polarised by the cadherin-mediated positioning of the oocyte. Development 125, 3635–44.Google ScholarPubMed
Gonze, D., and Goldbeter, A. (2001). A model for a network of phosphorylation–dephosphorylation cycles displaying the dynamics of dominoes and clocks. J. Theor. Biol. 210, 167–86.CrossRefGoogle ScholarPubMed
Gosden, R., Krapez, J., and Briggs, D. (1997). Growth and development of the mammalian oocyte. Bioessays 19, 875–82.CrossRefGoogle ScholarPubMed
Gossler, A., and Angelis, Hrabe M. (1998). Somitogenesis. Curr. Top. Dev. Biol. 38, 225–87.CrossRefGoogle ScholarPubMed
Gould, S. E., Upholt, W. B., and Kosher, R. A. (1992). Syndecan 3: a member of the syndecan family of membrane-intercalated proteoglycans that is expressed in high amounts at the onset of chicken limb cartilage differentiation. Proc. Nat. Acad. Sci. USA 89, 3271–75.CrossRefGoogle ScholarPubMed
Gould, S. J. (1977). Ontogeny and Phylogeny. Harvard University Press, Cambridge, MA.Google Scholar
Goulian, M., and Simon, S. M. (2000). Tracking single proteins within cells. Biophys. J. 79, 2188–98.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
Green, J. (2002). Morphogen gradients, positional information, and Xenopus: interplay of theory and experiment. Dev. Dyn. 225, 392–408.CrossRefGoogle ScholarPubMed
Greenwald, I. (1998). LIN-12/Notch signaling: lessons from worms and flies. Genes Dev. 12, 1751–62.CrossRefGoogle ScholarPubMed
Greenwald, I., and Rubin, G. M. (1992). Making a difference: the role of cell–cell interactions in establishing separate identities for equivalent cells. Cell 68, 271–81.CrossRefGoogle ScholarPubMed
Gregory, P. D., Wagner, K., and Horz, W. (2001). Histone acetylation and chromatin remodeling. Exp. Cell Res. 265, 195–202.CrossRefGoogle ScholarPubMed
Gritsman, K., Talbot, W. S., and Schier, A. F. (2000). Nodal signaling patterns the organizer. Development 127, 921–32.Google ScholarPubMed
Gullberg, D. E., and Lundgren-Akerlund, E. (2002). Collagen-binding I domain integrins – what do they do?Prog. Histochem. Cytochem. 37, 3–54.CrossRefGoogle Scholar
Gumbiner, B. M. (1996). Cell adhesion: the molecular basis of tissue architecture and morphogenesis. Cell 84, 345–57.CrossRefGoogle ScholarPubMed
Gumbiner, B. M. (2000). Regulation of cadherin adhesive activity. J. Cell. Biol. 148, 399–404.CrossRefGoogle ScholarPubMed
Gurdon, J. B. (1988). A community effect in animal development. Nature 336, 772–4.CrossRefGoogle ScholarPubMed
Guthrie, S., and Lumsden, A. (1991). Formation and regeneration of rhombomere boundaries in the developing chick hindbrain. Development 112, 221–9.Google ScholarPubMed
Haddon, C., Smithers, L., Schneider-Maunoury, S., Coche, T., Henrique, D., and Lewis, J. (1998). Multiple delta genes and lateral inhibition in zebrafish primary neurogenesis. Development 125, 359–70.Google ScholarPubMed
Hafner, M., Petzelt, C., Nobiling, R., Pawley, J. B., Kramp, D., and Schatten, G. (1988). Wave of free calcium at fertilization in the sea urchin egg visualized with fura-2. Cell Motil. Cytoskeleton 9, 271–7.CrossRefGoogle ScholarPubMed
Haga, H., Nagayama, M., Kawabata, K., Ito, E., Ushiki, T., and Sambongi, T. (2000). Time-lapse viscoelastic imaging of living fibroblasts using force modulation mode in AFM. J. Electron Microsc. (Tokyo) 49, 473–81.CrossRefGoogle ScholarPubMed
Hahn, H. S., Ortoleva, P. J., and Ross, J. (1973). Chemical oscillations and multiple steady states due to variable boundary permeability. J. Theor. Biol. 41, 503–21.CrossRefGoogle ScholarPubMed
Hall, B. K., and Miyake, T. (1995). Divide, accumulate, differentiate: cell condensation in skeletal development revisited. Int. J. Dev. Biol. 39, 881–93.Google ScholarPubMed
Hall, B. K., and Miyake, T. (2000). All for one and one for all: condensations and the initiation of skeletal development. Bioessays 22, 138–47.3.0.CO;2-4>CrossRefGoogle ScholarPubMed
Hall, D., and Minton, A. P. (2003). Macromolecular crowding: qualitative and semiquantitative successes, quantitative challenges. Biochim. Biophys. Acta 1649, 127–39.CrossRefGoogle ScholarPubMed
Hardin, J. (1996). The cellular basis of sea urchin gastrulation. Curr. Top. Dev. Biol. 33, 159–262.CrossRefGoogle ScholarPubMed
Harding, K., Hoey, T., Warrior, R., and Levine, M. (1989). Autoregulatory and gap gene response elements of the even-skipped promoter of Drosophila. EMBO J. 8, 1205–12.Google ScholarPubMed
Hardman, P., and Spooner, B. S. (1992). Salivary epithelium branching morphogenesis. In Epithelial Organization and Development. (Fleming, T. P., ed.), pp. 353–75. Chapman and Hall, London.CrossRefGoogle Scholar
Harland, R., and Gerhart, J. (1997). Formation and function of Spemann's organizer. Ann. Rev. Cell Dev. Biol. 13, 611–67.CrossRefGoogle ScholarPubMed
Harper, G. S., Comper, W. D., and Preston, B. N. (1984). Dissipative structures in proteoglycan solutions. J. Biol. Chem. 259, 10582–9.Google ScholarPubMed
Harrison, L. J. (1993). Kinetic Theory of Living Form. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Harrisson, F. (1989). The extracellular matrix and cell surface, mediators of cell interactions in chicken gastrulation. Int. J. Dev. Biol. 33, 417–38.Google ScholarPubMed
Hayashi, T., and Carthew, R. W. (2004). Surface mechanics mediate pattern formation in the developing retina. Nature 431, 647–52.CrossRefGoogle ScholarPubMed
He, X., and Dembo, M. (1997). On the mechanics of the first cleavage division of the sea urchin egg. Exp. Cell Res. 233, 252–73.CrossRefGoogle ScholarPubMed
Heintzelman, K. F., Phillips, H. M., and Davis, G. S. (1978). Liquid-tissue behavior and differential cohesiveness during chick limb budding. J. Embryol. Exp. Morphol. 47, 1–15.Google ScholarPubMed
Helfrich, W. (1973). Elastic properties of lipid bilayers: theory and possible experiments. Z. Naturforsch. C 28, 693–703.CrossRefGoogle ScholarPubMed
Hentschel, H. G., Glimm, T., Glazier, J. A., and Newman, S. A. (2004). Dynamical mechanisms for skeletal pattern formation in the vertebrate limb. Proc. R. Soc. Lond. B Biol. Sci. 271, 1713–22.CrossRefGoogle ScholarPubMed
Heyman, I., Faissner, A., and Lumsden, A. (1995). Cell and matrix specialisations of rhombomere boundaries. Dev. Dyn. 204, 301–15.CrossRefGoogle ScholarPubMed
Hieda, Y., and Nakanishi, Y. (1997). Epithelial morphogenesis in mouse embryonic submandibular gland: its relationships to the tissue organization of epithelium and mesenchyme. Dev. Growth Differ. 39, 1–8.CrossRefGoogle ScholarPubMed
Hiramoto, Y. (1956). Cell division without mitotic apparatus in sea urchin eggs. Exp. Cell Res. 11, 630–636.CrossRefGoogle ScholarPubMed
Hiramoto, Y. (1958). A quantitative description of protoplasmic movement during cleavage in the sea-urchin egg. J. Exp. Biol. 35, 407–424.Google Scholar
Hiramoto, Y. (1963). Mechanical properties of sea urchin eggs. I. Surface force and elastic modulus of the cell membrane. Exp. Cell Res. 32, 59–75.CrossRefGoogle ScholarPubMed
Hiramoto, Y. (1968). The mechanics and mechanism of cleavage in the sea-urchin egg. Symp. Soc. Exp. Biol. 22, 311–27.Google ScholarPubMed
Hiramoto, Y. (1978). Mechanical properties of the dividing sea urchin egg. In Cell Motility: Molecules and Organization (Hatano, S., Ishikawa, H., and Sato, H., eds.), pp. 653–63. University Park Press, Baltimore.Google Scholar
Hobbie, R. K. (1997). Intermediate Physics for Medicine and Biology. Springer-Verlag, New York.Google Scholar
Hofmeyr, J. H., and Cornish-Bowden, A. (1997). The reversible Hill equation: how to incorporate cooperative enzymes into metabolic models. Comput. Appl. Biosci. 13, 377–85.Google ScholarPubMed
Holley, S. A., Geisler, R., and Nusslein-Volhard, C. (2000). Control of her1 expression during zebrafish somitogenesis by a Delta-dependent oscillator and an independent wave-front activity. Genes Dev. 14, 1678–90.Google ScholarPubMed
Holley, S. A., Julich, D., Rauch, G. J., Geisler, R., and Nusslein-Volhard, C. (2002). her1 and the notch pathway function within the oscillator mechanism that regulates zebrafish somitogenesis. Development 129, 1175–83.Google ScholarPubMed
Holtzendorff, J., Hung, D., Brende, P., et al. (2004). Oscillating global regulator control the genetic circuit driving a bacterial cell cycle. Science 304, 983–7.CrossRefGoogle ScholarPubMed
Hörstadius, S., and Sellman, S. (1946). Experimentelle Untersuchungen uber die Determination des knorpeligen Kopfskelettes bei Urodelen. Nova Acta R. Soc. Scient. Upsal. Ser. 4 13, 1–170.Google Scholar
Hoshi, M. (1979). Exogastrulation induced by heavy water in sea urchin larvae. Cell Differ. 8, 431–5.CrossRefGoogle ScholarPubMed
Howard, J. (2001). Mechanics of Motor Proteins and the Cytoskeleton. Sinauer Associates, Inc., Sunderland.Google Scholar
Howard, K., and Ingham, P. (1986). Regulatory interactions between the segmentation genes fushi tarazu, hairy, and engrailed in the Drosophila blastoderm. Cell 44, 949–57.CrossRefGoogle ScholarPubMed
Huang, F. Z., Bely, A. E., and Weisblat, D. A. (2001). Stochastic WNT signaling between nonequivalent cells regulates adhesion but not fate in the two-cell leech embryo. Curr. Biol. 11, 1–7.CrossRefGoogle Scholar
Hunkapiller, T., and Hood, L. (1989). Diversity of the immunoglobulin gene superfamily. Adv. Immunol. 44, 1–63.CrossRefGoogle ScholarPubMed
Huppert, S. S., Jacobsen, T. L., and Muskavitch, M. A. (1997). Feedback regulation is central to Delta–Notch signalling required for Drosophila wing vein morphogenesis. Development 124, 3283–91.Google ScholarPubMed
Hutson, M. S., Tokutake, Y., Chang, M. S., Bloor, J. W., Venakides, S., Kiehart, D. P., and Edwards, G. S. (2003). Forces for morphogenesis investigated with laser microsurgery and quantitative modeling. Science 300, 145–9.CrossRefGoogle ScholarPubMed
Hynes, R. O. (1987). Integrins: a family of cell surface receptors. Cell 48, 349–54.CrossRefGoogle ScholarPubMed
Hynes, R. O. (1992). Integrins: versatility, modulation, and signaling in cell adhesion. Cell 69, 11–25.CrossRefGoogle ScholarPubMed
Hynes, R. O., and Lander, A. D. (1992). Contact and adhesive specificities in the associations, migrations, and targeting of cells and axons. Cell 68, 303–22.CrossRefGoogle ScholarPubMed
Ingber, D. E. (1991). Extracellular matrix and cell shape: potential control points for inhibition of angiogenesis. J. Cell. Biochem. 47, 236–41.CrossRefGoogle ScholarPubMed
Ingber, D. E., and Folkman, J. (1989a). How does extracellular matrix control capillary morphogenesis?Cell 58, 803–5.CrossRefGoogle ScholarPubMed
Ingber, D. E., and Folkman, J. (1989b). Mechanochemical switching between growth and differentiation during fibroblast growth factor-stimulated angiogenesis in vitro: role of extracellular matrix. J. Cell Biol. 109, 317–30.CrossRefGoogle ScholarPubMed
Ingham, P. W. (1988). The molecular genetics of embryonic pattern formation in Drosophila. Nature 335, 25–34.CrossRefGoogle ScholarPubMed
Ingolia, N. T. (2004). Topology and robustness in the Drosophila segment polarity network. PLoS Biol. 2, 805–15.CrossRefGoogle ScholarPubMed
International Human Genome Consortium (2004). Finishing the euchromatic sequence of the human genome. Nature 431, 931–45.CrossRef
Irvine, K. D., and Wieschaus, E. (1994). Cell intercalation during Drosophila germband extension and its regulation by pair-rule segmentation genes. Development 120, 827–41.Google ScholarPubMed
Ish-Horowicz, D., Pinchin, S. M., Ingham, P. W., and Gyurkovics, H. G. (1989). Autocatalytic ftz activation and instability induced by ectopic ftz expression. Cell 57, 223–232.CrossRefGoogle ScholarPubMed
Israelachvili, J. N. (1991). Intermolecular and Surface Forces. Academic Press, London, San Diego.Google Scholar
Itow, T. (1986). Inhibitors of DNA synthesis change the differentiation of body segments and increase the segment number in horseshoe crab embryos. Roux's Arch. Dev. Biol. 195, 323–33.CrossRefGoogle ScholarPubMed
Jacobson, A. G., Oster, G. F., Odell, G. M., and Cheng, L. Y. (1986). Neurulation and the cortical tractor model for epithelial folding. J. Embryol. Exp. Morphol. 96, 19–49.Google ScholarPubMed
Jacobson, K. A., Sheets, E. D., and Simson, R. (1995). Revisiting the fluid mosaic model of membranes. Science 268, 1441–2.CrossRefGoogle ScholarPubMed
Jacobson, K. A., Moore, S. E., Yang, B., Doherty, P., Gordon, G. W., and Walsh, F. S. (1997). Cellular determinants of the lateral mobility of neural cell adhesion molecules. Biochim. Biophys. Acta 1330, 138–44.CrossRefGoogle ScholarPubMed
Jaffe, L. A. (1976). Fast block to polyspermy in sea urchin eggs is electrically mediated. Nature 261, 68–71.CrossRefGoogle ScholarPubMed
Jaffe, L. A., and Cross, N. L. (1986). Electrical regulation of sperm–egg fusion. Ann. Rev. Physiol. 48, 191–200.CrossRefGoogle ScholarPubMed
Jaffe, L. A., Sharp, A. P., and Wolf, D. P. (1983). Absence of an electrical polyspermy block in the mouse. Dev. Biol. 96, 317–23.CrossRefGoogle ScholarPubMed
Jaffe, L. A., Giusti, A. F., Carroll, D. J., and Foltz, K. R. (2001). Ca2+ signalling during fertilization of echinoderm eggs. Semin. Cell Dev. Biol. 12, 45–51.CrossRefGoogle ScholarPubMed
Jakab, K., Neagu, A., Mironov, V., Markwald, R. R., and Forgacs, G. (2004). Engineering biological structures of prescribed shape using self-assembling multicellular systems. Proc. Nat. Acad. Sci. USA 101, 2864–9.CrossRefGoogle ScholarPubMed
Janmey, P. (1998). The cytoskeleton and cell signaling: component localization and mechanical coupling. Physiol. Rev. 78, 763–81.CrossRefGoogle ScholarPubMed
Janson, L. W., and Taylor, D. L. (1993). In vitro models of tail contraction and cytoplasmic streaming in amoeboid cells. J. Cell Biol. 123, 345–56.CrossRefGoogle ScholarPubMed
Jen, W. C., Gawantka, V., Pollet, N., Niehrs, C., and Kintner, C. (1999). Periodic repression of Notch pathway genes governs the segmentation of Xenopus embryos. Genes Dev. 13, 1486–99.CrossRefGoogle ScholarPubMed
Jiang, T., Jung, H., Widelitz, R. B., and Chuong, C. (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
Jiang, Y., Levine, H., and Glazier, J. A. (1998). Possible cooperation of differential adhesion and chemotaxis in mound formation of Dictyostelium. Biophys. J. 75, 2615–25.CrossRefGoogle ScholarPubMed
Jiang, Y. J., Aerne, B. L., Smithers, L., Haddon, C., Ish-Horowicz, D., and Lewis, J. (2000). Notch signalling and the synchronization of the somite segmentation clock. Nature 408, 475–9.Google ScholarPubMed
Jimenez, G., Griffiths, S. D., Ford, A. M., Greaves, M. F., and Enver, T. (1992). Activation of the beta-globin locus control region precedes commitment to the erythroid lineage. Proc. Nat. Acad. Sci. USA 89, 10618–22.CrossRefGoogle ScholarPubMed
Jones, K. T. (1998). Protein kinase C action at fertilization: overstated or undervalued?Rev. Reprod. 3, 7–12.CrossRefGoogle ScholarPubMed
Joos, T. O., Whittaker, C. A., Meng, F., DeSimone, D. W., Gnau, V., and Hausen, P. (1995). Integrin alpha 5 during early development of Xenopus laevis. Mech. Dev. 50, 187–99.CrossRefGoogle ScholarPubMed
Jouve, C., Palmeirim, I., Henrique, D., et al. (2000). Notch signalling is required for cyclic expression of the hairy-like gene HES1 in the presomitic mesoderm. Development 127, 1421–9.Google ScholarPubMed
Jouve, C., Iimura, T., and Pourquié, O. (2002). Onset of the segmentation clock in the chick embryo: evidence for oscillations in the somite precursors in the primitive streak. Development 129, 1107–17.Google ScholarPubMed
Juan, H., and Hamada, H. (2001). Roles of nodal–lefty regulatory loops in embryonic patterning of vertebrates. Genes Cells 6, 923–30.CrossRefGoogle ScholarPubMed
Kabata, H., Kurosawa, O., Arai, I., et al. (1993). Visualization of single molecules of RNA polymerase sliding along DNA. Science 262, 1561–3.CrossRefGoogle ScholarPubMed
Kadler, K. E., Holmes, D. F., Trotter, J. A., and Chapman, J. A. (1996). Collagen fibril formation. Biochem. J. 316, 1–11.CrossRefGoogle ScholarPubMed
Kalodimos, C. G., Biris, N., Bonvin, A. M., et al. (2004). Structure and flexibility adaptation in nonspecific and specific protein–DNA complexes. Science 305, 386–9.CrossRefGoogle ScholarPubMed
Kamawaki, Y., Raya, A., Raya, R. M., Rodriquez-Esteban, C., and Belmont, J. C. I. (2005). Retinoic acid signalling links, left–right asymmetric patterning and bilaterally symmetric somitogenesis in the zebrafish embryo. Nature 435, 165–71.CrossRefGoogle Scholar
Kamei, N., Swanson, W. J., and Glabe, C. G. (2000). A rapidly diverging EGF protein regulates species-specific signal transduction in early sea urchin development. Dev. Biol. 225, 267–76.CrossRefGoogle ScholarPubMed
Kaneko, K. (2003). Organization through intra–inter dynamics. In Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology (Müller, G. B. and Newman, S. A., eds.), pp. 195–220. MIT Press, Cambridge, MA.Google Scholar
Kaneko, K., and Yomo, T. (1994). Cell division, differentiation and dynamic clustering. Physica D 75, 89–102.CrossRefGoogle Scholar
Kaneko, K., and Yomo, T. (1997). Isologous diversification: a theory of cell differentiation. Bull. Math. Biol. 59, 139–96.CrossRefGoogle ScholarPubMed
Kaneko, K., and Yomo, T. (1999). Isologous diversification for robust development of cell society. J. Theor. Biol. 199, 243–56.CrossRefGoogle ScholarPubMed
Karr, T. L., Weir, M. P., Ali, Z., and Kornberg, T. (1989). Patterns of engrailed protein in early Drosophila embryos. Development 105, 605–612.Google ScholarPubMed
Kawaki, Y.,Raya, A.,Raya, R. M., Rodriguez-Esteban, C., and Belmonte, J. C. I. (2005). Retinoic acid signalling links left–right asymmetric patterning and bilaterally symmetric somitogenesis in the zebrafish embryo. Nature 435, 165–71.CrossRefGoogle Scholar
Keller, A. D. (1995). Model genetic circuits encoding autoregulatory transcription factors. J. Theor. Biol. 172, 169–85.CrossRefGoogle ScholarPubMed
Keller, R. (2000). The origin and morphogenesis of amphibian somites. Curr. Top. Dev. Biol. 47, 183–246.CrossRefGoogle ScholarPubMed
Keller, R. (2002). Shaping the vertebrate body plan by polarized embryonic cell movements. Science 298, 1950–4.CrossRefGoogle ScholarPubMed
Keller, R., and Danilchik, M. (1988). Regional expression, pattern and timing of convergence and extension during gastrulation of Xenopus laevis. Development 103, 193–209.Google ScholarPubMed
Keller, R. E., Danilchik, M., Gimlich, R., and Shih, J. (1985). The function and mechanism of convergent extension during gastrulation of Xenopus laevis. J. Embryol. Exp. Morphol. 89 Suppl., 185–209.Google ScholarPubMed
Keller, R., Cooper, M. S., Danilchik, M., Tibbetts, P., and Wilson, P. A. (1989). Cell intercalation during notochord development in Xenopus laevis. J. Exp. Zool. 251, 134–54.CrossRefGoogle ScholarPubMed
Keller, R., Shih, J., and Sater, A. (1992a). The cellular basis of the convergence and extension of the Xenopus neural plate. Dev. Dyn. 193, 199–217.CrossRefGoogle Scholar
Keller, R., Shih, J., Sater, A. K., and Moreno, C. (1992b). Planar induction of convergence and extension of the neural plate by the organizer of Xenopus. Dev. Dyn. 193, 218–34.CrossRefGoogle Scholar
Keller, R., Davidson, L., Edlund, A.et al. (2000). Mechanisms of convergence and extension by cell intercalation. Philos. Trans. R. Soc. Lond. B Biol. Sci. 355, 897–922.CrossRefGoogle ScholarPubMed
Kerszberg, M., and Changeux, J. P. (1998). A simple molecular model of neurulation. Bioessays 20, 758–70.3.0.CO;2-C>CrossRefGoogle ScholarPubMed
Kerszberg, M., and Wolpert, L. (1998). Mechanisms for positional signalling by morphogen transport: a theoretical study. J. Theor. Biol. 191, 103–14.CrossRefGoogle ScholarPubMed
Kimmins, S., and Sassone-Corsi, P. (2005). Chromatin remodeling and epigenetic features of germ cells. Nature 434, 583–9.CrossRefGoogle ScholarPubMed
Kiskowski, M. A., Alber, M. S., Thomas, G. L., et al. (2004). Interplay between activator–inhibitor coupling and cell–matrix adhesion in a cellular automaton model for chondrogenic patterning. Dev. Biol. 271, 372–87.CrossRefGoogle Scholar
Klein, C., and Hurlbut, C. S. (2002). Manual of Mineral Science. Wiley, New York.Google Scholar
Knoll, A. H. (2003). Life on a Young Planet: the First Three Billion Years of Evolution on Earth. Princeton University Press, Princeton, NJ.Google Scholar
Kofron, M., Spagnuolo, A., Klymkowsky, M., Wylie, C., and Heasman, J. (1997). The roles of maternal alpha-catenin and plakoglobin in the early Xenopus embryo. Development 124, 1553–60.Google ScholarPubMed
Kofron, M., Heasman, J., Lang, S. A., and Wylie, C. C. (2002). Plakoglobin is required for maintenance of the cortical actin skeleton in early Xenopus embryos and for cdc42-mediated wound healing. J. Cell Biol. 158, 695–708.CrossRefGoogle ScholarPubMed
Kohn, K. W. (1999). Molecular interaction map of the mammalian cell cycle control and DNA repair systems. Mol. Biol. Cell. 10, 2703–34.CrossRefGoogle ScholarPubMed
Kondo, S., and Asai, R. (1995). A reaction–diffusion wave on the skin of the marine angelfish Pomacanthus. Nature 376, 765–8.CrossRefGoogle ScholarPubMed
Kosher, R. A., Savage, M. P., and Chan, S. C. (1979). In vitro studies on the morphogenesis and differentiation of the mesoderm subjacent to the apical ectodermal ridge of the embryonic chick limb-bud. J. Embryol. Exp. Morphol. 50, 75–97.Google ScholarPubMed
Kosher, R. A., Walker, K. H., and Ledger, P. W. (1982). Temporal and spatial distribution of fibronectin during development of the embryonic chick limb bud. Cell. Differ. 11, 217–228.CrossRefGoogle ScholarPubMed
Kruse, K., Pantazis, P., Bollenbach, T., Julicher, F., and Gonzalez-Gaitan, M. (2004). Dpp gradient formation by dynamin-dependent endocytosis: receptor trafficking and the diffusion model. Development 131, 4843–56.CrossRefGoogle ScholarPubMed
Kubota, H. Y., Yoshimoto, Y., Yoneda, M., and Hiramoto, Y. (1987). Free calcium wave upon activation in Xenopus eggs. Dev. Biol. 119, 129–36.CrossRefGoogle ScholarPubMed
Kulesa, P. M., and Fraser, S. E. (2000). In ovo time-lapse analysis of chick hindbrain neural crest cell migration shows cell interactions during migration to the branchial arches. Development 127, 1161–72.Google ScholarPubMed
Kulesa, P. M., and Fraser, S. E. (2002). Cell dynamics during somite boundary formation revealed by time-lapse analysis. Science 298, 991–5.CrossRefGoogle ScholarPubMed
Kumano, G., and Smith, W. C. (2002). Revisions to the Xenopus gastrula fate map: implications for mesoderm induction and patterning. Dev. Dyn. 225, 409–21.CrossRefGoogle ScholarPubMed
Lallier, T. E., Whittaker, C. A., and DeSimone, D. W. (1996). Integrin alpha 6 expression is required for early nervous system development in Xenopus laevis. Development 122, 2539–54.Google ScholarPubMed
Lander, A. D., Nie, Q., and Wan, F. Y. (2002). Do morphogen gradients arise by diffusion?Dev. Cell 2, 785–96.CrossRefGoogle ScholarPubMed
Lane, M. C., and Sheets, M. D. (2000). Designation of the anterior/posterior axis in pregastrula Xenopus laevis. Dev. Biol. 225, 37–58.CrossRefGoogle ScholarPubMed
Lane, M. C., and Sheets, M. D. (2002). Rethinking axial patterning in amphibians. Dev. Dyn. 225, 434–47.CrossRefGoogle ScholarPubMed
Lane, M. C., and Smith, W. C. (1999). The origins of primitive blood in Xenopus: implications for axial patterning. Development 126, 423–34.Google ScholarPubMed
Lane, M. C., Koehl, M. A., Wilt, F., and Keller, R. (1993). A role for regulated secretion of apical extracellular matrix during epithelial invagination in the sea urchin. Development 117, 1049–60.Google ScholarPubMed
Langille, R. M., and Hall, B. K. (1993). Pattern formation and the neural crest. In The Vertebrate Skull (Hanken, J. and Hall, B. K., eds.), Vol. 1, Development, pp. 77–111. University of Chicago Press, Chicago.Google ScholarPubMed
Langman, J. (1981). Medical Embryology. Williams & Wilkins, Baltimore.Google Scholar
Lawrence, P. A. (1992). The Making of a Fly: the Genetics of Animal Design. Blackwell Scientific Publications, Oxford, Boston.Google Scholar
Lawson, K. A. (1983). Stage specificity in the mesenchyme requirement of rodent lung epithelium in vitro: a matter of growth control?J. Embryol. Exp. Morphol. 74, 183–206.Google ScholarPubMed
Lechleiter, J., Girard, S., Clapham, D., and Peralta, E. (1991). Subcellular patterns of calcium release determined by G protein-specific residues of muscarinic receptors. Nature 350, 505–8.CrossRefGoogle Scholar
Lechleiter, J. D., John, L. M., and Camacho, P. (1998). Ca2+ wave dispersion and spiral wave entrainment in Xenopus laevis oocytes overexpressing Ca2+ ATPases. Biophys. Chem. 72, 123–9.CrossRefGoogle ScholarPubMed
Leal, L. G. (1992). Laminar Flow and Convective Processes: Scaling Principles and Asymptotic Analysis. Butterworth-Heinemann, Boston.Google Scholar
Lengyel, I., and Epstein, I. R. (1991). Diffusion-induced instability in chemically reacting systems: steady-state multiplicity, oscillation, and chaos. Chaos 1, 69–76.CrossRefGoogle ScholarPubMed
Lengyel, I., and Epstein, I. R. (1992). A chemical approach to designing Turing patterns in reaction–diffusion systems. Proc. Nat. Acad. Sci. USA 89, 3977–9.CrossRefGoogle ScholarPubMed
Leonard, C. M., Fuld, H. M., Frenz, D. A., Downie, S. A., Massague, J., and Newman, S. A. (1991). Role of transforming growth factor-β in chondrogenic pattern formation in the embryonic limb: stimulation of mesenchymal condensation and fibronectin gene expression by exogenous TGF-beta and evidence for endogenous TGF-β-like activity. Dev. Biol. 145, 99–109.CrossRefGoogle Scholar
Lercher, M. J., Urrutia, A. O., and Hurst, L. D. (2002). Clustering of housekeeping genes provides a unified model of gene order in the human genome. Nat. Genet. 31, 180–3.CrossRefGoogle ScholarPubMed
Leslie, P. H. (1948). Some further notes on the use of matrices in population mathematics. Biometrika 35, 213–45.CrossRefGoogle Scholar
Levi, G., Ginsberg, D., Girault, J. M., Sabanay, I., Thiery, J. P., and Geiger, B. (1991). EP-cadherin in muscles and epithelia of Xenopus laevis embryos. Development 113, 1335–44.Google ScholarPubMed
Lewis, J. (2003). Autoinhibition with transcriptional delay: a simple mechanism for the zebrafish somitogenesis oscillator. Curr. Biol. 13, 1398–408.CrossRefGoogle ScholarPubMed
Li, S., Piotrowicz, R. S., Levin, E. G., Shyy, Y. J., and Chien, S. (1996). Fluid shear stress induces the phosphorylation of small heat shock proteins in vascular endothelial cells. Am. J. Physiol. 271, C994–1000.CrossRefGoogle ScholarPubMed
Li, S., Zhou, D., Lu, M. M., and Morrisey, E. E. (2004). Advanced cardiac morphogenesis does not require heart tube fusion. Science 305, 1619–22.CrossRefGoogle Scholar
Linsenmayer, T. F., Fitch, J. M., Gordon, M. K., Cai, C. X., Igoe, F., Marchant, J. K., and Birk, D. E. (1998). Development and roles of collagenous matrices in the embryonic avian cornea. Prog. Retin. Eye Res. 17, 231–65.Google ScholarPubMed
Lowery, L. A., and Sive, H. (2004). Strategies of vertebrate neurulation and a re-evaluation of teleost neural tube formation. Mech. Dev. 121, 1189–97.CrossRefGoogle Scholar
Lubarsky, B., and Krasnow, M. A. (2003). Tube morphogenesis: making and shaping biological tubes. Cell 112, 19–28.CrossRefGoogle ScholarPubMed
Lubkin, S. R., and Li, Z. (2002). Force and deformation on branching rudiments: cleaving between hypotheses. Biomech. Model Mechanobiol. 1, 5–16.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
Luo, Y., Kostetskii, I., and Radiche, G. L. (2005). N-cadherin is not essential for limb mesenchymal chondrogenesis. Dev. Dyn. 232, 336–44.CrossRefGoogle Scholar
Mandato, C. A., Benink, H. A., and Bement, W. M. (2000). Microtubule-actomyosin interactions in cortical flow and cytokinesis. Cell Motil. Cytoskeleton 45, 87–92.3.0.CO;2-0>CrossRefGoogle ScholarPubMed
Mandelbrot, B. B. (1983). The Fractal Geometry of Nature. W. H. Freeman, New York.
Maniatis, T., and Tasic, B. (2002). Alternative pre-mRNA splicing and proteome expansion in metazoans. Nature 418, 236–43.CrossRefGoogle ScholarPubMed
Manner, J. (2000). Cardiac looping in the chick embryo: a morphological review with special reference to terminological and biomechanical aspects of the looping process. Anat. Rec. 259, 248–62.3.0.CO;2-K>CrossRefGoogle ScholarPubMed
Mannervik, M., Nibu, Y., Zhang, H., and Levine, M. (1999). Transcriptional coregulators in development. Science 284, 606–9.CrossRefGoogle ScholarPubMed
Marom, K., Shapira, E., and Fainsod, A. (1997). The chicken caudal genes establish an anterior–posterior gradient by partially overlapping temporal and spatial patterns of expression. Mech. Dev. 64, 41–52.CrossRefGoogle Scholar
Marsden, M., and DeSimone, D. W. (2003). Integrin-ECM interactions regulate cadherin-dependent cell adhesion and are required for convergent extension in Xenopus. Curr. Biol. 13, 1182–91.CrossRefGoogle ScholarPubMed
Marshall, B. T., Long, M., Piper, J. W., Yago, T., McEver, R. P., and Zhu, C. (2003). Direct observation of catch bonds involving cell-adhesion molecules. Nature 423, 190–3.CrossRefGoogle ScholarPubMed
Martin, G. R. (1998). The roles of FGFs in the early development of vertebrate limbs. Genes Dev. 12, 1571–86.CrossRefGoogle ScholarPubMed
Martin, J. E., Adolf, D., and Wilcoxon, J. P. (1989). Rheology of the incipient gel: theory and data for epoxies. Polym. Prepr. Am. Chem. Soc. Div. Polym. Chem. 30, 83–84.Google Scholar
Martin, V. J., Littlefield, C. L., Archer, W. E., and Bode, H. R. (1997). Embryogenesis in hydra. Biol. Bull. 192, 345–63.CrossRefGoogle ScholarPubMed
Smith, Maynard J. (1978). Models in Ecology. Cambridge University Press, Cambridge, New York.Google Scholar
Mayr, E. (1982). The Growth of Biological Thought: Diversity, Evolution, and Inheritance. Belknap Press, Cambridge, MA.Google Scholar
McCarthy, R. A., and Hay, E. D. (1991). Collagen I, laminin, and tenascin: ultrastructure and correlation with avian neural crest formation. Int. J. Dev. Biol. 35, 437–52.Google ScholarPubMed
McDougall, A., Shearer, J., and Whitaker, M. (2000). The initiation and propagation of the fertilization wave in sea urchin eggs. Biol. Cell. 92, 205–14.CrossRefGoogle ScholarPubMed
McDowell, N., Gurdon, J. B., and Grainger, D. J. (2001). Formation of a functional morphogen gradient by a passive process in tissue from the early Xenopus embryo. Int. J. Dev. Biol. 45, 199–207.Google ScholarPubMed
McKim, K. S., Jang, J. K., and Manheim, E. A. (2002). Meiotic recombination and chromosome segregation in Drosophila females. Ann. Rev. Genet. 36, 205–32.CrossRefGoogle ScholarPubMed
McLaren, A. (1984). Meiosis and differentiation of mouse germ cells. Symp. Soc. Exp. Biol. 38, 7–23.Google ScholarPubMed
Medalia, O., Weber, I., Frangakis, A. S., Nicastro, D., Gerisch, G., and Baumeister, W. (2002). Macromolecular architecture in eukaryotic cells visualized by cryoelectron tomography. Science 298, 1209–13.CrossRefGoogle ScholarPubMed
Meek, K. M., and Fullwood, N. J. (2001). Corneal and scleral collagens – a microscopist's perspective. Micron 32, 261–72.CrossRefGoogle ScholarPubMed
Meier, S. (1984). Somite formation and its relationship to metameric patterning of the mesoderm. Cell Differ. 14, 235–43.CrossRefGoogle ScholarPubMed
Meinhardt, H. (1982). Models of Biological Pattern Formation. Academic Press, New York.
Meinhardt, H. (2001). Organizer and axes formation as a self-organizing process. Int. J. Dev. Biol. 45, 177–88.Google ScholarPubMed
Meinhardt, H., and Gierer, A. (2000). Pattern formation by local self-activation and lateral inhibition. Bioessays 22, 753–60.3.0.CO;2-Z>CrossRefGoogle ScholarPubMed
Meir, E., Dassow, G., Munro, E., and Odell, G. M. (2002). Robustness, flexibility, and the role of lateral inhibition in the neurogenic network. Curr. Biol. 12, 778–86.CrossRefGoogle ScholarPubMed
Melnick, M., and Jaskoll, T. (2000). Mouse submandibular gland morphogenesis: a paradigm for embryonic signal processing. Crit. Rev. Oral Biology and Medicine 11, 199–215.CrossRefGoogle ScholarPubMed
Merks, R. M. H., Newman, S. A., and Glazier, J. A. (2004). Cell-oriented modeling of in vitro capillary development. In Cellular Automata: Proc. 6th International Conf. on Cellular Automata for Research and Industry (Sloot, P. M. A., Chopard, B., and Hoekstra, A. G., eds.), pp. 425–34. Springer-Verlag, Amsterdam, The Netherlands.CrossRefGoogle Scholar
Metropolis, N., Rosenbluth, M. N., Rosenbluth, A., Teller, H., and Teller, E. (1953). Equations of state calculations by fast computing machines. J. Chem. Phys. 21, 1087–91.CrossRefGoogle Scholar
Minelli, A. (2003). The Development of Animal Form: Ontogeny, Morphology, and Evolution. Cambridge University Press, Cambridge, New York.CrossRefGoogle Scholar
Minelli, A., and Fusco, G. (2004). Evo-devo perspectives on segmentation: model organisms, and beyond. Trends Ecol. Evol. 19, 423–9.CrossRefGoogle ScholarPubMed
Miranti, C. K., and Brugge, J. S. (2002). Sensing the environment: a historical perspective on integrin signal transduction. Nat. Cell. Biol. 4, E83–E90.CrossRefGoogle ScholarPubMed
Misteli, T. (2001). Protein dynamics: implications for nuclear architecture and gene expression. Science 291, 843–7.CrossRefGoogle ScholarPubMed
Mittenthal, J. E., and Mazo, R. M. (1983). A model for shape generation by strain and cell–cell adhesion in the epithelium of an arthropod leg segment. J. Theor. Biol. 100, 443–83.CrossRefGoogle ScholarPubMed
Miura, T., and Maini, P. K. (2004). Speed of pattern appearance in reaction–diffusion models: implications in the pattern formation of limb bud mesenchyme cells. Bull. Math. Biol. 66, 627–49.CrossRefGoogle ScholarPubMed
Miura, T., and Shiota, K. (2000a). Extracellular matrix environment influences chondrogenic pattern formation in limb bud micromass culture: experimental verification of theoretical models. Anat. Rec. 258, 100–7.3.0.CO;2-3>CrossRefGoogle Scholar
Miura, T., and Shiota, K. (2000b). TGFβ2 acts as an “activator” molecule in reaction–diffusion model and is involved in cell sorting phenomenon in mouse limb micromass culture. Dev. Dyn. 217, 241–9.3.0.CO;2-K>CrossRefGoogle Scholar
Miura, T., and Shiota, K. (2000c). Time-lapse observation of branching morphogenesis of the lung bud epithelium in mesenchyme-free culture and its relationship with the localization of actin filaments. Int. J. Dev. Biol. 44, 899–902.Google Scholar
Miura, T., Komori, M., and Shiota, K. (2000). A novel method for analysis of the periodicity of chondrogenic patterns in limb bud cell culture: correlation of in vitro pattern formation with theoretical models. Anat. Embryol. (Berlin) 201, 419–28.CrossRefGoogle ScholarPubMed
Miyazaki, S., Shirakawa, H., Nakada, K., and Honda, Y. (1993). Essential role of the inositol 1,4,5-trisphosphate receptor/Ca2+ release channel in Ca2+ waves and Ca2+ oscillations at fertilization of mammalian eggs. Dev. Biol. 158, 62–78.CrossRefGoogle ScholarPubMed
Mlodzik, M. (2002). Planar cell polarization: do the same mechanisms regulate Drosophila tissue polarity and vertebrate gastrulation? Trends Genet. 18, 564–71.CrossRefGoogle ScholarPubMed
Moftah, M. Z., Downie, S. A., Bronstein, N. B., Mezentseva, N., Pu, J., Maher, P. A., and Newman, S. A. (2002). Ectodermal FGFs induce perinodular inhibition of limb chondrogenesis in vitro and in vivo via FGF receptor 2. Dev. Biol. 249, 270–82.CrossRefGoogle ScholarPubMed
Mombach, J. C., Glazier, J. A., Raphael, R. C., and Zajac, M. (1995). Quantitative comparison between differential adhesion models and cell sorting in the presence and absence of fluctuations. Phys. Rev. Lett. 75, 2244–7.CrossRefGoogle ScholarPubMed
Monk, N. A. (2003). Oscillatory expression of Hes1, p53, and NF-kappaB driven by transcriptional time delays. Curr. Biol. 13, 1409–13.CrossRefGoogle ScholarPubMed
Montalta-He, H., and Reichert, H. (2003). Impressive expressions: developing a systematic database of gene-expression patterns in Drosophila embryogenesis. Genome Biol. 4, 205.CrossRefGoogle ScholarPubMed
Montero, J. A., and Heisenberg, C. P. (2003). Adhesive crosstalk in gastrulation. Dev. Cell 5, 190–1.CrossRefGoogle ScholarPubMed
Morisco, C., Seta, K., Hardt, S. E., Lee, Y., Vatner, S. F., and Sadoshima, J. (2001). Glycogen synthase kinase 3β regulates GATA4 in cardiac myocytes. J. Biol. Chem. 276, 28586–97.CrossRefGoogle ScholarPubMed
Morrison, S. J., Perez, S. E., Qiao, Z.et al. (2000). Transient Notch activation initiates an irreversible switch from neurogenesis to gliogenesis by neural crest stem cells. Cell 101, 499–510.CrossRefGoogle ScholarPubMed
Muratov, C. B. (1997). Synchronization, chaos, and the breakdown of the collective domain oscillations in reaction–diffusion systems. Phys. Rev. E 55, 1463–77.CrossRefGoogle Scholar
Murray, A. W., and Hunt, T. (1993). The Cell Cycle: An Introduction. W. H. Freeman, New York.Google Scholar
Murray, A. W., and Kirschner, M. W. (1989). Dominoes and clocks: the union of two views of the cell cycle. Science 246, 614–21.CrossRefGoogle ScholarPubMed
Murray, J. D. (2002). Mathematical biology. Springer, New York.Google Scholar
Nagafuchi, A., and Takeichi, M. (1988). Cell binding function of E-cadherin is regulated by the cytoplasmic domain. EMBO J. 7, 3679–84.Google ScholarPubMed
Nagar, B., Overduin, M., Ikura, M., and Rini, J. M. (1996). Structural basis of calcium-induced E-cadherin rigidification and dimerization. Nature 380, 360–4.CrossRefGoogle ScholarPubMed
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
Nakanishi, Y., Morita, T., and Nogawa, H. (1987). Cell proliferation is not required for the initiation of early cleft formation in mouse embryonic submandibular epithelium in vitro. Development 99, 429–37.Google Scholar
Nakatsuji, N., Snow, M. H., and Wylie, C. C. (1986). Cinemicrographic study of the cell movement in the primitive-streak-stage mouse embryo. J. Embryol. Exp. Morphol. 96, 99–109.Google ScholarPubMed
Nakayama, T., Yakubo, K., and Orbach, R. (1994). Dynamical properties of fractal networks: scaling, numerical simulations and physical realizations. Rev. Mod. Phys. 66, 381–443.CrossRefGoogle Scholar
Nanjundiah, V. (2005). Mathematics and biology. Current Science 88, 388–93.Google Scholar
Narbonne, G. M. (2004). Modular construction of early Ediacaran complex life forms. Science 305, 1141–4.CrossRefGoogle ScholarPubMed
Needham, D., and Hochmuth, R. M. (1992). A sensitive measure of surface stress in the resting neutrophil. Biophys. J. 61, 1664–70.CrossRefGoogle ScholarPubMed
Neff, A. W., Malacinski, G. M., Wakahara, M., and Jurand, A. (1983). Pattern formation in amphibian embryos prevented from undergoing the classical “rotation response” to egg activation. Dev. Biol. 97, 103–12.CrossRefGoogle Scholar
Neff, A. W., Wakahara, M., Jurand, A., and Malacinski, G. M. (1984). Experimental analyses of cytoplasmic rearrangements which follow fertilization and accompany symmetrization of inverted Xenopus eggs. J. Embryol. Exp. Morphol. 80, 197–224.Google ScholarPubMed
Newgreen, D. F. (1989). Physical influences on neural crest cell migration in avian embryos: contact guidance and spatial restriction. Dev. Biol. 131, 136–48.CrossRefGoogle ScholarPubMed
Newgreen, D. F., and Minichiello, J. (1995). Control of epitheliomesenchymal transformation. I. Events in the onset of neural crest cell migration are separable and inducible by protein kinase inhibitors. Dev. Biol. 170, 91–101.CrossRefGoogle ScholarPubMed
Newman, S. A. (1977). Lineage and pattern in the developing wing bud. In Vertebrate Limb and Somite Morphogenesis (Ede, D. A., Hinchliffe, J. R., and Balls, M., eds.), pp. 181–97. Cambridge University Press, Cambridge.Google Scholar
Newman, S. A. (1988). Lineage and pattern in the developing vertebrate limb. Trends Genet. 4, 329–32.CrossRefGoogle ScholarPubMed
Newman, S. A. (1993). Is segmentation generic? BioEssays 15, 277–83.CrossRefGoogle ScholarPubMed
Newman, S. A. (1994). Generic physical mechanisms of tissue morphogenesis: a common basis for development and evolution. J. Evol. Biol. 7, 467–88.CrossRefGoogle Scholar
Newman, S. A. (1995). Interplay of genetics and physical processes of tissue morphogenesis in development and evolution: the biological fifth dimension. In “Interplay of Genetic and Physical Processes in the Development of Biological Form (Beysens, D., Forgacs, G., and Gaill, F., eds.), pp. 3–12. World Scientific, Singapore.CrossRefGoogle Scholar
Newman, S. A. (1998a). Epithelial morphogenesis: a physico-evolutionary interpretation. In Molecular Basis of Epithelial Appendage Morphogenesis (Chuong, C.-M., ed.), pp. 341–58. R. G. Landes, Austin, TX.Google Scholar
Newman, S. A. (1998b). Networks of extracellular fibers and the generation of morphogenetic forces. In Dynamical Networks in Physics and Biology (Beysens, D. and Forgacs, G., eds.), pp. 139–48. Springer-Verlag, Berlin.CrossRefGoogle Scholar
Newman, S. A. (2003a). The fall and rise of systems biology. GeneWatch 16, 8–12.Google Scholar
Newman, S. A. (2003b). From physics to development: the evolution of morphogenetic mechanisms. In Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology. (Müller, G. B. and Newman, S. A., eds.), pp. 221–39. MIT Press, Cambridge, MA.Google Scholar
Newman, S. A., and Comper, W. D. (1990). “Generic” physical mechanisms of morphogenesis and pattern formation. Development 110, 1–18.Google ScholarPubMed
Newman, S. A., and Frisch, H. L. (1979). Dynamics of skeletal pattern formation in developing chick limb. Science 205, 662–8.CrossRefGoogle ScholarPubMed
Newman, S. A., and Müller, G. B. (2000). Epigenetic mechanisms of character origination. J. Exp. Zool. (Mol. Evol. Dev.) 288, 304–17.3.0.CO;2-G>CrossRefGoogle ScholarPubMed
Newman, S. A., and Müller, G. B. (eds.) (2003). Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology. MIT Press, Cambridge, MA.Google Scholar
Newman, S. A., and Tomasek, J. J. (1996). Morphogenesis of connective tissues. In Extracellular Matrix (Comper, W. D., ed.), Vol. 2, Molecular Components and Interactions, pp. 335–69. Harwood Academic Publishers, Amsterdam.Google Scholar
Newman, S. A., Frisch, H. L., Perle, M. A., and Tomasek, J. J. (1981). Limb development: aspects of differentiation, pattern formation and morphogenesis. In Morphogenesis and Pattern Formation (Connolly, T. G., Brinkley, L. L., and Carlson, B. M., eds.), pp. 163–78. Raven Press, New York.Google Scholar
Newman, S. A., Frenz, D. A., Tomasek, J. J., and Rabuzzi, D. D. (1985). Matrix-driven translocation of cells and nonliving particles. Science 228, 885–9.CrossRefGoogle ScholarPubMed
Newman, S. A., Frenz, D. A., Hasegawa, E., and Akiyama, S. K. (1987). Matrix-driven translocation: dependence on interaction of amino-terminal domain of fibronectin with heparin-like surface components of cells or particles. Proc. Nat. Acad. Sci. USA 84, 4791–5.CrossRefGoogle ScholarPubMed
Newman, S. A., Frisch, H. L., and Percus, J. K. (1988). On the stationary state analysis of reaction–diffusion mechanisms for biological pattern formationJ. Theor. Biol. 134, 183–197 (published erratum appears in J. Theor. Biol. 135, 137 (1988)).CrossRefGoogle ScholarPubMed
Newman, S. A., Cloitre, M., Allain, C., Forgacs, G., and Beysens, D. (1997). Viscosity and elasticity during collagen assembly in vitro: relevance to matrix-driven translocation. Biopolymers 41, 337–47.3.0.CO;2-T>CrossRefGoogle ScholarPubMed
Newman, S. A., Forgacs, G., Hinner, B., Maier, C. W., and Sackmann, E. (2004). Phase transformations in a model mesenchymal tissue. Phys. Biol. 1, 100–9.CrossRefGoogle Scholar
Newport, J., and Kirschner, M. (1982a). A major developmental transition in early Xenopus embryos: I. Characterization and timing of cellular changes at the midblastula stage. Cell 30, 675–86.CrossRefGoogle Scholar
Newport, J., and Kirschner, M. (1982b). A major developmental transition in early Xenopus embryos: I. Control of the onset of transcription. Cell 30, 687–96.CrossRefGoogle Scholar
Nicklas, R. B., and Koch, C. A. (1969). Chromosome micromanipulation. 3. Spindle fiber tension and the reorientation of mal-oriented chromosomes. J. Cell Biol. 43, 40–50.CrossRefGoogle ScholarPubMed
Nieuwkoop, P. D. (1969). The formation of mesoderm in Urodelean amphibians. I. Induction by the endoderm. Wilhelm Roux' Arch. Entw. Mech. Org. 162, 341–73.CrossRefGoogle ScholarPubMed
Nieuwkoop, P. D. (1973). The “organization center” of the amphibian embryo: its origin, spatial organization and morphogenetic action. Adv. Morphogen. 10, 1–39.CrossRefGoogle ScholarPubMed
Nieuwkoop, P. D. (1992). The formation of the mesoderm in Urodelean amphibians VI. The self-organizing capacity of the induced meso-endoderm. Roux's Arch. Dev. Biol. 201, 18–29.CrossRefGoogle ScholarPubMed
Noden, D. M. (1984). Craniofacial development: new views on old problems. Anat. Rec. 208, 1–13.CrossRefGoogle ScholarPubMed
Noden, D. M. (1988). Interactions and fates of avian craniofacial mesenchyme. Development Suppl. 103, 121–40.Google ScholarPubMed
Nogawa, H., and Ito, T. (1995). Branching morphogenesis of embryonic mouse lung epithelium in mesenchyme-free culture. Development 121, 1015–22.Google ScholarPubMed
Nogawa, H., and Takahashi, Y. (1991). Substitution for mesenchyme by basement-membrane-like substratum and epidermal growth factor in inducing branching morphogenesis of mouse salivary epithelium. Development 112, 855–61.Google ScholarPubMed
Nonaka, S., Tanaka, Y., Okada, Y., et al. (1998). Randomization of left–right asymmetry due to loss of nodal cilia generating leftward flow of extraembryonic fluid in mice lacking KIF3B motor protein. Cell 95, 829–37.CrossRefGoogle ScholarPubMed
Nonaka, S., Shiratori, H., Saijoh, Y., and Hamada, H. (2002). Determination of left–right patterning of the mouse embryo by artificial nodal flow. Nature 418, 96–9.CrossRefGoogle ScholarPubMed
Norel, R., and Agur, Z. (1991). A model for the adjustment of the mitotic clock by cyclin and MPF levels. Science 251, 1076–8.CrossRefGoogle ScholarPubMed
Novak, B., and Tyson, J. J. (1993). Numerical analysis of a comprehensive model of M-phase control in Xenopus oocyte extracts and intact embryos. J. Cell Sci. 106, 1153–68.Google ScholarPubMed
Novak, B., Csikasz-Nagy, A., Gyorffy, B., Nasmyth, K., and Tyson, J. J. (1998). Model scenarios for evolution of the eukaryotic cell cycle. Philos. Trans. R. Soc. Lond. B Biol. Sci. 353, 2063–76.CrossRefGoogle ScholarPubMed
Novak, B., Toth, A., Csikasz-Nagy, A., Gyorffy, B., Tyson, J. J., and Nasmyth, K. (1999). Finishing the cell cycle. J. Theor. Biol. 199, 223–33.CrossRefGoogle ScholarPubMed
Oates, A. C., and Ho, R. K. (2002). Hairy/E (spl)-related (Her) genes are central components of the segmentation oscillator and display redundancy with the Delta/Notch signaling pathway in the formation of anterior segmental boundaries in the zebrafish. Development 129, 2929–46.Google ScholarPubMed
Oberlender, S. A., and Tuan, R. S. (1994). Expression and functional involvement of N-cadherin in embryonic limb chondrogenesis. Development 120, 177–87.Google ScholarPubMed
Oda, S., Deguchi, R., Mohri, T., Shikano, T., Nakanishi, S., and Miyazaki, S. (1999). Spatiotemporal dynamics of the [Ca2+]i rise induced by microinjection of sperm extract into mouse eggs: preferential induction of a Ca2+ wave from the cortex mediated by the inositol 1,4,5-trisphosphate receptor. Dev. Biol. 209, 172–85.CrossRefGoogle Scholar
Odell, G. M., Oster, G., Alberch, P., and Burnside, B. (1981). The mechanical basis of morphogenesis. I. Epithelial folding and invagination. Dev. Biol. 85, 446–62.CrossRefGoogle ScholarPubMed
Oheim, M., and Stuhmer, W. (2000). Tracking chromaffin granules on their way through the actin cortex. Eur. Biophys. J. 29, 67–89.CrossRefGoogle ScholarPubMed
Oliveri, P., and Davidson, E. H. (2004). Gene regulatory network controlling embryonic specification in the sea urchin. Curr. Opin. Genet. Dev. 14, 351–60.CrossRefGoogle ScholarPubMed
Oliveri, P., Carrick, D. M., and Davidson, E. H. (2002). A regulatory gene network that directs micromere specification in the sea urchin embryo. Dev. Biol. 246, 209–28.CrossRefGoogle ScholarPubMed
Opas, M., Davies, J. R., Zhou, Y., and Dziak, E. (2001). Formation of retinal pigment epithelium in vitro by transdifferentiation of neural retina cells. Int. J. Dev. Biol. 45, 633–42.Google ScholarPubMed
Ornitz, D. M., and Marie, P. J. (2002). FGF signaling pathways in endochondral and intramembranous bone development and human genetic disease. Genes Dev. 16, 1446–65.CrossRefGoogle ScholarPubMed
Ouyang, Q., and Swinney, H. (1991). Transition from a uniform state to hexagonal and striped Turing patterns. Nature 352, 610–12.CrossRefGoogle Scholar
Owen, M. R., andSherratt, J. A. (1998). Mathematical modelling of juxtacrine cell signalling. Math. Biosci. 153, 125–50.CrossRefGoogle ScholarPubMed
Owen, M. R., Sherratt, J. A., and Myers, S. R. (1999). How far can a juxtacrine signal travel? Proc. R. Soc. Lond. B Biol. Sci. 266, 579–85.CrossRefGoogle ScholarPubMed
Owen, M. R., Sherratt, J. A., and Wearing, H. J. (2000). Lateral induction by juxtacrine signaling is a new mechanism for pattern formation. Dev. Biol. 217, 54–61.CrossRefGoogle ScholarPubMed
Ozdamar, B., Bose, R., Barrios-Rodiles, M., Wang, H.-R., Zhang, Y., and Wrana, J. L. (2005). Regulation of the polarity protein Par6 by TGFβ receptors controls epithelial cell plasticity. Science 307, 1603–9.CrossRefGoogle ScholarPubMed
Pagan-Westphal, S. M., and Tabin, C. J. (1998). The transfer of left–right positional information during chick embryogenesis. Cell 93, 25–35.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
Pardanaud, L., and Dieterlen-Lievre, F. (2000). Ontogeny of the endothelial system in the avian model. Adv. Exp. Med. Biol. 476, 67–78.CrossRefGoogle ScholarPubMed
Pardanaud, L., Luton, D., Prigent, M., Bourcheix, L. M., Catala, M., and Dieterlen-Lievre, F. (1996). Two distinct endothelial lineages in ontogeny, one of them related to hemopoiesis. Development 122, 1363–71.Google ScholarPubMed
Parkinson, J., Kadler, K. E., and Brass, A. (1995). Simple physical model of collagen fibrillogenesis based on diffusion limited aggregation. J. Mol. Biol. 247, 823–31.CrossRefGoogle ScholarPubMed
Patan, S. (2000). Vasculogenesis and angiogenesis as mechanisms of vascular network formation, growth and remodeling. J. Neurooncol. 50, 1–15.CrossRefGoogle ScholarPubMed
Patel, N. H. (1994). Developmental evolution: insights from studies of insect segmentation. Science 266, 581–90.CrossRefGoogle ScholarPubMed
Patel, N. H., Kornberg, T. B., and Goodman, C. S. (1989). Expression of engrailed during segmentation in grasshopper and crayfish. Development 107, 201–12.Google ScholarPubMed
Patel, N. H., Ball, E. E., and Goodman, C. S. (1992). Changing role of even-skipped during the evolution of insect pattern formation. Nature 357, 339–42.CrossRefGoogle ScholarPubMed
Patel, N. H., Condron, B. G., and Zinn, K. (1994). Pair-rule expression patterns of even-skipped are found in both short- and long-germ beetles. Nature 367, 429–34.CrossRefGoogle Scholar
Pearson, J. E., and Ponce-Dawson, S. (1998). Crisis on skid row. Physica A 257, 141–48.CrossRefGoogle Scholar
Pecht, I., and Lancet, D. (1976). In Chemical Relaxation in Molecular Biology (Riegler, R. and Pecht, I., eds.) Springer-Verlag, Heidelberg.Google Scholar
Perrin, F. (1934). Mouvement Brownien d'un ellipsoide (I). Dispersion dielectrique pour des molecules ellipsoidales. J. Physique et la Radium Serie 7, 5, 497–511.CrossRefGoogle Scholar
Perrin, F. (1936). Mouvement Brownien d'un ellipsoide (II). Rotation libre et depolarisation des fluorescences. Translation et diffusion des molecules ellipsoidales. J. Physique et la Radium Serie 7, 7, 1–11.CrossRefGoogle Scholar
Perris, R., and Perissonotto, D. (2000). Role of the extracellular matrix during neural crest migration. Mech. Dev. 95, 3–21.CrossRefGoogle Scholar
Peters, K. G., Werner, S., Chen, G., and Williams, L. T. (1992). Two FGF receptor genes are differentially expressed in epithelial and mesenchymal tissues during limb formation and organogenesis in the mouse. Development 114, 233–43.Google ScholarPubMed
Phillips, H. M. (1969). Equilibrium measurements of embryonic cell adhesiveness: physical formulation and testing of the differential adhesion hypothesis. Ph.D. thesis, Johns Hopkins University, Baltimore.Google Scholar
Picton, H., Briggs, D., and Gosden, R. (1998). The molecular basis of oocyte growth and development. Mol. Cell. Endocrinol. 145, 27–37.CrossRefGoogle ScholarPubMed
Pollack, G. H. (2001). Is the gel a gel – and why does it matter? Jpn J. Physiol. 51, 649–60.CrossRefGoogle ScholarPubMed
Pourquié, O. (2000). Skin development: delta laid bare. Curr. Biol. 10, R425–8.CrossRefGoogle ScholarPubMed
Pourquié, O. (2001). Vertebrate somitogenesis. Ann. Rev. Cell Dev. Biol. 17, 311–50.CrossRefGoogle ScholarPubMed
Pourquié, O. (2003). The segmentation clock: converting embryonic time into spatial pattern. Science 301, 328–30.CrossRefGoogle ScholarPubMed
Pourquié, O., and Goldbeter, A. (2003). Segmentation clock: insights from computational models. Curr. Biol. 13, R632–4.CrossRefGoogle ScholarPubMed
Preston, B. N., Laurent, T. C., Comper, W. D., and Checkley, G. J. (1980). Rapid polymer transport in concentrated solutions through the formation of ordered structures. Nature 287, 499–503.CrossRefGoogle Scholar
Primmett, D. R., Norris, W. E., Carlson, G. J., Keynes, R. J., and Stern, C. D. (1989). Periodic segmental anomalies induced by heat shock in the chick embryo are associated with the cell cycle. Development 105, 119–30.Google ScholarPubMed
Purcell, E. M. (1977). Life at low Reynolds number. Am. J. Phys. 45, 3–11.CrossRefGoogle Scholar
Quaas, J., and Wylie, C. (2002). Surface contraction waves (SCWs) in the Xenopus egg are required for the localization of the germ plasm and are dependent upon maternal stores of the kinesin-like protein Xklp1. Dev. Biol. 243, 272–80.CrossRefGoogle ScholarPubMed
Raff, R. A. (1996). The Shape of Life: Genes, Development, and the Evolution of Animal Form. University of Chicago Press, Chicago.Google Scholar
Ramirez-Weber, F. A., and Kornberg, T. B. (1999). Cytonemes: cellular processes that project to the principal signaling center in Drosophila imaginal discs. Cell 97, 599–607.CrossRefGoogle ScholarPubMed
Rappaport, R. (1966). Experiments concerning the cleavage furrow in invertebrate eggs. J. Exp. Zool. 161, 1–8.CrossRefGoogle ScholarPubMed
Rappaport, R. (1986). Establishment of the mechanism of cytokinesis in animal cells. Int. Rev. Cytol. 105, 245–81.CrossRefGoogle ScholarPubMed
Rappaport, R. (1996). Cytokinesis in Animal Cells. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Rashevsky, N. (1960). Mathematical Biophysics: Physico-mathematical Foundations of Biology, Vol. 1, Dover, New York.Google Scholar
Raya, A., Kawakami, Y., Rodriguez-Esteban, C., et al.(2003). Notch activity induces Nodal expression and mediates the establishment of left–right asymmetry in vertebrate embryos. Genes Dev. 17, 1213–8.CrossRefGoogle ScholarPubMed
Raya, A., Kawakami, Y., Rodriguez-Esteban, C., et al. (2004). Notch activity acts as a sensor for extracellular calcium during vertebrate left–right determination. Nature 427, 121–8.CrossRefGoogle ScholarPubMed
Redner, S. (2001). A Guide to First Passage Processes.Cambridge University Press, Cambridge.
Reilly, K. M., and Melton, D. A. (1996). Short-range signaling by candidate morphogens of the TGF beta family and evidence for a relay mechanism of induction. Cell 86, 743–54.CrossRefGoogle ScholarPubMed
Reinitz, J., Mjolsness, E., and Sharp, D. H. (1995). Model for cooperative control of positional information in Drosophila by bicoid and maternal hunchback. J. Exp. Zool. 271, 47–56.CrossRefGoogle ScholarPubMed
Resnick, N., Collins, T., Atkinson, W., Bonthron, D. T., Dewey, C. F. Jr, and Gimbrone, M. A. Jr (1993). Platelet-derived growth factor B chain promoter contains a cis-acting fluid shear-stress-responsive element. Proc. Nat. Acad. Sci. USA 90, 4591–5 (published erratum appears in Proc. Nat. Acad. Sci. USA 90, 7908 (1993)).CrossRefGoogle ScholarPubMed
Riedl, R. (1977). A systems-analytical approach to macro-evolutionary phenomena. Q. Rev. Biol. 52, 351–70.CrossRefGoogle ScholarPubMed
Rieu, J. P., Kataoka, N., and Sawada, Y. (1998). Quantitative analysis of cell motion during sorting in 2D aggregates of dissociated hydra cells. Phys. Rev. E 57, 924–31.CrossRefGoogle Scholar
Rieu, J. P., Upadhyaya, A., Glazier, J. A., Ouchi, N. B., and Sawada, Y. (2000). Diffusion and deformations of single hydra cells in cellular aggregates. Biophys. J. 79, 1903–14.CrossRefGoogle ScholarPubMed
Robinson, E. E., Zazzali, K. M., Corbett, S. A., and Foty, R. A. (2003). Alpha5beta1 integrin mediates strong tissue cohesion. J. Cell. Sci. 116, 377–86.CrossRefGoogle ScholarPubMed
Rodriguez, I., and Basler, K. (1997). Control of compartmental affinity boundaries by hedgehog. Nature 389, 614–8.Google ScholarPubMed
Roegiers, F., McDougall, A., and Sardet, C. (1995). The sperm entry point defines the orientation of the calcium-induced contraction wave that directs the first phase of cytoplasmic reorganization in the ascidian egg. Development 121, 3457–66.Google ScholarPubMed
Roegiers, F., Djediat, C., Dumollard, R., Rouviere, C., and Sardet, C. (1999). Phases of cytoplasmic and cortical reorganizations of the ascidian zygote between fertilization and first division. Development 126, 3101–17.Google ScholarPubMed
Rooke, J. E., and Xu, T. (1998). Positive and negative signals between interacting cells for establishing neural fate. Bioessays 20, 209–14.3.0.CO;2-M>CrossRefGoogle ScholarPubMed
Ross, M. H., Kaye, G. I., and Pawlina, W. (2003). Histology: A Text and Atlas. Lippincott Williams & Wilkins, Philadelphia, PA.Google Scholar
Roy, P., Petroll, W. M., Cavanagh, H. D., and Jester, J. V. (1999). Exertion of tractional force requires the coordinated up-regulation of cell contractility and adhesion. Cell Motil. Cytoskeleton 43, 23–34.3.0.CO;2-M>CrossRefGoogle ScholarPubMed
Rubinstein, M., Colby, R. H., and Gillmor, J. R. (1989). Dynamic scaling for polymer gelation. Polym. Prepr. Am. Chem. Soc. Div. Polym. Chem. 30, 81–2.Google Scholar
Rudnick, J., and Gaspari, G. (2004). Elements of Random Walk. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Runft, L. L., Jaffe, L. A., and Mehlmann, L. M. (2002). Egg activation at fertilization: where it all begins. Dev. Biol. 245, 237–54.CrossRefGoogle ScholarPubMed
Rupp, R. A., Singhal, N., and Veenstra, G. J. (2002). When the embryonic genome flexes its muscles. Eur. J. Biochem 269, 2294–9.CrossRefGoogle ScholarPubMed
Rupp, P. A., Czirok, A., and Little, C. D. (2003). Novel approaches for the study of vascular assembly and morphogenesis in avian embryos. Trends Cardiovasc Med. 13, 283–8.CrossRefGoogle Scholar
Rustom, A., Saffrich, R., Markovic, I., Walther, P., and Gerdes, H. H. (2004). Nanotubular highways for intercellular organelle transport. Science 303, 1007–10.CrossRefGoogle ScholarPubMed
Ryan, P. L., Foty, R. A., Kohn, J., and Steinberg, M. S. (2001). Tissue spreading on implantable substrates is a competitive outcome of cell–cell vs. cell–substratum adhesivity. Proc. Nat. Acad. Sci. USA 98, 4323–7.CrossRefGoogle ScholarPubMed
Sahimi, M. (1994). Applications of Percolation Theory. Taylor & Francis, London.Google Scholar
Sakai, T., Larsen, M., and Yamada, K. M. (2003). Fibronectin requirement in branching morphogenesis. Nature 423, 876–81.CrossRefGoogle ScholarPubMed
Sakuma, R., Yi, Ohnishi Y., Meno, C.et al. (2002). Inhibition of Nodal signalling by Lefty mediated through interaction with common receptors and efficient diffusion. Genes Cells 7, 401–12.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I., and Jernvall, J. (2002). A gene network model accounting for development and evolution of mammalian teeth. Proc. Nat. Acad. Sci. USA 99, 8116–20.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I., Garcia-Fernandez, J., and Solé, R. V. (2000). Gene networks capable of pattern formation: from induction to reaction–diffusion. J. Theor. Biol. 205, 587–603.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I., Newman, S. A., and Solé, R. (2001a). Phenotypic and dynamical transitions in model genetic networks. I. Emergence of patterns and genotype–phenotype relationships. Evolution & Development 3, 84–94.CrossRefGoogle Scholar
Salazar-Ciudad, I., Solé, R., and Newman, S. A. (2001b). Phenotypic and dynamical transitions in model genetic networks. II. Application to the evolution of segmentation mechanisms. Evolution & Development 3, 95–103.CrossRefGoogle Scholar
Salazar-Ciudad, I., Jernvall, J., and Newman, S. A. (2003). Mechanisms of pattern formation in development and evolution. Development 130, 2027–37.CrossRefGoogle ScholarPubMed
Salthe, S. N. (1993). Development and Evolution: Complexity and Change in Biology. MIT Press, Cambridge, MA.Google Scholar
Sanchez, L., and Thieffry, D. (2001). A logical analysis of the Drosophila gap-gene system. J. Theor. Biol. 211, 115–41.CrossRefGoogle ScholarPubMed
Sanders, E. J. (1991). Embryonic cell invasiveness: an in vitro study of chick gastrulation. J. Cell Sci. 98, 403–7.Google Scholar
Sardet, C., Roegiers, F., Dumollard, R., Rouviere, C., and McDougall, A. (1998). Calcium waves and oscillations in eggs. Biophys. Chem. 72, 131–40.CrossRefGoogle ScholarPubMed
Sardet, C., Prodon, F., Dumollard, R., Chang, P., and Chenevert, J. (2002). Structure and function of the egg cortex from oogenesis through fertilization. Dev. Biol. 241, 1–23.CrossRefGoogle ScholarPubMed
Sato, Y., Yasuda, K., and Takahashi, Y. (2002). Morphological boundary forms by a novel inductive event mediated by Lunatic fringe and Notch during somitic segmentation. Development 129, 3633–44.Google ScholarPubMed
Saunders, J. W. Jr (1948). The proximo-distal sequence of origin of the parts of the chick wing and the role of the ectoderm. J. Exp. Zool. 108, 363–402.CrossRefGoogle ScholarPubMed
Savill, N. J., and Sherratt, J. A. (2003). Control of epidermal stem cell clusters by Notch-mediated lateral induction. Dev. Biol. 258, 141–53.CrossRefGoogle ScholarPubMed
Saxton, M. J., and Jacobson, K. (1997). Single-particle tracking: applications to membrane dynamics. Ann. Rev. Biophys. Biomol. Struct. 26, 373–99.CrossRefGoogle ScholarPubMed
Schier, A. F., and Gehring, W. J. (1993). Analysis of a fushi tarazu autoregulatory element: multiple sequence elements contribute to enhancer activity. EMBO J. 12, 1111–9.Google ScholarPubMed
Schlosser, G. and Wagner, G. P. (eds.) (2004). Modularity in Development and Evolution. University of Chicago Press, Chicago.Google Scholar
Schmalhausen, I. I. (1949). Factors of Evolution. Blakiston, Philadelphia.Google Scholar
Schroeder, T. E. (1975). Dynamics of the contractile ring. Soc. Gen. Physiol. Ser. 30, 305–34.Google ScholarPubMed
Schulte-Merker, S., and Smith, J. C. (1995). Mesoderm formation in response to Brachyury requires FGF signalling. Curr. Biol. 5, 62–7.CrossRefGoogle ScholarPubMed
Schuster, S., Marhl, M., and Hofer, T. (2002). Modelling of simple and complex calcium oscillations. From single-cell responses to intercellular signalling. Eur. J. Biochem. 269, 1333–55.CrossRefGoogle ScholarPubMed
Seifert, U. (1997). Configurations of fluid membranes and vesicles. Adv. Phys. 46, 13–137.CrossRefGoogle Scholar
Seilacher, A. (1992). Vendobionta and Psammocorallia – lost constructions of precambrian evolution. J. Geolog. Soc. Lond. 149, 607–13.CrossRefGoogle Scholar
Serini, G., Ambrosi, D., Giraudo, E., Gamba, A., Preziosi, L., and Bussolino, F. (2003). Modeling the early stages of vascular network assembly. EMBO J. 22, 1771–9.CrossRefGoogle ScholarPubMed
Shafrir, Y., and Forgacs, G. (2002). Mechanotransduction through the cytoskeleton. Am. J. Physiol. Cell Physiol. 282, C479–86.CrossRefGoogle ScholarPubMed
Shafrir, Y., ben-Avraham, D., and Forgacs, G. (2000). Trafficking and signaling through the cytoskeleton: a specific mechanism. J. Cell Sci. 113, 2747–57.Google ScholarPubMed
Shapiro, L., Fannon, A. M., Kwong, P. D., et al. (1995). Structural basis of cell–cell adhesion by cadherins. Nature 374, 327–37.CrossRefGoogle ScholarPubMed
Shav-Tal, Y., Darzacq, X., Shenoy, S. M., et al. (2004). Dynamics of single mRNPs in nuclei of living cells. Science 304, 1797–800.CrossRefGoogle ScholarPubMed
Sheetz, M. P. (2001). Cell control by membrane-cytoskeleton adhesion. Nat. Rev. Mol. Cell Biol. 2, 392–6.CrossRefGoogle ScholarPubMed
Sherwood, D. R., and McClay, D. R. (2001). LvNotch signaling plays a dual role in regulating the position of the ectoderm–endoderm boundary in the sea urchin embryo. Development 128, 2221–32.Google Scholar
Shih, J., and Keller, R. (1992). Cell motility driving mediolateral intercalation in explants of Xenopus laevis. Development 116, 901–14.Google ScholarPubMed
Shyy, J. Y., Li, Y. S., Lin, M. C., Chen, W., Yuan, S., Usami, S., and Chien, S. (1995a). Multiple cis-elements mediate shear stress-induced gene expression. J. Biomech. 28, 1451–7.CrossRefGoogle Scholar
Shyy, J. Y., Lin, M. C., Han, J., Lu, Y., Petrime, M., and Chien, S. (1995b). The cis-acting phorbol ester “12-O-tetradecanoylphorbol 13-acetate”- responsive element is involved in shear stress-induced monocyte chemotactic protein 1 gene expression. Proc. Nat. Acad. Sci. USA 92, 8069–73.CrossRefGoogle Scholar
Simonneau, L., Kitagawa, M., Suzuki, S., and Thiery, J. P. (1995). Cadherin 11 expression marks the mesenchymal phenotype: towards new functions for cadherins? Cell Adhes. Commun. 3, 115–30.CrossRefGoogle ScholarPubMed
Singer, S. J., and Nicolson, G. L. (1972). The fluid mosaic model of the structure of cell membranes. Science 175, 720–31.CrossRefGoogle ScholarPubMed
Sivasankar, S., Brieher, W., Lavrik, N., Gumbiner, B., and Leckband, D. (1999). Direct molecular force measurements of multiple adhesive interactions between cadherin ectodomains. Proc. Nat. Acad. Sci. USA 96, 11820–4.CrossRefGoogle ScholarPubMed
Sivasankar, S., Gumbiner, B., and Leckband, D. (2001). Direct measurements of multiple adhesive alignments and unbinding trajectories between cadherin extracellular domains. Biophys. J. 80, 1758–68.CrossRefGoogle ScholarPubMed
Small, S., Kraut, R., Hoey, T., Warrior, R., and Levine, M. (1991). Transcriptional regulation of a pair-rule stripe in Drosophila. Genes Dev. 5, 827–39.CrossRefGoogle ScholarPubMed
Small, S., Blair, A., and Levine, M. (1992). Regulation of even-skipped stripe 2 in the Drosophila embryo. EMBO J. 11, 4047–57.Google ScholarPubMed
Small, S., Blair, A., and Levine, M. (1996). Regulation of two pair-rule stripes by a single enhancer in the Drosophila embryo. Dev. Biol. 175, 314–24.CrossRefGoogle ScholarPubMed
Solnica-Krezel, L. (2003). Vertebrate development: taming the nodal waves. Curr. Biol. 13, R7–R9.CrossRefGoogle ScholarPubMed
Spehr, M., Gisselmann, G., Poplawski, A., et al. (2003). Identification of a testicular odorant receptor mediating human sperm chemotaxis. Science 299, 2054–8.CrossRefGoogle ScholarPubMed
Spemann, H., and Mangold, H. (1924). Über Induktion von Embryonalanlagen durch Implantation artfremder Organisatoren. Wilhelm Roux' Arch. Entw. Mech. Org. 100, 599–638.Google Scholar
Spooner, B. S., and Wessells, N. K. (1972). An analysis of salivary gland morphogenesis: role of cytoplasmic microfilaments and microtubules. Dev. Biol. 27, 38–54.CrossRefGoogle ScholarPubMed
St Johnston, D., and Nusslein-Volhard, C. (1992). The origin of pattern and polarity in the Drosophila embryo. Cell 68, 201–19.CrossRefGoogle ScholarPubMed
Standley, H. J., Zorn, A. M., and Gurdon, J. B. (2001). eFGF and its mode of action in the community effect during Xenopus myogenesis. Development 128, 1347–57.Google ScholarPubMed
Starz-Gaiano, M., and Lehmann, R. (2001). Moving towards the next generation. Mech. Dev. 105, 5–18.CrossRefGoogle ScholarPubMed
Stauber, M., Jackle, H., and Schmidt-Ott, U. (1999). The anterior determinant bicoid of Drosophila is a derived Hox class 3 gene. Proc. Nat. Acad. Sci. USA 96, 3786–9.CrossRefGoogle ScholarPubMed
Steinberg, M. S. (1963). Reconstruction of tissues by dissociated cells. Some morphogenetic tissue movements and the sorting out of embryonic cells may have a common explanation. Science 141, 401–8.CrossRefGoogle ScholarPubMed
Steinberg, M. S. (1978). Specific cell ligands and the differential adhesion hypothesis: how do they fit together? In Specificity of Embryological Interactions (Garrod, D. R., ed.), pp. 97–130. Chapman and Hall, London.CrossRefGoogle Scholar
Steinberg, M. S. (1998). Goal-directedness in embryonic development. Integrative Biology 1, 49–59.3.0.CO;2-Z>CrossRefGoogle Scholar
Steinberg, M. S., and Foty, R. A. (1997). Intercellular adhesions as determinants of tissue assembly and malignant invasion. J. Cell Physiol. 173, 135–9.3.0.CO;2-F>CrossRefGoogle ScholarPubMed
Steinberg, M. S., and Poole, T. J. (1982). Liquid behavior of embryonic tissues. In Cell Behavior (Bellairs, R. and Curtis, A. S. G., eds.), pp. 583–607. Cambridge University Press,Cambridge.Google Scholar
Steinberg, M. S., and Takeichi, M. (1994). Experimental specification of cell sorting, tissue spreading, and specific spatial patterning by quantitative differences in cadherin expression. Proc. Nat. Acad. Sci. USA 91, 206–9.CrossRefGoogle ScholarPubMed
Stern, C. D., and Bellairs, R. (1984). Mitotic activity during somite segmentation in the early chick embryo. Anat. Embryol. (Berlin) 169, 97–102.CrossRefGoogle ScholarPubMed
Stollewerk, A., Schoppmeier, M., and Damen, W. G. (2003). Involvement of Notch and Delta genes in spider segmentation. Nature 423, 863–5.CrossRefGoogle ScholarPubMed
Stopak, D., and Harris, A. K. (1982). Connective tissue morphogenesis by fibroblast traction. I. Tissue culture observations. Dev. Biol. 90, 383–398.CrossRefGoogle ScholarPubMed
Stossel, T. P. (2001). Manifesto for a cytoplasmic revolution. Science 293, 611.CrossRefGoogle Scholar
Stricker, S. A. (1999). Comparative biology of calcium signaling during fertilization and egg activation in animals. Dev. Biol. 211, 157–76.CrossRefGoogle ScholarPubMed
Strogatz, S. H. (1994). Nonlinear Dynamics and Chaos: With Applications to Physics, Biology, Chemistry, and Engineering. Perseus, Cambridge, MA.Google Scholar
Strogatz, S. H. (2003). Sync: The Emerging Science of Spontaneous Order. Theia, New York.Google Scholar
Subtelny, S., and Penkala, J. E. (1984). Experimental evidence for a morphogenetic role in the emergence of primordial germ cells from the endoderm in Rana pipiens. Differentiation 26, 211–19.CrossRefGoogle Scholar
Sun, B., Bush, S., Collins-Racie, L., et al. (1999). derriere: a TGF-beta family member required for posterior development in Xenopus. Development 126, 1467–82.Google ScholarPubMed
Sun, Q. Y. (2003). Cellular and molecular mechanisms leading to cortical reaction and polyspermy block in mammalian eggs. Microsc. Res. Tech. 61, 342–8.CrossRefGoogle ScholarPubMed
Supp, D. M., Witte, D. P., Potter, S. S., and Brueckner, M. (1997). Mutation of an axonemal dynein affects left–right asymmetry in inversus viscerum mice. Nature 389, 963–6.CrossRefGoogle ScholarPubMed
Supp, D. M., Potter, S. S., and Brueckner, M. (2000). Molecular motors: the driving force behind mammalian left–right development. Trends Cell Biol. 10, 41–5.CrossRefGoogle ScholarPubMed
Suzuki, K., Tanaka, Y., Nakajima, Y., et al. (1995). Spatiotemporal relationships among early events of fertilization in sea urchin eggs revealed by multiview microscopy. Biophys. J. 68, 739–48.CrossRefGoogle ScholarPubMed
Sveiczer, A., Csikasz-Nagy, A., Gyorffy, B., Tyson, J. J., and Novak, B. (2000). Modeling the fission yeast cell cycle: quantized cycle times in wee1– cdc25Delta mutant cells. Proc. Nat. Acad. Sci. USA 97, 7865–70.CrossRefGoogle ScholarPubMed
Szebenyi, G., Savage, M. P., Olwin, B. B., and Fallon, J. F. (1995). Changes in the expression of fibroblast growth factor receptors mark distinct stages of chondrogenesis in vitro and during chick limb skeletal patterning. Dev. Dyn. 204, 446–56.CrossRefGoogle ScholarPubMed
Takahashi, Y., and Nogawa, H. (1991). Branching morphogenesis of mouse salivary epithelium in basement membrane-like substratum separated from mesenchyme by the membrane filter. Development 111, 327–35.Google ScholarPubMed
Takeichi, M. (1991). Cadherin cell adhesion receptors as a morphogenetic regulator. Science 251, 1451–5.CrossRefGoogle ScholarPubMed
Takeichi, M. (1995). Morphogenetic roles of classic cadherins. Curr. Opin. Cell Biol. 7, 619–27.CrossRefGoogle ScholarPubMed
Takeichi, T., and Kubota, H. Y. (1984). Structural basis of the activation wave in the egg of Xenopus laevis. J. Embryol. Exp. Morphol. 81, 1–16.Google ScholarPubMed
Takke, C., and Campos-Ortega, J. A. (1999). her1, a zebrafish pair-rule like gene, acts downstream of notch signalling to control somite development. Development 126, 3005–14.Google ScholarPubMed
Tanaka, M., and Tickle, C. (2004). Tbx18 and boundary formation in chick somite and wing development. Dev. Biol. 268, 470–80.CrossRefGoogle ScholarPubMed
Taunton, J., Rowning, B. A., Coughlin, M. L., et al. (2000). Actin-dependent propulsion of endosomes and lysosomes by recruitment of N-WASP. J. Cell Biol. 148, 519–30.CrossRefGoogle ScholarPubMed
Tautz, D. (1992). Redundancies, development and the flow of information. Bioessays 14, 263–6.CrossRefGoogle Scholar
Technau, U., and Holstein, T. W. (1992). Cell sorting during the regeneration of Hydra from reaggregated cells. Dev. Biol. 151, 117–27.CrossRefGoogle ScholarPubMed
Teleman, A. A., and Cohen, S. M. (2000). Dpp gradient formation in the Drosophila wing imaginal disc. Cell 103, 971–80.CrossRefGoogle ScholarPubMed
Teleman, A. A., Strigini, M., and Cohen, S. M. (2001). Shaping morphogen gradients. Cell 105, 559–62.CrossRefGoogle ScholarPubMed
Terasaki, M. (1996). Actin filament translocations in sea urchin eggs. Cell Motil. Cytoskeleton 34, 48–56.3.0.CO;2-E>CrossRefGoogle ScholarPubMed
Thiebaud, C. H. (1979). Quantitative determination of amplified rDNA and its distribution during oogenesis in Xenopus laevis. Chromosoma 73, 37–44.CrossRefGoogle ScholarPubMed
Thompson, D. W. (1917). On Growth and Form.Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Tickle, C. (2003). Patterning systems – from one end of the limb to the other. Dev. Cell 4, 449–58.CrossRefGoogle ScholarPubMed
Tiedemann, H., Asashima, M., Grunz, H., and Knochel, W. (1998). Neural induction in embryos. Dev. Growth Differ. 40, 363–76.CrossRefGoogle ScholarPubMed
Tomasek, J. J., Mazurkiewicz, J. E., and Newman, S. A. (1982). Nonuniform distribution of fibronectin during avian limb development. Dev. Biol. 90, 118–126.CrossRefGoogle ScholarPubMed
Torza, S., and Mason, S. G. (1969). Coalescence of two immiscible liquid drops. Science 163, 813–14.CrossRefGoogle ScholarPubMed
Townes, P. L., and Holtfreter, J. (1955). Directed movements and selective adhesion of embryonic amphibian cells. J. Exp. Zool. 128, 53–120.CrossRefGoogle Scholar
Tritton, D. J. (1988). Physical Fluid Dynamics.Clarendon Press, Oxford.Google Scholar
Tsai, M. A., Frank, R. S., and Waugh, R. E. (1994). Passive mechanical behavior of human neutrophils: effect of cytochalasin B. Biophys. J. 66, 2166–72.CrossRefGoogle ScholarPubMed
Tsai, M. A., Waugh, R. E., and Keng, P. C. (1998). Passive mechanical behavior of human neutrophils: effects of colchicine and paclitaxel. Biophys. J. 74, 3282–91.CrossRefGoogle ScholarPubMed
Tsarfaty, I., Resau, J. H., Rulong, S., et al. (1992). The met proto-oncogene receptor and lumen formation. Science 257, 1258–61.CrossRefGoogle ScholarPubMed
Tsarfaty, I., Rong, S., Resau, J. H., et al. (1994). The Met proto-oncogene mesenchymal to epithelial cell conversion. Science 263, 98–101.CrossRefGoogle ScholarPubMed
Tseng, Y., Kole, T. P., and Wirtz, D. (2002). Micromechanical mapping of live cells by multiple-particle-tracking microrheology. Biophys. J. 83, 3162–76.CrossRefGoogle ScholarPubMed
Tsonis, P. A., Del Rio-Tsonis, K., Millan, J. L., and Wheelock, M. J. (1994). Expression of N-cadherin and alkaline phosphatase in chick limb bud mesenchymal cells: regulation by 1, 25-dihydroxyvitamin D3 or TGF-beta 1. Exp. Cell. Res. 213, 433–7.CrossRefGoogle ScholarPubMed
Turing, A. (1952). The chemical basis of morphogenesis. Phil. Trans. Roy. Soc. Lond. B 237, 37–72.CrossRefGoogle Scholar
Tyson, J. J. (1991). Modeling the cell division cycle: cdc2 and cyclin interactions. Proc. Nat. Acad. Sci. USA 88, 7328–32.CrossRefGoogle ScholarPubMed
Tyson, J. J., and Novak, B. (2001). Regulation of the eukaryotic cell cycle: molecular antagonism, hysteresis, and irreversible transitions. J. Theor. Biol. 210, 249–63.CrossRefGoogle ScholarPubMed
Umbhauer, M., Boucaut, J. C., and Shi, D. L. (2001). Repression of XMyoD expression and myogenesis by Xhairy-1 in Xenopus early embryo. Mech. Dev. 109, 61–8.CrossRefGoogle ScholarPubMed
Ushiki, T. (2002). Collagen fibers, reticular fibers and elastic fibers. A comprehensive understanding from a morphological viewpoint. Arch Histol. Cytol. 65, 109–26.CrossRefGoogle ScholarPubMed
Uyttendaele, H., Ho, J., Rossant, J., and Kitajewski, J. (2001). Vascular patterning defects associated with expression of activated Notch4 in embryonic endothelium. Proc. Nat. Acad. Sci. USA 98, 5643–8.CrossRefGoogle ScholarPubMed
Valberg, P. A., and Feldman, H. A. (1987). Magnetic particle motions within living cells. Measurement of cytoplasmic viscosity and motile activity. Biophys. J. 52, 551–61.CrossRefGoogle ScholarPubMed
Valentine, J. W. (2004). On the Origin of Phyla. University of Chicago Press, Chicago.Google Scholar
Valinsky, J. E., and Douarin, N. M. (1985). Production of plasminogen activator by migrating cephalic neural crest cells. EMBO J. 4, 1403–6.Google ScholarPubMed
Obberghen-Schilling, E., Roche, N. S., Flanders, K. C., Sporn, M. B., and Roberts, A. (1988). Transforming growth factor β1 positively regulates its own expression in normal and transformed cells. J. Biol. Chem. 263, 7741–6.Google Scholar
Oss, C. J., Gillman, C. F., and Neumann, A. W. (1975). Phagocytic Engulfment and Cell Adhesiveness.Marcel Dekker, New York.
Veis, A., and George, A. (1994). Fundamentals of interstitial collagen self-assembly. In Extracellular Matrix Assembly and Function (Yurchenco, P. D., Birk, D. E., and Mecham, R. P., eds.), pp. 15–45. Academic Press, San Diego.Google Scholar
Verheul, H. M., Voest, E. E., and Schlingemann, R. O. (2004). Are tumours angiogenesis-dependent? J. Pathol. 202, 5–13.CrossRefGoogle ScholarPubMed
Vermot, J., and Pourquié, O. (2005). Retinoic acid coordinates somitogenesis and left–right patterning in vertebrate embryos. Nature 435, 215–20.CrossRefGoogle ScholarPubMed
Vermot, J., Llamas, J. G., Fraulob, V., Niederreither, K., Chambon, P., and Dolle, P. (2005). Retinoic acid controls the bilateral symmetry of somite formation in the mouse embryo. Science 308, 563–6.CrossRefGoogle ScholarPubMed
Vernon, G. G., and Woolley, D. M. (1995). The propagation of a zone of activation along groups of flagellar doublet microtubules. Exp. Cell Res. 220, 482–94.CrossRefGoogle ScholarPubMed
Vernon, R. B., Angello, J. C., Iruela-Arispe, M. L., Lane, T. F., and Sage, E. H. (1992). Reorganization of basement membrane matrices by cellular traction promotes the formation of cellular networks in vitro. Lab. Invest. 66, 536–47.Google ScholarPubMed
Vernon, R. B., Lara, S. L., Drake, C. J., et al. (1995). Organized type I collagen influences endothelial patterns during “spontaneous angiogenesis in vitro”: planar cultures as models of vascular development. In Vitro Cell Dev. Biol. Anim. 31, 120–31.CrossRefGoogle ScholarPubMed
Vilar, J. M., Kueh, H. Y., Barkai, N., and Leibler, S. (2002). Mechanisms of noise resistance in genetic oscillators. Proc. Nat. Acad. Sci. USA 99, 5988–92.CrossRefGoogle ScholarPubMed
Dassow, G., and Munro, E. (1999). Modularity in animal development and evolution: elements of a conceptual framework for EvoDevo. J. Exp. Zool. (Mol. Dev. Evol.) 285, 307–25.3.0.CO;2-V>CrossRefGoogle Scholar
Dassow, G., Meir, E., Munro, E. M., and Odell, G. M. (2000). The segment polarity network is a robust developmental module. Nature 406, 188–92.CrossRefGoogle Scholar
Hippel, P. H., and Berg, O. G. (1989). Facilitated target location in biological systems. J. Biol. Chem. 264, 675–8.Google Scholar
Voronov, D. A., and Taber, L. A. (2002). Cardiac looping in experimental conditions: effects of extraembryonic forces. Dev. Dyn. 224, 413–21.CrossRefGoogle ScholarPubMed
Voronov, D. A., Alford, P. W., Xu, G., and Taber, L. A. (2004). The role of mechanical forces in dextral rotation during cardiac looping in the chick embryo. Dev. Biol. 272, 339–50.CrossRefGoogle ScholarPubMed
Waddington, C. H. (1942). Canalization of development and the inheritance of acquired characters. Nature 150, 563–5.CrossRefGoogle Scholar
Wagner, G. P., and Altenberg, L. (1996). Complex adaptations and the evolution of evolvability. Evolution 50, 967–6.CrossRefGoogle ScholarPubMed
Wakamatsu, Y., Maynard, T. M., and Weston, J. A. (2000). Fate determination of neural crest cells by NOTCH-mediated lateral inhibition and asymmetrical cell division during gangliogenesis. Development 127, 2811–21.Google ScholarPubMed
Wakely, J., and England, M. A. (1977). Scanning electron microscopy (SEM) of the chick embryo primitive streak. Differentiation 7, 181–6.CrossRefGoogle ScholarPubMed
Wallingford, J. B., and Harland, R. M. (2002). Neural tube closure requires Dishevelled-dependent convergent extension of the midline. Development 129, 5815–25.CrossRefGoogle Scholar
Wang, N., and Stamenovic, D. (2000). Contribution of intermediate filaments to cell stiffness, stiffening, and growth. Am. J. Physiol. Cell Physiol. 279, C188–94.CrossRefGoogle Scholar
Wang, N., Butler, J. P., and Ingber, D. E. (1993). Mechanotransduction across the cell surface and through the cytoskeleton. Science 260, 1124–7.CrossRefGoogle ScholarPubMed
Wassarman, P., Chen, J., Cohen, N., Litscher, E., Liu, C., Qi, H., and Williams, Z. (1999). Structure and function of the mammalian egg zona pellucida. J. Exp. Zool. 285, 251–8.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Wassarman, P. M. (1999). Mammalian fertilization: molecular aspects of gamete adhesion, exocytosis, and fusion. Cell 96, 175–83.CrossRefGoogle ScholarPubMed
Waters, C. M., Oberg, K. C., Carpenter, G., and Overholser, K. A. (1990). Rate constants for binding, dissociation, and internalization of EGF: effect of receptor occupancy and ligand concentration. Biochemistry 29, 3563–9.CrossRefGoogle ScholarPubMed
Wearing, H. J., Owen, M. R., and Sherratt, J. A. (2000). Mathematical modelling of juxtacrine patterning. Bull. Math. Biol. 62, 293–320.CrossRefGoogle ScholarPubMed
Webb, S. D., and Owen, M. R. (2004). Oscillations and patterns in spatially discrete models for developmental intercellular signalling. J. Math. Biol. 48, 444–76.CrossRefGoogle ScholarPubMed
Weidinger, G., Wolke, U., Koprunner, M., Thisse, C., Thisse, B., and Raz, E. (2002). Regulation of zebrafish primordial germ cell migration by attraction towards an intermediate target. Development 129, 25–36.Google ScholarPubMed
Weng, W., and Stemple, D. L. (2003). Nodal signaling and vertebrate germ layer formation. Birth Defects Res. Part C Embryo Today 69, 325–32.CrossRefGoogle ScholarPubMed
Wessel, G. M., and McClay, D. R. (1987). Gastrulation in the sea urchin embryo requires the deposition of crosslinked collagen within the extracellular matrix. Dev. Biol. 121, 149–65.CrossRefGoogle ScholarPubMed
West-Eberhard, M. J. (2003). Developmental Plasticity and Evolution. Oxford University Press, Oxford, New York.Google Scholar
Wheelock, M. J., and Johnson, K. R. (2003). Cadherins as modulators of cellular phenotype. Ann. Rev. Cell Dev. Biol. 19, 207–35.CrossRefGoogle ScholarPubMed
White, J. G., and Borisy, G. G. (1983). On the mechanisms of cytokinesis in animal cells. J. Theor. Biol. 101, 289–316.CrossRefGoogle ScholarPubMed
Wikramanayake, A. H., Huang, L., and Klein, W. H. (1998). beta-Catenin is essential for patterning the maternally specified animal–vegetal axis in the sea urchin embryo. Proc. Nat. Acad. Sci. USA 95, 9343–8.CrossRefGoogle ScholarPubMed
Wilkins, A. S. (1992). Genetic analysis of animal development. Wiley-Liss, New York.Google Scholar
Wilkins, A. S. (1997). Canalization: a molecular genetic perspective. BioEssays 19, 257–62.CrossRefGoogle ScholarPubMed
Wilkins, A. S. (2002). The Evolution of Developmental Pathways. Sinauer Associates, Sunderland, MA.Google Scholar
Williams, A. F., and Barclay, A. N. (1988). The immunoglobulin superfamily – domains for cell surface recognition. Ann. Rev. Immunol. 6, 381–405.CrossRefGoogle ScholarPubMed
Wilson, P. D. (2004). Polycystic kidney disease: new understanding in the pathogenesis. Int. J. Biochem. Cell Biol. 36, 1868–73.CrossRefGoogle ScholarPubMed
Wimmer, E. A., Carleton, A., Harjes, P., Turner, T., and Desplan, C. (2000). Bicoid-independent formation of thoracic segments in Drosophila. Science 287, 2476–9.CrossRefGoogle ScholarPubMed
Winther, R. G. (2001). Varieties of modules: kinds, levels, origins, and behaviors. J. Exp. Zool. 291, 116–29.CrossRefGoogle ScholarPubMed
Wolf, D. M., and Eeckman, F. H. (1998). On the relationship between genomic regulatory element organization and gene regulatory dynamics. J. Theor. Biol. 195, 167–86.CrossRefGoogle ScholarPubMed
Wolf, J., and Heinrich, R. (1997). Dynamics of two-component biochemical systems in interacting cells; synchronization and desynchronization of oscillations and multiple steady states. Biosystems 43, 1–24.CrossRefGoogle ScholarPubMed
Wolfram, S. (2002). A New Kind of Science. Wolfram Media, Champaign, IL.Google Scholar
Wolpert, L. (1969). Positional information and the spatial pattern of cellular differentiation. J. Theor. Biol. 25, 1–47.CrossRefGoogle ScholarPubMed
Wolpert, L. (2002). Principles of Development. Oxford University Press, Oxford, New York.Google Scholar
Wong, G. K., Allen, P. G., and Begg, D. A. (1997). Dynamics of filamentous actin organization in the sea urchin egg cortex during early cleavage divisions: implications for the mechanism of cytokinesis. Cell Motil. Cytoskeleton 36, 30–42.3.0.CO;2-L>CrossRefGoogle Scholar
Woolley, D. M., and Vernon, G. G. (2002). Functional state of the axonemal dyneins during flagellar bend propagation. Biophys. J. 83, 2162–9.CrossRefGoogle ScholarPubMed
Wylie, C. C., and Heasman, J. (1993). Migration, proliferation, and potency of primordial germ cells. Seminars in Dev. Biol. 4, 161–170.CrossRefGoogle Scholar
Xiao, S., and Knoll, A. H. (2000). Phosphatized animal embryos from the Neoproterozoic Doushantuo Formation in Weng'an, Guizhou, South China. Paleontology 74, 767–88.CrossRefGoogle Scholar
Xiao, S., Yuan, X., and Knoll, A. H. (2000). Eumetazoan fossils in terminal Proterozoic phosphorites? Proc. Nat. Acad. Sci. USA 97, 13684–9.CrossRefGoogle ScholarPubMed
Xu, Z., and Tung, V. W. (2001). Temporal and spatial variations in slow axonal transport velocity along peripheral motoneuron axons. Neuroscience 102, 193–200.CrossRefGoogle ScholarPubMed
Yamada, S., Wirtz, D., and Kuo, S. C. (2000). Mechanics of living cells measured by laser tracking microrheology. Biophys. J. 78, 1736–47.CrossRefGoogle ScholarPubMed
Yamamoto, K., and Yoneda, M. (1983). Cytoplasmic cycle in meiotic division of starfish oocytes. Dev. Biol. 96, 166–72.CrossRefGoogle ScholarPubMed
Yamamoto, M., Saijoh, Y., Perea-Gomez, A., et al. (2004). Nodal antagonists regulate formation of the anteroposterior axis of the mouse embryo. Nature 428, 387–92.CrossRefGoogle ScholarPubMed
Yanagimachi, R., and Noda, Y. D. (1970). Electron microscope studies of sperm incorporation into the golden hamster egg. Am. J. Anat. 128, 429–62.CrossRefGoogle ScholarPubMed
Yoneda, M. (1973). Tension at the surface of sea urchin eggs on the basis of “liquid-drop” concept. Adv. Biophys. 4, 153–90.Google ScholarPubMed
Yoshida, M., Inaba, K., and Morisawa, M. (1993). Sperm chemotaxis during the process of fertilization in the ascidians Ciona savignyi and Ciona intestinalis. Dev. Biol. 157, 497–506.CrossRefGoogle ScholarPubMed
Yurchenco, P. D., and O'Rear, J. J. (1994). Basal lamina assembly. Curr. Opin. Cell. Biol. 6, 674–681.CrossRefGoogle ScholarPubMed
Zachariae, W., and Nasmyth, K. (1999). Whose end is destruction: cell division and the anaphase-promoting complex. Genes Dev. 13, 2039–58.CrossRefGoogle ScholarPubMed
Zahalak, G. I., Wagenseil, J. E., Wakatsuki, T., and Elson, E. L. (2000). A cell-based constitutive relation for bio-artificial tissues. Biophys. J. 79, 2369–81.CrossRefGoogle ScholarPubMed
Zajac, M., Jones, G. L., and Glazier, J. A. (2000). Model of convergent extension in animal morphogenesis. Phys. Rev. Lett. 85, 2022–5.CrossRefGoogle ScholarPubMed
Zajac, M., Jones, G. L., and Glazier, J. A. (2003). Simulating convergent extension by way of anisotropic differential adhesion. J. Theor. Biol. 222, 247–59.CrossRefGoogle ScholarPubMed
Zeng, W., Thomas, G. L., Newman, S. A., and Glazier, J. A. (2003). A novel mechanism for mesenchymal condensation during limb chondrogenesis in vitro. In Mathematical Modelling and Computing in Biology and Medicine, Proc. Fifth ESMTB Conf 2002 (Capasso, V., ed.), pp. 80–6. Società Editrice Esculapio, Bologna, Italy.Google Scholar
Zhang, J., Houston, D. W., King, M. L., Payne, C., Wylie, C., and Heasman, J. (1998). The role of maternal VegT in establishing the primary germ layers in Xenopus embryos. Cell 94, 515–24.CrossRefGoogle ScholarPubMed
Adams, C. L., Nelson, W. J., and Smith, S. J. (1996). Quantitative analysis of cadherin–catenin–actin reorganization during development of cell–cell adhesion. J. Cell Biol. 135, 1899–911.CrossRefGoogle ScholarPubMed
Adams, C. L., Chen, Y. T., Smith, S. J., and Nelson, W. J. (1998). Mechanisms of epithelial cell–cell adhesion and cell compaction revealed by high-resolution tracking of E-cadherin-green fluorescent protein. J. Cell Biol. 142, 1105–19.CrossRefGoogle ScholarPubMed
Afzelius, B. A. (1985). The immotile-cilia syndrome: a microtubule-associated defect. CRC Crit. Rev. Biochem. 19, 63–87.CrossRefGoogle ScholarPubMed
Agius, E., Oegeschlager, M., Wessely, O., Kemp, C., and Roberts, E. M. (2000). Endodermal Nodal-related signals and mesoderm induction in Xenopus. Development 127, 1173–83.Google ScholarPubMed
Agutter, P. S., and Wheatley, D. N. (2000). Random walk and cell size. BioEssays 22, 1018–23.3.0.CO;2-Y>CrossRefGoogle Scholar
Akam, M. (1989). Making stripes inelegantly. Nature 341, 282–3.CrossRefGoogle ScholarPubMed
Akatiya, T., and Bronner-Fraser, M. (1992). Expression of cell adhesion molecules during initiation and cessation of neural crest cell migration. Dev. Dyn. 194, 12–20.CrossRefGoogle Scholar
Alber, M. S., Kiskowski, M. A., Glazier, J. A., and Jiang, Y. (2003). On cellular automaton approaches to modeling biological cells. In Mathematical Systems Theory in Biology, Communication, and Finance (Rosenthal, J. and Gilliam, D. S., eds.), Vol. 134, pp. 1–40. New York, Springer-Verlag.CrossRefGoogle Scholar
Alber, M., Hentschel, H. G. E., Kazmierczak, B., and Newman, S. A. (2005). Existence of solutions to a new model of biological pattern formation. Journal of Mathematical Analysis and Application 308, 175–194.CrossRefGoogle Scholar
Alberts, B., Johnson, A., Lewis, J., Raff, M., Roberts, K., and Walter, P. (2002). Molecular Biology of the Cell. Garland Science, New York.Google Scholar
Allaerts, W. (1991). On the role of gravity and positional information in embryological axis formation and tissue compartmentalization. Acta Biotheor. 39, 47–62.CrossRefGoogle ScholarPubMed
Amonlirdviman, K., Khare, N. A., Tree, D. R., Chen, W. S., Axelrod, J. D., and Tomlin, C. J. (2005).Mathematical modeling of planar cell polarity to understand domineering nonautonomy. Science 307, 423–6.CrossRefGoogle ScholarPubMed
Anderson, A. R., and Chaplain, M. A. (1998). Continuous and discrete mathematical models of tumor-induced angiogenesis. Bull. Math. Biol. 60, 857–99.CrossRefGoogle ScholarPubMed
Anderson, K. V., and Ingham, P. W. (2003). The transformation of the model organism: a decade of developmental genetics. Nat. Genet. 33 Suppl., 285–93.CrossRefGoogle ScholarPubMed
Angres, B., Barth, A., and Nelson, W. J. (1996). Mechanism for transition from initial to stable cell–cell adhesion: kinetic analysis of E-cadherin-mediated adhesion using a quantitative adhesion assay. J. Cell Biol. 134, 549–57.CrossRefGoogle ScholarPubMed
Armstrong, P. B. (1989). Cell sorting out: the self-assembly of tissues in vitro. Crit. Rev. Biochem. and Mol. Biol. 24, 119–49.CrossRefGoogle ScholarPubMed
Artavanis-Tsakonas, S., Rand, M. D., and Lake, R. J. (1999). Notch signaling: cell fate control and signal integration in development. Science 284, 770–6.CrossRefGoogle ScholarPubMed
Arthur, W. (1997). The Origin of Animal Body Plans: A Study in Evolutionary Developmental Biology. Cambridge, New York, Cambridge University Press.CrossRefGoogle Scholar
Atherton-Fessler, S., Hannig, G., and Piwnica-Worms, H. (1993). Reversible tyrosine phosphorylation and cell cycle control. Semin. Cell Biol. 4, 433–42.CrossRefGoogle ScholarPubMed
Augustin, H. G. (2001). Tubes, branches, and pillars: the many ways of forming a new vasculature. Circ. Res. 89, 645–7.Google ScholarPubMed
Aulehla, A., Wehrle, C., Brand-Saberi, B., Kemler, R., Gossler, A., Kanzler, B., and Herrmann, B. G. (2003). Wnt3a plays a major role in the segmentation clock controlling somitogenesis. Dev. Cell 4, 395–406.CrossRefGoogle Scholar
Aylsworth, A. S. (2001). Clinical aspects of defects in the determination of laterality. Am. J. Med. Genet. 101, 345–55.CrossRefGoogle ScholarPubMed
Bachvarova, R. (1985). Gene expression during oogenesis and oocyte development in mammals. Dev. Biol. 1, 453–524.Google ScholarPubMed
Ball, W. D. (1974). Development of the rat salivary glands. 3. Mesenchymal specificity in the morphogenesis of the embryonic submaxillary and sublingual glands of the rat. J. Exp. Zool. 188, 277–88.CrossRefGoogle ScholarPubMed
Ballaro, B., and Reas, P. G. (2000). Chemical and mechanical waves on the cortex of fertilized egg cells: a bioexcitability effect. Rev. Biol. 93, 83–101.Google ScholarPubMed
Baoal, D. (2002). Mechanics of the Cell. Cambridge University Press, Cambridge.Google Scholar
Barabási, A.-L. (2002). Linked: The New Science of Networks. Perseus Publications, Cambridge, MA.Google Scholar
Bard, J. B. (1999). A bioinformatics approach to investigating developmental pathways in the kidney and other tissues. Int. J. Dev. Biol. 43, 397–403.Google ScholarPubMed
Barkai, N., and Leibler, S. (2000). Circadian clocks limited by noise. Nature 403, 267–8.CrossRefGoogle ScholarPubMed
Basu, S., Gerchmann, Y., Collins, C. H., Arnold, F. H., Weiss, R. (2005). A synthetic multicellular system for programmed pattern formation. Nature 434, 1130–4.CrossRefGoogle ScholarPubMed
Bateman, E. (1998). Autoregulation of eukaryotic transcription factors. Prog. Nucleic Acid Res. Mol. Biol. 60, 133–68.CrossRefGoogle ScholarPubMed
Bateson, W. (1894). Materials for the Study of Variation. Macmillan. London.Google Scholar
Baumgartner, W., Hinterdorfer, P., Ness, W., et al. (2000). Cadherin interaction probed by atomic force microscopy. Proc. Nat. Acad. Sci. USA 97, 4005–10.CrossRefGoogle ScholarPubMed
Bausch, A. R., Moller, W., and Sackmann, E. (1999). Measurement of local viscoelasticity and forces in living cells by magnetic tweezers. Biophys. J. 76, 573–9.CrossRefGoogle ScholarPubMed
Becker, M., Baumann, C., John, S., et al. (2002). Dynamic behavior of transcription factors on a natural promoter in living cells. EMBO Rep. 3, 1188–94.CrossRefGoogle ScholarPubMed
Beddington, R. S., and Robertson, E. J. (1999). Axis development and early asymmetry in mammals. Cell 96, 195–209.CrossRefGoogle ScholarPubMed
Bejsovec, A., and Wieschaus, E. (1993). Segment polarity gene interactions modulate epidermal patterning in Drosophila embryos. Development 119, 501–17.Google ScholarPubMed
Bell, G. I. (1978). Models for the specific adhesion of cells to cells. Science 200, 618–27.CrossRefGoogle ScholarPubMed
Beloussov, L. (1998). The Dynamic Architecture of a Developing Organism.Kluwer Academic Publishers, Dordrecht.CrossRefGoogle Scholar
Ben-Avraham, D., and Havlin, S. (2000). Diffusion and Reactions in Fractals and Disordered Systems. Cambridge University Press, Cambridge, New York.CrossRefGoogle Scholar
Benink, H. A., Mandato, C. A., and Bement, W. M. (2000). Analysis of cortical flow models in vivo. Mol. Biol. Cell 11, 2553–63.CrossRefGoogle ScholarPubMed
Berg, H. C. (1993). Random Walks in Biology. Princeton University Press, Princeton.Google Scholar
Berridge, M. J., Lipp, P., and Bootman, M. D. (2000). The versatility and universality of calcium signalling. Nat. Rev. Mol. Cell Biol. 1, 11–21.CrossRefGoogle ScholarPubMed
Berry, L. D., and Gould, K. L. (1996). Regulation of Cdc2 activity by phosphorylation at T14/Y15. Prog. Cell. Cycle Res. 2, 99–105.CrossRefGoogle ScholarPubMed
Bertrand, N., Castro, D. S., and Guillemot, F. (2002). Proneural genes and the specification of neural cell types. Nat. Rev. Neurosci. 3, 517–30.CrossRefGoogle ScholarPubMed
Bevilacqua, M., Butcher, E., Furie, B., et al. (1991). Selectins: a family of adhesion receptors. Cell 67, 233.CrossRefGoogle ScholarPubMed
Beysens, D. A., Forgacs, G., and Glazier, J. A. (2000). Cell sorting is analogous to phase ordering in fluids. Proc. Nat. Acad. Sci. USA 97, 9467–71.CrossRefGoogle ScholarPubMed
Bhalla, U. S., and Iyengar, R. (1999). Emergent properties of networks of biological signaling pathways. Science 283, 381–7.CrossRefGoogle ScholarPubMed
Bissell, M. J., and Barcellos-Hoff, M. H. (1987). The influence of extracellular matrix on gene expression: is structure the message?J. Cell Sci. Suppl. 8, 327–43.CrossRefGoogle ScholarPubMed
Blair, S. S. (2003). Developmental biology: boundary lines. Nature 424, 379–81.CrossRefGoogle ScholarPubMed
Boal, D. H. (2002). Mechanics of the Cell. Cambridge University Press, Cambridge, New York.Google Scholar
Boggon, T. J., Murray, J., Chappuis-Flament, S., et al. (2002). C-cadherin ectodomain structure and implications for cell adhesion mechanisms. Science 296, 1308–13.CrossRefGoogle ScholarPubMed
Boissonade, J., Dulos, E., and DeKepper, P. (1994). Turing patterns: from myth to reality. In Chemical Waves and Patterns (Kapral, R. and Showalter, K., eds.), pp. 221–68. Kluwer, Boston.Google Scholar
Bolouri, H., and Davidson, E. H. (2003). Transcriptional regulatory cascades in development: initial rates, not steady state, determine network kinetics. Proc. Nat. Acad. Sci. USA 100, 9371–6.CrossRefGoogle Scholar
Bonner, J. T. (1998). The origins of multicellularity. Integrative Biology 1, 27–36.3.0.CO;2-6>CrossRefGoogle Scholar
Boring, L. (1989). Cell–cell interactions determine the dorsoventral axis in embryos of an equally cleaving opisthobranch mollusc. Dev. Biol. 136, 239–53.CrossRefGoogle ScholarPubMed
Borisuk, M. T., and Tyson, J. J. (1998). Bifurcation analysis of a model of mitotic control in frog eggs. J. Theor. Biol. 195, 69–85.CrossRefGoogle ScholarPubMed
Borkhvardt, V. G. (2000). The growth and form development of the limb buds in vertebrate animals. Ontogenez 31, 192–200.Google ScholarPubMed
Bouligand, Y. (1972). Twisted fibrous arrangements in biological materials and cholesteric mesophases. Tissue Cell 4, 189–217.CrossRefGoogle ScholarPubMed
Braat, A. K., Zandbergen, T., Water, S., Goos, H. J., and Zivkovic, D. (1999). Characterization of zebrafish primordial germ cells: morphology and early distribution of vasa RNA. Dev. Dyn. 216, 153–67.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Braga, V. M. (2002). Cell–cell adhesion and signalling. Curr. Opin. Cell. Biol. 14, 546–56.CrossRefGoogle ScholarPubMed
Branford, W. W., and Yost, H. J. (2002). Lefty-dependent inhibition of Nodal- and Wnt-responsive organizer gene expression is essential for normal gastrulation. Curr. Biol. 12, 2136–41.CrossRefGoogle ScholarPubMed
Brasier, M., and Antcliffe, J. (2004). Paleobiology. Decoding the Ediacaran enigma. Science 305, 1115–7.CrossRefGoogle ScholarPubMed
Breckenridge, R. A., Mohun, T. J., and Amaya, E. (2001). A role for BMP signalling in heart looping morphogenesis in Xenopus. Dev. Biol. 232, 191–203.CrossRefGoogle ScholarPubMed
Brinker, C. J., and Scherer, G. W. (1990). Sol–Gel Science. Academic Press, New York.Google Scholar
Brock, H. W., and Fisher, C. L. (2005). Maintenance of gene expression patterns. Dev. Dyn. 232, 633–55.CrossRefGoogle ScholarPubMed
Bronner-Fraser, M. (1982). Distribution of latex beads and retinal pigment epithelial cells along the ventral neural crest pathways. Dev. Biol. 91, 50–63.CrossRefGoogle Scholar
Bronner-Fraser, M. (1984). Latex beads as probes of a neural crest pathway: effects of laminin, collagen, and surface charge on bead translocation. J. Cell Biol. 98, 1947–60.CrossRefGoogle ScholarPubMed
Bronner-Fraser, M. (1985). Effects of different fragments of the fibronectin molecule on latex bead translocation along neural crest migratory pathways. Dev. Biol. 108, 131–45.CrossRefGoogle ScholarPubMed
Bronner-Fraser, M., Wolf, J. J., and Murray, B. A. (1992). Effects of antibodies against N-cadherin and N-CAM on the cranial neural crest and neural tube. Dev. Biol. 153, 291–301.CrossRef
Brown, S. J., Hilgenfeld, R. B., and Denell, R. E. (1994a). The beetle Tribolium castaneum has a fushi tarazu homolog expressed in stripes during segmentation. Proc. Nat. Acad. Sci. USA 91, 12922–6.CrossRefGoogle Scholar
Brown, S. J., Patel, N. H., and Denell, R. E. (1994b). Embryonic expression of the single Tribolium engrailed homolog. Dev. Genet. 15, 7–18.CrossRefGoogle Scholar
Browne, E. N. (1909). The production of new hydranths in Hydra by insertion of small grafts. J. Exp. Zool. 7, 1–23.CrossRefGoogle Scholar
Brunet, J. F., and Ghysen, A. (1999). Deconstructing cell determination: proneural genes and neuronal identity. Bioessays 21, 313–8.3.0.CO;2-C>CrossRefGoogle ScholarPubMed
Bryant, P. J. (1999). Filopodia: fickle fingers of cell fate?Curr. Biol. 9, R655–7.CrossRefGoogle ScholarPubMed
Buckley, C. D., Rainger, G. E., Bradfield, P. F., Nash, G. B., and Simmons, D. L. (1998). Cell adhesion: more than just glue. Mol. Membr. Biol. 15, 167–76.CrossRefGoogle ScholarPubMed
Bugrim, A., Fontanilla, R., Eutenier, B. B., Keizer, J., and Nuccitelli, R. (2003). Sperm initiate a Ca2+ wave in frog eggs that is more similar to Ca2+ waves initiated by IP3 than by Ca2+. Biophys. J. 84, 1580–90.CrossRefGoogle ScholarPubMed
Bugrim, A. E., Zhabotinsky, A. M., and Epstein, I. R. (1997). Calcium waves in a model with a random spatially discrete distribution of Ca2+ release sites. Biophys. J. 73, 2897–906.CrossRefGoogle Scholar
Callamaras, N., Marchant, J. S., Sun, X. P., and Parker, I. (1998). Activation and co-ordination of InsP3-mediated elementary Ca2+ events during global Ca2+ signals in Xenopus oocytes. J. Physiol. 509, 81–91.CrossRefGoogle ScholarPubMed
Campochiaro, P. A. (2000). Retinal and choroidal neovascularization. J. Cell Physiol. 184, 301–10.3.0.CO;2-H>CrossRefGoogle ScholarPubMed
Canman, J. C., and Bement, W. M. (1997). Microtubules suppress actomyosin-based cortical flow in Xenopus oocytes. J. Cell Sci. 110, 1907–17.Google ScholarPubMed
Capco, D. G., and McGaughey, R. W. (1986). Cytoskeletal reorganization during early mammalian development: analysis using embedment-free sections. Dev. Biol. 115, 446–58.CrossRefGoogle ScholarPubMed
Carnac, G., and Gurdon, J. B. (1997). The community effect in Xenopus myogenesis is promoted by dorsalizing factors. Int. J. Dev. Biol. 41, 521–4.Google ScholarPubMed
Castets, V., Dulos, E., Boissonade, J., and DeKepper, P. (1990). Experimental evidence of a sustained standing Turing-type nonequilibrium chemical pattern. Phys. Rev. Lett. 64, 2953–6.CrossRefGoogle ScholarPubMed
Cayan, S., Conaghan, J., Schriock, E. D., Ryan, I. P., Black, L. D., and Turek, P. J. (2001). Birth after intracytoplasmic sperm injection with use of testicular sperm from men with Kartagener/immotile cilia syndrome. Fertil. Steril. 76, 612–4.CrossRefGoogle ScholarPubMed
Chambon, F., and Winter, H. H. (1987). Linear viscoelasticity at the gel point of a crosslinking PDMS with imbalanced stoichiometry. J. Rheol. 31, 683–97.CrossRefGoogle Scholar
Chan, A. P., and Etkin, L. D. (2001). Patterning and lineage specification in the amphibian embryo. Curr. Top. Dev. Biol. 51, 1–67.CrossRefGoogle ScholarPubMed
Chaturvedi, R., Huang, C., Kazmierczak, B., Schneider, T., Izaguirre, J. A., Glimm, T., Hentschel, H. G. E., Newman, S. A., Glazier, J. A., and Alber, M. (2005). On multiscale approaches to three-dimensional modeling of morphogenesis. J. Roy. Soc. London Interface 2, 237–53.CrossRefGoogle Scholar
Cheer, A., Vincent, J. P., Nuccitelli, R., and Oster, G. (1987). Cortical activity in vertebrate eggs. I: The activation waves. J. Theor. Biol. 124, 377–404.CrossRefGoogle ScholarPubMed
Chen, J. N., Eeden, F. J., Warren, K. S., et al. (1997). Left–right pattern of cardiac BMP4 may drive asymmetry of the heart in zebrafish. Development 124, 4373–82.Google ScholarPubMed
Chen, J.-Y., Bottjer, D. J., Oliveri, P., et al. (2004). Small bilaterian fossils from 40 to 55 million years before the Cambrian. Science 305, 218–22.CrossRefGoogle ScholarPubMed
Chen, Y., and Schier, A. F. (2002). Lefty proteins are long-range inhibitors of squint-mediated nodal signaling. Curr. Biol. 12, 2124–8.CrossRefGoogle ScholarPubMed
Cheng, A., Ross, K. E., Kaldis, P., and Solomon, M. J. (1999). Dephosphorylation of cyclin-dependent kinases by type 2C protein phosphatases. Genes Dev. 13, 2946–57.CrossRefGoogle ScholarPubMed
Christen, B., and Slack, J. (1999). Spatial response to fibroblast growth factor signalling in Xenopus embryos. Development 126, 119–25.Google ScholarPubMed
Chuong, C. M. (1993). The making of a feather: homeoproteins, retinoids and adhesion molecules. BioEssays 15, 513–21.CrossRefGoogle ScholarPubMed
Cickovski, T., Huang, C., Chaturvedi, R., et al. (2005). A framework for three-dimensional simulation of morphogenesis. IEEE/ACM Trans. Computat. Biol. Bioinformatics, in press.CrossRefGoogle ScholarPubMed
Cinquin, O., and Demongeof, J. (2005). High-dimensional switches and the modelling of cellular differentiation. J. Theor. Biol. 233, 391–411.CrossRefGoogle ScholarPubMed
Clements, D., Friday, R. V., and Woodland, H. R. (1999). Mode of action of VegT in mesoderm and endoderm formation. Development 126, 4903–11.Google ScholarPubMed
Clerk, J. P., Giraud, G., Laugier, J. M., and Luck, J. M. (1990). The AC electrical conductance of binary disordered systems, percolation custers, fractals and related models. Adv. Phys. 39, 191–309.CrossRefGoogle Scholar
Cline, C. A., Schatten, H., Balczon, R., and Schatten, G. (1983). Actin-mediated surface motility during sea urchin fertilization. Cell Motil. 3, 513–24.CrossRefGoogle ScholarPubMed
Clyde, D. E., Corado, M. S., 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
Colas, J. F., and Schoenwolf, G. C. (2001). Towards a cellular and molecular understanding of neurulation. Dev. Dyn. 221, 117–45.CrossRefGoogle ScholarPubMed
Collier, J. R., Monk, N. A., Maini, P. K., and Lewis, J. H. (1996). Pattern formation by lateral inhibition with feedback: a mathematical model of Delta–Notch intercellular signalling. J. Theor. Biol. 183, 429–46.CrossRefGoogle ScholarPubMed
Comper, W. D. (1996). Extracellular Matrix. Vol. I. Tissue Function; Vol. II. Molecular Components and Interactions. Harwood Academic Publishers, Amsterdam.Google Scholar
Comper, W. D., Pratt, L., Handley, C. J., and Harper, G. S. (1987). Cell transport in model extracellular matrices. Arch. Biochem. Biophys. 252, 60–70.CrossRefGoogle ScholarPubMed
Conway Morris, S. (2003). The Cambrian “explosion” of metazoans. In Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology (Müller, G. B. and Newman, S. A., eds.), pp. 13–32. MIT Press, Cambridge, MA.Google Scholar
Cooke, J., Nowak, M. A., Boerlijst, M., and Maynard-Smith, J. (1997). Evolutionary origins and maintenance of redundant gene expression during metazoan development. Trends Genet. 13, 360–4.CrossRefGoogle ScholarPubMed
Cooke, J., and Zeeman, E. C. (1976). A clock and wavefront model for control of the number of repeated structures during animal morphogenesis. J. Theor. Biol. 58, 455–76.CrossRefGoogle ScholarPubMed
Cormack, D. H. (1987). Ham's Histology, ninth edn. Lippincott, Philadelphia.Google Scholar
Cornish-Bowden, A. (1995). Fundamentals of Enzyme Kinetics. Ashgate Publ. Co., London.Google Scholar
Coulombe, J. N., and Bronner-Fraser, M. (1984). Translocation of latex beads after laser ablation of the avian neural crest. Dev. Biol. 106, 121–34.CrossRefGoogle ScholarPubMed
Crawford, K., and Stocum, D. L. (1988). Retinoic acid coordinately proximalizes regenerate pattern and blastema differential affinity in axolotl limbs. Development 102, 687–98.Google ScholarPubMed
Crick, F. H. C. (1970). Diffusion in embryogenesis. Nature 225, 420–2.CrossRefGoogle ScholarPubMed
Crick, F. H. C., and Lawrence, P. A. (1975). Compartments and polyclones in insect development. Science 189, 340–7.CrossRefGoogle ScholarPubMed
Cross, N. L., and Elinson, R. P. (1980). A fast block to polyspermy in frogs mediated by changes in the membrane potential. Dev. Biol. 75, 187–98.CrossRef
Cunliffe, V. T. (2003). Memory by modification: the influence of chromatin structure on gene expression during vertebrate development. Gene 305, 141–50.CrossRefGoogle ScholarPubMed
Czirok, A., Rupp, P. A., Rongish, B. J., and Little, C. D. (2002). Multi-field 3D scanning light microscopy of early embryogenesis. J. Microsc. 206, 209–17.CrossRefGoogle ScholarPubMed
Dale, J. K., Maroto, M., Dequeant, M. L., Malapert, P., McGrew, M., and Pourquie, O. (2003). Periodic notch inhibition by Lunatic fringe underlies the chick segmentation clock. Nature 421, 275–8.CrossRefGoogle ScholarPubMed
Danos, M. C., and Yost, H. J. (1996). Role of notochord in specification of cardiac left–right orientation in zebrafish and Xenopus. Dev. Biol. 177, 96–103.CrossRefGoogle ScholarPubMed
Dathe, V., Gamel, A., Manner, J., Brand-Saberi, B., and Christ, B. (2002). Morphological left–right asymmetry of Hensen's node precedes the asymmetric expression of Shh and Fgf8 in the chick embryo. Anat. Embryol. (Berlin) 205, 343–54.CrossRefGoogle ScholarPubMed
Davidson, E. H. (2001). Genomic Regulatory Systems: Development and Evolution. Academic Press, San Diego.Google Scholar
Davidson, E. H., Rast, J. P., Oliveri, P., et al. (2002). A genomic regulatory network for development. Science 295, 1669–78.CrossRefGoogle Scholar
Davidson, L. A., Koehl, M. A., Keller, R., and Oster, G. F. (1995). How do sea urchins invaginate? Using biomechanics to distinguish between mechanisms of primary invagination. Development 121, 2005–18.Google ScholarPubMed
Davidson, L. A., Oster, G. F., Keller, R. E., and Koehl, M. A. (1999). Measurements of mechanical properties of the blastula wall reveal which hypothesized mechanisms of primary invagination are physically plausible in the sea urchin Strongylocentrotus purpuratus. Dev. Biol. 209, 221–38.CrossRefGoogle ScholarPubMed
Dawes, R., Dawson, I., Falciani, F., Tear, G., and Akam, M. (1994). Dax, a locust Hox gene related to fushi-tarazu but showing no pair-rule expression. Development 120, 1561–72.Google ScholarPubMed
Dawson, S. P., Keizer, J., and Pearson, J. E. (1999). Fire–diffuse–fire model of dynamics of intracellular calcium waves. Proc. Nat. Acad. Sci. USA 96, 6060–3.CrossRefGoogle ScholarPubMed
Felici, M. (2000). Regulation of primordial germ cell development in the mouse. Int. J. Dev. Biol. 44, 575–80.Google ScholarPubMed
Gennes, P. G. (1976a). Critical dimensionality for a special percolation problem (relevant to the gelation in polymers). J. Physique (Paris) 30, 1049–54.Google Scholar
Gennes, P. G. (1976b). On the relation between percolation and the elasticity of gels. J. Physique (Paris) 37, L1–2.CrossRefGoogle Scholar
Gennes, P. G. (1992). Soft matter. Science 256, 495–7.CrossRefGoogle ScholarPubMed
Gennes, P. G., and Prost, J. (1993). The Physics of Liquid Crystals. Clarendon Press, Oxford.Google Scholar
Deguchi, R., Shirakawa, H., Oda, S., Mohri, T., and Miyazaki, S. (2000). Spatiotemporal analysis of Ca2+ waves in relation to the sperm entry site and animal–vegetal axis during Ca2+ oscillations in fertilized mouse eggs. Dev. Biol. 218, 299–313.CrossRefGoogle Scholar
DeMarais, A. A., and Moon, R. T. (1992). The armadillo homologs beta-catenin and plakoglobin are differentially expressed during early development of Xenopus laevis. Dev. Biol. 153, 337–46.CrossRefGoogle ScholarPubMed
Smedt, V., Poulhe, R., Cayla, X., et al. (2002). Thr-161 phosphorylation of monomeric Cdc2. Regulation by protein phosphatase 2C in Xenopus oocytes. J. Biol. Chem. 277, 28592–600.CrossRefGoogle ScholarPubMed
Dembo, M., Glushko, V., Aberlin, M. E., and Sonenberg, M. (1979). A method for measuring membrane microviscosity using pyrene excimer formation. Application to human erythrocyte ghosts. Biochim. Biophys. Acta 552, 201–11.CrossRefGoogle ScholarPubMed
Derganc, J., Bozic, B., Svetina, S., and Zeks, B. (2000). Stability analysis of micropipette aspiration of neutrophils. Biophys. J. 79, 153–62.CrossRefGoogle ScholarPubMed
Dewar, H., Tanaka, K., Nasmyth, K., and Tanaka, T. U. (2004). Tension between two kinetochores suffices for their bi-orientation on the mitotic spindle. Nature 428, 93–7.CrossRefGoogle ScholarPubMed
Dickinson, R. B., and Tranquillo, R. T. (1993). A stochastic model for adhesion-mediated cell random motility and haptotaxis. J. Math. Biol. 31, 563–600.CrossRefGoogle ScholarPubMed
Djabourov, M., Leblond, J., and Papon, P. (1988). Gelation of aqueous gelatin solutions. II. Rheology of the sol–gel transition. J. Phys. (Paris) 49, 333–343.CrossRefGoogle Scholar
Djabourov, M., Lechaire, J. P., and Gaill, F. (1993). Structure and rheology of gelatin and collagen gels. Biorheology 30, 191–205.CrossRefGoogle ScholarPubMed
Doedel, E. J., and Wang, X. J. (1995). AUTO94: Software for continuation and bifurcation problems in ordinary differential equations. Center for Research on Parallel Computing, California Institute of Technology, Pasadena, CA.Google Scholar
Dolmetsch, R. E., Xu, K., and Lewis, R. S. (1998). Calcium oscillations increase the efficiency and specificity of gene expression. Nature 392, 933–6.CrossRefGoogle ScholarPubMed
Dosch, R., Gawantka, V., Delius, H., Blumenstock, C., and Niehrs, C. (1997). Bmp-4 acts as a morphogen in dorsoventral mesoderm patterning in Xenopus. Development 124, 2325–34.Google ScholarPubMed
Downie, S. A., and Newman, S. A. (1994). Morphogenetic differences between fore and hind limb precartilage mesenchyme: relation to mechanisms of skeletal pattern formation. Dev. Biol. 162, 195–208.CrossRefGoogle ScholarPubMed
Downie, S. A., and Newman, S. A. (1995). Different roles for fibronectin in the generation of fore and hind limb precartilage condensations. Dev. Biol. 172, 519–30.CrossRefGoogle ScholarPubMed
Drasdo, D., and Forgacs, G. (2000). Modeling the interplay of generic and genetic mechanisms in cleavage, blastulation, and gastrulation. Dev. Dyn. 219, 182–91.3.3.CO;2-1>CrossRefGoogle ScholarPubMed
Drasdo, D., Kree, R., and McCaskill, J. S. (1995). Monte Carlo approach to tissue-cell populations. Phys. Rev. E Statistical Physics, Plasmas, Fluids, and Related Interdisciplinary Topics 52, 6635–57.Google ScholarPubMed
Duband, J. L., Monier, F., Delannet, M., and Newgreen, D. (1995). Epithelium–mesenchyme transition during neural crest development. Acta Anat. (Basel) 154, 63–78.CrossRefGoogle ScholarPubMed
Dubrulle, J., McGrew, M. J., and Pourquié, O. (2001). FGF signaling controls somite boundary position and regulates segmentation clock control of spatiotemporal Hox gene activation. Cell 106, 219–32.CrossRefGoogle ScholarPubMed
Ducibella, T., Huneau, D., Angelichio, E., Xu, Z., Schultz, R. M., Kopf, G. S., Fissore, R., Madoux, S., and Ozil, J. P. (2002). Egg-to-embryo transition is driven by differential responses to Ca2+ oscillation number. Dev. Biol. 250, 280–91.CrossRefGoogle Scholar
Duguay, D., Foty, R. A., and Steinberg, M. S. (2003). Cadherin-mediated cell adhesion and tissue segregation: qualitative and quantitative determinants. Dev. Biol. 253, 309–23.CrossRefGoogle ScholarPubMed
Dumollard, R., Carroll, J., Dupont, G., and Sardet, C. (2002). Calcium wave pacemakers in eggs. J. Cell Sci. 115, 3557–64.CrossRefGoogle ScholarPubMed
Dumont, J. N., and Brummett, A. R. (1985). Egg envelopes in vertebrates. Dev. Biol. 1, 235–88.Google ScholarPubMed
Durand, D., Delsanti, M., Adam, M., and Luck, J. M. (1987). Frequency dependence of viscoelastic properties of branched polymers near gelation threshold. Europhys. Lett. 3, 297–301.CrossRefGoogle Scholar
Edelman, G. M. (1992). Morphoregulation. Dev. Dyn. 193, 2–10.CrossRefGoogle ScholarPubMed
Eichmann, A., Pardanaud, L., Yuan, L., and Moyon, D. (2002). Vasculogenesis and the search for the hemangioblast. J. Hematother. Stem Cell Res. 11, 207–14.CrossRefGoogle ScholarPubMed
Eidne, K. A., Zabavnik, J., Allan, W. T., Trewavas, A. J., Read, N. D., and Anderson, L. (1994). Calcium waves and dynamics visualized by confocal microscopy in Xenopus oocytes expressing cloned TRH receptors. J. Neuroendocrinol. 6, 173–8.CrossRefGoogle ScholarPubMed
Ekblom, P. (1992). Renal Development. Raven Press, New York.Google Scholar
Elinson, R. P., and Rowning, B. (1988). A transient array of parallel microtubules in frog eggs: potential tracks for a cytoplasmic rotation that specifies the dorso-ventral axis. Dev. Biol. 128, 185–97.CrossRefGoogle ScholarPubMed
Ellis, R. J. (2001). Macromolecular crowding: an important but neglected aspect of the intracellular environment. Curr. Opin. Struct. Biol. 11, 114–9.CrossRefGoogle ScholarPubMed
Ellis, R. J., and Minton, A. P. (2003). Cell biology: join the crowd. Nature 425, 27–8.CrossRefGoogle Scholar
Elowitz, M. B., and Leibler, S. (2000). A synthetic oscillatory network of transcriptional regulators. Nature 403, 335–8.CrossRefGoogle ScholarPubMed
Entchev, E. V., Schwabedissen, A., and Gonzalez-Gaitan, M. (2000). Gradient formation of the TGF-β homolog Dpp. Cell 103, 981–91.CrossRefGoogle ScholarPubMed
Erickson, C. A. (1985). Control of neural crest cell dispersion in the trunk of the avian embryo. Dev. Biol. 111, 138–57.CrossRefGoogle ScholarPubMed
Erickson, C. A. (1988). Control of pathfinding by the avian trunk neural crest. Development 103, 63–80.Google ScholarPubMed
Erickson, C. A., and Isseroff, R. R. (1989). Plasminogen activator activity is associated with neural crest cell motility in tissue culture. J. Exp. Zool. 251, 123–33.CrossRefGoogle ScholarPubMed
Erickson, C. A., and Perris, R. (1993). The role of cell–cell and cell–matrix interactions in the morphogenesis of the neural crest. Dev. Biol. 159, 60–74.CrossRefGoogle ScholarPubMed
Eshkind, L., Tian, Q., Schmidt, A., et al. (2002). Loss of desmoglein 2 suggests essential functions for early embryonic development and proliferation of embryonal stem cells of mice and models: improved animal models for biomedical research. Synaptic vesicle alterations in rod photoreceptors of synaptophysin-deficient mice. Eur. J. Cell. Biol. 81, 592–8.CrossRefGoogle Scholar
Espeseth, A., Johnson, E., and Kintner, C. (1995). Xenopus F-cadherin, a novel member of the cadherin family of cell adhesion molecules, is expressed at boundaries in the neural tube. Mol. Cell. Neurosci. 6, 199–211.CrossRefGoogle ScholarPubMed
Essner, J. J., Vogan, K. J., Wagner, M. K., Tabin, C. J., Yost, H. J., and Brueckner, M. (2002). Conserved function for embryonic nodal cilia. Nature 418, 37–8.CrossRefGoogle ScholarPubMed
Ettensohn, C. A. (1999). Cell movements in the sea urchin embryo. Curr. Opin. Genet. Dev. 9, 461–5.CrossRefGoogle ScholarPubMed
Ettinger, L., and Doljanski, F. (1992). On the generation of form by the continuous interactions between cells and their extracellular matrix. Biol. Rev. Camb. Philos. Soc. 67, 459–89.CrossRefGoogle ScholarPubMed
Evans, E., and Yeoung, A. (1989). Apparent viscosity and cortical tension of blood granulocytes determined by micropipet aspiration. Biophys. J. 56, 151–60.CrossRefGoogle ScholarPubMed
Fagotto, F., Guger, K., and Gumbiner, B. M. (1997). Induction of the primary dorsalizing center in Xenopus by the Wnt/GSK/beta-catenin signaling pathway, but not by Vg1, Activin or Noggin. Development 124, 453–60.Google ScholarPubMed
Ferrell, J. E. Jr, Wu, M., Gerhart, J. C., and Martin, G. S. (1991). Cell cycle tyrosine phosphorylation of p34cdc2 and a microtubule-associated protein kinase homolog in Xenopus oocytes and eggs. Mol. Cell. Biol. 11, 1965–71.CrossRefGoogle Scholar
Fitch, J., Fini, M. E., Beebe, D. C., and Linsenmayer, T. F. (1998). Collagen type IX and developmentally regulated swelling of the avian primary corneal stroma. Dev. Dyn. 212, 27–37.3.0.CO;2-4>CrossRefGoogle ScholarPubMed
Fleming, T. P., and Goodall, H. (1986). Endocytic traffic in trophectoderm and polarised blastomeres of the mouse preimplantation embryo. Anat. Rec. 216, 490–503.CrossRefGoogle ScholarPubMed
Folkman, J. (2003). Angiogenesis and proteins of the hemostatic system. J. Thromb. Haemost. 1, 1681–2.CrossRefGoogle ScholarPubMed
Folkman, J., and Moscona, A. (1978). Role of cell shape in growth control. Nature 273, 345–9.CrossRefGoogle ScholarPubMed
Fontanilla, R. A., and Nuccitelli, R. (1998). Characterization of the sperm-induced calcium wave in Xenopus eggs using confocal microscopy. Biophys. J. 75, 2079–87.CrossRefGoogle ScholarPubMed
Forgacs, G. (1995). On the possible role of cytoskeletal filamentous networks in intracellular signalling: an approach based on percolation. J. Cell Sci. 108, 2131–2143.Google Scholar
Forgacs, G., and Foty, R. A. (2004). Biological implications of tissue viscoelasticity. In Function and Regulation of Cellular Systems: Experiments and Models (Deutsch, A., Falke, M., Howard, J., and Zimmerman, W., eds.), pp. 269–77. Biskhauser Basel.CrossRefGoogle Scholar
Forgacs, G., and Newman, S. A. (1994). Phase transitions, interfaces, and morphogenesis in a network of protein fibers. Int. Rev. Cytol. 150, 139–48.CrossRefGoogle Scholar
Forgacs, G., Jaikaria, N. S., Frisch, H. L., and Newman, S. A. (1989). Wetting, percolation and morphogenesis in a model tissue system. J. Theor. Biol. 140, 417–430.CrossRefGoogle Scholar
Forgacs, G., Newman, S. A., Obukhov, S. P., and Birk, D. E. (1991). Phase transition and morphogenesis in a model biological system. Phys. Rev. Lett. 67, 2399–402.CrossRefGoogle Scholar
Forgacs, G., Foty, R. A., Shafrir, Y., and Steinberg, M. S. (1998). Viscoelastic properties of living embryonic tissues: a quantitative study. Biophys. J. 74, 2227–34.CrossRefGoogle ScholarPubMed
Forgacs, G., Newman, S. A., Hinner, B., Maier, C. W., and Sackmann, E. (2003). Assembly of collagen matrices as a phase transition revealed by structural and rheologic studies. Biophys. J. 84, 1272–80.CrossRefGoogle ScholarPubMed
Forgacs, G., Yook, S. H., Janmey, P. A., Jeong, H., and Burd, C. G. (2004). Role of the cytoskeleton in signaling networks. J. Cell. Sci. 117, 2769–75.CrossRefGoogle ScholarPubMed
Foty, R. A., and Steinberg, M. S. (1997). Measurement of tumor cell cohesion and suppression of invasion by E- or P-cadherin. Cancer Res. 57, 5033–6.Google ScholarPubMed
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., Forgacs, G., Pfleger, C. M., and Steinberg, M. S. (1994). Liquid properties of embryonic tissues: measurement of interfacial tensions. Phys. Rev. Lett. 72, 2298–301.CrossRefGoogle ScholarPubMed
Foty, R. A., Pfleger, C. M., Forgacs, G., and Steinberg, M. S. (1996). Surface tensions of embryonic tissues predict their mutual envelopment behavior. Development 122, 1611–20.Google ScholarPubMed
Frasch, M., and Levine, M. (1987). Complementary patterns of even-skipped and fushi tarazu expression involve their differential regulation by a common set of segmentation genes in Drosophila. Genes Dev. 1, 981–95.CrossRefGoogle ScholarPubMed
Freeman, M. (2002). Morphogen gradients, in theory. Dev. Cell 2, 689–90.CrossRefGoogle ScholarPubMed
Frenz, D. A., Akiyama, S. K., Paulsen, D. F., and Newman, S. A. (1989a). Latex beads as probes of cell surface–extracellular matrix interactions during chondrogenesis: evidence for a role for amino-terminal heparin-binding domain of fibronectin. Dev. Biol. 136, 87–96.CrossRefGoogle Scholar
Frenz, D. A., Jaikaria, N. S., and Newman, S. A. (1989b). The mechanism of precartilage mesenchymal condensation: a major role for interaction of the cell surface with the amino-terminal heparin-binding domain of fibronectin. Dev. Biol. 136, 97–103.CrossRefGoogle Scholar
Freyman, T. M., Yannas, I. V., Yokoo, R., and Gibson, L. J. (2001). Fibroblast contraction of a collagen-GAG matrix. Biomaterials 22, 2883–91.CrossRefGoogle ScholarPubMed
Friedlander, D. R., Mege, R.-M., Cunningham, B. A., and Edelman, G. M. (1989). Cell sorting-out is modulated by both the specificity and amount of different cell adhesion molecules (CAMs) expressed on cell surfaces. Proc. Nat. Acad. Sci. USA 86, 7043–7047.CrossRefGoogle ScholarPubMed
Fristrom, D., and Chihara, C. (1978). The mechanism of evagination of imaginal discs of Drosophila melanogaster. V. Evagination of disc fragments. Dev. Biol. 66, 564–570.CrossRefGoogle ScholarPubMed
Fujiwara, T., Ritchie, K., Murakoshi, H., Jacobson, K., and Kusumi, A. (2002). Phospholipids undergo hop diffusion in compartmentalized cell membrane. J. Cell Biol. 157, 1071–81.CrossRefGoogle ScholarPubMed
Fung, Y. C. (1993). Biomechanics: Mechanical Properties of Living Tissues. Springer-Verlag, New York.CrossRefGoogle Scholar
Furusawa, C., and Kaneko, K. (1998). Emergence of rules in cell society: differentiation, hierarchy, and stability. Bull. Math. Biol. 60, 659–87.CrossRefGoogle ScholarPubMed
Furusawa, C., and Kaneko, K. (2001). Theory of robustness of irreversible differentiation in a stem cell system: chaos hypothesis. J. Theor. Biol. 209, 395–416.CrossRefGoogle Scholar
Gaill, F., Lechaire, J. P., and Denefle, J. P. (1991). Fibrillar pattern of self-assembled and cell-assembled collagen: resemblance and analogy. Biol. Cell 72, 149–58.CrossRefGoogle ScholarPubMed
Gamba, A., Ambrosi, D., Coniglio, A., et al. (2003). Percolation, morphogenesis, and Burgers dynamics in blood vessels formation. Phys. Rev. Lett. 90, 118 101.CrossRefGoogle ScholarPubMed
Garcia-Bellido, A. (1975). Genetic control of wing disc development in Drosophila. Ciba Found. Sym. 29, 169–78.Google Scholar
Garcia-Bellido, A., Ripoll, P., and Morata, G. (1976). Developmental compartmentalization in the dorsal mesothoracic disc of Drosophila. Dev. Biol. 48, 132–47.CrossRefGoogle ScholarPubMed
Garcia-Perez, A. I., Lopez-Beltran, E. A., Kluner, P., Luque, J., Ballesteros, P., and Cerdan, S. (1999). Molecular crowding and viscosity as determinants of translational diffusion of metabolites in subcellular organelles. Arch. Biochem. Biophys. 362, 329–38.CrossRefGoogle ScholarPubMed
Gardner, R. L. (2001). The initial phase of embryonic patterning in mammals. Int. Rev. Cytol. 203, 233–90.CrossRefGoogle ScholarPubMed
Gerhart, J. (2002). Changing the axis changes the perspective. Dev. Dyn. 225, 380–3.CrossRefGoogle Scholar
Gerhart, J., Ubbels, G., Black, S., Hara, K., and Kirschner, M. (1981). A reinvestigation of the role of the grey crescent in axis formation in Xenopus laevis. Nature 292, 511–6.CrossRefGoogle ScholarPubMed
Ghosh, S., and Comper, W. D. (1988). Oriented fibrillogenesis of collagen in vitro by ordered convection. Connect. Tissue. Res. 17, 33–41.CrossRefGoogle ScholarPubMed
Giancotti, F. G., and Ruoslahti, E. (1999). Integrin signaling. Science 285, 1028–32.CrossRefGoogle ScholarPubMed
Giansanti, M. G., Bonaccorsi, S., Bucciarelli, E., and Gatti, M. (2001). Drosophila male meiosis as a model system for the study of cytokinesis in animal cells. Cell Struct. Funct. 26, 609–17.CrossRefGoogle Scholar
Gierer, A. (1977). Physical aspects of tissue evagination and biological form. Quart. Rev. Biophys. 10, 529–93.CrossRefGoogle ScholarPubMed
Gilbert, S. F. (2003). Developmental Biology. Sinauer Associates, Sunderland, MA.Google ScholarPubMed
Gilkey, J. C., Jaffe, L. F., Ridgway, E. B., and Reynolds, G. T. (1978). A free calcium wave traverses the activating egg of the medaka, Oryzias latipes. J. Cell Biol. 76, 448–66.CrossRefGoogle ScholarPubMed
Gimlich, R. L. (1985). Cytoplasmic localization and chordamesoderm induction in the frog embryo. J. Embryol. Exp. Morphol. 89 Suppl., 89–111.Google ScholarPubMed
Gimlich, R. L. (1986). Acquisition of developmental autonomy in the equatorial region of the Xenopus embryo. Dev. Biol. 115, 340–52.CrossRefGoogle ScholarPubMed
Ginsburg, M., Snow, M. H., and McLaren, A. (1990). Primordial germ cells in the mouse embryo during gastrulation. Development 110, 521–8.Google ScholarPubMed
Giraud-Guille, M. M. (1996). Twisted liquid crystalline supramolecular arrangements in morphogenesis. Int. Rev. Cytol. 166, 59–101.CrossRefGoogle ScholarPubMed
Glahn, D., and Nuccitelli, R. (2003). Voltage-clamp study of the activation currents and fast block to polyspermy in the egg of Xenopus laevis. Dev. Growth Differ. 45, 187–97.CrossRefGoogle ScholarPubMed
Glazier, J. A., and Graner, F. (1993). A simulation of the differential adhesion driven rearrangement of biological cells. Phys. Rev. E 47, 2128–54.CrossRefGoogle ScholarPubMed
Godt, D., and Tepass, U. (1998). Drosophila oocyte localization is mediated by differential cadherin-based adhesion. Nature 395, 387–91.CrossRefGoogle ScholarPubMed
Goldbeter, A. (1996). Biochemical Oscillations and Cellular Rhythms: the Molecular Bases of Periodic and Chaotic Behaviour. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Gomperts, M., Wylie, C., and Heasman, J. (1994). Primordial germ cell migration. Ciba Found. Symp. 182, 121–34; discussion 134–9.Google ScholarPubMed
Gong, Y., Mo, C., and Fraser, S. E. (2004). Planar cell polarity signalling controls cell division orientation during zebrafish gastrulation. Nature 430, 689–93.CrossRefGoogle ScholarPubMed
Gonzalez-Reyes, A., and St Johnston, D. (1998). The Drosophila AP axis is polarised by the cadherin-mediated positioning of the oocyte. Development 125, 3635–44.Google ScholarPubMed
Gonze, D., and Goldbeter, A. (2001). A model for a network of phosphorylation–dephosphorylation cycles displaying the dynamics of dominoes and clocks. J. Theor. Biol. 210, 167–86.CrossRefGoogle ScholarPubMed
Gosden, R., Krapez, J., and Briggs, D. (1997). Growth and development of the mammalian oocyte. Bioessays 19, 875–82.CrossRefGoogle ScholarPubMed
Gossler, A., and Angelis, Hrabe M. (1998). Somitogenesis. Curr. Top. Dev. Biol. 38, 225–87.CrossRefGoogle ScholarPubMed
Gould, S. E., Upholt, W. B., and Kosher, R. A. (1992). Syndecan 3: a member of the syndecan family of membrane-intercalated proteoglycans that is expressed in high amounts at the onset of chicken limb cartilage differentiation. Proc. Nat. Acad. Sci. USA 89, 3271–75.CrossRefGoogle ScholarPubMed
Gould, S. J. (1977). Ontogeny and Phylogeny. Harvard University Press, Cambridge, MA.Google Scholar
Goulian, M., and Simon, S. M. (2000). Tracking single proteins within cells. Biophys. J. 79, 2188–98.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
Green, J. (2002). Morphogen gradients, positional information, and Xenopus: interplay of theory and experiment. Dev. Dyn. 225, 392–408.CrossRefGoogle ScholarPubMed
Greenwald, I. (1998). LIN-12/Notch signaling: lessons from worms and flies. Genes Dev. 12, 1751–62.CrossRefGoogle ScholarPubMed
Greenwald, I., and Rubin, G. M. (1992). Making a difference: the role of cell–cell interactions in establishing separate identities for equivalent cells. Cell 68, 271–81.CrossRefGoogle ScholarPubMed
Gregory, P. D., Wagner, K., and Horz, W. (2001). Histone acetylation and chromatin remodeling. Exp. Cell Res. 265, 195–202.CrossRefGoogle ScholarPubMed
Gritsman, K., Talbot, W. S., and Schier, A. F. (2000). Nodal signaling patterns the organizer. Development 127, 921–32.Google ScholarPubMed
Gullberg, D. E., and Lundgren-Akerlund, E. (2002). Collagen-binding I domain integrins – what do they do?Prog. Histochem. Cytochem. 37, 3–54.CrossRefGoogle Scholar
Gumbiner, B. M. (1996). Cell adhesion: the molecular basis of tissue architecture and morphogenesis. Cell 84, 345–57.CrossRefGoogle ScholarPubMed
Gumbiner, B. M. (2000). Regulation of cadherin adhesive activity. J. Cell. Biol. 148, 399–404.CrossRefGoogle ScholarPubMed
Gurdon, J. B. (1988). A community effect in animal development. Nature 336, 772–4.CrossRefGoogle ScholarPubMed
Guthrie, S., and Lumsden, A. (1991). Formation and regeneration of rhombomere boundaries in the developing chick hindbrain. Development 112, 221–9.Google ScholarPubMed
Haddon, C., Smithers, L., Schneider-Maunoury, S., Coche, T., Henrique, D., and Lewis, J. (1998). Multiple delta genes and lateral inhibition in zebrafish primary neurogenesis. Development 125, 359–70.Google ScholarPubMed
Hafner, M., Petzelt, C., Nobiling, R., Pawley, J. B., Kramp, D., and Schatten, G. (1988). Wave of free calcium at fertilization in the sea urchin egg visualized with fura-2. Cell Motil. Cytoskeleton 9, 271–7.CrossRefGoogle ScholarPubMed
Haga, H., Nagayama, M., Kawabata, K., Ito, E., Ushiki, T., and Sambongi, T. (2000). Time-lapse viscoelastic imaging of living fibroblasts using force modulation mode in AFM. J. Electron Microsc. (Tokyo) 49, 473–81.CrossRefGoogle ScholarPubMed
Hahn, H. S., Ortoleva, P. J., and Ross, J. (1973). Chemical oscillations and multiple steady states due to variable boundary permeability. J. Theor. Biol. 41, 503–21.CrossRefGoogle ScholarPubMed
Hall, B. K., and Miyake, T. (1995). Divide, accumulate, differentiate: cell condensation in skeletal development revisited. Int. J. Dev. Biol. 39, 881–93.Google ScholarPubMed
Hall, B. K., and Miyake, T. (2000). All for one and one for all: condensations and the initiation of skeletal development. Bioessays 22, 138–47.3.0.CO;2-4>CrossRefGoogle ScholarPubMed
Hall, D., and Minton, A. P. (2003). Macromolecular crowding: qualitative and semiquantitative successes, quantitative challenges. Biochim. Biophys. Acta 1649, 127–39.CrossRefGoogle ScholarPubMed
Hardin, J. (1996). The cellular basis of sea urchin gastrulation. Curr. Top. Dev. Biol. 33, 159–262.CrossRefGoogle ScholarPubMed
Harding, K., Hoey, T., Warrior, R., and Levine, M. (1989). Autoregulatory and gap gene response elements of the even-skipped promoter of Drosophila. EMBO J. 8, 1205–12.Google ScholarPubMed
Hardman, P., and Spooner, B. S. (1992). Salivary epithelium branching morphogenesis. In Epithelial Organization and Development. (Fleming, T. P., ed.), pp. 353–75. Chapman and Hall, London.CrossRefGoogle Scholar
Harland, R., and Gerhart, J. (1997). Formation and function of Spemann's organizer. Ann. Rev. Cell Dev. Biol. 13, 611–67.CrossRefGoogle ScholarPubMed
Harper, G. S., Comper, W. D., and Preston, B. N. (1984). Dissipative structures in proteoglycan solutions. J. Biol. Chem. 259, 10582–9.Google ScholarPubMed
Harrison, L. J. (1993). Kinetic Theory of Living Form. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Harrisson, F. (1989). The extracellular matrix and cell surface, mediators of cell interactions in chicken gastrulation. Int. J. Dev. Biol. 33, 417–38.Google ScholarPubMed
Hayashi, T., and Carthew, R. W. (2004). Surface mechanics mediate pattern formation in the developing retina. Nature 431, 647–52.CrossRefGoogle ScholarPubMed
He, X., and Dembo, M. (1997). On the mechanics of the first cleavage division of the sea urchin egg. Exp. Cell Res. 233, 252–73.CrossRefGoogle ScholarPubMed
Heintzelman, K. F., Phillips, H. M., and Davis, G. S. (1978). Liquid-tissue behavior and differential cohesiveness during chick limb budding. J. Embryol. Exp. Morphol. 47, 1–15.Google ScholarPubMed
Helfrich, W. (1973). Elastic properties of lipid bilayers: theory and possible experiments. Z. Naturforsch. C 28, 693–703.CrossRefGoogle ScholarPubMed
Hentschel, H. G., Glimm, T., Glazier, J. A., and Newman, S. A. (2004). Dynamical mechanisms for skeletal pattern formation in the vertebrate limb. Proc. R. Soc. Lond. B Biol. Sci. 271, 1713–22.CrossRefGoogle ScholarPubMed
Heyman, I., Faissner, A., and Lumsden, A. (1995). Cell and matrix specialisations of rhombomere boundaries. Dev. Dyn. 204, 301–15.CrossRefGoogle ScholarPubMed
Hieda, Y., and Nakanishi, Y. (1997). Epithelial morphogenesis in mouse embryonic submandibular gland: its relationships to the tissue organization of epithelium and mesenchyme. Dev. Growth Differ. 39, 1–8.CrossRefGoogle ScholarPubMed
Hiramoto, Y. (1956). Cell division without mitotic apparatus in sea urchin eggs. Exp. Cell Res. 11, 630–636.CrossRefGoogle ScholarPubMed
Hiramoto, Y. (1958). A quantitative description of protoplasmic movement during cleavage in the sea-urchin egg. J. Exp. Biol. 35, 407–424.Google Scholar
Hiramoto, Y. (1963). Mechanical properties of sea urchin eggs. I. Surface force and elastic modulus of the cell membrane. Exp. Cell Res. 32, 59–75.CrossRefGoogle ScholarPubMed
Hiramoto, Y. (1968). The mechanics and mechanism of cleavage in the sea-urchin egg. Symp. Soc. Exp. Biol. 22, 311–27.Google ScholarPubMed
Hiramoto, Y. (1978). Mechanical properties of the dividing sea urchin egg. In Cell Motility: Molecules and Organization (Hatano, S., Ishikawa, H., and Sato, H., eds.), pp. 653–63. University Park Press, Baltimore.Google Scholar
Hobbie, R. K. (1997). Intermediate Physics for Medicine and Biology. Springer-Verlag, New York.Google Scholar
Hofmeyr, J. H., and Cornish-Bowden, A. (1997). The reversible Hill equation: how to incorporate cooperative enzymes into metabolic models. Comput. Appl. Biosci. 13, 377–85.Google ScholarPubMed
Holley, S. A., Geisler, R., and Nusslein-Volhard, C. (2000). Control of her1 expression during zebrafish somitogenesis by a Delta-dependent oscillator and an independent wave-front activity. Genes Dev. 14, 1678–90.Google ScholarPubMed
Holley, S. A., Julich, D., Rauch, G. J., Geisler, R., and Nusslein-Volhard, C. (2002). her1 and the notch pathway function within the oscillator mechanism that regulates zebrafish somitogenesis. Development 129, 1175–83.Google ScholarPubMed
Holtzendorff, J., Hung, D., Brende, P., et al. (2004). Oscillating global regulator control the genetic circuit driving a bacterial cell cycle. Science 304, 983–7.CrossRefGoogle ScholarPubMed
Hörstadius, S., and Sellman, S. (1946). Experimentelle Untersuchungen uber die Determination des knorpeligen Kopfskelettes bei Urodelen. Nova Acta R. Soc. Scient. Upsal. Ser. 4 13, 1–170.Google Scholar
Hoshi, M. (1979). Exogastrulation induced by heavy water in sea urchin larvae. Cell Differ. 8, 431–5.CrossRefGoogle ScholarPubMed
Howard, J. (2001). Mechanics of Motor Proteins and the Cytoskeleton. Sinauer Associates, Inc., Sunderland.Google Scholar
Howard, K., and Ingham, P. (1986). Regulatory interactions between the segmentation genes fushi tarazu, hairy, and engrailed in the Drosophila blastoderm. Cell 44, 949–57.CrossRefGoogle ScholarPubMed
Huang, F. Z., Bely, A. E., and Weisblat, D. A. (2001). Stochastic WNT signaling between nonequivalent cells regulates adhesion but not fate in the two-cell leech embryo. Curr. Biol. 11, 1–7.CrossRefGoogle Scholar
Hunkapiller, T., and Hood, L. (1989). Diversity of the immunoglobulin gene superfamily. Adv. Immunol. 44, 1–63.CrossRefGoogle ScholarPubMed
Huppert, S. S., Jacobsen, T. L., and Muskavitch, M. A. (1997). Feedback regulation is central to Delta–Notch signalling required for Drosophila wing vein morphogenesis. Development 124, 3283–91.Google ScholarPubMed
Hutson, M. S., Tokutake, Y., Chang, M. S., Bloor, J. W., Venakides, S., Kiehart, D. P., and Edwards, G. S. (2003). Forces for morphogenesis investigated with laser microsurgery and quantitative modeling. Science 300, 145–9.CrossRefGoogle ScholarPubMed
Hynes, R. O. (1987). Integrins: a family of cell surface receptors. Cell 48, 349–54.CrossRefGoogle ScholarPubMed
Hynes, R. O. (1992). Integrins: versatility, modulation, and signaling in cell adhesion. Cell 69, 11–25.CrossRefGoogle ScholarPubMed
Hynes, R. O., and Lander, A. D. (1992). Contact and adhesive specificities in the associations, migrations, and targeting of cells and axons. Cell 68, 303–22.CrossRefGoogle ScholarPubMed
Ingber, D. E. (1991). Extracellular matrix and cell shape: potential control points for inhibition of angiogenesis. J. Cell. Biochem. 47, 236–41.CrossRefGoogle ScholarPubMed
Ingber, D. E., and Folkman, J. (1989a). How does extracellular matrix control capillary morphogenesis?Cell 58, 803–5.CrossRefGoogle ScholarPubMed
Ingber, D. E., and Folkman, J. (1989b). Mechanochemical switching between growth and differentiation during fibroblast growth factor-stimulated angiogenesis in vitro: role of extracellular matrix. J. Cell Biol. 109, 317–30.CrossRefGoogle ScholarPubMed
Ingham, P. W. (1988). The molecular genetics of embryonic pattern formation in Drosophila. Nature 335, 25–34.CrossRefGoogle ScholarPubMed
Ingolia, N. T. (2004). Topology and robustness in the Drosophila segment polarity network. PLoS Biol. 2, 805–15.CrossRefGoogle ScholarPubMed
International Human Genome Consortium (2004). Finishing the euchromatic sequence of the human genome. Nature 431, 931–45.CrossRef
Irvine, K. D., and Wieschaus, E. (1994). Cell intercalation during Drosophila germband extension and its regulation by pair-rule segmentation genes. Development 120, 827–41.Google ScholarPubMed
Ish-Horowicz, D., Pinchin, S. M., Ingham, P. W., and Gyurkovics, H. G. (1989). Autocatalytic ftz activation and instability induced by ectopic ftz expression. Cell 57, 223–232.CrossRefGoogle ScholarPubMed
Israelachvili, J. N. (1991). Intermolecular and Surface Forces. Academic Press, London, San Diego.Google Scholar
Itow, T. (1986). Inhibitors of DNA synthesis change the differentiation of body segments and increase the segment number in horseshoe crab embryos. Roux's Arch. Dev. Biol. 195, 323–33.CrossRefGoogle ScholarPubMed
Jacobson, A. G., Oster, G. F., Odell, G. M., and Cheng, L. Y. (1986). Neurulation and the cortical tractor model for epithelial folding. J. Embryol. Exp. Morphol. 96, 19–49.Google ScholarPubMed
Jacobson, K. A., Sheets, E. D., and Simson, R. (1995). Revisiting the fluid mosaic model of membranes. Science 268, 1441–2.CrossRefGoogle ScholarPubMed
Jacobson, K. A., Moore, S. E., Yang, B., Doherty, P., Gordon, G. W., and Walsh, F. S. (1997). Cellular determinants of the lateral mobility of neural cell adhesion molecules. Biochim. Biophys. Acta 1330, 138–44.CrossRefGoogle ScholarPubMed
Jaffe, L. A. (1976). Fast block to polyspermy in sea urchin eggs is electrically mediated. Nature 261, 68–71.CrossRefGoogle ScholarPubMed
Jaffe, L. A., and Cross, N. L. (1986). Electrical regulation of sperm–egg fusion. Ann. Rev. Physiol. 48, 191–200.CrossRefGoogle ScholarPubMed
Jaffe, L. A., Sharp, A. P., and Wolf, D. P. (1983). Absence of an electrical polyspermy block in the mouse. Dev. Biol. 96, 317–23.CrossRefGoogle ScholarPubMed
Jaffe, L. A., Giusti, A. F., Carroll, D. J., and Foltz, K. R. (2001). Ca2+ signalling during fertilization of echinoderm eggs. Semin. Cell Dev. Biol. 12, 45–51.CrossRefGoogle ScholarPubMed
Jakab, K., Neagu, A., Mironov, V., Markwald, R. R., and Forgacs, G. (2004). Engineering biological structures of prescribed shape using self-assembling multicellular systems. Proc. Nat. Acad. Sci. USA 101, 2864–9.CrossRefGoogle ScholarPubMed
Janmey, P. (1998). The cytoskeleton and cell signaling: component localization and mechanical coupling. Physiol. Rev. 78, 763–81.CrossRefGoogle ScholarPubMed
Janson, L. W., and Taylor, D. L. (1993). In vitro models of tail contraction and cytoplasmic streaming in amoeboid cells. J. Cell Biol. 123, 345–56.CrossRefGoogle ScholarPubMed
Jen, W. C., Gawantka, V., Pollet, N., Niehrs, C., and Kintner, C. (1999). Periodic repression of Notch pathway genes governs the segmentation of Xenopus embryos. Genes Dev. 13, 1486–99.CrossRefGoogle ScholarPubMed
Jiang, T., Jung, H., Widelitz, R. B., and Chuong, C. (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
Jiang, Y., Levine, H., and Glazier, J. A. (1998). Possible cooperation of differential adhesion and chemotaxis in mound formation of Dictyostelium. Biophys. J. 75, 2615–25.CrossRefGoogle ScholarPubMed
Jiang, Y. J., Aerne, B. L., Smithers, L., Haddon, C., Ish-Horowicz, D., and Lewis, J. (2000). Notch signalling and the synchronization of the somite segmentation clock. Nature 408, 475–9.Google ScholarPubMed
Jimenez, G., Griffiths, S. D., Ford, A. M., Greaves, M. F., and Enver, T. (1992). Activation of the beta-globin locus control region precedes commitment to the erythroid lineage. Proc. Nat. Acad. Sci. USA 89, 10618–22.CrossRefGoogle ScholarPubMed
Jones, K. T. (1998). Protein kinase C action at fertilization: overstated or undervalued?Rev. Reprod. 3, 7–12.CrossRefGoogle ScholarPubMed
Joos, T. O., Whittaker, C. A., Meng, F., DeSimone, D. W., Gnau, V., and Hausen, P. (1995). Integrin alpha 5 during early development of Xenopus laevis. Mech. Dev. 50, 187–99.CrossRefGoogle ScholarPubMed
Jouve, C., Palmeirim, I., Henrique, D., et al. (2000). Notch signalling is required for cyclic expression of the hairy-like gene HES1 in the presomitic mesoderm. Development 127, 1421–9.Google ScholarPubMed
Jouve, C., Iimura, T., and Pourquié, O. (2002). Onset of the segmentation clock in the chick embryo: evidence for oscillations in the somite precursors in the primitive streak. Development 129, 1107–17.Google ScholarPubMed
Juan, H., and Hamada, H. (2001). Roles of nodal–lefty regulatory loops in embryonic patterning of vertebrates. Genes Cells 6, 923–30.CrossRefGoogle ScholarPubMed
Kabata, H., Kurosawa, O., Arai, I., et al. (1993). Visualization of single molecules of RNA polymerase sliding along DNA. Science 262, 1561–3.CrossRefGoogle ScholarPubMed
Kadler, K. E., Holmes, D. F., Trotter, J. A., and Chapman, J. A. (1996). Collagen fibril formation. Biochem. J. 316, 1–11.CrossRefGoogle ScholarPubMed
Kalodimos, C. G., Biris, N., Bonvin, A. M., et al. (2004). Structure and flexibility adaptation in nonspecific and specific protein–DNA complexes. Science 305, 386–9.CrossRefGoogle ScholarPubMed
Kamawaki, Y., Raya, A., Raya, R. M., Rodriquez-Esteban, C., and Belmont, J. C. I. (2005). Retinoic acid signalling links, left–right asymmetric patterning and bilaterally symmetric somitogenesis in the zebrafish embryo. Nature 435, 165–71.CrossRefGoogle Scholar
Kamei, N., Swanson, W. J., and Glabe, C. G. (2000). A rapidly diverging EGF protein regulates species-specific signal transduction in early sea urchin development. Dev. Biol. 225, 267–76.CrossRefGoogle ScholarPubMed
Kaneko, K. (2003). Organization through intra–inter dynamics. In Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology (Müller, G. B. and Newman, S. A., eds.), pp. 195–220. MIT Press, Cambridge, MA.Google Scholar
Kaneko, K., and Yomo, T. (1994). Cell division, differentiation and dynamic clustering. Physica D 75, 89–102.CrossRefGoogle Scholar
Kaneko, K., and Yomo, T. (1997). Isologous diversification: a theory of cell differentiation. Bull. Math. Biol. 59, 139–96.CrossRefGoogle ScholarPubMed
Kaneko, K., and Yomo, T. (1999). Isologous diversification for robust development of cell society. J. Theor. Biol. 199, 243–56.CrossRefGoogle ScholarPubMed
Karr, T. L., Weir, M. P., Ali, Z., and Kornberg, T. (1989). Patterns of engrailed protein in early Drosophila embryos. Development 105, 605–612.Google ScholarPubMed
Kawaki, Y.,Raya, A.,Raya, R. M., Rodriguez-Esteban, C., and Belmonte, J. C. I. (2005). Retinoic acid signalling links left–right asymmetric patterning and bilaterally symmetric somitogenesis in the zebrafish embryo. Nature 435, 165–71.CrossRefGoogle Scholar
Keller, A. D. (1995). Model genetic circuits encoding autoregulatory transcription factors. J. Theor. Biol. 172, 169–85.CrossRefGoogle ScholarPubMed
Keller, R. (2000). The origin and morphogenesis of amphibian somites. Curr. Top. Dev. Biol. 47, 183–246.CrossRefGoogle ScholarPubMed
Keller, R. (2002). Shaping the vertebrate body plan by polarized embryonic cell movements. Science 298, 1950–4.CrossRefGoogle ScholarPubMed
Keller, R., and Danilchik, M. (1988). Regional expression, pattern and timing of convergence and extension during gastrulation of Xenopus laevis. Development 103, 193–209.Google ScholarPubMed
Keller, R. E., Danilchik, M., Gimlich, R., and Shih, J. (1985). The function and mechanism of convergent extension during gastrulation of Xenopus laevis. J. Embryol. Exp. Morphol. 89 Suppl., 185–209.Google ScholarPubMed
Keller, R., Cooper, M. S., Danilchik, M., Tibbetts, P., and Wilson, P. A. (1989). Cell intercalation during notochord development in Xenopus laevis. J. Exp. Zool. 251, 134–54.CrossRefGoogle ScholarPubMed
Keller, R., Shih, J., and Sater, A. (1992a). The cellular basis of the convergence and extension of the Xenopus neural plate. Dev. Dyn. 193, 199–217.CrossRefGoogle Scholar
Keller, R., Shih, J., Sater, A. K., and Moreno, C. (1992b). Planar induction of convergence and extension of the neural plate by the organizer of Xenopus. Dev. Dyn. 193, 218–34.CrossRefGoogle Scholar
Keller, R., Davidson, L., Edlund, A.et al. (2000). Mechanisms of convergence and extension by cell intercalation. Philos. Trans. R. Soc. Lond. B Biol. Sci. 355, 897–922.CrossRefGoogle ScholarPubMed
Kerszberg, M., and Changeux, J. P. (1998). A simple molecular model of neurulation. Bioessays 20, 758–70.3.0.CO;2-C>CrossRefGoogle ScholarPubMed
Kerszberg, M., and Wolpert, L. (1998). Mechanisms for positional signalling by morphogen transport: a theoretical study. J. Theor. Biol. 191, 103–14.CrossRefGoogle ScholarPubMed
Kimmins, S., and Sassone-Corsi, P. (2005). Chromatin remodeling and epigenetic features of germ cells. Nature 434, 583–9.CrossRefGoogle ScholarPubMed
Kiskowski, M. A., Alber, M. S., Thomas, G. L., et al. (2004). Interplay between activator–inhibitor coupling and cell–matrix adhesion in a cellular automaton model for chondrogenic patterning. Dev. Biol. 271, 372–87.CrossRefGoogle Scholar
Klein, C., and Hurlbut, C. S. (2002). Manual of Mineral Science. Wiley, New York.Google Scholar
Knoll, A. H. (2003). Life on a Young Planet: the First Three Billion Years of Evolution on Earth. Princeton University Press, Princeton, NJ.Google Scholar
Kofron, M., Spagnuolo, A., Klymkowsky, M., Wylie, C., and Heasman, J. (1997). The roles of maternal alpha-catenin and plakoglobin in the early Xenopus embryo. Development 124, 1553–60.Google ScholarPubMed
Kofron, M., Heasman, J., Lang, S. A., and Wylie, C. C. (2002). Plakoglobin is required for maintenance of the cortical actin skeleton in early Xenopus embryos and for cdc42-mediated wound healing. J. Cell Biol. 158, 695–708.CrossRefGoogle ScholarPubMed
Kohn, K. W. (1999). Molecular interaction map of the mammalian cell cycle control and DNA repair systems. Mol. Biol. Cell. 10, 2703–34.CrossRefGoogle ScholarPubMed
Kondo, S., and Asai, R. (1995). A reaction–diffusion wave on the skin of the marine angelfish Pomacanthus. Nature 376, 765–8.CrossRefGoogle ScholarPubMed
Kosher, R. A., Savage, M. P., and Chan, S. C. (1979). In vitro studies on the morphogenesis and differentiation of the mesoderm subjacent to the apical ectodermal ridge of the embryonic chick limb-bud. J. Embryol. Exp. Morphol. 50, 75–97.Google ScholarPubMed
Kosher, R. A., Walker, K. H., and Ledger, P. W. (1982). Temporal and spatial distribution of fibronectin during development of the embryonic chick limb bud. Cell. Differ. 11, 217–228.CrossRefGoogle ScholarPubMed
Kruse, K., Pantazis, P., Bollenbach, T., Julicher, F., and Gonzalez-Gaitan, M. (2004). Dpp gradient formation by dynamin-dependent endocytosis: receptor trafficking and the diffusion model. Development 131, 4843–56.CrossRefGoogle ScholarPubMed
Kubota, H. Y., Yoshimoto, Y., Yoneda, M., and Hiramoto, Y. (1987). Free calcium wave upon activation in Xenopus eggs. Dev. Biol. 119, 129–36.CrossRefGoogle ScholarPubMed
Kulesa, P. M., and Fraser, S. E. (2000). In ovo time-lapse analysis of chick hindbrain neural crest cell migration shows cell interactions during migration to the branchial arches. Development 127, 1161–72.Google ScholarPubMed
Kulesa, P. M., and Fraser, S. E. (2002). Cell dynamics during somite boundary formation revealed by time-lapse analysis. Science 298, 991–5.CrossRefGoogle ScholarPubMed
Kumano, G., and Smith, W. C. (2002). Revisions to the Xenopus gastrula fate map: implications for mesoderm induction and patterning. Dev. Dyn. 225, 409–21.CrossRefGoogle ScholarPubMed
Lallier, T. E., Whittaker, C. A., and DeSimone, D. W. (1996). Integrin alpha 6 expression is required for early nervous system development in Xenopus laevis. Development 122, 2539–54.Google ScholarPubMed
Lander, A. D., Nie, Q., and Wan, F. Y. (2002). Do morphogen gradients arise by diffusion?Dev. Cell 2, 785–96.CrossRefGoogle ScholarPubMed
Lane, M. C., and Sheets, M. D. (2000). Designation of the anterior/posterior axis in pregastrula Xenopus laevis. Dev. Biol. 225, 37–58.CrossRefGoogle ScholarPubMed
Lane, M. C., and Sheets, M. D. (2002). Rethinking axial patterning in amphibians. Dev. Dyn. 225, 434–47.CrossRefGoogle ScholarPubMed
Lane, M. C., and Smith, W. C. (1999). The origins of primitive blood in Xenopus: implications for axial patterning. Development 126, 423–34.Google ScholarPubMed
Lane, M. C., Koehl, M. A., Wilt, F., and Keller, R. (1993). A role for regulated secretion of apical extracellular matrix during epithelial invagination in the sea urchin. Development 117, 1049–60.Google ScholarPubMed
Langille, R. M., and Hall, B. K. (1993). Pattern formation and the neural crest. In The Vertebrate Skull (Hanken, J. and Hall, B. K., eds.), Vol. 1, Development, pp. 77–111. University of Chicago Press, Chicago.Google ScholarPubMed
Langman, J. (1981). Medical Embryology. Williams & Wilkins, Baltimore.Google Scholar
Lawrence, P. A. (1992). The Making of a Fly: the Genetics of Animal Design. Blackwell Scientific Publications, Oxford, Boston.Google Scholar
Lawson, K. A. (1983). Stage specificity in the mesenchyme requirement of rodent lung epithelium in vitro: a matter of growth control?J. Embryol. Exp. Morphol. 74, 183–206.Google ScholarPubMed
Lechleiter, J., Girard, S., Clapham, D., and Peralta, E. (1991). Subcellular patterns of calcium release determined by G protein-specific residues of muscarinic receptors. Nature 350, 505–8.CrossRefGoogle Scholar
Lechleiter, J. D., John, L. M., and Camacho, P. (1998). Ca2+ wave dispersion and spiral wave entrainment in Xenopus laevis oocytes overexpressing Ca2+ ATPases. Biophys. Chem. 72, 123–9.CrossRefGoogle ScholarPubMed
Leal, L. G. (1992). Laminar Flow and Convective Processes: Scaling Principles and Asymptotic Analysis. Butterworth-Heinemann, Boston.Google Scholar
Lengyel, I., and Epstein, I. R. (1991). Diffusion-induced instability in chemically reacting systems: steady-state multiplicity, oscillation, and chaos. Chaos 1, 69–76.CrossRefGoogle ScholarPubMed
Lengyel, I., and Epstein, I. R. (1992). A chemical approach to designing Turing patterns in reaction–diffusion systems. Proc. Nat. Acad. Sci. USA 89, 3977–9.CrossRefGoogle ScholarPubMed
Leonard, C. M., Fuld, H. M., Frenz, D. A., Downie, S. A., Massague, J., and Newman, S. A. (1991). Role of transforming growth factor-β in chondrogenic pattern formation in the embryonic limb: stimulation of mesenchymal condensation and fibronectin gene expression by exogenous TGF-beta and evidence for endogenous TGF-β-like activity. Dev. Biol. 145, 99–109.CrossRefGoogle Scholar
Lercher, M. J., Urrutia, A. O., and Hurst, L. D. (2002). Clustering of housekeeping genes provides a unified model of gene order in the human genome. Nat. Genet. 31, 180–3.CrossRefGoogle ScholarPubMed
Leslie, P. H. (1948). Some further notes on the use of matrices in population mathematics. Biometrika 35, 213–45.CrossRefGoogle Scholar
Levi, G., Ginsberg, D., Girault, J. M., Sabanay, I., Thiery, J. P., and Geiger, B. (1991). EP-cadherin in muscles and epithelia of Xenopus laevis embryos. Development 113, 1335–44.Google ScholarPubMed
Lewis, J. (2003). Autoinhibition with transcriptional delay: a simple mechanism for the zebrafish somitogenesis oscillator. Curr. Biol. 13, 1398–408.CrossRefGoogle ScholarPubMed
Li, S., Piotrowicz, R. S., Levin, E. G., Shyy, Y. J., and Chien, S. (1996). Fluid shear stress induces the phosphorylation of small heat shock proteins in vascular endothelial cells. Am. J. Physiol. 271, C994–1000.CrossRefGoogle ScholarPubMed
Li, S., Zhou, D., Lu, M. M., and Morrisey, E. E. (2004). Advanced cardiac morphogenesis does not require heart tube fusion. Science 305, 1619–22.CrossRefGoogle Scholar
Linsenmayer, T. F., Fitch, J. M., Gordon, M. K., Cai, C. X., Igoe, F., Marchant, J. K., and Birk, D. E. (1998). Development and roles of collagenous matrices in the embryonic avian cornea. Prog. Retin. Eye Res. 17, 231–65.Google ScholarPubMed
Lowery, L. A., and Sive, H. (2004). Strategies of vertebrate neurulation and a re-evaluation of teleost neural tube formation. Mech. Dev. 121, 1189–97.CrossRefGoogle Scholar
Lubarsky, B., and Krasnow, M. A. (2003). Tube morphogenesis: making and shaping biological tubes. Cell 112, 19–28.CrossRefGoogle ScholarPubMed
Lubkin, S. R., and Li, Z. (2002). Force and deformation on branching rudiments: cleaving between hypotheses. Biomech. Model Mechanobiol. 1, 5–16.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
Luo, Y., Kostetskii, I., and Radiche, G. L. (2005). N-cadherin is not essential for limb mesenchymal chondrogenesis. Dev. Dyn. 232, 336–44.CrossRefGoogle Scholar
Mandato, C. A., Benink, H. A., and Bement, W. M. (2000). Microtubule-actomyosin interactions in cortical flow and cytokinesis. Cell Motil. Cytoskeleton 45, 87–92.3.0.CO;2-0>CrossRefGoogle ScholarPubMed
Mandelbrot, B. B. (1983). The Fractal Geometry of Nature. W. H. Freeman, New York.
Maniatis, T., and Tasic, B. (2002). Alternative pre-mRNA splicing and proteome expansion in metazoans. Nature 418, 236–43.CrossRefGoogle ScholarPubMed
Manner, J. (2000). Cardiac looping in the chick embryo: a morphological review with special reference to terminological and biomechanical aspects of the looping process. Anat. Rec. 259, 248–62.3.0.CO;2-K>CrossRefGoogle ScholarPubMed
Mannervik, M., Nibu, Y., Zhang, H., and Levine, M. (1999). Transcriptional coregulators in development. Science 284, 606–9.CrossRefGoogle ScholarPubMed
Marom, K., Shapira, E., and Fainsod, A. (1997). The chicken caudal genes establish an anterior–posterior gradient by partially overlapping temporal and spatial patterns of expression. Mech. Dev. 64, 41–52.CrossRefGoogle Scholar
Marsden, M., and DeSimone, D. W. (2003). Integrin-ECM interactions regulate cadherin-dependent cell adhesion and are required for convergent extension in Xenopus. Curr. Biol. 13, 1182–91.CrossRefGoogle ScholarPubMed
Marshall, B. T., Long, M., Piper, J. W., Yago, T., McEver, R. P., and Zhu, C. (2003). Direct observation of catch bonds involving cell-adhesion molecules. Nature 423, 190–3.CrossRefGoogle ScholarPubMed
Martin, G. R. (1998). The roles of FGFs in the early development of vertebrate limbs. Genes Dev. 12, 1571–86.CrossRefGoogle ScholarPubMed
Martin, J. E., Adolf, D., and Wilcoxon, J. P. (1989). Rheology of the incipient gel: theory and data for epoxies. Polym. Prepr. Am. Chem. Soc. Div. Polym. Chem. 30, 83–84.Google Scholar
Martin, V. J., Littlefield, C. L., Archer, W. E., and Bode, H. R. (1997). Embryogenesis in hydra. Biol. Bull. 192, 345–63.CrossRefGoogle ScholarPubMed
Smith, Maynard J. (1978). Models in Ecology. Cambridge University Press, Cambridge, New York.Google Scholar
Mayr, E. (1982). The Growth of Biological Thought: Diversity, Evolution, and Inheritance. Belknap Press, Cambridge, MA.Google Scholar
McCarthy, R. A., and Hay, E. D. (1991). Collagen I, laminin, and tenascin: ultrastructure and correlation with avian neural crest formation. Int. J. Dev. Biol. 35, 437–52.Google ScholarPubMed
McDougall, A., Shearer, J., and Whitaker, M. (2000). The initiation and propagation of the fertilization wave in sea urchin eggs. Biol. Cell. 92, 205–14.CrossRefGoogle ScholarPubMed
McDowell, N., Gurdon, J. B., and Grainger, D. J. (2001). Formation of a functional morphogen gradient by a passive process in tissue from the early Xenopus embryo. Int. J. Dev. Biol. 45, 199–207.Google ScholarPubMed
McKim, K. S., Jang, J. K., and Manheim, E. A. (2002). Meiotic recombination and chromosome segregation in Drosophila females. Ann. Rev. Genet. 36, 205–32.CrossRefGoogle ScholarPubMed
McLaren, A. (1984). Meiosis and differentiation of mouse germ cells. Symp. Soc. Exp. Biol. 38, 7–23.Google ScholarPubMed
Medalia, O., Weber, I., Frangakis, A. S., Nicastro, D., Gerisch, G., and Baumeister, W. (2002). Macromolecular architecture in eukaryotic cells visualized by cryoelectron tomography. Science 298, 1209–13.CrossRefGoogle ScholarPubMed
Meek, K. M., and Fullwood, N. J. (2001). Corneal and scleral collagens – a microscopist's perspective. Micron 32, 261–72.CrossRefGoogle ScholarPubMed
Meier, S. (1984). Somite formation and its relationship to metameric patterning of the mesoderm. Cell Differ. 14, 235–43.CrossRefGoogle ScholarPubMed
Meinhardt, H. (1982). Models of Biological Pattern Formation. Academic Press, New York.
Meinhardt, H. (2001). Organizer and axes formation as a self-organizing process. Int. J. Dev. Biol. 45, 177–88.Google ScholarPubMed
Meinhardt, H., and Gierer, A. (2000). Pattern formation by local self-activation and lateral inhibition. Bioessays 22, 753–60.3.0.CO;2-Z>CrossRefGoogle ScholarPubMed
Meir, E., Dassow, G., Munro, E., and Odell, G. M. (2002). Robustness, flexibility, and the role of lateral inhibition in the neurogenic network. Curr. Biol. 12, 778–86.CrossRefGoogle ScholarPubMed
Melnick, M., and Jaskoll, T. (2000). Mouse submandibular gland morphogenesis: a paradigm for embryonic signal processing. Crit. Rev. Oral Biology and Medicine 11, 199–215.CrossRefGoogle ScholarPubMed
Merks, R. M. H., Newman, S. A., and Glazier, J. A. (2004). Cell-oriented modeling of in vitro capillary development. In Cellular Automata: Proc. 6th International Conf. on Cellular Automata for Research and Industry (Sloot, P. M. A., Chopard, B., and Hoekstra, A. G., eds.), pp. 425–34. Springer-Verlag, Amsterdam, The Netherlands.CrossRefGoogle Scholar
Metropolis, N., Rosenbluth, M. N., Rosenbluth, A., Teller, H., and Teller, E. (1953). Equations of state calculations by fast computing machines. J. Chem. Phys. 21, 1087–91.CrossRefGoogle Scholar
Minelli, A. (2003). The Development of Animal Form: Ontogeny, Morphology, and Evolution. Cambridge University Press, Cambridge, New York.CrossRefGoogle Scholar
Minelli, A., and Fusco, G. (2004). Evo-devo perspectives on segmentation: model organisms, and beyond. Trends Ecol. Evol. 19, 423–9.CrossRefGoogle ScholarPubMed
Miranti, C. K., and Brugge, J. S. (2002). Sensing the environment: a historical perspective on integrin signal transduction. Nat. Cell. Biol. 4, E83–E90.CrossRefGoogle ScholarPubMed
Misteli, T. (2001). Protein dynamics: implications for nuclear architecture and gene expression. Science 291, 843–7.CrossRefGoogle ScholarPubMed
Mittenthal, J. E., and Mazo, R. M. (1983). A model for shape generation by strain and cell–cell adhesion in the epithelium of an arthropod leg segment. J. Theor. Biol. 100, 443–83.CrossRefGoogle ScholarPubMed
Miura, T., and Maini, P. K. (2004). Speed of pattern appearance in reaction–diffusion models: implications in the pattern formation of limb bud mesenchyme cells. Bull. Math. Biol. 66, 627–49.CrossRefGoogle ScholarPubMed
Miura, T., and Shiota, K. (2000a). Extracellular matrix environment influences chondrogenic pattern formation in limb bud micromass culture: experimental verification of theoretical models. Anat. Rec. 258, 100–7.3.0.CO;2-3>CrossRefGoogle Scholar
Miura, T., and Shiota, K. (2000b). TGFβ2 acts as an “activator” molecule in reaction–diffusion model and is involved in cell sorting phenomenon in mouse limb micromass culture. Dev. Dyn. 217, 241–9.3.0.CO;2-K>CrossRefGoogle Scholar
Miura, T., and Shiota, K. (2000c). Time-lapse observation of branching morphogenesis of the lung bud epithelium in mesenchyme-free culture and its relationship with the localization of actin filaments. Int. J. Dev. Biol. 44, 899–902.Google Scholar
Miura, T., Komori, M., and Shiota, K. (2000). A novel method for analysis of the periodicity of chondrogenic patterns in limb bud cell culture: correlation of in vitro pattern formation with theoretical models. Anat. Embryol. (Berlin) 201, 419–28.CrossRefGoogle ScholarPubMed
Miyazaki, S., Shirakawa, H., Nakada, K., and Honda, Y. (1993). Essential role of the inositol 1,4,5-trisphosphate receptor/Ca2+ release channel in Ca2+ waves and Ca2+ oscillations at fertilization of mammalian eggs. Dev. Biol. 158, 62–78.CrossRefGoogle ScholarPubMed
Mlodzik, M. (2002). Planar cell polarization: do the same mechanisms regulate Drosophila tissue polarity and vertebrate gastrulation? Trends Genet. 18, 564–71.CrossRefGoogle ScholarPubMed
Moftah, M. Z., Downie, S. A., Bronstein, N. B., Mezentseva, N., Pu, J., Maher, P. A., and Newman, S. A. (2002). Ectodermal FGFs induce perinodular inhibition of limb chondrogenesis in vitro and in vivo via FGF receptor 2. Dev. Biol. 249, 270–82.CrossRefGoogle ScholarPubMed
Mombach, J. C., Glazier, J. A., Raphael, R. C., and Zajac, M. (1995). Quantitative comparison between differential adhesion models and cell sorting in the presence and absence of fluctuations. Phys. Rev. Lett. 75, 2244–7.CrossRefGoogle ScholarPubMed
Monk, N. A. (2003). Oscillatory expression of Hes1, p53, and NF-kappaB driven by transcriptional time delays. Curr. Biol. 13, 1409–13.CrossRefGoogle ScholarPubMed
Montalta-He, H., and Reichert, H. (2003). Impressive expressions: developing a systematic database of gene-expression patterns in Drosophila embryogenesis. Genome Biol. 4, 205.CrossRefGoogle ScholarPubMed
Montero, J. A., and Heisenberg, C. P. (2003). Adhesive crosstalk in gastrulation. Dev. Cell 5, 190–1.CrossRefGoogle ScholarPubMed
Morisco, C., Seta, K., Hardt, S. E., Lee, Y., Vatner, S. F., and Sadoshima, J. (2001). Glycogen synthase kinase 3β regulates GATA4 in cardiac myocytes. J. Biol. Chem. 276, 28586–97.CrossRefGoogle ScholarPubMed
Morrison, S. J., Perez, S. E., Qiao, Z.et al. (2000). Transient Notch activation initiates an irreversible switch from neurogenesis to gliogenesis by neural crest stem cells. Cell 101, 499–510.CrossRefGoogle ScholarPubMed
Muratov, C. B. (1997). Synchronization, chaos, and the breakdown of the collective domain oscillations in reaction–diffusion systems. Phys. Rev. E 55, 1463–77.CrossRefGoogle Scholar
Murray, A. W., and Hunt, T. (1993). The Cell Cycle: An Introduction. W. H. Freeman, New York.Google Scholar
Murray, A. W., and Kirschner, M. W. (1989). Dominoes and clocks: the union of two views of the cell cycle. Science 246, 614–21.CrossRefGoogle ScholarPubMed
Murray, J. D. (2002). Mathematical biology. Springer, New York.Google Scholar
Nagafuchi, A., and Takeichi, M. (1988). Cell binding function of E-cadherin is regulated by the cytoplasmic domain. EMBO J. 7, 3679–84.Google ScholarPubMed
Nagar, B., Overduin, M., Ikura, M., and Rini, J. M. (1996). Structural basis of calcium-induced E-cadherin rigidification and dimerization. Nature 380, 360–4.CrossRefGoogle ScholarPubMed
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
Nakanishi, Y., Morita, T., and Nogawa, H. (1987). Cell proliferation is not required for the initiation of early cleft formation in mouse embryonic submandibular epithelium in vitro. Development 99, 429–37.Google Scholar
Nakatsuji, N., Snow, M. H., and Wylie, C. C. (1986). Cinemicrographic study of the cell movement in the primitive-streak-stage mouse embryo. J. Embryol. Exp. Morphol. 96, 99–109.Google ScholarPubMed
Nakayama, T., Yakubo, K., and Orbach, R. (1994). Dynamical properties of fractal networks: scaling, numerical simulations and physical realizations. Rev. Mod. Phys. 66, 381–443.CrossRefGoogle Scholar
Nanjundiah, V. (2005). Mathematics and biology. Current Science 88, 388–93.Google Scholar
Narbonne, G. M. (2004). Modular construction of early Ediacaran complex life forms. Science 305, 1141–4.CrossRefGoogle ScholarPubMed
Needham, D., and Hochmuth, R. M. (1992). A sensitive measure of surface stress in the resting neutrophil. Biophys. J. 61, 1664–70.CrossRefGoogle ScholarPubMed
Neff, A. W., Malacinski, G. M., Wakahara, M., and Jurand, A. (1983). Pattern formation in amphibian embryos prevented from undergoing the classical “rotation response” to egg activation. Dev. Biol. 97, 103–12.CrossRefGoogle Scholar
Neff, A. W., Wakahara, M., Jurand, A., and Malacinski, G. M. (1984). Experimental analyses of cytoplasmic rearrangements which follow fertilization and accompany symmetrization of inverted Xenopus eggs. J. Embryol. Exp. Morphol. 80, 197–224.Google ScholarPubMed
Newgreen, D. F. (1989). Physical influences on neural crest cell migration in avian embryos: contact guidance and spatial restriction. Dev. Biol. 131, 136–48.CrossRefGoogle ScholarPubMed
Newgreen, D. F., and Minichiello, J. (1995). Control of epitheliomesenchymal transformation. I. Events in the onset of neural crest cell migration are separable and inducible by protein kinase inhibitors. Dev. Biol. 170, 91–101.CrossRefGoogle ScholarPubMed
Newman, S. A. (1977). Lineage and pattern in the developing wing bud. In Vertebrate Limb and Somite Morphogenesis (Ede, D. A., Hinchliffe, J. R., and Balls, M., eds.), pp. 181–97. Cambridge University Press, Cambridge.Google Scholar
Newman, S. A. (1988). Lineage and pattern in the developing vertebrate limb. Trends Genet. 4, 329–32.CrossRefGoogle ScholarPubMed
Newman, S. A. (1993). Is segmentation generic? BioEssays 15, 277–83.CrossRefGoogle ScholarPubMed
Newman, S. A. (1994). Generic physical mechanisms of tissue morphogenesis: a common basis for development and evolution. J. Evol. Biol. 7, 467–88.CrossRefGoogle Scholar
Newman, S. A. (1995). Interplay of genetics and physical processes of tissue morphogenesis in development and evolution: the biological fifth dimension. In “Interplay of Genetic and Physical Processes in the Development of Biological Form (Beysens, D., Forgacs, G., and Gaill, F., eds.), pp. 3–12. World Scientific, Singapore.CrossRefGoogle Scholar
Newman, S. A. (1998a). Epithelial morphogenesis: a physico-evolutionary interpretation. In Molecular Basis of Epithelial Appendage Morphogenesis (Chuong, C.-M., ed.), pp. 341–58. R. G. Landes, Austin, TX.Google Scholar
Newman, S. A. (1998b). Networks of extracellular fibers and the generation of morphogenetic forces. In Dynamical Networks in Physics and Biology (Beysens, D. and Forgacs, G., eds.), pp. 139–48. Springer-Verlag, Berlin.CrossRefGoogle Scholar
Newman, S. A. (2003a). The fall and rise of systems biology. GeneWatch 16, 8–12.Google Scholar
Newman, S. A. (2003b). From physics to development: the evolution of morphogenetic mechanisms. In Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology. (Müller, G. B. and Newman, S. A., eds.), pp. 221–39. MIT Press, Cambridge, MA.Google Scholar
Newman, S. A., and Comper, W. D. (1990). “Generic” physical mechanisms of morphogenesis and pattern formation. Development 110, 1–18.Google ScholarPubMed
Newman, S. A., and Frisch, H. L. (1979). Dynamics of skeletal pattern formation in developing chick limb. Science 205, 662–8.CrossRefGoogle ScholarPubMed
Newman, S. A., and Müller, G. B. (2000). Epigenetic mechanisms of character origination. J. Exp. Zool. (Mol. Evol. Dev.) 288, 304–17.3.0.CO;2-G>CrossRefGoogle ScholarPubMed
Newman, S. A., and Müller, G. B. (eds.) (2003). Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology. MIT Press, Cambridge, MA.Google Scholar
Newman, S. A., and Tomasek, J. J. (1996). Morphogenesis of connective tissues. In Extracellular Matrix (Comper, W. D., ed.), Vol. 2, Molecular Components and Interactions, pp. 335–69. Harwood Academic Publishers, Amsterdam.Google Scholar
Newman, S. A., Frisch, H. L., Perle, M. A., and Tomasek, J. J. (1981). Limb development: aspects of differentiation, pattern formation and morphogenesis. In Morphogenesis and Pattern Formation (Connolly, T. G., Brinkley, L. L., and Carlson, B. M., eds.), pp. 163–78. Raven Press, New York.Google Scholar
Newman, S. A., Frenz, D. A., Tomasek, J. J., and Rabuzzi, D. D. (1985). Matrix-driven translocation of cells and nonliving particles. Science 228, 885–9.CrossRefGoogle ScholarPubMed
Newman, S. A., Frenz, D. A., Hasegawa, E., and Akiyama, S. K. (1987). Matrix-driven translocation: dependence on interaction of amino-terminal domain of fibronectin with heparin-like surface components of cells or particles. Proc. Nat. Acad. Sci. USA 84, 4791–5.CrossRefGoogle ScholarPubMed
Newman, S. A., Frisch, H. L., and Percus, J. K. (1988). On the stationary state analysis of reaction–diffusion mechanisms for biological pattern formationJ. Theor. Biol. 134, 183–197 (published erratum appears in J. Theor. Biol. 135, 137 (1988)).CrossRefGoogle ScholarPubMed
Newman, S. A., Cloitre, M., Allain, C., Forgacs, G., and Beysens, D. (1997). Viscosity and elasticity during collagen assembly in vitro: relevance to matrix-driven translocation. Biopolymers 41, 337–47.3.0.CO;2-T>CrossRefGoogle ScholarPubMed
Newman, S. A., Forgacs, G., Hinner, B., Maier, C. W., and Sackmann, E. (2004). Phase transformations in a model mesenchymal tissue. Phys. Biol. 1, 100–9.CrossRefGoogle Scholar
Newport, J., and Kirschner, M. (1982a). A major developmental transition in early Xenopus embryos: I. Characterization and timing of cellular changes at the midblastula stage. Cell 30, 675–86.CrossRefGoogle Scholar
Newport, J., and Kirschner, M. (1982b). A major developmental transition in early Xenopus embryos: I. Control of the onset of transcription. Cell 30, 687–96.CrossRefGoogle Scholar
Nicklas, R. B., and Koch, C. A. (1969). Chromosome micromanipulation. 3. Spindle fiber tension and the reorientation of mal-oriented chromosomes. J. Cell Biol. 43, 40–50.CrossRefGoogle ScholarPubMed
Nieuwkoop, P. D. (1969). The formation of mesoderm in Urodelean amphibians. I. Induction by the endoderm. Wilhelm Roux' Arch. Entw. Mech. Org. 162, 341–73.CrossRefGoogle ScholarPubMed
Nieuwkoop, P. D. (1973). The “organization center” of the amphibian embryo: its origin, spatial organization and morphogenetic action. Adv. Morphogen. 10, 1–39.CrossRefGoogle ScholarPubMed
Nieuwkoop, P. D. (1992). The formation of the mesoderm in Urodelean amphibians VI. The self-organizing capacity of the induced meso-endoderm. Roux's Arch. Dev. Biol. 201, 18–29.CrossRefGoogle ScholarPubMed
Noden, D. M. (1984). Craniofacial development: new views on old problems. Anat. Rec. 208, 1–13.CrossRefGoogle ScholarPubMed
Noden, D. M. (1988). Interactions and fates of avian craniofacial mesenchyme. Development Suppl. 103, 121–40.Google ScholarPubMed
Nogawa, H., and Ito, T. (1995). Branching morphogenesis of embryonic mouse lung epithelium in mesenchyme-free culture. Development 121, 1015–22.Google ScholarPubMed
Nogawa, H., and Takahashi, Y. (1991). Substitution for mesenchyme by basement-membrane-like substratum and epidermal growth factor in inducing branching morphogenesis of mouse salivary epithelium. Development 112, 855–61.Google ScholarPubMed
Nonaka, S., Tanaka, Y., Okada, Y., et al. (1998). Randomization of left–right asymmetry due to loss of nodal cilia generating leftward flow of extraembryonic fluid in mice lacking KIF3B motor protein. Cell 95, 829–37.CrossRefGoogle ScholarPubMed
Nonaka, S., Shiratori, H., Saijoh, Y., and Hamada, H. (2002). Determination of left–right patterning of the mouse embryo by artificial nodal flow. Nature 418, 96–9.CrossRefGoogle ScholarPubMed
Norel, R., and Agur, Z. (1991). A model for the adjustment of the mitotic clock by cyclin and MPF levels. Science 251, 1076–8.CrossRefGoogle ScholarPubMed
Novak, B., and Tyson, J. J. (1993). Numerical analysis of a comprehensive model of M-phase control in Xenopus oocyte extracts and intact embryos. J. Cell Sci. 106, 1153–68.Google ScholarPubMed
Novak, B., Csikasz-Nagy, A., Gyorffy, B., Nasmyth, K., and Tyson, J. J. (1998). Model scenarios for evolution of the eukaryotic cell cycle. Philos. Trans. R. Soc. Lond. B Biol. Sci. 353, 2063–76.CrossRefGoogle ScholarPubMed
Novak, B., Toth, A., Csikasz-Nagy, A., Gyorffy, B., Tyson, J. J., and Nasmyth, K. (1999). Finishing the cell cycle. J. Theor. Biol. 199, 223–33.CrossRefGoogle ScholarPubMed
Oates, A. C., and Ho, R. K. (2002). Hairy/E (spl)-related (Her) genes are central components of the segmentation oscillator and display redundancy with the Delta/Notch signaling pathway in the formation of anterior segmental boundaries in the zebrafish. Development 129, 2929–46.Google ScholarPubMed
Oberlender, S. A., and Tuan, R. S. (1994). Expression and functional involvement of N-cadherin in embryonic limb chondrogenesis. Development 120, 177–87.Google ScholarPubMed
Oda, S., Deguchi, R., Mohri, T., Shikano, T., Nakanishi, S., and Miyazaki, S. (1999). Spatiotemporal dynamics of the [Ca2+]i rise induced by microinjection of sperm extract into mouse eggs: preferential induction of a Ca2+ wave from the cortex mediated by the inositol 1,4,5-trisphosphate receptor. Dev. Biol. 209, 172–85.CrossRefGoogle Scholar
Odell, G. M., Oster, G., Alberch, P., and Burnside, B. (1981). The mechanical basis of morphogenesis. I. Epithelial folding and invagination. Dev. Biol. 85, 446–62.CrossRefGoogle ScholarPubMed
Oheim, M., and Stuhmer, W. (2000). Tracking chromaffin granules on their way through the actin cortex. Eur. Biophys. J. 29, 67–89.CrossRefGoogle ScholarPubMed
Oliveri, P., and Davidson, E. H. (2004). Gene regulatory network controlling embryonic specification in the sea urchin. Curr. Opin. Genet. Dev. 14, 351–60.CrossRefGoogle ScholarPubMed
Oliveri, P., Carrick, D. M., and Davidson, E. H. (2002). A regulatory gene network that directs micromere specification in the sea urchin embryo. Dev. Biol. 246, 209–28.CrossRefGoogle ScholarPubMed
Opas, M., Davies, J. R., Zhou, Y., and Dziak, E. (2001). Formation of retinal pigment epithelium in vitro by transdifferentiation of neural retina cells. Int. J. Dev. Biol. 45, 633–42.Google ScholarPubMed
Ornitz, D. M., and Marie, P. J. (2002). FGF signaling pathways in endochondral and intramembranous bone development and human genetic disease. Genes Dev. 16, 1446–65.CrossRefGoogle ScholarPubMed
Ouyang, Q., and Swinney, H. (1991). Transition from a uniform state to hexagonal and striped Turing patterns. Nature 352, 610–12.CrossRefGoogle Scholar
Owen, M. R., andSherratt, J. A. (1998). Mathematical modelling of juxtacrine cell signalling. Math. Biosci. 153, 125–50.CrossRefGoogle ScholarPubMed
Owen, M. R., Sherratt, J. A., and Myers, S. R. (1999). How far can a juxtacrine signal travel? Proc. R. Soc. Lond. B Biol. Sci. 266, 579–85.CrossRefGoogle ScholarPubMed
Owen, M. R., Sherratt, J. A., and Wearing, H. J. (2000). Lateral induction by juxtacrine signaling is a new mechanism for pattern formation. Dev. Biol. 217, 54–61.CrossRefGoogle ScholarPubMed
Ozdamar, B., Bose, R., Barrios-Rodiles, M., Wang, H.-R., Zhang, Y., and Wrana, J. L. (2005). Regulation of the polarity protein Par6 by TGFβ receptors controls epithelial cell plasticity. Science 307, 1603–9.CrossRefGoogle ScholarPubMed
Pagan-Westphal, S. M., and Tabin, C. J. (1998). The transfer of left–right positional information during chick embryogenesis. Cell 93, 25–35.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
Pardanaud, L., and Dieterlen-Lievre, F. (2000). Ontogeny of the endothelial system in the avian model. Adv. Exp. Med. Biol. 476, 67–78.CrossRefGoogle ScholarPubMed
Pardanaud, L., Luton, D., Prigent, M., Bourcheix, L. M., Catala, M., and Dieterlen-Lievre, F. (1996). Two distinct endothelial lineages in ontogeny, one of them related to hemopoiesis. Development 122, 1363–71.Google ScholarPubMed
Parkinson, J., Kadler, K. E., and Brass, A. (1995). Simple physical model of collagen fibrillogenesis based on diffusion limited aggregation. J. Mol. Biol. 247, 823–31.CrossRefGoogle ScholarPubMed
Patan, S. (2000). Vasculogenesis and angiogenesis as mechanisms of vascular network formation, growth and remodeling. J. Neurooncol. 50, 1–15.CrossRefGoogle ScholarPubMed
Patel, N. H. (1994). Developmental evolution: insights from studies of insect segmentation. Science 266, 581–90.CrossRefGoogle ScholarPubMed
Patel, N. H., Kornberg, T. B., and Goodman, C. S. (1989). Expression of engrailed during segmentation in grasshopper and crayfish. Development 107, 201–12.Google ScholarPubMed
Patel, N. H., Ball, E. E., and Goodman, C. S. (1992). Changing role of even-skipped during the evolution of insect pattern formation. Nature 357, 339–42.CrossRefGoogle ScholarPubMed
Patel, N. H., Condron, B. G., and Zinn, K. (1994). Pair-rule expression patterns of even-skipped are found in both short- and long-germ beetles. Nature 367, 429–34.CrossRefGoogle Scholar
Pearson, J. E., and Ponce-Dawson, S. (1998). Crisis on skid row. Physica A 257, 141–48.CrossRefGoogle Scholar
Pecht, I., and Lancet, D. (1976). In Chemical Relaxation in Molecular Biology (Riegler, R. and Pecht, I., eds.) Springer-Verlag, Heidelberg.Google Scholar
Perrin, F. (1934). Mouvement Brownien d'un ellipsoide (I). Dispersion dielectrique pour des molecules ellipsoidales. J. Physique et la Radium Serie 7, 5, 497–511.CrossRefGoogle Scholar
Perrin, F. (1936). Mouvement Brownien d'un ellipsoide (II). Rotation libre et depolarisation des fluorescences. Translation et diffusion des molecules ellipsoidales. J. Physique et la Radium Serie 7, 7, 1–11.CrossRefGoogle Scholar
Perris, R., and Perissonotto, D. (2000). Role of the extracellular matrix during neural crest migration. Mech. Dev. 95, 3–21.CrossRefGoogle Scholar
Peters, K. G., Werner, S., Chen, G., and Williams, L. T. (1992). Two FGF receptor genes are differentially expressed in epithelial and mesenchymal tissues during limb formation and organogenesis in the mouse. Development 114, 233–43.Google ScholarPubMed
Phillips, H. M. (1969). Equilibrium measurements of embryonic cell adhesiveness: physical formulation and testing of the differential adhesion hypothesis. Ph.D. thesis, Johns Hopkins University, Baltimore.Google Scholar
Picton, H., Briggs, D., and Gosden, R. (1998). The molecular basis of oocyte growth and development. Mol. Cell. Endocrinol. 145, 27–37.CrossRefGoogle ScholarPubMed
Pollack, G. H. (2001). Is the gel a gel – and why does it matter? Jpn J. Physiol. 51, 649–60.CrossRefGoogle ScholarPubMed
Pourquié, O. (2000). Skin development: delta laid bare. Curr. Biol. 10, R425–8.CrossRefGoogle ScholarPubMed
Pourquié, O. (2001). Vertebrate somitogenesis. Ann. Rev. Cell Dev. Biol. 17, 311–50.CrossRefGoogle ScholarPubMed
Pourquié, O. (2003). The segmentation clock: converting embryonic time into spatial pattern. Science 301, 328–30.CrossRefGoogle ScholarPubMed
Pourquié, O., and Goldbeter, A. (2003). Segmentation clock: insights from computational models. Curr. Biol. 13, R632–4.CrossRefGoogle ScholarPubMed
Preston, B. N., Laurent, T. C., Comper, W. D., and Checkley, G. J. (1980). Rapid polymer transport in concentrated solutions through the formation of ordered structures. Nature 287, 499–503.CrossRefGoogle Scholar
Primmett, D. R., Norris, W. E., Carlson, G. J., Keynes, R. J., and Stern, C. D. (1989). Periodic segmental anomalies induced by heat shock in the chick embryo are associated with the cell cycle. Development 105, 119–30.Google ScholarPubMed
Purcell, E. M. (1977). Life at low Reynolds number. Am. J. Phys. 45, 3–11.CrossRefGoogle Scholar
Quaas, J., and Wylie, C. (2002). Surface contraction waves (SCWs) in the Xenopus egg are required for the localization of the germ plasm and are dependent upon maternal stores of the kinesin-like protein Xklp1. Dev. Biol. 243, 272–80.CrossRefGoogle ScholarPubMed
Raff, R. A. (1996). The Shape of Life: Genes, Development, and the Evolution of Animal Form. University of Chicago Press, Chicago.Google Scholar
Ramirez-Weber, F. A., and Kornberg, T. B. (1999). Cytonemes: cellular processes that project to the principal signaling center in Drosophila imaginal discs. Cell 97, 599–607.CrossRefGoogle ScholarPubMed
Rappaport, R. (1966). Experiments concerning the cleavage furrow in invertebrate eggs. J. Exp. Zool. 161, 1–8.CrossRefGoogle ScholarPubMed
Rappaport, R. (1986). Establishment of the mechanism of cytokinesis in animal cells. Int. Rev. Cytol. 105, 245–81.CrossRefGoogle ScholarPubMed
Rappaport, R. (1996). Cytokinesis in Animal Cells. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Rashevsky, N. (1960). Mathematical Biophysics: Physico-mathematical Foundations of Biology, Vol. 1, Dover, New York.Google Scholar
Raya, A., Kawakami, Y., Rodriguez-Esteban, C., et al.(2003). Notch activity induces Nodal expression and mediates the establishment of left–right asymmetry in vertebrate embryos. Genes Dev. 17, 1213–8.CrossRefGoogle ScholarPubMed
Raya, A., Kawakami, Y., Rodriguez-Esteban, C., et al. (2004). Notch activity acts as a sensor for extracellular calcium during vertebrate left–right determination. Nature 427, 121–8.CrossRefGoogle ScholarPubMed
Redner, S. (2001). A Guide to First Passage Processes.Cambridge University Press, Cambridge.
Reilly, K. M., and Melton, D. A. (1996). Short-range signaling by candidate morphogens of the TGF beta family and evidence for a relay mechanism of induction. Cell 86, 743–54.CrossRefGoogle ScholarPubMed
Reinitz, J., Mjolsness, E., and Sharp, D. H. (1995). Model for cooperative control of positional information in Drosophila by bicoid and maternal hunchback. J. Exp. Zool. 271, 47–56.CrossRefGoogle ScholarPubMed
Resnick, N., Collins, T., Atkinson, W., Bonthron, D. T., Dewey, C. F. Jr, and Gimbrone, M. A. Jr (1993). Platelet-derived growth factor B chain promoter contains a cis-acting fluid shear-stress-responsive element. Proc. Nat. Acad. Sci. USA 90, 4591–5 (published erratum appears in Proc. Nat. Acad. Sci. USA 90, 7908 (1993)).CrossRefGoogle ScholarPubMed
Riedl, R. (1977). A systems-analytical approach to macro-evolutionary phenomena. Q. Rev. Biol. 52, 351–70.CrossRefGoogle ScholarPubMed
Rieu, J. P., Kataoka, N., and Sawada, Y. (1998). Quantitative analysis of cell motion during sorting in 2D aggregates of dissociated hydra cells. Phys. Rev. E 57, 924–31.CrossRefGoogle Scholar
Rieu, J. P., Upadhyaya, A., Glazier, J. A., Ouchi, N. B., and Sawada, Y. (2000). Diffusion and deformations of single hydra cells in cellular aggregates. Biophys. J. 79, 1903–14.CrossRefGoogle ScholarPubMed
Robinson, E. E., Zazzali, K. M., Corbett, S. A., and Foty, R. A. (2003). Alpha5beta1 integrin mediates strong tissue cohesion. J. Cell. Sci. 116, 377–86.CrossRefGoogle ScholarPubMed
Rodriguez, I., and Basler, K. (1997). Control of compartmental affinity boundaries by hedgehog. Nature 389, 614–8.Google ScholarPubMed
Roegiers, F., McDougall, A., and Sardet, C. (1995). The sperm entry point defines the orientation of the calcium-induced contraction wave that directs the first phase of cytoplasmic reorganization in the ascidian egg. Development 121, 3457–66.Google ScholarPubMed
Roegiers, F., Djediat, C., Dumollard, R., Rouviere, C., and Sardet, C. (1999). Phases of cytoplasmic and cortical reorganizations of the ascidian zygote between fertilization and first division. Development 126, 3101–17.Google ScholarPubMed
Rooke, J. E., and Xu, T. (1998). Positive and negative signals between interacting cells for establishing neural fate. Bioessays 20, 209–14.3.0.CO;2-M>CrossRefGoogle ScholarPubMed
Ross, M. H., Kaye, G. I., and Pawlina, W. (2003). Histology: A Text and Atlas. Lippincott Williams & Wilkins, Philadelphia, PA.Google Scholar
Roy, P., Petroll, W. M., Cavanagh, H. D., and Jester, J. V. (1999). Exertion of tractional force requires the coordinated up-regulation of cell contractility and adhesion. Cell Motil. Cytoskeleton 43, 23–34.3.0.CO;2-M>CrossRefGoogle ScholarPubMed
Rubinstein, M., Colby, R. H., and Gillmor, J. R. (1989). Dynamic scaling for polymer gelation. Polym. Prepr. Am. Chem. Soc. Div. Polym. Chem. 30, 81–2.Google Scholar
Rudnick, J., and Gaspari, G. (2004). Elements of Random Walk. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Runft, L. L., Jaffe, L. A., and Mehlmann, L. M. (2002). Egg activation at fertilization: where it all begins. Dev. Biol. 245, 237–54.CrossRefGoogle ScholarPubMed
Rupp, R. A., Singhal, N., and Veenstra, G. J. (2002). When the embryonic genome flexes its muscles. Eur. J. Biochem 269, 2294–9.CrossRefGoogle ScholarPubMed
Rupp, P. A., Czirok, A., and Little, C. D. (2003). Novel approaches for the study of vascular assembly and morphogenesis in avian embryos. Trends Cardiovasc Med. 13, 283–8.CrossRefGoogle Scholar
Rustom, A., Saffrich, R., Markovic, I., Walther, P., and Gerdes, H. H. (2004). Nanotubular highways for intercellular organelle transport. Science 303, 1007–10.CrossRefGoogle ScholarPubMed
Ryan, P. L., Foty, R. A., Kohn, J., and Steinberg, M. S. (2001). Tissue spreading on implantable substrates is a competitive outcome of cell–cell vs. cell–substratum adhesivity. Proc. Nat. Acad. Sci. USA 98, 4323–7.CrossRefGoogle ScholarPubMed
Sahimi, M. (1994). Applications of Percolation Theory. Taylor & Francis, London.Google Scholar
Sakai, T., Larsen, M., and Yamada, K. M. (2003). Fibronectin requirement in branching morphogenesis. Nature 423, 876–81.CrossRefGoogle ScholarPubMed
Sakuma, R., Yi, Ohnishi Y., Meno, C.et al. (2002). Inhibition of Nodal signalling by Lefty mediated through interaction with common receptors and efficient diffusion. Genes Cells 7, 401–12.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I., and Jernvall, J. (2002). A gene network model accounting for development and evolution of mammalian teeth. Proc. Nat. Acad. Sci. USA 99, 8116–20.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I., Garcia-Fernandez, J., and Solé, R. V. (2000). Gene networks capable of pattern formation: from induction to reaction–diffusion. J. Theor. Biol. 205, 587–603.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I., Newman, S. A., and Solé, R. (2001a). Phenotypic and dynamical transitions in model genetic networks. I. Emergence of patterns and genotype–phenotype relationships. Evolution & Development 3, 84–94.CrossRefGoogle Scholar
Salazar-Ciudad, I., Solé, R., and Newman, S. A. (2001b). Phenotypic and dynamical transitions in model genetic networks. II. Application to the evolution of segmentation mechanisms. Evolution & Development 3, 95–103.CrossRefGoogle Scholar
Salazar-Ciudad, I., Jernvall, J., and Newman, S. A. (2003). Mechanisms of pattern formation in development and evolution. Development 130, 2027–37.CrossRefGoogle ScholarPubMed
Salthe, S. N. (1993). Development and Evolution: Complexity and Change in Biology. MIT Press, Cambridge, MA.Google Scholar
Sanchez, L., and Thieffry, D. (2001). A logical analysis of the Drosophila gap-gene system. J. Theor. Biol. 211, 115–41.CrossRefGoogle ScholarPubMed
Sanders, E. J. (1991). Embryonic cell invasiveness: an in vitro study of chick gastrulation. J. Cell Sci. 98, 403–7.Google Scholar
Sardet, C., Roegiers, F., Dumollard, R., Rouviere, C., and McDougall, A. (1998). Calcium waves and oscillations in eggs. Biophys. Chem. 72, 131–40.CrossRefGoogle ScholarPubMed
Sardet, C., Prodon, F., Dumollard, R., Chang, P., and Chenevert, J. (2002). Structure and function of the egg cortex from oogenesis through fertilization. Dev. Biol. 241, 1–23.CrossRefGoogle ScholarPubMed
Sato, Y., Yasuda, K., and Takahashi, Y. (2002). Morphological boundary forms by a novel inductive event mediated by Lunatic fringe and Notch during somitic segmentation. Development 129, 3633–44.Google ScholarPubMed
Saunders, J. W. Jr (1948). The proximo-distal sequence of origin of the parts of the chick wing and the role of the ectoderm. J. Exp. Zool. 108, 363–402.CrossRefGoogle ScholarPubMed
Savill, N. J., and Sherratt, J. A. (2003). Control of epidermal stem cell clusters by Notch-mediated lateral induction. Dev. Biol. 258, 141–53.CrossRefGoogle ScholarPubMed
Saxton, M. J., and Jacobson, K. (1997). Single-particle tracking: applications to membrane dynamics. Ann. Rev. Biophys. Biomol. Struct. 26, 373–99.CrossRefGoogle ScholarPubMed
Schier, A. F., and Gehring, W. J. (1993). Analysis of a fushi tarazu autoregulatory element: multiple sequence elements contribute to enhancer activity. EMBO J. 12, 1111–9.Google ScholarPubMed
Schlosser, G. and Wagner, G. P. (eds.) (2004). Modularity in Development and Evolution. University of Chicago Press, Chicago.Google Scholar
Schmalhausen, I. I. (1949). Factors of Evolution. Blakiston, Philadelphia.Google Scholar
Schroeder, T. E. (1975). Dynamics of the contractile ring. Soc. Gen. Physiol. Ser. 30, 305–34.Google ScholarPubMed
Schulte-Merker, S., and Smith, J. C. (1995). Mesoderm formation in response to Brachyury requires FGF signalling. Curr. Biol. 5, 62–7.CrossRefGoogle ScholarPubMed
Schuster, S., Marhl, M., and Hofer, T. (2002). Modelling of simple and complex calcium oscillations. From single-cell responses to intercellular signalling. Eur. J. Biochem. 269, 1333–55.CrossRefGoogle ScholarPubMed
Seifert, U. (1997). Configurations of fluid membranes and vesicles. Adv. Phys. 46, 13–137.CrossRefGoogle Scholar
Seilacher, A. (1992). Vendobionta and Psammocorallia – lost constructions of precambrian evolution. J. Geolog. Soc. Lond. 149, 607–13.CrossRefGoogle Scholar
Serini, G., Ambrosi, D., Giraudo, E., Gamba, A., Preziosi, L., and Bussolino, F. (2003). Modeling the early stages of vascular network assembly. EMBO J. 22, 1771–9.CrossRefGoogle ScholarPubMed
Shafrir, Y., and Forgacs, G. (2002). Mechanotransduction through the cytoskeleton. Am. J. Physiol. Cell Physiol. 282, C479–86.CrossRefGoogle ScholarPubMed
Shafrir, Y., ben-Avraham, D., and Forgacs, G. (2000). Trafficking and signaling through the cytoskeleton: a specific mechanism. J. Cell Sci. 113, 2747–57.Google ScholarPubMed
Shapiro, L., Fannon, A. M., Kwong, P. D., et al. (1995). Structural basis of cell–cell adhesion by cadherins. Nature 374, 327–37.CrossRefGoogle ScholarPubMed
Shav-Tal, Y., Darzacq, X., Shenoy, S. M., et al. (2004). Dynamics of single mRNPs in nuclei of living cells. Science 304, 1797–800.CrossRefGoogle ScholarPubMed
Sheetz, M. P. (2001). Cell control by membrane-cytoskeleton adhesion. Nat. Rev. Mol. Cell Biol. 2, 392–6.CrossRefGoogle ScholarPubMed
Sherwood, D. R., and McClay, D. R. (2001). LvNotch signaling plays a dual role in regulating the position of the ectoderm–endoderm boundary in the sea urchin embryo. Development 128, 2221–32.Google Scholar
Shih, J., and Keller, R. (1992). Cell motility driving mediolateral intercalation in explants of Xenopus laevis. Development 116, 901–14.Google ScholarPubMed
Shyy, J. Y., Li, Y. S., Lin, M. C., Chen, W., Yuan, S., Usami, S., and Chien, S. (1995a). Multiple cis-elements mediate shear stress-induced gene expression. J. Biomech. 28, 1451–7.CrossRefGoogle Scholar
Shyy, J. Y., Lin, M. C., Han, J., Lu, Y., Petrime, M., and Chien, S. (1995b). The cis-acting phorbol ester “12-O-tetradecanoylphorbol 13-acetate”- responsive element is involved in shear stress-induced monocyte chemotactic protein 1 gene expression. Proc. Nat. Acad. Sci. USA 92, 8069–73.CrossRefGoogle Scholar
Simonneau, L., Kitagawa, M., Suzuki, S., and Thiery, J. P. (1995). Cadherin 11 expression marks the mesenchymal phenotype: towards new functions for cadherins? Cell Adhes. Commun. 3, 115–30.CrossRefGoogle ScholarPubMed
Singer, S. J., and Nicolson, G. L. (1972). The fluid mosaic model of the structure of cell membranes. Science 175, 720–31.CrossRefGoogle ScholarPubMed
Sivasankar, S., Brieher, W., Lavrik, N., Gumbiner, B., and Leckband, D. (1999). Direct molecular force measurements of multiple adhesive interactions between cadherin ectodomains. Proc. Nat. Acad. Sci. USA 96, 11820–4.CrossRefGoogle ScholarPubMed
Sivasankar, S., Gumbiner, B., and Leckband, D. (2001). Direct measurements of multiple adhesive alignments and unbinding trajectories between cadherin extracellular domains. Biophys. J. 80, 1758–68.CrossRefGoogle ScholarPubMed
Small, S., Kraut, R., Hoey, T., Warrior, R., and Levine, M. (1991). Transcriptional regulation of a pair-rule stripe in Drosophila. Genes Dev. 5, 827–39.CrossRefGoogle ScholarPubMed
Small, S., Blair, A., and Levine, M. (1992). Regulation of even-skipped stripe 2 in the Drosophila embryo. EMBO J. 11, 4047–57.Google ScholarPubMed
Small, S., Blair, A., and Levine, M. (1996). Regulation of two pair-rule stripes by a single enhancer in the Drosophila embryo. Dev. Biol. 175, 314–24.CrossRefGoogle ScholarPubMed
Solnica-Krezel, L. (2003). Vertebrate development: taming the nodal waves. Curr. Biol. 13, R7–R9.CrossRefGoogle ScholarPubMed
Spehr, M., Gisselmann, G., Poplawski, A., et al. (2003). Identification of a testicular odorant receptor mediating human sperm chemotaxis. Science 299, 2054–8.CrossRefGoogle ScholarPubMed
Spemann, H., and Mangold, H. (1924). Über Induktion von Embryonalanlagen durch Implantation artfremder Organisatoren. Wilhelm Roux' Arch. Entw. Mech. Org. 100, 599–638.Google Scholar
Spooner, B. S., and Wessells, N. K. (1972). An analysis of salivary gland morphogenesis: role of cytoplasmic microfilaments and microtubules. Dev. Biol. 27, 38–54.CrossRefGoogle ScholarPubMed
St Johnston, D., and Nusslein-Volhard, C. (1992). The origin of pattern and polarity in the Drosophila embryo. Cell 68, 201–19.CrossRefGoogle ScholarPubMed
Standley, H. J., Zorn, A. M., and Gurdon, J. B. (2001). eFGF and its mode of action in the community effect during Xenopus myogenesis. Development 128, 1347–57.Google ScholarPubMed
Starz-Gaiano, M., and Lehmann, R. (2001). Moving towards the next generation. Mech. Dev. 105, 5–18.CrossRefGoogle ScholarPubMed
Stauber, M., Jackle, H., and Schmidt-Ott, U. (1999). The anterior determinant bicoid of Drosophila is a derived Hox class 3 gene. Proc. Nat. Acad. Sci. USA 96, 3786–9.CrossRefGoogle ScholarPubMed
Steinberg, M. S. (1963). Reconstruction of tissues by dissociated cells. Some morphogenetic tissue movements and the sorting out of embryonic cells may have a common explanation. Science 141, 401–8.CrossRefGoogle ScholarPubMed
Steinberg, M. S. (1978). Specific cell ligands and the differential adhesion hypothesis: how do they fit together? In Specificity of Embryological Interactions (Garrod, D. R., ed.), pp. 97–130. Chapman and Hall, London.CrossRefGoogle Scholar
Steinberg, M. S. (1998). Goal-directedness in embryonic development. Integrative Biology 1, 49–59.3.0.CO;2-Z>CrossRefGoogle Scholar
Steinberg, M. S., and Foty, R. A. (1997). Intercellular adhesions as determinants of tissue assembly and malignant invasion. J. Cell Physiol. 173, 135–9.3.0.CO;2-F>CrossRefGoogle ScholarPubMed
Steinberg, M. S., and Poole, T. J. (1982). Liquid behavior of embryonic tissues. In Cell Behavior (Bellairs, R. and Curtis, A. S. G., eds.), pp. 583–607. Cambridge University Press,Cambridge.Google Scholar
Steinberg, M. S., and Takeichi, M. (1994). Experimental specification of cell sorting, tissue spreading, and specific spatial patterning by quantitative differences in cadherin expression. Proc. Nat. Acad. Sci. USA 91, 206–9.CrossRefGoogle ScholarPubMed
Stern, C. D., and Bellairs, R. (1984). Mitotic activity during somite segmentation in the early chick embryo. Anat. Embryol. (Berlin) 169, 97–102.CrossRefGoogle ScholarPubMed
Stollewerk, A., Schoppmeier, M., and Damen, W. G. (2003). Involvement of Notch and Delta genes in spider segmentation. Nature 423, 863–5.CrossRefGoogle ScholarPubMed
Stopak, D., and Harris, A. K. (1982). Connective tissue morphogenesis by fibroblast traction. I. Tissue culture observations. Dev. Biol. 90, 383–398.CrossRefGoogle ScholarPubMed
Stossel, T. P. (2001). Manifesto for a cytoplasmic revolution. Science 293, 611.CrossRefGoogle Scholar
Stricker, S. A. (1999). Comparative biology of calcium signaling during fertilization and egg activation in animals. Dev. Biol. 211, 157–76.CrossRefGoogle ScholarPubMed
Strogatz, S. H. (1994). Nonlinear Dynamics and Chaos: With Applications to Physics, Biology, Chemistry, and Engineering. Perseus, Cambridge, MA.Google Scholar
Strogatz, S. H. (2003). Sync: The Emerging Science of Spontaneous Order. Theia, New York.Google Scholar
Subtelny, S., and Penkala, J. E. (1984). Experimental evidence for a morphogenetic role in the emergence of primordial germ cells from the endoderm in Rana pipiens. Differentiation 26, 211–19.CrossRefGoogle Scholar
Sun, B., Bush, S., Collins-Racie, L., et al. (1999). derriere: a TGF-beta family member required for posterior development in Xenopus. Development 126, 1467–82.Google ScholarPubMed
Sun, Q. Y. (2003). Cellular and molecular mechanisms leading to cortical reaction and polyspermy block in mammalian eggs. Microsc. Res. Tech. 61, 342–8.CrossRefGoogle ScholarPubMed
Supp, D. M., Witte, D. P., Potter, S. S., and Brueckner, M. (1997). Mutation of an axonemal dynein affects left–right asymmetry in inversus viscerum mice. Nature 389, 963–6.CrossRefGoogle ScholarPubMed
Supp, D. M., Potter, S. S., and Brueckner, M. (2000). Molecular motors: the driving force behind mammalian left–right development. Trends Cell Biol. 10, 41–5.CrossRefGoogle ScholarPubMed
Suzuki, K., Tanaka, Y., Nakajima, Y., et al. (1995). Spatiotemporal relationships among early events of fertilization in sea urchin eggs revealed by multiview microscopy. Biophys. J. 68, 739–48.CrossRefGoogle ScholarPubMed
Sveiczer, A., Csikasz-Nagy, A., Gyorffy, B., Tyson, J. J., and Novak, B. (2000). Modeling the fission yeast cell cycle: quantized cycle times in wee1– cdc25Delta mutant cells. Proc. Nat. Acad. Sci. USA 97, 7865–70.CrossRefGoogle ScholarPubMed
Szebenyi, G., Savage, M. P., Olwin, B. B., and Fallon, J. F. (1995). Changes in the expression of fibroblast growth factor receptors mark distinct stages of chondrogenesis in vitro and during chick limb skeletal patterning. Dev. Dyn. 204, 446–56.CrossRefGoogle ScholarPubMed
Takahashi, Y., and Nogawa, H. (1991). Branching morphogenesis of mouse salivary epithelium in basement membrane-like substratum separated from mesenchyme by the membrane filter. Development 111, 327–35.Google ScholarPubMed
Takeichi, M. (1991). Cadherin cell adhesion receptors as a morphogenetic regulator. Science 251, 1451–5.CrossRefGoogle ScholarPubMed
Takeichi, M. (1995). Morphogenetic roles of classic cadherins. Curr. Opin. Cell Biol. 7, 619–27.CrossRefGoogle ScholarPubMed
Takeichi, T., and Kubota, H. Y. (1984). Structural basis of the activation wave in the egg of Xenopus laevis. J. Embryol. Exp. Morphol. 81, 1–16.Google ScholarPubMed
Takke, C., and Campos-Ortega, J. A. (1999). her1, a zebrafish pair-rule like gene, acts downstream of notch signalling to control somite development. Development 126, 3005–14.Google ScholarPubMed
Tanaka, M., and Tickle, C. (2004). Tbx18 and boundary formation in chick somite and wing development. Dev. Biol. 268, 470–80.CrossRefGoogle ScholarPubMed
Taunton, J., Rowning, B. A., Coughlin, M. L., et al. (2000). Actin-dependent propulsion of endosomes and lysosomes by recruitment of N-WASP. J. Cell Biol. 148, 519–30.CrossRefGoogle ScholarPubMed
Tautz, D. (1992). Redundancies, development and the flow of information. Bioessays 14, 263–6.CrossRefGoogle Scholar
Technau, U., and Holstein, T. W. (1992). Cell sorting during the regeneration of Hydra from reaggregated cells. Dev. Biol. 151, 117–27.CrossRefGoogle ScholarPubMed
Teleman, A. A., and Cohen, S. M. (2000). Dpp gradient formation in the Drosophila wing imaginal disc. Cell 103, 971–80.CrossRefGoogle ScholarPubMed
Teleman, A. A., Strigini, M., and Cohen, S. M. (2001). Shaping morphogen gradients. Cell 105, 559–62.CrossRefGoogle ScholarPubMed
Terasaki, M. (1996). Actin filament translocations in sea urchin eggs. Cell Motil. Cytoskeleton 34, 48–56.3.0.CO;2-E>CrossRefGoogle ScholarPubMed
Thiebaud, C. H. (1979). Quantitative determination of amplified rDNA and its distribution during oogenesis in Xenopus laevis. Chromosoma 73, 37–44.CrossRefGoogle ScholarPubMed
Thompson, D. W. (1917). On Growth and Form.Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Tickle, C. (2003). Patterning systems – from one end of the limb to the other. Dev. Cell 4, 449–58.CrossRefGoogle ScholarPubMed
Tiedemann, H., Asashima, M., Grunz, H., and Knochel, W. (1998). Neural induction in embryos. Dev. Growth Differ. 40, 363–76.CrossRefGoogle ScholarPubMed
Tomasek, J. J., Mazurkiewicz, J. E., and Newman, S. A. (1982). Nonuniform distribution of fibronectin during avian limb development. Dev. Biol. 90, 118–126.CrossRefGoogle ScholarPubMed
Torza, S., and Mason, S. G. (1969). Coalescence of two immiscible liquid drops. Science 163, 813–14.CrossRefGoogle ScholarPubMed
Townes, P. L., and Holtfreter, J. (1955). Directed movements and selective adhesion of embryonic amphibian cells. J. Exp. Zool. 128, 53–120.CrossRefGoogle Scholar
Tritton, D. J. (1988). Physical Fluid Dynamics.Clarendon Press, Oxford.Google Scholar
Tsai, M. A., Frank, R. S., and Waugh, R. E. (1994). Passive mechanical behavior of human neutrophils: effect of cytochalasin B. Biophys. J. 66, 2166–72.CrossRefGoogle ScholarPubMed
Tsai, M. A., Waugh, R. E., and Keng, P. C. (1998). Passive mechanical behavior of human neutrophils: effects of colchicine and paclitaxel. Biophys. J. 74, 3282–91.CrossRefGoogle ScholarPubMed
Tsarfaty, I., Resau, J. H., Rulong, S., et al. (1992). The met proto-oncogene receptor and lumen formation. Science 257, 1258–61.CrossRefGoogle ScholarPubMed
Tsarfaty, I., Rong, S., Resau, J. H., et al. (1994). The Met proto-oncogene mesenchymal to epithelial cell conversion. Science 263, 98–101.CrossRefGoogle ScholarPubMed
Tseng, Y., Kole, T. P., and Wirtz, D. (2002). Micromechanical mapping of live cells by multiple-particle-tracking microrheology. Biophys. J. 83, 3162–76.CrossRefGoogle ScholarPubMed
Tsonis, P. A., Del Rio-Tsonis, K., Millan, J. L., and Wheelock, M. J. (1994). Expression of N-cadherin and alkaline phosphatase in chick limb bud mesenchymal cells: regulation by 1, 25-dihydroxyvitamin D3 or TGF-beta 1. Exp. Cell. Res. 213, 433–7.CrossRefGoogle ScholarPubMed
Turing, A. (1952). The chemical basis of morphogenesis. Phil. Trans. Roy. Soc. Lond. B 237, 37–72.CrossRefGoogle Scholar
Tyson, J. J. (1991). Modeling the cell division cycle: cdc2 and cyclin interactions. Proc. Nat. Acad. Sci. USA 88, 7328–32.CrossRefGoogle ScholarPubMed
Tyson, J. J., and Novak, B. (2001). Regulation of the eukaryotic cell cycle: molecular antagonism, hysteresis, and irreversible transitions. J. Theor. Biol. 210, 249–63.CrossRefGoogle ScholarPubMed
Umbhauer, M., Boucaut, J. C., and Shi, D. L. (2001). Repression of XMyoD expression and myogenesis by Xhairy-1 in Xenopus early embryo. Mech. Dev. 109, 61–8.CrossRefGoogle ScholarPubMed
Ushiki, T. (2002). Collagen fibers, reticular fibers and elastic fibers. A comprehensive understanding from a morphological viewpoint. Arch Histol. Cytol. 65, 109–26.CrossRefGoogle ScholarPubMed
Uyttendaele, H., Ho, J., Rossant, J., and Kitajewski, J. (2001). Vascular patterning defects associated with expression of activated Notch4 in embryonic endothelium. Proc. Nat. Acad. Sci. USA 98, 5643–8.CrossRefGoogle ScholarPubMed
Valberg, P. A., and Feldman, H. A. (1987). Magnetic particle motions within living cells. Measurement of cytoplasmic viscosity and motile activity. Biophys. J. 52, 551–61.CrossRefGoogle ScholarPubMed
Valentine, J. W. (2004). On the Origin of Phyla. University of Chicago Press, Chicago.Google Scholar
Valinsky, J. E., and Douarin, N. M. (1985). Production of plasminogen activator by migrating cephalic neural crest cells. EMBO J. 4, 1403–6.Google ScholarPubMed
Obberghen-Schilling, E., Roche, N. S., Flanders, K. C., Sporn, M. B., and Roberts, A. (1988). Transforming growth factor β1 positively regulates its own expression in normal and transformed cells. J. Biol. Chem. 263, 7741–6.Google Scholar
Oss, C. J., Gillman, C. F., and Neumann, A. W. (1975). Phagocytic Engulfment and Cell Adhesiveness.Marcel Dekker, New York.
Veis, A., and George, A. (1994). Fundamentals of interstitial collagen self-assembly. In Extracellular Matrix Assembly and Function (Yurchenco, P. D., Birk, D. E., and Mecham, R. P., eds.), pp. 15–45. Academic Press, San Diego.Google Scholar
Verheul, H. M., Voest, E. E., and Schlingemann, R. O. (2004). Are tumours angiogenesis-dependent? J. Pathol. 202, 5–13.CrossRefGoogle ScholarPubMed
Vermot, J., and Pourquié, O. (2005). Retinoic acid coordinates somitogenesis and left–right patterning in vertebrate embryos. Nature 435, 215–20.CrossRefGoogle ScholarPubMed
Vermot, J., Llamas, J. G., Fraulob, V., Niederreither, K., Chambon, P., and Dolle, P. (2005). Retinoic acid controls the bilateral symmetry of somite formation in the mouse embryo. Science 308, 563–6.CrossRefGoogle ScholarPubMed
Vernon, G. G., and Woolley, D. M. (1995). The propagation of a zone of activation along groups of flagellar doublet microtubules. Exp. Cell Res. 220, 482–94.CrossRefGoogle ScholarPubMed
Vernon, R. B., Angello, J. C., Iruela-Arispe, M. L., Lane, T. F., and Sage, E. H. (1992). Reorganization of basement membrane matrices by cellular traction promotes the formation of cellular networks in vitro. Lab. Invest. 66, 536–47.Google ScholarPubMed
Vernon, R. B., Lara, S. L., Drake, C. J., et al. (1995). Organized type I collagen influences endothelial patterns during “spontaneous angiogenesis in vitro”: planar cultures as models of vascular development. In Vitro Cell Dev. Biol. Anim. 31, 120–31.CrossRefGoogle ScholarPubMed
Vilar, J. M., Kueh, H. Y., Barkai, N., and Leibler, S. (2002). Mechanisms of noise resistance in genetic oscillators. Proc. Nat. Acad. Sci. USA 99, 5988–92.CrossRefGoogle ScholarPubMed
Dassow, G., and Munro, E. (1999). Modularity in animal development and evolution: elements of a conceptual framework for EvoDevo. J. Exp. Zool. (Mol. Dev. Evol.) 285, 307–25.3.0.CO;2-V>CrossRefGoogle Scholar
Dassow, G., Meir, E., Munro, E. M., and Odell, G. M. (2000). The segment polarity network is a robust developmental module. Nature 406, 188–92.CrossRefGoogle Scholar
Hippel, P. H., and Berg, O. G. (1989). Facilitated target location in biological systems. J. Biol. Chem. 264, 675–8.Google Scholar
Voronov, D. A., and Taber, L. A. (2002). Cardiac looping in experimental conditions: effects of extraembryonic forces. Dev. Dyn. 224, 413–21.CrossRefGoogle ScholarPubMed
Voronov, D. A., Alford, P. W., Xu, G., and Taber, L. A. (2004). The role of mechanical forces in dextral rotation during cardiac looping in the chick embryo. Dev. Biol. 272, 339–50.CrossRefGoogle ScholarPubMed
Waddington, C. H. (1942). Canalization of development and the inheritance of acquired characters. Nature 150, 563–5.CrossRefGoogle Scholar
Wagner, G. P., and Altenberg, L. (1996). Complex adaptations and the evolution of evolvability. Evolution 50, 967–6.CrossRefGoogle ScholarPubMed
Wakamatsu, Y., Maynard, T. M., and Weston, J. A. (2000). Fate determination of neural crest cells by NOTCH-mediated lateral inhibition and asymmetrical cell division during gangliogenesis. Development 127, 2811–21.Google ScholarPubMed
Wakely, J., and England, M. A. (1977). Scanning electron microscopy (SEM) of the chick embryo primitive streak. Differentiation 7, 181–6.CrossRefGoogle ScholarPubMed
Wallingford, J. B., and Harland, R. M. (2002). Neural tube closure requires Dishevelled-dependent convergent extension of the midline. Development 129, 5815–25.CrossRefGoogle Scholar
Wang, N., and Stamenovic, D. (2000). Contribution of intermediate filaments to cell stiffness, stiffening, and growth. Am. J. Physiol. Cell Physiol. 279, C188–94.CrossRefGoogle Scholar
Wang, N., Butler, J. P., and Ingber, D. E. (1993). Mechanotransduction across the cell surface and through the cytoskeleton. Science 260, 1124–7.CrossRefGoogle ScholarPubMed
Wassarman, P., Chen, J., Cohen, N., Litscher, E., Liu, C., Qi, H., and Williams, Z. (1999). Structure and function of the mammalian egg zona pellucida. J. Exp. Zool. 285, 251–8.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Wassarman, P. M. (1999). Mammalian fertilization: molecular aspects of gamete adhesion, exocytosis, and fusion. Cell 96, 175–83.CrossRefGoogle ScholarPubMed
Waters, C. M., Oberg, K. C., Carpenter, G., and Overholser, K. A. (1990). Rate constants for binding, dissociation, and internalization of EGF: effect of receptor occupancy and ligand concentration. Biochemistry 29, 3563–9.CrossRefGoogle ScholarPubMed
Wearing, H. J., Owen, M. R., and Sherratt, J. A. (2000). Mathematical modelling of juxtacrine patterning. Bull. Math. Biol. 62, 293–320.CrossRefGoogle ScholarPubMed
Webb, S. D., and Owen, M. R. (2004). Oscillations and patterns in spatially discrete models for developmental intercellular signalling. J. Math. Biol. 48, 444–76.CrossRefGoogle ScholarPubMed
Weidinger, G., Wolke, U., Koprunner, M., Thisse, C., Thisse, B., and Raz, E. (2002). Regulation of zebrafish primordial germ cell migration by attraction towards an intermediate target. Development 129, 25–36.Google ScholarPubMed
Weng, W., and Stemple, D. L. (2003). Nodal signaling and vertebrate germ layer formation. Birth Defects Res. Part C Embryo Today 69, 325–32.CrossRefGoogle ScholarPubMed
Wessel, G. M., and McClay, D. R. (1987). Gastrulation in the sea urchin embryo requires the deposition of crosslinked collagen within the extracellular matrix. Dev. Biol. 121, 149–65.CrossRefGoogle ScholarPubMed
West-Eberhard, M. J. (2003). Developmental Plasticity and Evolution. Oxford University Press, Oxford, New York.Google Scholar
Wheelock, M. J., and Johnson, K. R. (2003). Cadherins as modulators of cellular phenotype. Ann. Rev. Cell Dev. Biol. 19, 207–35.CrossRefGoogle ScholarPubMed
White, J. G., and Borisy, G. G. (1983). On the mechanisms of cytokinesis in animal cells. J. Theor. Biol. 101, 289–316.CrossRefGoogle ScholarPubMed
Wikramanayake, A. H., Huang, L., and Klein, W. H. (1998). beta-Catenin is essential for patterning the maternally specified animal–vegetal axis in the sea urchin embryo. Proc. Nat. Acad. Sci. USA 95, 9343–8.CrossRefGoogle ScholarPubMed
Wilkins, A. S. (1992). Genetic analysis of animal development. Wiley-Liss, New York.Google Scholar
Wilkins, A. S. (1997). Canalization: a molecular genetic perspective. BioEssays 19, 257–62.CrossRefGoogle ScholarPubMed
Wilkins, A. S. (2002). The Evolution of Developmental Pathways. Sinauer Associates, Sunderland, MA.Google Scholar
Williams, A. F., and Barclay, A. N. (1988). The immunoglobulin superfamily – domains for cell surface recognition. Ann. Rev. Immunol. 6, 381–405.CrossRefGoogle ScholarPubMed
Wilson, P. D. (2004). Polycystic kidney disease: new understanding in the pathogenesis. Int. J. Biochem. Cell Biol. 36, 1868–73.CrossRefGoogle ScholarPubMed
Wimmer, E. A., Carleton, A., Harjes, P., Turner, T., and Desplan, C. (2000). Bicoid-independent formation of thoracic segments in Drosophila. Science 287, 2476–9.CrossRefGoogle ScholarPubMed
Winther, R. G. (2001). Varieties of modules: kinds, levels, origins, and behaviors. J. Exp. Zool. 291, 116–29.CrossRefGoogle ScholarPubMed
Wolf, D. M., and Eeckman, F. H. (1998). On the relationship between genomic regulatory element organization and gene regulatory dynamics. J. Theor. Biol. 195, 167–86.CrossRefGoogle ScholarPubMed
Wolf, J., and Heinrich, R. (1997). Dynamics of two-component biochemical systems in interacting cells; synchronization and desynchronization of oscillations and multiple steady states. Biosystems 43, 1–24.CrossRefGoogle ScholarPubMed
Wolfram, S. (2002). A New Kind of Science. Wolfram Media, Champaign, IL.Google Scholar
Wolpert, L. (1969). Positional information and the spatial pattern of cellular differentiation. J. Theor. Biol. 25, 1–47.CrossRefGoogle ScholarPubMed
Wolpert, L. (2002). Principles of Development. Oxford University Press, Oxford, New York.Google Scholar
Wong, G. K., Allen, P. G., and Begg, D. A. (1997). Dynamics of filamentous actin organization in the sea urchin egg cortex during early cleavage divisions: implications for the mechanism of cytokinesis. Cell Motil. Cytoskeleton 36, 30–42.3.0.CO;2-L>CrossRefGoogle Scholar
Woolley, D. M., and Vernon, G. G. (2002). Functional state of the axonemal dyneins during flagellar bend propagation. Biophys. J. 83, 2162–9.CrossRefGoogle ScholarPubMed
Wylie, C. C., and Heasman, J. (1993). Migration, proliferation, and potency of primordial germ cells. Seminars in Dev. Biol. 4, 161–170.CrossRefGoogle Scholar
Xiao, S., and Knoll, A. H. (2000). Phosphatized animal embryos from the Neoproterozoic Doushantuo Formation in Weng'an, Guizhou, South China. Paleontology 74, 767–88.CrossRefGoogle Scholar
Xiao, S., Yuan, X., and Knoll, A. H. (2000). Eumetazoan fossils in terminal Proterozoic phosphorites? Proc. Nat. Acad. Sci. USA 97, 13684–9.CrossRefGoogle ScholarPubMed
Xu, Z., and Tung, V. W. (2001). Temporal and spatial variations in slow axonal transport velocity along peripheral motoneuron axons. Neuroscience 102, 193–200.CrossRefGoogle ScholarPubMed
Yamada, S., Wirtz, D., and Kuo, S. C. (2000). Mechanics of living cells measured by laser tracking microrheology. Biophys. J. 78, 1736–47.CrossRefGoogle ScholarPubMed
Yamamoto, K., and Yoneda, M. (1983). Cytoplasmic cycle in meiotic division of starfish oocytes. Dev. Biol. 96, 166–72.CrossRefGoogle ScholarPubMed
Yamamoto, M., Saijoh, Y., Perea-Gomez, A., et al. (2004). Nodal antagonists regulate formation of the anteroposterior axis of the mouse embryo. Nature 428, 387–92.CrossRefGoogle ScholarPubMed
Yanagimachi, R., and Noda, Y. D. (1970). Electron microscope studies of sperm incorporation into the golden hamster egg. Am. J. Anat. 128, 429–62.CrossRefGoogle ScholarPubMed
Yoneda, M. (1973). Tension at the surface of sea urchin eggs on the basis of “liquid-drop” concept. Adv. Biophys. 4, 153–90.Google ScholarPubMed
Yoshida, M., Inaba, K., and Morisawa, M. (1993). Sperm chemotaxis during the process of fertilization in the ascidians Ciona savignyi and Ciona intestinalis. Dev. Biol. 157, 497–506.CrossRefGoogle ScholarPubMed
Yurchenco, P. D., and O'Rear, J. J. (1994). Basal lamina assembly. Curr. Opin. Cell. Biol. 6, 674–681.CrossRefGoogle ScholarPubMed
Zachariae, W., and Nasmyth, K. (1999). Whose end is destruction: cell division and the anaphase-promoting complex. Genes Dev. 13, 2039–58.CrossRefGoogle ScholarPubMed
Zahalak, G. I., Wagenseil, J. E., Wakatsuki, T., and Elson, E. L. (2000). A cell-based constitutive relation for bio-artificial tissues. Biophys. J. 79, 2369–81.CrossRefGoogle ScholarPubMed
Zajac, M., Jones, G. L., and Glazier, J. A. (2000). Model of convergent extension in animal morphogenesis. Phys. Rev. Lett. 85, 2022–5.CrossRefGoogle ScholarPubMed
Zajac, M., Jones, G. L., and Glazier, J. A. (2003). Simulating convergent extension by way of anisotropic differential adhesion. J. Theor. Biol. 222, 247–59.CrossRefGoogle ScholarPubMed
Zeng, W., Thomas, G. L., Newman, S. A., and Glazier, J. A. (2003). A novel mechanism for mesenchymal condensation during limb chondrogenesis in vitro. In Mathematical Modelling and Computing in Biology and Medicine, Proc. Fifth ESMTB Conf 2002 (Capasso, V., ed.), pp. 80–6. Società Editrice Esculapio, Bologna, Italy.Google Scholar
Zhang, J., Houston, D. W., King, M. L., Payne, C., Wylie, C., and Heasman, J. (1998). The role of maternal VegT in establishing the primary germ layers in Xenopus embryos. Cell 94, 515–24.CrossRefGoogle ScholarPubMed

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
  • Gabor Forgacs, University of Missouri, Columbia, Stuart A. Newman, New York Medical College
  • Book: Biological Physics of the Developing Embryo
  • Online publication: 24 May 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511755576.013
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
  • Gabor Forgacs, University of Missouri, Columbia, Stuart A. Newman, New York Medical College
  • Book: Biological Physics of the Developing Embryo
  • Online publication: 24 May 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511755576.013
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
  • Gabor Forgacs, University of Missouri, Columbia, Stuart A. Newman, New York Medical College
  • Book: Biological Physics of the Developing Embryo
  • Online publication: 24 May 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511755576.013
Available formats
×