Skip to main content Accessibility help
×
Hostname: page-component-cb9f654ff-k7rjm Total loading time: 0 Render date: 2025-08-02T21:19:34.355Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 June 2014

Lewis I. Held, Jr
Affiliation:
Texas Tech University
Get access

Information

Type
Chapter
Information
How the Snake Lost its Legs
Curious Tales from the Frontier of Evo-Devo
, pp. 157 - 278
Publisher: Cambridge University Press
Print publication year: 2014

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.)

Book purchase

Temporarily unavailable

References

Abbott, A. and other members of the C. elegans Sequencing Consortium (1998). Genome sequence of the nematode C. elegans: a platform for investigating biology. Science 282, 2012–2018.Google Scholar
Abbott, A.L. (2003). Heterochronic genes. Curr. Biol. 13, R824–R825.CrossRefGoogle ScholarPubMed
Abbott, M.K. and Lengyel, J.A. (1991). Embryonic head involution and rotation of male terminalia require the Drosophila locus head involution defective. Genetics 129, 783–789.Google ScholarPubMed
Abmayr, S.M. and Pavlath, G.K. (2012). Myoblast fusion: lessons from flies and mice. Development 139, 641–656.CrossRefGoogle ScholarPubMed
Abouheif, E. (1997). Developmental genetics and homology: a hierarchical approach. Trends Ecol. Evol. 12, 405–408.CrossRefGoogle ScholarPubMed
Abouheif, E. and Wray, G.A. (2002). Evolution of the gene network underlying wing polyphenism in ants. Science 297, 249–252.CrossRefGoogle ScholarPubMed
Abu-Shaar, M. and Mann, R.S. (1998). Generation of multiple antagonistic domains along the proximodistal axis during Drosophila leg development. Development 125, 3821–3830.Google ScholarPubMed
Abzhanov, A. and Kaufman, T.C. (2000). Homologs of Drosophila appendage genes in the patterning of arthropod limbs. Dev. Biol. 227, 673–689.CrossRefGoogle ScholarPubMed
Abzhanov, A., Kuo, W.P., Hartmann, C., Grant, B.R., Grant, P.R., and Tabin, C. (2006). The calmodulin pathway and evolution of elongated beak morphology in Darwin’s finches. Nature 442, 563–567.CrossRefGoogle ScholarPubMed
Abzhanov, A., Protas, M., Grant, B.R., Grant, P.R., and Tabin, C.J. (2004). Bmp4 and morphological variation of beaks in Darwin’s finches. Science 305, 1462–1465.CrossRefGoogle ScholarPubMed
Acampora, D., Annino, A., Tuorto, F., Puelles, E., Lucchesi, W., Papalia, A., and Simeone, A. (2005). Otx genes in the evolution of the vertebrate brain. Brain Res. Bull. 66, 410–420.CrossRefGoogle ScholarPubMed
Achenbach, J. (2010). Lost giants. Natl. Geogr. 218 #4, 90–109.Google Scholar
Adami, C. (2006). Reducible complexity. Science 312, 61–63.CrossRefGoogle ScholarPubMed
Adams, J.K. (1989). Pteronotus davyi. Mamm. Species 346, 1–5.CrossRef
Adams, R.A. and Pedersen, S.C. (1994). Wings on their fingers. Nat. Hist. 103 #1, 48–55.Google Scholar
Adamska, M., Larroux, C., Adamski, M., Green, K., Lovas, E., Koop, D., Richards, G.S., Zwafink, C., and Degnan, B.M. (2010). Structure and expression of conserved Wnt pathway components in the demosponge Amphimedon queenslandica. Evol. Dev. 12, 494–518.CrossRefGoogle ScholarPubMed
Adamska, M., Matus, D.Q., Adamski, M., Green, K., Rokhsar, D.S., Martindale, M.Q., and Degnan, B.M. (2007). The evolutionary origin of hedgehog proteins. Curr. Biol. 17, R836–R837.CrossRefGoogle ScholarPubMed
Adler, P.N. (1992). The genetic control of tissue polarity in Drosophila. BioEssays 14, 735–741.CrossRefGoogle ScholarPubMed
Affolter, M., Pyrowolakis, G., Weiss, A., and Basler, K. (2008). Signal-induced repression: the exception or the rule in developmental signaling? Dev. Cell 15, 11–22.CrossRefGoogle ScholarPubMed
Agrawal, S. and Riffell, J.A. (2011). Behavioral neurobiology: The bitter life of male flies. Curr. Biol. 21, R470–R472.CrossRefGoogle ScholarPubMed
Aguinaldo, A.M.A., Turbeville, J.M., Linford, L.S., Rivera, M.C., Garey, J.R., Raff, R.A., and Lake, J.A. (1997). Evidence for a clade of nematodes, arthropods and other moulting animals. Nature 387, 489–493.CrossRefGoogle ScholarPubMed
Ahnesjö, I. and Craig, J.F. (2011). The biology of Syngnathidae: pipefishes, seadragons and seahorses. J. Fish Biol. 78, 1597–1602.CrossRefGoogle ScholarPubMed
Ahuja, A. and Singh, R.S. (2008). Variation and evolution of male sex combs in Drosophila: nature of selection response and theories of genetic variation for sexual traits. Genetics 179, 503–509.CrossRefGoogle ScholarPubMed
Aiello, L.C. and Andrews, P. (2006). The australopithecines in review. In The Human Evolution Source Book (Ciochon, R.L. and Fleagle, J.G., eds.). Advances in Human Evolution series. Pearson Prentice Hall, Upper Saddle River, NJ, pp. 76–89.Google Scholar
Akam, M. (1998). Hox genes, homeosis and the evolution of segment identity: no need for hopeless monsters. Int. J. Dev. Biol. 42, 445–451.Google ScholarPubMed
Akam, M. (1998). Hox genes: from master genes to micromanagers. Curr. Biol. 8, R676–R678.CrossRefGoogle ScholarPubMed
Akam, M. (1998). The yin and yang of evo/devo. Cell 92, 153–155.CrossRefGoogle ScholarPubMed
Akam, M. (2000). Arthropods: developmental diversity within a (super) phylum. PNAS 97 #9, 4438–4441.CrossRefGoogle ScholarPubMed
Akiyama-Oda, Y. and Oda, H. (2006). Axis specification in the spider embryo: dpp is required for radial-to-axial symmetry transformation and sog for ventral patterning. Development 133, 2347–2357.CrossRefGoogle ScholarPubMed
Albalat, R. (2009). The retinoic acid machinery in invertebrates: ancestral elements and vertebrate innovations. Mol. Cell. Endocrinol. 313, 23–35.CrossRefGoogle ScholarPubMed
Alberch, P. (1980). Ontogenesis and morphological diversification. Am. Zool. 20, 653–667.CrossRefGoogle Scholar
Alberch, P. (1985). Developmental constraints: why St. Bernards often have an extra digit and poodles never do. Am. Nat. 126, 430–433.CrossRefGoogle Scholar
Alberch, P. (1986). Possible dogs. Nat. Hist. 95 #12, 4–8.Google Scholar
Alberch, P., Gould, S.J., Oster, G.F., and Wake, D.B. (1979). Size and shape in ontogeny and phylogeny. Paleobiology 5, 296–317.CrossRefGoogle Scholar
Alberts, B., Johnson, A., Lewis, J., Raff, M., Roberts, K., and Walter, P. (2002). Molecular Biology of the Cell, 4th edn. Garland, New York, NY.Google Scholar
Alexander, B., Baggaley, A., Dennis-Bryan, K., McDonald, F., Munsey, E., Preston, P., Tuson, C., Yelland, A., Hamilton, J., Heilman, C., and Perlmutter, J., eds. (2010). Smithsonian Natural History: The Ultimate Visual Guide to Everything on Earth. Dorling Kindersley, New York, NY.
Alexander, R.M. (1975). The Chordates, 2nd edn. Cambridge University Press, New York, NY.Google Scholar
Alexander, R.M. (1985). Body support, scaling, and allometry. In Functional Vertebrate Morphology (Hildebrand, M., Bramble, D.M., Liem, K.F., and Wake, D.B., eds.). Harvard University Press, Cambridge, MA, pp. 26–37.Google Scholar
Alibardi, L. (2012). Perspectives on hair evolution based on some comparative studies on vertebrate cornification. J. Exp. Zool. (Mol. Dev. Evol.) 318B, 325–343.CrossRefGoogle Scholar
Aliee, M., Röper, J.-C., Landsberg, K.P., Pentzold, C., Widmann, T.J., Jülicher, F., and Dahmann, C. (2012). Physical mechanisms shaping the Drosophila dorsovental compartment boundary. Curr. Biol. 22, 967–976.CrossRefGoogle Scholar
Allen, C.E. (2008). The “eyespot module” and eyespots as modules: development, evolution, and integration of a complex phenotype. J. Exp. Zool. (Mol. Dev. Evol.) 310B, 179–190.CrossRefGoogle Scholar
Allen, C.E., Beldade, P., Zwaan, B.J., and Brakefield, P.M. (2008). Differences in the selection response of serially repeated color pattern characters: standing variation, development, and evolution. BMC Evol. Biol. 8, Article 94 (13 pp.).CrossRefGoogle ScholarPubMed
Allen, W.L., Cuthill, I.C., Scott-Samuel, N.E., and Baddeley, R. (2011). Why the leopard got its spots: relating pattern development to ecology in felids. Proc. R. Soc. Lond. B 278, 1373–1380.CrossRefGoogle ScholarPubMed
Altig, R. (2006). Tadpoles evolved and frogs are the default. Herpetologica 62, 1–10.CrossRefGoogle Scholar
Amemiya, C.T., Miyake, T., and Rast, J.P. (2005). Echinoderms. Curr. Biol. 15, R944–R946.CrossRefGoogle ScholarPubMed
Amundson, R. (2005). The Changing Role of the Embryo in Evolutionary Thought: Roots of Evo-Devo. Cambridge University Press, New York, NY.CrossRefGoogle Scholar
Anderson, E., Peluso, S., Lettice, L.A., and Hill, R.E. (2012). Human limb abnormalities caused by disruption of hedgehog signaling. Trends Genet. 28, 364–373.CrossRefGoogle ScholarPubMed
Andersson, E.R., Sandberg, R., and Lendahl, U. (2011). Notch signaling: simplicity in design, versatility in function. Development 138, 3593–3612.CrossRefGoogle Scholar
Andersson, K., Norman, D., and Werdelin, L. (2011). Sabretoothed carnivores and the killing of large prey. PLoS ONE 6 #10, e24971.CrossRefGoogle ScholarPubMed
Andersson, L.S., Larhammar, M., Memic, F., Wootz, H., Schwochow, D., Rubin, C.-J., Patra, K., Arnason, T., Wellbring, L., Hjälm, G., Imsland, F., Petersen, J.L., McCue, M.E., Mickelson, J.R., Cothran, G., Ahituv, N., Roepstorff, L., Mikko, S., Vallstedt, A., Lindgren, G., Andersson, L., and Kullander, K. (2012). Mutations in DMRT3 affect locomotion in horses and spinal circuit function in mice. Nature 488, 642–646.CrossRefGoogle ScholarPubMed
Andersson, M. and Simmons, L.W. (2006). Sexual selection and mate choice. Trends Ecol. Evol. 21, 296–302.CrossRefGoogle ScholarPubMed
Andersson, O., Reissmann, E., and Ibáñez, C.F. (2006). Growth differentiation factor 11 signals through the transforming growth factor-β receptor ALK5 to regionalize the anterior–posterior axis. EMBO Reports 7, 831–837.Google ScholarPubMed
Andres, B. (2012). The early evolutionary history and adaptive radiation of the pterosauria. Acta Geol. Sinica (English Ed.) 86, 1356–1365. [See also Witton, M.P. (2013). Pterosaurs: Natural History, Evolution, Anatomy. Princeton University Press, Princeton, NJ.]CrossRefGoogle Scholar
Andrew, D.J. and Ewald, A.J. (2010). Morphogenesis of epithelial tubes: insights into tube formation, elongation, and elaboration. Dev. Biol. 341, 34–55.CrossRefGoogle ScholarPubMed
Angelini, D.R. and Kaufman, T.C. (2004). Functional analyses in the hemipteran Oncopeltus fasciatus reveal conserved and derived aspects of appendage patterning in insects. Dev. Biol. 271, 306–321.CrossRefGoogle ScholarPubMed
Angelini, D.R. and Kaufman, T.C. (2005). Comparative developmental genetics and the evolution of arthropod body plans. Annu. Rev. Genet. 39, 95–119.CrossRefGoogle ScholarPubMed
Angelini, D.R. and Kaufman, T.C. (2005). Functional analyses in the milkweed bug Oncopeltus fasciatus (Hemiptera) support a role for Wnt signaling in body segmentation but not appendage development. Dev. Biol. 283, 409–423.CrossRefGoogle Scholar
Angelini, D.R. and Kaufman, T.C. (2005). Insect appendages and comparative ontogenetics. Dev. Biol. 286, 57–77.CrossRefGoogle ScholarPubMed
Angelini, D.R., Liu, P.Z., Hughes, C.L., and Kaufman, T.C. (2005). Hox gene function and interaction in the milkweed bug Oncopeltus fasciatus (Hemiptera). Dev. Biol. 287, 440–455.CrossRefGoogle Scholar
Aoyama, H. and Asamoto, K. (2000). The developmental fate of the rostral/caudal half of a somite for vertebra and rib formation: experimental confirmation of the resegmentation theory using chick-quail chimeras. Mechs. Dev. 99, 71–82.CrossRefGoogle ScholarPubMed
Apesteguía, S. and Zaher, H. (2006). A Cretaceous terrestrial snake with robust hindlimbs and a sacrum. Nature 440, 1037–1040.CrossRefGoogle ScholarPubMed
Apter, M.J. and Wolpert, L. (1965). Cybernetics and development. I. Information theory. J. Theor. Biol. 8, 244–257.CrossRefGoogle ScholarPubMed
Aragón, J.L., Torres, M., Gil, D., Barrio, R.A., and Maini, P.K. (2002). Turing patterns with pentagonal symmetry. Phys. Rev. E (Stat. Nonlin. Soft Matter Phys.) 65, 051913 (9 pp.).CrossRefGoogle ScholarPubMed
Arakane, Y., Lomakin, J., Gehrke, S.H., Hiromasa, Y., Tomich, J.M., Muthukrishnan, S., Beeman, R.W., Kramer, K.J., and Kanost, M.R. (2012). Formation of rigid, non-flight forewings (elytra) of a beetle requires two major cuticular proteins. PLoS Genet. 8 #4, e1002682.CrossRefGoogle ScholarPubMed
Arbib, M.A. (1972). Automata theory in the context of theoretical embryology. In Foundations of Mathematical Biology, Vol. 2 (Rosen, R., ed.). Academic Press, New York, NY, pp. 141–215.CrossRefGoogle Scholar
Arendt, D. and Nübler-Jung, K. (1996). Common ground plans in early brain development in mice and flies. BioEssays 18, 255–259.CrossRefGoogle ScholarPubMed
Arendt, D. and Nübler-Jung, K. (1999). Comparison of early nerve cord development in insects and vertebrates. Development 126, 2309–2325.Google ScholarPubMed
Arendt, D., Technau, U., and Wittbrodt, J. (2001). Evolution of the bilaterian larval foregut. Nature 409, 81–85.CrossRefGoogle ScholarPubMed
Armstrong, J.R. and Ferguson, M.W.J. (1995). Ontogeny of the skin and the transition from scar-free to scarring phenotype during wound healing in the pouch young of a marsupial, Monodelphis domestica. Dev. Biol. 169, 242–260.CrossRefGoogle ScholarPubMed
Arnoult, L., Su, K.F.Y., Manoel, D., Minervino, C., Magriña, J., Gompel, N., and Prud’homme, B. (2013). Emergence and diversification of fly pigmentation through evolution of a gene regulatory module. Science 339, 1423–1426.CrossRefGoogle ScholarPubMed
Arthur, W. (2000). The concept of developmental reprogramming and the quest for an inclusive theory of evolutionary mechanisms. Evol. Dev. 2, 49–57.CrossRefGoogle ScholarPubMed
Arthur, W. (2002). The emerging conceptual framework of evolutionary developmental biology. Nature 415, 757–764.CrossRefGoogle ScholarPubMed
Arthur, W. (2004). Biased Embryos and Evolution. Cambridge University Press, New York, NY.CrossRefGoogle Scholar
Arthur, W. (2006). D’Arcy Thompson and the theory of transformations. Nature Rev. Genet. 7, 401–406.CrossRefGoogle Scholar
Arthur, W. (2011). Evolution: A Developmental Approach. Wiley-Blackwell, Chichester.Google Scholar
Arthur, W. (2011). Searching for evo-devo’s Holy Grail: the nature of developmental variation. Evol. Dev. 13, 405–407.CrossRefGoogle ScholarPubMed
Arthur, W. and Chipman, A.D. (2005). How does arthropod segment number evolve? Some clues from centipedes. Evol. Dev. 7, 600–607.CrossRefGoogle ScholarPubMed
Arthur, W. and Farrow, M. (1999). The pattern of variation in centipede segment number as an example of developmental constraint in evolution. J. Theor. Biol. 200, 183–191.CrossRefGoogle Scholar
Asai, R., Taguchi, E., Kume, Y., Saito, M., and Kondo, S. (1999). Zebrafish Leopard gene as a component of the putative reaction-diffusion system. Mechs. Dev. 89, 87–92.CrossRefGoogle ScholarPubMed
Ashburner, M. (1989). Drosophila: A Laboratory Handbook. CSH Press, Cold Spring Harbor, NY.Google Scholar
Ashburner, M. (2006). Won for All: How the Drosophila Genome Was Sequenced. Cold Spring Harbor Laboratory Press, Plainview, NY.Google Scholar
Asher, R.J., Lin, K.H., Kardjilov, N., and Hautier, L. (2011). Variability and constraint in the mammalian vertebral column. J. Evol. Biol. 24, 1080–1090.CrossRefGoogle ScholarPubMed
Aspiras, A.C., Smith, F.W., and Angelini, D.R. (2011). Sex-specific gene interactions in the patterning of insect genitalia. Dev. Biol. 360, 369–380.CrossRefGoogle ScholarPubMed
Atallah, J., Liu, N.H., Dennis, P., Hon, A., Godt, D., and Larsen, E.W. (2009). Cell dynamics and developmental bias in the ontogeny of a complex sexually dimorphic trait in Drosophila melanogaster. Evol. Dev. 11, 191–204.CrossRefGoogle ScholarPubMed
Atallah, J., Liu, N.H., Dennis, P., Hon, A., and Larsen, E.W. (2009). Developmental constraints and convergent evolution in Drosophila sex comb formation. Evol. Dev. 11, 205–218.CrossRefGoogle ScholarPubMed
Atallah, J., Watabe, H., and Kopp, A. (2012). Many ways to make a novel structure: a new mode of sex comb development in Drosophilidae. Evol. Dev. 14, 476–483.CrossRefGoogle ScholarPubMed
Attenborough, D. (2002). The Life of Mammals. Princeton University Press, Princeton, NJ.Google Scholar
Attenborough, D. (2005). Life in the Undergrowth. Princeton University Press, Princeton, NJ.Google Scholar
Attenborough, D. (2008). Life in Cold Blood. Princeton University Press, Princeton, NJ.Google Scholar
Attisano, L. and Wrana, J.L. (1998). Mads and Smads in TGFb signalling. Curr. Opin. Cell Biol. 10, 188–194.CrossRefGoogle Scholar
Aubret, F., Bonnet, X., Pearson, D., and Shine, R. (2005). How can blind tiger snakes (Notechis scutatus) forage successfully?Aust. J. Zool. 53, 283–288.CrossRefGoogle Scholar
Aulehla, A. and Herrmann, B.G. (2004). Segmentation in vertebrates: clock and gradient finally joined. Genes Dev. 18, 2060–2067.CrossRefGoogle ScholarPubMed
Averof, M. (2002). Arthropod Hox genes: insights on the evolutionary forces that shape gene functions. Curr. Opin. Gen. Dev. 12, 386–392.CrossRefGoogle ScholarPubMed
Averof, M. and Cohen, S.M. (1997). Evolutionary origin of insect wings from ancestral gills. Nature 385, 627–630.CrossRefGoogle ScholarPubMed
Avise, J.C. (2006). Evolutionary Pathways in Nature: A Phylogenetic Approach. Cambridge University Press, New York, NY.CrossRefGoogle Scholar
Aw, S. and Levin, M. (2008). What’s left in asymmetry? Dev. Dynamics 237, 3453–3463.CrossRefGoogle ScholarPubMed
Aw, S. and Levin, M. (2009). Is left-right asymmetry a form of planar cell polarity? Development 136, 355–366.CrossRefGoogle ScholarPubMed
Ayers, K.L., Gallet, A., Staccini-Lavenant, L., and Thérond, P.P. (2010). The long-range activity of Hedgehog is regulated in the apical extracellular space by the glypican Dally and the hydrolase Notum. Dev. Cell 18, 605–620. [See also Sanders, T.A., et al. (2013). Specialized filopodia direct long-range transport of SHH during vertebrate tissue patterning. Nature 497, 628–632.]CrossRefGoogle ScholarPubMed
Ayers, K.L., Mteirek, R., Cervantes, A., Lavenant-Staccini, L., Thérond, P.P., and Gallet, A. (2012). Dally and Notum regulate the switch between low and high level Hedgehog pathway signalling. Development 139, 3168–3179.CrossRefGoogle ScholarPubMed
Ayres, J.M. (1990). Scarlet faces of the Amazon. Nat. Hist. 99 #3, 32–41.Google Scholar
Azevedo, A.S., Sousa, S., Jacinto, A., and Saúde, L. (2012). An amputation resets positional information to a proximal identity in the regenerating zebrafish caudal fin. BMC Dev. Biol. 12, Article 24 (10 pp.).CrossRefGoogle ScholarPubMed
Bada, J.L., Mitchell, E., and Kemper, B. (1983). Aspartic acid racemization in narwhal teeth. Nature 303, 418–420.CrossRefGoogle Scholar
Badyaev, A.V. (2011). How do precise adaptive features arise in development? Examples with evolution of context-specific sex ratios and perfect beaks. The Auk 128, 467–474.CrossRefGoogle Scholar
Baena-Lopez, L.A., Nojima, H., and Vincent, J.-P. (2012). Integration of morphogen signalling within the growth regulatory network. Curr. Opin. Cell Biol. 24, 166–172.CrossRefGoogle ScholarPubMed
Baer, M.M., Chanut-Delalande, H., and Affolter, M. (2009). Cellular and molecular mechanisms underlying the formation of biological tubes. Curr. Topics Dev. Biol. 89, 137–162.CrossRefGoogle ScholarPubMed
Baguñà, J., Martinez, P., Paps, J., and Riutort, M. (2008). Unravelling body plan and axial evolution in the Bilateria with molecular phylogenetic markers. In Evolving Pathways: Key Themes in Evolutionary Developmental Biology (Minelli, A. and Fusco, G., eds.). Cambridge University Press, New York, NY, pp. 217–238.CrossRefGoogle Scholar
Bailly, X., Reichert, H., and Hartenstein, V. (2013). The urbilaterian brain revisited: novel insights into old questions from new flatworm clades. Dev. Genes Evol. 223, 149–157.CrossRefGoogle ScholarPubMed
Bak, P., Chen, K., and Creutz, M. (1989). Self-organized criticality in the “Game of Life”. Nature 342, 780–782.CrossRefGoogle Scholar
Baker, R.E. and Murray, P.J. (2012). Understanding hair follicle cycling: a systems approach. Curr. Opin. Gen. Dev. 22, 607–612.CrossRefGoogle ScholarPubMed
Baker, R.E., Schnell, S., and Maini, P.K. (2006). A clock and wavefront mechanism for somite formation. Dev. Biol. 293, 116–126.CrossRefGoogle ScholarPubMed
Baker, R.E., Schnell, S., and Maini, P.K. (2009). Waves and patterning in developmental biology: vertebrate segmentation and feather bud formation as case studies. Int. J. Dev. Biol. 53, 783–794.CrossRefGoogle ScholarPubMed
Baker, R.H. and Wilkinson, G.S. (2001). Phylogenetic analysis of sexual dimorphism and eye-span allometry in stalk-eyed flies (Diopsidae). Evolution 55, 1373–1385.CrossRefGoogle Scholar
Ball, E.E., de Jong, D.M., Schierwater, B., Shinzato, C., Hayward, D.C., and Miller, D.J. (2011). Cnidarian gene expression patterns and the origins of bilaterality: are cnidarians reading the same game plan as “higher” animals? In Key Transitions in Animal Evolution (DeSalle, R. and Schierwater, B., eds.). Science Publishers, Enfield, NH, pp. 197–216.Google Scholar
Ball, P. (1999). The Self-Made Tapestry: Pattern Formation in Nature. Oxford University Press, New York, NY.Google Scholar
Ball, P. (2012). Pattern formation in nature: physical constraints and self-organising characteristics. Architectural Design 216 #SI (Special Issue), 22–27.CrossRefGoogle Scholar
Ball, P. (2013). Celebrate the unknowns. Nature 496, 419–420.CrossRefGoogle ScholarPubMed
Ballard, J.W.O., Olsen, G.J., Faith, D.P., Odgers, W.A., Rowell, D.M., and Atkinson, P.W. (1992). Evidence from 12S ribosomal RNA sequences that onychophorans are modified arthropods. Science 258, 1345–1348.CrossRefGoogle ScholarPubMed
Balter, M. (2013). Dramatic fossils suggest early birds were biplanes. Science 339, 1261.CrossRefGoogle ScholarPubMed
Baratte, S., Andouche, A., and Bonnaud, L. (2007). Engrailed in cephalopods: a key gene related to the emergence of morphological novelties. Dev. Genes Evol. 217, 353–362.CrossRefGoogle ScholarPubMed
Bard, J. and Lauder, I. (1974). How well does Turing’s theory of morphogenesis work? J. Theor. Biol. 45, 501–531.CrossRefGoogle ScholarPubMed
Bard, J.B.L. (1977). A unity underlying the different zebra striping patterns. J. Zool. Lond. 183, 527–539.Google Scholar
Bard, J.B.L. (1981). A model for generating aspects of zebra and other mammalian coat patterns. J. Theor. Biol. 93, 363–385.CrossRefGoogle ScholarPubMed
Bard, J.B.L. (1984). Butterfly wing patterns: how good a determining mechanism is the simple diffusion of a single morphogen? J. Embryol. Exp. Morph. 84, 255–274.Google Scholar
Bard, J.B.L. (1990). Traction and the formation of mesenchymal condensations in vivo. BioEssays 12, 389–395.CrossRefGoogle ScholarPubMed
Bardach, J.E. (1967). The chemical senses and food intake in the lower vertebrates. In The Chemical Senses and Nutrition (Kare, M.R. and Maller, O., eds.). Johns Hopkins Press, Baltimore, MD, pp. 19–43.Google Scholar
Barmina, O. and Kopp, A. (2007). Sex-specific expression of a HOX gene associated with rapid morphological evolution. Dev. Biol. 311, 277–286.CrossRefGoogle ScholarPubMed
Barnes, J., ed. (1984). The Complete Works of Aristotle: The Revised Oxford Translation. Vol. 1. Princeton University Press, Princeton, NJ.
Barnett, A.A. and Brandon-Jones, D. (1997). The ecology, biogeography and conservation of the uakaris, Cacajao (PItheciinae). Folia Primatol. 68, 223–235.CrossRefGoogle Scholar
Barnett, R., Barnes, I., Phillips, M.J., Martin, L.D., Harington, C.R., Leonard, J.A., and Cooper, A. (2005). Evolution of the extinct sabretooths and the American cheetah-like cat. Curr. Biol. 15, R589–R590.CrossRefGoogle ScholarPubMed
Barolo, S. and Posakony, J.W. (2002). Three habits of highly effective signaling pathways: principles of transcriptional control by developmental cell signaling. Genes Dev. 16, 1167–1181.CrossRefGoogle ScholarPubMed
Barrett, R.D.H. and Hoekstra, H.E. (2011). Molecular spandrels: tests of adaptation at the genetic level. Nature Rev. Genet. 12, 767–780.CrossRefGoogle ScholarPubMed
Barrio, R.A., Baker, R.E., Vaughan, B., Tribuzy, K., de Carvalho, M.R., Bassanezi, R., and Maini, P.K. (2009). Modeling the skin pattern of fishes. Phys. Rev. E 79, Article 031908 (11 pp.).CrossRefGoogle ScholarPubMed
Barrio, R.A., Varea, C., Aragón, J.L., and Maini, P.K. (1999). A two-dimensional numerical study of spatial pattern formation in interacting Turing systems. Bull. Math. Biol. 61, 483–505.CrossRefGoogle ScholarPubMed
Barsbold, R., Currie, P.J., Myhrvold, N.P., Osmólska, H., Tsogtbaatar, K., and Watabe, M. (2000). A pygostyle from a non-avian theropod. Nature 403, 155.CrossRefGoogle ScholarPubMed
Barton, R.A. and Harvey, P.H. (2000). Mosaic evolution of brain structure in mammals. Nature 405, 1055–1058.CrossRefGoogle ScholarPubMed
Basler, K. and Struhl, G. (1994). Compartment boundaries and the control of Drosophila limb pattern by hedgehog protein. Nature 368, 208–214.CrossRefGoogle ScholarPubMed
Bastida, M.F. and Ros, M.A. (2008). How do we get a perfect complement of digits? Curr. Opin. Gen. Dev. 18, 374–380.CrossRefGoogle ScholarPubMed
Bateson, P. (2004). The active role of behaviour in evolution. Biol. Philos. 19, 283–298.CrossRefGoogle Scholar
Bateson, W. (1894). Materials for the Study of Variation Treated with Especial Regard to Discontinuity in the Origin of Species. Macmillan, London.CrossRefGoogle Scholar
Baum, B. (2006). Left-right asymmetry: actin-myosin through the looking glass. Curr. Biol. 16, R502–R504.CrossRefGoogle ScholarPubMed
Bautista, A., Mendoza-Degante, M., Coureaud, G., Martínez-Gómez, M., and Hudson, R. (2005). Scramble competition in newborn domestic rabbits for an unusually restricted milk supply. Anim. Behav. 70, 1011–1021.CrossRefGoogle Scholar
Bavendam, F. (1995). The giant cuttlefish: chameleon of the reef. Natl. Geogr. 188 #3, 94–107.Google Scholar
Beachy, P.A., Helfand, S.L., and Hogness, D.S. (1985). Segmental distribution of bithorax complex proteins during Drosophila development. Nature 313, 545–551.CrossRefGoogle ScholarPubMed
Bechstedt, S. and Howard, J. (2008). Hearing mechanics: a fly in your ear. Curr. Biol. 18, R869–R870.CrossRefGoogle ScholarPubMed
Bechtel, H.B. (1995). Reptile and Amphibian Variants: Colors, Patterns, and Scales. Krieger, Malabar, FL.Google Scholar
Becker, J.B., Berkley, K.J., Geary, N., Hampson, E., Herman, J.P., and Young, E.A., eds. (2008). Sex Differences in the Brain: From Genes to Behavior. Oxford University Press, New York, NY.
Beckman, J., Banks, S.C., Sunnucks, P., Lill, A., and Taylor, A.C. (2007). Phylogeography and environmental correlates of a cap on reproduction: teat number in a small marsupial, Antechinus agilis. Mol. Ecol. 16, 1069–1083.CrossRefGoogle Scholar
Bejder, L. and Hall, B.K. (2002). Limbs in whales and limblessness in other vertebrates: mechanisms of evolutionary and developmental transformation and loss. Evol. Dev. 4, 445–458.CrossRefGoogle ScholarPubMed
Belalcazar, A.D., Doyle, K., Hogan, J., Neff, D., and Collier, S. (2013). Insect wing membrane topography is determined by the dorsal wing epithelium. G3 (Genes, Genomes, Genetics) 3, 5–8.Google ScholarPubMed
Beldade, P. and Brakefield, P.M. (2002). The genetics and evo-devo of butterfly wing patterns. Nature Rev. Gen. 3, 442–452.CrossRefGoogle ScholarPubMed
Beldade, P. and Brakefield, P.M. (2003). Concerted evolution and developmental integration in modular butterfly wing patterns. Evol. Dev. 5, 169–179.CrossRefGoogle ScholarPubMed
Beldade, P. and Brakefield, P.M. (2003). The difficulty of agreeing about constraints. Evol. Dev. 5, 119–120.CrossRefGoogle ScholarPubMed
Beldade, P., Brakefield, P.M., and Long, A.D. (2005). Generating phenotypic variation: prospects from “evo-devo” research on Bicyclus anynana wing patterns. Evol. Dev. 7, 101–107.CrossRefGoogle ScholarPubMed
Beldade, P., French, V., and Brakefield, P.M. (2008). Developmental and genetic mechanisms for evolutionary diversification of serial repeats: eyespot size in Bicyclus anynana butterflies. J. Exp. Zool. (Mol. Dev. Evol.) 310B, 191–201.CrossRefGoogle Scholar
Beldade, P., Koops, K., and Brakefield, P.M. (2002). Developmental constraints versus flexibility in morphological evolution. Nature 416, 844–847.CrossRefGoogle ScholarPubMed
Beldade, P., Koops, K., and Brakefield, P.M. (2002). Modularity, individuality, and evo-devo in butterfly wings. PNAS 99 #22, 14262–14267.CrossRefGoogle ScholarPubMed
Beldade, P., McMillan, W.O., and Papanicolaou, A. (2008). Butterfly genomics eclosing. Heredity 100, 150–157.CrossRefGoogle ScholarPubMed
Bell, E., Andres, B., and Goswami, A. (2011). Integration and dissociation of limb elements in flying vertebrates: a comparison of pterosaurs, birds and bats. J. Evol. Biol. 24, 2586–2599.CrossRefGoogle ScholarPubMed
Bellairs, A.D. (1948). The eyelids and spectacle in geckos. Proc. Zool. Soc. Lond. 118, 420–425.CrossRefGoogle Scholar
Bellairs, A.D. and Underwood, G. (1951). The origin of snakes. Biol. Rev. 26, 193–237.CrossRefGoogle ScholarPubMed
Bellone, R.R., Forsyth, G., Leeb, T., Archer, S., Sigurdsson, S., Imsland, F., Mauceli, E., Engensteiner, M., Bailey, E., Sandmeyer, L., Grahn, B., Lindblad-Toh, K., and Wade, C.M. (2010). Fine-mapping and mutation analysis of TRPM1: a candidate gene for leopard complex (LP) spotting and congenital stationary night blindness in horses. Briefings Funct. Genomics 9, 193–207.CrossRefGoogle ScholarPubMed
Belote, J.M. and Baker, B.S. (1982). Sex determination in Drosophila melanogaster: analysis of transformer-2, a sex-transforming locus. PNAS 79 #5, 1568–1572.CrossRefGoogle ScholarPubMed
Beloussov, L.V. (2006). Direct physical formation of anatomical structures by cell traction forces: an interview with Albert Harris. Int. J. Dev. Biol. 50, 93–101.CrossRefGoogle Scholar
Beloussov, L.V. and Grabovsky, V.I. (2006). Morphomechanics: goals, basic experiments and models. Int. J. Dev. Biol. 50, 81–92.CrossRefGoogle ScholarPubMed
Bely, A.E. (2010). Evolutionary loss of animal regeneration: pattern and process. Integr. Comp. Biol. 50, 515–527.CrossRefGoogle Scholar
Ben-Jacob, E. and Levine, H. (1998). The artistry of microorganisms. Sci. Am. 279 #4, 82–87.CrossRefGoogle Scholar
Ben-Zvi, D., Shilo, B.-Z., and Barkai, N. (2011). Scaling of morphogen gradients. Curr. Opin. Gen. Dev. 21, 704–710. [See also Inomata, H., et al. (2013). Scaling of dorsal-ventral patterning by embryo size-dependent degradation of Spemann’s organizer signals. Cell 153, 1296–1311.]CrossRefGoogle ScholarPubMed
Bengtson, S., Cunningham, J.A., Yin, C., and Donoghue, P.C.J. (2012). A merciful death for the “earliest bilaterian,” Vernanimalcula. Evol. Dev. 14, 421–427.CrossRefGoogle ScholarPubMed
Bennett, D.K. (1980). Stripes do not a zebra make, Part 1: A cladistic analysis of Equus. Syst. Zool. 29, 272–287.CrossRefGoogle Scholar
Bensoussan-Trigano, V., Lallemand, Y., Cloment, C.S., and Robert, B. (2012). Msx1 and Msx2 in limb mesenchyme modulate digit number and identity. Dev. Dynamics 240, 1190–1202.CrossRefGoogle Scholar
Benton, M.J. (2010). Studying function and behavior in the fossil record. PLoS Biol. 8 #3, e1000321.CrossRefGoogle ScholarPubMed
Berger, L.R. (2013). The mosaic nature of Australopithecus sediba. Science 340, 163.CrossRefGoogle ScholarPubMed
Bergmann, P.J. and Berk, C.P. (2012). The evolution of positive allometry of weaponry in horned lizards (Phrynosoma). Evol. Biol. 39, 311–323.CrossRefGoogle Scholar
Bergmann, P.J. and Irschick, D.J. (2010). Alternate pathways of body shape evolution translate into common patterns of locomotor evolution in two clades of lizards. Evolution 64, 1569–1582.CrossRefGoogle ScholarPubMed
Bernard, F., Dutriaux, A., Silber, J., and Lalouette, A. (2006). Notch pathway repression by vestigial is required to promote indirect flight muscle differentiation in Drosophila melanogaster. Dev. Biol. 295, 164–177.CrossRefGoogle ScholarPubMed
Bernard, F., Lalouette, A., Gullaud, M., Jeantet, A.Y., Cossard, R., Zider, A., Ferveur, J.F., and Silber, J. (2003). Control of apterous by vestigial drives indirect flight muscle development in Drosophila. Dev. Biol. 260, 391–403.CrossRefGoogle ScholarPubMed
Berta, A., Ray, C.E., and Wyss, A.R. (1989). Skeleton of the oldest known pinniped, Enaliarctos mealsi. Science 244, 60–62.CrossRefGoogle ScholarPubMed
Berta, A., Sumich, J.L., and Kovacs, K.M. (2006). Marine Mammals: Evolutionary Biology, 2nd edn. Elsevier, New York, NY.Google Scholar
Bertrand, S., Brunet, F.G., Escriva, H., Parmentier, G., Laudet, V., and Robinson-Rechavi, M. (2004). Evolutionary genomics of nuclear receptors: from 25 ancestral genes to derived endocrine systems. Mol. Biol. Evol. 21, 1923–1937.CrossRefGoogle ScholarPubMed
Bertrand, S. and Escriva, H. (2011). Evolutionary crossroads in developmental biology: amphioxus. Development 138, 4819–4830.CrossRefGoogle ScholarPubMed
Bessho, Y., Hirata, H., Masamizu, Y., and Kageyama, R. (2003). Periodic repression by the bHLH factor Hes7 is an essential mechanism for the somite segmentation clock. Genes Dev. 17, 1451–1456.CrossRefGoogle ScholarPubMed
Bessho, Y., Miyoshi, G., Sakata, R., and Kageyama, R. (2001). Hes7: a bHLH-type repressor gene regulated by Notch and expressed in the presomitic mesoderm. Genes to Cells 6, 175–185.CrossRefGoogle ScholarPubMed
Bessodes, N., Haillot, E., Duboc, V., Röttinger, E., Lahaye, F., and Lepage, T. (2012). Reciprocal signaling between the ectoderm and a mesendodermal left-right organizer directs left-right determination in the sea urchin embryo. PLoS Genet. 8 #12, e1003121.CrossRefGoogle Scholar
Best, R.C. (1981). The tusk of the narwhal (Monodon monoceros L.): interpretation of its function (Mammalia: Cetacea). Can. J. Zool. 59, 2386–2393.CrossRefGoogle Scholar
Beverdam, A., Merlo, G.R., Paleari, L., Mantero, S., Genova, F., Barbieri, O., Janvier, P., and Levi, G. (2002). Jaw transformation with gain of symmetry after Dlx5/Dlx6 inactivation: mirror of the past?Genesis 34, 221–227.CrossRefGoogle ScholarPubMed
Beveridge, W.I.B. (1950). The Art of Scientific Investigation. Random House, New York, NY.Google Scholar
Bhalla, U.S. and Iyengar, R. (1999). Emergent properties of networks of biological signaling pathways. Science 283, 381–387.CrossRefGoogle ScholarPubMed
Bhattacharyya, R.P., Reményi, A., Yeh, B.J., and Lim, W.A. (2006). Domains, motifs, and scaffolds: the role of modular interactions in the evolution and wiring of cell signaling circuits. Annu. Rev. Biochem. 75, 655–680.CrossRefGoogle ScholarPubMed
Bhullar, B.-A.S. (2012). A phylogenetic approach to ontogeny and heterochrony in the fossil record: cranial evolution and development in anguimorphan lizards (Reptilia: Squamata). J. Exp. Zool. (Mol. Dev. Evol.) 318B, 521–530.CrossRefGoogle Scholar
Bhullar, B.-A.S., Marugán-Lobón, J., Racimo, F., Bever, G.S., Rowe, T.B., Norell, M.A., and Abzhanov, A. (2012). Birds have paedomorphic dinosaur skulls. Nature 487, 223–226.CrossRefGoogle ScholarPubMed
Bickelmann, C., Mitgutsch, C., Richardson, M.K., Jiménez, R., de Bakker, M.A.G., and Sánchez-Villagra, M.R. (2012). Transcriptional heterochrony in talpid mole autopods. EvoDevo 3, Article 16 (5 pp.).CrossRefGoogle ScholarPubMed
Bickford, D., Iskandar, D., and Barlian, A. (2008). A lungless frog discovered on Borneo. Curr. Biol. 18, R374–R375.CrossRefGoogle ScholarPubMed
Bier, E. (1997). Anti-neural-inhibition: a conserved mechanism for neural induction. Cell 89, 681–684.CrossRefGoogle ScholarPubMed
Bier, E. (2011). Evolution of development: diversified dorsoventral patterning. Curr. Biol. 21, R591–R594.CrossRefGoogle ScholarPubMed
Billet, G., Hautier, L., Asher, R.J., Schwarz, C., Crumpton, N., Martin, T., and Ruf, I. (2012). High morphological variation of vestibular system accompanies slow and infrequent locomotion in three-toed sloths. Proc. R. Soc. Lond. B 279, 3932–3939.CrossRefGoogle ScholarPubMed
Bishop, K.L. (2008). The evolution of flight in bats: narrowing the field of plausible hypotheses. Q. Rev. Biol. 83, 153–169.CrossRefGoogle ScholarPubMed
Bitan, A., Rosenbaum, I., and Abdu, U. (2012). Stable and dynamic microtubules coordinately determine and maintain Drosophila bristle shape. Development 139, 1987–1996.CrossRefGoogle ScholarPubMed
Blackstone, N.W. (2006). Charles Manning Child (1869–1954): the past, present, and future of metabolic signaling. J. Exp. Zool. (Mol. Dev. Evol.) 306B, 1–7.CrossRefGoogle Scholar
Blair, P. (1710). Osteographia elephantina: Or, a full and exact description of all the bones of an elephant, which died near Dundee, April the 27th, 1706. With their several dimensions. Communicated in a letter to Dr. Hans Sloane, R. S. Secr. by Mr Patrick Blair, Surgeon, &c. Phil. Trans. 27, 53–116.CrossRefGoogle Scholar
Blair, S.S. (2009). Segmentation in animals. Curr. Biol. 18, R991–R995.CrossRefGoogle Scholar
Blanco, M.J., Misof, B.Y., and Wagner, G.P. (1998). Heterochronic differences of Hoxa-11 expression in Xenopus fore- and hind limb development: Evidence for lower limb identity of the anural ankle bones. Dev. Genes Evol. 208, 175–187.CrossRefGoogle Scholar
Blaxter, M. and Sunnucks, P. (2011). Velvet worms. Curr. Biol. 21, R238–R240.CrossRefGoogle ScholarPubMed
Boaz, N.T. (1997). Eco Homo: How the Human Being Emerged from the Cataclysmic History of the Earth. Basic Books, New York, NY.Google Scholar
Bobkova, M.V., Gál, J., Zhukov, V.V., Shepeleva, I.P., and Meyer-Rochow, V.B. (2004). Variations in the retinal designs of pulmonate snails (Mollusca, Gastropoda): squaring phylogenetic background and ecophysiological needs (I). Invert. Biol. 123, 101–115.CrossRefGoogle Scholar
Bock, G. and Goode, J., eds. (2007). Tinkering: The Microevolution of Development. Wiley, Chichester.
Bodmann, B.E.J. and Mombach, J.C.M. (2000). On the role of probability amplitudes in cell aggregation: an approach study towards morphogenesis. Physica A 278, 243–259.CrossRefGoogle Scholar
Bodmer, R. and Venkatesh, T.V. (1998). Heart development in Drosophila and vertebrates: conservation of molecular mechanisms. Dev. Genet. 22, 181–186.3.0.CO;2-2>CrossRefGoogle ScholarPubMed
Boell, L. (2013). Lines of least resistance and genetic architecture of house mouse (Mus musculus) mandible shape. Evol. Dev. 15, 197–204.CrossRefGoogle ScholarPubMed
Boero, F. and Piraino, S. (2011). From cnidaria to “higher metazoa” in one step. In Key Transitions in Animal Evolution (DeSalle, R. and Schierwater, B., eds.). Science Publishers, Enfield, NH, pp. 162–174.Google Scholar
Boero, F., Schierwater, B., and Piraino, S. (2007). Cnidarian milestones in metazoan evolution. Integr. Comp. Biol. 47, 693–700.CrossRefGoogle ScholarPubMed
Boisserie, J.-R., Fisher, R.E., Lihoreau, F., and Weston, E.M. (2011). Evolving between land and water: key questions on the emergence and history of the Hippopotamidae (Hippopotamoidea, Cetancodonta, Cetartiodactyla). Biol. Rev. 86, 601–625.CrossRefGoogle Scholar
Bökel, C. and Brand, M. (2013). Generation and interpretation of FGF morphogen gradients in vertebrates. Curr. Opin. Gen. Dev. 23, 415–422.CrossRefGoogle ScholarPubMed
Boletzky, S.v. (1988). Cephalopod development and evolutionary concepts. In The Mollusca, Vol. 12: Paleontology and Neontology of Cephalopods (Wilbur, K.M., Clarke, M.R., and Trueman, E.R., eds.). Academic Press, New York, NY, pp. 185–202.CrossRefGoogle Scholar
Boletzky, S.v. (1988). Characteristics of cephalopod embryogenesis. In Cephalopods: Present and Past (Wiedmann, J. and Kullmann, J., eds.). O. H. Schindewolf-Symposium Tübingen 1985 (2nd International Cephalopod Symposium). Schweizerbart, Stuttgart, pp. 167–179.Google Scholar
Bolker, J.A. and Raff, R.A. (1996). Developmental genetics and traditional homology. BioEssays 18, 489–494.CrossRefGoogle ScholarPubMed
Bomtorin, A.D., Barchuk, A.R., Moda, L.M., and Simoes, Z.L.P. (2012). Hox gene expression leads to differential hind leg development between honeybee castes. PLoS ONE 7 #7, e40111.CrossRefGoogle ScholarPubMed
Bonato, L., Foddai, D., and Minelli, A. (2003). Evolutionary trends and patterns in centipede segment number based on a cladistic analysis of Mecistocephalidae (Chilopoda: Geophilomorpha). Syst. Entomol. 28, 539–579.CrossRefGoogle Scholar
Bonnet, X., Bradshaw, D., Shine, R., and Pearson, D. (1999). Why do snakes have eyes? The (non-)effect of blindness in island tiger snakes (Notechis scutatus). Behav. Ecol. Sociobiol. 46, 267–272.CrossRefGoogle Scholar
Borowsky, R. (2008). Restoring sight in blind cavefish. Curr. Biol. 18, R23–R24.CrossRefGoogle ScholarPubMed
Bosch, T.C.G. (2003). Ancient signals: peptides and the interpretation of positional information in ancestral metazoans. Comp. Biochem. Physiol. B 136, 185–196.CrossRefGoogle Scholar
Bourlat, S.J., Juliusdottir, T., Lowe, C.J., Freeman, R., Aronowicz, J., Kirschner, M., Lander, E.S., Thorndyke, M., Nakano, H., Kohn, A.B., Heyland, A., Moroz, L.L., Copley, R.R., and Telford, M.J. (2006). Deuterostome phylogeny reveals monophyletic chordates and the new phylum Xenoturbellida. Nature 444, 85–88.CrossRefGoogle ScholarPubMed
Bouzaffour, M., Dufourcq, P., Lecaudey, V., Haas, P., and Vriz, S. (2009). Fgf and Sdf-1 pathways interact during zebrafish fin regeneration. PLoS ONE 4 #6, e5824.CrossRefGoogle ScholarPubMed
Bowsher, J.H. and Nijhout, H.F. (2009). Partial co-option of the appendage patterning pathway in the development of abdominal appendages in the sepsid fly Themira biloba. Dev. Genes Evol. 219, 577–587.CrossRefGoogle ScholarPubMed
Bradley, D.G. (2003). Genetic hoofprints. Nat. Hist. 112 #1, 36–41.Google Scholar
Bragulla, H. and Hirschberg, R.M. (2003). Horse hooves and bird feathers: two model systems for studying the structure and development of highly adapted integumentary accessory organs – the role of the dermo-epidermal interface for the micro-architecture of complex epidermal structures. J. Exp. Zool. (Mol. Dev. Evol.) 298B, 140–151.CrossRefGoogle Scholar
Bragulla, H.H. and Homberger, D.G. (2009). Structure and functions of keratin proteins in simple, stratified, keratinized and cornified epithelia. J. Anat. 214, 516–559.CrossRefGoogle ScholarPubMed
Braiman, Y., Lindner, J.F., and Ditto, W.L. (1995). Taming spatiotemporal chaos with disorder. Nature 378, 465–467.CrossRefGoogle Scholar
Brakefield, P.M. (1998). The evolution-development interface and advances with the eyespot patterns of Bicyclus butterflies. Heredity 80, 265–272.CrossRefGoogle Scholar
Brakefield, P.M. (2007). Butterfly eyespot patterns and how evolutionary tinkering yields diversity. In Tinkering: The Microevolution of Development. Novartis Foundation Symposium 284. Wiley, Chichester, pp. 90–109.CrossRefGoogle Scholar
Brakefield, P.M. and French, V. (1999). Butterfly wings: the evolution of development of colour patterns. BioEssays 21, 391–401.3.0.CO;2-Q>CrossRefGoogle Scholar
Brakefield, P.M., Gates, J., Keys, D., Kesbeke, F., Wijngaarden, P.J., Monteiro, A., French, V., and Carroll, S.B. (1996). Development, plasiticity and evolution of butterfly eyespot patterns. Nature 384, 236–242.CrossRefGoogle Scholar
Brandley, M.C., Huelsenbeck, J.P., and Wiens, J.J. (2008). Rates and patterns in the evolution of snake-like body form in squamate reptiles: evidence for repeated re-evolution of lost digits and long-term persistence of intermediate body forms. Evolution 62, 2042–2064.CrossRefGoogle ScholarPubMed
Branham, M.A. and Wenzel, J.W. (2003). The origin of photic behavior and the evolution of sexual communication in fireflies (Coleoptera: Lampyridae). Cladistics 19, 1–22.CrossRefGoogle Scholar
Brear, K., Currey, J.D., Kingsley, M.C.S., and Ramsay, M. (1993). The mechanical design of the tusk of the narwhal (Monodon monocerous: Cetacea). J. Zool. Lond. 230, 411–423.CrossRefGoogle Scholar
Brena, C. and Akam, M. (2012). The embryonic development of the centipede Strigamia maritima. Dev. Biol. 363, 290–307.CrossRefGoogle ScholarPubMed
Brena, C., Chipman, A.D., Minelli, A., and Akam, M. (2006). Expression of trunk Hox genes in the centipede Strigamia maritima: sense and anti-sense transcripts. Evol. Dev. 8, 252–265.CrossRefGoogle ScholarPubMed
Breuker, C.J., Debat, V., and Klingenberg, C.P. (2006). Functional evo-devo. Trends Ecol. Evol. 21, 488–492.CrossRefGoogle ScholarPubMed
Bridges, C.B. and Morgan, T.H. (1923). The third-chromosome group of mutant characters of Drosophila melanogaster. Carnegie Inst. Wash. Publ. 327, 1–251.Google Scholar
Bridgham, J.T., Carroll, S.B., and Thornton, J.W. (2006). Evolution of hormone-receptor complexity by molecular exploitation. Science 312, 97–101.CrossRefGoogle ScholarPubMed
Brigandt, I. and Love, A.C. (2012). Conceptualizing evolutionary novelty: moving beyond definitional debates. J. Exp. Zool. (Mol. Dev. Evol.) 318B, 417–427.CrossRefGoogle Scholar
Bright, M., Klose, J., and Nussbaumer, A.D. (2013). Giant tubeworms. Curr. Biol. 23, R224–R225.CrossRefGoogle ScholarPubMed
Brischoux, F., Pizzatto, L., and Shine, R. (2010). Insights into the adaptive significance of vertical pupil shape in snakes. J. Evol. Biol. 23, 1878–1885.CrossRefGoogle ScholarPubMed
Brischoux, F. and Shine, R. (2011). Morphological adaptations to marine life in snakes. J. Morph. 272, 566–572.CrossRefGoogle ScholarPubMed
Briscoe, J., Lawrence, P.A., and Vincent, J.-P., eds. (2010). Generation and Interpretation of Morphogen Gradients. Cold Spring Harbor Laboratory Press, Woodbury, NY.
Bristow, C.M. (1985). Sugar nannies. Nat. Hist. 94 #9, 63–69.Google Scholar
Britz, R. and Johnson, G.D. (2011). Comments on the establishment of the one to one relationship between characters as a prerequisite for homology assessment in phylogenetic studies. Zootaxa 2946, 65–72.Google Scholar
Britz, R. and Johnson, G.D. (2012). Ontogeny and homology of the skeletal elements that form the sucking disc of remoras (Teleostei, Echeneoidei, Echeneidae). J. Morph. 273, 1353–1366. [See also Friedman, M., et al. (2013). An early fossil remora (Echeneoideda) reveals the evolutionary assembly of the adhesion disc. Proc. R. Soc. Lond. B 280, .]CrossRefGoogle Scholar
Brockes, J.P. and Kumar, A. (2005). Appendage regeneration in adult vertebrates and implications for regenerative medicine. Science 310, 1919–1922.CrossRefGoogle ScholarPubMed
Brockes, J.P., Kumar, A., and Velloso, C.P. (2001). Regeneration as an evolutionary variable. J. Anat. 199, 3–11.CrossRefGoogle ScholarPubMed
Brodie, C. (2007). At home in the dark. Am. Sci. 95, 460–461.Google Scholar
Brodie, E.D. (1992). Correlational selection for color pattern and antipredator behavior in the garter snake Thamnophis ordinoides. Evolution 46, 1284–1298.CrossRefGoogle ScholarPubMed
Brodie, E.D. (2009). Toxins and venoms. Curr. Biol. 19, R931–R935.CrossRefGoogle ScholarPubMed
Brodie, E.D. (2010). Convergent evolution: pick your poison carefully. Curr. Biol. 20, R152–R154.CrossRefGoogle ScholarPubMed
Brodsky, A.K. (1994). The Evolution of Insect Flight. Oxford University Press, New York, NY.Google Scholar
Brogiolo, W., Stocker, H., Ikeya, T., Rintelen, F., Fernandez, R., and Hafen, E. (2001). An evolutionarily conserved function of the Drosophila insulin receptor and insulin-like peptides in growth control. Curr. Biol. 11, 213–221.CrossRefGoogle ScholarPubMed
Bronner, M.E. and LeDouarin, N.M. (2012). Development and evolution of the neural crest: an overview. Dev. Biol. 366, 2–9.CrossRefGoogle ScholarPubMed
Brooke, M.D.L., Hanley, S., and Laughlin, S.B. (1999). The scaling of eye size with body mass in birds. Proc. R. Soc. Lond. B 266, 405–412.CrossRefGoogle Scholar
Brower, D.L. (1987). Ultrabithorax gene expression in Drosophila imaginal discs and larval nervous system. Development 101, 83–92.Google ScholarPubMed
Brown, C., Garwood, M.P., and Williamson, J.E. (2012). It pays to cheat: tactical deception in a cephalopod social signalling system. Biol. Lett. 8, 729–732.CrossRefGoogle Scholar
Brown, D.M., Brenneman, R.A., Koepfli, K.-P., Pollinger, J.P., Milá, B., Georgiadis, N.J., Louis, E.E., Grether, G.F., Jacobs, D.K., and Wayne, R.K. (2007). Extensive population genetic structure in the giraffe. BMC Biol. 5, Article 57 (13 pp.).CrossRefGoogle ScholarPubMed
Brown, J.L. (1982). The adaptationist program. Science 217, 884–886.CrossRefGoogle Scholar
Brown, L. and Rockwood, L.L. (1986). On the dilemma of horns. Nat. Hist. 95 #7, 54–61.Google Scholar
Brown, N.A. and Wolpert, L. (1990). The development of handedness in left/right asymmetry. Development 109, 1–9.Google ScholarPubMed
Brown, N.L., Patel, S., Brzezinski, J., and Glaser, T. (2001). Math5 is required for retinal ganglion cell and optic nerve formation. Development 128, 2497–2508.Google ScholarPubMed
Bruce, A.E.E. and Shankland, M. (1998). Expression of the head gene Lox22-Otx in the leech Helobdella and the origin of the bilaterian body plan. Dev. Biol. 201, 101–112.CrossRefGoogle ScholarPubMed
Bruemmer, F. (1993). The Narwhal: Unicorn of the Sea. Swan Hill Press, Shrewsbury.Google Scholar
Brunetti, C.R., Selegue, J.E., Monteiro, A., French, V., Brakefield, P.M., and Carroll, S.B. (2001). The generation and diversification of butterfly eyespot color patterns. Curr. Biol. 11, 1578–1585.CrossRefGoogle ScholarPubMed
Brusca, R.C. and Brusca, G.J. (1990). Invertebrates. Sinauer, Sunderland, MA.Google Scholar
Bryant, H.N. and Churcher, C.S. (1987). All sabretoothed carnivores aren’t sharks. Nature 325, 488. [See also Wroe, S., et al. (2013). Comparative biomechanical modeling of metatherian and placental saber-tooths: a different kind of bite for an extreme pouched predator. PLoS ONE 8 #6, e66888.]CrossRefGoogle Scholar
Bryant, P.J. (1978). Pattern formation in imaginal discs. In The Genetics and Biology of Drosophila, Vol. 2c (Ashburner, M. and Wright, T.R.F., eds.). Academic Press, New York, NY, pp. 229–335.Google Scholar
Bryant, P.J. (1993). The Polar Coordinate Model goes molecular. Science 259, 471–472.CrossRefGoogle ScholarPubMed
Bryant, P.J., Bryant, S.V., and French, V. (1977). Biological regeneration and pattern formation. Sci. Am. 237 #1, 66–81.CrossRefGoogle ScholarPubMed
Buchholtz, E.A., Bailin, H.G., Laves, S.A., Yang, J.T., Chan, M.-Y., and Drozd, L.E. (2012). Fixed cervical count and the origin of the mammalian diaphram. Evol. Dev. 14, 399–411.CrossRefGoogle Scholar
Buchholtz, E.A. and Stepien, C.C. (2009). Anatomical transformation in mammals: developmental origin of aberrant cervical anatomy in tree sloths. Evol. Dev. 11, 69–79.CrossRefGoogle ScholarPubMed
Buchtová, M., Handrigan, G.R., Tucker, A.S., Lozanoff, S., Town, L., Fu, K., Diewert, V.M., Wicking, C., and Richman, J.M. (2008). Initiation and patterning of the snake dentition are dependent on Sonic Hedgehog signaling. Dev. Biol. 319, 132–145.CrossRefGoogle ScholarPubMed
Buckley-Beason, V.A., Johnson, W.E., Nash, W.G., Stanyon, R., Menninger, J.C., Driscoll, C.A., Howard, J., Bush, M., Page, J.E., Roelke, M.E., Stone, G., Martelli, P.P., Wen, C., Ling, L., Duraisingam, R.K., Lam, P.V., and O’Brien, S.J. (2006). Molecular evidence for species-level distinctions in clouded leopards. Curr. Biol. 16, 2371–2376.CrossRefGoogle ScholarPubMed
Budd, G.E. (1999). Does evolution in body patterning genes drive morphological change – or vice versa? BioEssays 21, 326–332.3.0.CO;2-0>CrossRefGoogle Scholar
Budd, G.E. (2002). A palaeontological solution to the arthropod head problem. Nature 417, 271–275.CrossRefGoogle ScholarPubMed
Budd, G.E. (2006). On the origin and evolution of major morphological characters. Biol. Rev. 81, 609–628.CrossRefGoogle ScholarPubMed
Budd, G.E. (2008). The earliest fossil record of the animals and its significance. Phil. Trans. R. Soc. Lond. B 363, 1425–1434.CrossRefGoogle ScholarPubMed
Budd, G.E. and Telford, M.J. (2009). The origin and evolution of arthropods. Nature 457, 812–817.CrossRefGoogle ScholarPubMed
Budi, E.H., Patterson, L.B., and Parichy, D.M. (2011). Post-embryonic nerve-associated precursors to adult pigment cells: genetic requirements and dynamics of morphogenesis and differentiation. PLoS Genet. 7 #5, e1002044.CrossRefGoogle Scholar
Budrene, E.O. and Berg, H.C. (1991). Complex patterns formed by motile cells of Escherichia coli. Nature 349, 630–633.CrossRefGoogle ScholarPubMed
Buecker, C. and Wysocka, J. (2012). Enhancers as information integration hubs in development: lessons from genomics. Trends Genet. 28, 276–284.CrossRefGoogle ScholarPubMed
Bull, J.J. and Charnov, E.L. (1985). On irreversible evolution. Evolution 39, 1149–1155.CrossRefGoogle ScholarPubMed
Burke, A.C., Nelson, C.E., Morgan, B.A., and Tabin, C. (1995). Hox genes and the evolution of vertebrate axial morphology. Development 121, 333–346.Google ScholarPubMed
Burke, R.D. (2011). Deuterostome neuroanatomy and the body plan paradox. Evol. Dev. 13, 110–115.CrossRefGoogle ScholarPubMed
Burnie, D. and Wilson, D.E., eds. (2011). Animal: The Definitive Visual Guide. Dorling Kindersley, New York, NY.
Burtis, K.C. and Baker, B.S. (1989). Drosophila doublesex gene controls somatic sexual differentiation by producing alternatively spliced mRNAs encoding related sex-specific polypeptides. Cell 56, 997–1010.CrossRefGoogle ScholarPubMed
Buschbeck, E.K. and Friedrich, M. (2008). Evolution of insect eyes: tales of ancient heritage, deconstruction, reconstruction, remodeling, and recycling. Evo. Edu. Outreach 1, 448–462.CrossRefGoogle Scholar
Büschges, A., Scholz, H., and El Manira, A. (2011). New moves in motor control. Curr. Biol. 21, R513–R524.CrossRefGoogle ScholarPubMed
Bush, J.O. and Jiang, R. (2012). Palatogenesis: morphogenetic and molecular mechanisms of secondary palate development. Development 139, 231–243.CrossRefGoogle ScholarPubMed
Buss, D.M., Haselton, M.G., Shackelford, T.K., Bleske, A.L., and Wakefield, J.C. (1998). Adaptations, exaptations, and spandrels. Am. Psychol. 53, 533–548.CrossRefGoogle ScholarPubMed
Butts, T., Holland, P.W.H., and Ferrier, D.E.K. (2008). The Urbilaterian Super-Hox cluster. Trends Genet. 24, 259–262. [See also Mallo, M. and Alonso, C.R. (2013). The regulation of Hox gene expression during animal development. Development 140, 3951–3963.]CrossRefGoogle ScholarPubMed
Byrnes, G. and Spence, A.J. (2011). Ecological and biomechanical insights into the evolution of gliding in mammals. Integr. Comp. Biol. 51, 991–1001.CrossRefGoogle ScholarPubMed
Caballero, L., Benítez, M., Alvarez-Buylla, E.R., Hernández, S., Arzola, A.V., and Cocho, G. (2012). An epigenetic model for pigment patterning based on mechanical and cellular interactions. J. Exp. Zool. (Mol. Dev. Evol.) 318B, 209–223.CrossRefGoogle Scholar
Cabrera, C.V., Botas, J., and Garcia-Bellido, A. (1985). Distribution of Ultrabithorax proteins in mutants of Drosophila bithorax complex and its transregulatory genes. Nature 318, 569–571.CrossRefGoogle Scholar
Caburet, S., Cocquet, J., Vaiman, D., and Veitia, R.A. (2005). Coding repeats and evolutionary “agility”. BioEssays 27, 581–587.CrossRefGoogle ScholarPubMed
Cadigan, K.M. (2012). TCFs and Wnt/â-catenin signaling: more than one way to throw the switch. Curr. Top. Dev. Biol. 98, 1–34.CrossRefGoogle Scholar
Cadigan, K.M. and Nusse, R. (1997). Wnt signaling: a common theme in animal development. Genes Dev. 11, 3286–3305.CrossRefGoogle ScholarPubMed
Calder, N. (1984). Timescale: An Atlas of the Fourth Dimension. Chatto & Windus, London.Google Scholar
Calder, W.A., III (1978). The kiwi. Sci. Am. 239 #1, 132–142.CrossRefGoogle Scholar
Calder, W.A., III (1979). The kiwi and egg design: evolution as a package deal. BioScience 29 #8, 461–467.CrossRefGoogle Scholar
Caldwell, M.W. (2003). “Without a leg to stand on”: on the evolution and development of axial elongation and limblessness in tetrapods. Can. J. Earth Sci. 40, 573–588.CrossRefGoogle Scholar
Caldwell, M.W. and Lee, M.S.Y. (1997). A snake with legs from the marine Cretaceous of the Middle East. Nature 386, 705–709.CrossRefGoogle Scholar
Caldwell, R.L. and Dingle, H. (1976). Stomatopods. Sci. Am. 234 #1, 80–89.CrossRefGoogle Scholar
Camazine, S. (2003). Patterns in nature. Nat. Hist. 112 #5, 34–41.Google Scholar
Cameron, C.B. (2005). A phylogeny of the hemichordates based on morphological characters. Can. J. Zool. 83, 196–215.CrossRefGoogle Scholar
Campbell, G. (2002). Distalization of the Drosophila leg by graded EGF-receptor activity. Nature 418, 781–785.CrossRefGoogle ScholarPubMed
Campbell, G. and Tomlinson, A. (1995). Initiation of the proximodistal axis in insect legs. Development 121, 619–628.Google ScholarPubMed
Campbell, G. and Tomlinson, A. (1998). The roles of the homeobox genes aristaless and Distal-less in patterning the legs and wings of Drosophila. Development 125, 4483–4493.Google ScholarPubMed
Campbell, J.H. (1985). An organizational interpretation of evolution. In Evolution at a Crossroads: The New Biology and the New Philosophy of Science (Depew, D.J. and Weber, B.H., eds.). MIT Press, Cambridge, MA, pp. 133–167.Google Scholar
Campo-Paysaa, F., Marlétaz, F., Laudet, V., and Schubert, M. (2008). Retinoic acid signaling in development: tissue-specific functions and evolutionary origins. Genesis 46, 640–656.CrossRefGoogle ScholarPubMed
Cañestro, C., Albalat, R., Irimia, M., and Garcia-Fernàndez, J. (2013). Impact of gene gains, losses and duplication modes on the origin and diversification of vertebrates. Semin. Cell Dev. Biol. 24, 83–94.CrossRefGoogle ScholarPubMed
Cañestro, C. and Postlethwait, J.H. (2007). Development of a chordate anterior–posterior axis without classical retinoic acid signaling. Dev. Biol. 305, 522–538.CrossRefGoogle ScholarPubMed
Cañestro, C., Postlethwait, J.H., Gonzàlez-Duarte, R., and Albalat, R. (2006). Is retinoic acid genetic machinery a chordate innovation? Evol. Dev. 8, 394–406.CrossRefGoogle ScholarPubMed
Cañestro, C., Yokoi, H., and Postlethwait, J.H. (2007). Evolutionary developmental biology and genomics. Nature Rev. Genet. 8, 932–942.CrossRefGoogle ScholarPubMed
Cantú, J.M. and Ruiz, C. (1985). On atavisms and atavistic genes. Ann. Génét. 28 #3, 141–142.Google ScholarPubMed
Caporale, L. (2005). Darwin in the genome. BioEssays 27, 984.CrossRefGoogle ScholarPubMed
Caporale, L.H. (2003). Darwin in the Genome: Molecular Strategies in Biological Evolution. McGraw-Hill, New York, NY.Google Scholar
Caporale, L.H. (2003). Foresight in genome evolution. Am. Sci. 91, 234–241.CrossRefGoogle Scholar
Caporale, L.H. (2008). It’s not random anymore. BioEssays 30, 400–402.CrossRefGoogle Scholar
Cappellari, S.C., Schaefer, H., and Davis, C.C. (2013). Evolution: pollen or pollinators – which came first? Curr. Biol. 23, R316–R318.CrossRefGoogle ScholarPubMed
Caprette, C.L., Lee, M.S.Y., Shine, R., Mokany, A., and Downhower, J.F. (2004). The origin of snakes (Serpentes) as seen through eye anatomy. Biol. J. Linnean Soc. 81, 469–482.CrossRefGoogle Scholar
Carapuço, M., Nóvoa, A., Bobola, N., and Mallo, M. (2005). Hox genes specify vertebral types in the presomitic mesoderm. Genes Dev. 19, 2116–2121.CrossRefGoogle ScholarPubMed
Carlson, B.M. (1994). Human Embryology and Developmental Biology. Mosby, St. Louis, MO.Google Scholar
Carlson, B.M. (2007). Principles of Regenerative Biology. Academic Press, New York, NY.Google Scholar
Carmona-Fontaine, C., Matthews, H.K., Kuriyama, S., Moreno, M., Dunn, G.A., Parsons, M., Stern, C.D., and Mayor, R. (2008). Contact inhibition of locomotion in vivo controls neural crest directional migration. Nature 456, 957–961.CrossRefGoogle ScholarPubMed
Caro, T. (2009). Contrasting coloration in terrestrial mammals. Phil. Trans. R. Soc. Lond. B 364, 537–548.CrossRefGoogle ScholarPubMed
Caron, J.-B., Conway Morris, S., and Cameron, C.B. (2013). Tubicolous enteropneusts from the Cambrian period. Nature 495, 503–506.CrossRefGoogle ScholarPubMed
Carrier, D.R. (1991). Conflict in the hypaxial musculo-skeletal system: documenting an evolutionary constraint. Am. Zool. 31, 644–654.CrossRefGoogle Scholar
Carroll, C. (2005). Following the stealth hunter. Natl. Geogr. 208 #5, 66–77.Google Scholar
Carroll, L. and Gardner, M. (1960). The Annotated Alice: Alice’s Adventures in Wonderland & Through the Looking Glass. Meridian, New York, NY.Google Scholar
Carroll, R.L. and Holmes, R.B. (2007). Evolution of the appendicular skeleton of amphibians. In Fins into Limbs: Evolution, Development, and Transformation (Hall, B.K., ed.). University of Chicago Press, Chicago, IL, pp. 185–224.Google Scholar
Carroll, S. (1997). Genetics on the wing or how the butterfly got its spots. Nat. Hist. 106 #1, 28–32.Google Scholar
Carroll, S.B. (1994). Developmental regulatory mechanisms in the evolution of insect diversity. Development 1994 Suppl., 217–223.Google Scholar
Carroll, S.B. (1995). Homeotic genes and the evolution of arthropods and chordates. Nature 376, 479–485.CrossRefGoogle ScholarPubMed
Carroll, S.B. (2001). Chance and necessity: the evolution of morphological complexity and diversity. Nature 409, 1102–1109.CrossRefGoogle ScholarPubMed
Carroll, S.B. (2005). Endless Forms Most Beautiful: The New Science of Evo Devo and the Making of the Animal Kingdom. Norton, New York, NY.Google Scholar
Carroll, S.B. (2006). The Making of the Fittest: DNA and the Ultimate Forensic Record of Evolution. Norton, New York, NY.Google Scholar
Carroll, S.B. (2008). Evo-devo and an expanding evolutionary synthesis: a genetic theory of morphological evolution. Cell 134, 25–36.CrossRefGoogle ScholarPubMed
Carroll, S.B., DiNardo, S., O’Farrell, P.H., White, R.A.H., and Scott, M.P. (1988). Temporal and spatial relationships between segmentation and homeotic gene expression in Drosophila embryos: distributions of the fushi tarazu, engrailed, Sex combs reduced, Antennapedia, and Ultrabithorax proteins. Genes Dev. 2, 350–360.CrossRefGoogle ScholarPubMed
Carroll, S.B., Gates, J., Keys, D.N., Paddock, S.W., Panganiban, G.E., Selegue, J.E., and Williams, J.A. (1994). Pattern formation and eyespot determination in butterfly wings. Science 265, 109–114.CrossRefGoogle ScholarPubMed
Carroll, S.B., Grenier, J.K., and Weatherbee, S.D. (2005). From DNA to Diversity: Molecular Genetics and the Evolution of Animal Design, 2nd edn. Blackwell, Malden, MA.Google Scholar
Carroll, S.B., Weatherbee, S.D., and Langeland, J.A. (1995). Homeotic genes and the regulation and evolution of insect wing number. Nature 375, 58–61.CrossRefGoogle ScholarPubMed
Carroll, S.B. and Whyte, J.S. (1989). The role of the hairy gene during Drosophila morphogenesis: stripes in imaginal discs. Genes Dev. 3, 905–916.CrossRefGoogle Scholar
Carson, H.L., Hardy, D.E., Spieth, H.T., and Stone, W.S. (1970). The evolutionary biology of the Hawaiian Drosophilidae. In Essays in Evolution and Genetics (Hecht, M.K. and Steere, W.C., eds.). Appleton-Century-Crofts, New York, NY, pp. 437–543.Google Scholar
Carstanjen, B., Abitbol, M., and Desbois, C. (2007). Bilateral polydactyly in a foal. J. Vet. Sci. 8, 201–203.CrossRefGoogle Scholar
Casares, F. and Mann, R.S. (2001). The ground state of the ventral appendage in Drosophila. Science 293, 1477–1480.CrossRefGoogle ScholarPubMed
Casci, T. (2001). How the butterfly got its spots. Nature Rev. Gen. 2, 911.CrossRefGoogle Scholar
Cass, A.N., Servetnick, M.D., and McCune, A.R. (2013). Expression of a lung developmental cassette in the adult and developing zebrafish swimbladder. Evol. Dev. 15, 119–132.CrossRefGoogle ScholarPubMed
Castelli-Gair Hombría, J. (2011). Butterfly eyespot serial homology: enter the Hox genes. BMC Biol. 9, Article 26 (3 pp.).Google Scholar
Castelli-Gair Hombría, J. and Lovegrove, B. (2003). Beyond homeosis: HOX function in morphogenesis and organogenesis. Differentiation 71, 461–476.CrossRefGoogle Scholar
Castelli-Gair, J. (1998). Implications of the spatial and temporal regulation of Hox genes on development and evolution. Int. J. Dev. Biol. 42, 437–444.Google ScholarPubMed
Castelli-Gair, J. and Akam, M. (1995). How the Hox gene Ultrabithorax specifies two different segments: the significance of spatial and temporal regulation within metameres. Development 121, 2973–2982.Google ScholarPubMed
Castelli-Gair, J., Greig, S., Micklem, G., and Akam, M. (1994). Dissecting the temporal requirements for homeotic gene function. Development 120, 1983–1995.Google ScholarPubMed
Castro, B., Barolo, S., Bailey, A.M., and Posakony, J.W. (2005). Lateral inhibition in proneural clusters: cis-regulatory logic and default repression by Suppressor of Hairless. Development 132, 3333–3344.CrossRefGoogle ScholarPubMed
Catania, K.C. (2002). Barrels, stripes, and fingerprints in the brain: implications for theories of cortical organization. J. Neurocytol. 31, 347–358.CrossRefGoogle ScholarPubMed
Cavin, L. (2010). On giant filter feeders. Science 327, 968–969.CrossRefGoogle ScholarPubMed
Chadwick, D.H. (2001). Evolution of whales. Natl. Geogr. 200 #5, 64–77.Google Scholar
Chahda, J.S., Sousa-Neves, R., and Mizutani, C.M. (2013). Variation in the Dorsal gradient distribution is a source for modified scaling of germ layers in Drosophila. Curr. Biol. 23, 710–716. [See also De Robertis, E.M. and Colozza, G. (2013). Development: scaling to size by protease inhibition. Curr. Biol. 23, R652–R654.]CrossRefGoogle ScholarPubMed
Chan, Y.F., Marks, M.E., Jones, F.C., Villarreal, G., Jr., Shapiro, M.D., Brady, S.D., Southwick, A.M., Absher, D.M., Grimwood, J., Schmutz, J., Myers, R.M., Petrov, D., Jónsson, B., Schluter, D., Bell, M.A., and Kingsley, D.M. (2010). Adaptive evolution of pelvic reduction in sticklebacks by recurrent deletion of a Pitx1 enhancer. Science 327, 302–305.CrossRefGoogle ScholarPubMed
Chandebois, R. (1980). Cell sociology and the problem of automation in the development of pluricellular animals. Acta Biotheor. 29, 1–35.CrossRefGoogle ScholarPubMed
Chandebois, R. and Faber, J. (1983). Automation in animal development. Monographs in Developmental Biology, Vol. 16. Karger, Basel.Google Scholar
Chandler, C.H., Chari, S., and Dworkin, I. (2013). Does your gene need a background check? How genetic background impacts the analysis of mutations, genes, and evolution. Trends Genet. 29, 358–366.CrossRefGoogle Scholar
Chang, C., Wu, P., Baker, R.E., Maini, P.K., Alibardi, L., and Chuong, C.-M. (2009). Reptile scale paradigm: evo-devo, pattern formation and regeneration. Int. J. Dev. Biol. 53, 813–826.CrossRefGoogle ScholarPubMed
Chang, K.C., Wang, C., and Wang, H. (2012). Balancing self-renewal and differentiation by asymmetric division: insights from brain tumor suppressors in Drosophila neural stem cells. BioEssays 34, 301–310.CrossRefGoogle ScholarPubMed
Chapman, T., Pomiankowski, A., and Fowler, K. (2005). Stalk-eyed flies. Curr. Biol. 15, R533–R535.CrossRefGoogle ScholarPubMed
Charles, C., Hovorakova, M., Ahn, Y., Lyons, D.B., Marangoni, P., Churava, S., Biehs, B., Jheon, A., Lesot, H., Balooch, G., Krumlauf, R., Viriot, L., Peterkova, R., and Klein, O.D. (2011). Regulation of tooth number by fine-tuning levels of receptor-tyrosine kinase signaling. Development 138, 4063–4073.CrossRefGoogle ScholarPubMed
Chatterjee, S. and Templin, R.J. (2004). Posture, locomotion, and paleoecology of pterosaurs. Geol. Soc. Am. Spec. Ppr. 376, 1–64.Google Scholar
Chea, H.K., Wright, C.V., and Swalla, B.J. (2005). Nodal signaling and the evolution of deuterostome gastrulation. Dev. Dynamics 234, 269–278.CrossRefGoogle ScholarPubMed
Chen, H., Xu, Z., Mei, C., Yu, D., and Small, S. (2012). A system of repressor gradients spatially organizes the boundaries of Bicoid-dependent target genes. Cell 149, 618–629.CrossRefGoogle ScholarPubMed
Chen, J. and Chuong, C.-M. (2011). Patterning skin by planar cell polarity: the multi-talented hair designer. Exp. Dermatol. 21, 81–85.CrossRefGoogle Scholar
Chen, M.-H., Wilson, C.W., and Chuang, P.-T. (2007). Snapshot: Hedgehog signaling pathway. Cell 130, 386.CrossRefGoogle ScholarPubMed
Chen, S., Lee, A.Y., Bowens, N.M., Huber, R., and Kravitz, E.A. (2002). Fighting fruit flies: a model system for the study of aggression. PNAS 99 #8, 5664–5668.CrossRefGoogle Scholar
Chen, S., Lin, B.-Z., Baig, M., Mitra, B., Lopes, R.J., Santos, A.M., Magee, D.A., Azevedo, M., Tarroso, P., Sasazaki, S., Ostrowski, S., Mahgoub, O., Chaudhuri, T.K., Zhang, Y.-P., Costa, V., Royo, L.J., Goyache, F., Luikart, G., Boivin, N., Fuller, D.Q., Mannen, H., Bradley, D.G., and Beja-Pereira, A. (2010). Zebu cattle are an exclusive legacy of the South Asia Neolithic. Mol. Biol. Evol. 27 #1, 1–6.CrossRefGoogle ScholarPubMed
Chen, X., McClusky, R., Chen, J., Beaven, S.W., Tontonoz, P., Arnold, A.P., and Reue, K. (2012). The number of X chromosomes causes sex differences in adiposity in mice. PLoS Genet. 8 #5, e1002709.CrossRefGoogle ScholarPubMed
Chen, Y., Knezevic, V., Ervin, V., Hutson, R., Ward, Y., and Mackem, S. (2004). Direct interaction with Hoxd proteins reverses Gli3-repressor function to promote digit formation downstream of Shh. Development 131, 2339–2347.CrossRefGoogle ScholarPubMed
Cherry, A.B.C. and Daley, G.Q. (2012). Reprogramming cellular identity for regenerative medicine. Cell 148, 1110–1122.CrossRefGoogle ScholarPubMed
Chesebro, J., Hrycaj, S., Mahfooz, N., and Popadic, A. (2009). Diverging functions of Scr between embryonic and post-embryonic development in a hemimetabolous insect, Oncopeltus fasciatus. Dev. Biol. 329, 142–151.CrossRefGoogle Scholar
Chiappe, L.M., Codorniú, L., Grellet-Tinner, G., and Rivarola, D. (2004). Argentinian unhatched pterosaur fossil. Nature 432, 571.CrossRefGoogle ScholarPubMed
Chiappe, L.M. and Rivarola, D. (1996). Pterodaustro’s smile. Nat. Hist. 105 #11, 34–35.Google Scholar
Child, C.M. (1941). Patterns and Problems of Development. University of Chicago Press, Chicago, IL.CrossRefGoogle Scholar
Chipman, A.D. (2009). On making a snake. Evol. Dev. 11, 3–5.CrossRefGoogle ScholarPubMed
Chipman, A.D. (2010). Parallel evolution of segmentation by co-option of ancestral gene regulatory networks. BioEssays 32, 60–70.CrossRefGoogle ScholarPubMed
Chipman, A.D. and Akam, M. (2008). The segmentation cascade in the centipede Strigamia maritima: involvement of the Notch pathway and pair-rule gene homologues. Dev. Biol. 319, 160–169.CrossRefGoogle ScholarPubMed
Chipman, A.D., Arthur, W., and Akam, M. (2004). A double segment periodicity underlies segment generation in centipede development. Curr. Biol. 14, 1250–1255.CrossRefGoogle ScholarPubMed
Cho, E.H. and Nijhout, H.F. (2013). Development of polyploidy of scale-building cells in the wings of Manduca sexta. Arthropod Struct. Dev. 42, 37–46.CrossRefGoogle ScholarPubMed
Chouard, T. (2010). Revenge of the hopeful monster. Nature 463, 864–867.CrossRefGoogle ScholarPubMed
Christiaen, L., Jaszczyszyn, Y., Kerfant, M., Kano, S., Thermes, V., and Joly, J.-S. (2007). Evolutionary modification of mouth position in deuterostomes. Semin. Cell Dev. Biol. 18, 502–511.CrossRefGoogle ScholarPubMed
Christiansen, A.E., Keisman, E.L., Ahmad, S.M., and Baker, B.S. (2002). Sex comes in from the cold: the integration of sex and pattern. Trends Genet. 18, 510–516.CrossRefGoogle ScholarPubMed
Christiansen, P. (2008). Evolution of skull and mandible shape in cats (Carnivora: Felidae). PLoS ONE 3 #7, e2807.CrossRefGoogle Scholar
Christiansen, P. (2008). Species distinction and evolutionary differences in the clouded leopard (Neofelis nebulosa) and Diard’s clouded leopard (Neofelis diardi). J. Mammalogy 89, 1435–1446.CrossRefGoogle Scholar
Christiansen, P. (2012). The making of a monster: postnatal ontogenetic changes in craniomandibular shape in the great sabercat Smilodon. PLoS ONE 7 #1, e29699.CrossRefGoogle ScholarPubMed
Christin, P.-A., Weinreich, D.M., and Besnard, G. (2010). Causes and evolutionary significance of genetic convergence. Trends Genet. 26, 400–405.CrossRefGoogle ScholarPubMed
Chuong, C.-M., Patel, N., Lin, J., Jung, H.-S., and Widelitz, R.B. (2000). Sonic hedgehog signaling pathway in vertebrate epithelial appendage morphogenesis: perspectives in development and evolution. Cell. Mol. Life Sci. 57, 1672–1681.CrossRefGoogle ScholarPubMed
Chuong, C.-M., Randall, V.A., Widelitz, R.B., Wu, P., and Jiang, T.-X. (2012). Physiological regeneration of skin appendages and implications for regenerative medicine. Physiology 27, 61–72.CrossRefGoogle ScholarPubMed
Chuong, C.M., Hou, L., Chen, P.J., Wu, P., Patel, N., and Chen, Y. (2001). Dinosaur’s feather and chicken’s tooth? Tissue engineering of the integument. Eur. J. Dermatol. 11, 286–292.Google ScholarPubMed
Chuong, C.M., Nickoloff, B.J., Elias, P.M., Goldsmith, L.A., Macher, E., Maderson, P.A., Sundberg, J.P., Tagami, H., Plonka, P.M., Thestrup-Pedersen, K., Bernard, B.A., Schröder, J.M., Dotto, P., Chang, C.H., Williams, M.L., Feingold, K.R., King, L.E., Kligman, A.M., Rees, J.L., and Christophers, E. (2002). What is the ‘true’ function of skin? Exp. Dermatol. 11, 159–187.Google Scholar
Cieslak, M., Reissmann, M., Hofreiter, M., and Ludwig, A. (2011). Colours of domestication. Biol. Rev. 86, 885–899.CrossRefGoogle ScholarPubMed
Cisneros, J.C., Abdala, F., Rubidge, B.S., Dentzien-Dias, P.C., and de Oliveira Bueno, A. (2011). Dental occlusion in a 260-million-year-old therapsid with saber canines from the Permian of Brazil. Science 331, 1603–1605.CrossRefGoogle Scholar
Citerne, H., Jabbour, F., Nadot, S., and Damerval, C. (2010). The evolution of floral symmetry. Adv. Bot. Res. 54, 85–137.CrossRefGoogle Scholar
Clark, J.W. (1871). On the skeleton of a narwhal (Monodon monoceros) with two fully developed tusks. Proc. Zool. Soc. Lond. 1871 #4, 42–53.Google Scholar
Clarke, J. (2013). Feathers before flight. Science 340, 690–692.CrossRefGoogle ScholarPubMed
Clarke, S.L., VanderMeer, J.E., Wenger, A.M., Schaar, B.T., Ahituv, N., and Bejerano, G. (2012). Human developmental enhancers conserved between deuterostomes and protostomes. PLoS Genet. 8 #8, e1002852.CrossRefGoogle ScholarPubMed
Clevers, H. and Nusse, R. (2012). Wnt/β-catenin signaling and disease. Cell 149, 1192–1205.CrossRefGoogle ScholarPubMed
Clyne, J.D. and Miesenböck, G. (2008). Sex-specific control and tuning of the pattern generator for courtship song in Drosophila. Cell 133, 354–363.CrossRefGoogle Scholar
Coates, M. and Ruta, M. (2000). Nice snake, shame about the legs. Trends Ecol. Evol. 15, 503–507.CrossRefGoogle ScholarPubMed
Coates, M.I. and Clack, J.A. (1990). Polydactyly in the earliest known tetrapod limbs. Nature 347, 66–69.CrossRefGoogle Scholar
Cobb, J. and Duboule, D. (2005). Comparative analysis of genes downstream of the Hoxd cluster in developing digits and external genitalia. Development 132, 3055–3067.CrossRefGoogle ScholarPubMed
Cobourne, M.T. and Mitsiadis, T. (2006). Neural crest cells and patterning of the mammalian dentition. J. Exp. Zool. (Mol. Dev. Evol.) 306B, 251–260.CrossRefGoogle Scholar
Coen, E. (1999). The Art of Genes: How Organisms Make Themselves. Oxford University Press, New York, NY.Google Scholar
Cohen, M. (2011). The importance of structured noise in the generation of self-organizing tissue patterns through contact-mediated cell-cell signaling. J. Roy. Soc. Interface 8, 787–798.CrossRefGoogle Scholar
Cohen, M., Georgiou, M., Stevenson, N.L., Miodownik, M., and Baum, B. (2010). Dynamic filopodia transmit intermittent Delta-Notch signaling to drive pattern refinement during lateral inhibition. Dev. Cell 19, 78–89.CrossRefGoogle ScholarPubMed
Cohen, M.M., Jr. (2012). Perspectives on asymmetry: The Erickson lecture. Am. J. Med. Genet. A 158A, 2981–2998.CrossRefGoogle ScholarPubMed
Cohen, S.M. and Jürgens, G. (1989). Proximal-distal pattern formation in Drosophila: cell autonomous requirement for Distal-less gene activity in limb development. EMBO J. 8, 2045–2055.Google ScholarPubMed
Cohen, S.M. and Jürgens, G. (1989). Proximal-distal pattern formation in Drosophila: graded requirement for Distal-less gene activity during limb development. Roux’s Arch. Dev. Biol. 198, 157–169.CrossRefGoogle ScholarPubMed
Cohn, M.J. (2011). Development of the external genitalia: conserved and divergent mechanisms of appendage patterning. Dev. Dynamics 240, 1108–1115.CrossRefGoogle ScholarPubMed
Cohn, M.J., Lovejoy, C.O., Wolpert, L., and Coates, M.I. (2002). Branching, segmentation and the metapterygial axis: pattern versus process in the vertebrate limb. BioEssays 24, 460–465.CrossRefGoogle ScholarPubMed
Cohn, M.J., Patel, K., Krumlauf, R., Wilkinson, D.G., Clarke, J.D.W., and Tickle, C. (1997). Hox9 genes and vertebrate limb specification. Nature 387, 97–101.CrossRefGoogle ScholarPubMed
Cohn, M.J. and Tickle, C. (1999). Developmental basis of limblessness and axial patterning in snakes. Nature 399, 474–479.CrossRefGoogle ScholarPubMed
Colas, J.-F. and Schoenwolf, G.C. (2001). Towards a cellular and molecular understanding of neurulation. Dev. Dynamics 221, 117–145.CrossRefGoogle ScholarPubMed
Cole, E.S. and Palka, J. (1982). The pattern of campaniform sensilla on the wing and haltere of Drosophila melanogaster and several of its homeotic mutants. J. Embryol. Exp. Morph. 71, 41–61.Google ScholarPubMed
Collin, R. and Miglietta, M.P. (2008). Reversing opinions on Dollo’s Law. Trends Ecol. Evol. 23, 602–609.CrossRefGoogle ScholarPubMed
Collins, M.A. (2003). The genus Grimpoteuthis (Octopoda: Grimpoteuthidae) in the north-east Atlantic, with descriptions of three new species. Zool. J. Linnean Soc. 139, 93–127.CrossRefGoogle Scholar
Comeron, J.M., Ratnappan, R., and Bailin, S. (2012). The many landscapes of recombination in Drosophila melanogaster. PLoS Genet. 8 #10, e1002905.CrossRefGoogle ScholarPubMed
Compagnucci, C., Debiais-Thibaud, M., Coolen, M., Fish, J., Griffin, J.N., Bertocchini, F., Minoux, M., Rijli, F.M., Borday-Birraux, V., Casane, D., Mazan, S., and Depew, M.J. (2013). Pattern and polarity in the development and evolution of the gnathostome jaw: both conservation and heterotopy in the branchial arches of the shark, Scyliorhinus canicula. Dev. Biol. 377, 428–448.CrossRefGoogle ScholarPubMed
Conant, G.C. and Wolfe, K.H. (2008). Turning a hobby into a job: how duplicated genes find new functions. Nature Rev. Genet. 9, 938–950.CrossRefGoogle ScholarPubMed
Conceição, I.C., Long, A.D., Gruber, J.D., and Beldade, P. (2011). Genomic sequence around butterfly wing development genes: annotation and comparative analysis. PLoS ONE 6 #8, e23778.CrossRefGoogle ScholarPubMed
Conniff, R. (1999). Cheetahs: ghosts of the grasslands. Natl. Geogr. 196 #6, 2–31.Google Scholar
Conway, B.R. and Rehding, A. (2013). Neuroaesthetics and the trouble with beauty. PLoS Biol. 11 #3, e1001504.CrossRefGoogle ScholarPubMed
Conway Morris, S. (2000). Evolution: Bringing molecules into the fold. Cell 100, 1–11.CrossRefGoogle Scholar
Conway-Morris, S. (2003). The Cambrian “explosion” of metazoans and molecular biology: would Darwin be satisfied?Int. J. Dev. Biol. 47, 505–515.Google ScholarPubMed
Conway-Morris, S. (2006). Evolutionary convergence. Curr. Biol. 16, R826–R827.CrossRefGoogle Scholar
Cook, O., Biehs, B., and Bier, E. (2004). brinker and optomotor-blind act coordinately to initiate development of the L5 wing vein primoridium in Drosophila. Development 131, 2113–2124.CrossRefGoogle Scholar
Cook, T. and Desplan, C. (2001). Photoreceptor subtype specification: from flies to humans. Semin. Cell Dev. Biol. 12, 509–518.CrossRefGoogle Scholar
Cook, T.A. (1914). The Curves of Life. Constable, London.Google Scholar
Cooke, J. (1975). The emergence and regulation of spatial organization in early animal development. Annu. Rev. Biophys. Bioeng. 4, 185–217.CrossRefGoogle ScholarPubMed
Cooke, J. (1981). The problem of periodic patterns in embryos. Phil. Trans. Roy. Soc. Lond. B 295, 509–524.CrossRefGoogle ScholarPubMed
Cooke, J. (1981). Scale of body pattern adjusts to available cell number in amphibian embryos. Nature 290, 775–778.CrossRefGoogle ScholarPubMed
Cooke, J. (1982). The relation between scale and the completeness of pattern in vertebrate embryogenesis: models and experiments. Am. Zool. 22, 91–104.CrossRefGoogle Scholar
Cooke, J. (1998). A gene that resuscitates a theory: somitogenesis and a molecular oscillator. Trends Genet. 14, 85–88.CrossRefGoogle Scholar
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, 455–476.CrossRefGoogle ScholarPubMed
Coombs, W.P., Jr. (1978). Theoretical aspects of cursorial adaptations in dinosaurs. Q. Rev. Biol. 53, 393–418.CrossRefGoogle Scholar
Cooper, C.D. and Raible, D.W. (2009). Mechanisms for reaching the differentiated state: insights from neural crest-derived melanocytes. Semin. Cell Dev. Biol. 20, 105–110.CrossRefGoogle ScholarPubMed
Cooper, H.M., Herbin, M., and Nevo, E. (1993). Ocular regression conceals adaptive progression of the visual system in a blind subterranean mammal. Nature 361, 156–159.CrossRefGoogle Scholar
Cooper, K.L. (2011). The lesser Egyptian jerboa, Jaculus jaculus: a unique rodent model for evolution and development. Cold Spr. Harb. Protoc. 2011 #12, 1451–1456.Google ScholarPubMed
Cooper, K.L., Oh, S., Sung, Y., Dasari, R.R., Kirschner, M.W., and Tabin, C.J. (2013). Multiple phases of chondrocyte enlargement underlie differences in skeletal proportions. Nature 495, 375–378.CrossRefGoogle ScholarPubMed
Cooper, K.L. and Tabin, C.J. (2008). Understanding of bat wing evolution takes flight. Genes Dev. 22, 121–124.CrossRefGoogle ScholarPubMed
Cooper, L.N., Armfield, B.A., and Thewissen, J.G.M. (2011). Evolution of the apical ectoderm in the developing vertebrate limb. In Epigenetics: Linking Genotype and Phenotype in Development and Evolution (Hallgrímsson, B. and Hall, B.K., eds.). University of California Press, Berkeley, CA, pp. 238–255.Google Scholar
Cooper, L.N., Berta, A., Dawson, S.D., and Reidenberg, J.S. (2007). Evolution of hyperphalangy and digit reduction in the cetacean manus. Anat. Rec. 290, 654–672.CrossRefGoogle ScholarPubMed
Corbet, P.S. (1999). Dragonflies: Behavior and Ecology of Odonata. Cornell University Press, Ithaca, NY.Google Scholar
Cordes, R., Schuster-Gossler, K., Serth, K., and Gossier, A. (2004). Specification of vertebral identity is coupled to Notch signalling and the segmentation clock. Development 131, 1221–1233.CrossRefGoogle ScholarPubMed
Corfield, J.R., Gsell, A.C., Brunton, D., Heesy, C.P., Hall, M.I., Acosta, M.L., and Iwaniuk, A.N. (2011). Anatomical specializations for nocturnality in a critically endangered parrot, the Kakapo (Strigops habroptilus). PLoS ONE 6 #8, e22945.CrossRefGoogle Scholar
Cornwallis, C.K. and Uller, T. (2008). Towards an evolutionary ecology of sexual traits. Trends Ecol. Evol. 25, 145–152.CrossRefGoogle Scholar
Costanzo, K. and Monteiro, A. (2007). The use of chemical and visual cues in female choice in the butterfly Bicyclus anynana. Proc. Roy. Soc. Lond. B 274, 845–851.CrossRefGoogle ScholarPubMed
Couso, J.P. (2009). Segmentation, metamerism and the Cambrian explosion. Int. J. Dev. Biol. 53, 1305–1316.CrossRefGoogle ScholarPubMed
Couso, J.P., Bate, M., and Martínez-Arias, A. (1993). A wingless-dependent polar coordinate system in Drosophila imaginal discs. Science 259, 484–489.CrossRefGoogle ScholarPubMed
Couso, J.P. and Bishop, S.A. (1998). Proximo-distal development in the legs of Drosophila. Int. J. Dev. Biol. 42, 345–352.Google ScholarPubMed
Coutelis, J.-B., Géminard, C., Spéder, P., Suzanne, M., Petzoldt, A.G., and Noselli, S. (2013). Drosophila left/right asymmetry establishment is controlled by the Hox gene Abdominal-B. Dev. Cell 24, 89–97.CrossRefGoogle ScholarPubMed
Coutelis, J.B., Petzoldt, A.G., Spéder, P., Suzanne, M., and Noselli, S. (2008). Left-right asymmetry in Drosophila. Semin. Cell Dev. Biol. 19, 252–262.CrossRefGoogle ScholarPubMed
Couturier, L., Vodovar, N., and Schweisguth, F. (2012). Endocytosis by Numb breaks Notch symmetry at cytokinesis. Nat. Cell Biol. 14, 131–139.CrossRefGoogle ScholarPubMed
Cox, P.G., Rayfield, E.J., Fagan, M.J., Herrel, A., Pataky, T.C., and Jeffery, N. (2012). Functional evolution of the feeding system in rodents. PLoS ONE 7 #4, e36299.CrossRefGoogle ScholarPubMed
Crampton, G. (1916). The phylogenetic origin and the nature of the wings of insects according to the paranotal theory. N. Y. Entomol. Soc. 24, 1–39 (plus 2 plates).Google Scholar
Cretekos, C.J., Deng, J.-M., Green, E.D., Rasweiler, J.J., IV, and Behringer, R.R. (2007). Isolation, genomic structure and developmental expression of Fgf8 in the short-tailed fruit bat, Carollia perspicillata. Int. J. Dev. Biol. 51, 333–338.CrossRefGoogle ScholarPubMed
Cretekos, C.J., Wang, Y., Green, E.D., Martin, J.F., Rasweiler, J.J., IV, and Behringer, R.R. (2008). Regulatory divergence modifies limb length between mammals. Genes Dev. 22, 141–151.CrossRefGoogle ScholarPubMed
Cretekos, C.J., Weatherbee, S.D., Chen, C.-H., Badwaik, N.K., Niswander, L., Behringer, R.R., and Rasweiler, J.J., IV (2005). Embryonic staging system for the short-tailed fruit bat, Carollia perspicillata, a model organism for the mammalian order Chiroptera, based upon timed pregnancies in captive-bred animals. Dev. Dynamics 233, 721–738.CrossRefGoogle Scholar
Crew, B. (2012). Zombie Birds, Astronaut Fish, and Other Weird Animals. Adams Media, Avon, MA. [See also Henderson, C. (2013). The Book of Barely Imagined Beings: A 21st Century Bestiary. University of Chicago Press, Chicago, IL, which contains its own alphabetized listing of exotic animals. N.B.: The author learned of this book after submitting his own manuscript and was shocked to see the same animals for several letters. When he informed Mr. Henderson of this coincidence, the latter kindly dismissed any concern about overlap. [See also Piper, R. (2013). Animal Earth. Thames & Hudson, New York, NY.]Google Scholar
Crickmore, M.A. and Mann, R.S. (2006). Hox control of organ size by regulation of morphogen production and mobility. Science 313, 63–68.CrossRefGoogle ScholarPubMed
Crickmore, M.A. and Mann, R.S. (2007). Hox control of morphogen mobility and organ development through regulation of glypican expression. Development 134, 327–334.CrossRefGoogle ScholarPubMed
Crickmore, M.A. and Mann, R.S. (2008). The control of size in animals: insights from selector genes. BioEssays 30, 843–853.CrossRefGoogle ScholarPubMed
Crispo, E. (2007). The Baldwin effect and genetic assimilation: revisiting two mechanisms of evolutionary change mediated by phenotypic plasticity. Evolution 61, 2469–2479.CrossRefGoogle ScholarPubMed
Crocker, J., Tamori, Y., and Erives, A. (2008). Evolution acts on enhancer organization to fine-tune gradient threshold readouts. PLoS Biol. 6 #11, e263.CrossRefGoogle ScholarPubMed
Cuadrado, M., Martín, J., and López, P. (2001). Camouflage and escape decisions in the common chameleon Chamaeleo chamaeleon. Biol. J. Linnean Soc. 72, 547–554.CrossRefGoogle Scholar
Cullen, K.E. (2011). The neural encoding of self-motion. Curr. Opin. Neurobiol. 21, 587–595.CrossRefGoogle ScholarPubMed
Cully, M. and Downward, J. (2008). Snapshot: Ras signaling. Cell 133, 1292.CrossRefGoogle ScholarPubMed
Cummins, H. and Midlo, C. (1943). Finger Prints, Palms and Soles. An Introduction to Dermatoglyphics. Dover, New York, NY.Google Scholar
Cundall, D. (2009). Viper fangs: functional limitations of extreme teeth. Physiol. Biochem. Zool. 82, 63–79.CrossRefGoogle ScholarPubMed
Cundall, D. and Deufel, A. (2006). Influence of the venom delivery system on intraoral prey transport in snakes. Zool. Anz. 245, 193–210.CrossRefGoogle Scholar
Cunningham, S.J., Alley, M.R., and Castro, I. (2011). Facial bristle feather histology and morphology in New Zealand birds: implications for function. J. Morph. 272, 118–128.CrossRefGoogle ScholarPubMed
Cuthill, I.C. (2007). Animal behaviour: strategic signalling by cephalopods. Curr. Biol. 17, R1059–R1060.CrossRefGoogle ScholarPubMed
Dahmann, C., Oates, A.C., and Brand, M. (2011). Boundary formation and maintenance in tissue development. Nature Rev. Genet. 12, 43–55. [See also Bard, J. (2013). Driving developmental and evolutionary change: a systems biology view. Prog. Biophys. Mol. Biol. 111, 83–91.]CrossRefGoogle ScholarPubMed
Dalton, S. (1975). Borne on the Wind: The Extraordinary World of Insects in Flight. Reader’s Digest Press, New York, NY.Google Scholar
Damen, W.G.M. (2004). Arthropod segmentation: why centipedes are odd. Curr. Biol. 14, R557–R559.CrossRefGoogle ScholarPubMed
Damen, W.G.M., Saridaki, T., and Averof, M. (2002). Diverse adaptations of an ancestral gill: a common evolutionary origin for wings, breathing organs, and spinnerets. Curr. Biol. 12, 1711–1716.CrossRefGoogle ScholarPubMed
Darwin, C. (1859). On the Origin of Species by Means of Natural Selection, or the Preservation of Favoured Races in the Struggle for Life. John Murray, London.Google Scholar
Darwin, C. (1868). The Variation of Animals and Plants Under Domestication. John Murray, London.Google Scholar
Darwin, C. (1871). The Descent of Man, and Selection in Relation to Sex. John Murray, London.Google Scholar
Darwin, C. (1877). The Various Contrivances by which Orchids are Fertilised by Insects, 2nd (revised) edn. John Murray, London.Google Scholar
Davidson, E.H. (1990). How embryos work: a comparative view of diverse modes of cell fate specification. Development 108, 365–389.Google ScholarPubMed
Davidson, E.H., Cameron, R.A., and Ransick, A. (1998). Specification of cell fate in the sea urchin embryo: summary and some proposed mechansisms. Development 125, 3269–3290.Google Scholar
Davidson, E.H. and Erwin, D.H. (2006). Gene regulatory networks and the evolution of animal body plans. Science 311, 796–800.CrossRefGoogle ScholarPubMed
Davidson, E.H., Peterson, K.J., and Cameron, R.A. (1995). Origin of bilaterian body plans: evolution of developmental regulatory mechanisms. Science 270, 1319–1325.CrossRefGoogle ScholarPubMed
Davis, B.W., Li, G., and Murphy, W.J. (2010). Supermatrix and species tree methods resolve phylogenetic relationships within the big cats, Panthera (Carnivora: Felidae). Mol. Phylogenet. Evol. 56, 64–76.CrossRefGoogle Scholar
Davis, D.D. (1964). The Giant Panda: A Morphological Study of Evolutionary Mechanisms. Fieldiana: Zoology Memoirs, Vol. 3. Chicago Natural History Museum, Chicago, IL.CrossRefGoogle Scholar
Davis, G.K. and Patel, N.H. (2003). Playing by pair-rules? BioEssays 25, 425–429.CrossRefGoogle ScholarPubMed
Davis, G.K., Srinivasan, D.G., Wittkopp, P.J., and Stern, D.L. (2007). The function and regulation of Ultrabithorax in the legs of Drosophila melanogaster. Dev. Biol. 308, 621–631.CrossRefGoogle ScholarPubMed
Davis, M.C., Dahn, R.D., and Shubin, N.H. (2007). An autopodial-like pattern of Hox expression in the fins of a basal actinopterygian fish. Nature 447, 473–476.CrossRefGoogle ScholarPubMed
Davit-Béal, T., Tucker, A.S., and Sire, J.-Y. (2009). Loss of teeth and enamel in tetrapods: fossil record, genetic data and morphological adaptations. J. Anat. 214, 477–501.CrossRefGoogle ScholarPubMed
Dawkins, R. (1982). The Extended Phenotype. Oxford University Press, New York, NY.Google Scholar
Dawkins, R. (1986). The Blind Watchmaker: Why the Evidence of Evolution Reveals a Universe Without Design. Norton, New York, NY.Google Scholar
Dawkins, R. (1989). The evolution of evolvability. In Artificial Life, Santa Fe Institute Studies in the Sciences of Complexity, Vol. 6 (Langton, C.G., ed.). Addison-Wesley, New York, NY, pp. 201–220.Google Scholar
Dawkins, R. (1995). The evolved imagination: animals as models of their world. Nat. Hist. 104 #9, 8–11, 22–24.Google Scholar
Dawkins, R. (2004). The Ancestor’s Tale: A Pilgrimage to the Dawn of Evolution. Houghton Mifflin, New York, NY.Google Scholar
Dawkins, R. (2009). The Greatest Show on Earth: The Evidence for Evolution. Free Press, New York, NY. [N.B.: All of Richard’s books are worth reading, especially for evo-devotees, and this book is no exception, but my personal favorite is his Unweaving the Rainbow: Science, Delusion and the Appetite for Wonder (1998).]Google Scholar
de Bruijn, S., Angenent, G.C., and Kaufmann, K. (2012). Plant “evo-devo” goes genomic: from candidate genes to regulatory networks. Trends Plant Sci. 17, 441–447.CrossRefGoogle ScholarPubMed
de Celis, J.F. (2004). The Notch signaling module. In Modularity in Development and Evolution (Schlosser, G. and Wagner, G.P., eds.). University of Chicago Press, Chicago, IL, pp. 81–100.Google Scholar
de Celis, J.F. and Diaz-Benjumea, J. (2003). Developmental basis for vein pattern variations in insect wings. Int. J. Dev. Biol. 47, 653–663.Google ScholarPubMed
de Duve, C. (1996). The gospel of contingency. Nature 383, 771–772.CrossRefGoogle Scholar
De Jong, M.C.M. and Sabelis, M.W. (1991). Limits to runaway sexual selection: the wallflower paradox. J. Evol. Biol. 4, 637–655.CrossRefGoogle Scholar
de Joussineau, C., Soulé, J., Martin, M., Anguille, C., Montcourrier, P., and Alexandre, D. (2003). Delta-promoted filopodia mediate long-range lateral inhibition in Drosophila. Nature 426, 555–559.CrossRefGoogle ScholarPubMed
de Kok-Mercado, F., Habib, M., Phelps, T., Gregg, L., and Gailloud, P. (2013). Adaptations of the owl’s cervical and cephalic arteries in relation to extreme neck rotation. Science 339, 514.Google Scholar
de Lussanet, M.H.E. (2011). A hexamer origin of the echinoderms’ five rays. Evol. Dev. 13, 228–238.CrossRefGoogle ScholarPubMed
de Muizon, C. (1993). Walrus-like feeding adaptation in a new cetacean from the Pliocene of Peru. Nature 365, 745–748.CrossRefGoogle Scholar
de Navas, L.F., Garaulet, D.L., and Sánchez-Herrero, E. (2006). The Ultrabithorax Hox gene of Drosophila controls haltere size by regulating the Dpp pathway. Development 133, 4495–4506.CrossRefGoogle ScholarPubMed
De Robertis, E.M. (2008). Evo-devo: variations on ancestral themes. Cell 132, 185–195.CrossRefGoogle ScholarPubMed
De Robertis, E.M. (2008). The molecular ancestry of segmentation mechanisms. PNAS 105 #43, 16411–16412.CrossRefGoogle ScholarPubMed
De Robertis, E.M. and Sasai, Y. (1996). A common plan for dorsoventral patterning in Bilateria. Nature 380, 37–40.CrossRefGoogle ScholarPubMed
Decraene, L.P.R. and Smets, E.F. (1994). Merosity in flowers: definition, origin, and taxonomic significance. Plant Syst. Evol. 191, 83–104.CrossRefGoogle Scholar
Deforest, M.E. and Basrur, P.K. (1979). Malformations and the Manx syndrome in cats. Can. Vet. J. 20, 304–314.Google ScholarPubMed
del Álamo, D., Terriente, J., and Díaz-Benjumea, F.J. (2002). Spitz/EGFr signalling via the Ras/MAPK pathway mediates the induction of bract cells in Drosophila legs. Development 129, 1975–1982.Google ScholarPubMed
Delanoue, R. and Léopold, P. (2010). Developmental biology: a DOR connecting growth and clocks. Curr. Biol. 20, R884–R886.CrossRefGoogle ScholarPubMed
DeLaurier, A., Schweitzer, R., and Logan, M. (2006). Pitx1 determines the morphology of muscle, tendon, and bones of the hindlimb. Dev. Biol. 299, 22–34.CrossRefGoogle ScholarPubMed
Delpretti, S., Zakany, J., and Duboule, D. (2012). A function for all posterior Hoxd genes during digit development?Dev. Dynamics 241, 792–802.CrossRefGoogle ScholarPubMed
Delsuc, F., Superina, M., Tilak, M.-K., Douzery, E.J.P., and Hassanin, A. (2012). Molecular phylogenetics unveils the ancient evolutionary origins of the enigmatic fairy armadillos. Mol. Phylogenet. Evol. 62, 673–680.CrossRefGoogle ScholarPubMed
Denes, A.S., Jékely, G., Steinmetz, P.R.H., Raible, F., Snyman, H., Prud’homme, B., Ferrier, D.E.K., Balavoine, G., and Arendt, D. (2007). Molecular architecture of annelid nerve cord supports common origin of nervous system centralization in Bilateria. Cell 129, 277–288.CrossRefGoogle ScholarPubMed
Denver, R.J. (2008). Chordate metamorphosis: ancient control by iodothyronines. Curr. Biol. 18, R567–R569.CrossRefGoogle ScholarPubMed
Depew, M.J., Simpson, C.A., Morasso, M., and Rubenstein, J.L.R. (2005). Reassessing the Dlx code: the genetic regulation of branchial arch skeletal pattern and development. J. Anat. 207, 501–561.CrossRefGoogle ScholarPubMed
Dequéant, M.-L. and Pourquié, O. (2008). Segmental patterning of the vertebrate embryonic axis. Nature Rev. Gen. 9, 370–382.CrossRefGoogle ScholarPubMed
Derby, C.D. (2007). Escape by inking and secreting: marine molluscs avoid predators through a rich array of chemicals and mechanisms. Biol. Bull. 213, 274–289.CrossRefGoogle ScholarPubMed
Derelle, R., Lopez, P., Le Guyader, H., and Manuel, M. (2007). Homeodomain proteins belong to the ancestral molecular toolkit of eukaryotes. Evol. Dev. 9, 212–219.CrossRefGoogle ScholarPubMed
Derynck, R. and Miyazono, K. (2008). The TGF- β Family. Cold Spring Harbor Laboratory Press (Monograph 50), Cold Spring Harbor, NY.
DeSalle, R. and Tattersall, I. (2008). Human Origins: What Bones and Genomes Tell Us About Ourselves. Texas A&M University Press, College Station, TX.Google Scholar
Deschamps, J. (2007). Ancestral and recently recruited global control of the Hox genes in development. Curr. Opin. Gen. Dev. 17, 422–427.CrossRefGoogle ScholarPubMed
Deschamps, J. and van Nes, J. (2005). Developmental regulation of the Hox genes during axial morphogenesis in the mouse. Development 132, 2931–2942.CrossRefGoogle ScholarPubMed
Deufel, A. and Cundall, D. (2006). Functional plasticity of the venom delivery system in snakes with a focus on the poststrike prey release behavior. Zool. Anz. 245, 249–267.CrossRefGoogle Scholar
Di-Poï, N., Montoya-Burgos, J.I., Miller, H., Pourquié, O., Milinkovitch, M.C., and Duboule, D. (2010). Changes in Hox genes’ structure and function during the evolution of the squamate body plan. Nature 464, 99–103.CrossRefGoogle ScholarPubMed
Diamond, J.M. (1996). Competition for brain space. Nature 382, 756–757.CrossRefGoogle ScholarPubMed
Diaz-Benjumea, F.J., Cohen, B., and Cohen, S.M. (1994). Cell interaction between compartments establishes the proximal-distal axis of Drosophila legs. Nature 372, 175–179.CrossRefGoogle ScholarPubMed
Dichtel-Danjoy, M.-L. and Félix, M.-A. (2004). Phenotypic neighborhood and micro-evolvability. Trends Genet. 20, 268–276.CrossRefGoogle ScholarPubMed
Dickerson, A.K., Shankles, P.G., Madhavan, N.M., and Hu, D.L. (2012). Mosquitoes survive raindrop collisions by virtue of their low mass. PNAS 109 #25, 9822–9827.CrossRefGoogle ScholarPubMed
Dickinson, M.H. (1999). Haltere-mediated equilibrium reflexes of the fruit fly, Drosophila melanogaster. Phil. Trans. R. Soc. Lond. B 354, 903–916.CrossRefGoogle ScholarPubMed
Dickinson, M.H. (2005). The initiation and control of rapid flight maneuvers in fruit flies. Integr. Comp. Biol. 45, 274–281.CrossRefGoogle ScholarPubMed
Dickinson, M.H. (2006). Insect flight. Curr. Biol. 16, R309–R314.CrossRefGoogle ScholarPubMed
Dickinson, M.H., Farley, C.T., Full, R.J., Koehl, M.A.R., Kram, R., and Lehman, S. (2000). How animals move: an integrative view. Science 288, 100–106.CrossRefGoogle Scholar
Dierick, H.A. (2008). Fly fighting: octopamine modulates aggression. Curr. Biol. 18, R161–R163.CrossRefGoogle ScholarPubMed
Dietrich, M.R. (2000). From hopeful monsters to homeotic effects: Richard Goldschmidt’s integration of development, evolution, and genetics. Am. Zool. 40, 738–747.Google Scholar
Dietrich, M.R. (2003). Richard Goldschmidt: hopeful monsters and other ‘heresies’. Nature Rev. Genet. 4, 68–74.CrossRefGoogle Scholar
Dimitropoulos, A. (1985). First records of Orsini’s viper, Vipera ursinii (Viperidae), in Greece. Ann. Mus. Goulandris 7, 319–323.Google Scholar
DiNardo, S. and O’Farrell, P.H. (1987). Establishment and refinement of segmental pattern in the Drosophila embryo: spatial control of engrailed expression by pair-rule genes. Genes Dev. 1, 1212–1225.CrossRefGoogle ScholarPubMed
Diogo, R., Linde-Medina, M., Abdala, V., and Ashley-Ross, M.A. (2013). New, puzzling insights from comparative myological studies on the old and unsolved forelimb/hindlimb enigma. Biol. Rev. 88, 196–214.CrossRefGoogle ScholarPubMed
Dittrich-Reed, D.R. and Fitzpatrick, B.M. (2013). Transgressive hybrids as hopeful monsters. Evol. Biol. 40, 310–315.CrossRefGoogle ScholarPubMed
Dixon, D. (1998). After Man: A Zoology of the Future. St. Martin’s Griffin, New York, NY.Google Scholar
Dodson, P. (1991). Life styles of the huge and famous. Nat. Hist. 100 #12, 30–35.Google Scholar
Doe, C.Q. (2006). Chinmo and neuroblast temporal identity. Cell 127, 254–256.CrossRefGoogle ScholarPubMed
Domning, D.P. (2001). The earliest known fully quadrupedal sirenian. Nature 413, 625–627.CrossRefGoogle ScholarPubMed
Don, E.K., Currie, P.D., and Cole, N.J. (2013). The evolutionary history of the development of the pelvic fin/hindlimb. J. Anat. 222, 114–133.CrossRefGoogle ScholarPubMed
Dong, J., Feldmann, G., Huang, J., Wu, S., Zhang, N., Comerford, S.A., Gayyed, M.F., Anders, R.A., Maitra, A., and Pan, D. (2007). Elucidation of a universal size-control mechanism in Drosophila and mammals. Cell 130, 1120–1133.CrossRefGoogle ScholarPubMed
Donlan, C.J. (2007). Restoring America’s big, wild animals. Sci. Am. 296 #6, 70–77.CrossRefGoogle ScholarPubMed
Donoghue, P.C.J., Graham, A., and Kelsh, R.N. (2008). The origin and evolution of the neural crest. BioEssays 30, 530–541.CrossRefGoogle ScholarPubMed
Doolittle, R.F. (1994). Convergent evolution: the need to be explicit. TiBS 19, 15–18.Google ScholarPubMed
Dornan, A.J. and Goodwin, S.F. (2008). Fly courtship song: triggering the light fantastic. Cell 133, 210–212.CrossRefGoogle ScholarPubMed
Douglas, R.H., Collett, T.S., and Wagner, H.-J. (1986). Accommodation in anuran Amphibia and its role in depth vision. J. Comp. Physiol. A 158, 133–143.CrossRefGoogle Scholar
Doupé, D.P. and Jones, P.H. (2013). Cycling progenitors maintain epithelia while diverse cell types contribute to repair. BioEssays 35, 443–451.CrossRefGoogle Scholar
Drake, A.G. (2011). Dispelling dog dogma: an investigation of heterochrony in dogs using 3D geometric morphometric analysis of skull shape. Evol. Dev.13, 204–213.CrossRefGoogle ScholarPubMed
Drake, A.G. and Klingenberg, C.P. (2010). Large-scale diversification of skull shape in domestic dogs: disparity and modularity. Am. Nat. 175, 289–301.CrossRefGoogle ScholarPubMed
Drimmer, F. (1973). Very Special People. Amjon, New York, NY.Google Scholar
Duboc, V. and Lepage, T. (2008). A conserved role for the nodal signaling pathway in the establishment of dorso-ventral and left-right axes in deuterostomes. J. Exp. Zool. (Mol. Dev. Evol.) 310B, 41–53.CrossRefGoogle Scholar
Duboc, V. and Logan, M.P.O. (2009). Building limb morphology through integration of signaling modules. Curr. Opin. Gen. Dev. 19, 497–503.CrossRefGoogle Scholar
Duboc, V. and Logan, M.P.O. (2011). Regulation of limb bud initiation and limb-type morphology. Dev. Dynamics 240, 1017–1027.CrossRefGoogle ScholarPubMed
Duboc, V., Röttinger, E., Lapraz, F., Besnardeau, L., and Lepage, T. (2005). Left-right asymmetry in the sea urchin embryo is regulated by Nodal signaling on the right side. Dev. Cell 9, 147–158.CrossRefGoogle ScholarPubMed
Duboule, D. (2007). The rise and fall of Hox gene clusters. Development 134, 2549–2560.CrossRefGoogle ScholarPubMed
Dubrulle, J. and Pourquié, O. (2004). Coupling segmentation to axis formation. Development 131, 5783–5793.CrossRefGoogle ScholarPubMed
Dudley, R. (2000). The Biomechanics of Insect Flight: Form, Function, Evolution. Princeton University Press, Princeton, NJ.Google Scholar
Dudley, R. and Yanoviak, S.P. (2011). Animal aloft: the origins of aerial behavior and flight. Integr. Comp. Biol. 51, 926–936.CrossRefGoogle Scholar
Dugon, M.M. and Arthur, W. (2012). Comparative studies on the structure and development of the venom-delivery system of centipedes, and a hypothesis on the origin of this evolutionary novelty. Evol. Dev. 14, 128–137.CrossRefGoogle Scholar
Dugon, M.M., Hayden, L., Black, A., and Arthur, W. (2012). Development of the venom ducts in the centipede Scolopendra: an example of recapitulation. Evol. Dev. 14, 515–521.CrossRefGoogle ScholarPubMed
Duncan, E.J., Leask, M.P., and Dearden, P.K. (2013). The pea aphid (Acyrthosiphon pisum) genome encodes two divergent early developmental programs. Dev. Biol. 377, 262–274.CrossRefGoogle ScholarPubMed
Dunn, R.R. (2006). Dig it! Nat. Hist. 115 #10, 36–41.Google Scholar
Duret, L. (2009). Mutation patterns in the human genome: more variable than expected. PLoS Biol. 7 #2, e1000028.CrossRefGoogle ScholarPubMed
Dutko, J.A. and Mullins, M.C. (2011). Snapshot: BMP signaling in development. Cell 145, 636.CrossRefGoogle ScholarPubMed
Dworkin, I. (2005). Canalization, cryptic variation, and developmental buffering: a critical examination and analytical perspective. In Variation: A Central Concept in Biology (Hallgrímsson, B. and Hall, B.K., eds.). Elsevier Academic Press, New York, NY, pp. 131–158.CrossRefGoogle Scholar
Eames, B.F. (2008). The genesis of cartilage size and shape during development and evolution. Development 135, 3947–3958.CrossRefGoogle ScholarPubMed
Eberhard, W.G. (1980). Horned beetles. Sci. Am. 242 #3, 166–181.CrossRefGoogle Scholar
Eberhard, W.G. (1987). Runaway sexual selection. Nat. Hist. 96 #12, 4–8.Google Scholar
Eberlein, S. and Russell, M.A. (1983). Effects of deficiencies in the engrailed region of Drosophila melanogaster. Dev. Biol. 100, 227–237.CrossRefGoogle ScholarPubMed
Eckalbar, W.L., Lasku, E., Infante, C.R., Elsey, R.M., Markov, G.J., Allen, A.N., Corneveaux, J.J., Losos, J.B., DeNardo, D.F., Huentelman, M.J., Wilson-Rawls, J., Rawls, A., and Kusumi, K. (2012). Somitogenesis in the anole lizard and alligator reveals evolutionary convergence and divergence in the amniote segmentation clock. Dev. Biol. 363, 308–319.CrossRefGoogle ScholarPubMed
Economides, K.D., Zeltser, L., and Capecchi, M.R. (2003). Hoxb13 mutations cause overgrowth of caudal spinal cord and tail vertebrae. Dev. Biol. 256, 317–330.CrossRefGoogle ScholarPubMed
Economou, A.D., Ohazama, A., Porntaveetus, T., Sharpe, P.T., Kondo, S., Basson, M.A., Gritli-Linde, A., Cobourne, M.T., and Green, J.B.A. (2012). Periodic stripe formation by a Turing mechanism operating at growth zones in the mammalian palate. Nature Genet. 44, 348–352.CrossRefGoogle ScholarPubMed
Eddison, M., Le Roux, I., and Lewis, J. (2000). Notch signaling in the development of the inner ear: lessons from Drosophila. PNAS 97 #22, 11692–11699.CrossRefGoogle ScholarPubMed
Ede, D.A. (1972). Cell behaviour and embryonic development. Internat. J. Neurosci. 3, 165–174.CrossRefGoogle ScholarPubMed
Edelman, G.M. (1987). Neural Darwinism: The Theory of Neuronal Group Selection. Basic Books, New York, NY.Google Scholar
Edelman, G.M. (1988). Topobiology: An Introduction to Molecular Embryology. Basic Books, New York, NY.Google Scholar
Edelman, G.M. (1993). Neural Darwinism: selection and reentrant signaling in higher brain function. Neuron 10, 115–125.CrossRefGoogle ScholarPubMed
Edelman, G.M. and Gallin, W.J. (1987). Cell adhesion as a basis of pattern in embryonic development. Am. Zool. 27, 645–656.CrossRefGoogle Scholar
Edgecombe, G.D. (2009). Palaeontological and molecular evidence linking arthropods, onychophorans, and other ecdysozoa. Evo. Edu. Outreach 2, 178–190.CrossRefGoogle Scholar
Edwards, J.S. (1997). The evolution of insect flight: implications for the evolution of the nervous system. Brain Behav. Evol. 50, 8–12.CrossRefGoogle ScholarPubMed
Egri, A., Blahó, M., Kriska, G., Farkas, R., Gyurkovszky, M., Åkesson, S., and Horváth, G. (2012). Polarotactic tabanids find striped patterns with brightness and/or polarization modulation least attractive: an advantage of zebra stripes. J. Exp. Biol. 215, 736–745.CrossRefGoogle ScholarPubMed
Eibner, C., Pittlik, S., Meyer, A., and Begemann, G. (2008). An organizer controls the development of the “sword,” a sexually selected trait in swordtail fish. Evol. Dev. 10, 403–412.CrossRefGoogle ScholarPubMed
Eilam, D. (1997). Postnatal development of body architecture and gait in several rodent species. J. Exp. Biol. 200, 1339–1350.Google ScholarPubMed
Eilam, D. and Shefer, G. (1997). The developmental order of bipedal locomotion in the jerboa (Jaculus orientalis): pivoting, creeping, quadrupedalism, and bipedalism. Dev. Psychobiol. 31, 137–142.3.0.CO;2-L>CrossRefGoogle ScholarPubMed
Eisner, T. (2003). For Love of Insects. Harvard University Press, Cambridge, MA.Google Scholar
Eizirik, E., David, V.A., Buckley-Beason, V., Roelke, M.E., Schäffer, A.A., Hannah, S.S., Narfström, K., O’Brien, S.J., and Menotti-Raymond, M. (2010). Defining and mapping mammalian coat pattern genes: multiple genomic regions implicated in domestic cat stripes and spots. Genetics 184, 267–275.CrossRefGoogle ScholarPubMed
El-Sherif, E., Averof, M., and Brown, S.J. (2012). A segmentation clock operating in blastoderm and germband stages of Tribolium development. Development 139, 4341–4346.CrossRefGoogle ScholarPubMed
Eldredge, N. and Eldredge, G. (2012). Editorial. Evo. Edu. Outreach 5, 179–180.CrossRefGoogle Scholar
Elgin, R.A., Hone, D.W.E., and Frey, E. (2011). The extent of the pterosaur flight membrane. Acta Palaeontol. Pol. 56, 99–111.CrossRefGoogle Scholar
Elsdale, T. and Wasoff, F. (1976). Fibroblast cultures and dermatoglyphics: the topology of two planar patterns. Wilhelm Roux’s Arch. 180, 121–147.CrossRefGoogle ScholarPubMed
Emerald, B.S. and Cohen, S.M. (2001). Limb development: getting down to the ground state. Curr. Biol. 11, R1025–R1027.CrossRefGoogle Scholar
Emerson, S.B. (1985). Jumping and leaping. In Functional Vertebrate Morphology (Hildebrand, M., Bramble, D.M., Liem, K.F., and Wake, D.B., eds.). Harvard University Press, Cambridge, MA, pp. 58–72.Google Scholar
Emlen, D. (2011). Diversity in the weapons of sexual selection: horn evolution in dung beetles. In In the Light of Evolution: Essays from the Laboratory and Field (Losos, J.B., ed.). Roberts, Greenwood Village, CO, pp. 149–170.Google Scholar
Emlen, D.J. (2000). Integrating development with evolution: a case study with beetle horns. BioScience 50, 403–418.CrossRefGoogle Scholar
Emlen, D.J. (2001). Costs and the diversification of exaggerated animal structures. Science 291, 1534–1536.CrossRefGoogle ScholarPubMed
Emlen, D.J. (2007). On the origin and evolutionary diversification of beetle horns. PNAS 104 (Suppl. 1), 8661–8668.CrossRefGoogle ScholarPubMed
Emlen, D.J. (2008). The evolution of animal weapons. Annu. Rev. Ecol. Evol. Syst. 39, 387–413.CrossRefGoogle Scholar
Emlen, D.J. and Allen, C.E. (2004). Genotype to phenotype: physiological control of trait size and scaling in insects. Integr. Comp. Biol. 43, 617–634.CrossRefGoogle Scholar
Emlen, D.J., Hunt, J., and Simmons, L.W. (2005). Evolution of sexual dimorphism and male dimorphism in the expression of beetle horns: phylogenetic evidence for modularity, evolutionary lability, and constraint. Am. Nat. 166 (Suppl.), S42–S68.CrossRefGoogle ScholarPubMed
Emlen, D.J. and Nijhout, H.F. (2000). The development and evolution of exaggerated morphologies in insects. Annu. Rev. Entomol. 45, 661–708.CrossRefGoogle ScholarPubMed
Emlen, D.J., Szafran, Q., Corley, L.S., and Dworkin, I. (2006). Insulin signaling and limb-patterning: candidate pathways for the origin and evolutionary diversification of beetle ‘horns’. Heredity 97, 179–191.CrossRefGoogle Scholar
Emlen, D.J., Warren, I.A., Johns, A., Dworkin, I., and Lavine, L.C. (2012). A mechanism of extreme growth and reliable signaling in sexually selected ornaments and weapons. Science 337, 860–864. [See also Warren, I.A., et al. (2013). A general mechanism for conditional expression of exaggerated sexually-selected traits. BioEssays 35, 889–899.]CrossRefGoogle ScholarPubMed
Emsley, M.G. (1975). Butterfly Magic. Viking Press, New York, NY.Google Scholar
Endo, H., Yamagiwa, D., Hayashi, Y., Koie, H., Yamaya, Y., and Kimura, J. (1999). Role of the giant panda’s “pseudo-thumb”. Nature 397, 309–310.CrossRefGoogle ScholarPubMed
Enquist, M. and Arak, A. (1993). Selection of exaggerated male traits by female esthetic senses. Nature 361, 446–448.CrossRefGoogle Scholar
Eom, D.S., Inoue, S., Patterson, L.B., Gordon, T.N., Slingwine, R., Kondo, S., Watanabe, M., and Parichy, D.M. (2012). Melanophore migration and survival during zebrafish adult pigment stripe development require the immunoglobulin superfamily adhesion molecule Igsf11. PLoS Genet. 8 #8, e1002899.CrossRefGoogle ScholarPubMed
Erclik, T., Hartenstein, V., Lipshitz, H.D., and McInnes, R.R. (2008). Conserved role of the Vsx genes supports a monophyletic origin for bilaterian visual systems. Curr. Biol. 18, 1278–1287.CrossRefGoogle ScholarPubMed
Erickson, G.M., Krick, B.A., Hamilton, M., Bourne, G.R., Norell, M.A., Lilleodden, E., and Sawyer, W.G. (2012). Complex dental structure and wear biomechanics in hadrosaurid dinosaurs. Science 338, 98–101.CrossRefGoogle ScholarPubMed
Erickson, G.M., Makovicky, P.J., Currie, P.J., Norell, M.A., Yerby, S.A., and Brochu, C.A. (2004). Gigantism and comparative life-history parameters of tyrannosaurid dinosaurs. Nature 430, 772–775.CrossRefGoogle ScholarPubMed
Erwin, D.H. (2006). Evolutionary contingency. Curr. Biol. 16, R825–R826.CrossRefGoogle ScholarPubMed
Erwin, D.H. (2007). Disparity: morphological pattern and developmental context. Palaeontology 50, 57–73.CrossRefGoogle Scholar
Erwin, D.H. (2009). Early origin of the bilaterian developmental toolkit. Phil. Trans. R. Soc. Lond. B 364, 2253–2261. [See also Wilkins, A.S. (2014). The genetic tool-kit: the life-history of an important metaphor. In Advances in Evolutionary Developmental Biology (J.T. Streelman, ed.). Wiley, New York, NY.]CrossRefGoogle ScholarPubMed
Erwin, D.H. and Davidson, E.H. (2002). The last common bilaterian ancestor. Development 129, 3021–3032.Google ScholarPubMed
Erwin, D.H., Laflamme, M., Tweedt, S.M., Sperling, E.A., Pisani, D., and Peterson, K.J. (2011). The Cambrian conundrum: early divergence and later ecological success in the early history of animals. Science 334, 1091–1097.CrossRefGoogle ScholarPubMed
Erwin, D.H. and Valentine, J.W. (2013). The Cambrian Explosion: The Construction of Animal Biodiversity. Roberts, Greenwood Village, CO.Google Scholar
Essalmani, R., Zaid, A., Marcinkiewicz, J., Chamberland, A., Pasquato, A., Seidah, N.G., and Prat, A. (2008). In vivo functions of the proprotein convertase PC5/6 during mouse development: Gdf11 is a likely substrate. PNAS 105 #15, 5750–5755.CrossRefGoogle ScholarPubMed
Estella, C. and Mann, R.S. (2008). Logic of Wg and Dpp induction of distal and medial fates in the Drosophila leg. Development 135, 627–636.CrossRefGoogle ScholarPubMed
Estella, C., McKay, D.J., and Mann, R.S. (2008). Molecular integration of Wingless, Decapentaplegic, and autoregulatory inputs into Distalless during Drosophila leg development. Dev. Cell 14, 86–96.CrossRefGoogle ScholarPubMed
Estella, C., Voutev, R., and Mann, R.S. (2012). A dynamic network of morphogens and transcription factors patterns the fly leg. Curr. Top. Dev. Biol. 98, 173–198.CrossRefGoogle ScholarPubMed
Estrada, B., Casares, F., and Sánchez-Herrero, E. (2003). Development of the genitalia in Drosophila melanogaster. Differentiation 71, 299–310.CrossRefGoogle ScholarPubMed
Estrada, B. and Sánchez-Herrero, E. (2001). The Hox gene Abdominal-B antagonizes appendage development in the genital disc of Drosophila. Development 128, 331–339.Google ScholarPubMed
Evans, C.J., Hartenstein, V., and Banerjee, U. (2003). Thicker than blood: conserved mechanisms in Drosophila and vertebrate hematopoiesis. Dev. Cell 5, 673–690.CrossRefGoogle ScholarPubMed
Evans, D.J.R., Valasek, P., Schmidt, C., and Patel, K. (2006). Skeletal muscle translocation in vertebrates. Anat. Embryol. 211 (Suppl. 1), S43–S50.CrossRefGoogle ScholarPubMed
Evans, J.W., Borton, A., HIntz, H.F., and Van Vleck, L.D. (1990). The Horse, 2nd edn. W. H. Freeman, New York, NY.Google Scholar
Evans, S.E. (2003). At the feet of the dinosaurs: the early history and radiation of lizards. Biol. Rev. 78, 513–551.CrossRefGoogle ScholarPubMed
Evans, S.M. (1999). Vertebrate tinman homologues and cardiac differentiation. Semin. Cell Dev. Biol. 10, 73–83.CrossRefGoogle ScholarPubMed
Ewart, J.C. (1894). The development of the skeleton of the limbs of the horse, with observations on polydactyly. J. Anat. Physiol. 28, 236–256 plus 21 figs.Google ScholarPubMed
Ewart, J.C. (1894). The second and fourth digits in the horse: their development and subsequent degeneration. Proc. Roy. Soc. Edinburgh 1894, 185–191.Google Scholar
Fain, G.L., Hardie, R., and Laughlin, S.B. (2010). Phototransduction and the evolution of photoreceptors. Curr. Biol. 20, R114–R124.CrossRefGoogle ScholarPubMed
Fan, J.-Y., Preuss, F., Muskus, M.J., Bjes, E.S., and Price, J.L. (2009). Drosophila and vertebrate casein kinase Iδ exhibits evolutionary conservation of circadian function. Genetics 181, 139–152.CrossRefGoogle ScholarPubMed
Farnum, C.E., Tinsley, M., and Hermanson, J.W. (2008). Forelimb versus hindlimb skeletal development in the big brown bat, Eptesicus fuscus: functional divergence is reflected in chondrocytic performance in autopodial growth plates. Cells Tissues Organs 187, 35–47.CrossRefGoogle ScholarPubMed
Faucheux, C., Nicholls, B.M., Allen, S., Danks, J.A., Horton, M.A., and Price, J.S. (2004). Recapitulation of the Parathyroid Hormone-related Peptide–Indian Hedgehog pathway in the regenerating deer antler. Dev. Dynamics 231, 88–97.CrossRefGoogle ScholarPubMed
Fausto-Sterling, A. and Smith-Schiess, H. (1982). Interactions between fused and engrailed, two mutations affecting pattern formation in Drosophila melanogaster. Genetics 101, 71–80.Google ScholarPubMed
Fayyazuddin, A. and Dickinson, M.H. (1996). Haltere afferents provide direct, electrotonic input to a steering motor neuron in the blowfly, Calliphora. J. Neurosci. 16, 5225–5232.CrossRefGoogle ScholarPubMed
Fedak, T.J. and Hall, B.K. (2004). Perspectives on hyperphalangy: patterns and processes. J. Anat. 204, 151–163.CrossRefGoogle ScholarPubMed
Fédrigo, O. and Wray, G.A. (2010). Developmental evolution: how beetles evolved their shields. Curr. Biol. 20, R64–R66.CrossRefGoogle ScholarPubMed
Feldhamer, G.A., Drickamer, L.C., Vessey, S.H., Merritt, J.F., and Krajewski, C. (2007). Mammalogy: Adaptation, Diversity, Ecology, 3rd edn. Johns Hopkins University Press, Baltimore, MD.Google Scholar
Félix, M.-A. (2012). Evolution in developmental phenotype space. Curr. Opin. Gen. Dev. 22, 593–599.CrossRefGoogle ScholarPubMed
Fenton, M.B. and Ratcliffe, J.M. (2010). Bats. Curr. Biol. 20, R1060–R1062.CrossRefGoogle ScholarPubMed
Ferkowicz, M.J. and Raff, R.A. (2001). Wnt gene expression in sea urchin development: heterochronies associated with the evolution of developmental mode. Evol. Dev. 3, 24–33.CrossRefGoogle ScholarPubMed
Fernandes, J., Bate, M., and VijayRaghavan, K. (1991). Development of the indirect flight muscles of Drosophila. Development 113, 67–77.Google ScholarPubMed
Feuda, R., Hamilton, S.C., McInerney, J.O., and Pisani, D. (2012). Metazoan opsin evolution reveals a simple route to animal vision. PNAS 109 #46, 18868–18872.CrossRefGoogle ScholarPubMed
Fichelson, P. and Gho, M. (2003). The glial cell undergoes apoptosis in the microchaete lineage of Drosophila. Development 130, 123–133.CrossRefGoogle ScholarPubMed
Filoramo, N.I. and Schwenk, K. (2009). The mechanism of chemical delivery to the vomeronasal organs in squamate reptiles: a comparative morphological approach. J. Exp. Zool. 311A, 20–34.CrossRefGoogle Scholar
Fincham, E.F. (1955). The proportion of ciliary muscular force required for accommodation. J. Physiol. 128, 99–112.CrossRefGoogle ScholarPubMed
Findlater, G.S., McDougall, R.D., and Kaufman, M.H. (1993). Eyelid development, fusion and subsequent reopening in the mouse. J. Anat. 183, 121–129.Google ScholarPubMed
Finnerty, J.R. (2003). The origins of axial patterning in the metazoa: how old is bilateral symmetry? Int. J. Dev. Biol. 47, 523–529.Google ScholarPubMed
Finnerty, J.R., Pang, K., Burton, P., Paulson, D., and Martindale, M.Q. (2004). Origins of bilateral symmetry: Hox and Dpp expression in a sea anemone. Science 304, 1335–1337.CrossRefGoogle Scholar
Fitch, D.H.A. and Sudhaus, W. (2002). One small step for worms, one giant leap for “Bauplan”?Evol. Dev. 4, 243–246.CrossRefGoogle ScholarPubMed
Fitch, W.M. (2000). Homology: a personal view on some of the problems. Trends Genet. 16, 227–231.CrossRefGoogle ScholarPubMed
Fitch, W.T. (2012). Evolutionary developmental biology and human language evolution: constraints on adaptation. Evol. Biol. 39, 613–637.CrossRefGoogle ScholarPubMed
Fjällbrant, T.T., Manger, P.R., and Pettigrew, J.D. (1998). Some related aspects of platypus electroreception: temporal integration behaviour, electroreceptive thresholds and directionality of the bill acting as an antenna. Phil. Trans. R. Soc. Lond. B 353, 1211–1219.CrossRefGoogle ScholarPubMed
Flatt, T. (2005). The evolutionary genetics of canalization. Q. Rev. Biol. 80, 287–316.CrossRefGoogle ScholarPubMed
Flinn, M.V., Geary, D.C., and Ward, C.V. (2005). Ecological dominance, social competition, and coalitionary arms races: why humans evolved extraordinary intelligence. Evol. Human Behav. 26, 10–46.CrossRefGoogle Scholar
Flynn, J.J. (2009). Splendid isolation. Nat. Hist. 118 #5, 26–32.Google Scholar
Fomenou, M.D., Scaal, M., Stockdale, F.E., Christ, B., and Huang, R. (2005). Cells of all somitic compartments are determined with respect to segmental identity. Dev. Dynamics 233, 1386–1393.CrossRefGoogle ScholarPubMed
Fondon, J.W., III and Garner, H.R. (2004). Molecular origins of rapid and continuous morphological evolution. PNAS 101 #52, 18058–18063.CrossRefGoogle ScholarPubMed
Ford, J. and Ford, D. (1986). Narwhal: unicorn of the Arctic seas. Natl. Geogr. 169 #3, 354–363.Google Scholar
Fordyce, R.E. and Ksepka, D.T. (2012). The strangest bird. Sci. Am. 307 #5, 56–61.CrossRefGoogle ScholarPubMed
Forgacs, G. and Newman, S.A. (2005). Biological Physics of the Developing Embryo. Cambridge University Press, New York, NY.CrossRefGoogle Scholar
Foronda, D., Martin, P., and Sánchez-Herrero, E. (2012). Drosophila Hox and sex-determination genes control segment elimination through EGFR and extramacrochaete activity. PLoS Genet. 8 #8, e1002874.CrossRefGoogle Scholar
Forsman, A. (1995). Opposing fitness consequences of colour pattern in male and female snakes. J. Evol. Biol. 8, 53–70.CrossRefGoogle Scholar
Fortey, R. (2012). Horseshoe Crabs and Velvet Worms: The Story of the Animals and Plants That Time Has Left Behind. Knopf, New York, NY.Google Scholar
Fortey, R. and Thomas, R.H. (1993). The case of the velvet worm. Nature 361, 205–206.CrossRefGoogle Scholar
Foster, S.J. and Vincent, A.C.J. (2004). Life history and ecology of seahorses: implications for conservation and management. J. Fish Biol. 65, 1–61.CrossRefGoogle Scholar
Foureaux, G., Egami, M.I., Jared, C., Antoniazzi, M.M., Gutierre, R.C., and Smith, R.L. (2010). Rudimentary eyes of squamate fossorial reptiles (Amphisbaenia and Serpentes). Anat. Rec. Adv. Integr. Anat. Evol. Biol. 293, 351–357.CrossRefGoogle Scholar
Fox, C.S., Liu, Y., White, C.C., Feitosa, M., Smith, A.V., Heard-Costa, N., Lohman, K., GIANT Consortium, MAGIC Consortium, GLGC Consortium, Johnson, A.D., Foster, M.C., Greenawalt, D.M., Griffin, P., Ding, J., Newman, A.B., Tylavsky, F., Miljkovic, I., Kritchevsky, S.B., Launer, L., Garcia, M., Eiriksdottir, G., Carr, J.J., Gudnason, V., Harris, T.B., Cupples, L.A., and Borecki, I.B. (2012). Genome-wide association for abdominal subcutaneous and visceral adipose reveals a novel locus for visceral fat in women. PLoS Genet. 8 #5, e1002695.CrossRefGoogle ScholarPubMed
Fox, C.S., White, C.C., Lohman, K., Heard-Costa, N., Cohen, P., Zhang, Y., Johnson, A.D., Emilsson, V., Liu, C.-T., Chen, Y.-D.I., Taylor, K.D., Allison, M., Budoff, M., CARDIoGRAM Consortium, Rotter, J.I., Carr, J.J., Hoffmann, U., Ding, J., Cupples, L.A., and Liu, Y. (2012). Genome-wide association of pericardial fat identifies a unique locus for ectopic fat. PLoS Genet. 8 #5, e1002705.CrossRefGoogle ScholarPubMed
Fox, C.W., Scheibly, K.L., and Reed, D.H. (2008). Experimental evolution of the genetic load and its implications for the genetic basis of inbreeding depression. Evolution 62, 2236–2249.CrossRefGoogle ScholarPubMed
Fox, D.T. and Duronio, R.J. (2012). Endoreplication and polyploidy: insights into development and disease. Development 140, 3–12.CrossRefGoogle Scholar
Fox, R.C. and Scott, C.S. (2005). First evidence of a venom delivery apparatus in extinct mammals. Nature 435, 1091–1093.CrossRefGoogle ScholarPubMed
Fraenkel, G. and Pringle, J.W.S. (1938). Halteres of flies as gyroscopic organs of equilibrium. Nature 141, 919–920.CrossRefGoogle Scholar
Francino, M.P. (2005). An adaptive radiation model for the origin of new gene functions. Nat. Genet. 37, 573–577.CrossRefGoogle ScholarPubMed
Francis, V.A., Zorzano, A., and Teleman, A.A. (2010). dDOR is an EcR coactivator that forms a feed-forward loop connecting insulin and ecdysone signaling. Curr. Biol. 20, 1799–1808.CrossRefGoogle ScholarPubMed
Francis-West, P.H., Robertson, K.E., Ede, D.A., C. Rodriguez, C., Izpisúa-Belmonte, J.C., Houston, B., Burt, D.W., Gribbin, C., Brickell, P.M., and Tickle, C. (1995). Expression of genes encoding bone morphogenetic proteins and Sonic hedgehog in talpid (ta3) limb buds: their relationships in the signalling cascade involved in limb patterning. Dev. Dynamics 203, 187–197.CrossRefGoogle ScholarPubMed
Frankino, W.A., Zwaan, B.J., Stern, D.L., and Brakefield, P.M. (2005). Natural selection and developmental constraints in the evolution of allometries. Science 307, 718–720.CrossRefGoogle ScholarPubMed
Frantsevich, L. (2012). Indirect closing of elytra by the prothorax in beetles (Coleoptera): general observations and exceptions. Zoology 115, 12–21.CrossRefGoogle ScholarPubMed
Fraser, F.C. (1938). Vestigial teeth in the narwhal. Proc. Linnean Soc. London 150, 155–162.CrossRefGoogle Scholar
Fraser, G.J., Bloomquist, R.F., and Streelman, J.T. (2013). Common developmental pathways link tooth shape to regeneration. Dev. Biol. 377, 399–414.CrossRefGoogle ScholarPubMed
Fraser, G.J., Hulsey, C.D., Bloomquist, R.F., Uyesugi, K., Manley, N.R., and Streelman, J.T. (2009). An ancient gene network is co-opted for teeth on old and new jaws. PLoS Biol. 7 #2, 233–247 (e1000031).CrossRefGoogle ScholarPubMed
Fraser, M.J., Jr. (2012). Insect transgenesis: current applications and future prospects. Annu. Rev. Entomol. 57, 267–289.CrossRefGoogle ScholarPubMed
Frazzetta, T.H. (1970). From hopeful monsters to bolyerine snakes? Am. Nat. 104, 55–72.CrossRefGoogle Scholar
Frazzetta, T.H. (2012). Flatfishes, turtles, and bolyerine snakes: evolution by small steps or large, or both? Evol. Biol. 39, 30–60.CrossRefGoogle Scholar
Freitas, R., Gómez-Marín, C., Wilson, J.M., Casares, F., and Gómez-Skarmeta, J.L. (2012). Hoxd13 contribution to the evolution of vertebrate appendages. Dev. Cell 23, 1219–1229.CrossRefGoogle ScholarPubMed
Freitas, R., Zhang, G., and Cohn, M.J. (2007). Biphasic Hoxd gene expression in shark paired fins reveals an ancient origin of the distal limb domain. PLoS ONE 2 #8, e754.CrossRefGoogle ScholarPubMed
French, V. (1997). Pattern formation in colour on butterfly wings. Curr. Opin. Gen. Dev. 7, 524–529.CrossRefGoogle ScholarPubMed
French, V. and Brakefield, P.M. (1992). The development of eyespot patterns on butterfly wings: morphogen sources or sinks? Development 116, 103–109.Google Scholar
French, V. and Brakefield, P.M. (1995). Eyespot determination on butterfly wings: the focal signal. Dev. Biol. 168, 112–123.CrossRefGoogle Scholar
French, V. and Brakefield, P.M. (2004). Pattern formation: a focus on Notch in butterfly eyespots. Curr. Biol. 14, R663–R665.CrossRefGoogle ScholarPubMed
French, V., Bryant, P.J., and Bryant, S.V. (1976). Pattern regulation in epimorphic fields. Science 193, 969–981.CrossRefGoogle ScholarPubMed
French, V. and Monteiro, A. (1994). Butterfly wings: colour patterns and now gene expression patterns. BioEssays 16, 789–791.CrossRefGoogle Scholar
Friberg, U. and Rice, W.R. (2008). Cut thy neighbor: cyclic birth and death of recombination hotspots via genetic conflict. Genetics 179, 2229–2238.CrossRefGoogle ScholarPubMed
Friedman, M. (2008). The evolutionary origin of flatfish asymmetry. Nature 454, 209–212.CrossRefGoogle ScholarPubMed
Fritz, A.E., Ikmi, A., Seidel, C., Paulson, A., and Gibson, M.C. (2013). Mechanisms of tentacle morphogenesis in the sea anemone Nematostella vectensis. Development 140, 2212–2223.CrossRefGoogle ScholarPubMed
Fritz, J., Hummel, J., Kienzle, E., Wings, O., Streich, W.J., and Clauss, M. (2011). Gizzard vs. teeth, it’s a tie: food-processing efficiency in herbivorous birds and mammals and implications for dinosaur feeding strategies. Paleobiology 37, 577–586.CrossRefGoogle Scholar
Fröbisch, J. (2011). On dental occlusion and saber teeth. Science 331, 1525–1528.CrossRefGoogle ScholarPubMed
Fry, B.G. (2005). From genome to “venome”: Molecular origin and evolution of the snake venom proteome inferred from phylogenetic analysis of toxin sequences and related body proteins. Genome Res. 15, 403–420.CrossRefGoogle ScholarPubMed
Fry, B.G., Vidal, N., Norman, J.A., Vonk, F.J., Scheib, H., Ramjan, S.F.R., Kuruppu, S., Fung, K., Hedges, S.B., Richardson, M.K., Hodgson, W.C., Ignjatovic, V., Summerhayes, R., and Kochva, E. (2006). Early evolution of the venom system in lizards and snakes. Nature 439, 584–588.CrossRefGoogle ScholarPubMed
Fry, C.L. (2006). Juvenile hormone mediates a trade-off between primary and secondary sexual traits in stalk-eyed flies. Evol. Dev. 8, 191–201.CrossRefGoogle ScholarPubMed
Fuchs, E. (2009). The tortoise and the hair: slow-cycling cells in the stem cell race. Cell 137, 811–819.CrossRefGoogle ScholarPubMed
Fuchs, Y. and Steller, H. (2011). Programmed cell death in animal development and disease. Cell 147, 742–758.CrossRefGoogle ScholarPubMed
Fusco, G. (2005). Trunk segment numbers and sequential segmentation in myriapods. Evol. Dev. 7, 608–617.CrossRefGoogle ScholarPubMed
Gad, J.M. and Tam, P.P.L. (1999). Axis development: the mouse becomes a dachshund. Curr. Biol. 9, R783–R786.CrossRefGoogle ScholarPubMed
Galant, R. and Carroll, S.B. (2002). Evolution of a transcriptional repression domain in an insect Hox protein. Nature 415, 910–913.CrossRefGoogle Scholar
Galant, R., Skeath, J.B., Paddock, S., Lewis, D.L., and Carroll, S.B. (1998). Expression pattern of a butterfly achaete-scute homolog reveals the homology of butterfly wing scales and insect sensory bristles. Curr. Biol. 8, 807–813.CrossRefGoogle ScholarPubMed
Galant, R., Walsh, C.M., and Carroll, S.B. (2002). Hox repression of a target gene: extradenticle-independent, additive action through multiple monomer binding sites. Development 129, 3115–3126.Google ScholarPubMed
Galindo, M.I., Bishop, S.A., and Couso, J.P. (2005). Dynamic EGFR-Ras signalling in Drosophila leg development. Dev. Dynamics 233, 1496–1508.CrossRefGoogle ScholarPubMed
Galindo, M.I., Bishop, S.A., Greig, S., and Couso, J.P. (2002). Leg patterning driven by proximal-distal interactions and EGFR signaling. Science 297, 256–259.CrossRefGoogle ScholarPubMed
Galindo, M.I., Fernández-Garza, D., Phillips, R., and Couso, J.P. (2011). Control of Distal-less expression in the Drosophila appendages by functional 3ʹ enhancers. Dev. Biol. 353, 396–410.CrossRefGoogle Scholar
Galindo, M.I., Pueyo, J.I., Fouix, S., Bishop, S.A., and Couso, J.P. (2007). Peptides encoded by short ORFs control development and define a new eukaryotic gene family. PLoS Biol. 5 #5, e106. [See also Lleras-Forero, L. (2013). Neuropeptides: developmental signals in placode progenitor formation. Dev. Cell 26, 195–203.]CrossRefGoogle ScholarPubMed
Galis, F. (1999). Why do almost all mammals have seven cervical vertebrae? Developmental constraints, Hox genes, and cancer. J. Exp. Zool. (Mol. Dev. Evol.) 285, 19–26.3.0.CO;2-Z>CrossRefGoogle ScholarPubMed
Galis, F. and Metz, J.A.J. (2003). Anti-cancer selection as a source of developmental and evolutionary constraints. BioEssays 25, 1035–1039.CrossRefGoogle ScholarPubMed
Galis, F. and Metz, J.A.J. (2007). Evolutionary novelties: the making and breaking of pleiotropic constraints. Integr. Comp. Biol. 47, 409–419.CrossRefGoogle ScholarPubMed
Galis, F., van Alphen, J.J.M., and Metz, J.A.J. (2001). Why five fingers? Evolutionary constraints on digit numbers. Trends Ecol. Evol. 16, 637–646.CrossRefGoogle Scholar
Galis, F., Van Dooren, T.J.M., Feuth, J.D., Metz, J.A.J., Witkam, A., Ruinard, S., Steigenga, M.J., and Wijnaendts, L.C.D. (2006). Extreme selection in humans against homeotic transformations of cervical vertebrae. Evolution 60, 2643–2654.CrossRefGoogle ScholarPubMed
Galis, F., Wagner, G.P., and Jockusch, E.L. (2003). Why is limb regeneration possible in amphibians but not in reptiles, birds, and mammals? Evol. Dev. 5, 208–220.CrossRefGoogle Scholar
Galton, F. (1869). Hereditary Genius: An Inquiry into Its Laws and Consequences. Macmillan, London.CrossRefGoogle Scholar
Galton, F. (1889). Natural Inheritance. Macmillan, London.CrossRefGoogle Scholar
Gamble, T. and Zarkower, D. (2012). Sex determination. Curr. Biol. 22, 257–262.CrossRefGoogle ScholarPubMed
Gañan, Y., Macias, D., Basco, R.D., Merino, R., and Hurle, J.M. (1998). Morphological diversity of the avian foot is related with the pattern of msx gene expression in the developing autopod. Dev. Biol. 196, 33–41.CrossRefGoogle ScholarPubMed
Ganfornina, M.D. and Sánchez, D. (1999). Generation of evolutionary novelty by functional shift. BioEssays 21, 432–439.3.0.CO;2-T>CrossRefGoogle ScholarPubMed
Gans, C. (1975). Tetrapod limblessness: Evolution and functional corollaries. Am. Zool. 15, 455–467.CrossRefGoogle Scholar
Gans, C. (1984). Slide-pushing: A transitional locomotor method of elongate squamates. Symp. Zool. Soc. Lond. 52, 13–26.Google Scholar
Garcia-Bellido, A. (1998). The engrailed story. Genetics 148, 539–544.Google ScholarPubMed
García-Bellido, A. (1977). Homoeotic and atavic mutations in insects. Am. Zool. 17, 613–629.CrossRefGoogle Scholar
García-Bellido, A. and de Celis, J.F. (2009). The complex tale of the achaete-scute complex: a paradigmatic case in the analysis of gene organization and function during development. Genetics 182, 631–639.CrossRefGoogle ScholarPubMed
Garcia-Bellido, A. and Santamaria, P. (1972). Developmental analysis of the wing disc in the mutant engrailed of Drosophila melanogaster. Genetics 72, 87–104.Google ScholarPubMed
Garcia-Fernàndez, J. (2005). The genesis and evolution of homeobox gene clusters. Nature Rev. Genet. 6, 881–892.CrossRefGoogle ScholarPubMed
Garcia-Fernàndez, J. and Benito-Gutiérrez, É. (2009). It’s a long way from amphioxus: descendants of the earliest chordate. BioEssays 31, 665–675.CrossRefGoogle ScholarPubMed
Gardner, M. (1970). The fantastic combinations of John Conway’s new solitare game “life”. Sci. Am. 223 #4, 120–123.CrossRefGoogle Scholar
Gardner, M. (1971). On cellular automata, self-reproduction, the Garden of Eden and the game “life”. Sci. Am. 224 #2, 112–117.Google Scholar
Garrouste, R., Clément, G., Nel, P., Engel, M.S., Grandcolas, P., D’Haese, C., Lagebro, L., Denayer, J., Gueriau, P., Lafaite, P., Olive, S., Prestianni, C., and Nel, A. (2012). A complete insect from the Late Devonian period. Nature 488, 82–85.CrossRefGoogle ScholarPubMed
Garvie, C.W. and Wolberger, C. (2001). Recognition of specific DNA sequences. Mol. Cell 8, 937–946.CrossRefGoogle ScholarPubMed
Garza-Garcia, A.A., Driscoll, P.C., and Brockes, J.P. (2010). Evidence for the local evolution of mechanisms underlying limb regeneration in salamanders. Integr. Comp. Biol. 50, 528–535.CrossRefGoogle ScholarPubMed
Gatesy, S.M. and Dial, K.P. (1996). From frond to fan: Archaeopteryx and the evolution of short-tailed birds. Evolution 50, 2037–2048.CrossRefGoogle ScholarPubMed
Gaunt, S.J. (1994). Conservation in the Hox code during morphological evolution. Int. J. Dev. Biol. 38, 549–552.Google ScholarPubMed
Gaviño, M.A. and Reddien, P.W. (2011). A Bmp/Admp regulatory circuit controls maintenance and regeneration of dorsal–ventral polarity in planarians. Curr. Biol. 21, 294–299.CrossRefGoogle ScholarPubMed
Gazave, E., Lapébie, P., Richards, G.S., Brunet, F., Ereskovsky, A.V., Degnan, B.M., Borchiellini, C., Vervoort, M., and Renard, E. (2009). Origin and evolution of the Notch signalling pathway: an overview from eukaryotic genomes. BMC Evol. Biol. 9, Article 249 (27 pp.).CrossRefGoogle ScholarPubMed
Gebelein, B., Culi, J., Ryoo, H.D., Zhang, W., and Mann, R.S. (2002). Specificity of Distalless repression and limb primordia development by abdominal Hox proteins. Dev. Cell 3, 487–498.CrossRefGoogle ScholarPubMed
Gebelein, B., McKay, D.J., and Mann, R.S. (2004). Direct integration of Hox and segmentation gene inputs during Drosophila development. Nature 431, 653–659.CrossRefGoogle ScholarPubMed
Gebo, D.L. (1987). Functional anatomy of the tarsier foot. Am. J. Phys. Anthrop. 73, 9–31.CrossRefGoogle Scholar
Gee, H. (2008). The amphioxus unleashed. Nature 453, 999–1000.CrossRefGoogle ScholarPubMed
Geertsema, A. (1991). The servals of Gorigor. Nat. Hist. 100 #2, 52–61.Google Scholar
Gehring, W. (2012). The animal body plan, the prototypic body segment, and eye evolution. Evol. Dev. 14, 34–46.CrossRefGoogle ScholarPubMed
Gehring, W.J. (1998). Master Control Genes in Development and Evolution: The Homeobox Story. Yale University Press, New Haven, CT.Google Scholar
Gehring, W.J. (2002). The genetic control of eye development and its implications for the evolution of the various eye-types. Int. J. Dev. Biol. 46, 65–73.Google ScholarPubMed
Geist, V. (1986). The paradox of the great Irish stags. Nat. Hist. 95 #3, 54–65.Google Scholar
Geist, V. (1994). Why antlers branched out. Nat. Hist. 103 #4, 78–83.Google Scholar
Gerber, S. (2013). On the relationship between the macroevolutionary trajectories of morphological integration and morphological disparity. PLoS ONE 8 #5, e63913.CrossRefGoogle ScholarPubMed
Gerhart, J. (1999). Signaling pathways in development (1998 Warkany lecture). Teratology 60, 226–239.3.0.CO;2-W>CrossRefGoogle Scholar
Gerhart, J. and Kirschner, M. (1997). Cells, Embryos, and Evolution. Blackwell Science, Malden, MA.Google Scholar
Gerhart, J. and Kirschner, M. (2007). The theory of facilitated variation. PNAS 104 (Suppl. 1), 8582–8589.CrossRefGoogle ScholarPubMed
Gerhart, J., Lowe, C., and Kirschner, M. (2005). Hemichordates and the origin of chordates. Curr. Opin. Gen. Dev. 15, 461–467.CrossRefGoogle ScholarPubMed
Ghazi, A., Anant, S., and VijayRaghavan, K. (2000). Apterous mediates development of direct flight muscles autonomously and indirect flight muscles through epidermal cues. Development 127, 5309–5318.Google ScholarPubMed
Ghazi, A., Paul, L., and VijayRaghavan, K. (2003). Prepattern genes and signaling molecules regulate stripe expression to specify Drosophila flight muscle attachment sites. Mechs. Dev. 120, 519–528.CrossRefGoogle ScholarPubMed
Gherman, A., Chen, P.E., Teslovich, T.M., Stankiewicz, P., Withers, M., Kashuk, C.S., Chakravarti, A., Lupski, J.R., Cutler, D.J., and Katsanis, N. (2007). Population bottlenecks as a potential major shaping force of human genome architecture. PLoS Genet. 3 #7, 1223–1231 (e119).CrossRefGoogle ScholarPubMed
Ghiradella, H. (2010). Insect cuticular surface modifications: scales and other structural formations. Adv. Insect Physiol. 38, 135–180.CrossRefGoogle Scholar
Ghiradella, H. and Schmidt, J.T. (2004). Fireflies at one hundred plus: a new look at flash control. Integr. Comp. Biol. 44, 203–212.CrossRefGoogle Scholar
Ghiselin, M.T. (1995). A movable feaster. Nat. Hist. 94 #9, 54–61.Google Scholar
Ghysen, A. (2003). The origin and evolution of the nervous system. Int. J. Dev. Biol. 47, 555–562.Google ScholarPubMed
Gibson, G. (1999). Developmental evolution: Going beyond the “just so”. Curr. Biol. 9, R942–R945.CrossRefGoogle Scholar
Gibson, G. (2000). Evolution: Hox genes and the cellared wine principle. Curr. Biol. 10, R452–R455.CrossRefGoogle ScholarPubMed
Gibson, G. and Honeycutt, E. (2002). The evolution of developmental regulatory pathways. Curr. Opin. Gen. Dev. 12, 695–700.CrossRefGoogle ScholarPubMed
Gierer, A. and Meinhardt, H. (1974). Biological pattern formation involving lateral inhibition. In Lectures on Mathematics in the Life Sciences, Vol. 7. American Mathematical Society, Providence, RI, pp. 163–183.Google Scholar
Gilbert, A.N. (1986). Mammary number and litter size in Rodentia: the “one-half rule”. PNAS 83 #13, 4828–4830.CrossRefGoogle ScholarPubMed
Gilbert, S.F. (2003). Opening Darwin’s black box: teaching evolution through developmental genetics. Nature Rev. Genet. 4, 735–741.CrossRefGoogle ScholarPubMed
Gilbert, S.F. (2010). Developmental Biology, 9th edn. Sinauer, Sunderland, MA.Google Scholar
Gilbert, S.F. and Bolker, J.A. (2001). Homologies of process and modular elements of embryonic construction. J. Exp. Zool. (Mol. Dev. Evol.) 291, 1–12.CrossRefGoogle ScholarPubMed
Gilbert, S.F. and Epel, D. (2008). Ecological Developmental Biology. Sinauer, Sunderland, MA.Google Scholar
Gill, F.B. (2007). Ornithology, 3rd edn. W. H. Freeman, New York, NY.Google Scholar
Gillespie, R.G. (2013). Adaptive radiation: convergence and non-equilibrium. Curr. Biol. 23, R71–R74. [See also Irschick, D.J., et al. (2013). Evo-devo beyond morphology: from genes to resource use. Trends Ecol. Evol. 28, 267–273.]CrossRefGoogle ScholarPubMed
Gillham, N.W. (2001). Evolution by jumps: Francis Galton and William Bateson and the mechanism of evolutionary change. Genetics 159, 1383–1392.Google Scholar
Gingerich, P.D., ul Haq, M., Zalmout, I.S., Khan, I.H., and Malkani, M.S. (2001). Origin of whales from early artiodactyls: hands and feet of Eocene protocetidae from Pakistan. Science 293, 2239–2242.CrossRefGoogle ScholarPubMed
Giorgianni, M. and Patel, N.H. (2005). Conquering land, air and water: the evolution and development of arthropod appendages. In Evolving Form and Function: Fossils and Development (Briggs, D.E.G., ed.). Yale Peabody Museum of Natural History, New Haven, CT, pp. 159–180.Google Scholar
Giorgianni, M.W. and Mann, R.S. (2011). Establishment of medial fates along the proximodistal axis of the Drosophila leg through direct activation of dachshund by Distalless. Dev. Cell 20, 455–468.CrossRefGoogle ScholarPubMed
Gleichauf, R. (1936). Anatomie und Variabilität des Geschlechtsapparates von Drosophila melanogaster (Meigen). Z. wiss. Zool. 148, 1–66.Google Scholar
Glimm, T., Zhang, J., Shen, Y.-Q., and Newman, S.A. (2012). Reaction-diffusion systems and external morphogen gradients: the two-dimensional case, with an application to skeletal pattern formation. Bull. Math. Biol. 74, 666–687.CrossRefGoogle ScholarPubMed
Gmitro, J.I. and Scriven, L.E. (1969). A physicochemical basis for pattern and rhythm. In Towards a Theoretical Biology. II. Sketches. (Waddington, C.H., ed.). Edinburgh University Press, Edinburgh, pp. 184–203.Google Scholar
Gnatzy, W., Grünert, U., and Bender, M. (1987). Campaniform sensilla of Calliphora vicina (Insecta, Diptera). I. Typography. Zoomorphology 106, 312–319.CrossRefGoogle Scholar
Goberdhan, D.C.I. and Wilson, C. (2002). Insulin receptor-mediated organ overgrowth in Drosophila is not restricted by body size. Dev. Genes Evol. 212, 196–202.CrossRefGoogle Scholar
Goel, N.S. and Thompson, R.L. (1989). Movable finite automata (MFA): a new tool for computer modeling of living systems. In Artificial Life (Langton, C.G., ed.). Addison-Wesley, New York, NY, pp. 317–340.Google Scholar
Goldbeter, A., Gonze, D., and Pourquié, O. (2007). Sharp developmental thresholds defined through bistability by antagonistic gradients of retinoic acid and FGF signaling. Dev. Dynamics 236, 1495–1508.CrossRefGoogle Scholar
Goldschmidt, R. (1938). Physiological Genetics. McGraw-Hill, New York, NY.Google Scholar
Goldschmidt, R. (1940). The Material Basis of Evolution. Yale University Press, New Haven, CT.Google Scholar
Goldschmidt, R.B. (1949). Phenocopies. Sci. Am. 181 #10, 46–49.CrossRefGoogle ScholarPubMed
Goldschmidt, R.B. (1952). Homoeotic mutants and evolution. Acta Biotheor. 10, 87–104.CrossRefGoogle Scholar
Goldsmith, T.H. (1990). Optimization, constraint, and history in the evolution of eyes. Q. Rev. Biol. 65, 281–322.CrossRefGoogle Scholar
Gomez, C., Özbudak, E.M., Wunderlich, J., Baumann, D., Lewis, J., and Pourquié, O. (2008). Control of segment number in vertebrate embryos. Nature 454, 335–339.CrossRefGoogle ScholarPubMed
Gomez, C. and Pourquié, O. (2009). Developmental control of segment numbers in vertebrates. J. Exp. Zool. (Mol. Dev. Evol.) 312B, 533–544.CrossRefGoogle Scholar
Gómez-Skarmeta, J.L., Rodríguez, I., Martínez, C., Culí, J., Ferrés-Marcó, D., Beamonte, D., and Modolell, J. (1995). Cis-regulation of achaete and scute: shared enhancer-like elements drive their coexpression in proneural clusters of the imaginal discs. Genes Dev. 9, 1869–1882.CrossRefGoogle ScholarPubMed
Gompel, N. and Prud’homme, B. (2009). The causes of repeated genetic evolution. Dev. Biol. 332, 36–47.CrossRefGoogle ScholarPubMed
Goodden, R. (1977). The Wonderful World of Butterflies and Moths. Hamlyn, New York, NY.Google Scholar
Goode, D.K., Callaway, H.A., Cerda, G.A., Lewis, K.E., and Elgar, G. (2011). Minor change, major difference: divergent functions of highly conserved cis-regulatory elements subsequent to whole genome duplication events. Development 138, 879–884.CrossRefGoogle ScholarPubMed
Goodfield, J. (1974). Changing strategies: a comparison of reductionist attitudes in biological and medical research in the nineteenth and twentieth centuries. In Studies in the Philosophy of Biology. Reduction and Related Problems (Ayala, F.J. and Dobzhansky, T., eds.). University of California Press, Berkeley, CA, pp. 65–86.CrossRefGoogle Scholar
Goodrich, L.V. and Strutt, D. (2011). Principles of planar polarity in animal development. Development 138, 1877–1892.CrossRefGoogle ScholarPubMed
Goodwin, B. (1994). How the Leopard Changed Its Spots. Charles Scribner’s Sons, New York, NY.Google Scholar
Goodwin, B.C. (1985). Developing organisms as self-organizing fields. In Mathematical Essays on Growth and the Emergence of Form (Antonelli, P.L., ed.). University of Alberta Press, Edmonton, pp. 185–200.Google Scholar
Goodwin, B.C. and Cohen, M.H. (1969). A phase-shift model for the spatial and temporal organization of developing systems. J. Theor. Biol. 25, 49–107.CrossRefGoogle ScholarPubMed
Gordo, I. and Campos, P.R.A. (2008). Sex and deleterious mutations. Genetics 179, 621–626.CrossRefGoogle ScholarPubMed
Gordo, I., Navarro, A., and Charlesworth, B. (2002). Muller’s ratchet and the pattern of variation at a neutral locus. Genetics 161, 835–848.Google Scholar
Gordon, R. (1999). The Hierarchical Genome and Differentiation Waves: Novel Unification of Development, Genetics and Evolution. World Scientific, Singapore.CrossRefGoogle Scholar
Gordon, R. (2001). Making waves: the paradigms of developmental biology and their impact on artificial life and embryonics. Cybernet. Syst. 32, 443–458.CrossRefGoogle Scholar
Gordon, R. and Beloussov, L. (2006). From observations to paradigms; the importance of theories and models: an interview with Hans Meinhardt. Int. J. Dev. Biol. 50, 103–111.CrossRefGoogle Scholar
Gorfinkiel, N., Morata, G., and Guerrero, I. (1997). The homeobox gene Distal-less induces ventral appendage development in Drosophila. Genes Dev. 11, 2259–2271.CrossRefGoogle ScholarPubMed
Gorfinkiel, N., Sánchez, L., and Guerrero, I. (1999). Drosophila terminalia as an appendage-like structure. Mechs. Dev. 86, 113–123.CrossRefGoogle ScholarPubMed
Gottlieb, A. (2012). It ain’t necessarily so. The New Yorker Sept. 17, 2012, 84–89.Google Scholar
Gould, G.C. and MacFadden, B.J. (2004). Gigantism, dwarfism, and Cope’s Rule: “Nothing in evolution makes sense without a phylogeny”. Bull. Am. Mus. Nat. Hist. 285, 219–237.2.0.CO;2>CrossRefGoogle Scholar
Gould, S.J. (1966). Allometry and size in ontogeny and phylogeny. Biol. Rev. 41, 587–640.CrossRefGoogle ScholarPubMed
Gould, S.J. (1970). Dollo on Dollo’s law: irreversibility and the status of evolutionary laws. J. Hist. Biol. 3, 189–212.CrossRefGoogle ScholarPubMed
Gould, S.J. (1971). D’Arcy Thompson and the science of form. New Lit. Hist. 2, 229–258.CrossRefGoogle Scholar
Gould, S.J. (1977). Ontogeny and Phylogeny. Harvard University Press, Cambridge, MA.Google Scholar
Gould, S.J. (1977). The return of hopeful monsters. Nat. Hist. 86 #6, 22–30.Google Scholar
Gould, S.J. (1980). The evolutionary biology of constraint. Proc. Am. Acad. Arts Sci. 109 #2, 39–52.Google Scholar
Gould, S.J. (1980). Is a new and general theory of evolution emerging? Paleobiol. 6, 119–130.CrossRefGoogle Scholar
Gould, S.J. (1980). The Panda’s Thumb: More Reflections in Natural History. Norton, New York, NY.Google Scholar
Gould, S.J. (1981). Kingdoms without wheels. Nat. Hist. 90 #3, 42–48. [See also Burrows, M. and Sutton, G. (2013). Interacting gears synchronize propulsive leg movements in a jumping insect. Science 341, 1254–1256.]Google Scholar
Gould, S.J. (1981). Quaggas, coiled oysters, and flimsy facts. Nat. Hist. 90 #9, 16–26.Google Scholar
Gould, S.J. (1981). What color is a zebra? Nat. Hist. 90 #8, 16–22.Google Scholar
Gould, S.J. (1981). What, if anything, is a zebra? Nat. Hist. 90 #7, 6–12.Google Scholar
Gould, S.J. (1982). Darwinism and the expansion of evolutionary theory. Science 216, 380–387.CrossRefGoogle ScholarPubMed
Gould, S.J. (1983). Hen’s Teeth and Horse’s Toes. Norton, New York, NY.Google Scholar
Gould, S.J. (1985). The flamingo’s smile. Nat. Hist. 94 #3, 6–19.Google Scholar
Gould, S.J. (1985). Geoffroy and the homeobox. Nat. Hist. 94 #11, 12–23.Google Scholar
Gould, S.J. (1985). Nasty little facts. Nat. Hist. 94 #2, 14–25.Google Scholar
Gould, S.J. (1985). To be a platypus. Nat. Hist. 94 #8, 10–15.Google Scholar
Gould, S.J. (1986). Archetype and adaptation. Nat. Hist. 95 #10, 16–27.Google Scholar
Gould, S.J. (1986). The egg-a-day barrier. Nat. Hist. 95 #7, 16–24.Google Scholar
Gould, S.J. (1986). Play it again, life. Nat. Hist. 95 #2, 18–26.Google Scholar
Gould, S.J. (1988). The heart of terminology. Nat. Hist. 97 #2, 24–31.Google Scholar
Gould, S.J. (1989). Full of hot air. Nat. Hist. 98 #10, 28–38.Google Scholar
Gould, S.J. (1989). Through a lens darkly. Nat. Hist. 98 #9, 16–24.Google Scholar
Gould, S.J. (1990). Bent out of shape. Nat. Hist. 99 #5, 12–27.Google Scholar
Gould, S.J. (1990). Everlasting legends. Nat. Hist. 99 #6, 12–17.Google Scholar
Gould, S.J. (1990). Wonderful Life: The Burgess Shale and the Nature of History. Norton, New York, NY.Google Scholar
Gould, S.J. (1991). Eight (or fewer) little piggies. Nat. Hist. 100 #1, 22–29.Google Scholar
Gould, S.J. (1991). Exaptation: a crucial tool for an evolutionary psychology. J. Social Issues 47 #3, 43–65.CrossRefGoogle Scholar
Gould, S.J. (1994). Common pathways of illumination. Nat. Hist. 103 #12, 10–20.Google Scholar
Gould, S.J. (1994). Hooking Leviathan by its past. Nat. Hist. 103 #5, 8–15.Google Scholar
Gould, S.J. (1995). Not necessarily a wing. Nat. Hist. 94 #10, 12–25.Google Scholar
Gould, S.J. (1996). The tallest tale: Is the textbook version of giraffe evolution a bit of a stretch? Nat. Hist. 105 #5, 18–23, 54–57.Google Scholar
Gould, S.J. (1997). Unanswerable questions. In A Glorious Accident (Kayzer, W., ed.). W. H. Freeman, New York, NY, pp. 75–104.Google Scholar
Gould, S.J. (2000). Abscheulich! (atrocious!): Haeckel’s distortions did not help Darwin. Nat. Hist. 109 #2, 42–49.Google Scholar
Gould, S.J. and Lewontin, R.C. (1979). The spandrels of San Marco and the Panglossian paradigm: a critique of the adaptationist programme. Proc. Roy. Soc. Lond. B 205, 581–598.CrossRefGoogle Scholar
Gould, S.J. and Vrba, E.S. (1982). Exaptation: a missing term in the science of form. Paleobiology 8, 4–15.CrossRefGoogle Scholar
Gracheva, E.O., Cordero-Morales, J.F., González-Carcacía, J.A., Ingolia, N.T., Manno, C., Aranguren, C.I., Weissman, J.S., and Julius, D. (2011). Ganglion-specific splicing of TRPV1 underlies infrared sensation in vampire bats. Nature 476, 88–91.CrossRefGoogle ScholarPubMed
Gracheva, E.O., Ingolia, N.T., Kelly, Y.M., Cordero-Morales, J.F., Hollopeter, G., Chesler, A.T., Sánchez, E.E., Perez, J.C., Weissman, J.S., and Julius, D. (2010). Molecular basis of infrared detection by snakes. Nature 464, 1006–1011.CrossRefGoogle ScholarPubMed
Graf, W. and Baker, R. (1983). Adaptive changes of the vestibulo-ocular reflex in flatfish are achieved by reorganization of central nervous pathways. Science 221, 777–779.CrossRefGoogle ScholarPubMed
Graham, A. and McGonnell, I. (1999). Developmental evolution: This side of paradise. Curr. Biol. 9, R630–R632.CrossRefGoogle Scholar
Graham, A. and McGonnell, I. (1999). Limb development: farewell to arms. Curr. Biol. 9, R368–R370.CrossRefGoogle Scholar
Grant, P.R., Grant, B.R., and Abzhanov, A. (2006). A developing paradigm for the development of bird beaks. Biol. J. Linnean Soc. 88, 17–22.CrossRefGoogle Scholar
Grant, R. (2012). How tigers get their stripes. The Scientist Mag. 2012 Feb. 22, Article 31728 (1 p.).Google Scholar
Grantham, T.A. (2004). Constraints and spandrels in Gould’s Structure of Evolutionary Theory. Biol. Philos. 19, 29–43.CrossRefGoogle Scholar
Graves, J.A.M. (2013). How to evolve new vertebrate sex determining genes. Dev. Dynamics 242, 354–359.CrossRefGoogle ScholarPubMed
Gray, R.S., Roszko, I., and Solnica-Krezel, L. (2011). Planar cell polarity: coordinating morphogenetic cell behaviors with embryonic polarity. Dev. Cell 21, 120–133.CrossRefGoogle ScholarPubMed
Greenberg, L. and Hatini, V. (2009). Essential roles for lines in mediating leg and antennal proximodistal patterning and generating a stable Notch signaling interface at segment borders. Dev. Biol. 330, 93–104.CrossRefGoogle ScholarPubMed
Greene, H.W. (1997). Snakes: The Evolution of Mystery in Nature. University of California Press, Berkeley, CA.Google Scholar
Greene, H.W. and Cundall, D. (2000). Limbless tetrapods and snakes with legs. Science 287, 1939–1941.CrossRefGoogle ScholarPubMed
Greer, J.M., Puetz, J., Thomas, K.R., and Capecchi, M.R. (2000). Maintenance of functional equivalence during paralogous Hox gene evolution. Nature 403, 661–665.CrossRefGoogle ScholarPubMed
Gregor, T., McGregor, A.P., and Wieschaus, E.F. (2008). Shape and function of the Bicoid morphogen gradient in dipteran species with different sized embryos. Dev. Biol. 316, 350–358.CrossRefGoogle ScholarPubMed
Grenier, J.K. and Carroll, S.B. (2000). Functional evolution of the Ultrabithorax protein. PNAS 97 #2, 704–709.CrossRefGoogle ScholarPubMed
Griffin, K.J.P., Stoller, J., Gibson, M., Chen, S., Yelon, D., Stainier, D.Y.R., and Kimelman, D. (2000). A conserved role for H15-related T-box transcription factors in zebrafish and Drosophila heart formation. Dev. Biol. 218, 235–247.CrossRefGoogle ScholarPubMed
Griffiths, M. (1988). The platypus. Sci. Am. 258 #5, 84–91.CrossRefGoogle Scholar
Grimaldi, D. and Engel, M.S. (2005). Evolution of the Insects. Cambridge University Press, New York, NY.Google Scholar
Griswold, C.K. (2006). Pleiotropic mutation, modularity and evolvability. Evol. Dev. 8, 81–93.CrossRefGoogle ScholarPubMed
Gross, J.C. and Boutros, M. (2013). Secretion and extracellular space travel of Wnt proteins. Curr. Opin. Gen. Dev. 23, 385–390.CrossRefGoogle ScholarPubMed
Grünbaum, B. and Shephard, G.C. (1987). Tilings and Patterns. W.H. Freeman, New York, NY.Google Scholar
Grus, W.E. and Zhang, J. (2006). Origin and evolution of the vertebrate vomeronasal system viewed through system-specific genes. BioEssays 28, 709–718.CrossRefGoogle ScholarPubMed
Gubb, D. (1985). Further studies on engrailed mutants in Drosophila melanogaster. Wilhelm Roux’s Arch. 194, 236–246.CrossRefGoogle Scholar
Guerreiro, I., Nunes, A., Woltering, J.M., Casaca, A., Nóvoa, A., Vinagre, T., Hunter, M.E., Duboule, D., and Mallo, M. (2013). Role of a polymorphism in a Hox/Pax-responsive enhancer in the evolution of the vertebrate spine. PNAS 110 #26, 10682–10686.CrossRefGoogle Scholar
Guichard, C., Harricane, M.-C., Lafitte, J.-J., Godard, P., Zaegel, M., Tack, V., Lalau, G., and Bouvagnet, P. (2001). Axonemal dynein intermediate-chain gene (DNAI1) mutations result in situs inversus and primary ciliary diskinesia (Kartagener syndrome). Am. J. Hum. Genet. 68, 1030–1035.CrossRefGoogle Scholar
Guild, G.M., Connelly, P.S., Ruggiero, L., Vranich, K.A., and Tilney, L.G. (2005). Actin filament bundles in Drosophila wing hairs: hairs and bristles use different strategies for assembly. Mol. Biol. Cell 16, 3620–3631.CrossRefGoogle ScholarPubMed
Guo, Z.V. and Mahadevan, L. (2008). Limbless undulatory propulsion on land. PNAS 105 #9, 3179–3184.CrossRefGoogle ScholarPubMed
Guruharsha, K.G., Kankel, M.W., and Artavanis-Tsakonas, S. (2012). The Notch signalling system: recent insights into the complexity of a conserved pathway. Nature Rev. Gen. 13, 654–666.CrossRefGoogle ScholarPubMed
Gustavson, E., Goldsborough, A.S., Ali, Z., and Kornberg, T.B. (1996). The Drosophila engrailed and invected genes: partners in regulation, expression and function. Genetics 142, 893–906.Google ScholarPubMed
Gutowitz, H., ed. Cellular Automata: Theory and Experiment. MIT Press, Cambridge, MA.
Gwin, P. (2012). Rhino wars. Natl. Geogr. 221 #3, 106–125.Google Scholar
Haag, E.S. and Lenski, R.E. (2011). L’enfant terrible at 30: the maturation of evolutionary developmental biology. Development 138, 2633–2637.CrossRefGoogle ScholarPubMed
Haag, E.S. and True, J.R. (2001). From mutants to mechanisms? Assessing the candidate gene paradigm in evolutionary biology. Evolution 55, 1077–1084.Google ScholarPubMed
Haas, M.S., Brown, S.J., and Beeman, R.W. (2001). Homeotic evidence for the appendicular origin of the labrum in Tribolium castaneum. Dev. Genes Evol. 211, 96–102.CrossRefGoogle Scholar
Haas, M.S., Brown, S.J., and Beeman, R.W. (2001). Pondering the procephalon: the segmental origin of the labrum. Dev. Genes Evol. 211, 89–95.CrossRefGoogle ScholarPubMed
Hack, M.A. and Rubenstein, D.I. (1998). Zebra zones. Nat. Hist. 107 #2, 26–33.Google Scholar
Hadorn, E. (1961). Developmental Genetics and Lethal Factors. Methuen, London (translated from 1955 German original, Thieme Verlag, Stuttgart, by Mittwoch, U.).Google Scholar
Halanych, K.M. (2004). The new view of animal phylogeny. Annu. Rev. Ecol. Evol. Syst. 35, 229–256.CrossRefGoogle Scholar
Haldane, J.B.S. (1928). On being the right size. In Possible Worlds and Other Papers. Harper, New York, NY, pp. 20–28.Google Scholar
Haldane, J.B.S. (1928). Possible Worlds and Other Papers. Harper, New York, NY.Google Scholar
Halder, G., Callaerts, P., and Gehring, W.J. (1995). Induction of ectopic eyes by targeted expression of the eyeless gene in Drosophila. Science 267, 1788–1792.CrossRefGoogle ScholarPubMed
Hall, B.K. (1984). Developmental mechanisms underlying the formation of atavisms. Biol. Rev. 59, 89–124.CrossRefGoogle Scholar
Hall, B.K. (2003). Descent with modification: the unity underlying homology and homoplasy as seen through an analysis of development and evolution. Biol. Rev. 78, 409–433.CrossRefGoogle ScholarPubMed
Hall, B.K., ed. (2007) Fins into Limbs: Evolution, Development, and Transformation. University of Chicago Press, Chicago, IL. [See also the amazing video of a walking shark by going to the Los Angeles Times website () and typing “walking shark” in the search window.]Google Scholar
Hall, B.K. (2009). The Neural Crest and Neural Crest Cells in Vertebrate Development and Evolution. Springer, New York, NY.CrossRefGoogle Scholar
Hall, B.K. (2012). Parallelism, deep homology, and evo-devo. Evol. Dev. 14, 29–33.CrossRefGoogle ScholarPubMed
Hall, B.K. and Kerney, R. (2012). Levels of biological organization and the origin of novelty. J. Exp. Zool. (Mol. Dev. Evol.) 318B, 428–437.CrossRefGoogle Scholar
Hall, B.K. and Olson, W.M., eds. (2003) Keywords and Concepts in Evolutionary Developmental Biology. Harvard University Press, Cambridge, MA.
Hallgrímsson, B. and Hall, B.K., eds. (2005). Variation: A Central Concept in Biology. Elsevier Academic Press, New York, NY.
Hallgrímsson, B., Jamniczky, H.A., Young, N.M., Rolian, C., Schmidt-Ott, U., and Marcucio, R.S. (2012). The generation of variation and the developmental basis for evolutionary novelty. J. Exp. Zool. (Mol. Dev. Evol.) 318B, 501–517.CrossRefGoogle Scholar
Halloy, J., Bernard, B.A., Loussouarn, G., and Goldbeter, A. (2000). Modeling the dynamics of human hair cycles by a follicular automaton. PNAS 97 #15, 8328–8333.CrossRefGoogle ScholarPubMed
Hallsson, J.H., Haflidadóttir, B.S., Stivers, C., Odenwald, W., Arnheiter, H., Pignoni, F., and Steingrímsson, E. (2004). The basic helix-loop-helix leucine zipper transcription factor Mitf is conserved in Drosophila and functions in eye development. Genetics 167, 233–241.CrossRefGoogle ScholarPubMed
Hamilton, C. (2010). Requiem for a Species: Why We Resist the Truth about Climate Change. Earthscan, New York, NY.Google Scholar
Hammer, M.F. (2013). Human hybrids. Sci. Am. 308 #5, 66–71.CrossRefGoogle ScholarPubMed
Hamrick, M.W. (2001). Development and evolution of the mammalian limb: adaptive diversification of nails, hooves, and claws. Evol. Dev. 3, 355–363.CrossRefGoogle ScholarPubMed
Hamrick, M.W. (2012). The developmental origins of mosaic evolution in the primate limb skeleton. Evol. Biol. 39, 447–455.CrossRefGoogle Scholar
Han, J., Lee, J.-E., Jin, J., Lim, J.S., Oh, N., Kim, K., Chang, S.-I., Shibuya, M., Kim, H., and Koh, G.Y. (2011). The spatiotemporal development of adipose tissue. Development 138, 5027–5037.CrossRefGoogle ScholarPubMed
Han, K.-A. and Kim, Y.-C. (2010). Courtship behavior: the right touch stimulates the proper song. Curr. Biol. 20, R25–R28.CrossRefGoogle ScholarPubMed
Hancock, J.M. (2005). Gene factories, microfunctionalization and the evolution of gene families. Trends Genet. 21, 591–595.CrossRefGoogle ScholarPubMed
Handrigan, G.R. (2003). Concordia discors: duality in the origin of the vertebrate tail. J. Anat. 202, 255–267.CrossRefGoogle ScholarPubMed
Handrigan, G.R. and Richman, J.M. (2011). Unicuspid and bicuspid tooth crown formation in squamates. J. Exp. Zool. (Mol. Dev. Evol.) 316B, 598–608.CrossRefGoogle Scholar
Handrigan, G.R. and Wassersug, R.J. (2007). The anuran Bauplan: a review of the adaptive, developmental, and genetic underpinnings of frog and tadpole morphology. Biol. Rev. 82, 1–25.CrossRefGoogle ScholarPubMed
Hanlon, R. (2007). Cephalopod dynamic camouflage. Curr. Biol. 17, R400–R404.CrossRefGoogle ScholarPubMed
Hansen, T.F. (2011). Epigenetics: adaptation or contingency? In Epigenetics: Linking Genotype and Phenotype in Development and Evolution (Hallgrímsson, B. and Hall, B.K., eds.). University of California Press, Berkeley, CA, pp. 357–376.Google Scholar
Harding, K. and Levine, M. (1988). Gap genes define the limits of Antennapedia and Bithorax gene expression during early development in Drosophila. EMBO J. 7, 205–214.Google ScholarPubMed
Hare, E.E., Peterson, B.K., and Eisen, M.B. (2008). A careful look at binding site reorganization in the even-skipped enhancers of Drosophila and sepsids. PLoS Genet. 4 #11, e1000268.CrossRefGoogle Scholar
Harima, Y. and Kageyama, R. (2013). Oscillatory links of Fgf signaling and Hes7 in the segmentation clock. Curr. Opin. Gen. Dev. 23, 484–490.CrossRefGoogle ScholarPubMed
Harjunmaa, E., Kallonen, A., Voutilainen, M., Hämäläinen, K., Mikkola, M.L., and Jernvall, J. (2012). On the difficulty of increasing dental complexity. Nature 483, 324–327.CrossRefGoogle ScholarPubMed
Harley, E.H., Knight, M.H., Lardner, C., Wooding, B., and Gregor, M. (2009). The Quagga project: progress over 20 years of selective breeding. S. Afr. J. Wildlife Res. 39 #2, 155–163.CrossRefGoogle Scholar
Harper, C.J., Swartz, S.M., and Brainerd, E.L. (2013). Specialized bat tongue is a hemodynamic nectar mop. PNAS 110 #22, 8852–8857.CrossRefGoogle ScholarPubMed
Harris, A.K., Stopak, D., and Warner, P. (1984). Generation of spatially periodic patterns by a mechanical instability: a mechanical alternative to the Turing model. J. Embryol. Exp. Morph. 80, 1–20.Google ScholarPubMed
Harris, M.L. and Erickson, C.A. (2007). Lineage specification in neural crest cell pathfinding. Dev. Dynamics 236, 1–19.CrossRefGoogle ScholarPubMed
Harris, M.P., Hasso, S.M., Ferguson, M.W.J., and Fallon, J.F. (2006). The development of archosaurian first-generation teeth in a chicken mutant. Curr. Biol. 16, 371–377.CrossRefGoogle Scholar
Hart, T.B. and Hart, J.A. (1992). Between sun and shadow. Nat. Hist. 101 #11, 28–35.Google Scholar
Hartenstein, V. (1993). Atlas of Drosophila Development. Cold Spring Harbor Laboratory Press, Cold Spring Harbor , NY.Google Scholar
Harvey, P.H. and Arnold, S.J. (1982). Female mate choice and runaway sexual selection. Nature 297, 533–534.CrossRefGoogle ScholarPubMed
Hasenfuss, I. (2002). A possible evolutionary pathway to insect flight starting from lepismatid organization. J. Zool. Syst. Evol. Research 40, 65–81.CrossRefGoogle Scholar
Hashimoto, H., Mizuta, A., Okada, N., Suzuki, T., Tagawa, M., Tabata, K., Yokoyama, Y., Sakaguchi, M., Tanaka, M., and Toyohara, H. (2002). Isolation and characterization of a Japanese flounder clonal line, reversed, which exhibits reversal of metamorphic left-right asymmetry. Mechs. Dev. 111, 17–24.CrossRefGoogle ScholarPubMed
Haskel-Ittah, M., Ben-Zvi, D., Branski-Arieli, M., Schejter, E.D., Shilo, B.-Z., and Barkai, N. (2012). Self-organized shuttling: generating sharp dorsoventral polarity in the early Drosophila embryo. Cell 150, 1016–1028.CrossRefGoogle ScholarPubMed
Hauswirth, R., Haase, B., Blatter, M., Brooks, S.A., Burger, D., Drögemüller, C., Gerber, V., Henke, D., Janda, J., Jude, R., Magdesian, K.G., Matthews, J.M., Poncet, P.-A., Svansson, V., Tozaki, T., Wilkinson-White, L., Penedo, M.C.T., Rieder, S., and Leeb, T. (2012). Mutations in MITF and PAX3 cause “splashed white” and other white spotting phenotypes in horses. PLoS Genet. 8 #4, e1002653.CrossRefGoogle Scholar
Hautier, L., Weisbecker, V., Sánchez-Villagra, M.R., Goswami, A., and Asher, R.J. (2010). Skeletal development in sloths and the evolution of mammalian vertebral patterning. PNAS 107 #44, 18903–18908.CrossRefGoogle ScholarPubMed
Hayashi, T. and Murakami, R. (2001). Left-right asymmetry in Drosophila melanogaster gut development. Develop. Growth Differ. 43, 239–246.CrossRefGoogle ScholarPubMed
Hayden, T. (2011). How to hatch a dinosaur. Wired 19 #10, 150–157, 186.Google Scholar
Hayes, B. (1984). Computer recreations: The cellular automaton offers a model of the world and a world unto itself. Sci. Am. 250 #3, 12–21.CrossRefGoogle Scholar
Haynie, J.L. (1982). Homologies of positional information in thoracic imaginal discs of Drosophila melanogaster. Wilhelm Roux’s Arch. 191, 293–300.CrossRefGoogle Scholar
Hays, R., Buchanan, K.T., Neff, C., and Orenic, T.V. (1999). Patterning of Drosophila leg sensory organs through combinatorial signaling by Hedgehog, Decapentaplegic and Wingless. Development 126, 2891–2899.Google ScholarPubMed
Head, J.J., Bloch, J.I., Hastings, A.K., Bourque, J.R., Cadena, E.A., Herrera, F.A., Polly, P.D., and Jaramillo, C.A. (2009). Giant boid snake from the Palaeocene neotropics reveals hotter past equatorial temperatures. Nature 457, 715–717.CrossRefGoogle ScholarPubMed
Head, J.J. and Polly, P.D. (2007). Dissociation of somatic growth from segmentation drives gigantism in snakes. Biol. Lett. 3, 296–298.CrossRefGoogle ScholarPubMed
Headland, T.N. and Greene, H.W. (2011). Hunter-gatherers and other primates as prey, predators, and competitors of snakes. PNAS 108 #52, E1470–E1474.CrossRefGoogle ScholarPubMed
Hedges, S.B. (2012). Amniote phylogeny and the position of turtles. BMC Biol. 10, Article 64 (2 pp.).CrossRefGoogle ScholarPubMed
Heeren, F. (2011). Rise of the titans. Nature 475, 159–161.CrossRefGoogle ScholarPubMed
Heers, A.M. and Dial, K.P. (2012). From extant to extinct: locomotor ontogeny and the evolution of avian flight. Trends Ecol. Evol. 27, 296–305.CrossRefGoogle ScholarPubMed
Heffer, A. and Pick, L. (2013). Conservation and variation in Hox genes: how insect models pioneered the evo-devo field. Annu. Rev. Entomol. 58, 161–179.CrossRefGoogle ScholarPubMed
Heine, P., Dohle, E., Bumsted-O’Brien, K., Engelkamp, D., and Schulte, D. (2008). Evidence for an evolutionary conserved role of homothorax/Meis1/2 during vertebrate retina development. Development 135, 805–811.CrossRefGoogle ScholarPubMed
Hejnol, A. and Martindale, M.Q. (2008). Acoel development indicates the independent evolution of the bilaterian mouth and anus. Nature 456, 382–386.CrossRefGoogle ScholarPubMed
Held, L.I. (1979). Pattern as a function of cell number and cell size on the second-leg basitarsus of Drosophila. Wilhelm Roux’s Arch. 187, 105–127.CrossRefGoogle Scholar
Held, L.I. (1990). Sensitive periods for abnormal patterning on a leg segment in Drosophila melanogaster. Roux’s Arch. Dev. Biol. 199, 31–47.CrossRefGoogle Scholar
Held, L.I. (1991). Bristle patterning inDrosophila. BioEssays 13, 633–640.CrossRefGoogle Scholar
Held, L.I. (1992). Models for Embryonic Periodicity. Monographs in Developmental Biology, Vol. 24. Karger, Basel.Google ScholarPubMed
Held, L.I. (1993). Segment-polarity mutations cause stripes of defects along a leg segment in Drosophila. Dev. Biol. 157, 240–250.CrossRefGoogle ScholarPubMed
Held, L.I. (1995). Axes, boundaries and coordinates: the ABCs of fly leg development. BioEssays 17, 721–732.CrossRefGoogle ScholarPubMed
Held, L.I. (2002). Bristles induce bracts via the EGFR pathway on Drosophila legs. Mechs. Dev. 117, 225–234.CrossRefGoogle ScholarPubMed
Held, L.I. (2002). Imaginal Discs: The Genetic and Cellular Logic of Pattern Formation. Developmental and Cell Biology Series, Vol. 39. Cambridge University Press, New York, NY.CrossRefGoogle Scholar
Held, L.I. (2002). Why should transverse rows need the EGFR pathway to align properly on Drosophila legs?Drosophila Info. Serv. 85, 17–20.Google Scholar
Held, L.I. (2009). Quirks of Human Anatomy: An Evo-Devo Look at the Human Body. Cambridge University Press, New York, NY.CrossRefGoogle Scholar
Held, L.I. (2010). The evo-devo puzzle of human hair patterning. Evol. Biol. 37, 113–122.CrossRefGoogle Scholar
Held, L.I. (2010). The evolutionary geometry of human anatomy: discovering our inner fly. Evol. Anthrop. 19, 227–235.CrossRefGoogle Scholar
Held, L.I. (2010). How does Scr cause first legs to deviate from second legs?Dros. Info. Serv. 93, 132–146.Google Scholar
Held, L.I. (2013). Rethinking butterfly eyespots. Evol. Biol. 40, 158–168.CrossRefGoogle Scholar
Held, L.I., Duarte, C.M., and Derakhshanian, K. (1986). Extra tarsal joints and abnormal cuticular polarities in various mutants of Drosophila melanogaster. Roux’s Arch. Dev. Biol. 195, 145–157.CrossRefGoogle Scholar
Held, L.I., Grimson, M.J., and Du, Z. (2004). Proving an old prediction: The sex comb rotates at 16 to 24 hours after pupariation. Drosophila Info. Serv. 87, 76–78.Google Scholar
Held, L.I. and Heup, M. (1996). Genetic mosaic analysis of decapentaplegic and wingless gene function in the Drosophila leg. Dev. Genes Evol. 206, 180–194.CrossRefGoogle ScholarPubMed
Held, L.I, Heup, M.A., Sappington, J.M., and Peters, S.D. (1994). Interactions of decapentaplegic, wingless, and Distal-less in the Drosophila leg. Roux’s Arch. Dev. Biol. 203, 310–319.CrossRefGoogle ScholarPubMed
Helfman, G.S., Collette, B.B., Facey, D.E., and Bowen, B.W. (2009). The Diversity of Fishes: Biology, Evolution and Ecology, 2nd edn. Wiley-Blackwell, Oxford.Google Scholar
Heller, K. (2011). How bird necks get naked. PLoS Biol. 9 #3, e1001029.CrossRefGoogle ScholarPubMed
Helms, J.A. and Brugmann, S.A. (2007). The origins of species-specific facial morphology: the proof is in the pigeon. Integr. Comp. Biol. 47, 338–342.CrossRefGoogle ScholarPubMed
Hendrikse, J.L., Parsons, T.E., and Hallgrímsson, B. (2007). Evolvability as the proper focus of evolutionary developmental biology. Evol. Dev. 9, 393–401.CrossRefGoogle ScholarPubMed
Hersh, B.M., Nelson, C.E., Stoll, S.J., Norton, J.E., Albert, T.J., and Carroll, S.B. (2007). The UBX-regulated network in the haltere imaginal disc of D. melanogaster. Dev. Biol. 302, 717–727.CrossRefGoogle ScholarPubMed
Hershkovitz, P. (1987). Uacaries, New World monkeys of the genus Cacajao (Cebidae, Platyrrhini): a preliminary taxonomic review with the description of a new subspecies. Am. J. Primatol. 12, 1–53.CrossRefGoogle Scholar
Hey, J. (2004). What’s so hot about recombination hotspots? PLoS Biol. 2 #6, 0730.CrossRefGoogle ScholarPubMed
Hibino, T., Ishii, Y., Levin, M., and Nishino, A. (2006). Ion flow regulates left-right asymmetry in sea urchin development. Dev. Genes Evol. 216, 265–276.CrossRefGoogle ScholarPubMed
Hibino, T., Nishino, A., and Amemiya, S. (2006). Phylogenetic correspondence of the body axes in bliaterians is revealed by the right-sided expression of Pitx genes in echinoderm larvae. Develop. Growth Differ. 48, 587–595.CrossRefGoogle ScholarPubMed
Hieronymus, T.L., Witmer, L.M., and Ridgely, R.C. (2006). Structure of white rhinoceros (Ceratotherium simum) horn investigated by X-ray computed tomography and histology with implications for growth and external form. J. Morph. 267, 1172–1176.CrossRefGoogle ScholarPubMed
Higgie, M. and Blows, M.W. (2008). The evolution of reproductive character displacement conflicts with how sexual selection operates within a species. Evolution 62, 1192–1203.CrossRefGoogle ScholarPubMed
Higuchi, R., Bowman, B., Freiberger, M., Ryder, O.A., and Wilson, A.C. (1984). DNA sequences from the quagga, an extinct member of the horse family. Nature 312, 282–284.CrossRefGoogle ScholarPubMed
Higuchi, R.G., Wrischnik, L.A., Oakes, E., George, M., Tong, B., and Wilson, A.C. (1987). Mitochondrial DNA of the extinct quagga: relatedness and extent of postmortem change. J. Mol. Evol. 25, 283–287.CrossRefGoogle ScholarPubMed
Hildebrand, J.G. and Shepherd, G.M. (1997). Mechanisms of olfactory discrimination: converging evidence for common principles across phyla. Annu. Rev. Neurosci. 20, 595–631.CrossRefGoogle ScholarPubMed
Hildebrand, M. (1974). Analysis of Vertebrate Structure. Wiley, New York, NY.Google Scholar
Hildebrand, M. (1985). Walking and running. In Functional Vertebrate Morphology (Hildebrand, M., Bramble, D.M., Liem, K.F., and Wake, D.B., eds.). Harvard University Press, Cambridge, MA, pp. 38–57.CrossRefGoogle Scholar
Hilgetag, C.C. and Barbasthis, H. (2009). Sculpting the brain. Sci. Am. 300 #2, 66–71.CrossRefGoogle Scholar
Hill, R.V. (2006). Comparative anatomy and histology of Xenarthran osteoderms. J. Morph. 267, 1441–1460.CrossRefGoogle ScholarPubMed
Hillis, D.M. (2007). Making evolution relevant and exciting to biology students. Evolution 61, 1261–1264.CrossRefGoogle ScholarPubMed
Hillmer, A.M., Flaquer, A., Hanneken, S., Eigelshoven, S., Kortüm, A.-K., Brockschmidt, F.F., Golla, A., Metzen, C., Thiele, H., Kolberg, S., Reinartz, R., Betz, R.C., Ruzicka, T., Hennies, H.C., Kruse, R., and Nöthen, M.M. (2008). Genome-wide scan and fine-mapping linkage study of androgenetic alopecia reveals a locus on chromosome 3q26. Am. J. Hum. Genet. 82, 737–743.CrossRefGoogle ScholarPubMed
Hirata, M., Nakamura, K.-i., and Kondo, S. (2005). Pigment cell distributions in different tissues of the zebrafish, with special reference to the striped pigment pattern. Dev. Dynamics 234, 293–300.CrossRefGoogle ScholarPubMed
Hironaka, K.-i. and Morishita, Y. (2012). Encoding and decoding of positional information in morphogen-dependent patterning. Curr. Opin. Gen. Dev. 22, 553–561.CrossRefGoogle ScholarPubMed
Hirth, F. (2010). On the origin and evolution of the tripartite brain. Brain Behav. Evol. 76, 3–10.CrossRefGoogle ScholarPubMed
Hirth, F., Kammermeier, L., Frei, E., Walldorf, U., Noll, M., and Reichert, H. (2003). An urbilaterian origin of the tripartite brain: developmental genetic insights from Drosophila. Development 130, 2365–2373.CrossRefGoogle ScholarPubMed
Hochner, B. (2012). An embodied view of octopus neurobiology. Curr. Biol. 22, R887–R892.CrossRefGoogle ScholarPubMed
Hockman, D., Cretekos, C.J., Mason, M.K., Behringer, R.R., Jacobs, D.S., and Illing, N. (2008). A second wave of Sonic hedgehog expression during the development of the bat limb. PNAS 105 #44, 16982–16987.CrossRefGoogle ScholarPubMed
Hockman, D., Mason, M.K., Jacobs, D.K., and Illing, N. (2009). The role of early development in mammalian limb diversification: a descriptive comparison of early limb development between the natal long-fingered bat (Miniopterus natalensis) and the mouse (Mus musculus). Dev. Dynamics 238, 965–979.CrossRefGoogle Scholar
Hodges, A. (1983). Alan Turing: The Enigma. Simon & Schuster, New York.Google Scholar
Hodgkin, J. (1998). Seven types of pleiotropy. Int. J. Dev. Biol. 42, 501–505.Google ScholarPubMed
Hodgkinson, A. and Eyre-Walker, A. (2011). Variation in the mutation rate across mammalian genomes. Nature Rev. Genet. 12, 756–766.CrossRefGoogle ScholarPubMed
Hodin, J. (2000). Plasticity and constraints in development and evolution. J. Exp. Zool. (Mol. Dev. Evol.) 288, 1–20.3.0.CO;2-7>CrossRefGoogle ScholarPubMed
Hoekstra, H.E. (2006). Genetics, development and evolution of adaptive pigmentation in vertebrates. Heredity 97, 222–234.CrossRefGoogle ScholarPubMed
Höfer, T., Maini, P.K., Kondo, S., and Asai, R. (1996). Turing patterns in fish skin? Nature 380, 678.CrossRefGoogle Scholar
Hoffstetter, R. and Gasc, J.-P. (1969). Vertebrae and ribs of modern reptiles. In Biology of the Reptilia, Vol. 1: Morphology, Part A (Gans, C., Bellairs, A.D., and Parsons, T.S., eds.). Academic Press, New York, NY, pp. 201–310.Google Scholar
Hofreiter, M. and Schöneberg, T. (2010). The genetic and evolutionary basis of color variation in vertebrates. Cell. Mol. Life Sci. 67, 2591–2603.CrossRefGoogle ScholarPubMed
Hofstadter, D. and Sander, E. (2013). Surfaces and Essences: Analogy as the Fuel and Fire of Thinking. Basic Books, New York, NY.Google Scholar
Hogan, B.L.M. and Kolodziej, P.A. (2002). Molecular mechanisms of tubulogenesis. Nature Rev. Genet. 3, 513–523.CrossRefGoogle ScholarPubMed
Holder, N. (1983). Developmental constraints and the evolution of vertebrate digit patterns. J. Theor. Biol. 104, 451–471.CrossRefGoogle ScholarPubMed
Holland, J.S. (2010). Hard hit. Natl. Geogr. 218 #2.Google Scholar
Holland, L.Z., Kene, M., Williams, N.A., and Holland, N.D. (1997). Sequence and embryonic expression of the amphioxus engrailed gene (AmphiEn): the metameric pattern of transcription resembles that of its segment-polarity homolog in Drosophila. Development 124, 1723–1732.Google ScholarPubMed
Holland, P.W.H. (1998). Major transitions in animal evolution: a developmental genetic perspective. Am. Zool. 38, 829–842.CrossRefGoogle Scholar
Hölldobler, B. and Wilson, E.O. (1990). The Ants. Harvard University Press, Cambridge, MA.CrossRefGoogle Scholar
Holley, S.A., Jackson, P.D., Sasai, Y., Lu, B., De Robertis, E.M., Hoffmann, F.M., and Ferguson, E.L. (1995). A conserved system for dorsal–ventral patterning in insects and vertebrates involving sog and chordin. Nature 376, 249–253.CrossRefGoogle ScholarPubMed
Holloway, M. (2000). Cuttlefish say it with skin. Nat. Hist. 109 #3, 70–79.Google Scholar
Holsinger, J.R. (1988). Troglobites: The evolution of cave-dwelling organisms. Am. Sci. 76, 146–153.Google Scholar
Holstein, T.W., Watanabe, H., and Özbek, S. (2011). Signaling pathways and axis formation in the lower metazoa. Curr. Top. Dev. Biol. 97, 137–177.CrossRefGoogle ScholarPubMed
Holt, R.D. (2000). Use it or lose it. Nature 407, 689–690.CrossRefGoogle Scholar
Horder, T.J. (2006). Gavin Rylands de Beer: how embryology foreshadowed the dilemmas of the genome. Nature Rev. Gen. 7, 892–898.CrossRefGoogle Scholar
Horne-Badovinac, S. and Munro, E. (2011). Tubular transformations. Science 333, 294–295.CrossRefGoogle ScholarPubMed
Horner, J.R., Padian, K., and de Ricqlès, A. (2005). How dinosaurs grew so large – and so small. Sci. Am. 293 #1, 56–63.CrossRefGoogle ScholarPubMed
Hosken, D.J. and Stockley, P. (2004). Sexual selection and genital evolution. Trends Ecol. Evol. 19, 87–93.CrossRefGoogle ScholarPubMed
Hou, L.-h., Zhou, Z., Martin, L.D., and Feduccia, A. (1995). A beaked bird from the Jurassic of China. Nature 377, 616–618.CrossRefGoogle Scholar
Houde, P. (1986). Ostrich ancestors found in the Northern Hemisphere suggest new hypothesis of ratite origins. Nature 324, 563–565.CrossRefGoogle ScholarPubMed
Houssaye, A., Xu, F., Helfen, L., de Buffrénil, V., Baumbach, T., and Tafforeau, P. (2011). Three-dimensional pelvis and limb anatomy of the Cenomanian hind-limbed snake Eupodophis descouensi (Squamata, Ophidia) revealed by synchrotron-radiation computed laminography. J. Vert. Paleo. 31, 2–7.CrossRefGoogle Scholar
Houston, D.C. (1994). To the vultures belong the spoils. Nat. Hist. 103 #9, 34–41.Google Scholar
Howarth, D.G., Martins, T., Chimney, E., and Donoghue, M.J. (2011). Diversification of CYCLOIDEA expression in the evolution of bilateral flower symmetry in Caprifoliaceae and Lonicera (Dipsacales). Ann. Bot. 107, 1521–1532.CrossRefGoogle Scholar
Howland, H.C., Merola, S., and Basarab, J.R. (2004). The allometry and scaling of the size of vertebrate eyes. Vision Res. 44, 2043–2065.CrossRefGoogle ScholarPubMed
Hoyer, S.C., Eckart, A., Herrel, A., Zars, T., Fischer, S.A., Hardie, S.L., and Heisenberg, M. (2008). Octopamine in male aggression of Drosophila. Curr. Biol. 18, 159–167.CrossRefGoogle ScholarPubMed
Hozumi, S., Maeda, R., Taniguchi, K., Kanai, M., Shirakabe, S., Sasamura, T., Spéder, P., Noselli, S., Aigaki, T., Murakami, R., and Matsuno, K. (2006). An unconventional myosin in Drosophila reverses the default handedness in visceral organs. Nature 440, 798–802.CrossRefGoogle ScholarPubMed
Hrycaj, S., Chesebro, J., and Popadic, A. (2010). Functional analysis of Scr during embryonic and post-embryonic development in the cockroach, Periplaneta americana. Dev. Biol. 341, 324–334.CrossRefGoogle ScholarPubMed
Hrycaj, S., Mihajlovic, M., Mahfooz, N., Couso, J.P., and Popadic, A. (2008). RNAi analysis of nubbin embryonic functions in a hemimetabolous insect, Oncopeltus fasciatus. Evol. Dev. 10, 705–716.CrossRefGoogle Scholar
Hsia, C.C. and McGinnis, W. (2003). Evolution of transcription factor function. Curr. Opin. Gen. Dev. 13, 199–206.CrossRefGoogle ScholarPubMed
Hsia, C.C., Paré, A.C., Hannon, M., Ronshaugen, M., and McGinnis, W. (2010). Silencing of an abdominal Hox gene during early development is correlated with limb development in a crustacean trunk. Evol. Dev. 12, 131–143. [See also Mallo, M. and Alonso, C.R. (2013). The regulation of Hox gene expression during animal development. Development 140, 3951–3963.]CrossRefGoogle Scholar
Hu, D. and Marcucio, R.S. (2009). Unique organization of the frontonasal ectodermal zone in birds and mammals. Dev. Biol. 325, 200–210.CrossRefGoogle ScholarPubMed
Hu, D.L., Nirody, J., Scott, T., and Shelley, M.J. (2009). The mechanics of slithering locomotion. PNAS 106 #25, 10081–10085.CrossRefGoogle ScholarPubMed
Hu, J. and He, L. (2008). Patterning mechanisms controlling digit development. J. Genet. Genomics 35, 517–524.CrossRefGoogle ScholarPubMed
Huang, R., Zhi, Q., Schmidt, C., Wilting, J., Brand-Saberi, B., and Christ, B. (2000). Sclerotomal origin of the ribs. Development 127, 527–532.Google ScholarPubMed
Hubbard, J.K., Uy, J.A.C., Hauber, M.E., Hoekstra, H.E., and Safran, R.J. (2010). Vertebrate pigmentation: from underlying genes to adaptive function. Trends Genet. 26, 231–239.CrossRefGoogle ScholarPubMed
Huber, B.A., Sinclair, B.J., and Schmitt, M. (2007). The evolution of asymmetric genitalia in spiders and insects. Biol. Rev. 82, 647–698.CrossRefGoogle ScholarPubMed
Hueber, S.D., Weiller, G.F., Djordjevic, M.A., and Frickey, T. (2010). Improving Hox protein classification across the major model organisms. PLoS ONE 5 #5, e10820.CrossRefGoogle ScholarPubMed
Huffard, C.L., Boneka, F., and Full, R.J. (2005). Underwater bipedal locomotion by octopuses in disguise. Science 307, 1927.CrossRefGoogle ScholarPubMed
Hughes, A.L. and Friedman, R. (2005). Loss of ancestral genes in the genomic evolution of Ciona intestinalis. Evol. Dev. 7, 196–200.CrossRefGoogle ScholarPubMed
Hughes, C.L. and Kaufman, T.C. (2002). Exploring the myriapod body plan: expression patterns of the ten Hox genes in a centipede. Development 129, 1225–1238.Google Scholar
Hughes, C.L. and Kaufman, T.C. (2002). Hox genes and the evolution of the arthropod body plan. Evol. Dev. 4, 459–499.CrossRefGoogle ScholarPubMed
Hughes, M.W., Wu, P., Jiang, T.-X., Lin, S.-J., Dong, C.-Y., Li, A., Hsieh, F.-J., Widelitz, R.B., and Chuong, C.M. (2011). In search of the Golden Fleece: unraveling principles of morphogenesis by studying the integrative biology of skin appendages. Integr. Biol. 3, 388–407.CrossRefGoogle ScholarPubMed
Hunt, P. and Krumlauf, R. (1992). Hox codes and positional specification in vertebrate embryonic axes. Annu. Rev. Cell Biol. 8, 227–256.CrossRefGoogle ScholarPubMed
Hunter, C.M., Caswell, H., Runge, M.C., Regehr, E.V., Amstrup, S.C., and Stirling, I. (2010). Climate change threatens polar bear populations: a stochastic demographic analysis. Ecology 91, 2883–2897.CrossRefGoogle ScholarPubMed
Hunter, L. (2011). Carnivores of the World. Princeton University Press, Princeton, NJ.Google Scholar
Hurle, J.M. and Fernandezteran, M.A. (1984). Fine-structure of the interdigital membranes during the morphogenesis of the digits of the webbed foot of the duck embryo. J. Morph. 79, 201–210.Google ScholarPubMed
Huszar, D., Sharpe, A., Hashmi, S., Bouchard, B., Houghton, A., and Jaenisch, R. (1991). Generation of pigmented stripes in albino mice by retroviral marking of neural crest melanoblasts. Development 113, 653–660.Google ScholarPubMed
Hutchinson, J.R. and Allen, V. (2009). The evolutionary continuum of limb function from early theropods to birds. Naturwissenschaften 96, 423–448.CrossRefGoogle ScholarPubMed
Hutchinson, J.R., Bates, K.T., Molnar, J., Allen, V., and Makovicky, P.J. (2011). A computational analysis of limb and body dimensions in Tyrannosaurus rex with implications for locomotion, ontogeny, and growth. PLoS ONE 6 #10, e26037.CrossRefGoogle Scholar
Hutchinson, J.R., Delmer, C., Miller, C.E., Hildebrandt, T., Pitsillides, A.A., and Boyde, A. (2011). From flat foot to fat foot: structure, ontogeny, function, and evolution of elephant “sixth toes”. Science 334, 1699–1703.CrossRefGoogle ScholarPubMed
Hutchison, V.H. (2008). Amphibians: lungs’ lift lost. Curr. Biol. 18, R392–R393.CrossRefGoogle ScholarPubMed
Ide, H. (2012). Bone pattern formation in mouse limbs after amputation at the forearm level. Dev. Dynamics 241, 435–441.CrossRefGoogle ScholarPubMed
Inaba, M., Yamanaka, H., and Kondo, S. (2012). Pigment pattern formation by contact-dependent depolarization. Science 335, 677.CrossRefGoogle ScholarPubMed
Indrebo, A., Langeland, M., Juul, H.M., Skogmo, H.K., Rengmark, A.H., and Lingaas, F. (2008). A study of inherited short tail and taillessness in Pembroke Welsh corgi. J. Small Anim. Pract. 49, 220–224.CrossRefGoogle ScholarPubMed
Infante, C.R., Park, S., Mihala, A.G., Kingsley, D.M., and Menke, D.B. (2013). Pitx broadly associates with limb enhancers and is enriched on hindlimb cis-regulatory elements. Dev. Biol. 374, 234–244.CrossRefGoogle Scholar
Ingber, D.E. and Levin, M. (2007). What lies at the interface of regenerative medicine and developmental biology? Development 134, 2541–2547.CrossRefGoogle ScholarPubMed
Ingham, P.W., Nakano, Y., and Seger, C. (2011). Mechanisms and functions of Hedgehog signalling across the metazoa. Nature Rev. Gen. 12, 393–406.CrossRefGoogle ScholarPubMed
Innan, H. and Kondrashov, F. (2010). The evolution of gene duplications: classifying and distinguishing between models. Nature Rev. Genet. 11, 97–108.CrossRefGoogle ScholarPubMed
Irish, F.J. (1989). The role of heterochrony in the origin of a novel bauplan: evolution of the ophidian skull. Geobios 22 (Suppl. 2, Ontogenèse et Evolution), 227–233.CrossRefGoogle Scholar
Irish, V.F., Martinez-Arias, A., and Akam, M. (1989). Spatial regulation of the Antennapedia and Ultrabithorax homeotic genes during Drosophila early development. EMBO J. 8, 1527–1537.Google ScholarPubMed
Iulianella, A., Melton, K.R., and Trainor, P.A. (2003). Somitogenesis: breaking new boundaries. Neuron 40, 11–14.CrossRefGoogle ScholarPubMed
Iwasato, T., Katoh, H., Nishimaru, H., Ishikawa, Y., Inoue, H., Saito, Y.M., Ando, R., Iwama, M., Takahashi, R., Negishi, M., and Itohara, S. (2007). Rac-GAP a-chimerin regulates motor-circuit formation as a key mediator of ephrinB3/EphA4 forward signaling. Cell 130, 742–753. [See also Talpalar, A.E., et al. (2013). Dual-mode operation of neuronal networks involved in left-right alternation. Nature 500, 85–88.]CrossRefGoogle ScholarPubMed
Iwashita, M., Watanabe, M., Ishii, M., Chen, T., Johnson, S.L., Kurachi, Y., Okada, N., and Kondo, S. (2006). Pigment pattern in jaguar/obelix zebrafish is caused by a Kir7.1 mutation: implications for the regulation of melanosome movement. PLoS Genet. 2 #11, e197.CrossRefGoogle ScholarPubMed
Jablonski, N.G. (2006). Skin: A Natural History. University of California Press, Berkeley, CA.Google Scholar
Jackman, W.R., Davies, S.H., Lyons, D.B., Stauder, C.K., Denton-Schneider, B.R., Jowdry, A., Aigler, S.R., Vogel, S.A., and Stock, D.W. (2013). Manipulation of Fgf and Bmp signaling in teleost fishes suggests potential pathways for the evolutionary origin of multicuspid teeth. Evol. Dev. 15, 107–118.CrossRefGoogle ScholarPubMed
Jackson, D.J., Meyer, N.P., Seaver, E., Pang, K., McDougall, C., Moy, V.N., Gordon, K., Degnan, B.M., Martindale, M.Q., Burke, R.D., and Peterson, K.J. (2010). Developmental expression of COE across the Metazoa supports a conserved role in neuronal cell-type specification and mesodermal development. Dev. Genes Evol. 220, 221–234.CrossRefGoogle ScholarPubMed
Jackson, K. (2003). The evolution of venom-delivery systems in snakes. Zool. J. Linnean Soc. 137, 337–354.CrossRefGoogle Scholar
Jackson, K. (2007). The evolution of venom-conducting fangs: insights from developmental biology. Toxicon 49, 975–981.CrossRefGoogle ScholarPubMed
Jacob, F. (1977). Evolution and tinkering. Science 196, 1161–1166.CrossRefGoogle ScholarPubMed
Jacob, F. (1982). The Possible and the Actual. University of Washington Press, Seattle, WA.Google Scholar
Jacobs, D.K., Hughes, N.C., Fitz-Gibbon, S.T., and Winchell, C.J. (2005). Terminal addition, the Cambrian radiation and the Phanerozoic evolution of bilaterian form. Evol. Dev. 7, 498–514.CrossRefGoogle ScholarPubMed
Jaeger, J., Manu, and Reinitz, J. (2012). Drosophila blastoderm patterning. Curr. Opin. Gen. Dev. 22, 533–541.CrossRefGoogle ScholarPubMed
Jain, A.K., Prabhakar, S., and Pankanti, S. (2002). On the similarity of identical twin fingerprints. Patt. Recog. 35, 2653–2663.CrossRefGoogle Scholar
Jameson, N.M., Xu, K., Yi, S.V., and Wildman, D.E. (2012). Development and annotation of shotgun sequence libraries from New World monkeys. Mol. Ecol. Resources 12, 950–955.CrossRefGoogle ScholarPubMed
Janda, C.Y., Waghray, D., Levin, A.M., Thomas, C., and García, K.C. (2012). Structural basis of Wnt recognition by Frizzled. Science 337, 59–64.CrossRefGoogle ScholarPubMed
Jane, S.M., Ting, S.B., and Cunningham, J.M. (2005). Epidermal impermeable barriers in mouse and fly. Curr. Opin. Gen. Dev. 15, 447–453.CrossRefGoogle ScholarPubMed
Janecka, J.E., Helgen, K.M., Lim, N.T.-L., Baba, M., Izawa, M., Boeadi, N.i., and Murphy, W.J. (2009). Evidence for multiple species of Sunda colugo. Curr. Biol. 18, R1001–R1002.CrossRefGoogle Scholar
Janis, C. (1994). The sabertooth’s repeat performances. Nat. Hist. 103 #4, 78–83.Google Scholar
Janssen, J.M., Monteiro, A., and Brakefield, P.M. (2001). Correlations between scale structure and pigmentation in butterfly wings. Evol. Dev. 3, 415–423.CrossRefGoogle ScholarPubMed
Janssen, R., Eriksson, B.J., Budd, G.E., Akam, M., and Prpic, N.-M. (2010). Gene expression patterns in an onychophoran reveal that regionalization predates limb segmentation in pan-arthropods. Evol. Dev. 12, 363–372.CrossRefGoogle Scholar
Janvier, P. (2008). Squint of the fossil flatfish. Nature 454, 169–170.CrossRefGoogle ScholarPubMed
Jarman, A.P. (2000). Developmental genetics: vertebrates and insects see eye to eye. Curr. Biol. 10, R857–R859.CrossRefGoogle ScholarPubMed
Jeffery, J.E., Bininda-Emonds, O.R.P., Coates, M.I., and Richardson, M.K. (2002). Analyzing evolutionary patterns in amniote embryonic development. Evol. Dev. 4, 292–302.CrossRefGoogle ScholarPubMed
Jeffery, W.R. (2009). Regressive evolution in Astyanax cavefish. Annu. Rev. Genet. 43, 25–47.CrossRefGoogle ScholarPubMed
Jegalian, B.G. and De Robertis, E.M. (1992). Homeotic transformation in the mouse induced by overexpression of a human Hox3.3 transgene. Cell 71, 901–910.CrossRefGoogle ScholarPubMed
Jegalian, B.G., Miller, R.W., Wright, C.V.E., Blum, M., and De Robertis, E.M. (1992). A Hox 3.3-lacZ transgene expressed in developing limbs. Mechs. Dev. 39, 171–180.CrossRefGoogle ScholarPubMed
Jenkins, F.A., Jr. and Walsh, D.M. (1993). An early Jurassic caecilian with limbs. Nature 365, 246–250.CrossRefGoogle Scholar
Jenner, R.A. (2000). Evolution of animal body plans: the role of metazoan phylogeny at the interface between pattern and process. Evol. Dev. 2, 208–221.CrossRefGoogle ScholarPubMed
Jenner, R.A. (2006). Unburdening evo-devo: ancestral attractions, model organisms, and basal baloney. Dev. Genes Evol. 216, 385–394.CrossRefGoogle ScholarPubMed
Jernvall, J. (2000). Linking development with generation of novelty in mammalian teeth. PNAS 97 #6, 2641–2645.CrossRefGoogle ScholarPubMed
Jernvall, J. and Thesleff, I. (2012). Tooth shape formation and tooth renewal: evolving with the same signals. Development 139, 3487–3497.CrossRefGoogle ScholarPubMed
Jheon, A.H. and Schneider, R.A. (2009). The cells that fill the bill: neural crest and the evolution of craniofacial development. J. Dent. Res. 88, 12–21.CrossRefGoogle ScholarPubMed
Ji, C., Wu, L., Zhao, W., Wang, S., and Lv, J. (2012). Echinoderms have bilateral tendencies. PLoS ONE 7 #1, e28978.CrossRefGoogle ScholarPubMed
Ji, Q., Ji, S.-A., Cheng, Y.-N., You, H.-L., , J.-C., Liu, Y.-Q., and Yuan, C.-X. (2004). Pterosaur egg with a leathery shell. Nature 432, 572.CrossRefGoogle ScholarPubMed
Jiang, J. and Chi-chung, H. (2008). Hedgehog signaling in development and cancer. Dev. Cell 15, 801–812.CrossRefGoogle ScholarPubMed
Jiménez-Guri, E., Philippe, H., Okamura, B., and Holland, P.W.H. (2007). Buddenbrockia is a cnidarian worm. Science 317, 116–118.CrossRefGoogle ScholarPubMed
Jockusch, E.L., Nulsen, C., Newfeld, S.J., and Nagy, L.M. (2000). Leg development in flies versus grasshoppers: differences in dpp expression do not lead to differences in the expression of downstream components of the leg patterning pathway. Development 127, 1617–1626.Google Scholar
Jockusch, E.L. and Ober, K.A. (2004). Hypothesis testing in evolutionary developmental biology: a case study from insect wings. J. Hered. 95, 382–396.CrossRefGoogle ScholarPubMed
Jockusch, E.L., Williams, T.A., and Nagy, L. (2004). The evolution of patterning of serially homologous appendages in insects. Dev. Genes Evol. 214, 324–338.CrossRefGoogle ScholarPubMed
Johanson, Z., Joss, J., Boisvert, C.A., Ericsson, R., Sutija, M., and Ahlberg, P.E. (2007). Fish fingers: Digit homologues in Sarcopterygian fish fins. J. Exp. Zool. (Mol. Dev. Evol.) 308B, 757–768.CrossRefGoogle Scholar
Jones, F.C., Grabherr, M.G., Chan, Y.F., Russell, P., Mauceli, E., Johnson, J., Swofford, R., Pirun, M., Zody, M.C., White, S., Birney, E., Searle, S., Schmutz, J., Grimwood, J., Dickson, M.C., Myers, R.M., Miller, C.T., Summers, B.R., Knecht, A.K., Brady, S.D., Zhang, H., Pollen, A.A., Howes, T., Amemiya, C., Broad Institutre Genome Sequencing Platform & Whole Genome Assembly Team, Lander, E.S., Di Palma, F., Lindblad-Toh, K., and Kingsley, D.M. (2012). The genomic basis of adaptive evolution in threespine sticklebacks. Nature 484, 55–61.CrossRefGoogle ScholarPubMed
Jones, G. (2010). Molecular evolution: gene convergence in echolocating mammals. Curr. Biol. 20, R62–R64. [See also Parker, J., et al. (2013). Genome-wide signatures of convergent evolution in echolocating mammals. Nature 502, 228–231.]CrossRefGoogle ScholarPubMed
Jordan, B., Vercammen, P., and Cooper, K.L. (2011). Husbandry and breeding of the lesser Egyptian jerboa, Jaculus jaculus. Cold Spr. Harb. Protoc. 2011 #12, 1457–1461.Google ScholarPubMed
Jordan, S.A. and Jackson, I.J. (2000). MGF (KIT ligand) is a chemokinetic factor for melanoblast migration into hair follicles. Dev. Biol. 225, 424–436.CrossRefGoogle ScholarPubMed
Josef, N., Amodio, P., Fiorito, G., and Shashar, N. (2012). Camouflaging in a complex environment – octopuses use specific features of their surroundings for background matching. PLoS ONE 7 #5, e37579. [See also Courage, K.H. (2013). Octopus! Penguin, New York, NY.]CrossRefGoogle Scholar
Joshi, M., Buchanan, K.T., Shroff, S., and Orenic, T.V. (2006). Delta and Hairy establish a periodic prepattern that positions sensory bristles in Drosophila legs. Dev. Biol. 293, 64–76.CrossRefGoogle ScholarPubMed
Kaelin, C.B., Xu, X., Hong, L.Z., David, V.A., McGowan, K.A., Schmitdt-Küntzel, A., Roelke, M.E., Pino, J., Pontius, J., Cooper, G.M., Manuel, H., Swanson, W.F., Marker, L., Harper, C.K., van Dyk, A., Yue, B., Mullikin, J.C., Warren, W.C., Eizirik, E., Kos, L., O’Brien, S.J., Barsh, G.S., and Menotti-Raymond, M. (2012). Specifiying and sustaining pigmentation patterns in domestic and wild cats. Science 337, 1536–1541.CrossRefGoogle Scholar
Kajiura, S.M. and Holland, K.N. (2002). Electroreception in juvenile scalloped hammerhead and sandbar sharks. J. Exp. Biol. 205, 3609–3621.Google ScholarPubMed
Kalay, G. and Wittkopp, P.J. (2010). Nomadic enhancers: Tissue-specific cis-reguatory elements of yellow have divergent genomic positions among Drosophila species. PLoS Genet. 6 #11, e1001222.CrossRefGoogle ScholarPubMed
Kalinka, A.T., Varga, K.M., Gerrard, D.T., Preibisch, S., Corcoran, D.L., Jarrells, J., Ohler, U., Bergman, C.M., and Tomancak, P. (2010). Gene expression divergence recapitulates the developmental hourglass model. Nature 468, 811–818.CrossRefGoogle ScholarPubMed
Kallman, K.D. and Kazianis, S. (2006). The genus Xiphophorus in Mexico and Central America. Zebrafish 3, 271–285.CrossRefGoogle ScholarPubMed
Kang, K., Pulver, S.R., Panzano, V.C., Chang, E.C., Griffith, L.C., Theobald, D.L., and Garrity, P.A. (2010). Analysis of Drosophila TRPA1 reveals an ancient origin for human chemical noniception. Nature 464, 597–600.CrossRefGoogle Scholar
Kankel, D.R., Ferrús, A., Garen, S.H., Harte, P.J., and Lewis, P.E. (1980). The structure and development of the nervous system. In The Genetics and Biology of Drosophila, Vol. 2d (Ashburner, M. and Wright, T.R.F., eds.). Academic Press, New York, NY, pp. 295–368.Google Scholar
Kardong, K.V. (1979). “Protovipers” and the evolution of snake fangs. Evolution 33, 433–443.Google ScholarPubMed
Kardong, K.V. (2005). An Introduction to Biological Evolution. McGraw-Hill, New York, NY.Google Scholar
Karleskint, G., Jr., Turner, R., and Small, J.W., Jr. (1999). Introduction to Marine Biology, 3rd edn. Brooks/Cole, Belmont, CA. [See also Loxton, D. (2013). Mermaids. Skeptic 18 #3, 65–71.]Google Scholar
Kauffman, S.A. (1983). Developmental constraints: internal factors in evolution. In Development and Evolution, Symp. Brit. Soc. Dev. Biol., Vol. 6 (Goodwin, B.C., Holder, N., and Wylie, C.C., eds.). Cambridge University Press, Cambridge, pp. 195–225.Google Scholar
Kauffman, S.A. (1993). The Origins of Order: Self-Organization and Selection in Evolution. Oxford University Press, Oxford.Google Scholar
Kavanagh, K.D., Evans, A.R., and Jernvall, J. (2007). Predicting evolutionary patterns of mammalian teeth from development. Nature 449, 427–432.CrossRefGoogle ScholarPubMed
Kawasaki, K. and Weiss, K.M. (2006). Evolutionary genetics of vertebrate tissue mineralization: the origin and evolution of the secretory calcium-binding phosphoprotein family. J. Exp. Zool. (Mol. Dev. Evol.) 306B, 295–316.CrossRefGoogle Scholar
Kay, E.H. and Hoekstra, H.E. (2008). Rodents. Curr. Biol. 18, R406–R410.CrossRefGoogle ScholarPubMed
Kearney, M. and Stuart, B.L. (2004). Repeated evolution of limblessness and digging heads in worm lizards revealed by DNA from old bones. Proc. R. Soc. Lond. B 271, 1677–1683.CrossRefGoogle ScholarPubMed
Kearny, M. and Stuart, B.L. (2004). Repeated evolution of limblessness and digging heads in worm lizards revealed by DNA from old bones. Proc. R. Soc. Lond. B 271, 1677–1683.CrossRefGoogle Scholar
Keays, D.A., Tian, G., Poirier, K., Huang, G.-J., Siebold, C., Cleak, J., Oliver, P.L., Fray, M., Harvey, R.J., Molnár, Z., Piñon, M.C., Dear, N., Valdar, W., Brown, S.D.M., Davies, K.E., Rawlins, J.N.P., Cowan, N.J., Nolan, P., Chelly, J., and Flint, J. (2007). Mutations in α-tubulin cause abnormal neuronal migration in mice and lissencephaly in humans. Cell 128, 45–57.CrossRefGoogle ScholarPubMed
Keddy-Hector, A.C. (1992). Mate choice in non-human primates. Am. Zool. 32, 62–70.CrossRefGoogle Scholar
Keil, T.A. and Steinbrecht, R.A. (1984). Mechanosensitive and olfactory sensilla of insects. In Insect Ultrastructure, Vol. 2 (King, R.C. and Akai, H., eds.). Plenum, New York, NY, pp. 477–516.CrossRefGoogle Scholar
Kelley, J.L., Fitzpatrick, J.L., and Merilaita, S. (2013). Spots and stripes: ecology and colour pattern evolution in butterflyfishes. Proc. R. Soc. Lond. B 280, in press. .CrossRefGoogle ScholarPubMed
Kellogg, V.L. and Bell, R.G. (1904). Studies of variation in insects. Proc. Wash. Acad. Sci. 6, 203–332.Google Scholar
Kelsh, R.N. and Barsh, G.S. (2011). A nervous origin for fish stripes. PLoS Genet. 7 #5, e1002081.CrossRefGoogle ScholarPubMed
Kelsh, R.N., Harris, M.L., Colanesi, S., and Erickson, C.A. (2009). Stripes and belly-spots: a review of pigment cell morphogenesis in vertebrates. Semin. Cell Dev. Biol. 20, 90–104.CrossRefGoogle ScholarPubMed
Kenward, B., Wachtmeister, C.-A., Ghirlanda, S., and Enquist, M. (2004). Spots and stripes: the evolution of repetition in visual signal form. J. Theor. Biol. 230, 407–419.CrossRefGoogle ScholarPubMed
Keogh, J.S., Scott, I.A.W., and Hayes, C. (2005). Rapid and repeated origin of insular gigantism and dwarfism in Australian tiger snakes. Evolution 59, 226–233.CrossRefGoogle ScholarPubMed
Kerschensteiner, D. (2011). Circuit assembly: the repulsive side of lamination. Curr. Biol. 21, R163–R166.CrossRefGoogle ScholarPubMed
Keynes, R.J. and Stern, C.D. (1988). Mechanisms of vertebrate segmentation. Development 103, 413–429.Google ScholarPubMed
Keys, D.N., Lewis, D.L., Selegue, J.E., Pearson, B.J., Goodrich, L.V., Johnson, R.L., Gates, J., Scott, M.P., and Carroll, S.B. (1999). Recruitment of a hedgehog regulatory circuit in butterfly eyespot evolution. Science 283, 532–534.CrossRefGoogle ScholarPubMed
Khadjeh, S., Turetzek, N., Pechmann, M., Schwager, E.E., Wimmer, E.A., Damen, W.G.M., and Prpic, N.-M. (2012). Divergent role of the Hox gene Antennapedia in spiders is responsible for the convergent evolution of abdominal limb repression. PNAS 109 #13, 4921–4926.CrossRefGoogle ScholarPubMed
Khila, A., Abouheif, E., and Rowe, L. (2012). Function, developmental genetics, and fitness consequences of a sexually antagonistic trait. Science 336, 585–589.CrossRefGoogle ScholarPubMed
Kicheva, A., Bollenbach, T., Wartlick, O., Jülicher, F., and Gonzalez-Gaitan, M. (2012). Investigating the principles of morphogen gradient formation: from tissues to cells. Curr. Opin. Gen. Dev. 22, 527–532.CrossRefGoogle ScholarPubMed
Kiefer, J.C. (2006). Emerging developmental model systems. Dev. Dynamics 235, 2895–2899.CrossRefGoogle ScholarPubMed
Kijimoto, T., Andrews, J., and Moczek, A.P. (2010). Programmed cell death shapes the expression of horns within and between species of horned beetles. Evol. Dev.12, 449–458.CrossRefGoogle ScholarPubMed
Kim, J., Sebring, A., Esch, J.J., Kraus, M.E., Vorwerk, K., Magee, J., and Carroll, S.B. (1996). Integration of positional signals and regulation of wing formation and identity by Drosophila vestigial gene. Nature 382, 133–138.CrossRefGoogle ScholarPubMed
Kim, J.S., Jin, D.I., Lee, J.H., Son, D.S., Lee, S.H., Yi, Y.J., and Park, C.S. (2005). Effects of teat number on litter size in gilts. Anim. Reprod. Sci. 90, 111–116.CrossRefGoogle ScholarPubMed
Kim, S.Y., Paylor, S.W., Magnuson, T., and Schumacher, A. (2007). Juxtaposed Polycomb complexes co-regulate vertebral identity. Development 133, 4957–4968.CrossRefGoogle Scholar
Kimm, M.A. and Prpic, N.-M. (2006). Formation of the arthropod labrum by fusion of paired and rotated limb-bud-like primordia. Zoomorphology 125, 147–155.CrossRefGoogle Scholar
King, N. and Carroll, S.B. (2001). A receptor tyrosine kinase from choanoflagellates: molecular insights into early animal evolution. PNAS 98 #26, 15032–15037.CrossRefGoogle ScholarPubMed
King, N., Hittinger, C.T., and Carroll, S.B. (2003). Evolution of key cell signaling and adhesion protein families predates animal origins. Science 301, 361–363.CrossRefGoogle ScholarPubMed
Kingdon, J. (1977). East African Mammals: An Atlas of Evolution in Africa: Carnivores, Vol. IIIA. University of Chicago Press, Chicago, IL.Google Scholar
Kingsley, M.C.S. and Ramsay, M.A. (1988). The spiral in the tusk of the narwhal. Artic 41, 236–238.Google Scholar
Kingsolver, J.G. and Koehl, M.A.R. (1985). Aerodynamics, thermoregulation, and the evolution of insect wings: differential scaling and evolutionary change. Evolution 39, 488–504.CrossRefGoogle ScholarPubMed
Kingsolver, J.G. and Koehl, M.A.R. (1994). Selective factors in the evolution of insect wings. Annu. Rev. Entomol. 39, 425–451.CrossRefGoogle Scholar
Kirschner, M. and Gerhart, J. (1998). Evolvability. PNAS 95 #15, 8420–8427.CrossRefGoogle ScholarPubMed
Kirschner, M.W. and Gerhart, J.C. (2005). The Plausibility of Life: Resolving Darwin’s Dilemma. Yale University Press, New Haven, CT.Google Scholar
Kitchener, A.C., Beaumont, M.A., and Richardson, D. (2006). Geographical variation in the clouded leopard, Neofelis nebulosa, reveals two species. Curr. Biol. 16, 2377–2383.CrossRefGoogle ScholarPubMed
Klauber, L.M. (1997). Rattlesnakes: Their Habits, Life Histories, and Infuence on Mankind. 2nd edn., Vol. 2. University of California Press, Berkeley, CA.Google Scholar
Klauer, G., Burda, H., and Nevo, E. (1997). Adaptive differentiations of the skin of the head in a subterranean rodent, Spalax ehrenbergi. J. Morph. 233, 53–66.3.0.CO;2-P>CrossRefGoogle Scholar
Klein, O.D., Lyons, D.B., Balooch, G., Marshall, G.W., Basson, M.A., Peterka, M., Boran, T., Peterkova, R., and Martin, G.R. (2008). An FGF signaling loop sustains the generation of differentiated progeny from stem cells in mouse incisors. Development 135, 377–385.CrossRefGoogle ScholarPubMed
Klein, O.D., Minowada, G., Peterkova, R., Kangas, A., Yu, B.D., Lesot, H., Peterka, M., Jernvall, J., and Martin, G. (2006). Sprouty genes control diastema tooth development via bidirectional antagonism of epithelial-mesenchymal FGF signaling. Dev. Cell 11, 181–190.CrossRefGoogle ScholarPubMed
Klein, R. (2012). Eph/ephrin signalling during development. Development 139, 4105–4109.CrossRefGoogle ScholarPubMed
Kley, N.J. and Brainerd, E.L. (1999). Feeding by mandibular raking in a snake. Nature 402, 369–370.CrossRefGoogle Scholar
Kley, N.J. and Kearney, M. (2007). Adaptations for digging and burrowing. In Fins into Limbs: Evolution, Development, and Transformation (Hall, B.K., ed.). University of Chicago Press, Chicago, IL, pp. 284–309.Google Scholar
Klingenberg, C.P. (1998). Heterochrony and allometry: the analysis of evolutionary change in ontogeny. Biol. Rev. 73, 79–123.CrossRefGoogle ScholarPubMed
Klingenberg, C.P. (2005). Developmental constraints, modules, and evolvability. In Variation: A Central Concept in Biology (Hallgrímsson, B. and Hall, B.K., eds.). Elsevier Academic Press, New York, NY, pp. 219–247.CrossRefGoogle Scholar
Kodandaramaiah, U. (2009). Eyespot evolution: phylogenetic insights from Junonia and related butterfly genera (Nymphalidae: Junoniini). Evol. Dev. 11, 489–497.CrossRefGoogle Scholar
Kodandaramaiah, U. (2009). Fixed eyespot display in a butterfly thwarts attacking birds. Anim. Behav. 77, 1415–1419.CrossRefGoogle Scholar
Kodandaramaiah, U. (2011). The evolutionary significance of butterfly eyespots. Behav. Ecol. 22, 1264–1271.CrossRefGoogle Scholar
Koestler, A. (1969). Beyond atomism and holism: the concept of the holon. In Beyond Reductionism. New Perspectives in the Life Sciences (Koestler, A. and Smythies, J.R., eds.). Macmillan, New York, NY, pp. 192–227.Google Scholar
Koganezawa, M., Haba, D., Matsuo, T., and Yamamoto, D. (2010). The shaping of male courtship posture by lateralized gustatory inputs to male-specific interneurons. Curr. Biol. 20, 1–8.CrossRefGoogle ScholarPubMed
Koh, T.-W. and Carlson, J.R. (2011). Chemoreception: Identifying friends and foes. Curr. Biol. 21, R998–R999.CrossRefGoogle ScholarPubMed
Kohlsdorf, T., Cummings, M.P., Lynch, V.J., Stopper, G.F., Takahashi, K., and Wagner, G.P. (2008). A molecular footprint of limb loss: sequence variation of the autopodial identity gene Hoxa-13. J. Mol. Evol. 67, 581–593.CrossRefGoogle ScholarPubMed
Kojima, T. (2004). The mechanism of Drosophila leg development along the proximodistal axis. Develop. Growth Differ. 46, 115–129.CrossRefGoogle ScholarPubMed
Kojima, T., Sato, M., and Saigo, K. (2000). Formation and specification of distal leg segments in Drosophila by dual Bar homeobox genes, BarH1and BarH2. Development 127, 769–778.Google ScholarPubMed
Kolbe, J.J., Leal, M., Schoener, T.W., Spiller, D.A., and Losos, J.B. (2012). Founder effects persist despite adaptive differentiation: a field experiment with lizards. Science 335, 1086–1089.CrossRefGoogle ScholarPubMed
Kollar, E.J. and Fisher, C. (1980). Tooth induction in chick epithelium: expression of quiescent genes for enamel synthesis. Science 207, 993–995.CrossRefGoogle ScholarPubMed
Kondo, S. (2002). The reaction-diffusion system: a mechanism for autonomous pattern formation in the animal skin. Genes to Cells 7, 535–541.CrossRefGoogle ScholarPubMed
Kondo, S. (2005). Cell-cell interaction network that generates the skin pattern of animal. Genome Informatics 16, 287–291.Google ScholarPubMed
Kondo, S. and Asai, R. (1995). A reaction-diffusion wave on the skin of the marine angelfish Pomacanthus. Nature 376, 765–768.CrossRefGoogle ScholarPubMed
Kondo, S., Iwashita, M., and Yamaguchi, M. (2009). How animals get their skin patterns: fish pigment pattern as a live Turing wave. Int. J. Dev. Biol. 53, 851–856.CrossRefGoogle ScholarPubMed
Kondo, S. and Miura, T. (2010). Reaction-diffusion model as a framework for understanding biological pattern formation. Science 329, 1616–1620.CrossRefGoogle ScholarPubMed
Kondo, S. and Shirota, H. (2009). Theoretical analysis of mechanisms that generate the pigmentation pattern of animals. Semin. Cell Dev. Biol. 20, 82–89.CrossRefGoogle ScholarPubMed
Koonin, E.V. (2005). Orthologs, paralogs, and evolutionary genomics. Annu. Rev. Genet. 39, 309–338.CrossRefGoogle ScholarPubMed
Koop, D., Holland, N.D., Sémon, M., Alvarez, S., Rodriguez de Lera, A., Laudet, V., Holland, L.Z., and Schubert, M. (2010). Retinoic acid signaling targets Hox genes during the amphioxus gastrula stage: Insights into early anterior–posterior patterning of the chordate body plan. Dev. Biol. 338, 98–106.CrossRefGoogle ScholarPubMed
Kopp, A. (2009). Metamodels and phylogenetic replication: a systematic approach to the evolution of developmental pathways. Evolution 63, 2771–2789.CrossRefGoogle ScholarPubMed
Kopp, A. (2011). Drosophila sex combs as a model of evolutionary innovations. Evol. Dev. 13, 504–522.CrossRefGoogle Scholar
Kopp, A. (2012). Dmrt genes in the development and evolution of sexual dimorphism. Trends Genet. 28, 175–184.CrossRefGoogle ScholarPubMed
Kopp, A. and True, J.R. (2002). Evolution of male sexual characters in the Oriental Drosophila melanogaster species group. Evol. Dev. 4, 278–291.CrossRefGoogle ScholarPubMed
Kornberg, T.B. and Guha, A. (2007). Understanding morphogen gradients: a problem of dispersion and containment. Curr. Opin. Gen. Dev. 17, 264–271.CrossRefGoogle ScholarPubMed
Kotiaho, J.S. (2001). Costs of sexual traits: a mismatch between theoretical and empirical evidence. Biol. Rev. 76, 365–376.CrossRefGoogle ScholarPubMed
Koyanagi, M., Kubokawa, K., Tsukamoto, H., Shichida, Y., and Terakita, A. (2005). Cephalochordate melanopsin: evolutionary linkage between invertebrate visual cells and vertebrate photosensitive retinal ganglion cells. Curr. Biol. 15, 1065–1069.CrossRefGoogle ScholarPubMed
Kozmikova, I., Smolikova, J., Vlcek, C., and Kozmik, Z. (2011). Conservation and diversification of an ancestral chordate gene regulatory network for dorsoventral patterning. PLoS One 6 #2, e14650.CrossRefGoogle ScholarPubMed
Kozopas, K.M. and Nusse, R. (2002). Direct flight muscles in Drosophila develop from cells with characteristics of founders and depend on DWnt-2for their correct patterning. Dev. Biol. 243, 312–325.CrossRefGoogle ScholarPubMed
Krajick, K. (2007). Discoveries in the dark. Natl. Geogr. 212 #3, 134–147.Google Scholar
Krauss, G. (2008). Biochemistry of Signal Transduction and Regulation, 4th edn. Wiley-VCH, Weinheim.Google Scholar
Krochmal, A.R., Bakken, G.S., and LaDuc, T.J. (2004). Heat in evolution’s kitchen: evolutionary perspectives on the functions and origin of the facial pit of pitvipers (Viperidae: Crotalinae). J. Exp. Biol. 207, 4231–4238.CrossRefGoogle Scholar
Kröger, B., Vinther, J., and Fuchs, D. (2011). Cephalopod origin and evolution: a congruent picture emerging from fossils, development and molecules. BioEssays 33, 602–613.CrossRefGoogle ScholarPubMed
Krol, A.J., Roellig, D., Dequéant, M.-L., Tassy, O., Glynn, E., Hattem, G., Mushegian, A., Oates, A.C., and Pourquié, O. (2011). Evolutionary plasticity of segmentation clock networks. Development 138, 2783–2792.CrossRefGoogle ScholarPubMed
Kronforst, M.R., Barsh, G.S., Kopp, A., Mallet, J., Monteiro, A., Mullen, S.P., Protas, M., Rosenblum, E.B., Schneider, C.J., and Hoekstra, H.E. (2012). Unraveling the thread of nature’s tapestry: the genetics of diversity and convergence in animal pigmentation. Pigment Cell Melanoma Res. 25, 411–433.CrossRefGoogle ScholarPubMed
Krumlauf, R. (1994). Hox genes in vertebrate development. Cell 78, 191–201.CrossRefGoogle ScholarPubMed
Kuch, U., Müller, J., Mödden, C., and Mebs, D. (2006). Snake fangs from the Lower Miocene of Germany: evolutionary stability of perfect weapons. Naturwissenschaften 93, 84–87.CrossRefGoogle ScholarPubMed
Kühn, A. (1971). Lectures on Developmental Physiology, 2nd edn. Springer-Verlag, Berlin.CrossRefGoogle Scholar
Kukalová-Peck, J. (1978). Origin and evolution of insect wings and their relation to metamorphosis, as documented by the fossil record. J. Morph. 156, 53–125.CrossRefGoogle Scholar
Kukalová-Peck, J. (1985). Ephemeroid wing venation based upon new gigantic Carboniferous mayflies and basic morphology, phylogeny, and metamorphosis of pterygote insects (Insecta, Ephemerida). Can. J. Zool. 63, 933–955.CrossRefGoogle Scholar
Kukalová-Peck, J. (1987). New Carboniferous Diplura, Monura, and Thysanura, the hexapod ground plan, and the role of thoracic side lobes in the origin of wings (Insecta). Can. J. Zool. 65, 2327–2345.CrossRefGoogle Scholar
Kukalová-Peck, J. (2008). Phylogeny of higher taxa in insecta: finding synapomorphies in the extant fauna and separating them from homoplasies. Evol. Biol. 35, 4–51.CrossRefGoogle Scholar
Kulesa, P.M. and Fraser, S.E. (2002). Cell dynamics during somite boundary formation revealed by time-lapse analysis. Science 298, 991–995.CrossRefGoogle ScholarPubMed
Kundrát, M. (2009). Heterochronic shift between early organogenesis and migration of cephalic neural crest cells in two divergent evolutionary phenotypes of archosaurs: crocodile and ostrich. Evol. Dev. 11, 535–546.CrossRefGoogle ScholarPubMed
Kunhardt, P.B., Jr., Kunhardt, P.B., III, and Kunhardt, P.W. (1995). P. T. Barnum: America’s Greatest Showman. Knopf, New York, NY.Google Scholar
Kuo, D.-H. and Weisblat, D.A. (2011). A new molecular logic for BMP-mediated dorsoventral patterning in the leech Helobdella. Curr. Biol. 21, 1282–1288.CrossRefGoogle ScholarPubMed
Kuranaga, E., Matsunuma, T., Kanuka, H., Takemoto, K., Koto, A., Kimura, K.-i., and Miura, M. (2011). Apoptosis controls the speed of looping morphogenesis in Drosophila male terminalia. Development 138, 1493–1499.CrossRefGoogle ScholarPubMed
Kuratani, S. (2012). Evolution of the vertebrate jaw from developmental perspectives. Evol. Dev. 14, 76–92.CrossRefGoogle ScholarPubMed
Kuratani, S., Kuraku, S., and Nagashima, H. (2011). Evolutionary developmental perspective for the origin of turtles: the folding theory for the shell based on the developmental nature of the carapacial ridge. Evol. Dev. 13, 1–14.CrossRefGoogle ScholarPubMed
Kurshan, P.T. and Shen, K. (2012). Dendritic patterning: three-dimensional position determines dendritic avoidance capability. Curr. Biol. 22, R192–R194.CrossRefGoogle ScholarPubMed
Kusumi, K., May, C.M., and Eckalbar, W.L. (2013). A large-scale view of the evolution of amniote development: insights from somitogenesis in reptiles. Curr. Opin. Gen. Dev. 23, 491–497.CrossRefGoogle ScholarPubMed
Kuzniar, A., van Ham, R.C.H.J., Pongor, S., and Leunissen, J.A.M. (2008). The quest for orthologs: finding the corresponding gene across genomes. Trends Genet. 24, 539–551.CrossRefGoogle ScholarPubMed
Kwon, C., Hays, R., Fetting, J., and Orenic, T.V. (2004). Opposing inputs by Hedgehog and Brinker define a stripe of hairy expression in the Drosophila leg imaginal disc. Development 131, 2681–2692.CrossRefGoogle ScholarPubMed
Lacalli, T. (1996). Dorsoventral axis inversion: a phylogenetic perspective. BioEssays 18, 251–254.CrossRefGoogle Scholar
Lacalli, T. (2008). Head organization and the head/trunk relationship in protochordates: problems and prospects. Integr. Comp. Biol. 48, 620–629.CrossRefGoogle ScholarPubMed
Lacalli, T. (2010). The emergence of the chordate body plan: some puzzles and problems. Acta Zool. (Stockholm) 91, 4–10.CrossRefGoogle Scholar
Lacalli, T. (2012). The middle Cambrian fossil Pikaia and the evolution of chordate swimming. EvoDevo 3, Article 12 (6 pp.).CrossRefGoogle ScholarPubMed
Lacalli, T.C. (2005). Protochordate body plan and the evolutionary role of larvae: old controversies resolved? Can. J. Zool. 83, 216–224.CrossRefGoogle Scholar
Lacalli, T.C. (2008). Mucus secretion and transport in amphioxus larvae: organization and ultrastructure of the food trapping system, and implications for head evolution. Acta Zool. (Stockholm) 89, 219–230.CrossRefGoogle Scholar
Lai, E.C. (2004). Notch signaling: control of cell communication and cell fate. Development 131, 965–973.CrossRefGoogle ScholarPubMed
Lai, E.C. and Orgogozo, V. (2004). A hidden program in Drosophila peripheral neurogenesis revealed: fundamental principles underlying sensory organ diversity. Dev. Biol. 269, 1–17.CrossRefGoogle ScholarPubMed
Lakoff, G. and Johnson, M. (2003). Metaphors We Live By, 2nd edn. University of Chicago Press, Chicago, IL.CrossRefGoogle Scholar
Laman, T. (2000). Wild gliders: the creatures of Borneo’s rain forest go airborne. Natl. Geogr. 198 #4, 68–85.Google Scholar
Lamar, W.W., Carmichael, P., and Shumway, G. (2002). The World’s Most Spectacular Reptiles & Amphibians. World Publications, Tampa, FL.Google Scholar
Lamb, T.D. (2011). Evolution of the eye. Sci. Am. 305 #1, 64–69.CrossRefGoogle Scholar
Lanctôt, C., Moreau, A., Chamberland, M., Tremblay, M.L., and Drouin, J. (1999). Hindlimb patterning and mandible development require the Ptx1 gene. Development 126, 1805–1810.Google ScholarPubMed
Land, M.F. (2006). Visual optics: the shapes of pupils. Curr. Biol. 16, R167–R168.CrossRefGoogle ScholarPubMed
Land, M.F. and Nilsson, D.-E. (2002). Animal Eyes. Oxford University Press, New York, NY.Google Scholar
Landberg, T., Mailhot, J.D., and Brainerd, E.L. (2003). Lung ventilation during treadmill locomotion in a terrestrial turtle, Terrapene carolina. J. Exp. Biol. 206, 3391–3404.CrossRefGoogle Scholar
Lande, R. (1978). Evolutionary mechanisms of limb loss in tetrapods. Evolution 32, 73–92.CrossRefGoogle ScholarPubMed
Lander, A.D. (2007). Morpheus unbound: Reimagining the morphogen gradient. Cell 128, 245–256.CrossRefGoogle ScholarPubMed
Lander, A.D. (2011). Pattern, growth, and control. Cell 144, 955–969.CrossRefGoogle Scholar
Lander, A.D. (2013). How cells know where they are. Science 339, 923–927.CrossRefGoogle Scholar
Lane, N. and Martin, W.F. (2012). The origin of membrane bioenergetics. Cell 151, 1406–1416.CrossRefGoogle ScholarPubMed
Langdon, J.H. (2005). The Human Strategy: An Evolutionary Perspective on Human Anatomy. Oxford University Press, New York, NY.Google Scholar
Langridge, K.V., Broom, M., and Osorio, D. (2007). Selective signalling by cuttlefish to predators. Curr. Biol. 17, R1044–R1045.CrossRefGoogle ScholarPubMed
Langston, W., Jr. (1981). Pterosaurs. Sci. Am. 244 #2, 122–136.CrossRefGoogle Scholar
Langton, C.G., ed. Artificial Life. Addison-Wesley, New York, NY.CrossRef
LaPolla, J.S., Dlussky, G.M., and Perrichot, V. (2013). Ants and the fossil record. Annu. Rev. Entomol. 58, 609–630. [See also Johnson, B.R., et al. (2013). Phylogenomics resolves evolutionary relationships among ants, bees, and wasps. Curr. Biol. 23, 2058–2062].CrossRefGoogle ScholarPubMed
Lapraz, F., Besnardeau, L., and Lepage, T. (2009). Patterning of the dorsal–ventral axis in echinoderms: insights into the evolution of the BMP-Chordin signaling network. PLoS Biol. 7 #11, e1000248.CrossRefGoogle ScholarPubMed
Larsen, C., Bardet, P.-L., Vincent, J.-P., and Alexandre, C. (2008). Specification and positioning of parasegment grooves in Drosophila. Dev. Biol. 321, 310–318.CrossRefGoogle ScholarPubMed
Larsen, E. and McLaughlin, H.M.G. (1987). The morphogenetic alphabet: lessons for simple-minded genes. BioEssays 7, 130–132.CrossRefGoogle ScholarPubMed
Larson, G. (1991). Unnatural Selections: A Far Side Collection. Andrews & McMeel (Universal Press Syndicate), Kansas City, MO.Google Scholar
Lau, F.H., Xia, F., Kaplan, A., Cerrato, F., Greene, A.K., Taghinia, A., Cowan, C.A., and Labow, B.I. (2012). Expression analysis of macrodactyly identifies pleiotrophin upregulation. PLoS ONE 7 #7, e40423.CrossRefGoogle ScholarPubMed
Lawrence, P.A. (1966). Development and determination of hairs and bristles in the milkweed bug Oncopeltus fasciatus (Lygaeidae, Hemiptera). J. Cell Sci. 1, 475–498.Google Scholar
Lawrence, P.A. (1984). Homoeotic selector genes: a working definition. BioEssays 1, 227–229.CrossRefGoogle Scholar
Lawrence, P.A. and Struhl, G. (1982). Further studies of the engrailed phenotype in Drosophila. EMBO J. 1, 827–833.Google ScholarPubMed
Lazzari, V., Charles, C., Tafforeau, P., Vianey-Liaud, M., Aguilar, J.-P., Jaeger, J.-J., Michaux, J., and Viriot, L. (2008). Mosaic convergence of rodent dentitions. PLoS ONE 3 #10, e3607.CrossRefGoogle ScholarPubMed
Le Douarin, N.M., Creuzet, S., Couly, G., and Dupin, E. (2004). Neural crest plasticity and its limits. Development 131, 4637–4650.CrossRefGoogle ScholarPubMed
Le Gros Clark, W.E. (1945). Deformation patterns in the cerebral cortex. In Essays on Growth and Form (Le Gros Clark, W.E. and Medawar, P.B., eds.). Clarendon Press, Oxford, pp. 1–22.Google Scholar
Le Rouzic, A., Álvarez-Castro, J.M., and Hansen, T.F. (2013). The evolution of canalization and evolvability in stable and fluctuating environments. Evol. Biol. 40, 317–340.CrossRefGoogle Scholar
Lecuit, T. and Cohen, S.M. (1997). Proximal-distal axis formation in the Drosophila leg. Nature 388, 139–145.CrossRefGoogle ScholarPubMed
Lee, H.-G., Kim, Y.-C., Dunning, J.S., and Han, K.-A. (2008). Recurring ethanol exposure induces disinhibited courtship in Drosophila. PLoS ONE #1, e1391.CrossRefGoogle ScholarPubMed
Lee, J. and Tumbar, T. (2012). Hairy tale of signaling in hair follicle development and cycling. Semin. Cell Dev. Biol. 23, 906–916.CrossRefGoogle ScholarPubMed
Lee, M.S.Y. (2009). Hidden support from unpromising data sets strongly unites snakes with anguimorph ‘lizards’. J. Exp. Biol. 22, 1308–1316.Google Scholar
Lee, M.S.Y., Bell, G.L., Jr., and Caldwell, M.W. (1999). The origin of snake feeding. Nature 400, 656–659.CrossRefGoogle Scholar
Lee, M.S.Y., Jago, J.B., García-Bellido, D.C., Edgecombe, G.D., Gehling, J.G., and Paterson, J.R. (2011). Modern optics in exceptionally preserved eyes of early Cambrian arthropods from Australia. Nature 474, 631–634.CrossRefGoogle ScholarPubMed
Lee, P.N., Callaerts, P., de Couet, H.G., and Martindale, M.Q. (2003). Cephalopod Hox genes and the origin of morphological novelties. Nature 424, 1061–1065. [See also Hoving, H.J.T., et al. (2013). First in situ observations of the deep-sea squid Grimalditeuthis bonplandi reveal unique use of tentacles. Proc.R.Soc.Lond.B 280, .]CrossRefGoogle ScholarPubMed
Lee, R.T.H., Thiery, J.P., and Carney, T.J. (2013). Dermal fin rays and scales derive from mesoderm, not neural crest. Curr. Biol. 23, R336–R337.CrossRefGoogle Scholar
Legué, E. and Nicolas, J.-F. (2005). Hair follicle renewal: organization of stem cells in the matrix and the role of stereotyped lineages and behaviors. Development 132, 4143–4154.CrossRefGoogle ScholarPubMed
Leichty, A.R., Pfennig, D.W., Jones, C.D., and Pfennig, K.S. (2012). Relaxed genetic constraint is ancestral to the evolution of phenotypic plasiticity. Integr. Comp. Biol. 52, 16–30.CrossRefGoogle Scholar
Leigh, S.R. (2012). Brain size growth and life history in human evolution. Evol. Biol. 39, 587–599.CrossRefGoogle Scholar
Lelli, K.M., Slattery, M., and Mann, R.S. (2012). Disentangling the many layers of eukaryotic transcriptional regulation. Annu. Rev. Genet. 46, 43–68.CrossRefGoogle ScholarPubMed
Lemaire, P. (2011). Evolutionary crossroads in developmental biology: the tunicates. Development 138, 2143–2152.CrossRefGoogle ScholarPubMed
Lemaire, P., Smith, W.C., and Nishida, H. (2008). Ascidians and the plasticity of the chordate developmental program. Curr. Biol. 18, R620–R631.CrossRefGoogle ScholarPubMed
Lemmon, M.A. and Schlessinger, J. (2010). Cell signaling by receptor tyrosine kinases. Cell 141, 1117–1134.CrossRefGoogle ScholarPubMed
Lemons, D. and McGinnis, W. (2006). Genomic evolution of Hox gene clusters. Science 313, 1918–1922.CrossRefGoogle ScholarPubMed
Lemus, D. (1995). Contributions of heterospecific tissue recombinations to odontogenesis. Int. J. Dev. Biol. 39, 291–297.Google ScholarPubMed
Lennox, J.G. (1991). Darwinian thought experiments: a function for just-so stories. In Thought Experiments in Science and Philosophy (Horowitz, T. and Massey, G.J., eds.). Rowman & Littlefield, Savage, MD, pp. 223–245.Google Scholar
Leonard, J.A., Rohland, N., Glaberman, S., Fleischer, R.C., Caccone, A., and Hofreiter, M. (2005). A rapid loss of stripes: the evolutionary history of the extint quagga. Biol. Lett. 1, 291–295.CrossRefGoogle ScholarPubMed
Lerner, H.R.L., Meyer, M., James, H.F., Hofreiter, M., and Fleischer, R.C. (2011). Multilocus resolution of phylogeny and timescale in the extant adaptive radiation of Hawaiian honeycreepers. Curr. Biol. 21, 1838–1844.CrossRefGoogle ScholarPubMed
Lettice, L.A., Hill, A.E., Devenney, P.S., and Hill, R.E. (2008). Point mutations in a distant sonic hedgehog cis-regulator generate a variable regulatory output responsible for preaxial polydactyly. Hum. Mol. Genet. 17, 978–985.CrossRefGoogle Scholar
Levin, M. (2005). Left-right asymmetry in embryonic development: a comprehensive review. Mechs. Dev. 122, 3–25.CrossRefGoogle ScholarPubMed
Levine, M. (2002). How insects lose their limbs. Nature 415, 848–849.CrossRefGoogle ScholarPubMed
Levine, M. (2010). Transcriptional enhancers in animal development and evolution. Curr. Biol. 20, R754–R763.CrossRefGoogle ScholarPubMed
Levinton, J.S. (1986). Developmental constraints and evolutionary saltations: a discussion and critique. In Genetics, Development, and Evolution (Gustafson, J.P., Stebbins, G.L., and Ayala, F.J., eds.). 17th Stadler Genetics Symposium. Plenum, New York, NY, pp. 253–288.CrossRefGoogle Scholar
Lewin, R. (1982). Adaptation can be a problem for evolutionists. Science 216, 1212–1213.CrossRefGoogle ScholarPubMed
Lewis, D.L., DeCamillis, M., and Bennett, R.L. (2000). Distinct roles of the homeotic genes Ubx and abd-A in beetle embryonic abdominal appendage development. PNAS 97 #9, 4504–4509.CrossRefGoogle ScholarPubMed
Lewis, E.B. (1994). Homeosis: the first 100 years. Trends Genet. 10, 341–343.CrossRefGoogle ScholarPubMed
Lewis, J. (2003). Autoinhibition with transcriptional delay: a simple mechanism for the zebrafish somitogenesis oscillator. Curr. Biol. 13, 1398–1408.CrossRefGoogle ScholarPubMed
Lewis, J.H. and Wolpert, L. (1976). The principle of non-equivalence in development. J. Theor. Biol. 62, 479–490.CrossRefGoogle ScholarPubMed
Lewis, S.M. and Cratsley, C.K. (2008). Flash signal evolution, mate choice, and predation in fireflies. Annu. Rev. Entomol. 53, 293–321.CrossRefGoogle ScholarPubMed
Leysen, H., Roos, G., and Adriaens, D. (2011). Morphological variation in head shape of pipefishes and seahorses in relation to snout length and developmental growth. J. Morph. 272, 1259–1270.CrossRefGoogle ScholarPubMed
Li, B.-W., Zhao, H.-P., and Feng, X.-Q. (2011). Static and dynamic mechanical properties of cattle horns. Materials Sci. Eng. C 31, 179–183.CrossRefGoogle Scholar
Li, C., Wu, X.-C., Rieppel, O., Wang, L.-T., and Zhao, L.-J. (2008). An ancestral turtle from the Late Triassic of southwestern China. Nature 456, 497–501.CrossRefGoogle ScholarPubMed
Li, C., Yang, F., Haines, S., Zhao, H., Wang, W., Xing, X., Sun, H., Chu, W., Lu, X., Liu, L., and McMahon, C. (2012). Stem cells responsible for deer antler regeneration are unable to recapitulate the process of first antler development – revealed through intradermal and subcutaneous tissue transplantation. J. Exp. Zool. (Mol. Dev. Evol.) 314B, 552–570.CrossRefGoogle Scholar
Li, C.-Y., Cha, W., Luder, H.-U., Charles, R.-P., McMahon, M., Mitsiadis, T.A., and Klein, O.D. (2012). E-cadherin regulates the behavior and fate of epithelial stem cells and their progeny in the mouse incisor. Dev. Biol. 366, 357–366.CrossRefGoogle ScholarPubMed
Li, H. and Popadic, A. (2004). Analysis of nubbin expression patterns in insects. Evol. Dev. 6, 310–324.CrossRefGoogle ScholarPubMed
Liang, H.-L., Xu, M., Chuang, Y.-C., and Rushlow, C. (2012). Response to the BMP gradient requires highly combinatorial inputs from multiple patterning systems in the Drosophila embryo. Development 139, 1956–1964.CrossRefGoogle ScholarPubMed
Lichtneckert, R. and Reichert, H. (2005). Insights into the urbilaterian brain: conserved genetic patterning mechanisms in insect and vertebrate brain development. Heredity 94, 465–477.CrossRefGoogle ScholarPubMed
Lieberman, B.S. (2012). Adaptive radiations in the context of macroevolutionary theory: a paleontological perspective. Evol. Biol. 39, 181–191.CrossRefGoogle Scholar
Lim, M.M., Wang, Z., Olazábal, D.E., Ren, X., Terwilliger, E.F., and Young, L.J. (2004). Enhanced partner preference in a promiscuous species by manipulating the expression of a single gene. Nature 429, 754–757.CrossRefGoogle Scholar
Lim, W.A. and Pawson, T. (2010). Phosphotyrosine signaling: evolving a new cellular communication system. Cell 142, 661–667.CrossRefGoogle ScholarPubMed
Lin, B., Wang, S.W., and Masland, R.H. (2004). Retinal ganglion cell type, size, and spacing can be specified independent of homotypic dendritic contacts. Neuron 43, 475–485.CrossRefGoogle ScholarPubMed
Lin, C., Yin, Y., Bell, S.M., Veith, G.M., Chen, H., Huh, S.-H., Ornitz, D.M., and Ma, L. (2013). Delineating a conserved genetic cassette promoting outgrowth of body appendages. PLoS Genet. 9 #1, e1003231.CrossRefGoogle ScholarPubMed
Lin, C.-M., Jiang, T.X., Baker, R.E., Maini, P.K., Widelitz, R.B., and Chuong, C.-M. (2009). Spots and stripes: pleomorphic patterning of stem cells via p-ERK-dependent cell chemotaxis shown by feather morphogenesis and mathematical simulation. Dev. Biol. 334, 369–382.CrossRefGoogle ScholarPubMed
Lin, C.H., Liu, J.H., Osterburg, J.W., and Nicol, J.D. (1982). Fingerprint comparison. I. Similarity of fingerprints. J. Forensic Sci. 27, 290–304.CrossRefGoogle ScholarPubMed
Lin, C.H. and Rankin, C.H. (2012). Alcohol addiction: chronic ethanol leads to cognitive dependence in Drosophila. Curr. Biol. 22, R1043–R1044.CrossRefGoogle ScholarPubMed
Lin, G., Chen, Y., and Slack, J.M.W. (2013). Imparting regenerative capacity to limbs by progenitor cell transplantation. Dev. Cell 24, 41–51.CrossRefGoogle ScholarPubMed
Lin, J.Y. and Fisher, D.E. (2007). Melanocyte biology and skin pigmentation. Nature 445, 843–850.CrossRefGoogle ScholarPubMed
Lindgren, J., Caldwell, M.W., Konishi, T., and Chiappe, L. (2010). Convergent evolution in aquatic tetrapods: insights from an exceptional fossil mosasaur. PLoS ONE 5 #8, e11998.CrossRefGoogle ScholarPubMed
Lindsley, D.L. and Zimm, G.G. (1992). The Genome of Drosophila melanogaster. Academic Press, New York, NY.Google Scholar
Lister, A.M., Edwards, C.J., Nock, D.A.W., Bunce, M., van Pijlen, I.A., Bradley, D.G., Thomas, M.G., and Barnes, I. (2005). The phylogenetic position of the “giant deer” Megalocerous giganteus. Nature 438, 850–853.CrossRefGoogle ScholarPubMed
Litingtung, Y., Dahn, R.D., Li, Y., Fallon, J.F., and Chiang, C. (2002). Shh and Gli3 are dispensable for limb skeleton formation but regulate digit number and identity. Nature 418, 979–983.CrossRefGoogle ScholarPubMed
Little, J.W., Byrd, C.A., and Brower, D.L. (1990). Effect of abx, bx and pbx mutations on expression of homeotic genes in Drosophila larvae. Genetics 124, 899–908.Google ScholarPubMed
Liu, B., Rooker, S.M., and Helms, J.A. (2010). Molecular control of facial morphology. Semin. Cell Dev. Biol. 21, 309–313.CrossRefGoogle ScholarPubMed
Liu, C., Fu, X., Liu, L., Ren, X., Chau, C.K.L., Li, S., Xiang, L., Zeng, H., Chen, G., Tang, L.-H., Lenz, P., Cui, X., Huang, W., Hwa, T., and Huang, J.-D. (2011). Sequential establishment of stripe patterns in an expanding cell population. Science 334, 238–241.CrossRefGoogle Scholar
Lloyd, S. (1996). Complexity simplified. Sci. Am. 274 #5, 104–108.Google Scholar
Locke, M. (2008). Structure of ivory. J. Morph. 269, 423–450.CrossRefGoogle ScholarPubMed
Loehlin, D.W. and Werren, J.H. (2012). Evolution of shape by multiple regulatory changes to a growth gene. Science 335, 943–947.CrossRefGoogle ScholarPubMed
Loftus, R.T., MacHugh, D.E., Bradley, D.G., Sharp, P.M., and Cunningham, P. (1994). Evidence for two independent domestications of cattle. PNAS 91 #7, 2757–2761.CrossRefGoogle ScholarPubMed
Logan, M. (2003). Finger or toe: the molecular basis of limb identity. Development 130, 6401–6410.CrossRefGoogle ScholarPubMed
Logan, M. and Tabin, C.J. (1999). Role of Pitx1 upstream of Tbx4 in specification of hindlimb identity. Science 283, 1736–1739.CrossRefGoogle ScholarPubMed
Logan, M.A. and Vetter, M.L. (2004). Do-it-yourself tiling: dendritic growth in the absence of homotypic contacts. Neuron 43, 439–440.CrossRefGoogle ScholarPubMed
Lohmann, I., McGinnis, N., Bodmer, M., and McGinnis, W. (2002). The Drosophila Hox gene Deformed sculpts head morphology via direct regulation of the apoptosis activator reaper. Cell 110, 457–466.CrossRefGoogle ScholarPubMed
Lohmann, I. and McGinnis, W. (2002). Hox genes: it’s all a matter of context. Curr. Biol. 12, R514–R516.CrossRefGoogle Scholar
Longrich, N.R., Bhullar, B.-A.S., and Gauthier, J.A. (2012). A transitional snake from the Late Cretaceous period of North America. Nature 488, 205–208.CrossRefGoogle ScholarPubMed
Looso, M., Preussner, J., Sousounis, K., Bruckskotten, M., Michel, C.S., Lignelli, E., Reinhardt, R., Hoeffner, S., Krueger, M., Tsonis, P.A., Borchardt, T., and Braun, T. (2013). A de novo assembly of the newt transcriptome combined with proteomic validation identifies new protein families expressed during tissue regeneration. Genome Biol. 14, R16.CrossRefGoogle Scholar
Losos, J.B. and Ricklefs, R.E. (2009). Adaptation and diversification on islands. Nature 457, 830–836.CrossRefGoogle ScholarPubMed
Louchart, A. and Viriot, L. (2011). From snout to beak: the loss of teeth in birds. Trends Ecol. Evol. 26, 663–673.CrossRefGoogle ScholarPubMed
Love, A.C. (2010). Darwin’s “imaginary illustrations”: Creatively teaching evolutionary concepts & the nature of science. Am. Biol. Teacher 72, 82–89.CrossRefGoogle Scholar
Lowe, C.J., Terasaki, M., Wu, M., Freeman, R.M., Runft, L., Kwan, K., Haigo, S., Aronowicz, J., Lander, E., Gruber, C., Smith, M., Kirschner, M., and Gerhart, J. (2006). Dorsoventral patterning in hemichordates: insights into early chordate evolution. PLoS Biol. 4 #9, 1603–1619 (e291).CrossRefGoogle ScholarPubMed
Lowenstein, J.M. and Ryder, O.A. (1985). Immunological systematics of the extinct quagga (Equidae). Experientia 41, 1192–1193.CrossRefGoogle Scholar
Lowery, L.A. and Sive, H. (2009). Totally tubular: the mystery behind function and origin of the brain ventricular system. BioEssays 31, 446–458.CrossRefGoogle ScholarPubMed
Lu, B., LaMora, A., Sun, Y., Welsh, M.J., and Ben-Shahar, Y. (2012). ppk23-dependent chemosensory functions contribute to courtship behavior in Drosophila melanogaster. PLoS Genet. 8 #3, e1002587.CrossRefGoogle ScholarPubMed
Lu, C.-H., Rincón-Limas, D.E., and Botas, J. (2000). Conserved overlapping and reciprocal expression of msh/Msx1 and apterous/Lhx2 in Drosophila and mice. Mechs. Dev. 99, 177–181.CrossRefGoogle ScholarPubMed
Lubarsky, B. and Krasnow, M.A. (2003). Tube morphogenesis: making and shaping biological tubes. Cell 112, 19–28.CrossRefGoogle ScholarPubMed
Luo, Y.-J. and Su, Y.-H. (2012). Opposing Nodal and BMP signals regulate left-right asymmetry in the sea urchin larva. PLoS Biol. 10 #10, e1001402.CrossRefGoogle ScholarPubMed
Lynch, L.J., Robinson, V., and Anderson, C.A. (1973). A scanning electron microscope study of the morphology of rhinoceros horn. Aust. J. Biol. Sci. 26, 395–399.CrossRefGoogle ScholarPubMed
Lynch, M. (2007). The evolution of genetic networks by non-adaptive processes. Nature Rev. Genet. 8, 803–813.CrossRefGoogle ScholarPubMed
Lynch, M. (2007). The frailty of adaptive hypotheses for the origins of organismal complexity. PNAS 104 (Suppl. 1), 8597–8604.CrossRefGoogle ScholarPubMed
Lyons, M.J. and Harrison, L.G. (1992). Stripe selection: an intrinsic property of some pattern-formation models with nonlinear dynamics. Dev. Dynamics 195, 201–215.CrossRefGoogle ScholarPubMed
Lyson, T.R., Bever, G.S., Scheyer, T.M., Hsiang, A.Y., and Gauthier, J.A. (2013). Evolutionary origin of the turtle shell. Curr. Biol. 23, 1113–1119.CrossRefGoogle ScholarPubMed
Lyson, T.R. and Joyce, W.G. (2012). Evolution of the turtle bauplan: the topological relationship of the scapula relative to the ribcage. Biol. Lett. 8, 1028–1031.CrossRefGoogle Scholar
Ma, X., Hou, X., Edgecombe, G.D., and Stausfeld, N.J. (2012). Complex brain and optic lobes in an early Cambrian arthropod. Nature 490, 258–261.CrossRefGoogle Scholar
Ma, Y., Li, A., Faller, W.J., Libertini, S., Fiorito, F., Gillespie, D.A., Sansom, O.J., Yamashiro, S., and Machesky, L.M. (2013). Fascin 1 is transiently expressed in mouse melanoblasts during development and promotes migration and proliferation. Development 140, 2203–2211.CrossRefGoogle ScholarPubMed
Maas, A.-H. (1948). Über die Auslösbarkeit von Temperatur-Modifikationen während der Embryonal-Entwicklung von Drosophila melanogaster Meigen. W. Roux’s Arch. Entw.-Mech. Org. 143, 515–572.CrossRefGoogle Scholar
MacArthur, J.W. and Ford, N. (1937). A Biological Study of the Dionne Quintuplets: An Identical Set. University of Toronto Press, Toronto.Google Scholar
MacDonald, B.T., Semenov, M.V., and He, X. (2007). Snapshot: Wnt/β-catenin signaling. Cell 131, 1204.CrossRefGoogle ScholarPubMed
Macdonald, W.P., Martin, A., and Reed, R.D. (2010). Butterfly wings shaped by a molecular cookie cutter: evolutionary radiation of lepidopteran wing shapes associated with a derived Cut/wingless wing margin boundary system. Evol. Dev. 12, 296–304.CrossRefGoogle Scholar
MacHugh, D.E., Shriver, M.D., Loftus, R.T., Cunningham, P., and Bradley, D.G. (1997). Microsatellite DNA variation and the evolution, domestication and phylogeography of taurine and zebu cattle (Bos taurus and Bos indicus). Genetics 146, 1071–1086.Google Scholar
MacLaurin, J. (2003). The good, the bad and the impossible. Biol. Philos. 18, 463–476.CrossRefGoogle Scholar
Macpherson, E., Jones, W., and Segonzac, M. (2005). A new squat lobster family of Galatheoidea (Crustacea, Decapoda, Anomura) from the hydrothermal vents of the Pacific-Antarctic Ridge. Zoosystema 27, 709–723.Google Scholar
Maderspacher, F. (2012). Colour patterns: channelling Turing. Curr. Biol. 22, R266–R268.CrossRefGoogle ScholarPubMed
Maderspacher, F. and Nüsslein-Volhard, C. (2003). Formation of the adult pigment pattern in zebrafish requires leopard and obelix dependent cell interactions. Development 130, 3447–3457.CrossRefGoogle ScholarPubMed
Maderspacher, F. and Stensmyr, M. (2011). Myrmecomorphomania. Curr. Biol. 21, R291–R293.CrossRefGoogle ScholarPubMed
Madore, B.F. and Freedman, W.L. (1983). Computer simulations of the Belousov-Zhabotinsky reaction. Science 222, 437–438.CrossRefGoogle ScholarPubMed
Madore, B.F. and Freedman, W.L. (1987). Self-organizing structures. Am. Sci. 75, 252–259.Google Scholar
Mahfooz, N., Turchyn, N., Mihajlovic, M., Hrycaj, S., and Popadic, A. (2007). Ubx regulates differential enlargement and diversification of insect hind legs. PLoS ONE 2 #9, e866.CrossRefGoogle ScholarPubMed
Mahler, D.L. and Kearney, M. (2006). The palatal dentition in squamate reptiles: morphology, development, attachment, and replacement. Fieldiana, Zoology New Series, No. 108, 1–61.Google Scholar
Mainguy, G., In der Rieden, P.M.J., Berezikov, E., Woltering, J.M., Plasterk, R.H.A., and Durston, A.J. (2003). A position-dependent organisation of retinoid response elements is conserved in the vertebrate Hox clusters. Trends Genet. 19, 476–479.CrossRefGoogle ScholarPubMed
Maini, P.K. (2003). How the mouse got its stripes. PNAS 100 #17, 9656–9657.CrossRefGoogle ScholarPubMed
Maini, P.K., Baker, R.E., and Chuong, C.-M. (2006). The Turing Model comes of molecular age. Science 314, 1397–1398.CrossRefGoogle ScholarPubMed
Maître, J.-L., Berthoumieux, H., Krens, S.F.G., Salbreux, G., Jülicher, F., Paluch, E., and Heisenberg, C.-P. (2012). Adhesion functions in cell sorting by mechanically coupling the cortices of adhering cells. Science 338, 253–256.CrossRefGoogle ScholarPubMed
Makanya, A.N. and Mortola, J.P. (2007). The structural design of the bat wing web and its possible role in gas exchange. J. Anat. 211, 687–697.CrossRefGoogle ScholarPubMed
Makhijani, K., Kalyani, C., Srividya, T., and Shashidhara, L.S. (2007). Modulation of Decapentaplegic gradient during haltere specification in Drosophila. Dev. Biol. 302, 243–255.CrossRefGoogle ScholarPubMed
Maladen, R.D., Ding, Y., Li, C., and Goldman, D.I. (2009). Undulatory swimming in sand: subsurface locomotion of the sandfish lizard. Science 325 #314–318.CrossRefGoogle ScholarPubMed
Mallarino, R. and Abzhanov, A. (2012). Paths less traveled: Evo-devo approaches to investigating animal morphological evolution. Annu. Rev. Cell Dev. Biol. 28, 743–763.CrossRefGoogle ScholarPubMed
Mallo, M., Vinagre, T., and Carapuço, M. (2009). The road to the vertebral formula. Int. J. Dev. Biol. 53, 1469–1481.CrossRefGoogle ScholarPubMed
Mallo, M., Wellik, D.M., and Deschamps, J. (2010). Hox genes and regional patterning of the vertebrate body plan. Dev. Biol. 344, 7–15. [See also Mallo, M. and Alonso, C.R. (2013). The regulation of Hox gene expression during animal development. Development 140, 3951–3963.]CrossRefGoogle ScholarPubMed
Malmström, T. and Kröger, R.H.H. (2006). Pupil shapes and lens optics in the eyes of terrestrial vertebrates. J. Exp. Biol. 209, 18–25.CrossRefGoogle ScholarPubMed
Mandal, L., Banerjee, U., and Hartenstein, V. (2004). Evidence for a fruit fly hemangioblast and similarities between lymph-gland hematopoiesis in fruit fly and mammal aorta-gonadal-mesonephros mesoderm. Nat. Genet. 36, 1019–1023.CrossRefGoogle ScholarPubMed
Manger, P.R., Hall, L.S., and Pettigrew, J.D. (1998). The development of the external features of the platypus (Ornithorhynchus anatinus). Phil. Trans. R. Soc. Lond. B 353, 1115–1125.CrossRefGoogle Scholar
Mann, R.S. and Carroll, S.B. (2002). Molecular mechanisms of selector gene function and evolution. Curr. Opin. Gen. Dev. 12, 592–600.CrossRefGoogle ScholarPubMed
Manry, D.E. (1985). Birds of fire. Nat. Hist. 94 #1, 38–44.Google Scholar
Manuel, M. and Forêt, S. (2012). Searching for Eve: basal metazoans and the evolution of multicellular complexity. BioEssays 34, 247–251.CrossRefGoogle ScholarPubMed
Mapalad, K.S., Leu, D., and Nieh, J.C. (2008). Bumble bees heat up for high quality pollen. J. Exp. Biol. 211, 2239–2242.CrossRefGoogle ScholarPubMed
Marcil, A., Dumontier, É., Chamberland, M., Camper, S.A., and Drouin, J. (2003). Pitx1 and Pitx2 are required for development of hindlimb buds. Development 130, 45–55.CrossRefGoogle ScholarPubMed
Marcon, L. and Sharpe, J. (2012). Turing patterns in development: what about the horse part? Curr. Opin. Gen. Dev. 22, 578–584.CrossRefGoogle ScholarPubMed
Marcus, J.M. (2001). The development and evolution of crossveins in insect wings. J. Anat. 199, 211–216.CrossRefGoogle ScholarPubMed
Marcus, J.M., Ramos, D.M., and Monteiro, A. (2004). Germline transformation of the butterfly Bicyclus anynana. Proc. Roy. Soc. Lond. B (Suppl.) 271, S263–S265.CrossRefGoogle ScholarPubMed
Marden, J.H. (1995). Flying lessons from a flightless insect. Nat. Hist. 104 #2, 4–8.Google Scholar
Marden, J.H. and Kramer, M.G. (1995). Locomotor performance of insects with rudimentary wings. Nature 377, 332–334.CrossRefGoogle Scholar
Marek, P.E. and Bond, J.E. (2006). Rediscovery of the world’s leggiest animal. Nature 441, 707.CrossRefGoogle ScholarPubMed
Maricich, S.M. and Zoghbi, H.Y. (2006). Getting back to basics. Cell 126, 11–15.CrossRefGoogle ScholarPubMed
Mark, R. (1996). Architecture and evolution. Am. Sci. 84, 383–389.Google Scholar
Markow, T.A. and O’Grady, P.M. (2005). Evolutionary genetics of reproductive behavior in Drosophila: Connecting the dots. Annu. Rev. Genet. 39, 263–291.CrossRefGoogle Scholar
Marlétaz, F., Holland, L.Z., Laudet, V., and Schubert, M. (2006). Retinoic acid signaling and the evolution of chordates. Int. J. Biol. Sci. 2, 38–47.CrossRefGoogle ScholarPubMed
Marquardt, T., Ashery-Padan, R., Andrejewski, N., Scardigli, R., Guillemot, F., and Gruss, P. (2001). Pax6 is required for the multipotent state of retinal progenitor cells. Cell 105, 43–55.CrossRefGoogle ScholarPubMed
Marsh, O.C. (1892). Recent polydactyle horses. Am J. Sci. 43 (3rd Series) or 143 (Combined Series), 339–355.CrossRefGoogle Scholar
Marshall, C.R., Raff, E.C., and Raff, R.A. (1994). Dollo’s law and the death and resurrection of genes. PNAS 91 #25, 12283–12287.CrossRefGoogle ScholarPubMed
Marshall, C.R. and Valentine, J.W. (2010). The importance of preadapted genomes in the origin of the animal bodyplans and the Cambrian Explosion. Evolution 64, 1189–1201.Google ScholarPubMed
Marshall, L.G. (1994). The terror birds of South America. Sci. Am. 270 #2, 90–95.CrossRefGoogle Scholar
Martin, A. and Reed, R.D. (2010). wingless and aristaless2 define a developmental ground plan for moth and butterfly wing pattern evolution. Mol. Biol. Evol. 27, 2864–2878.CrossRefGoogle ScholarPubMed
Martin, G.R., Wilson, K.-J., Wild, J.M., Parsons, S., Kubke, M.F., and Corfield, J. (2007). Kiwi forego vision in the guidance of their nocturnal activities. PLoS ONE 2 #2, e198.CrossRefGoogle ScholarPubMed
Martindale, M.Q. (2005). The evolution of metazoan axial properties. Nature Rev. Genet. 6, 917–927.CrossRefGoogle ScholarPubMed
Martindale, M.Q., Finnerty, J.R., and Henry, J.Q. (2002). The Radiata and the evolutionary origins of the bilaterian body plan. Mol. Phylogenet. Evol. 24, 358–365.CrossRefGoogle ScholarPubMed
Martindale, M.Q. and Hejnol, A. (2009). A developmental perspective: changes in the position of the blastopore during bilaterian evolution. Dev. Cell 17, 162–174.CrossRefGoogle ScholarPubMed
Martindale, M.Q. and Henry, J.Q. (1998). The development of radial and biradial symmetry: the evolution of bilaterality. Am. Zool. 38, 672–684.CrossRefGoogle Scholar
Masel, J. and Trotter, M.V. (2010). Robustness and evolvability. Trends Genet. 26, 406–414.CrossRefGoogle ScholarPubMed
Massare, J.A. (1992). Ancient mariners. Nat. Hist. 101 #9, 48–53.Google Scholar
Masumoto, M., Yaginuma, T., and Niimi, T. (2009). Functional analysis of Ultrabithorax in the silkworm, Bombyx mori, using RNAi. Dev. Genes Evol. 219, 437–444.CrossRefGoogle ScholarPubMed
Mateus, O., Maidment, S.C.R., and Christiansen, N.A. (2009). A new long-necked ‘sauropod-mimic’ stegosaur and the evolution of the plated dinosaurs. Proc. Roy. Soc. Lond. B 276, 1815–1821.CrossRefGoogle ScholarPubMed
Mäthger, L.M., Bell, G.R.R., Kuzirian, A.M., Allen, J.J., and Hanlon, R.T. (2012). How does the blue-ringed octopus (Hapalochlaena lunulata) flash its blue rings?J. Exp. Biol. 215, 3752–3757.CrossRefGoogle ScholarPubMed
Matson, C.K. and Zarkower, D. (2012). Sex and the singular DM domain: insights into sexual regulation, evolution and plasiticity. Nature Rev. Gen. 13, 163–174.CrossRefGoogle Scholar
Matsuda, S. and Shimmi, O. (2012). Directional transport and active retention of Dpp/BMP create wing vein patterns in Drosophila. Dev. Biol. 366, 153–162.CrossRefGoogle ScholarPubMed
Matsuoka, T., Ahlberg, P.E., Kessaris, N., Iannarelli, P., Dennehy, U., Richardson, W.D., McMahon, A.P., and Koentges, G. (2005). Neural crest origins of the neck and shoulder. Nature 436, 347–355.CrossRefGoogle ScholarPubMed
Mattison, C. (2007). The New Encyclopedia of Snakes. Princeton University Press, Princeton, NJ.Google Scholar
Maxmen, A. (2011). A can of worms. Nature 470, 161–162.CrossRefGoogle Scholar
Maxwell, E.E. and Harrison, L.B. (2009). Methods for the analysis of developmental sequence data. Evol. Dev. 11, 109–119.CrossRefGoogle ScholarPubMed
May, R.M. (1978). The evolution of ecological systems. Sci. Am. 239 #3, 160–175.CrossRefGoogle Scholar
Mayer, J.A., Foley, J., De la Cruz, D., Chuong, C.-M., and Widelitz, R. (2008). Conversion of the nipple to hair-bearing epithelia by lowering bone morphogenetic protein pathway activity at the dermal-epidermal interface. Am. J. Pathol. 173, 1339–1348.CrossRefGoogle ScholarPubMed
Maynard Smith, J. (1960). Continuous, quantized and modal variation. Proc. Roy. Soc. Lond. B 152, 397–409.CrossRefGoogle Scholar
Maynard Smith, J. (1968). The counting problem. In Towards a Theoretical Biology. I. Prolegomena (Waddington, C.H., ed.). Aldine, Chicago, IL, pp. 120–124.Google Scholar
Maynard Smith, J., Burian, R., Kauffman, S., Alberch, P., Campbell, J., Goodwin, B., Lande, R., Raup, D., and Wolpert, L. (1985). Developmental constraints and evolution. Q. Rev. Biol. 60, 265–287.CrossRefGoogle Scholar
Mayor, R. and Theveneau, E. (2013). The neural crest. Development 140, 2247–2251.CrossRefGoogle ScholarPubMed
Mayr, E. (1976). Evolution and the Diversity of Life. Harvard University Press, Cambridge, MA.Google Scholar
Mayr, E. (1985). How biology differs from the physical sciences. In Evolution at a Crossroads: The New Biology and the New Philosophy of Science (Depew, D.J. and Weber, B.H., eds.). MIT Press, Cambridge, MA, pp. 43–63.Google Scholar
Mayr, E. (1991). One Long Argument: Charles Darwin and the Genesis of Modern Evolutionary Thought. Harvard University Press, Cambridge, MA.Google Scholar
Mayr, E. (1994). Recapitulation reinterpreted: the somatic program. Q. Rev. Biol. 69, 223–232.CrossRefGoogle Scholar
Mayr, E. and Provine, W. (1980). The Evolutionary Synthesis. Harvard University Press, Cambridge, MA.CrossRefGoogle Scholar
Mazák, J.H., Christiansen, P., and Kitchener, A.C. (2011). Oldest known pantherine skull and evolution of the tiger. PLoS ONE 6 #10, e25483. [Note added in proof: The tiger genome has just been sequenced! See Cho, Y.S., et al. (2013). The tiger genome and comparative analysis with lion and snow leopard genomes. Nature Comm. ].CrossRefGoogle Scholar
McAlpine, J.F. (1981). Morphology and terminology: adults. In Manual of Nearctic Diptera, Vol. 1 (McAlpine, J.F., Peterson, B.V., Shewell, G.E., Teskey, H.J., Vockeroth, J.R., and Wood, D.M., eds.). Canadian Government Publishing Centre, Hull, Quebec, pp. 9–63.Google Scholar
McClure, M. and McCune, A.R. (2003). Evidence for developmental linkage of pigment patterns with body size and shape in Danios (Teleostei: Cyprinidae). Evolution 57, 1863–1875.CrossRefGoogle Scholar
McCollum, M. and Sharpe, P.T. (2001). Evolution and development of teeth. J. Anat. 199, 153–159.CrossRefGoogle ScholarPubMed
McConnell, S.K. and Kaznowski, C.E. (1991). Cell cycle dependence of laminar determination in developing neocortex. Science 254, 282–285.CrossRefGoogle ScholarPubMed
McCue, M.D. (2007). Prey envenomation does not improve digestive performance in western diamondback rattlesnakes (Crotalus atrox). J. Exp. Zool. 307A, 568–577.CrossRefGoogle Scholar
McCue, M.E., Bannasch, D.L., Petersen, J.L., Gurr, J., Bailey, E., Binns, M.M., Distl, O., Guérin, G., Hasegawa, T., Hill, E.W., Leeb, T., Lindgren, G., Penedo, M.C.T., Røed, K.H., Ryder, O.A., Swinburne, J.E., Tozaki, T., Valberg, S.J., Vaudin, M., Lindblad-Toh, K., Wade, C.M., and Mickelson, J.R. (2012). A high density SNP array for the domestic horse and extant perissodactyla: utility for association mapping, genetic diversity, and phylogeny studies. PLoS Genet. 8 #1, e1002451.CrossRefGoogle ScholarPubMed
McCullough, E.L., Weingarden, P.R., and Emlen, D.J. (2012). Costs of elaborate weapons in a rhinocerous beetle: how difficult is it to fly with a big horn? Behav. Ecol. 23, 1042–1048.CrossRefGoogle Scholar
McCune, A.R. and Carlson, R.L. (2004). Twenty ways to lose your bladder: common natural mutants in zebrafish and widespread convergence of swim bladder loss among teleost fishes. Evol. Dev. 6, 246–259.CrossRefGoogle ScholarPubMed
McGhee, G.R., Jr. (1999). Theoretical Morphology: The Concept and Its Applications. Columbia University Press, New York, NY.Google Scholar
McGhee, G.R., Jr. (2011). Convergent Evolution: Limited Forms Most Beautiful. Vienna Series in Theoretical Biology. MIT Press, Cambridge, MA.Google Scholar
McGinnis, W. (1994). A century of homeosis, a decade of homeoboxes. Genetics 137, 607–611.Google ScholarPubMed
McGregor, A.P. (2005). How to get ahead: the origin, evolution and function of bicoid. BioEssays 27, 904–913.CrossRefGoogle ScholarPubMed
McHenry, M.J. (2005). The morphology, behavior, and biomechanics of swimming in ascidian larvae. Can. J. Zool. 83, 62–74.CrossRefGoogle Scholar
McIntyre, D.C., Rakshit, S., Yallowitz, A.R., Loken, L., Jeannotte, L., Capecchi, M.R., and Wellik, D.M. (2007). Hox patterning of the vertebrate rib cage. Development 134, 2981–2989.CrossRefGoogle ScholarPubMed
McKay, D.J., Estella, C., and Mann, R.S. (2009). The origins of the Drosophila leg revealed by the cis-regulatory architecture of the Distalless gene. Development 136, 61–71.CrossRefGoogle ScholarPubMed
McLean, W.H.I. (2008). Combing the genome for the root cause of baldness. Nature Genet. 11, 1270–1271.CrossRefGoogle Scholar
McLeish, T. (2013). Narwhals: Arctic Whales in a Melting World. University of Washington Press, Seattle, WA.Google Scholar
McLellan, J.S., Zheng, X., Hauk, G., Ghirlando, R., Beachy, P.A., and Leahy, D.J. (2008). The mode of Hedgehog binding to Ihog homologues is not conserved across different phyla. Nature 455, 979–983.CrossRefGoogle Scholar
McLennan, D.A. (2008). The concept of co-option: why evolution often looks miraculous. Evo. Edu. Outreach 1, 247–258.CrossRefGoogle Scholar
McMillan, W.O., Monteiro, A., and Kapan, D.D. (2002). Development and evolution on the wing. Trends Ecol. Evol. 17, 125–133.CrossRefGoogle Scholar
McNamara, K.J. (1995). Sexual dimorphism: the role of heterochrony. In Evolutionary Change and Heterochrony (McNamara, K.J., ed.). Wiley, Chichester, pp. 65–89.Google Scholar
McNamara, K.J. (2012). Heterochrony: the evolution of development. Evo. Edu. Outreach 5, 203–218.CrossRefGoogle Scholar
McNamara, K.J. and McKinney, M.L. (2005). Heterochrony, disparity, and macroevolution. Paleobiology 31, 17–26.CrossRefGoogle Scholar
McNulty, K.P. (2010). Apes and tricksters: the evolution and diversification of humans’ closest relatives. Evo. Edu. Outreach 3, 322–332. [See also Stevens, N.J., et al. (2013). Palaeontological evidence for an Oligocene divergence between Old World monkeys and apes. Nature 497, 611–614.]CrossRefGoogle Scholar
McNulty, K.P. (2012). Evolutionary development in Australopithecus africanus. Evol. Biol. 39, 488–498.CrossRefGoogle Scholar
McPherron, A.C., Lawler, A.M., and Lee, S.-J. (1999). Regulation of anterior/posterior patterning of the axial skeleton by growth/differentiation factor 11. Nature Genet. 22, 260–264.CrossRefGoogle ScholarPubMed
McShea, D.W. and Hordijk, W. (2013). Complexity by subtraction. Evol. Biol. 40, in press. .CrossRefGoogle Scholar
Meachen-Samuels, J.A. and van Valkenburgh, B. (2010). Radiographs reveal exceptional forelimb strength in the sabertooth cat, Smilodon fatalis. PLoS ONE 5 #7, e11412.CrossRefGoogle ScholarPubMed
Mead, J.G. (1975). Anatomy of the external nasal passages and facial complex in the Delphinidae (Mammalia: Cetacea). Smithsonian Contrib. Zool. 207, 72 pp.Google Scholar
Medeiros, D.M. and Crump, J.G. (2012). New perspectives on pharyngeal dorsoventral patterning in development and evolution of the vertebrate jaw. Dev. Biol. 371, 121–135.CrossRefGoogle ScholarPubMed
Méhes, E., Mones, E., Németh, V., and Vicsek, T. (2012). Collective motion of cells mediates segregation and pattern formation in co-cultures. PLoS ONE 7 #2, e31711.CrossRefGoogle ScholarPubMed
Mehta, R.S. and Wainwright, P.C. (2007). Raptorial jaws in the throat help moray eels swallow large prey. Nature 449, 79–82.CrossRefGoogle ScholarPubMed
Mehta, R.S., Ward, A.B., Alfaro, M.E., and Wainwright, P.C. (2010). Elongation of the body in eels. Integr. Comp. Biol. 50, 1091–1105.CrossRefGoogle ScholarPubMed
Meier, P., Finch, A., and Evan, G. (2000). Apoptosis in development. Nature 407, 796–801.CrossRefGoogle ScholarPubMed
Meik, J.M. and Pires-daSilva, A. (2009). Evolutionary morphology of the rattlesnake style. BMC Evol. Biol. 9, Article 35 (9 pp.).CrossRefGoogle ScholarPubMed
Meinhardt, H. (1982). Models of Biological Pattern Formation. Academic Press, New York, NY.Google Scholar
Meinhardt, H. (1995). Dynamics of stripe formation. Nature 376, 722–723.CrossRefGoogle Scholar
Meinhardt, H. (2004). Different strategies for midline formation in bilaterians. Nature Rev. Neurosci. 5, 502–510.CrossRefGoogle ScholarPubMed
Meinhardt, H. (2004). Models for the generation of the embryonic body axes: ontogenetic and evolutionary aspects. Curr. Opin. Gen. Dev. 14, 446–454.CrossRefGoogle ScholarPubMed
Meinhardt, H. and Gierer, A. (1974). Applications of a theory of biological pattern formation based on lateral inhibition. J. Cell Sci. 15, 321–346.Google ScholarPubMed
Meinhardt, H. and Gierer, A. (2000). Pattern formation by local self-activation and lateral inhibition. BioEssays 22, 753–760.3.0.CO;2-Z>CrossRefGoogle ScholarPubMed
Mendelson, T.C. and Shaw, K.L. (2012). The (mis)concept of species recognition. Trends Ecol. Evol. 27, 421–427.CrossRefGoogle ScholarPubMed
Menegaz, R.A. and Kirk, E.C. (2009). Septa and processes: convergent evolution of the orbit in haplorhine primates and strigiform birds. J. Hum. Evol. 57, 672–687.CrossRefGoogle ScholarPubMed
Merabet, S. and Hudry, B. (2013). Hox transcriptional specificity despite a single class of cofactors: are flexible interaction modes the key? BioEssays 35, 88–92.CrossRefGoogle ScholarPubMed
Mercader, N., Leonardo, E., Azpiazu, N., Serrano, A., Morata, G., Martínez, C., and Torres, M. (1999). Conserved regulation of proximodistal limb axis development by Meis1/Hth. Nature 402, 425–429.CrossRefGoogle ScholarPubMed
Mercola, M. and Levin, M. (2001). Left-right asymmetry determination in vertebrates. Annu. Rev. Cell Dev. Biol. 17, 779–805.CrossRefGoogle ScholarPubMed
Meredith, R.W., Gatesy, J., Emerling, C.A., York, V.M., and Springer, M.S. (2013). Rod monochromacy and the coevolution of cetacean retinal opsins. PLoS Genet. 9 #4, e1003432.CrossRefGoogle ScholarPubMed
Merino, R., Rodriguez-Leon, J., Macias, D., Gañan, Y., Economides, A.N., and Hurle, J.M. (1999). The BMP antagonist Gremlin regulates outgrowth, chondrogenesis and programmed cell death in the developing limb. Development 126, 5515–5522.Google ScholarPubMed
Merritt, D.J. (2007). The organule concept of insect sense organs: sensory transduction and organule evolution. Adv. Insect Physiol. 33, 192–241.CrossRefGoogle Scholar
Meulemans, D. and Bronner-Fraser, M. (2005). Central role of gene cooption in neural crest evolution. J. Exp. Zool. (Mol. Dev. Evol.) 304B, 298–303.CrossRefGoogle Scholar
Michaelson, J. (1987). Cell selection in development. Biol. Rev. 62, 115–139.CrossRefGoogle ScholarPubMed
Mikhailov, A.T. (2005). Putting evo-devo into focus. Int. J. Dev. Biol. 49, 9–16.CrossRefGoogle Scholar
Mikó, I., Friedrich, F., Yoder, M.J., Hines, H.M., Deitz, L.L., Bertone, M.A., Seltmann, K.C., Wallace, M.S., and Deans, A.R. (2012). On dorsal prothoracic appendages in treehoppers (Hemiptera: Membracidae) and the nature of morphological evidence. PLoS ONE 7 #1, e30137.CrossRefGoogle ScholarPubMed
Milewski, A.V. and Dierenfeld, E.S. (2013). Structural and functional comparison of the proboscis between tapirs and other extant and extinct vertebrates. Integr. Zool. 8, 84–94.CrossRefGoogle ScholarPubMed
Milinkovitch, M.C., Manukyan, L., Debry, A., Di-Poï, N., Martin, S., Singh, D., Lambert, D., and Zwicker, M. (2013). Crocodile head scales are not developmental units but emerge from physical cracking. Science 339, 78–81.CrossRefGoogle Scholar
Miller, G. (2007). Fruit fly fight club. Science 315, 180–182.CrossRefGoogle ScholarPubMed
Miller, G. (2009). On the origin of the nervous system. Science 325, 24–26.CrossRefGoogle Scholar
Miller, G.S., Jr. (1931). Human hair and primate patterning. Smithsonian Misc. Coll. (Publ. No. 3130) 85 #10, 1–13 (plus 5 plates).Google Scholar
Mills, M.G. and Patterson, L.B. (2009). Not just black and white: pigment pattern development and evolution in vertebrates. Semin. Cell Dev. Biol. 20, 72–81.CrossRefGoogle ScholarPubMed
Milner, M.J. and Haynie, J.L. (1979). Fusion of Drosophila eye-antennal imaginal discs during differentiation in vitro. Wilhelm Roux’s Arch. 185, 363–370.CrossRefGoogle ScholarPubMed
Min, M.S., Yang, S.Y., Bonett, R.M., Vieites, D.R., Brandon, R.A., and Wake, D.B. (2005). Discovery of the first Asian plethodontid salamander. Nature 435, 87–90.CrossRefGoogle ScholarPubMed
Minelli, A. (2000). Limbs and tail as evolutionarily diverging duplicates of the main body axis. Evol. Dev. 2, 157–165.CrossRefGoogle ScholarPubMed
Minelli, A. (2002). Homology, limbs, and genitalia. Evol. Dev. 4, 127–132.CrossRefGoogle ScholarPubMed
Minelli, A. (2003). The Development of Animal Form: Ontogeny, Morphology, and Evolution. Cambridge University Press, New York, NY.CrossRefGoogle Scholar
Minelli, A. (2003). The origin and evolution of appendages. Int. J. Dev. Biol. 47, 573–581.Google ScholarPubMed
Minelli, A. (2005). A morphologist’s perspective on terminal growth and segmentation. Evol. Dev. 7, 568–573.CrossRefGoogle ScholarPubMed
Minelli, A. (2009). Forms of Becoming: The Evolutionary Biology of Development. Princeton University Press, Princeton, NJ.Google Scholar
Minelli, A. (2011). A principle of developmental inertia. In Epigenetics: Linking Genotype and Phenotype in Development and Evolution (Hallgrímsson, B. and Hall, B.K., eds.). University of California Press, Berkeley, CA, pp. 116–133.Google Scholar
Minelli, A. and Bortoletto, S. (1988). Myriapod metamerism and arthropod segmentation. Biol. J. Linnean Soc. 33, 323–343.CrossRefGoogle Scholar
Minelli, A., Brena, C., Deflorian, G., Maruzzo, D., and Fusco, G. (2006). From embryo to adult: beyond the conventional periodization of arthropod development. Dev. Genes Evol. 216, 373–383.CrossRefGoogle ScholarPubMed
Minelli, A. and Fusco, G. (2005). Conserved versus innovative features in animal body organization. J. Exp. Zool. (Mol. Dev. Evol.) 304B, 520–525.CrossRefGoogle Scholar
Minguillon, C., Del Buono, J., and Logan, M.P. (2005). Tbx5 and Tbx4 are not sufficient to determine limb-specific morphologies but have common roles in initiating limb outgrowth. Dev. Cell 8, 75–84.CrossRefGoogle Scholar
Minsuk, S.B. and Raff, R.A. (2002). Pattern formation in a pentameral animal: induction of early adult rudiment development in sea urchins. Dev. Biol. 247, 335–350.CrossRefGoogle Scholar
Minsuk, S.B., Turner, F.R., Andrews, M.E., and Raff, R.A. (2009). Axial patterning of the pentaradial adult echinoderm body plan. Dev. Genes Evol. 219, 89–101.CrossRefGoogle ScholarPubMed
Mitchell, G. and Skinner, J.D. (2003). On the origin, evolution and phylogeny of giraffes Giraffa camelopardalis. Trans. Roy. Soc. S. Afr. 58 #1, 51–73.CrossRefGoogle Scholar
Mitchell, G., van Sittert, S.J., and Skinner, J.D. (2009). Sexual selection is not the origin of long necks in giraffes. J. Zool. 278, 281–286.CrossRefGoogle Scholar
Mitchell, J.S., Heckert, A.B., and Sues, H.-D. (2010). Grooves to tubes: evolution of the venom delivery system in a Late Triassic “reptile”. Naturwissenschaften 97, 1117–1121.CrossRefGoogle Scholar
Mitgutsch, C., Richardson, M.K., Jiménez, R., Martin, J.E., Kondrashov, P., de Bakker, M.A.G., and Sánchez-Villagra, M.R. (2011). Circumventing the polydactyly “constraint”: the mole’s ‘thumb’. Biol. Lett. 8, 74–77.CrossRefGoogle Scholar
Mito, T., Shinmyo, Y., Kurita, K., Nakamura, T., Ohuchi, H., and Noji, S. (2011). Ancestral functions of Delta/Notch signaling in the formation of body and leg segments in the cricket Gryllus bimaculatus. Development 138, 3823–3833.CrossRefGoogle ScholarPubMed
Mitrophanov, A.Y. and Groisman, E.A. (2008). Positive feedback in cellular control systems. BioEssays 30, 542–555.CrossRefGoogle ScholarPubMed
Mitsiadis, T.A., Caton, J., and Cobourne, M. (2006). Waking-up Sleeping Beauty: recovery of the ancestral bird odontogenetic program. J. Exp. Zool. (Mol. Dev. Evol.) 306B, 227–233.CrossRefGoogle Scholar
Mitsiadis, T.A., Caton, J., De Bari, C., and Bluteau, G. (2008). The large functional spectrum of the heparin-binding cytokines MK and HB-GAM in continuously growing organs: the rodent incisor as a model. Dev. Biol. 320, 256–266.CrossRefGoogle Scholar
Mitsiadis, T.A., Chéraud, Y., Sharpe, P., and Fontaine-Pérus, J. (2003). Development of teeth in chick embryos after mouse neural crest transplantations. PNAS 100 #1, 6541–6545.CrossRefGoogle ScholarPubMed
Mitsiadis, T.A. and Smith, M.M. (2006). How do genes make teeth to order through development? J. Exp. Zool. (Mol. Dev. Evol.) 306B, 177–182.CrossRefGoogle Scholar
Mitteroecker, P., Gunz, P., Neubauer, S., and Müller, G. (2012). How to explore morphological integration in human evolution and development? Evol. Biol. 39, 536–553.CrossRefGoogle Scholar
Miura, T. and Shiota, K. (2000). TGFb2 acts as an “activator” molecule in reaction-diffusion model and is involved in cell sorting phenomenon in mouse limb micromass culture. Dev. Dynamics 217, 241–249.3.0.CO;2-K>CrossRefGoogle Scholar
Miyazawa, S., Okamoto, M., and Kondo, S. (2010). Blending of animal colour patterns by hybridization. Nature Commun. 1, Article 66 (6 pp.).CrossRefGoogle Scholar
Mizutani, C.M. and Bier, E. (2008). EvoD/Vo: the origins of BMP signalling in the neuro-ectoderm. Nature Rev. Genet. 9, 663–677.CrossRefGoogle Scholar
Mobley, K.B., Small, C.M., and Jones, A.G. (2011). The genetics and genomics of Syngnathidae: pipefishes, seahorses and seadragons. J. Fish 78, 1624–1646.Google ScholarPubMed
Moczek, A.P. (2006). Integrating micro- and macroevolution of development through the study of horned beetles. Heredity 97, 168–178.CrossRefGoogle Scholar
Moczek, A.P. (2007). Developmental capacitance, genetic accommodation, and adaptive evolution. Evol. Dev. 9, 299–305.CrossRefGoogle ScholarPubMed
Moczek, A.P. (2008). On the origins of novelty in development and evolution. BioEssays 30, 432–447.CrossRefGoogle ScholarPubMed
Moczek, A.P. (2011). The origins of novelty. Nature 473, 34–35.CrossRefGoogle ScholarPubMed
Moczek, A.P. (2012). The nature of nurture and the future of evodevo: toward a theory of developmental evolution. Integr. Comp. Biol. 52, 108–119.CrossRefGoogle Scholar
Moczek, A.P., Andrews, J., Kijimoto, T., Yerushalmi, Y., and Rose, D.J. (2007). Emerging model systems in evo-devo: horned beetles and the origins of diversity. Evol. Dev. 9, 323–328.CrossRefGoogle ScholarPubMed
Moczek, A.P. and Rose, D.J. (2009). Differential recruitment of limb patterning genes during development and diversification of beetle horns. PNAS 106 #22, 8992–8997.CrossRefGoogle ScholarPubMed
Moffett, M.W. (2006). Mantids: armed and dangerous. Natl. Geogr. 209 #1, 102–113.Google Scholar
Mohit, P., Bajpai, R., and Shashidhara, L.S. (2003). Regulation of Wingless and Vestigial expression in wing and haltere discs of Drosophila. Development 130, 1537–1547.Google Scholar
Mohit, P., Makhijani, K., Madhavi, M.B., Bharathi, V., Lal, A., Sirdesai, G., Reddy, V.R., Ramesh, P., Kannan, R., Dhawan, J., and Shashidhara, L.S. (2006). Modulation of AP and DV signaling pathways by the homeotic gene Ultrabithorax during haltere development in Drosophila. Dev. Biol. 291, 356–367.CrossRefGoogle ScholarPubMed
Molina, M.D., Neto, A., Maeso, I., Gómez-Skarmeta, J.L., Saló, E., and Cebrià, F. (2011). Noggin and noggin-like genes control dorsoventral axis regeneration in planarians. Curr. Biol. 21, 300–305.CrossRefGoogle ScholarPubMed
Monastersky, R. (2001). Pterosaurs: lords of the ancient skies. Natl. Geogr. 199 #5, 86–105.Google Scholar
Mongera, A. and Nüsslein-Volhard, C. (2013). Scales of fish arise from mesoderm. Curr. Biol. 23, R338–R339.CrossRefGoogle ScholarPubMed
Monod, J. (1974). On chance and necessity. In Studies in the Philosophy of Biology: Reduction and Related Problems (Ayala, F.J. and Dobzhansky, T., eds.). University of California Press, Berkeley, CA, pp. 357–375.CrossRefGoogle Scholar
Montagne, J., Groppe, J., Guillemin, K., Krasnow, M.A., Gehring, W.J., and Affolter, M. (1996). The Drosophila Serum Response Factor gene is required for the formation of intervein tissue of the wing and is allelic to blistered. Development 122, 2589–2597.Google ScholarPubMed
Montavon, T., Le Garrec, J.-F., Kerszberg, M., and Duboule, D. (2008). Modeling Hox gene regulation in digits: reverse collinearity and the molecular origin of thumbness. Genes Dev. 22, 346–359.CrossRefGoogle ScholarPubMed
Montavon, T., Soshnikova, N., Mascrez, B., Joye, E., Thevenet, L., Splinter, E., De Laat, W., Spitz, F., and Duboule, D. (2011). A regulatory archipelago controls Hox genes transcription in digits. Cell 147, 1132–1145.CrossRefGoogle ScholarPubMed
Monteiro, A. (2008). Alternative models for the evolution of eyespots and of serial homology on lepidopteran wings. BioEssays 30, 358–366.CrossRefGoogle ScholarPubMed
Monteiro, A. (2011). Gene regulatory networks reused to build novel traits. BioEssays 34, 181–186.CrossRefGoogle Scholar
Monteiro, A., Brakefield, P.M., and French, V. (1997). Butterfly eyespots: the genetics and development of the color rings. Evolution 51, 1207–1216.CrossRefGoogle ScholarPubMed
Monteiro, A., Brakefield, P.M., and French, V. (1997). The genetics and development of an eyespot pattern in the butterfly Bicyclus anynana: response to selection for eyespot shape. Genetics 146, 287–294.Google ScholarPubMed
Monteiro, A., Brakefield, P.M., and French, V. (1997). The relationship between eyespot shape and wing shape in the butterfly Bicyclus anynana: A genetic and morphometrical approach. J. Evol. Biol. 10, 787–802.CrossRefGoogle Scholar
Monteiro, A., French, V., Smit, G., Brakefield, P.M., and Metz, J.A.J. (2001). Butterfly eyespot patterns: evidence for specification by a morphogen diffusion gradient. Acta Biotheor. 49, 77–88.CrossRefGoogle ScholarPubMed
Monteiro, A., Glaser, G., Stockslager, S., Glansdorp, N., and Ramos, D. (2006). Comparative insights into questions of lepidopteran wing pattern homology. BMC Dev. Biol. 6, Article 52 (13 pp.).CrossRefGoogle ScholarPubMed
Monteiro, A. and Podlaha, O. (2009). Wings, horns, and butterfly eyespots: How do complex traits evolve? PLoS Biol. 7 #2, e1000037.CrossRefGoogle ScholarPubMed
Monteiro, A., Prijs, J., Bax, M., Hakkaart, T., and Brakefield, P.M. (2003). Mutants highlight the modular control of butterfly eyespot patterns. Evol. Dev. 5, 180–187.CrossRefGoogle Scholar
Monteiro, A. and Prudic, K.L. (2010). Multiple approaches to study color pattern evolution in butterflies. Trends Evol. Biol. 2, Article e2 (7 pp.).CrossRefGoogle Scholar
Monteiro, A.F., Brakefield, P.M., and French, V. (1994). The evolutionary genetics and developmental basis of wing pattern variation in the butterfly Bicyclus anynana. Evolution 48, 1147–1157.CrossRefGoogle ScholarPubMed
Mooallem, J. (2013). Wild Ones: A Sometimes Dismaying, Weirdly Reassuring Story About Looking at People Looking at Animals in America. Penguin, New York, NY.Google Scholar
Moore, J.A. (1993). Science as a Way of Knowing. The Foundations of Modern Biology. Harvard University Press, Cambridge, MA.Google Scholar
Mooseker, M.S. and Cheney, R.E. (1995). Unconventional myosins. Annu. Rev. Cell Dev. Biol. 11, 633–675.CrossRefGoogle ScholarPubMed
Morata, G. (1975). Analysis of gene expression during development in the homeotic mutant Contrabithorax of Drosophila melanogaster. J. Embryol. Exp. Morph. 34, 19–31.Google ScholarPubMed
Morata, G., Macías, A., Urquía, N., and González-Reyes, A. (1990). Homoeotic genes. Semin. Cell Biol. 1, 219–227.Google ScholarPubMed
Morimura, S., Maves, L., and Hoffmann, F.M. (1996). decapentaplegic overexpression affects Drosophila wing and leg imaginal disc development and wingless expression. Dev. Biol. 177, 136–151.CrossRefGoogle ScholarPubMed
Morin-Kensicki, E.M., Melancon, E., and Eisen, J.S. (2002). Segmental relationship between somites and vertebral column in zebrafish. Development 129, 3851–3860.Google ScholarPubMed
Morris, V.B. (2012). Early development of coelomic structures in an echinoderm larva and a similarity with coelomic structures in a chordate embryo. Dev. Genes Evol. 222, 313–323.CrossRefGoogle Scholar
Moss, E.G. (2007). Heterochronic genes and the nature of developmental time. Curr. Biol. 17, R425–R434.CrossRefGoogle ScholarPubMed
Motani, R. (2000). Rulers of the Jurassic seas. Sci. Am. 283 #6, 52–59.CrossRefGoogle ScholarPubMed
Motani, R. (2009). The evolution of marine reptiles. Evo. Edu. Outreach 2, 224–235.CrossRefGoogle Scholar
Motani, R., Rothschild, B.M., and Wahl, W., Jr. (1999). Large eyeballs in diving ichthyosaurs. Nature 402, 747. [See also Schmitz, L., et al. (2013). Allometry indicates giant eyes of giant squid are not exceptional. BMC Evol. Biol. 13, Article 45 (9 pp.).]CrossRefGoogle Scholar
Mount, J.G., Muzylak, M., Allen, S., Althnaian, T., McGonnell, I.M., and Price, J.S. (2006). Evidence that the canonical Wnt signalling pathway regulates deer antler regeneration. Dev. Dynamics 235, 1390–1399.CrossRefGoogle ScholarPubMed
Moussian, B. and Uv, A.E. (2005). An ancient control of epithelial barrier formation and wound healing. BioEssays 27, 987–990.CrossRefGoogle ScholarPubMed
Moustakas, A. and Heldin, C.-H. (2009). The regulation of TGFβ signal transduction. Development 136, 3699–3714.CrossRefGoogle ScholarPubMed
Muchhala, N. (2006). Nectar bat stows huge tongue in its rib cage. Nature 444, 701–702.CrossRefGoogle ScholarPubMed
Muchhala, N. and Thomson, J.D. (2009). Going to great lengths: selection for long corolla tubes in an extremely specialized bat–flower mutualism. Proc. R. Soc. Lond. B 276, 2147–2152.CrossRefGoogle Scholar
Mueller, T. (2010). Valley of the whales. Natl. Geogr. 218 #2, 118–137.Google Scholar
Müller, G.B. (1990). Developmental mechanisms at the origin of morphological novelty: a side-effect hypothesis. In Evolutionary Innovations (Nitecki, M.H., ed.). University of Chicago Press, Chicago, IL, pp. 99–130.Google Scholar
Müller, G.B. (2007). Evo-devo: extending the evolutionary synthesis. Nature Rev. Genet. 8, 943–949.CrossRefGoogle ScholarPubMed
Müller, G.B. (2008). Evo-devo as a discipline. In Evolving Pathways: Key Themes in Evolutionary Developmental Biology (Minelli, A. and Fusco, G., eds.). Cambridge University Press, New York, NY, pp. 5–30.CrossRefGoogle Scholar
Müller, G.B. and Newman, S.A., eds. (2003). Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology. MIT Press, Cambridge, MA.
Müller, G.B. and Newman, S.A. (2005). The innovation triad: an EvoDevo agenda. J. Exp. Zool. (Mol. Dev. Evol.) 304B, 487–503.CrossRefGoogle Scholar
Müller, G.B. and Wagner, G.P. (1991). Novelty in evolution: restructuring the concept. Annu. Rev. Ecol. Syst. 22, 229–256.CrossRefGoogle Scholar
Muller, H.J. (1939). Reversibility in evolution considered from the standpoint of genetics. Biol. Rev. 14, 261–280.CrossRefGoogle Scholar
Müller, J., Hipsley, C.A., Head, J.J., Kardjilov, N., Hilger, A., Wuttke, M., and Reisz, R.R. (2011). Eocene lizard from Germany reveals amphisbaenian origins. Nature 473, 364–367.CrossRefGoogle ScholarPubMed
Müller, J., Scheyer, T.M., Head, J.J., Barrett, P.M., Werneburg, I., Ericson, P.G.P., Pol, D., and Sánchez-Villagra, M.R. (2010). Homeotic effects, somitogenesis and the evolution of vertebral numbers in recent and fossil amniotes. PNAS 107 #5, 2118–2123.CrossRefGoogle ScholarPubMed
Müller, P., Rogers, K.W., Yu, S.R., Brand, M., and Schier, A.F. (2013). Morphogen transport. Development 140, 1621–1638.CrossRefGoogle Scholar
Müller, P. and Schier, A.F. (2011). Extracellular movement of signaling molecules. Dev. Cell 21, 145–158.CrossRefGoogle ScholarPubMed
Muneoka, K., Han, M., and Gardiner, D.M. (2008). Regrowing human limbs. Sci. Am. 298 #4, 56–63.CrossRefGoogle ScholarPubMed
Muñoz Descalzo, S. and Martinez Arias, A. (2012). The structure of Wntch signalling and the resolution of transition states in development. Semin. Cell Dev. Biol. 23, 443–449.CrossRefGoogle ScholarPubMed
Murata, Y., Tamura, M., Aita, Y., Fujimura, K., Murakami, Y., Okabe, M., Okada, N., and Tanaka, M. (2010). Allometric growth of the trunk leads to the rostral shift of the pelvic fin in teleost fishes. Dev. Biol. 347, 236–245.CrossRefGoogle ScholarPubMed
Murawala, P., Tanaka, E.M., and Currie, J.D. (2012). Regeneration: the ultimate example of wound healing. Semin. Cell Dev. Biol. 23, 954–962.CrossRefGoogle ScholarPubMed
Murray, J.D. (1981). On pattern formation mechanisms for lepidopteran wing patterns and mammalian coat markings. Phil. Trans. Roy. Soc. Lond. B 295, 473–496.CrossRefGoogle ScholarPubMed
Murray, J.D. (1981). A pre-pattern formation mechanism for animal coat markings. J. Theor. Biol. 88, 161–199.CrossRefGoogle Scholar
Murray, J.D. (1988). How the leopard gets its spots. Sci. Am. 258 #3, 80–87.CrossRefGoogle Scholar
Murray, J.D. (1989). Mathematical Biology. Springer-Verlag, Berlin.CrossRefGoogle Scholar
Murray, J.D. (1990). Turing’s theory of morphogenesis: its influence on modelling biological pattern and form. Bull. Math. Biol. 52, 119–152.CrossRefGoogle Scholar
Murray, J.D. (2012). Vignettes from the field of mathematical biology: the application of mathematics to biology and medicine. Interface Focus 2, 397–406.CrossRefGoogle ScholarPubMed
Murray, J.D., Deeming, D.C., and Ferguson, M.W.J. (1990). Size-dependent pigmentation-pattern formation in embryos of Alligator mississippiensis: time of initiation of pattern generation mechanism. Proc. Roy. Soc. Lond. B 239, 279–293.CrossRefGoogle ScholarPubMed
Murray, J.D. and Maini, P.K. (1989). Pattern formation mechanisms–a comparison of reaction-diffusion and mechanochemical models. In Cell to Cell Signalling: From Experiments to Theoretical Models (Goldbeter, A., ed.). Academic Press, New York, NY, pp. 159–170.CrossRefGoogle Scholar
Murray, J.D. and Myerscough, M.R. (1991). Pigmentation pattern formation on snakes. J. Theor. Biol. 149, 339–360.CrossRefGoogle ScholarPubMed
Murren, C.J. (2012). The integrated phenotype. Integr. Comp. Biol. 52, 64–76.CrossRefGoogle ScholarPubMed
Nacu, E. and Tanaka, E.M. (2011). Limb regeneration: a new development? Annu. Rev. Cell Dev. Biol. 27, 409–440.CrossRefGoogle ScholarPubMed
Nadrowski, B., Albert, J.T., and Göpfert, M.C. (2008). Transducer-based force generation explains active process in Drosophila hearing. Curr. Biol. 18, 1365–1372.CrossRefGoogle ScholarPubMed
Nagaraj, R. and Adler, P.N. (2012). Dusky-like functions as a Rab11 effector for the deposition of cuticle during Drosophila bristle development. Development 139, 906–916.CrossRefGoogle ScholarPubMed
Nagashima, H., Kuraku, S., Uchida, K., Kawashima-Ohya, Y., Narita, Y., and Kuratani, S. (2012). Body plan of turtles: an anatomical, developmental and evolutionary perspective. Anat. Sci. Int. 87, 1–13.CrossRefGoogle ScholarPubMed
Nagorcka, B.N. (1989). Wavelike isomorphic prepatterns in development. J. Theor. Biol. 137, 127–162.CrossRefGoogle ScholarPubMed
Nagorcka, B.N. and Mooney, J.R. (1992). From stripes to spots: prepatterns which can be produced in the skin by a reaction-diffusion system. Math. Med. Biol. 9, 249–267.CrossRefGoogle ScholarPubMed
Nagy, L.M. and Grbic, M. (1999). Cell lineages in larval development and evolution of holometabolous insects. In The Origin and Evolution of Larval Forms (Hall, B.K. and Wake, M.H., eds.). Academic Press, New York, NY, pp. 275–300.CrossRefGoogle Scholar
Nahmad, M., Glass, L., and Abouheif, E. (2008). The dynamics of developmental system drift in the gene network underlying wing polyphenism in ants: a mathematical model. Evol. Dev. 10, 360–374.CrossRefGoogle ScholarPubMed
Nahmad, M. and Lander, A.D. (2011). Spatiotemporal mechanisms of morphgen gradient interpretation. Curr. Opin. Gen. Dev. 21, 726–731.CrossRefGoogle ScholarPubMed
Nakamasu, A., Takahashi, G., Kanbe, A., and Kondo, S. (2009). Interactions between zebrafish pigment cells responsible for the generation of Turing patterns. PNAS 106 #21, 8429–8434.CrossRefGoogle ScholarPubMed
Nakamura, T. and Hamada, H. (2012). Left-right patterning: conserved and divergent mechanisms. Development 139, 3257–3262.CrossRefGoogle ScholarPubMed
Naples, V.L. (1999). Morphology, evolution and function of feeding in the giant anteater (Myrmecophaga tridactyla). J. Zool. Lond. 249, 19–41.CrossRefGoogle Scholar
Nardi, F., Spinsanti, G., Boore, J.L., Carapelli, A., Dallai, R., and Frati, F. (2003). Hexapod origins: monophyletic or paraphyletic? Science 299, 1887–1889.CrossRefGoogle ScholarPubMed
Nardi, J.B. (1994). Rearrangement of epithelial cell types in an insect wing monolayer is accompanied by differential expression of a cell surface protein. Dev. Dynamics 199, 315–325.CrossRefGoogle Scholar
Nardi, J.B. and Magee-Adams, S.M. (1986). Formation of scale spacing patterns in a moth wing. I. Epithelial feet may mediate cell rearrangement. Dev. Biol. 116, 265–277.CrossRefGoogle Scholar
Narita, Y. and Kuratani, S. (2005). Evolution of the vertebral formulae in mammals: a perspective on developmental constraints. J. Exp. Zool. (Mol. Dev. Evol.) 304B, 91–106.CrossRefGoogle Scholar
Natori, K., Tajiri, R., Furukawa, S., and Kojima, T. (2012). Progressive tarsal patterning in the Drosophila by temporally dynamic regulation of transcription factor genes. Dev. Biol. 361, 450–462.CrossRefGoogle ScholarPubMed
Needham, J. (1933). On the dissociability of the fundamental processes in ontogenesis. Biol. Rev. 8, 180–223.CrossRefGoogle Scholar
Negre, B. and Simpson, P. (2009). Evolution of the achaete-scute complex in insects: convergent duplication of proneural genes. Trends Genet. 25, 147–152.CrossRefGoogle ScholarPubMed
Nelson, C. (2004). Selector genes and the genetic control of developmental modules. In Modularity in Development and Evolution (Schlosser, G. and Wagner, G.P., eds.). University of Chicago Press, Chicago, IL, pp. 17–33.Google Scholar
Nelson, C.E. (2012). Why don’t undergraduates really “get” evolution? In Evolution Challenges: Integrating Research and Practice in Teaching and Learning about Evolution (Rosengren, K.S., Evans, E.M., Brem, S., and Sinatra, G., eds.). Oxford University Press, Oxford, pp. 311–347.CrossRefGoogle Scholar
Nelson, C.E., Morgan, B.A., Burke, A.C., Laufer, E., DiMambro, E., Murtaugh, L.C., Gonzales, E., Tessarollo, L., Parada, L.F., and Tabin, C. (1996). Analysis of Hox gene expression in the chick limb bud. Development 122, 1449–1466.Google ScholarPubMed
Nelson, J. and Gemmell, R. (2004). Implications of marsupial births for an understanding of behavioural development. Int. J. Compar. Psychol. 17, 53–70.Google Scholar
Nelson, J.E. and Gemmell, R.T. (2003). Birth in the northern quoll, Dasyurus hallucatus (Marsupialia: Dasyuridae). Aust. J. Zool. 51, 187–198.CrossRefGoogle Scholar
Nelson, W.J. and Nusse, R. (2004). Convergence of Wnt, b-catenin, and cadherin pathways. Science 303, 1483–1487.CrossRefGoogle Scholar
Neubauer, S. and Hublin, J.-J. (2012). The evolution of human brain development. Evol. Biol. 39, 568–586.CrossRefGoogle Scholar
Neumann, C.J. and Cohen, S.M. (1997). Long-range action of Wingless organizes the dorsal–ventral axis of the Drosophila wing. Development 124, 871–880.Google ScholarPubMed
Neuweiler, G. (2000). The Biology of Bats. Oxford University Press, New York, NY.Google Scholar
Newman, C. (1997). Cats: nature’s masterwork. Natl. Geogr. 191 #6, 54–85.Google Scholar
Newman, C., Buesching, C.D., and Wolff, J.O. (2005). The function of facial masks in “midguild” carnivores. Oikos 108, 623–633.CrossRefGoogle Scholar
Newman, S.A. and Comper, W.D. (1990). “Generic” physical mechanisms of morphogeneiss and pattern formation. Development 110, 1–18.Google Scholar
Newman, S.A., Forgacs, G., and Müller, G.B. (2006). Before programs: The physical origination of multicellular forms. Int. J. Dev. Biol. 50, 289–299.CrossRefGoogle ScholarPubMed
Newman, S.A. and Frisch, H.L. (1979). Dynamics of skeletal pattern formation in developing chick limb. Science 205, 662–668.CrossRefGoogle ScholarPubMed
Newman, S.A. and Müller, G.B. (2000). Epigenetic mechanisms of character origination. J. Exp. Zool. (Mol. Dev. Evol.) 288, 304–317.3.0.CO;2-G>CrossRefGoogle ScholarPubMed
Newton, A. (1896). A Dictionary of Birds. Adam & Charles Black, London.Google Scholar
Ng, C.S. and Kopp, A. (2008). Sex combs are important for male mating success in Drosophila melanogaster. Behav. Genet. 38, 195–201.CrossRefGoogle ScholarPubMed
Ng, M., Diaz-Benjumea, F.J., Vincent, J.P., Wu, J., and Cohen, S.M. (1996). Specification of the wing by localized expression of wingless protein. Nature 381, 316–318.CrossRefGoogle ScholarPubMed
Nicklen, P. (2007). Arctic ivory: hunting the narwhal. Natl. Geogr. 212 #2, 110–129.Google Scholar
Nicolson, S.W. and Human, H. (2008). Bees get a head start on honey production. Biol. Lett. 4, 299–301.CrossRefGoogle ScholarPubMed
Nie, J., Mahato, S., Mustill, W., Tipping, C., Bhattacharya, S.S., and Zelhof, A.C. (2012). Cross species analysis of Prominin reveals a conserved cellular role in invertebrate and vertebrate photoreceptor cells. Dev. Biol. 371, 312–320.CrossRefGoogle ScholarPubMed
Niehrs, C. (2010). On growth and form: a Cartesian coordinate system of Wnt and BMP signaling specifies bilaterian body axes. Development 137, 845–857.CrossRefGoogle ScholarPubMed
Niehuis, O., Hartig, G., Grath, S., Pohl, H., Lehmann, J., Tafer, H., Donath, A., Krauss, V., Eisenhardt, C., Hertel, J., Petersen, M., Mayer, C., Meusemann, K., Peters, R.S., Stadler, P.F., Beutel, R.G., Bornberg-Bauer, E., McKenna, D.D., and Misof, B. (2012). Genomic and morphological evidence converge to resolve the enigma of Strepsiptera. Curr. Biol. 22, 1309–1313.CrossRefGoogle ScholarPubMed
Nielsen, C. (2001). Animal Evolution: Interrelationships of the Living Phyla, 2nd edn. Oxford University Press, New York, NY.Google Scholar
Nielsen, C. (2008). Six major steps in animal evolution: are we derived sponge larvae? Evol. Dev. 10, 241–257.CrossRefGoogle ScholarPubMed
Niimura, Y. and Nei, M. (2007). Extensive gains and losses of olfactory receptor genes in mammalian evolution. PLoS ONE 8, e708.Google Scholar
Nijhout, H.F. (1978). Wing pattern formation in Lepidoptera: a model. J. Exp. Zool. 206, 119–136.CrossRefGoogle Scholar
Nijhout, H.F. (1980). Ontogeny of the color pattern on the wings of Precis coenia (Lepidoptera: Nymphalidae). Dev. Biol. 80, 275–288.CrossRefGoogle Scholar
Nijhout, H.F. (1980). Pattern formation on lepidopteran wings: determination of an eyespot. Dev. Biol. 80, 267–274.CrossRefGoogle ScholarPubMed
Nijhout, H.F. (1981). The color patterns of butterflies and moths. Sci. Am. 245 #5, 140–151.CrossRefGoogle Scholar
Nijhout, H.F. (1985). Independent development of homologous pattern elements in the wing patterns of butterflies. Dev. Biol. 108, 146–151.CrossRefGoogle ScholarPubMed
Nijhout, H.F. (1986). Pattern and pattern diversity on lepidopteran wings. BioScience 36, 527–533.CrossRefGoogle Scholar
Nijhout, H.F. (1991). The Development and Evolution of Butterfly Wing Patterns. Smithsonian Press, Washington, DC.Google Scholar
Nijhout, H.F. (1996). Focus on butterfly eyespot development. Nature 384, 209–210.CrossRefGoogle ScholarPubMed
Nijhout, H.F. (2001). Elements of butterfly wing patterns. J. Exp. Zool. (Mol. Dev. Evol.) 291, 213–225.CrossRefGoogle ScholarPubMed
Nijhout, H.F. (2010). Molecular and physiological basis of colour pattern formation. Adv. Insect Physiol. 38, 219–265.CrossRefGoogle Scholar
Nijhout, H.F. and Emlen, D.J. (1998). Competition among body parts in the development and evolution of insect morphology. PNAS 95 #7, 3685–3689.CrossRefGoogle ScholarPubMed
Nijhout, H.F. and German, R.Z. (2012). Developmental causes of allometry: new models and implications for phenotypic plasticity and evolution. Integr. Comp. Biol. 52, 43–52.CrossRefGoogle ScholarPubMed
Nilson, G. and Andren, C. (1988). A new subspecies of the subalpine meadow viper, Vipera ursinii Bonaparte (Reptilia, Viperidae), from Greece. Zool. Scripta 17, 311–314.CrossRefGoogle Scholar
Nilsson, D.-E. (2004). Eye evolution: a question of genetic promiscuity. Curr. Opin. Neurobiol. 14, 407–414.CrossRefGoogle ScholarPubMed
Nilsson, D.-E. and Arendt, D. (2009). Eye evolution: the blurry beginning. Curr. Biol. 18, R1096–R1098.CrossRefGoogle Scholar
Nilsson, D.-E., Warrant, E.J., Johnsen, S., Hanlon, R., and Shashar, N. (2012). A unique advantage for giant eyes in giant squid. Curr. Biol. 22, 683–688.CrossRefGoogle ScholarPubMed
Niskanen, M. and Mappes, J. (2005). Significance of the dorsal zigzag pattern of Vipera latastei gaditana against avian predators. J. Anim. Ecol. 74, 1091–1101.CrossRefGoogle Scholar
Nitecki, M.H. (1990). The plurality of evolutionary innovations. In Evolutionary Innovations (Nitecki, M.H., ed.). University of Chicago Press, Chicago, IL, pp. 3–18.Google Scholar
Nitzan, E., Krispin, S., Pfaltzgraff, E.R., Klar, A., Labosky, P.A., and Kalcheim, C. (2013). A dynamic code of dorsal neural tube genes regulates the segregation between neurogenic and melanogenic neural crest cells. Development 140, 2269–2279.CrossRefGoogle ScholarPubMed
Niven, J.E. and Chittka, L. (2010). Reuse of identified neurons in multiple neural circuits. Behav. Brain Sci. 33, 285.CrossRefGoogle Scholar
Niwa, N., Akimoto-Kato, A., Niimi, T., Tojo, K., Machida, R., and Hayashi, S. (2010). Evolutionary origin of the insect wing via integration of two developmental modules. Evol. Dev. 12, 168–176.CrossRefGoogle ScholarPubMed
Niwa, N., Inoue, Y., Nozawa, A., Saito, M., Misumi, Y., Ohuchi, H., Yoshioka, H., and Noji, S. (2000). Correlation of diversity of leg morphology in Gryllus bimaculatus (cricket) with divergence in dpp expression pattern during leg development. Development 127, 4373–4381.Google ScholarPubMed
Nolte, M.J., Hockman, D., Cretekos, C.J., Behringer, R.R., and Rasweiler, J.J., IV (2009). Embryonic staging system for the black mastiff bat, Molossus rufus (Molossidae), correlated with structure-function relationships in the adult. Anat. Rec. 292, 155–168.CrossRefGoogle Scholar
Noordermeer, D. and Duboule, D. (2013). Chromatin architectures and Hox gene collinearity. Curr. Top. Dev. Biol. 104, 113–148.CrossRefGoogle ScholarPubMed
Noordermeer, D., Leleu, M., Splinter, E., Rougemont, J., De Laat, W., and Duboule, D. (2011). The dynamic architecture of Hox gene clusters. Science 334, 222–225.CrossRefGoogle ScholarPubMed
Norell, M., Chiappe, L., and Clark, J. (1993). New limb on the avian family tree. Nat. Hist. 102 #9, 38–42.Google Scholar
Northcutt, R.G. (2012). Evolution of centralized nervous systems: two schools of evolutionary thought. PNAS 109 (Suppl. 1), 10626–10633.CrossRefGoogle ScholarPubMed
Nottebohm, E., Usui, A., Therianos, S., Kimura, K.-i., Dambly-Chaudière, C., and Ghysen, A. (1994). The gene poxn controls different steps of the formation of chemosensory organs in Drosophila. Neuron 12, 25–34.CrossRefGoogle ScholarPubMed
Nowak, R.M. (1999). Walker’s Mammals of the World, 6th edn. Johns Hopkins University Press, Baltimore, MD.Google Scholar
Nowicki, J.L. and Burke, A.C. (2000). Hox genes and morphological identity: axial versus lateral patterning in the vertebrate mesoderm. Development 127, 4265–4275.Google ScholarPubMed
Nussbaumer, U., Halder, G., Groppe, J., Affolter, M., and Montagne, J. (2000). Expression of the blistered/DSRF gene is controlled by different morphogens during Drosophila trachea and wing development. Mechs. Dev. 96, 27–36.CrossRefGoogle ScholarPubMed
Nusse, R. (2003). Wnts and Hedgehogs: lipid-modified proteins and similarities in signaling mechanisms at the cell surface. Development 130, 5297–5305.CrossRefGoogle ScholarPubMed
Nüsslein-Volhard, C. (1996). Gradients that organize embryo development. Sci. Am. 275 #2, 54–61.CrossRefGoogle ScholarPubMed
Nweeia, M.T., Eichmiller, F.C., Hauschka, P.V., Tyler, E., Mead, J.G., Potter, C.W., Angnatsiak, D.P., Richard, P.R., Orr, J.R., and Black, S.R. (2012). Vestigial tooth anatomy and tusk nomenclature for Monodon monoceros. Anat. Rec. 295, 1006–1016.CrossRefGoogle ScholarPubMed
Nweeia, M.T., Eichmiller, F.C., Nutarak, C., Eidelman, N., Giuseppetti, A.A., Quinn, J., Mead, J.G., K’issuk, K., Hauschka, P.V., Tyler, E.M., Potter, C., Orr, J.R., Avike, R., Nielsen, P., and Angnatsiak, D. (2009). Considerations of anatomy, morphology, evolution, and function for narwhal dentition. In Smithsonian at the Poles: Contributions to International Polar Year Science (Krupnik, I., Lang, M.A., and Miller, S.E., eds.). Smithsonian Institution Scholarly Press, Washington, DC, pp. 223–240.CrossRefGoogle Scholar
Nyakatura, J.A., Petrovitch, A., and Fischer, M.S. (2010). Limb kinematics during locomotion in the two-toed sloth (Choloepus didactylus, Xenarthra) and its implications for the evolution of the sloth locomotor apparatus. Zoology 113, 221–234.CrossRefGoogle ScholarPubMed
Nyholt, D.R., Gillespie, N.A., Heath, A.C., and Martin, N.G. (2003). Genetic basis of male pattern baldness. J. Invest. Dermatol. 121, 1561–1564.Google ScholarPubMed
O’Connor, J.K., Zhang, Y., Chiappe, L.M., Meng, Q., Quanguo, L., and Di, L. (2013). A new enantiornithine from the Yixian Formation with the first recognized avian enamel specialization. J. Vert. Paleo. 33, 1–12.CrossRefGoogle Scholar
O’Dor, R., Stewart, J., Gilly, W., Payne, J., Borges, T.C., and Thys, T. (2012). Squid rocket science: how squid launch into air. Deep-Sea Res. II, in press. . [See also O’Dor, R.K. (2013). How squid swim and fly. Can. J. Zool. 91, 413–419.]Google Scholar
O’Leary, M.A., Bloch, J.I., Flynn, J.J., Gaudin, T.J., Giallombardo, A., Giannini, N.P., Goldgerb, S.L., Kraatz, B.P., Luo, Z.-X., Meng, J., Ni, X., Novacek, M.J., Perini, F.A., Randall, Z.S., Rougier, G.W., Sargis, E.J., Silcox, M.T., Simmons, N.B., Spaulding, M., Velazco, P.M., Weksler, M., Wible, J.R., and Cirranello, A.L. (2013). The placental mammal ancestor and the post-K-Pg radiation of placentals. Science 339, 662–667.CrossRefGoogle ScholarPubMed
O’Toole, B. (2002). Phylogeny of the species of the superfamily Echeneoidea (Perciformes: Carangoidei: Echeneidae, Rachycentridae, and Coryphaenidae), with an interpretation of echeneid hitchhiking behaviour. Can. J. Zool. 80, 596–623.CrossRefGoogle Scholar
Oates, A.C., Morelli, L.G., and Ares, S. (2012). Patterning embryos with oscillations: structure, function, and dynamics of the vertebrate segmentation clock. Development 139, 625–639.CrossRefGoogle ScholarPubMed
Ober, K.A. and Jockusch, E.L. (2006). The roles of wingless and decapentaplegic in axis and appendage development in the red flour beetle, Tribolium castaneum. Dev. Biol. 294, 391–405.CrossRefGoogle Scholar
Oetting, W.S. and King, R.A. (1999). Molecular basis of albinism: mutations and polymorphisms of pigmentation genes associated with albinism. Hum. Mut. 13, 99–115.3.0.CO;2-C>CrossRefGoogle ScholarPubMed
Offen, N., Blum, N., Meyer, A., and Begemann, G. (2008). Fgfr1 signalling in the development of a sexually selected trait in vertebrates, the sword of the swordtail fish. BMC Dev. Biol. 8, Article 98 (18 pp.).CrossRefGoogle ScholarPubMed
Offen, N., Meyer, A., and Begemann, G. (2009). Identification of novel genes involved in the development of the sword and gonopodium in swordtail fish. Dev. Dynamics 238, 1674–1687.CrossRefGoogle ScholarPubMed
Ogura, A., Ikeo, K., and Gojobori, T. (2005). Estimation of ancestral gene set of bilaterian animals and its implication to dynamic change of gene content in bilaterian evolution. Gene 345, 65–71.CrossRefGoogle ScholarPubMed
Ohde, T., Yaginuma, T., and Niimi, T. (2013). Insect morphological diversification through the modification of wing serial homologs. Science 340, 495–498.CrossRefGoogle ScholarPubMed
Oka, K., Yoshiyama, N., Tojo, K., Machida, R., and Hatakeyama, M. (2010). Characterization of abdominal appendages in the sawfly, Athalia rosae (Hymenoptera), by morphological and gene expression analyses. Dev. Genes Evol. 220, 53–59.CrossRefGoogle Scholar
Okamoto, K.W. and Grether, G.F. (2013). The evolution of species recognition in competitive and mating contexts: the relative efficacy of alternative mechanisms of character displacement. Ecol. Letters 16, 670–678.CrossRefGoogle ScholarPubMed
Oldham, S., Stocker, H., Laffargue, M., Wittwer, F., Wymann, M., and Hafen, E. (2002). The Drosophila insulin/IGF receptor controls growth and size by modulating PtdInsP3 levels. Development 129, 4103–4109.Google Scholar
Oliver, G., Wright, C.V.E., Hardwicke, J., and De Robertis, E.M. (1988). Differential antero-posterior expression of two proteins encoded by a homeobox gene in Xenopus and mouse embryos. EMBO J. 7, 3199–3209.Google ScholarPubMed
Oliver, G., Wright, C.V.E., Hardwicke, J., and De Robertis, E.M. (1988). A gradient of homeodomain protein in developing forelimbs of Xenopus and mouse embryos. Cell 55, 1017–1024.CrossRefGoogle ScholarPubMed
Oliver, J.C. and Monteiro, A. (2011). On the origins of sexual dimorphism in butterflies. Proc. R. Soc. Lond. B 278, 1981–1988.CrossRefGoogle ScholarPubMed
Oliver, J.C., Tong, X.-L., Gall, L.F., Piel, W.H., and Monteiro, A. (2012). A single origin for Nymphalid butterfly eyespots followed by widespread loss of associated gene expression. PLoS Genet. 8 #8, e1002893.CrossRefGoogle ScholarPubMed
Olofsson, M., Jakobsson, S., and Wiklund, C. (2013). Bird attacks on a butterfly with marginal eyespots and the role of prey concealment against the background. Biol. J. Linnean Soc. 109, 290–297.CrossRefGoogle Scholar
Olofsson, M., Vallin, A., Jakobsson, S., and Wiklund, C. (2010). Marginal eyespots on butterfly wings deflect bird attacks under low light intensities with UV wavelengths. PLoS ONE 5 #5, e10798.CrossRefGoogle ScholarPubMed
Olson, M.E. (2012). The developmental renaissance in adaptationism. Trends Ecol. Evol. 27, 278–287.CrossRefGoogle ScholarPubMed
Olson, S.L. and Feduccia, A. (1980). Relationships and evolution of flamingos (Aves: Phoenicopteridae). Smithson. Contr. Zool. No. 316, 1–73.Google Scholar
Olson-Manning, C.F., Wagner, M.R., and Mitchell-Olds, T. (2012). Adaptive evolution: evaluating empirical support for theoretical predictions. Nature Rev. Genet. 13, 867–877.CrossRefGoogle ScholarPubMed
Olsson, L. (2011). Morphogenesis of pigment patterns. In Epigenetics: Linking Genotype and Phenotype in Development and Evolution (Hallgrímsson, B. and Hall, B.K., eds.). University of California Press, Berkeley, CA, pp. 164–180.Google Scholar
Olsson, M., Meadows, J.R.S., Truvé, K., Pielberg, G.R., Puppo, F., Mauceli, E., Quilez, J., Tonomura, N., Zanna, G., Docampo, M.J., Bassols, A., Avery, A.C., Karlsson, E.K., Thomas, A., Kastner, D.L., Bongcam-Rudloff, E., Webster, M.T., Sanchez, A., Hedhammar, A., Remmers, E.F., Andersson, L., Ferrer, L., Tintle, L., and Lindblad-Toh, K. (2011). A novel unstable duplication upstream of HAS2 predisposes to a breed-defining skin phenotype and a periodic fever syndrome in Chinese Shar-Pei dogs. PLoS Genet. 7 #3, e1001332.CrossRefGoogle Scholar
Oosterveen, T., Kurdija, S., Alekseenko, Z., Uhde, C.W., Bergsland, M., Sandberg, M., Andersson, E.R., Dias, J.M., Muhr, J., and Ericson, J. (2012). Mechanistic differences in the transcriptional interpretation of local and long-range Shh morphogen signaling. Dev. Cell 23, 1006–1019.CrossRefGoogle ScholarPubMed
Oppenheimer, P. (1989). The artificial menagerie. In Artificial Life (Langton, C.G., ed.). Addison-Wesley, New York, NY, pp. 251–274.Google Scholar
Orenic, T.V., Held, L.I., Jr., Paddock, S.W., and Carroll, S.B. (1993). The spatial organization of epidermal structures: hairy establishes the geometrical pattern of Drosophila leg bristles by delimiting the domains of achaete expression. Development 118, 9–20.Google ScholarPubMed
Ortolani, A. (1999). Spots, stripes, tail tips and dark eyes: Predicting the function of carnivore colour patterns using the comparative method. Biol. J. Linnean Soc. 67, 433–476.CrossRefGoogle Scholar
Oster, G. (1988). Lateral inhibition models of developmental processes. Math. Biosci. 90, 265–286.CrossRefGoogle Scholar
Oster, G. (2005). George Oster (interview). Curr. Biol. 15, R5–R7.CrossRefGoogle Scholar
Oster, G.F., Murray, J.D., and Harris, A.K. (1983). Mechanical aspects of mesenchymal morphogenesis. J. Embryol. Exp. Morph. 78, 83–125.Google ScholarPubMed
Oster, G.F., Murray, J.D., and Maini, P.K. (1985). A model for chondrogenic condensations in the developing limb: the role of extracellular matrix and cell tractions. J. Embryol. Exp. Morph. 89, 93–112.Google ScholarPubMed
Oster, G.F., Shubin, N., Murray, J.D., and Alberch, P. (1988). Evolution and morphogenetic rules: the shape of the vertebrate limb in ontogeny and phylogeny. Evolution 42, 862–884.CrossRefGoogle ScholarPubMed
Osterauer, R., Marschner, L., Betz, O., Gerberding, M., Sawasdee, B., Cloetens, P., Haus, N., Sures, B., Triebskorn, R., and Köhler, H.-R. (2010). Turning snails into slugs: induced body plan changes and formation of an internal shell. Evol. Dev. 12, 474–483.CrossRefGoogle ScholarPubMed
Othmer, H.G., Painter, K., Umulis, D., and Xue, C. (2009). The intersection of theory and application in elucidating pattern formation in developmental biology. Math. Model. Nat. Phenom. 4 #4, 3–82.CrossRefGoogle ScholarPubMed
Outomuro, D., Adams, D.C., and Johansson, F. (2013). The evolution of wing shape in ornamented-winged damselflies (Calopterygidae, Odonata). Evol. Biol. 40, 300–309.CrossRefGoogle Scholar
Ouweneel, W.J. (1976). Developmental genetics of homoeosis. Adv. Genet. 18, 179–248.Google ScholarPubMed
Pabo, C.O. and Sauer, R.T. (1992). Transcription factors: structural families and principles of DNA recognition. Annu. Rev. Biochem. 61, 1053–1095.CrossRefGoogle ScholarPubMed
Packard, A. (1972). Cephalopods and fish: the limits of convergence. Biol. Rev. 47, 241–307.CrossRefGoogle Scholar
Packer, C. (2010). Lions. Curr. Biol. 20, R590–R591.CrossRefGoogle ScholarPubMed
Page, D.T. (2002). Inductive patterning of the embryonic brain in Drosophila. Development 129, 2121–2128.Google ScholarPubMed
Pajni-Underwood, S., Wilson, C.P., Elder, C., Mishina, Y., and Lewandoski, M. (2007). BMP signals control limb bud interdigital programmed cell death by regulating FGF signaling. Development 134, 2359–2368.CrossRefGoogle ScholarPubMed
Palmer, A.R. (2005). Antisymmetry. In Variation: A Central Concept in Biology (Hallgrímsson, B. and Hall, B.K., eds.). Elsevier Academic Press, New York, NY, pp. 359–397.CrossRefGoogle Scholar
Palmer, A.R. (2011). Developmental plasticity and the origin of novel forms: unveiling cryptic genetic variation via “use” and “disuse”. J. Exp. Zool. (Mol. Dev. Evol.) 318B, 466–479.Google Scholar
Palmer, C. and Dyke, G.J. (2010). Biomechanics of the unique pterosaur pteroid. Proc. R. Soc. Lond. B 277, 1121–1127.CrossRefGoogle ScholarPubMed
Palopoli, M.F. and Patel, N.H. (1998). Evolution of the interaction between Hox genes and a downstream target. Curr. Biol. 8, 587–590.CrossRefGoogle Scholar
Panchen, A.L. (2001). Étienne Geoffroy St.-Hilaire: father of “evo-devo”?Evol. Dev. 3, 41–46.CrossRefGoogle Scholar
Pang, K., Ryan, J.F., Baxevanis, A.D., and Martindale, M.Q. (2011). Evolution of the TGF-β signaling pathway and its potential role in the ctenophore, Mnemiopsis leidyi. PLoS ONE 6 #9, e24152.CrossRefGoogle ScholarPubMed
Panganiban, G. (2000). Distal-less function during Drosophila appendage and sense organ development. Dev. Dynamics 218, 554–562.3.0.CO;2-#>CrossRefGoogle ScholarPubMed
Panganiban, G., Nagy, L., and Carroll, S.B. (1994). The role of the Distal-less gene in the development and evolution of insect limbs. Curr. Biol. 4, 671–675.CrossRefGoogle ScholarPubMed
Panganiban, G. and Rubenstein, J.L.R. (2002). Developmental functions of the Distal-less/Dlx homeobox genes. Development 129, 4371–4386.Google ScholarPubMed
Panhuis, T.M., Butlin, R., Zuk, M., and Tregenza, T. (2001). Sexual selection and speciation. Trends Ecol. Evol. 16, 364–371.CrossRefGoogle ScholarPubMed
Pantalacci, S., Sémon, M., Martin, A., Chevret, P., and Laudet, V. (2009). Heterochronic shifts explain variations in a sequentially developing repeated pattern: palatal ridges of muroid rodents. Evol. Dev. 11, 422–433.CrossRefGoogle Scholar
Pantin, C.F.A. (1951). Organic design. Advancement of Science (London) 8, 138–150.Google Scholar
Papa, R., Martin, A., and Reed, R.D. (2009). Genomic hotspots of adaptation in butterfly wing pattern evolution. Curr. Opin. Gen. Dev. 18, 559–564.CrossRefGoogle Scholar
Papatsenko, D. (2009). Stripe formation in the early fly embryo: principles, models, and networks. BioEssays 31, 1172–1180.CrossRefGoogle Scholar
Parichy, D.M. (2003). Pigment patterns: fish in stripes and spots.Curr.Biol.13,R947–R950.CrossRefGoogle ScholarPubMed
Paris, M., Escriva, H., Schubert, M., Brunet, F., Brtko, J., Ciesielski, F., Roecklin, D., Vivat-Hannah, V., Jamin, E.L., Cravedi, J.-P., Scanlan, T.S., Renaud, J.-P., Holland, N.D., and Laudet, V. (2008). Amphioxus postembryonic development reveals the homology of chordate metamorphosis. Curr. Biol. 18, 825–830.CrossRefGoogle ScholarPubMed
Parker, G.H. (1928). Vestigial organs. In Creation by Evolution: A Consensus of Present-day Knowledge as Set Forth by Leading Authorities in Non-Technical Language That All May Understand (Mason, F., ed.). Macmillan, New York, pp. 34–48.Google Scholar
Parsons, K.J. and Albertson, R.C. (2009). Roles for BMP4 and CaM1 in shaping the jaw: evo-devo and beyond. Annu. Rev. Genet. 43, 369–388.CrossRefGoogle ScholarPubMed
Parsons, R., Aldous-Mycock, C., and Perrin, M.R. (2007). A genetic index for stripe-pattern reduction in the zebra: the quagga project. S. Afr. J. Wildlife Res. 37 #2, 105–116.CrossRefGoogle Scholar
Partridge, J.C. (2012). Sensory ecology: giant eyes for giant predators? Curr. Biol. 22, R268–R270.CrossRefGoogle ScholarPubMed
Passalacqua, K.D., Hrycaj, S., Mahfooz, N., and Popadic, A. (2010). Evolving expression patterns of the homeotic gene Scr in insects. Int. J. Dev. Biol. 54, 897–904.CrossRefGoogle ScholarPubMed
Paterson, J.R., García-Bellido, D.C., Lee, M.S.Y., Brock, G.A., Jago, J.B., and Edgecombe, G.D. (2011). Acute vision in the giant Cambrian predator Anomalocaris and the origin of compound eyes. Nature 480, 237–240.CrossRefGoogle ScholarPubMed
Patwari, P., Emilsson, V., Schadt, E.E., Chutkow, W.A., Lee, S., Marsili, A., Zhang, Y., Dobrin, R., Cohen, D.E., Larsen, P.R., Zavacki, A.M., Fong, L.G., Young, S.G., and Lee, R.T. (2011). The arrestin domain-containing 3 protein regulates body mass and energy expenditure. Cell Metab. 14, 671–683.CrossRefGoogle ScholarPubMed
Paulsen, S.M. and Nijhout, H.F. (1993). Phenotypic correlation structure among elements of the color pattern in Precis coenia (Lepidoptera: Nymphalidae). Evolution 47, 593–618.Google Scholar
Pavan, W.J. and Raible, D.W. (2012). Specification of neural crest into sensory neuron and melanocyte lineages. Dev. Biol. 366, 55–63.CrossRefGoogle ScholarPubMed
Pavlicev, M. and Wagner, G.P. (2012). Coming to grips with evolvability. Evo. Edu. Outreach 5, 231–244.CrossRefGoogle Scholar
Pavlopoulos, A. and Akam, M. (2011). Hox gene Ultrabithorax regulates distinct sets of target genes at successive stages of Drosophila haltere morphogenesis. PNAS 108 #7, 2855–2860.CrossRefGoogle ScholarPubMed
Pavlopoulos, A. and Averof, M. (2002). Developmental evolution: Hox proteins ring the changes. Curr. Biol. 12, R291–R293.CrossRefGoogle ScholarPubMed
Payankaulam, S., Li, L.M., and Arnosti, D.N. (2010). Transcriptional repression: conserved and evolved features. Curr. Biol. 20, R764–R771.CrossRefGoogle ScholarPubMed
Pearl, R. (1913). On the correlation between the number of mammae of the dam and size of litter in mammals. I. Interracial correlation. Proc. Soc. Exp. Biol. Med. 11, 27–30.CrossRefGoogle Scholar
Pearl, R. (1913). On the correlation between the number of mammae of the dam and size of litter in mammals. II. Intraracial correlation in swine. Proc. Soc. Exp. Biol. Med. 11, 31–32.CrossRefGoogle Scholar
Pearson, J.C., Lemons, D., and McGinnis, W. (2005). Modulating Hox gene functions during animal body patterning. Nature Rev. Genet. 6, 893–904.CrossRefGoogle ScholarPubMed
Pearson, J.E. (1993). Complex patterns in a simple system. Science 261, 189–192.CrossRefGoogle Scholar
Pecoits, E., Konhauser, K.O., Aubet, N.R., Heaman, L.M., Veroslavsky, G., Stern, R.A., and Gingras, M.K. (2012). Bilaterian burrows and grazing behavior at >585 million years ago. Science 336, 1693–1696.CrossRefGoogle ScholarPubMed
Peery, M.Z. and Pauli, J.N. (2012). The mating system of a “lazy” mammal, Hoffmann’s two-toed sloth. Anim. Behav. 84, 555–562.CrossRefGoogle Scholar
Peichl, L., Behrmann, G., and Kröger, R.H.H. (2001). For whales and seals the ocean is not blue: a visual pigment loss in marine mammals. Eur. J. Neurosci. 13, 1520–1528.CrossRefGoogle Scholar
Pelaz, S., Urquía, N., and Morata, G. (1993). Normal and ectopic domains of the homeotic gene Sex combs reduced of Drosophila. Development 117, 917–923.Google ScholarPubMed
Peng, Y., Han, C., and Axelrod, J.D. (2012). Planar polarized protrusions break the symmetry of EGFR signaling during Drosophila bract cell fate induction. Dev. Cell 23, 507–518.CrossRefGoogle ScholarPubMed
Pennacchio, L.A., Bickmore, W., Dean, A., Nobrega, M.A., and Bejerano, G. (2013). Enhancers: five essential questions. Nature Rev. Genet. 14, 288–295.CrossRefGoogle ScholarPubMed
Perry, S.F., Similowski, T., Klein, W., and Codd, J.R. (2010). The evolutionary origin of the mammalian diaphragm. Respir. Physiol. Neurobiol. 171, 1–16.CrossRefGoogle ScholarPubMed
Peterkova, R., Lesot, H., and Peterka, M. (2006). Phylogenetic memory of developing mammalian dentition. J. Exp. Zool. (Mol. Dev. Evol.) 306B, 234–250.CrossRefGoogle Scholar
Peterson, A. (1962). Larvae of Insects: An Introduction to Nearctic Species. Part I. Lepidoptera and Plant Infesting Hymenoptera, 4th edn. Edwards Bros., Columbus, OH.Google Scholar
Peterson, A.A. and Peterson, A.T. (1992). Aztec exploitation of cloud forests: tributes of liquidambar resin and quetzal feathers. Global Ecol. Biogeogr. Lett. 2, 165–173.CrossRefGoogle Scholar
Petzoldt, A.G., Coutelis, J.-B., Géminard, C., Spéder, P., Suzanne, M., Cerezo, D., and Noselli, S. (2012). DE-Cadherin regulates unconventional Myosin ID and Myosin IC in Drosophila left-right asymmetry establishment. Development 139, 1874–1884.CrossRefGoogle ScholarPubMed
Pfennig, D.W. and Pfennig, K.S. (2012). Evolution’s Wedge: Competition and the Origins of Diversity. University of California Press, Berkeley, CA.CrossRefGoogle Scholar
Philippe, H., Brinkmann, H., Copley, R.R., Moroz, L.L., Nakano, H., Poustka, A.J., Wallberg, A., Peterson, K.J., and Telford, M.J. (2011). Acoelomorph flatworms are deuterostomes related to Xenoturbella. Nature 470, 255–258.CrossRefGoogle ScholarPubMed
Philippe, H., Derelle, R., Lopez, P., Pick, K., Borchiellini, C., Boury-Esnault, N., Vacelet, J., Renard, E., Houliston, E., Quéinnec, E., Da Silva, C., Wincker, P., Le Guyader, H., Leys, S., Jackson, D.J., Schreiber, F., Erpenbeck, D., Morgenstern, B., Wörheide, G., and Manuel, M. (2009). Phylogenomics revives traditional views on deep animal relationships. Curr. Biol. 19, 706–712.CrossRefGoogle ScholarPubMed
Phinchongsakuldit, J., MacArthur, S., and Brookfield, J.F.Y. (2004). Evolution of developmental genes: Molecular microevolution of enhancer sequences at the Ubx locus in Drosophila and its impact on developmental phenotypes. Mol. Biol. Evol. 21, 348–363.CrossRefGoogle ScholarPubMed
Piasecka, B., Lichocki, P., Moretti, S., Bergmann, S., and Robinson-Rechavi, M. (2013). The hourglass and the early conservation models: co-existing patterns of developmental constraints in vertebrates. PLoS Genet. 9 #4, e1003476.CrossRefGoogle ScholarPubMed
Pick, L. and Heffer, A. (2012). Hox gene evolution: multiple mechanisms contributing to evolutionary novelties. Ann. N. Y. Acad. Sci. 1256, 15–32.CrossRefGoogle ScholarPubMed
Picken, L.E.R. (1949). Shape and molecular orientation in lepidopteran scales. Phil. Trans. R. Soc. Lond. B 234 #608, 1–28.CrossRefGoogle Scholar
Pigliucci, M. (2008). Is evolvability evolvable? Nature Rev. Gen. 9, 75–82.CrossRefGoogle ScholarPubMed
Pinney, R. (1981). The Snake Book. Doubleday, Garden City, NY.Google Scholar
Pires-daSilva, A. and Sommer, R.J. (2003). The evolution of signalling pathways in animal development. Nature Rev. Genet. 4, 39–49.CrossRefGoogle ScholarPubMed
Pitsouli, C. and Perrimon, N. (2008). Our fly cousins’ gut. Nature 454, 592–593.CrossRefGoogle ScholarPubMed
Platt, J.R. (1964). Strong inference. Science 146, 347–353.CrossRefGoogle ScholarPubMed
Plikus, M. and Chuong, C.-M. (2004). Making waves with hairs. J. Invest. Dermatol. 122, vii–ix.CrossRefGoogle ScholarPubMed
Plikus, M.V., Baker, R.E., Chen, C.-C., Fare, C., de la Cruz, D., Andl, T., Maini, P.K., Millar, S.E., Widelitz, R., and Chuong, C.-M. (2011). Self-organizing and stochastic behaviors during the regeneration of hair stem cells. Science 332, 586–589.CrossRefGoogle ScholarPubMed
Plikus, M.V. and Chuong, C.-M. (2008). Complex hair cycle domain patterns and regenerative hair waves in living rodents. J. Invest. Dermatol. 128, 1071–1080.CrossRefGoogle ScholarPubMed
Polak, M., Starmer, W.T., and Wolf, L.L. (2004). Sexual selection for size and symmetry in a diversifying secondary sexual character in Drosophila bipectinata Duda (Diptera: Drosophilidae). Evolution 58, 597–607.CrossRefGoogle Scholar
Policansky, D. (1982). The asymmetry of flounders. Sci. Am. 246 #5, 116–122.CrossRefGoogle Scholar
Polly, P.D. (2007). Development with a bite. Nature 449, 413–415.CrossRefGoogle ScholarPubMed
Polly, P.D., Head, J.J., and Cohn, M.J. (2001). Testing modularity and dissociation: the evolution of regional proportions in snakes. In Beyond Heterochrony: The Evolution of Development (Zelditch, M.L., ed.). Wiley-Liss, New York, NY, pp. 305–335.Google Scholar
Pool, R. (1991). Did Turing discover how the leopard got its spots? Science 251, 627.CrossRefGoogle ScholarPubMed
Popadic, A., Abzhanov, A., Rusch, D., and Kaufman, T.C. (1998). Understanding the genetic basis of morphological evolution: the role of homeotic genes in the diversification of the arthropod bauplan. Int. J. Dev. Biol. 42, 453–461.Google ScholarPubMed
Popadic, A., Panganiban, G., Rusch, D., Shear, W.A., and Kaufman, T.C. (1998). Molecular evidence for the gnathobasic derivation of arthropod mandibles and for the appendicular origin of the labrum and other structures. Dev. Genes Evol. 208, 142–150.Google ScholarPubMed
Porcher, A. and Dostatni, N. (2010). The Bicoid morphogen system. Curr. Biol. 20, R249–R254.CrossRefGoogle ScholarPubMed
Porter, M.L. and Crandall, K.A. (2003). Lost along the way: the significance of evolution in reverse. Trends Ecol. Evol. 18, 541–547.CrossRefGoogle Scholar
Portmann, A. (1967). Animal Forms and Patterns: A Study of the Appearance of Animals. Schocken Books, New York, NY.Google Scholar
Posnien, N., Bashasab, F., and Bucher, G. (2009). The insect upper lip (labrum) is a nonsegmental appendage-like structure. Evol. Dev. 11, 480–488.CrossRefGoogle ScholarPubMed
Poss, K.D. (2010). Advances in understanding tissue regenerative capacity and mechanisms in animals. Nature Rev. Genet. 11, 710–722.CrossRefGoogle ScholarPubMed
Pottin, K., Hinaux, H., and Rétaux, S. (2011). Restoring eye size in Astyanax mexicanus blind cavefish embryos through modulation of the Shh and Fgf8 forebrain organising centers. Development 138, 2467–2476.CrossRefGoogle Scholar
Pough, F.H., Janis, C.M., and Heiser, J.B. (2009). Vertebrate Life. Benjamin Cummings, New York, NY.Google Scholar
Pourquié, O. (2011). Vertebrate segmentation: from cyclic gene networks to scoliosis. Cell 145, 650–663. [See also Bénazéraf, B. and Pourquié, O. (2013). Formation and segmentation of the vertebrate body axis. Annu. Rev. Cell Dev. Biol. 29, 1–26.]CrossRefGoogle ScholarPubMed
Powell, B.C. and Rogers, G.E. (1990). Cyclic hair-loss and regrowth in transgenic mice overexpressing an intermediate filament gene. EMBO J. 9, 1485–1493.Google ScholarPubMed
Preston, J.C. and Hileman, L.C. (2012). Parallel evolution of TCP and B-class genes in Commelinaceae flower bilateral symmetry. EvoDevo 3, Article 6 (14 pp.).CrossRefGoogle ScholarPubMed
Prince, V.E. and Pickett, F.B. (2002). Splitting pairs: the diverging fates of duplicated genes. Nature Rev. Genet. 3, 827–837.CrossRefGoogle ScholarPubMed
Pringle, J.W.S. (1948). The gyroscopic mechanism of the halteres of diptera. Phil. Trans. Roy. Soc. Lond. B 233, 347–384.CrossRefGoogle Scholar
Pringle, J.W.S. (1957). Insect Flight. Cambridge Monographs in Experimental Biology, Vol. 9. Cambridge University Press, New York, NY.Google Scholar
Pringle, J.W.S. (1975). Insect flight. Oxford Biol. Readers 52, 16 pp.Google Scholar
Proske, U. and Gregory, E. (2003). Electrolocation in the platypus–some speculations. Comp. Biochem. Physiol. A: Mol. & Integr. Physiol. 136, 821–825.CrossRefGoogle ScholarPubMed
Protas, M.E. and Patel, N.H. (2008). Evolution of color patterns. Annu. Rev. Cell Dev. Biol. 24, 425–446. [See also Ruxton, G.D. (2013). Hide and seek and other sensory games. Curr. Biol. 23, R465–R467.]CrossRefGoogle Scholar
Prothero, D.R. (1987). The rise and fall of the American rhino. Nat. Hist. 96 #8, 26–33.Google Scholar
Prothero, D.R. (2009). Evolutionary transitions in the fossil record of terrestrial hoofed mammals. Evo. Edu. Outreach 2, 289–302.CrossRefGoogle Scholar
Prothero, D.R. and Schoch, R.M. (2002). Horns, Tusks, and Flippers: The Evolution of Hoofed Mammals. Johns Hopkins University Press, Baltimore, MD.Google Scholar
Prpic, N.-M. (2008). Arthropod appendages: a prime example for the evolution of morphological diversity and innovation. In Evolving Pathways: Key Themes in Evolutionary Developmental Biology (Minelli, A. and Fusco, G., eds.). Cambridge University Press, New York, NY, pp. 381–398.CrossRefGoogle Scholar
Prud’homme, B. and Gompel, N. (2010). Genomic hourglass. Nature 468, 768–769.CrossRefGoogle ScholarPubMed
Prud’homme, B., Gompel, N., and Carroll, S.B. (2007). Emerging principles of regulatory evolution. PNAS 104 (Suppl. 1) 8605–8612.CrossRefGoogle ScholarPubMed
Prud’homme, B., Gompel, N., Rokas, A., Kassner, V.A., Williams, T.M., Yeh, S.-D., True, J.R., and Carroll, S.B. (2006). Repeated morphological evolution through cis-regulatory changes in a pleiotropic gene. Nature 440, 1050–1053.CrossRefGoogle Scholar
Prud’homme, B., Minervino, C., Hocine, M., Cande, J.D., Aouane, A., Dufour, H.D., Kassner, V.A., and Gompel, N. (2011). Body plan innovation in treehoppers through the evolution of an extra wing-like appendage. Nature 473, 83–86.CrossRefGoogle ScholarPubMed
Prudic, K.L., Jeon, C., Cao, H., and Monteiro, A. (2011). Developmental plasiticity in sexual roles of butterfly species drives mutual sexual ornamentation. Science 331, 73–75.CrossRefGoogle Scholar
Pruvosta, M., Bellonec, R., Beneckeb, N., Sandoval-Castellanosd, E., Cieslaka, M., Kuznetsovae, T., Morales-Muñizf, A., O’Connorg, T., Reissmannh, M., Hofreiteri, M., and Ludwiga, A. (2011). Genotypes of predomestic horses match phenotypes painted in Paleolithic works of cave art. PNAS 108 #46, 18626–18630.CrossRefGoogle Scholar
Prykhozhij, S.V. and Neumann, C.J. (2008). Distinct roles of Shh and Fgf signaling in regulating cell proliferation during zebrafish pectoral fin development. BMC Dev. Biol. 8, Article 91 (11 pp.).CrossRefGoogle ScholarPubMed
Pueyo, J.I. and Couso, J.P. (2005). Parallels between the proximal-distal development of vertebrate and arthropod appendages: homology without an ancestor? Curr. Opin. Gen. Dev. 15, 439–446.CrossRefGoogle ScholarPubMed
Pueyo, J.I. and Couso, J.P. (2011). Tarsal-less peptides control Notch signalling through the Shavenbaby transcription factor. Dev. Biol. 355, 183–193.CrossRefGoogle ScholarPubMed
Purves, W.K., Orians, G.H., and Heller, H.C. (1992). Life: The Science of Biology. Sinauer, Sunderland, MA.Google Scholar
Pyron, R.A., Burbrink, F.T., and Wiens, J.J. (2013). A phylogeny and revised classification of Squamata, including 4161 species of lizards and snakes. BMC Evol. Biol. 13, Article 93 (53 pp.).CrossRefGoogle ScholarPubMed
Quan, X.-J. and Hassan, B.A. (2005). From skin to nerve: flies, vertebrates and the first helix. Cell. Mol. Life Sci. 62, 2036–2049.CrossRefGoogle ScholarPubMed
Quigley, I.K., Manuel, J.L., Roberts, R.A., Nuckels, R.J., Herrington, E.R., MacDonald, E.L., and Parichy, D.M. (2004). Evolutionary diversification of pigment pattern in Danio fishes: differential fms dependence and stripe loss in D. albolineatus. Development 132, 89–104.CrossRefGoogle ScholarPubMed
Quigley, I.K., Turner, J.M., Nuckels, R.J., Manuel, J.L., Budi, E.H., MacDonald, E.L., and Parichy, D.M. (2004). Pigment pattern evolution by differential deployment of neural crest and post-embryonic melanophore lineages in Danio fishes. Development 131, 6053–6069.CrossRefGoogle ScholarPubMed
Rachlow, J.L. and Berger, J. (1997). Conservation implications of patterns of horn regeneration in dehorned white rhinos. Conserv. Biol. 11, 84–91.CrossRefGoogle Scholar
Radinsky, L. and Emerson, S. (1982). The late, great sabertooths. Nat. Hist. 91 #4, 50–57.Google Scholar
Raff, E.C. and Raff, R.A. (2000). Dissociability, modularity, evolvability. Evol. Dev. 2, 235–237.CrossRefGoogle ScholarPubMed
Raff, R.A. (1996). The Shape of Life: Genes, Development, and the Evolution of Animal Form. University of Chicago Press, Chicago, IL.Google Scholar
Raff, R.A. and Kaufman, T.C. (1983). Embryos, Genes, and Evolution: The Developmental-Genetic Basis of Evolutionary Change. Macmillan, New York, NY.Google Scholar
Raff, R.A. and Sly, B.J. (2000). Modularity and dissociation in the evolution of gene expression territories in development. Evol. Dev. 2, 102–113.CrossRefGoogle ScholarPubMed
Raible, F. and Arendt, D. (2004). Metazoan evolution: some animals are more equal than others. Curr. Biol. 14, R106–R108.CrossRefGoogle ScholarPubMed
Ramachandran, V.S., Tyler, C.W., Gregory, R.L., Rogers-Ramachandran, D., Duensing, S., Pillsbury, C., and Ramachandran, C. (1996). Rapid adaptive camouflage in tropical flounders. Nature 379, 815–818.CrossRefGoogle ScholarPubMed
Ramel, M.-C. and Hill, C.S. (2013). The ventral to dorsal BMP activity gradient in the early zebrafish embryo is determined by graded expression of BMP ligands. Dev. Biol. 378, 170–182.CrossRefGoogle ScholarPubMed
Ramos, D.M. and Monteiro, A. (2007). Transgenic approaches to study wing color pattern development in Lepidoptera. Mol. BioSyst. 3, 530–535.CrossRefGoogle ScholarPubMed
Ramsden, C.A., Bankier, A., Brown, T.J., Cowen, P.S.J., Frost, G.I., McCallum, D.D., Studdert, V.P., and Fraser, J.R.E. (2000). A new disorder of hyaluronan metabolism associated with generalized folding and thickening of the skin. J. Ped. 136, 62–68.CrossRefGoogle Scholar
Rancourt, D.E., Tsuzuki, T., and Capecchi, M.R. (1995). Genetic interaction between hoxb-5 and hoxb-6 is revealed by nonallelic noncomplementation. Genes Dev. 9, 108–122.CrossRefGoogle ScholarPubMed
Randall, V.A. (2007). Hormonal regulation of hair follicles exhibits a biological paradox. Semin. Cell Dev. Biol. 18, 274–285.CrossRefGoogle ScholarPubMed
Randsholt, N.B. and Santamaria, P. (2008). How Drosophila change their combs: the Hox gene Sex combs reduced and sex comb variation among Sophophora species. Evol. Dev. 10, 121–133.CrossRefGoogle ScholarPubMed
Rasmussen, A.R., Murphy, J.C., Ompi, M., Gibbons, J.W., and Uetz, P. (2011). Marine reptiles. PLoS ONE 6 #11, e27373.CrossRefGoogle ScholarPubMed
Rastegar, S., Hess, I., Dickmeis, T., Nicod, J.C., Ertzer, R., Hadzhiev, Y., Thies, W.-G., Scherer, G., and Strähle, U. (2008). The words of the regulatory code are arranged in a variable manner in highly conserved enhancers. Dev. Biol. 318, 366–377.CrossRefGoogle Scholar
Raubenheimer, E.J. (2000). Early development of the tush and the tusk of the African elephant (Loxodonta africana). Arch. Oral Biol. 45, 983–986.CrossRefGoogle Scholar
Raup, D.M. (1966). Geometric analysis of shell coiling: general problems. J. Paleontol. 40, 1178–1190.Google Scholar
Raup, D.M. and Michelson, A. (1965). Theoretical morphology of the coiled shell. Science 147, 1294–1295.CrossRefGoogle ScholarPubMed
Rawls, J.F., Mellgren, E.M., and Johnson, S.L. (2001). How the zebrafish gets its stripes. Dev. Biol. 240, 301–314.CrossRefGoogle ScholarPubMed
Raynaud, A. (1985). Development of limbs and embryonic limb reduction. In Biology of the Reptilia, Vol. 15: Development, Part B (Gans, C. and Billett, F., eds.). Wiley, New York, NY, pp. 60–148.Google Scholar
Razzell, W. and Martin, P. (2012). Embryonic clutch control. Science 335, 1181–1182.CrossRefGoogle ScholarPubMed
Reaka, M.L. (1981). The hole shrimp story. Nat. Hist. 90 #7, 36–43.Google Scholar
Rebeiz, M., Stone, T., and Posakony, J.W. (2005). An ancient transcriptional regulatory linkage. Dev. Biol. 281, 299–308.CrossRefGoogle ScholarPubMed
Rebollo, R., Romanish, M.T., and Mager, D.L. (2012). Transposable elements: an abundant and natural source of regulatory sequences for host genes. Annu. Rev. Genet. 46, 21–42.CrossRefGoogle ScholarPubMed
Reed, R.D. (2004). Evidence for Notch-mediated lateral inhibition in organizing butterfly wing scales. Dev. Genes Evol. 214, 43–46.CrossRefGoogle ScholarPubMed
Reed, R.D. and Nagy, L.M. (2005). Evolutionary redeployment of a biosynthetic module: expression of eye pigment genes vermilion, cinnabar, and white in butterfly wing development. Evol. Dev. 7, 301–311.CrossRefGoogle Scholar
Reed, R.D., Papa, R., Martin, A., Hines, H.M., Counterman, B.A., Pardo-Diaz, C., Jiggins, C.D., Chamberlain, N.L., Kronforst, M.R., Chen, R., Halder, G., Nijhout, H.F., and McMillan, W.O. (2011). optix drives the repeated convergent evolution of butterfly wing pattern mimicry. Science 333, 1137–1141.CrossRefGoogle ScholarPubMed
Reed, R.D. and Serfas, M.S. (2004). Butterfly wing pattern evolution is associated with changes in a Notch/Distal-less temporal pattern formation process. Curr. Biol. 14, 1159–1166.CrossRefGoogle Scholar
Reeder, D.M., Helgen, K.M., Vodzak, M.E., Lunde, D.P., and Ejotre, I. (2013). A new genus for a rare African vespertilionid bat: insights from South Sudan. ZooKeys 285, 89–115.CrossRefGoogle Scholar
Reeves, R.R. and Mitchell, E. (1981). The whale behind the tusk. Nat. Hist. 90 #8, 50–57.Google Scholar
Rehorn, K.-P., Thelen, H., Michelson, A.M., and Reuter, R. (1996). A molecular aspect of hematopoiesis and endoderm development common to vertebrates and Drosophila. Development 122, 4023–4031.Google ScholarPubMed
Reik, E.F. (1976). Four-winged diptera from the upper Permian of Australia. Proc. Linnean Soc. New South Wales 101 #4, 250–255.Google Scholar
Reinius, B., Saetre, P., Leonard, J.A., Blekhman, R., Merino-Martinez, R., Gilad, Y., and Jazin, E. (2008). An evolutionarily conserved sexual signature in the primate brain. PLoS Genet. 4 #6, e1000100.CrossRefGoogle ScholarPubMed
Reiss, J.O. (2003). Time. In Keywords and Concepts in Evolutionary Developmental Biology (Hall, B.K. and Olson, W.M., eds.). Harvard University Press, Cambridge, MA, pp. 358–368.Google Scholar
Reisz, R.R. (2006). Origin of dental occlusion in tetrapods: signal for terrestrial vertebrate evolution? J. Exp. Zool. (Mol. Dev. Evol.) 306B, 261–277.CrossRefGoogle Scholar
Reisz, R.R. and Head, J.J. (2008). Turtle origins out to sea. Nature 456, 450–451.CrossRefGoogle Scholar
Renaud, S., Pantalacci, S., and Auffray, J.-C. (2011). Differential evolvability along lines of least resistance of upper and lower molars in island house mice. PLoS ONE 6 #5, e18951.CrossRefGoogle ScholarPubMed
Renoult, J.P., Schaefer, H.M., Sallé, B., and Charpentier, M.J.E. (2011). The evolution of the multicolored face of mandrills: insights from the perceptual space of colour vision. PLoS ONE 6 #12, e29117.CrossRefGoogle Scholar
Rhinn, M. and Dollé, P. (2012). Retinoic acid signalling during development. Development 139, 843–858.CrossRefGoogle ScholarPubMed
Rice, S.H. (2002). The role of heterochrony in primate brain evolution. In Human Evolution Through Developmental Change (Minugh-Purvis, N. and McNamara, K.J., eds.). Johns Hopkins University Press, Baltimore, MD, pp. 154–170.Google Scholar
Richards, A.G. (1951). The Integument of Arthropods. University of Minnesota Press, Minneapolis, MN.Google Scholar
Richardson, M.K. and Chipman, A.D. (2003). Developmental constraints in a comparative framework: a test case using variations in phalanx number during amniote evolution. J. Exp. Zool. (Mol. Dev. Evol.) 296B, 8–22.CrossRefGoogle Scholar
Richardson, M.K., Gobes, S.M.H., van Leeuwen, A.C., Polman, J.A.E., Pieau, C., and Sánchez-Villagra, M.R. (2009). Heterochrony in limb evolution: developmental mechanisms and natural selection. J. Exp. Zool. (Mol. Dev. Evol.) 312B, 639–664.CrossRefGoogle Scholar
Richardson, M.K., Hornbruch, A., and Wolpert, L. (1990). Mechanisms of pigment pattern formation in the quail embryo. Development 109, 81–89.Google Scholar
Richardson, M.K., Jeffery, J.E., and Tabin, C.J. (2004). Proximodistal patterning of the limb: insights from evolutionary morphology. Evol. Dev. 6, 1–5.CrossRefGoogle ScholarPubMed
Richardson, M.K. and Keuck, G. (2002). Haeckel’s ABC of evolution and development. Biol. Rev. 77, 495–528.CrossRefGoogle ScholarPubMed
Richardson, M.K. and Oelschläger, H.H.A. (2002). Time, pattern, and heterochrony: a study of hyperphalangy in the dolphin embryo flipper. Evol. Dev. 4, 435–444.CrossRefGoogle ScholarPubMed
Richmond, D.L. and Oates, A.C. (2012). The segmentation clock: inherited trait or universal design principle? Curr. Opin. Gen. Dev. 22, 600–606.CrossRefGoogle ScholarPubMed
Ridley, M. (2004). Evolution, 3rd edn. Blackwell, Malden, MA.Google Scholar
Riedl, R. (1978). Order in Living Organisms: A Systems Analysis of Evolution. Wiley, New York, NY.Google Scholar
Rieppel, O. (2001). Turtles as hopeful monsters. BioEssays 23, 987–991.CrossRefGoogle ScholarPubMed
Rieppel, O. (2009). How did the turtle get its shell? Science 325, 154–155.CrossRefGoogle ScholarPubMed
Rincón-Limas, D.E., Lu, C.-H., Canal, I., Calleja, M., Rodríguez-Esteban, C., Izpisúa-Belmonte, J.C., and Botas, J. (1999). Conservation of the expression and function of apterous orthologs in Drosophila and mammals. PNAS 96 #5, 2165–2170.CrossRefGoogle ScholarPubMed
Ripple, J. (1999). Manatees and Dugongs of the World. Voyageur Press, Stillwater, MN.Google Scholar
Riskin, D.K. and Hermanson, J.W. (2005). Independent evolution of running in vampire bats. Nature 434, 292.CrossRefGoogle ScholarPubMed
Riskin, D.K., Parsons, S., Schutt, W.A., Jr., Carter, G.G., and Hermanson, J.W. (2006). Terrestrial locomotion of the New Zealand short-tailed bat Mystacina tuberculata and the common vampire bat Desmodus rotundus. J. Exp. Biol. 209, 1725–1736.CrossRefGoogle ScholarPubMed
Rivera, A.S. and Weisblat, D.A. (2009). And Lophotrochozoa makes three: Notch/Hes signaling in annelid segmentation. Dev. Genes Evol. 219, 37–43.CrossRefGoogle ScholarPubMed
Rizo, J. and Gierasch, L.M. (1992). Constrained peptides: models of bioactive peptides and protein substructures. Annu. Rev. Biochem. 61, 387–418.CrossRefGoogle ScholarPubMed
Robbins, D.L. (2012). KISS and the veterans. Boomer Magazine 6 #2, 92–97.Google Scholar
Robert, B. and Lallemand, Y. (2006). Anteroposterior patterning in the limb and digit specification: contributions of mouse genetics. Dev. Dynamics 235, 2337–2352.CrossRefGoogle Scholar
Robertson, K.A. and Monteiro, A. (2005). Female Bicyclus anynana butterflies choose males on the basis of their dorsal UV-reflective eyespot pupils. Proc. R. Soc. Lond. B 272, 1541–1546.CrossRefGoogle ScholarPubMed
Robertson, R.M., Pearson, K.G., and Reichert, H. (1982). Flight interneurons in the locust and the origin of insect wings. Science 217, 177–179.CrossRefGoogle ScholarPubMed
Robinett, C.C., Vaughan, A.G., Knapp, J.-M., and Baker, B.S. (2010). Sex and the single cell. II. There is a time and place for sex. PLoS Biol. 8 #5, e1000365.CrossRefGoogle Scholar
Robinson, B.G., Khurana, S., Kuperman, A., and Atkinson, N.S. (2012). Neural adaptation leads to cognitive ethanol dependence. Curr. Biol. 22, 2338–2341.CrossRefGoogle ScholarPubMed
Roch, F. and Akam, M. (2000). Ultrabithorax and the control of cell morphology in Drosophila halteres. Development 127, 97–107.Google ScholarPubMed
Roch, F., Baonza, A., Martín-Blanco, E., and García-Bellido, A. (1998). Genetic interactions and cell behaviour in blistered mutants during proliferation and differentiation of the Drosophila wing. Development 125, 1823–1832.Google ScholarPubMed
Roellig, D., Morelli, L.G., Ares, S., Jülicher, F., and Oates, A.C. (2011). Snapshot: The segmentation clock. Cell 145, 800.CrossRefGoogle ScholarPubMed
Roemmich, D. (2007). Twice bitten. Nature 449, 33–34.Google Scholar
Roff, D.A. (1994). The evolution of flightlessness: is history important? Evol. Ecol. 8, 639–657.CrossRefGoogle Scholar
Rogers, B.T., Peterson, M.D., and Kaufman, T.C. (1997). Evolution of the insect body plan as revealed by the Sex combs reduced expression pattern. Development 124, 149–157.Google ScholarPubMed
Rogers, B.T., Peterson, M.D., and Kaufman, T.C. (2002). The development and evolution of insect mouthparts as revealed by the expression patterns of gnathocephalic genes. Evol. Dev. 4, 96–110.CrossRefGoogle ScholarPubMed
Rogers, K.A.C. and D’Emic, M.D. (2012). Triumph of the titans. Sci. Am. 306 #5, 48–55.CrossRefGoogle ScholarPubMed
Rogers, K.W. and Schier, A.F. (2011). Morphogen gradients: From generation to interpretation. Annu. Rev. Cell Dev. Biol. 27, 377–407.CrossRefGoogle ScholarPubMed
Rokas, A. (2008). The origins of multicellularity and the early history of the genetic toolkit for animal development. Annu. Rev. Genet. 42, 235–251.CrossRefGoogle ScholarPubMed
Rolf, H.J., Kierdorf, U., Kierdorf, H., Schulz, J., Seymour, N., Schliephake, H., Napp, J., Niebert, S., Wölfel, H., and Wiese, K.G. (2008). Localization and characterization of STRO-1+ cells in the deer pedicle and regenerating antler. PLoS ONE 3 #4, e2064.CrossRefGoogle ScholarPubMed
Rolian, C. (2008). Developmental basis of limb length in rodents: evidence for multiple divisions of labor in mechanisms of endochondral bone growth. Evol. Dev. 10, 15–28.CrossRefGoogle ScholarPubMed
Rolian, C. and Willmore, K.E. (2009). Morphological integration at 50: patterns and processes of integration in biological anthropology. Evol. Biol. 36, 1–4.CrossRefGoogle Scholar
Rome, L.C. (1997). Testing a muscle’s design. Am. Sci. 85, 356–363.Google Scholar
Romer, A.S. (1933). Vertebrate Paleontology. University of Chicago Press, Chicago, IL.Google Scholar
Romer, A.S. (1959). The Vertebrate Story. University of Chicago Press, Chicago, IL. [See also Knüsel, J., et al. (2013). A salamander’s flexible spinal network for locomotion, modeled at two levels of abstraction. Integr. Comp. Biol. 53, 269–282.]Google Scholar
Romer, A.S. (1970). The Vertebrate Body, 4th edn. W. B. Saunders, Philadelphia, PA.Google Scholar
Ronshaugen, M., McGinnis, N., and McGinnis, W. (2002). Hox protein mutation and macroevolution of the insect body plan. Nature 415, 914–917.CrossRefGoogle ScholarPubMed
Roos, G., Van Wassenbergh, S., Aerts, P., Herrel, A., and Adriaens, D. (2011). Effects of snout dimensions on the hydrodynamics of suction feeding in juvenile and adult seahorses. J. Theor. Biol. 269, 307–317.CrossRefGoogle ScholarPubMed
Roos, G., Van Wassenbergh, S., Herrel, A., Adriaens, D., and Aerts, P. (2010). Snout allometry in seahorses: insights on optimisation of pivot feeding performance during ontogeny. J. Exp. Biol. 213, 2184–2193.CrossRefGoogle ScholarPubMed
Roos, G., Van Wassenbergh, S., Herrel, A., and Aerts, P. (2009). Kinematics of suction feeding in the seahorse Hippocampus reidi. J. Exp. Biol. 212, 3490–3498.CrossRefGoogle ScholarPubMed
Rosenberg, M.I. and Desplan, C. (2010). Hiding in plain sight. Science 329, 284–285.CrossRefGoogle ScholarPubMed
Rosenberger, A.L. and Preuschoft, H. (2012). Evolutionary morphology, cranial biomechanics and the origins of tarsiers and anthropoids. Palaeobiol. Palaeoenv. 92, 507–525.CrossRefGoogle Scholar
Rosenqvist, G. and Berglund, A. (2011). Sexual signals and mating patterns in Syngnathidae. J. Fish Biol. 78, 1647–1661.CrossRefGoogle ScholarPubMed
Ross, E.S. (1984). Mantids: the praying predators. Natl. Geogr. 165 #2, 268–279.Google Scholar
Roth, C., Rastogi, S., Arvestad, L., Dittmar, K., Light, S., Ekman, D., and Liberles, D.A. (2007). Evolution after gene duplication: models, mechanisms, sequences, systems, and organisms. J. Exp. Zool. (Mol. Dev. Evol.) 308B, 58–73.CrossRefGoogle Scholar
Roth, S. and Lynch, J. (2012). Axis formation: microtubules push in the right direction. Curr. Biol. 22, R537–R539.CrossRefGoogle ScholarPubMed
Roth, V.L. (1984). On homology. Biol. J. Linnean Soc. 22, 13–29.CrossRefGoogle Scholar
Röttinger, E. and Lowe, C.J. (2012). Evolutionary crossroads in developmental biology: hemichordates. Development 139, 2463–2475.CrossRefGoogle ScholarPubMed
Roush, W. (1995). Wing scales may help beat the heat. Science 269, 1816.CrossRefGoogle Scholar
Rousso, T., Lynch, J., Yogev, S., Roth, S., Schejter, E.D., and Shilo, B.-Z. (2010). Generation of distinct signaling modes via diversification of the EGFR ligand-processing cassette. Development 137, 3427–3427.CrossRefGoogle ScholarPubMed
Roy, S. and VijayRaghavan, K. (2012). Developmental biology: taking flight. Curr. Biol. 22, R63–R65.CrossRefGoogle ScholarPubMed
Rozowski, M. (2002). Establishing character correspondence for sensory organ traits in flies: sensory organ development provides insights for reconstructing character evolution. Mol. Phylogenet. Evol. 24, 400–411.CrossRefGoogle Scholar
Rozowski, M. and Akam, M. (2002). Hox gene control of segment-specific bristle patterns in Drosophila. Genes Dev. 16, 1150–1162.CrossRefGoogle ScholarPubMed
Rubenstein, M., Sai, Y., Chuong, C.-M., and Shen, W.-M. (2009). Regenerative patterning in Swarm Robots: mutual benefits of research in robotics and stem cell biology. Int. J. Dev. Biol. 53, 869–881.CrossRefGoogle ScholarPubMed
Rudel, D. and Sommer, R.J. (2003). The evolution of developmental mechanisms. Dev. Biol. 264, 15–37.CrossRefGoogle ScholarPubMed
Rundshagen, U., Zühlke, C., Opitz, S., Schwinger, E., and Käsmann-Kellner, B. (2004). Mutations in the MATP gene in five German patients affected by oculocutaneous albinism type 4. Hum. Mut. 23, 106–110.CrossRefGoogle ScholarPubMed
Ruppert, E.E. (2005). Key characters uniting hemichordates and chordates: homologies or homoplasies? Can. J. Zool. 83, 8–23.CrossRefGoogle Scholar
Rushlow, C.A. and Shvartsman, S.Y. (2012). Temporal dynamics, spatial range, and transcriptional interpretation of the Dorsal morphogen gradient. Curr. Opin. Gen. Dev. 22, 542–546.CrossRefGoogle ScholarPubMed
Russell, F.E. (1984). Snake venoms. Symp. Zool. Soc. Lond. 52, 469–480.Google Scholar
Russo, G.A. and Shapiro, L.J. (2011). Morphological correlates of tail length in the catarrhine sacrum. J. Hum. Evol. 61, 223–232.CrossRefGoogle ScholarPubMed
Russwurm, A.D.A. (1978). Aberrations of British Butterflies. Classey,Faringdon.Google Scholar
Rusten, T.E., Cantera, R., Kafatos, F.C., and Barrio, R. (2002). The role of TGFb signaling in the formation of the dorsal nervous system is conserved between Drosophila and chordates. Development 129, 3575–3584.Google Scholar
Rusting, R.L. (2001). Hair: Why it grows, why it stops. Sci. Am. 284 #6, 70–79.CrossRefGoogle ScholarPubMed
Ruvinsky, I. and Gibson-Brown, J.J. (2000). Genetic and developmental bases of serial homology in vertebrate limb evolution. Development 127, 5233–5244.Google ScholarPubMed
Ruxton, G.D. (2002). The possible fitness benefits of striped coat coloration for zebra. Mammal Rev. 32 #4, 237–244.CrossRefGoogle Scholar
Saele, Ø., Smáradóttir, H., and Pittman, K. (2006). Twisted story of eye migration in flatfish. J. Morph. 267, 730–738.CrossRefGoogle ScholarPubMed
Saenko, S.V., Brakefield, P.M., and Beldade, P. (2010). Single locus affects embryonic segment polarity and multiple aspects of an adult evolutionary novelty. BMC Biol. 8, Article 111 (13 pp.).CrossRefGoogle ScholarPubMed
Saenko, S.V., French, V., Brakefield, P.M., and Beldade, P. (2008). Conserved developmental processes and the formation of evolutionary novelties: examples from butterfly wings. Phil. Trans. R. Soc. Lond. B 363, 1549–1555.CrossRefGoogle ScholarPubMed
Saenko, S.V., Marialva, M.S.P., and Beldade, P. (2011). Involvement of the conserved Hox gene Antennapedia in the development and evolution of a novel trait. EvoDevo 2, Article 9 (9 pp.).CrossRefGoogle ScholarPubMed
Saffo, M.B. (2005). Accidental elegance. Am. Scholar 74, 18–27. [See also Gee, H. (2013). The Accidental Species. University of Chicago Press, Chicago, IL.]Google Scholar
Salazar-Ciudad, I. (2006). On the origins of morphological disparity and its diverse developmental bases. BioEssays 28, 1112–1122.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I. (2007). On the origins of morphological variation, canalization, robustness, and evolvability. Integr. Comp. Biol. 47, 390–400.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I. (2010). Morphological evolution and embryonic developmental diversity in metazoa. Development 137, 531–539.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I. (2012). Tooth patterning and evolution. Curr. Opin. Gen. Dev. 22, 585–592.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I. and Jernvall, J. (2002). A gene network model accounting for development and evolution of mammalian teeth. PNAS 99 #12, 8116–8120.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I. and Jernvall, J. (2004). How different types of pattern formation mechanisms affect the evolution of form and development. Evol. Dev. 6, 6–16.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I. and Jernvall, J. (2010). A computational model of teeth and the developmental origins of morphological variation. Nature 464, 583–586.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I., Jernvall, J., and Newman, S.A. (2003). Mechanisms of pattern formation in development and evolution. Development 130, 2027–2037.CrossRefGoogle ScholarPubMed
Salazar-Ciudad, I. and Marín-Riera, M. (2013). Adaptive dynamics under development-based genotype-phenotype maps. Nature 497, 361–364.CrossRefGoogle ScholarPubMed
Sallé, J., Campbell, S.D., Gho, M., and Audibert, A. (2012). CycA is involved in the control of endoreplication dynamics in the Drosophila bristle lineage. Development 139, 547–557.CrossRefGoogle ScholarPubMed
Salzberg, A. and Bellen, H.J. (1996). Invertebrate versus vertebrate neurogenesis: variations on the same theme? Dev. Genet. 18, 1–10.3.0.CO;2-D>CrossRefGoogle ScholarPubMed
Sambrani, N., Hudry, B., Maurel-Zaffran, C., Zouaz, A., Mishra, R., Merabet, S., and Graba, Y. (2013). Distinct molecular strategies for Hox-mediated limb suppression in Drosophila: from cooperativity to dispensability/antagonism in TALE partnership. PLoS Genet. 9 #3, e1003307.CrossRefGoogle ScholarPubMed
Sampson, S.D. (1995). Horns, herds, and hierarchies. Nat. Hist. 104 #6, 36–40.Google Scholar
Sánchez, L. and Guerrero, I. (2001). The development of the Drosophila genital disc. BioEssays 23, 698–707.CrossRefGoogle ScholarPubMed
Sander, K. (2002). Ernst Haeckel’s ontogenetic recapitulation: irritation and incentive from 1866 to our time. Ann. Anat. 184, 523–533.CrossRefGoogle ScholarPubMed
Sander, P.M., Christian, A., Clauss, M., Fechner, R., Gee, C.T., Griebeler, E.-M., Gunga, H.-C., Hummel, J., Mallison, H., Perry, S.F., Preuschoft, H., Rauhut, O.W.M., Remes, K., Tütken, T., Wings, O., and Witzel, U. (2011). Biology of the sauropod dinosaurs: the evolution of gigantism. Biol. Rev. 86, 117–155.CrossRefGoogle ScholarPubMed
Sanderson, S.L. and Wasserug, R.J. (1990). Suspension-feeding vertebrates. Sci. Am. 262 #3, 96–101.CrossRefGoogle Scholar
Sane, S.P. and McHenry, M.J. (2009). The biomechanics of sensory organs. Integr. Comp. Biol. 49 #6, i8–i23.CrossRefGoogle Scholar
Sanger, T.J. (2012). The emergence of squamates as model systems for integrative biology. Evol. Dev. 14, 231–233.CrossRefGoogle ScholarPubMed
Sanger, T.J. and Gibson-Brown, J.J. (2004). The developmental basis of limb reduction and body elongation in squamates. Evolution 58, 2103–2106.CrossRefGoogle Scholar
Santana, S.E., Alfaro, J.L., and Alfaro, M.E. (2012). Adaptive evolution of facial colour patterns in Neotropical primates. Proc. R. Soc. Lond. B 279, 2204–2211.CrossRefGoogle ScholarPubMed
Sato, T., Hasegawa, Y., and Manabe, M. (2006). A new elasmosaurid plesiosaur from the Upper Cretaceous of Fukushima, Japan. Palaeontology 49, 467–484.CrossRefGoogle Scholar
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, 3633–3644.Google ScholarPubMed
Saunders, F. (2009). What’s all the flap about? Am. Sci. 97, 23–24.CrossRefGoogle Scholar
Savard, J., Marques-Souza, H., Aranda, M., and Tautz, D. (2006). A segmentation gene in Tribolium produces a polycistronic mRNA that codes for multiple conserved peptides. Cell 126, 559–569.CrossRefGoogle ScholarPubMed
Savic, D. (1995). Model of pattern formation in animal coatings. J. Theor. Biol. 172, 299–303.CrossRefGoogle Scholar
Savriama, Y. and Klingenberg, C.P. (2011). Beyond bilateral symmetry: geometric morphometric methods for any type of symmetry. BMC Evol. Biol. 11, Article 280 (24 pp.).CrossRefGoogle ScholarPubMed
Sawyer, G.J. and Deak, V. (2007). The Last Human: A Guide to Twenty-two Species of Extinct Humans. Yale University Press, New Haven, CT.Google Scholar
Sayed-Ahmed, A., Rudas, P., and Bartha, T. (2005). Partial cloning and localisation of leptin and its receptor in the one-humped camel (Camelus dromedarius). Vet. J. 170, 264–269.CrossRefGoogle Scholar
Scanlon, J.D. and Lee, M.S.Y. (2000). The Pleistocene serpent Wonambi and the early evolution of snakes. Nature 403, 416–420.CrossRefGoogle ScholarPubMed
Schier, A.F. and Needleman, D. (2009). Rise of the source-sink model. Nature 461, 480–481.CrossRefGoogle ScholarPubMed
Schierwater, B. and DeSalle, R. (2007). Can we ever identify the Urmetazoan? Integr. Comp. Biol. 47, 670–676.CrossRefGoogle ScholarPubMed
Schilling, T.F. and Knight, R.D. (2001). Origins of anteroposterior patterning and Hox gene regulation during chordate evolution. Phil. Trans. R. Soc. Lond. B 356, 1599–1613.CrossRefGoogle ScholarPubMed
Schilling, T.F., Nie, Q., and Lander, A.D. (2012). Dynamics and precision in retinoic acid morphogen gradients. Curr. Opin. Gen. Dev. 22, 562–569.CrossRefGoogle ScholarPubMed
Schilman, P.E., Kaiser, A., and Lighton, J.R.B. (2008). Breathe softly, beetle: Continuous gas exchange, water loss and the role of the subelytral space in the tenebrionid beetle, Eleodes obscura. J. Insect Physiol. 54, 192–203.CrossRefGoogle ScholarPubMed
Schlosser, G. and Wagner, G.P. (2004). Introduction: The modularity concept in developmental and evolutionary biology. In Modularity in Development and Evolution (Schlosser, G. and Wagner, G.P., eds.). University of Chicago Press, Chicago, IL, pp. 1–15.Google Scholar
Schmidt, E.R. (1978). Chromatophore development and cell interactions in the skin of Xiphophorine fish. Wilhelm Roux’s Arch. 184, 115–134.CrossRefGoogle ScholarPubMed
Schmidt, J., Francois, V., Bier, E., and Kimelman, D. (1995). Drosophila short gastrulation induces an ectopic axis in Xenopus: evidence for conserved mechanisms of dorsal–ventral patterning. Development 121, 4319–4328.Google ScholarPubMed
Schmidt-Nielsen, K. (1972). How Animals Work. Cambridge University Press, New York, NY.CrossRefGoogle Scholar
Schmidt-Nielsen, K. (1984). Scaling: Why Is Animal Size So Important?Cambridge University Press, New York, NY.CrossRefGoogle Scholar
Schmidt-Nielsen, K. (1998). The Camel’s Nose: Memoirs of a Curious Scientist. Island Press, Washington, DC.Google Scholar
Schneider, I. and Shubin, N. (2012). Making limbs from fins. Dev. Cell 23, 1121–1122.CrossRefGoogle ScholarPubMed
Schneider, R.A. (2005). Developmental mechanisms facilitating the evolution of bills and quills. J. Anat. 207, 563–573.CrossRefGoogle ScholarPubMed
Schneider, R.A. (2006). How to tweak a beak: molecular techniques for studying the evolution of size and shape in Darwin’s finches and other birds. BioEssays 29, 1–6.CrossRefGoogle Scholar
Schneider, R.A. and Helms, J.A. (2003). The cellular and molecular origins of beak morphology. Science 299, 565–568.CrossRefGoogle ScholarPubMed
Schoch, R.R. (2010). Riedl’s burden and the body plan: selection, constraint, and deep time. J. Exp. Zool. (Mol. Dev. Evol.) 314B, 1–10.CrossRefGoogle Scholar
Schoenebeck, J.J. and Ostrander, E.A. (2013). The genetics of canine skull shape variation. Genetics 193, 317–325.CrossRefGoogle ScholarPubMed
Schoenemann, B., Castellani, C., Clarkson, E.N.K., Haug, J.T., Maas, A., Haug, C., and Waloszek, D. (2012). The sophisticated visual system of a tiny Cambrian crustacean: analysis of a stalked fossil compound eye. Proc. Roy. Soc. Lond. B 279, 1335–1340.CrossRefGoogle ScholarPubMed
Schoenwolf, G.C., Bleyl, S.B., Brauer, P.R., and Francis-West, P.H. (2009). Larsen’s Human Embryology, 4th edn. Churchill Livingstone, Philadelphia, PA.Google Scholar
Schoenwolf, G.C. and Smith, J.L. (1990). Mechanisms of neurulation: traditional viewpoint and recent advances. Development 109, 243–270.Google ScholarPubMed
Scholtz, G. (2005). Homology and ontogeny: pattern and process in comparative developmental biology. Theory Biosci. 124, 121–143.CrossRefGoogle ScholarPubMed
Scholtz, G. and Edgecombe, G.D. (2006). The evolution of arthropod heads: reconciling morphological, developmental and palaentological evidence. Dev. Genes Evol. 216, 395–415.CrossRefGoogle Scholar
Schönbauer, C., Distler, J., Jährling, N., Radolf, M., Dodt, H.-U., Frasch, M., and Schnorrer, F. (2011). Spalt mediates an evolutionarily conserved switch to fibrillar muscle fate in insects. Nature 479, 406–409.CrossRefGoogle ScholarPubMed
Schoppmeier, M., Fischer, S., Schmitt-Engel, C., Löhr, U., and Klingler, M. (2009). An ancient anterior patterning system promotes caudal repression and head formation in Ecdysozoa. Curr. Biol. 19, 1811–1815.CrossRefGoogle ScholarPubMed
Schreiber, A.M. (2006). Asymmetric craniofacial remodeling and lateralized behavior in larval flatfish. J. Exp. Biol. 209, 610–621.CrossRefGoogle ScholarPubMed
Schreiber, A.M. (2013). Flatfish: an asymmetric perspective on metamorphosis. Curr. Top. Dev. Biol. 103, 167–194.CrossRefGoogle ScholarPubMed
Schröter, C. and Oates, A.C. (2010). Segment number and axial identity in a segmentation clock period mutant. Curr. Biol. 20, 1254–1258.CrossRefGoogle Scholar
Schubert, M., Yu, J.-K., Holland, N.D., Escriva, H., Laudet, V., and Holland, L.Z. (2005). Retinoic acid signaling acts via Hox1 to establish the posterior limit of the pharynx in the chordate amphioxus. Development 132, 61–73.CrossRefGoogle ScholarPubMed
Schulp, A.S., Mulder, E.W.A., and Schwenk, K. (2005). Did mosasaurs have forked tongues? Netherlands J. Geosci. 84, 359–371.CrossRefGoogle Scholar
Schutt, B. (2008). The curious, bloody lives of vampire bats. Nat. Hist. 117 #9, 22–27.Google Scholar
Schwanwitsch, B.N. (1924). On the ground-plan of wing-pattern in Nymphalids and certain other families of the rhopalocerous Lepidoptera. Proc. Zool. Soc. Lond. 34, 509–528.Google Scholar
Schwanwitsch, B.N. (1926). On the modes of evolution of the wing-pattern in Nymphalids and certain other families of the Rhopalocerous Lepidoptera. Proc. Zool. Soc. Lond. 33, 493–508 (plus 3 plates).Google Scholar
Schwanwitsch, B.N. (1929). Two schemes of the wing-pattern of butterflies. Zoomorphology 14, 36–58.Google Scholar
Schwanwitsch, B.N. (1930). Studies upon the wing-pattern of Prepona and Agrias, two genera of South-American Nyphalid butterflies. Acta Zool. 11, 289–410.CrossRefGoogle Scholar
Schwartz, J.H. (2007). Recognizing William Bateson’s contributions. Science 315, 1077.CrossRefGoogle ScholarPubMed
Schweitzer, R., Zelzer, E., and Volk, T. (2010). Connecting muscles to tendons: tendons and musculoskeletal development in flies and vertebrates. Development 137, 2807–2817.CrossRefGoogle ScholarPubMed
Schwenk, K. (1994). A utilitarian approach to evolutionary constraint. Zoology 98, 251–262.Google Scholar
Schwenk, K. (1994). Why snakes have forked tongues. Science 263, 1573–1577.CrossRefGoogle ScholarPubMed
Schwenk, K. (1995). Of tongues and noses: chemoreception in lizards and snakes. Trends Ecol. Evol. 10, 7–12.CrossRefGoogle ScholarPubMed
Schwenk, K. (1995). The serpent’s tongue. Nat. Hist. 104 #4, 48–55.Google Scholar
Schwenk, K., ed. Feeding: Form, Function and Evolution in Tetrapod Vertebrates. Academic Press, New York, NY.
Schwenk, K. (2001). Functional units and their selection. In The Character Concept in Evolutionary Biology (Wagner, G.P., ed.). Academic Press, San Diego, CA, pp. 165–198.CrossRefGoogle Scholar
Schwenk, K. and Wagner, G.P. (2004). The relativism of constraints on phenotypic evolution. In Phenotypic Integration. Studying the Ecology and Evolution of Complex Phenotypes. (Pigliucci, M. and Preston, K., eds.). Oxford University Press, New York, NY, pp. 390–408.Google Scholar
Scotland, R.W. (2010). Deep homology: a view from systematics. BioEssays 32, 438–449.CrossRefGoogle ScholarPubMed
Sears, K.E. (2011). Novel insights into the regulation of limb development from ‘natural’ mammalian mutants. BioEssays 33, 327–331.CrossRefGoogle ScholarPubMed
Sears, K.E., Behringer, R.R., Rasweiler, J.J., IV, and Niswander, L.A. (2006). Development of bat flight: morphologic and molecular evolution of bat wing digits. PNAS 103 #17, 6581–6586.CrossRefGoogle ScholarPubMed
Sears, K.E., Bormet, A.K., Rockwell, A., Powers, L.E., Cooper, L.N., and Wheeler, M.B. (2011). Developmental basis of mammalian digit reduction: a case study in pigs. Evol. Dev. 13, 533–541.CrossRefGoogle ScholarPubMed
Seher, T.D., Ng, C.S., Signor, S.A., Podlaha, O., Barmina, O., and Kopp, A. (2012). Genetic basis of a violation of Dollo’s Law: re-evolution of rotating sex combs in Drosophila bipectinata. Genetics 192, 1465–1475.CrossRefGoogle ScholarPubMed
Sehnal, F., Svácha, P., and Zrzavy, J. (1996). Evolution of insect metamorphosis. In Metamorphosis: Postembryonic Reprogramming of Gene Expression in Amphibian and Insect Cells (Gilbert, L.I., Tata, J.R., and Atkinson, B.G., eds.). Academic Press, New York, NY, pp. 3–58.CrossRefGoogle Scholar
Seifert, A.W., Kiama, S.G., Seifert, M.G., Goheen, J.R., Palmer, T.M., and Maden, M. (2012). Skin shedding and tissue regeneration in African spiny mice (Acomys). Nature 489, 561–565.CrossRefGoogle Scholar
Seifert, A.W., Monaghan, J.R., Smith, M.D., Pasch, B., Stier, A.C., Michonneau, F., and Maden, M. (2012). The influence of fundamental traits on mechanisms controlling appendage regeneration. Biol. Rev. 87, 330–345.CrossRefGoogle ScholarPubMed
Seifert, A.W., Monaghan, J.R., Voss, S.R., and Maden, M. (2012). Skin regeneration in adult axolotls: a blueprint for scar-free healing in vertebrates. PLoS ONE 7 #4, e32875.CrossRefGoogle ScholarPubMed
Sekimura, T., Maini, P.K., Nardi, J.B., Zhu, M., and Murray, J.D. (1998). Pattern formation in lepidopteran wings. Comments Theor. Biol. 5, 69–87.Google Scholar
Semenov, M.V., Habas, R., MacDonald, B.T., and He, X. (2007). Snapshot: Noncanonical Wnt signaling pathways. Cell 131, 1378.CrossRefGoogle ScholarPubMed
Senn, D.G. and Northcutt, R.G. (1973). The forebrain and midbrain of some squamates and their bearing on the origin of snakes. J. Morph. 140, 135–151.CrossRefGoogle ScholarPubMed
Senter, P. (2007). Necks for sex: sexual selection as an explanation for sauropod dinosaur neck elongation. J. Zool. 271, 45–53.CrossRefGoogle Scholar
Serena, M. (2000). Duck-billed platypus: Australia’s urban oddity. Natl. Geogr. 197 #4, 118–129.Google Scholar
Sereno, P.C. (1999). The evolution of dinosaurs. Science 284, 2137–2147.CrossRefGoogle ScholarPubMed
Shankland, M. and Seaver, E.C. (2000). Evolution of the bilaterian body plan: What have we learned from annelids? PNAS 97 #9, 4434–4437.CrossRefGoogle ScholarPubMed
Shapiro, M.D., Bell, M.A., and Kingsley, D.M. (2006). Parallel genetic origins of pelvic reduction in vertebrates. PNAS 103 #37, 13753–13758.CrossRefGoogle ScholarPubMed
Shapiro, M.D. and Domyan, E.T. (2013). Domestic pigeons. Curr. Biol. 23, R302–R303.CrossRefGoogle ScholarPubMed
Shapiro, M.D., Kronenberg, Z., Li, C., Domyan, E.T., Pan, H., Campbell, M., Tan, H., Huff, C.D., Hu, H., Vickrey, A.I., Nielsen, S.C.A., Stringham, S.A., Hu, H., Willerslev, E., Thomas, M., Gilbert, P., Yandell, M., Zhang, G., and Wang, J. (2013). Genomic diversity and evolution of the head crest in the rock pigeon. Science 339, 1063–1067.CrossRefGoogle ScholarPubMed
Shapiro, M.D., Marks, M.E., Peichel, C.L., Blackman, B.K., Nereng, K.S., Jónsson, B., Schluter, D., and Kingsley, D.M. (2004). Genetic and developmental basis of evolutionary pelvic reduction in threespine sticklebacks. Nature 428, 717–723.CrossRefGoogle ScholarPubMed
Shapiro, M.D., Shubin, N.H., and Downs, J.P. (2007). Limb diversity and digit reduction in reptilian evolution. In Fins into Limbs: Evolution, Development, and Transformation (Hall, B.K., ed.). University of Chicago Press, Chicago, IL, pp. 225–244.Google Scholar
Sharon, E., Marder, M., and Swinney, H.L. (2004). Leaves, flowers and garbage bags: making waves. Am. Sci. 92, 254–261.CrossRefGoogle Scholar
Shashidhara, L.S., Agrawal, N., Bajpai, R., Bharathi, V., and Sinha, P. (1999). Negative regulation of dorsoventral signaling by the homeotic gene Ultrabithorax during haltere development in Drosophila. Dev. Biol. 212, 491–502.CrossRefGoogle ScholarPubMed
Shaywitz, D.A. and Melton, D.A. (2005). The molecular biography of the cell. Cell 120, 729–731.CrossRefGoogle ScholarPubMed
Shbailat, S.J. and Abouheif, E. (2013). The wing-patterning network in the wingless castes of myrmicine and formicine ant species is a mix of evolutionarily labile and non-labile genes. J. Exp. Zool. (Mol. Dev. Evol.) 320B, 74–83.CrossRefGoogle Scholar
Shbailat, S.J., Khila, A., and Abouheif, E. (2010). Correlations between spatiotemporal changes in gene expression and apoptosis underlie wing polyphenism in the ant Pheidole morrisi. Evol. Dev. 12, 580–591.CrossRefGoogle ScholarPubMed
Shear, W.A. (1999). Millipeds. Am. Sci. 87, 232–239.CrossRefGoogle Scholar
Shear, W.A. (2012). An insect to fill the gap. Nature 488, 34–35.CrossRefGoogle ScholarPubMed
Sherman, P.W., Braude, S., and Jarvis, J.U.M. (1999). Litter sizes and mammary numbers of naked mole-rats: Breaking the one-half rule. J. Mammalogy 80, 720–733.CrossRefGoogle Scholar
Sheth, R., Grégoire, D., Dumouchel, A., Scotti, M., Pham, J.M.T., Nemec, S., Bastida, M.F., Ros, M.A., and Kmita, M. (2013). Decoupling the function of Hox and Shh in developing limb reveals multiple inputs of Hox genes on limb growth. Development 140, 2130–2138.CrossRefGoogle ScholarPubMed
Sheth, R., Marcon, L., Bastida, M.F., Junco, M., Quintana, L., Dahn, R., Kmita, M., Sharpe, J., and Ros, M.A. (2012). Hox genes regulate digit patterning by controlling the wavelength of a Turing-type mechanism. Science 338, 1476–1480.CrossRefGoogle ScholarPubMed
Shevtsova, E., Hansson, C., Janzen, D.H., and Kjaerandsen, J. (2011). Stable structural color patterns displayed on transparent insect wings. PNAS 108 #2, 668–673.CrossRefGoogle ScholarPubMed
Shidlovskiy, F.K., Kirillova, I.V., and Wood, J. (2011). Horns of the woolly rhinoceros Coelodonta antiquitatis (Blumenbach, 1799) in the Ice Age Museum collection (Moscow, Russia). Quarternary Internat. 255, 125–129.CrossRefGoogle Scholar
Shilo, B.-Z., Haskel-Ittah, M., Ben-Zvi, D., Schejter, E.D., and Barkai, N. (2013). Creating gradients by morphogen shuttling. Trends Genet. 29, 339–347.CrossRefGoogle ScholarPubMed
Shimeld, S.M., Gaunt, S.J., Coletta, P.L., Geada, A.M.C., and Sharpe, P.T. (1993). Spatial localisation of transcripts of the Hox-C6 gene. J. Anat. 183, 515–523.Google ScholarPubMed
Shimozono, S., Iimura, T., Kitaguchi, T., Higashijima, S., and Miyawaki, A. (2013). Visualization of an endogenous retinoic acid gradient across embryonic development. Nature 496, 363–366.CrossRefGoogle ScholarPubMed
Shine, R. and Wall, M. (2008). Interactions between locomotion, feeding, and bodily elongation during the evolution of snakes. Biol. J. Linnean Soc. 95, 293–304.CrossRefGoogle Scholar
Shingleton, A.W. and Frankino, W.A. (2013). New perspectives on the evolution of exaggerated traits. BioEssays 35, 100–107.CrossRefGoogle ScholarPubMed
Shingleton, A.W., Frankino, W.A., Flatt, T., Nijhout, H.F., and Emlen, D.J. (2007). Size and shape: the developmental regulation of static allometry in insects. BioEssays 29, 536–548.CrossRefGoogle ScholarPubMed
Shirai, L.T., Saenko, S.V., Keller, R.A., Jeronimo, M.A., Brakefield, P.M., Descimon, H., Wahlberg, N., and Beldade, P. (2012). Evolutionary history of the recruitment of conserved developmental genes in association to the formation and diversification of a novel trait. BMC Evol. Biol. 12, Article 21 (11 pp.).CrossRefGoogle ScholarPubMed
Shohat-Ophir, G., Kaun, K.R., Azanchi, R., and Heberlein, U. (2012). Sexual deprivation increases ethanol intake in Drosophila. Science 335, 1351–1355.CrossRefGoogle ScholarPubMed
Shoji, H. and Iwasa, Y. (2005). Labyrinthine versus stright-striped patterns generated by two-dimensional Turing systems. J. Theor. Biol. 237, 104–116.CrossRefGoogle ScholarPubMed
Shoji, H., Iwasa, Y., and Kondo, S. (2003). Stripes, spots, or reversed spots in two-dimensional Turing systems. J. Theor. Biol. 224, 339–350.CrossRefGoogle ScholarPubMed
Shoji, H., Mochizuki, A., Iwasa, Y., Hirata, M., Watanabe, T., Hioki, S., and Kondo, S. (2003). Origin of directionality in the fish stripe pattern. Dev. Dynamics 226, 627–633.CrossRefGoogle ScholarPubMed
Sholtis, S.J. and Noonan, J.P. (2010). Gene regulation and the origins of human biological uniqueness. Trends Genet. 26, 110–118.CrossRefGoogle ScholarPubMed
Shorrocks, B. and Croft, D.P. (2009). Necks and networks: a preliminary study of population structure in the reticulated giraffe (Giraffa camelopardalis reticulata de Winston). Afr. J. Ecol. 47, 374–381.CrossRefGoogle Scholar
Shou, S., Carlson, H.L., Perez, W.D., and Stadler, H.S. (2013). HOXA13 regulates Aldh1a2 expression in the autopod to facilitate interdigital programmed cell death. Dev. Dynamics 242, 687–698.CrossRefGoogle ScholarPubMed
Shroff, S., Joshi, M., and Orenic, T.V. (2007). Differential Delta expression underlies the diversity of sensory organ patterns among the legs of the Drosophila adult. Mechs. Dev. 124, 43–58.CrossRefGoogle ScholarPubMed
Shubin, N., Tabin, C., and Carroll, S. (1997). Fossils, genes and the evolution of animal limbs. Nature 388, 639–648.CrossRefGoogle ScholarPubMed
Shubin, N., Tabin, C., and Carroll, S. (2009). Deep homology and the origins of evolutionary novelty. Nature 457, 818–823.CrossRefGoogle ScholarPubMed
Shubin, N.H. and Alberch, P. (1986). A morphogenetic approach to the origin and basic organization of the tetrapod limb. Evol. Biol. 20, 319–387.Google Scholar
Shubin, N.H. and Dahn, R.D. (2004). Lost and found. Nature 428, 703–704.CrossRefGoogle ScholarPubMed
Sikes, N.E. (1999). Plio-Pleistocene floral context and habitat preferences of sympatric hominid species in East Africa. In African Biogeography, Climate Change, and Human Evolution (Bromage, T.G. and Schrenk, F., eds.). Oxford University Press, New York, NY, pp. 301–315.Google Scholar
Silverman, H.B. and Dunbar, M.J. (1980). Aggressive tusk use by the narwhal (Monodon monocerous L.). Nature 284, 57–58.CrossRefGoogle Scholar
Simmonds, A.J. and Bell, J.B. (1998). A genetic and molecular analysis of an invectedDominant mutation in Drosophila melanogaster. Genome 41, 381–390.Google ScholarPubMed
Simmonds, A.J., Brook, W.J., Cohen, S.M., and Bell, J.B. (1995). Distinguishable functions for engrailed and invected in anterior–posterior patterning in the Drosophila wing. Nature 376, 424–427.CrossRefGoogle ScholarPubMed
Simmons, N.B. (2008). Taking wing. Sci. Am. 299 #6, 96–103.CrossRefGoogle ScholarPubMed
Simmons, N.B., Seymour, K.L., Habersetzer, J., and Gunnell, G.F. (2008). Primitve Early Eocene bat from Wyoming and the evolution of flight and echolocation. Nature 451, 818–821.CrossRefGoogle Scholar
Simmons, R.E. and Scheepers, L. (1996). Winning by a neck: sexual selection in the evolution of giraffe. Am. Nat. 148, 771–786.CrossRefGoogle Scholar
Simões-Costa, M.S., Azambuja, A.P., and Xavier-Neto, J. (2008). The search for non-chordate retinoic acid signaling: lessons from chordates. J. Exp. Zool. (Mol. Dev. Evol.) 310B, 54–72.CrossRefGoogle Scholar
Simon, C.A. (1982). Masters of the tongue flick. Nat. Hist. 91 #9, 58–67.Google Scholar
Simonnet, F. and Moczek, A.P. (2011). Conservation and diversification of gene function during mouthpart development in Onthophagus beetles. Evol. Dev. 13, 280–289.CrossRefGoogle ScholarPubMed
Simpson, S.J., Sword, G.A., and Lo, N. (2011). Polyphenism in insects. Curr. Biol. 21, R738–R749.CrossRefGoogle ScholarPubMed
Sinclair, R. (1998). Male pattern androgenetic alopecia. BMJ 317, 865–869.CrossRefGoogle ScholarPubMed
Sinclair, R. (2007). Female pattern hair loss, dandruff and greying of hair in twins. J. Investig. Dermatol. 127, 2680.Google Scholar
Singh, A., Kango-Singh, M., Parthasarathy, R., and Gopinathan, K.P. (2007). Larval legs of mulberry silkworm Bombyx mori are prototypes for the adult legs. Genesis 45, 169–176.CrossRefGoogle ScholarPubMed
Singh, N.D., Larracuente, A.M., Sackton, T.B., and Clark, A.G. (2009). Comparative genomics on the Drosophila phylogenetic tree. Annu. Rev. Ecol. Evol. Syst. 40, 459–480.CrossRefGoogle Scholar
Singh, N.P. and Mishra, R.K. (2008). A double-edged sword to force posterior dominance of Hox genes. BioEssays 30, 1058–1061.CrossRefGoogle ScholarPubMed
Sinigaglia, C., Busengdal, H., Leclère, L., Technau, U., and Rentzsch, F. (2013). The bilaterian head patterning gene six3/6 controls aboral domain development in a cnidarian. PLoS Biol. 11 #2, e1001488.CrossRefGoogle Scholar
Sipper, M. (1995). Studying artificial life using a simple, general cellular model. Artif. Life 2, 1–35.CrossRefGoogle Scholar
Sire, J.-Y., Delgado, S.C., and Girondot, M. (2008). Hen’s teeth with enamel cap: from dream to impossibility. BMC Evol. Biol. 8, Article 246 (11 pp.).CrossRefGoogle ScholarPubMed
Sites, J.W., Jr., Reeder, T.W., and Wiens, J.J. (2011). Phylogenetic insights on evolutionary novelties in lizards and snakes: sex, birth, bodies, niches, and venom. Annu. Rev. Ecol. Evol. Syst. 42, 227–244.CrossRefGoogle Scholar
Sivak, J.G. (1977). The role of the spectacle in the visual optics of the snake eye. Vision Res. 17, 293–298.CrossRefGoogle ScholarPubMed
Sivinski, J. (1997). Ornamentation in the Diptera. Florida Entomologist 80, 142–164.CrossRefGoogle Scholar
Skinner, A., Lee, M.S.Y., and Hutchinson, M.N. (2008). Rapid and repeated limb loss in a clade of scincid lizards. BMC Evol. Biol. 8, e310.CrossRefGoogle Scholar
Slack, F. and Ruvkun, G. (1997). Temporal pattern formation by heterochronic genes. Annu. Rev. Genet. 31, 611–634.CrossRefGoogle ScholarPubMed
Slack, J.M.W., Holland, P.W.H., and Graham, C.F. (1993). The zootype and the phylotypic stage. Nature 361, 490–492.CrossRefGoogle ScholarPubMed
Smith, A.B. (2008). Deuterostomes in a twist: the origins of a radical new body plan. Evol. Dev. 10, 493–503.CrossRefGoogle Scholar
Smith, C.F., Schwenk, K., Earley, R.L., and Schuett, G.W. (2008). Sexual size dimorphism of the tongue in a North American pitviper. J. Zool. 274, 367–374.CrossRefGoogle Scholar
Smith, J.V., Braun, E.L., and Kimball, R.T. (2013). Ratite nonmonophyly: independent evidence from 40 novel loci. Syst. Biol. 62, 35–49.CrossRefGoogle ScholarPubMed
Smith, K.K. (2003). Time’s arrow: heterochrony and the evolution of development. Int. J. Dev. Biol. 47, 613–621.Google ScholarPubMed
Smith, R. (2012). Cheetahs on the edge. Natl. Geogr. 222 #5, 110–123.Google Scholar
Smits, P.D. and Evans, A.R. (2012). Functional constraints on tooth morphology in carnivorous mammals. BMC Evol. Biol. 12, Article 146 (11 pp.).CrossRefGoogle ScholarPubMed
Sniegowski, P.D. and Murphy, H.A. (2006). Evolvability. Curr. Biol. 16, R831–R834.CrossRefGoogle ScholarPubMed
Snodgrass, R.E. (1935). Principles of Insect Morphology. McGraw-Hill, New York, NY.Google Scholar
Snodgrass, R.E. (1954). Insect metamorphosis. Smithsonian Misc. Coll. Pub. 4144 122 #9, iii–124.Google Scholar
Snodgrass, R.E. (1958). Evolution of arthropod mechanisms. Smithsonian Misc. Coll. Pub. 4347 138 #2, i–77.Google Scholar
Snodgrass, R.E. (1961). The caterpillar and the butterfly. Smithsonian Misc. Coll. Pub. 4472 143 #6, 1–51.Google Scholar
Sokol, N.S. (2012). Small temporal RNAs in animal development. Curr. Opin. Gen. Dev. 22, 368–373.CrossRefGoogle ScholarPubMed
Soligo, C. and Müller, A.E. (1999). Nails and claws in primate evolution. J. Hum. Evol. 36, 97–114.CrossRefGoogle ScholarPubMed
Solounias, N. (1999). The remarkable anatomy of the giraffe’s neck. J. Zool. 247, 257–268.CrossRefGoogle Scholar
Sommer, R.J. (2008). Homology and the hierarchy of biological systems. BioEssays 30, 653–658.CrossRefGoogle ScholarPubMed
Sommer, R.J. (2009). The future of evo-devo: model systems and evolutionary theory. Nature Rev. Genet. 10, 416–422.CrossRefGoogle ScholarPubMed
Sommer, S., Whittington, C.M., and Wilson, A.B. (2012). Standardised classification of pre-release development in male-brooding pipefish, seahorses, and seadragons (Family Syngnathidae). BMC Dev. Biol. 12, Article 39 (6 pp.).CrossRefGoogle Scholar
Soshnikova, N., Dewaele, R., Janvier, P., Krumlauf, R., and Duboule, D. (2013). Duplications of Hox gene clusters and the emergence of vertebrates. Dev. Biol. 378, 194–199.CrossRefGoogle ScholarPubMed
Soto, I.M., Carreira, V.P., Soto, E.M., Márquez, F., Lipko, P., and Hasson, E. (2013). Rapid divergent evolution of male genitalia among populations of Drosophila buzzatii. Evol. Biol. 40, 395–407.CrossRefGoogle Scholar
Spéder, P., Ádám, G., and Noselli, S. (2006). Type ID unconventional myosin controls left-right asymmetry in Drosophila. Nature 440, 803–807.CrossRefGoogle ScholarPubMed
Spéder, P. and Noselli, S. (2007). Left-right asymmetry: class I myosins show the direction. Curr. Opin. Cell Biol. 19, 82–87.CrossRefGoogle Scholar
Spitz, F. and Furlong, E.E.M. (2012). Transcription factors: from enhancer binding to developmental control. Nature Rev. Gen. 13, 613–626.CrossRefGoogle ScholarPubMed
Spoon, J.M. (2001). Situs inversus totalis. Neonatal Netw. 20, 59–63.CrossRefGoogle ScholarPubMed
Springer, K., Brown, M., and Stulberg, D.L. (2003). Common hair loss disorders. Am. Fam. Phys. 68, 93–102.Google ScholarPubMed
Springer, M.S., Kirsch, J.A.W., and Case, J.A. (1997). The chronicle of marsupial evolution. In Molecular Evolution and Adaptive Radiation (Givnish, T.J. and Sytsma, K.J., eds.). Cambridge University Press, New York, NY, pp. 129–157.Google Scholar
Sprinzak, D., Lakhanpal, A., LeBon, L., Santat, L.A., Fontes, M.E., Anderson, G.A., Garcia-Ojalvo, J., and Elowitz, M.B. (2010). Cis-interactions between Notch and Delta generate mutually exclusive signalling states. Nature 465, 86–90.CrossRefGoogle ScholarPubMed
Srygley, R.B. and Thomas, A.L.R. (2002). Unconventional lift-generating mechanisms in free-flying butterflies. Nature 420, 660–664.CrossRefGoogle ScholarPubMed
Stafford, P. (2000). Snakes. Smithsonian Institution Press, Washington, DC.Google Scholar
Stahl, R., Walcher, T., Romero, C.D.J., Pilz, G.A., Cappello, S., Irmler, M., Sanz-Aquela, J.M., Beckers, J., Blum, R., Borrell, V., and Götz, M. (2013). Trnp1 regulates expansion and folding of the mammalian cerebral cortex by control of radial glial fate. Cell 153, 535–549.CrossRefGoogle ScholarPubMed
Standen, E.M. (2008). Pelvic fin locomotion function in fishes: three-dimensional kinematics in rainbow trout (Oncorhynchus mykiss). J. Exp. Biol. 211, 2931–2942.CrossRefGoogle Scholar
Stanek, C. and Hantak, E. (1986). Bilateral atavistic polydactyly in a colt and its dam. Equine Vet. J. 18, 76–79.CrossRefGoogle Scholar
Stauber, M., Prell, A., and Schmidt-Ott, U. (2002). A single Hox3 gene with composite bicoid and zerknüllt expression characteristics in non-Cyclorrhaphan flies. PNAS 99 #1, 274–279.CrossRefGoogle ScholarPubMed
Stearns, S.C. (2002). Less would have been more. Evolution 56, 2339–2345.Google Scholar
Stebbins, G.L. (1983). Mosaic evolution: an integrating principle for the modern synthesis. Experientia 39, 823–834.CrossRefGoogle ScholarPubMed
Stegmann, U.E. (2005). John Maynard Smith’s notion of animal signals. Biol. Philos. 20, 1011–1025.CrossRefGoogle Scholar
Steinberg, M.S. (2007). Differential adhesion in morphogenesis: a modern view. Curr. Opin. Gen. Dev. 17, 281–286.CrossRefGoogle ScholarPubMed
Stern, C. (1968). Genetic Mosaics and Other Essays. Harvard University Press, Cambridge, MA.CrossRefGoogle Scholar
Stern, D. (2006). Morphing into shape. Science 313, 50–51.CrossRefGoogle ScholarPubMed
Stern, D.L. (1998). A role of Ultrabithorax in morphological differences between Drosophila species. Nature 396, 463–466.CrossRefGoogle ScholarPubMed
Stern, D.L. (2000). Evolutionary developmental biology and the problem of variation. Evolution 54, 1079–1091.CrossRefGoogle ScholarPubMed
Stern, D.L. (2003). The Hox gene Ultrabithorax modulates the shape and size of the third leg of Drosophila by influencing diverse mechanisms. Dev. Biol. 256, 355–366.CrossRefGoogle ScholarPubMed
Stern, D.L. and Emlen, D.J. (1999). The developmental basis for allometry in insects. Development 126, 1091–1101.Google ScholarPubMed
Stevens, M. (2005). The role of eyespots as anti-predator mechanisms, principally demonstrated in the Lepidoptera. Biol. Rev. 80, 573–588.CrossRefGoogle ScholarPubMed
Stevens, M. (2007). Predator perception and the interrelation between different forms of protective coloration. Proc. R. Soc. Lond. B 274, 1457–1464.CrossRefGoogle ScholarPubMed
Stevens, M., Hardman, C.J., and Stubbins, C.L. (2008). Conspicuousness, not eye mimicry, makes “eyespots” effective antipredator signals. Behav. Ecol. 19, 525–531.CrossRefGoogle Scholar
Stevens, M. and Merilaita, S., eds. (2011). Animal Camouflage: Mechanisms and Function. Cambridge University Press, New York, NY.CrossRef
Stevens, M., Winney, I.S., Cantor, A., and Graham, J. (2009). Outline and surface disruption in animal camouflage. Proc. R. Soc. Lond. B 276, 781–786.CrossRefGoogle ScholarPubMed
Stevens, M., Yule, D.H., and Ruxton, G.D. (2008). Dazzle coloration and prey movement. Proc. R. Soc. Lond. B 275, 2639–2643.CrossRefGoogle ScholarPubMed
Stewart, A.J. and Plotkin, J.B. (2012). Why transcription factor binding sites are ten nucleotides long. Genetics 192, 973–985.CrossRefGoogle ScholarPubMed
Stiassny, M.L.J. (2003). Atavism. In Keywords and Concepts in Evolutionary Developmental Biology (Hall, B.K. and Olson, W.M., eds.). Harvard University Press, Cambridge, MA, pp. 10–14.Google Scholar
Stock, D.W. (2001). The genetic basis of modularity in the development and evolution of the vertebrate dentition. Phil. Trans. Roy. Soc. Lond. B 356, 1633–1653.CrossRefGoogle ScholarPubMed
Stock, G.B. and Bryant, S.V. (1981). Studies of digit regeneration and their implications for theories of development and evolution of vertebrate limbs. J. Exp. Zool. 216, 423–433.CrossRefGoogle ScholarPubMed
Stockard, C.R. (1930). The presence of a factorial basis for characters lost in evolution: the atavistic reappearance of digits in mammals. Am. J. Anat. 45, 345–375.CrossRefGoogle Scholar
Stocum, D.L. (2012). Regenerative Biology and Medicine, 2nd edn. Academic Press, New York, NY.Google Scholar
Stocum, D.L. and Cameron, J.A. (2011). Looking proximally and distally: 100 years of limb regeneration and beyond. Dev. Dynamics 240, 943–968.CrossRefGoogle ScholarPubMed
Stoehr, A.M., Walker, J.F., and Monteiro, A. (2013). Spalt expression and the development of melanic color patterns in pierid butterflies. EvoDevo 4, Article 6 (11 pp.).CrossRefGoogle ScholarPubMed
Stoick-Cooper, C.L., Moon, R.T., and Weidinger, G. (2007). Advances in signaling in vertebrate regeneration as a prelude to regenerative medicine. Genes Dev. 21, 1292–1315.CrossRefGoogle ScholarPubMed
Stokely, P.S. (1947). Limblessness and correlated changes in the girdles of a comparative morphological series of lizards. Am. Midland Nat. 38, 725–754.CrossRefGoogle Scholar
Stokes, M.D. and Holland, N. (1998). The lancelet. Am. Sci. 86 #6, 552–560.CrossRefGoogle Scholar
Stokstad, E. (2001). Early tyrannosaur was small but well armed. Science 292, 1278–1279.CrossRefGoogle ScholarPubMed
Stokstad, E. (2004). T. rex clan evolved head first. Science 306, 211.CrossRefGoogle Scholar
Stokstad, E. (2007). Jaw shows platypus goes way back. Science 318, 1237.Google ScholarPubMed
Stopper, G.F. and Wagner, G.P. (2005). Of chicken wings and frog legs: a smorgasbord of evolutionary variation in mechanisms of tetrapod limb development. Dev. Biol. 288, 21–39.CrossRefGoogle ScholarPubMed
Strausfeld, N.J. and Hirth, F. (2013). Deep homology of arthropod central complex and vertebrate basal ganglia. Science 340, 157–161.CrossRefGoogle ScholarPubMed
Strugnell, J., Norman, M., Drummond, A.J., and Cooper, A. (2004). Neotenous origins for pelagic octopuses. Curr. Biol. 14, R300–R301.CrossRefGoogle ScholarPubMed
Stuart-Fox, D. and Moussalli, A. (2008). Selection for social signalling drives the evolution of chameleon colour change. PLoS Biol. 6 #1, e25.CrossRefGoogle ScholarPubMed
Stuart-Fox, D. and Moussalli, A. (2009). Camouflage, communication and thermoregulation: lessons from colour changing organisms. Phil. Trans. R. Soc. Lond. B 364, 463–470.CrossRefGoogle ScholarPubMed
Stumpke, H. (1981). The Snouters: Form and Life of the Rhinogrades. University of Chicago Press, Chicago, IL.Google Scholar
Sudarsan, V., Anant, S., Guptan, P., VijayRaghavan, K., and Skaer, H. (2001). Myoblast diversification and ectodermal signaling in Drosophila. Dev. Cell 1, 829–839.CrossRefGoogle ScholarPubMed
Sues, H.-D. (1991). Venom-conducting teeth in a Triassic reptile. Nature 351, 141–143.CrossRefGoogle Scholar
Sumbre, G., Fiorito, G., Flash, T., and Hochner, B. (2005). Motor control of flexible octopus arms. Nature 433, 595–596.CrossRefGoogle ScholarPubMed
Sumbre, G., Fiorito, G., Flash, T., and Hochner, B. (2006). Octopuses use a human-like strategy to control precise point-to-point arm movements. Curr. Biol. 16, 767–772.CrossRefGoogle ScholarPubMed
Summers, A. (2002). Fast food joints. Nat. Hist. 111 #4, 84–85.Google Scholar
Summers, A. (2008). Jaws two. Nat. Hist. 117 #1, 22–23.Google Scholar
Sun, Y., Kanekar, S.L., Vetter, M.L., Gorski, S., Jan, Y.-N., Glaser, T., and Brown, N.L. (2003). Conserved and divergent functions of Drosophila atonal, amphibian, and mammalian Ath5 genes. Evol. Dev. 5, 532–541. [See also Jarman, A.P., and Groves, A.K. (2013). The role of Atonal transcription factors in the development of mechanosensitive cells. Semin. Cell Dev. Biol. 24, 438–447.]CrossRefGoogle ScholarPubMed
Sundaram, M.V. (2005). The love-hate relationship between Ras and Notch. Genes Dev. 19, 1825–1839.CrossRefGoogle ScholarPubMed
Superina, M. and Loughry, W.J. (2012). Life on the half-shell: consequences of a carapace in the evolution of armadillos (Xenarthra: Cingulata). J. Mamm. Evol. 19, 217–224.CrossRefGoogle Scholar
Sustaita, D., Pouydebat, E., Manzano, A., Abdala, V., Hertel, F., and Herrel, A. (2013). Getting a grip on tetrapod grasping: form, function, and evolution. Biol. Rev. 88, 380–405.CrossRefGoogle ScholarPubMed
Suzanne, M., Petzoldt, A.G., Spéder, P., Coutelis, J.-B., Steller, H., and Noselli, S. (2010). Coupling of apoptosis and L/R patterning controls stepwise organ looping. Curr. Biol. 20, 1773–1778.CrossRefGoogle ScholarPubMed
Suzuki, N., Hirata, M., and Kondo, S. (2003). Traveling stripes on the skin of a mutant mouse. PNAS 100 #17, 9680–9685.CrossRefGoogle ScholarPubMed
Suzuki, T., Washio, Y., Aritaki, M., Fujinami, Y., Shimizu, D., Uji, S., and Hashimoto, H. (2009). Metamorphic pitx2 expression in the left habenula correlated with lateralization of eye-sidedness in flounder. Develop. Growth Differ. 51, 797–808.CrossRefGoogle ScholarPubMed
Suzuki, Y. and Palopoli, M.F. (2001). Evolution of insect abdominal appendages: are prolegs homologous or convergent traits? Dev. Genes Evol. 211, 486–492.CrossRefGoogle ScholarPubMed
Swalla, B.J. (2006). Building divergent body plans with similar genetic pathways. Heredity 97, 235–243.CrossRefGoogle ScholarPubMed
Swanson, C.I., Schwimmer, D.B., and Barolo, S. (2011). Rapid evolutionary rewiring of a structurally constrained eye enhancer. Curr. Biol. 21, 1186–1196.CrossRefGoogle ScholarPubMed
Szeto, D.P., Rodriguez-Esteban, C., Ryan, A.K., O’Connell, S.M., Liu, F., Kioussi, C., Gleiberman, A.S., Izpisúa-Belmonte, J.C., and Rosenfeld, M.G. (1999). Role of the Bicoid-related homeodomain factor Pitx1 in specifying hindlimb morphogenesis and pituitary development. Genes Dev. 13, 484–494.CrossRefGoogle ScholarPubMed
Tabata, T. and Takei, Y. (2004). Morphogens, their identification and regulation. Development 131, 703–712.CrossRefGoogle ScholarPubMed
Tabin, C.J. (1992). Why we have (only) five fingers per hand: Hox genes and the evolution of paired limbs. Development 116, 289–296.Google ScholarPubMed
Tabin, C.J., Carroll, S.B., and Panganiban, G. (1999). Out on a limb: parallels in vertebrate and invertebrate limb patterning and the origin of appendages. Am. Zool. 39, 650–663.CrossRefGoogle Scholar
Taborsky, M. and Taborsky, B. (1993). The kiwi’s parental burden. Nat. Hist. 102 #12, 50–57.Google Scholar
Tajiri, R., Misaki, K., Yonemura, S., and Hayashi, S. (2011). Joint morphology in the insect leg: evolutionary history inferred from Notch loss-of-function phenotypes in Drosophila. Development 138, 4621–4626.CrossRefGoogle ScholarPubMed
Takahashi, G. and Kondo, S. (2008). Melanophores in the stripes of adult zebrafish do not have the nature to gather, but disperse when they have the space to move. Pigment Cell Melanoma Res. 21, 677–686.CrossRefGoogle ScholarPubMed
Takahashi, M., Arita, H., Hiraiwa-Hasegawa, M., and Hasegawa, T. (2008). Peahens do not prefer peacocks with more elaborate trains. Anim. Behav. 75, 1209–1219. [See also Yorzinski, J.L., et al. (2013). Through their eyes: selective attention in peahens during courtship. J. Exp. Biol. 216: 3035–3046.]CrossRefGoogle Scholar
Takashima, S., Mkrtchyan, M., Younossi-Hartenstein, A., Merriam, J.R., and Hartenstein, V. (2008). The behaviour of Drosophila adult hindgut stem cells is controlled by Wnt and Hh signalling. Nature 454, 651–655.CrossRefGoogle ScholarPubMed
Takashima, Y., Ohtsuka, T., González, A., Miyachi, H., and Kageyama, R. (2011). Intronic delay is essential for oscillatory expression in the segmentation clock. PNAS 108 #8, 3300–3305.CrossRefGoogle ScholarPubMed
Tanaka, E.M. (2003). Regeneration: If they can do it, why can’t we? Cell 113, 559–562.CrossRefGoogle Scholar
Tanaka, E.M. (2012). Skin, heal thyself. Nature 489, 508–510.CrossRefGoogle Scholar
Tanaka, E.M. and Reddien, P.W. (2011). The cellular basis for animal regeneration. Dev. Cell 21, 172–185.CrossRefGoogle ScholarPubMed
Tanaka, K., Barmina, O., and Kopp, A. (2009). Distinct developmental mechanisms underlie the evolutionary diversification of Drosophila sex combs. PNAS 106 #12, 4764–4769.CrossRefGoogle ScholarPubMed
Tanaka, K., Barmina, O., Sanders, L.E., Arbeitman, M.N., and Kopp, A. (2011). Evolution of sex-specific traits through changes in HOX-dependent doublesex expression. PLoS Biol. 9 #8, e1001131.CrossRefGoogle ScholarPubMed
Tanaka, K. and Truman, J.W. (2005). Development of the adult leg epidermis in Manduca sexta: contribution of different larval cell populations. Dev. Genes Evol. 215, 78–89.CrossRefGoogle ScholarPubMed
Tanaka, K. and Truman, J.W. (2007). Molecular patterning mechanism underlying metamorphosis of the thoracic leg in Manduca sexta. Dev. Biol. 305, 539–550.CrossRefGoogle ScholarPubMed
Tanaka, M. (2011). Revealing the mechanisms of the rostral shift of pelvic fins among teleost fishes. Evol. Dev. 13, 382–390.CrossRefGoogle ScholarPubMed
Taniguchi, K., Maeda, R., Ando, T., Okumura, T., Nakazawa, N., Hatori, R., Nakamura, M., Hozumi, S., Fujiwara, H., and Matsuno, K. (2011). Chirality in planar cell shape contributes to left-right asymmetric epithelial morphogenesis. Science 333, 339–341.CrossRefGoogle ScholarPubMed
Taylor, G.K. (2001). Mechanics and aerodynamics of insect flight control. Biol. Rev. 76, 449–471.CrossRefGoogle ScholarPubMed
Taylor, H.L., Cole, C.J., Dessauer, H.C., and Parker, E.D., Jr. (2003). Congruent patterns of genetic and morphological variation in the parthenogenetic lizard Aspidoscelis tesselata (Squamata: Teiidae) and the origins of color pattern classes and genotypic clones in eastern New Mexico. Am. Mus. Novitates 3424, 1–40.2.0.CO;2>CrossRefGoogle Scholar
Taylor, I.W. (2008). Snapshot: The TGFβ pathway interactome. Cell 133, 378.CrossRefGoogle ScholarPubMed
Taylor, M.P., Hone, D.W.E., Wedel, M.J., and Naish, D. (2011). The long necks of sauropods did not evolve primarily through sexual selection. J. Zool. 285, 150–161.CrossRefGoogle Scholar
Taylor, M.P. and Wedel, M.J. (2013). Why sauropods had long necks; and why giraffes have short necks. PeerJ 1, e36. .CrossRefGoogle ScholarPubMed
Tchernov, E., Rieppel, O., Zaher, H., Polcyn, M.J., and Jacobs, L.L. (2000). A fossil snake with legs. Science 287, 2010–2012.CrossRefGoogle Scholar
Technau, U. and Steele, R.E. (2011). Evolutionary crossroads in developmental biology: Cnidaria. Development 138, 1447–1458.CrossRefGoogle ScholarPubMed
Telford, M.J. (2007). A single origin of the central nervous system? Cell 129, 237–239.CrossRefGoogle ScholarPubMed
Telford, M.J. (2009). Animal evolution: once upon a time. Curr. Biol. 19, R339–R341.CrossRefGoogle Scholar
ten Berge, D., Brouwer, A., El Bahi, S., Guenet, J.-L., Robert, B., and Meijlink, F. (1998). Mouse Alx3: an aristaless-like homeobox gene expressed during embryogenesis in ectomesenchyme and lateral plate mesoderm. Dev. Biol. 199, 11–25.CrossRefGoogle ScholarPubMed
ten Broek, C.M.A., Bakker, A.J., Varela-Lasheras, I., Bugiani, M., Van Dongen, S., and Galis, F. (2012). Evo-devo of the human vertebral column: on homeotic transformations, pathologies and prenatal selection. Evol. Biol. 39, 456–471.CrossRefGoogle ScholarPubMed
Tennant, A. (1984). The Snakes of Texas. Texas Monthly Press, Austin, TX.Google Scholar
Teotónio, H. and Rose, M.R. (2001). Reversible evolution. Evolution 55, 653–660.CrossRefGoogle Scholar
Terra, W.R. (1990). Evolution of digestive systems of insects. Annu. Rev. Entomol. 35, 181–200.CrossRefGoogle Scholar
Teske, P.R. and Beheregaray, L.B. (2009). Evolution of seahorses’ upright posture was linked to Oligocene expansion of seagrass habitats. Biol. Lett. 5, 521–523.CrossRefGoogle ScholarPubMed
Theissen, G. (2006). The proper place of hopeful monsters in evolutionary biology. Theory Biosci. 124, 349–369.CrossRefGoogle ScholarPubMed
Théry, M. and Gomez, D. (2010). Insect colours and visual appearance in the eyes of their predators. Adv. Insect Physiol. 38, 267–353.CrossRefGoogle Scholar
Theveneau, E. and Mayor, R. (2012). Neural crest delamination and migration: from epithelium-to-mesenchyme transition to collective cell migration. Dev. Biol. 366, 34–54.CrossRefGoogle ScholarPubMed
Thewissen, J.G.M. (2007). Aquatic adaptations in the limbs of amniotes. In Fins into Limbs: Evolution, Development, and Transformation (Hall, B.K., ed.). University of Chicago Press, Chicago, IL, pp. 310–322.Google Scholar
Thewissen, J.G.M., Cohn, M.J., Stevens, L.S., Bajpai, S., Heyning, J., and Horton, W.E., Jr. (2006). Developmental basis for hind-limb loss in dolphins and origin of the cetacean bodyplan. PNAS 103 #22, 8414–8418.CrossRefGoogle ScholarPubMed
Thewissen, J.G.M., Cooper, L.N., Clementz, M.T., Bajpai, S., and Tiwari, B.N. (2007). Whales originated from aquatic artiodactyls in the Eocene epoch of India. Nature 450, 1190–1194.CrossRefGoogle Scholar
Thewissen, J.G.M., Cooper, L.N., George, J.C., and Bajpai, S. (2009). From land to water: the origin of whales, dolphins, and porpoises. Evo. Edu. Outreach 2, 272–288.CrossRefGoogle Scholar
Thewissen, J.G.M., Williams, E.M., Roe, L.J., and Hussain, S.T. (2001). Skeletons of terrestrial cetaceans and the relationship of whales to artiodactyls. Nature 413, 277–281.CrossRefGoogle ScholarPubMed
Thistle, R., Cameron, P., Ghorayshi, A., Dennison, L., and Scott, K. (2012). Contact chemoreceptors mediate male-male repulsion and male-female attraction during Drosophila courtship. Cell 149, 1140–1151.CrossRefGoogle ScholarPubMed
Thomas, R.D.K. and Reif, W.-E. (1993). The skeleton space: a finite set of organic designs. Evolution 47, 341–360.CrossRefGoogle ScholarPubMed
Thompson, B.J. (2013). Cell polarity: models and mechanisms from yeast, worms, and flies. Development 140, 13–21.CrossRefGoogle ScholarPubMed
Thompson, C.W. (2013). Implications of hybridization between the Rio Grande ground squirrel (Ictidomys parvidens) and the thirteen-lined ground squirrel (I. tridecemlineatus). Ph.D. dissertation. Department of Biological Sciences, Texas Tech University, Lubbock, TX.Google Scholar
Thompson, D.W. (1917). On Growth and Form. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Thor, S. and Thomas, J.B. (2002). Motor neuron specification in worms, flies and mice: conserved and “lost” mechanisms. Curr. Opin. Gen. Dev. 12, 558–564.CrossRefGoogle ScholarPubMed
Thorington, R.W., Jr., Koprowski, J.L., Steele, M.A., and Whatton, J.F. (2012). Squirrels of the World. Johns Hopkins University Press, Baltimore, MD.Google Scholar
Thorington, R.W., Jr., Pitassy, D., and Jansa, S.A. (2002). Phylogenies of flying squirrels (Pteromyinae). J. Mamm. Evol. 9, 99–135.CrossRefGoogle Scholar
Thorington, R.W., Jr. and Santana, E.M. (2007). How to make a flying squirrel: Glaucomys anatomy in phylogenetic perspective. J. Mammalogy 88, 882–896.CrossRefGoogle Scholar
Thurber, A.R., Jones, W.J., and Schnabel, K. (2011). Dancing for food in the deep sea: bacterial farming by a new species of yeti crab. PLoS ONE 6 #11, e26243.CrossRefGoogle ScholarPubMed
Tobias, J.A., Montgomerie, R., and Lyon, B.E. (2012). The evolution of female ornaments and weaponry: social selection, sexual selection and ecological competition. Phil. Trans. R. Soc. Lond. B 367, 2274–2293.CrossRefGoogle ScholarPubMed
Tobin, A.J. and Dusheck, J. (2004). Asking About Life, 3rd edn. Brooks/Cole, Belmont, CA.Google Scholar
Toda, H., Zhao, X., and Dickson, B.J. (2012). The Drosophila female aphrodisiac pheromone activates ppk23+ sensory neurons to elicit male courtship behavior. Cell Reports 1, 599–607.CrossRefGoogle ScholarPubMed
Toh, Y. (1985). Structure of campaniform sensilla on the haltere of Drosophila prepared by cryofixation. J. Ultrastruct. Res. 93, 92–100.CrossRefGoogle Scholar
Tokita, M., Chaeychomsri, W., and Siruntawineti, J. (2012). Developmental basis of toothlessness in turtles: insight into convergent evolution of vertebrate morphology. Evolution 67, 260–273.CrossRefGoogle ScholarPubMed
Tokita, M., Kiyoshi, T., and Armstrong, K.N. (2007). Evolution of craniofacial novelty in parrots through developmental modularity and heterochrony. Evol. Dev. 9, 590–601.CrossRefGoogle ScholarPubMed
Tokita, M. and Schneider, R.A. (2009). Developmental origins of species-specific muscle pattern. Dev. Biol. 331, 311–325.CrossRefGoogle ScholarPubMed
Tokunaga, C. (1961). The differentiation of a secondary sex comb under the influence of the gene engrailed in Drosophila melanogaster. Genetics 46, 157–176.Google ScholarPubMed
Tokunaga, C. (1962). Cell lineage and differentiation on the male foreleg of Drosophila melanogaster. Dev. Biol. 4, 489–516.CrossRefGoogle ScholarPubMed
Tokunaga, C. (1978). Genetic mosaic studies of pattern formation in Drosophila melanogaster, with special reference to the prepattern hypothesis. In Genetic Mosaics and Cell Differentiation (Gehring, W.J., ed.). Results and Problems in Cell Differentiation, Vol. 9. Springer-Verlag, Berlin, pp. 157–204.CrossRefGoogle Scholar
Tomita, S. and Kikuchi, A. (2009). Abd-B suppresses lepidopteran proleg development in posterior abdomen. Dev. Biol. 328, 403–409.CrossRefGoogle ScholarPubMed
Tomoyasu, Y., Arakane, Y., Kramer, K.J., and Denell, R.E. (2009). Repeated co-options of exoskeleton formation during wing-to-elytron evolution in beetles. Curr. Biol. 19, 2057–2065.CrossRefGoogle ScholarPubMed
Tomoyasu, Y., Wheeler, S.R., and Denell, R.E. (2005). Ultrabithorax is required for membranous wing identity in the beetle Tribolium casteneum. Nature 433, 643–647.CrossRefGoogle Scholar
Tong, X., Lindemann, A., and Monteiro, A. (2012). Differential involvement of Hedgehog signaling in butterfly wing and eyespot development. PLoS ONE 7 #12, e51087.CrossRefGoogle ScholarPubMed
Toro, R. (2012). On the possible shapes of the human brain. Evol. Biol. 39, 600–612.CrossRefGoogle Scholar
Tour, E. and McGinnis, W. (2006). Gap peptides? A new way to control embryonic patterning? Cell 126, 448–449.CrossRefGoogle ScholarPubMed
Trainor, P. (2003). The bills of qucks and duails. Science 299, 523–524.CrossRefGoogle ScholarPubMed
Treisman, J.E. (2004). How to make an eye. Development 131, 3823–3827.CrossRefGoogle Scholar
True, J.R. (2003). Insect melanism: the molecules matter. Trends Ecol. Evol. 18, 640–647.CrossRefGoogle Scholar
True, J.R. and Carroll, S.B. (2002). Gene co-option in physiological and morphological evolution. Annu. Rev. Cell Dev. Biol. 18, 53–80.CrossRefGoogle ScholarPubMed
True, J.R. and Haag, E.S. (2001). Developmental system drift and flexibility in evolutionary trajectories. Evol. Dev. 3, 109–119.CrossRefGoogle ScholarPubMed
Trueb, L. (1973). Bones, frogs, and evolution. In Evolutionary Biology of the Anurans: Contemporary Research on Major Problems (Vial, J.L., ed.). University of Missouri Press, Columbia, MO, pp. 65–132.Google Scholar
Trueman, J.W.H., Pfeil, B.E., Kelchner, S.A., and Yeates, D.K. (2004). Did stick insects really regain their wings? Syst. Entomol. 29, 138–139.CrossRefGoogle Scholar
Truman, J.W. and Riddiford, L.M. (1999). The origins of insect metamorphosis. Nature 401, 447–452.CrossRefGoogle ScholarPubMed
Tschopp, P., Frandeau, N., Béna, F., and Duboule, D. (2011). Reshuffling genomic landscapes to study the regulatory evolution of Hox gene clusters. PNAS 108 #26, 10632–10637.CrossRefGoogle Scholar
Tsubota, T., Saigo, K., and Kojima, T. (2008). Hox genes regulate the same character by different strategies in each segment. Mechs. Dev. 125, 894–905.CrossRefGoogle ScholarPubMed
Tsuihiji, T., Kearney, M., and Rieppel, O. (2006). First report of a pectoral girdle muscle in snakes, with comments on the snake cervico-dorsal boundary. Copeia 2006, 206–215.CrossRefGoogle Scholar
Tsuihiji, T., Kearney, M., and Rieppel, O. (2012). Finding the neck-trunk boundary in snakes: anteroposterior dissociation of myological characteristics in snakes and its implications for their neck and trunk body regionalization. J. Morph. 273, 992–1009.CrossRefGoogle ScholarPubMed
Tsujimoto, M. and Hattori, A. (2005). The oxytocinase subfamily of M1 aminopeptidases. Biochim. Biophys. Acta 1751, 9–18.CrossRefGoogle ScholarPubMed
Tucker, A. and Sharpe, P. (2004). The cutting-edge of mammalian development; how the embryo makes teeth. Nature Rev. Genet. 5, 499–508.CrossRefGoogle ScholarPubMed
Tucker, R.P. and Erickson, C.A. (1986). The control of pigment cell pattern formation in the California newt, Taricha torosa. J. Embryol. Exp. Morph. 97, 141–168.Google ScholarPubMed
Tucker, R.P. and Erickson, C.A. (1986). Pigment cell pattern formation in Taricha torosa: the role of the extracellular matrix in controlling pigment cell migration and differentiation. Dev. Biol. 118, 268–285.CrossRefGoogle Scholar
Tummers, M. and Thesleff, I. (2003). Root or crown: a developmental choice orchestrated by the differential regulation of the epithelial stem cell niche in the tooth of two rodent species. Development 130, 1049–1057.CrossRefGoogle ScholarPubMed
Tunnicliffe, V. (1992). Hydrothermal-vent communities of the deep sea. Am. Sci. 80, 336–349.Google Scholar
Turchyn, N., Chesebro, J., Hrycaj, S., Couso, J.P., and Popadic, A. (2011). Evolution of nubbin function in hemimetabolous and holometabolous insect appendages. Dev. Biol. 357, 83–95.CrossRefGoogle ScholarPubMed
Turing, A.M. (1952). The chemical basis of morphogenesis. Phil. Trans. Roy. Soc. Lond., B 237, 37–72.CrossRefGoogle Scholar
Tuttle, R.H. (1990). Apes of the world. Am. Sci. 78, 115–125.Google Scholar
Twitty, V.C. (1944). Chromatophore migration as a response to mutual influences of the developing pigment cells. J. Exp. Zool. 95, 259–290.CrossRefGoogle Scholar
Twitty, V.C. (1945). The developmental analysis of specific pigment patterns. J. Exp. Zool. 100, 141–178.CrossRefGoogle Scholar
Twitty, V.C. and Niu, M.C. (1948). Causal analysis of chromatophore migration. J. Exp. Zool. 108, 405–437.CrossRefGoogle ScholarPubMed
Twitty, V.C. and Niu, M.C. (1954). The motivation of cell migration, studied by isolation of embryonic pigment cells singly and in small groups in vitro. J. Exp. Zool. 125, 541–573.CrossRefGoogle Scholar
Underwood, G. (1970). The eye. In Biology of the Reptilia, Vol. 2: Morphology, Part B (Gans, C. and Parsons, T.S., eds.). Academic Press, New York, NY, pp. 1–97.Google Scholar
Unwin, D.M. (1999). Pterosaurs: back to the traditional model? Trends Ecol. Evol. 14, 263–268.CrossRefGoogle ScholarPubMed
Unwin, M. (2011). Southern African Wildlife, 2nd edn. Bradt Wildlife Explorer. Bradt Travel Guides, Chalfont St. Peter.Google Scholar
Updike, J. and Gwin, P. (2007). Extreme dinosaurs. Natl. Geogr. 212 #6, 32–57.Google Scholar
Urdy, S. (2012). On the evolution of morphogenetic models: mechano-chemical interactions and an integrated view of cell differentiation, growth, pattern formation and morphogenesis. Biol. Rev. 87, 786–803.CrossRefGoogle ScholarPubMed
Vachon, G., Cohen, B., Pfeifle, C., McGuffin, M.E., Botas, J., and Cohen, S.M. (1992). Homeotic genes of the Bithorax Complex repress limb development in the abdomen of the Drosophila embryo through the target gene Distal-less. Cell 71, 437–450.CrossRefGoogle ScholarPubMed
Valentine, J.W. (2004). On the Origin of Phyla. University of Chicago Press, Chicago, IL.Google Scholar
Valkonen, J., Niskanen, M., Björklund, M., and Mappes, J. (2011). Disruption or aposematism? Significance of dorsal zigzag pattern of European vipers. Evol. Ecol. 25, 1047–1063.CrossRefGoogle Scholar
Vallin, A., Dimitrova, M., Kodandaramaiah, U., and Merilaita, S. (2011). Deflective effect and the effect of prey detectability on anti-predator function of eyespots. Behav. Ecol. Sociobiol. 65, 1629–1636.CrossRefGoogle Scholar
Vallin, A., Jakobsson, S., Lind, J., and Wiklund, C. (2006). Crypsis versus intimidation: anti-predation defence in three closely related butterflies. Behav. Ecol. Sociobiol. 59, 455–459.CrossRefGoogle Scholar
Vallin, A., Jakobsson, S., and Wiklund, C. (2007). “An eye for an eye?” On the generality of the intimidating quality of eyespots in a butterfly and a hawkmoth. Behav. Ecol. Sociobiol. 61, 1419–1424.CrossRefGoogle Scholar
van Amerongen, R. and Nusse, R. (2009). Towards an integrated view of Wnt signaling in development. Development 136, 3205–3214.CrossRefGoogle ScholarPubMed
Van de Peer, Y., Maere, S., and Meyer, A. (2009). The evolutionary significance of ancient genome duplications. Nature Rev. Genet. 10, 725–732.CrossRefGoogle ScholarPubMed
van Deusen, H.M. (1966). The seventh Archbold expedition. BioScience 16, 449–455.CrossRefGoogle Scholar
van Doorn, K. (2012). Investigations on the reptilian spectacle. Ph.D. dissertation. School of Optometry and Vision Science, University of Waterloo, Canada.
Van Essen, D.C. (1997). A tension-based theory of morphogenesis and compact wiring in the central nervous system. Nature 385, 313–318.CrossRefGoogle ScholarPubMed
van Gelder, R.G. (1959). A taxonomic revision of the spotted skunks (genus Spilogale). Bull. Am. Mus. Nat. Hist. 117 #5, 229–392.Google Scholar
Van Sittert, S.J., Skinner, J.D., and Mitchell, G. (2010). From fetus to adult: an allometric analysis of the giraffe vertebral column. J. Exp. Zool. (Mol. Dev. Evol.) 314B, 469–479.CrossRefGoogle Scholar
Van Valkenburgh, B. and Wayne, R.K. (2010). Carnivores. Curr. Biol. 20, R915–R919.CrossRefGoogle ScholarPubMed
Van Wassenbergh, S., Leysen, H., Adriaens, D., and Aerts, P. (2013). Mechanics of snout expansion in suction-feeding seahorses: musculoskeletal force transmission. J. Exp. Biol. 216, 407–417.CrossRefGoogle ScholarPubMed
Van Wassenbergh, S., Roos, G., and Ferry, L. (2011). An adaptive explanation for the horse-like shape of seahorses. Nature Commun. 2, Article 164 (5 pp.).CrossRefGoogle ScholarPubMed
Van Wassenbergh, S., Roos, G., Genbrugge, A., Leysen, H., Aerts, P., Adriaens, D., and Herrel, A. (2009). Suction is kid’s play: extremely fast suction in newborn seahorses. Biol. Lett. 5, 200–203.CrossRefGoogle ScholarPubMed
Vandenberg, L.N. and Levin, M. (2010). Far from solved: a perspective on what we know about early mechanisms of left-right asymmetry. Dev. Dynamics 239, 3131–3146.CrossRefGoogle ScholarPubMed
Vandenberg, L.N. and Levin, M. (2013). A unified model for left-right asymmetry? Comparison and synthesis of molecular models of embryonic laterality. Dev. Biol. 379, 1–15.CrossRefGoogle ScholarPubMed
Vandervorst, P. and Ghysen, A. (1980). Genetic control of sensory connections in Drosophila. Nature 286, 65–67.CrossRefGoogle ScholarPubMed
Varela-Lasheras, I., Bakker, A.J., van der Mije, S.D., Metz, J.A.J., van Alphen, J., and Galis, F. (2011). Breaking evolutionary and pleiotropic constraints in mammals: on sloths, manatees and homeotic mutations. EvoDevo 2, Article 11 (27 pp.).CrossRefGoogle ScholarPubMed
Varjosalo, M. and Taipale, J. (2008). Hedgehog: functions and mechanisms. Genes Dev. 22, 2454–2472.CrossRefGoogle ScholarPubMed
Varricchio, D.J. (2001). Gut contents from a Cretaceous Tyrannosaurid: implications for theropod dinosaur digestive tracts. J. Paleont. 75, 401–406.CrossRefGoogle Scholar
Vaughn, T.A., Ryan, J.M., and Czaplewski, N.J. (2000). Mammalogy, 4th edn. Saunders, New York, NY.Google Scholar
Vecchione, M., Young, R.E., Guerra, A., Lindsay, D.J., Clague, D.A., Bernhard, J.M., Sager, W.W., Gonzalez, A.F., Rocha, F.J., and Segonzac, M. (2001). Worldwide observations of remarkable deep-sea squids. Science 294, 2505–2506.CrossRefGoogle ScholarPubMed
Veeman, M.T., Newman-Smith, E., El-Nachef, D., and Smith, W.C. (2010). The ascidian mouth opening is derived from the anterior neuropore: reassessing the mouth/neural tube relationship in chordate evolution. Dev. Biol. 344, 138–149.CrossRefGoogle ScholarPubMed
Venditti, C. and Pagel, M. (2008). Speciation and bursts of evolution. Evo. Edu. Outreach 1, 274–280.CrossRefGoogle Scholar
Veraksa, A., Del Campo, M., and McGinnis, W. (2000). Developmental patterning genes and their conserved functions: from model organisms to humans. Mol. Genet. Metab. 69, 85–100.CrossRefGoogle Scholar
Verhulst, J. (1996). Atavisms in Homo sapiens: a Bolkian heterodoxy revisited. Acta Biotheor. 44, 59–73.CrossRefGoogle ScholarPubMed
Vickaryous, M. and Olson, W.M. (2007). Sesamoids and ossicles in the appendicular skeleton. In Fins into Limbs: Evolution, Development, and Transformation (Hall, B.K., ed.). University of Chicago Press, Chicago, IL, pp. 323–341.Google Scholar
Vickaryous, M.K. (2009). The integumentary skeleton of tetrapods: origin, evolution, and development. J. Anat. 214, 441–464.CrossRefGoogle ScholarPubMed
Vidal, N. and Hedges, S.B. (2004). Molecular evidence for a terrestrial origin of snakes. Proc. R. Soc. Lond. B (Suppl.) 271, S226–S229.CrossRefGoogle ScholarPubMed
Viguerie, N., Montastier, E., Maoret, J.-J., Roussel, B., Combes, M., Valle, C., Villa-Vialaneix, N., Iacovoni, J.S., Martinez, J.A., Holst, C., Astrup, A., Vidal, H., Clément, K., Hager, J., Saris, W.H.M., and Langin, D. (2012). Determinants of human adipose tissue gene expression: impact of diet, sex, metabolic status, and cis genetic regulation. PLoS Genet. 8 #9, e1002959.CrossRefGoogle ScholarPubMed
Villee, C.A. (1942). The phenomenon of homeosis. Am. Nat. 76, 494–506.CrossRefGoogle Scholar
Villella, A. and Hall, J.C. (2008). Neurogenetics of courtship and mating in Drosophila. Adv. Genet. 62, 67–184.Google ScholarPubMed
Villmoare, B. (2013). Morphological integration, evolutionary constraints, and extinction: a computer simulation-based study. Evol. Biol. 40, 76–83.CrossRefGoogle Scholar
Vinagre, T., Moncaut, N., Carapuço, M., Nóvoa, A., Bom, J., and Mallo, M. (2010). Evidence for a myotomal Hox/Myf cascade governing nonautonomous control of rib specification within global vertebral domains. Dev. Cell 18, 655–661.CrossRefGoogle ScholarPubMed
Vincent, J.-P. and Dubois, L. (2002). Morphogen transport along epithelia, an integrated trafficking problem. Dev. Cell 3, 615–623.CrossRefGoogle ScholarPubMed
Vitt, L.J. and Pianka, E.R. (2006). The scaly ones. Nat. Hist. 115 #6, 28–35.Google Scholar
Vogel, G. (2012). Turing pattern fingered for digit formation. Science 338, 1406.CrossRefGoogle ScholarPubMed
von Neumann, J. (1966). Theory of Self-Reproducing Automata. University of Illinois Press, Urbana, IL.Google Scholar
Vonk, F.J., Admiraal, J.F., Jackson, K., Reshef, R., de Bakker, M.A.G., Vanderschoot, K., van den Berge, I., van Atten, M., Burgerhout, E., Beck, A., Mirtschin, P.J., Kochva, E., Witte, F., Fry, B.G., Woods, A.E., and Richardson, M.K. (2008). Evolutionary origin and development of snake fangs. Nature 454, 630–633.CrossRefGoogle ScholarPubMed
Vonk, F.J., Jackson, K., Doley, R., Madaras, F., Mirtschin, P.J., and Vidal, N. (2011). Snake venom: from fieldwork to the clinic. BioEssays 33, 269–279.CrossRefGoogle ScholarPubMed
Vonk, F.J. and Richardson, M.K. (2008). Serpent clocks tick faster. Nature 454, 282–283.CrossRefGoogle ScholarPubMed
Vreede, B.M.I., Lynch, J.A., Roth, S., and Sucena, E. (2013). Co-option of a coordinate system defined by the EGFr and Dpp pathways in the evolution of a morphological novelty. EvoDevo 4, Article 7 (12 pp.).CrossRefGoogle ScholarPubMed
Waage, J.K. (1981). How the zebra got its stripes: biting flies as selective agents in the evolution of zebra colouration. J. Entomol. Soc. S. Afr. 44, 351–358.Google Scholar
Waddington, C.H. (1969). The theory of evolution today. In Beyond Reductionism: New Perspectives in the Life Sciences (Koestler, A. and Smythies, J.R., eds.). Macmillan, New York, NY, pp. 357–395.Google Scholar
Wagner, A. (2008). Gene duplications, robustness and evolutionary innovations. BioEssays 30, 367–373.CrossRefGoogle ScholarPubMed
Wagner, A. (2011). The molecular origins of evolutionary innovations. Trends Genet. 27, 397–410.CrossRefGoogle ScholarPubMed
Wagner, D.L. and Liebherr, J.K. (1992). Flightlessness in insects. Trends Ecol. Evol. 7, 216–220.CrossRefGoogle ScholarPubMed
Wagner, G.P. (1989). The biological homology concept. Annu. Rev. Ecol. Syst. 20, 51–69.CrossRefGoogle Scholar
Wagner, G.P. (1989). The origin of morphological characters and the biological basis of homology. Evolution 43, 1157–1171.CrossRefGoogle ScholarPubMed
Wagner, G.P. (1993). How can a character be developmentally constrained despite variation in developmental pathways? J. Evol. Biol. 6, 449–455.CrossRefGoogle Scholar
Wagner, G.P. (1996). Homologues, natural kinds and the evolution of modularity. Am. Zool. 36, 36–43.CrossRefGoogle Scholar
Wagner, G.P. (2007). The developmental genetics of homology. Nature Rev. Genet. 8, 473–479.CrossRefGoogle ScholarPubMed
Wagner, G.P. and Altenberg, L. (1996). Complex adaptations and the evolution of evolvability. Evolution 50, 967–976.CrossRefGoogle ScholarPubMed
Wagner, G.P., Chiu, C.-H., and Laubichler, M. (2000). Developmental evolution as a mechanistic science: the inference from developmental mechanisms to evolutionary processes. Am. Zool. 40, 819–831.Google Scholar
Wagner, G.P. and Lynch, V.J. (2008). The gene regulatory logic of transcription factor evolution. Trends Ecol. Evol. 23, 377–385.CrossRefGoogle ScholarPubMed
Wagner, G.P. and Lynch, V.J. (2010). Evolutionary novelties. Curr. Biol. 20, R48–R52.CrossRefGoogle ScholarPubMed
Wagner, G.P., Pavlicev, M., and Cheverud, J.M. (2007). The road to modularity. Nature Rev. Genet. 8, 921–931.CrossRefGoogle ScholarPubMed
Wagner, G.P. and Zhang, J. (2011). The pleiotropic structure of the genotype-phenotype map: the evolvability of complex organisms. Nature Rev. Gen. 12, 204–213.CrossRefGoogle ScholarPubMed
Wagner, R.A., Tabibiazar, R., Liao, A., and Quertermous, T. (2005). Genome-wide expression dynamics during mouse embryonic development reveal similarities to Drosophila development. Dev. Biol. 288, 595–611.CrossRefGoogle ScholarPubMed
Wake, D.B. and Larson, A. (1987). Multidimensional analysis of an evolving lineage. Science 238, 42–48.CrossRefGoogle ScholarPubMed
Wake, D.B., Wake, M.H., and Specht, C.D. (2011). Homoplasy: from detecting pattern to determining process and mechanism of evolution. Science 331, 1032–1035.CrossRefGoogle Scholar
Walker, E.P., Warnick, F., Hamlet, S.E., Lange, K.I., Davis, M.A., Uible, H.E., Wright, P.F., and Paradiso, J.L. (1975). Mammals of the World. Johns Hopkins University Press, Baltimore, MD.Google Scholar
Wallace, B. (1985). Reflections on the still-“hopeful monster”. Quart. Rev. Biol. 60, 31–42.CrossRefGoogle Scholar
Walls, G.L. (1942). The Vertebrate Eye. Cranbrook Press, Bloomfield Hills, MI.Google Scholar
Wang, L., Han, X., Mehren, J., Hiroi, M., Billeter, J.-C., Miyamoto, T., Amrein, H., Levine, J.D., and Anderson, D.J. (2011). Hierarchical chemosensory regulation of male-male social interactions in Drosophila. Nat. Neurosci. 14, 757–762.CrossRefGoogle ScholarPubMed
Wang, S. and Samakovlis, C. (2012). Grainy head and its target genes in epithelial morphogenesis and wound healing. Curr. Top. Dev. Biol. 98, 35–63.CrossRefGoogle ScholarPubMed
Wang, X. (2012). Passing the smell test. Nat. Hist. 120 #5, 22–29.Google Scholar
Wang, X. and Zhou, Z. (2004). Pterosaur embryo from the Early Cretaceous. Nature 429, 621.CrossRefGoogle ScholarPubMed
Wang, X.-P., Suomalainen, M., Felszeghy, S., Zelarayan, L.C., Alonso, M.T., Plikus, M.V., Maas, R.L., Chuong, C.-M., Schimmang, T., and Thesleff, I. (2007). An integrated gene regulatory network controls stem cell proliferation in teeth. PLoS Biol. 5 #6, e159.CrossRefGoogle ScholarPubMed
Wang, X.-P., Suomalainen, M., Jorgez, C.J., Matzuk, M.M., Werner, S., and Thesleff, I. (2004). Follistatin regulates enamel patterning in mouse incisors by asymmetrically inhibiting BMP signaling and ameloblast differentiation. Dev. Cell 7, 719–730.CrossRefGoogle ScholarPubMed
Wang, Y., Gao, Y., Imsland, F., Gu, X., Feng, C., Liu, R., Song, C., Tixier-Boichard, M., Gourichon, D., Li, Q., Chen, K., Li, H., Andersson, L., Hu, X., and Li, N. (2012). The Crest phenotype in chicken is associated with ectopic expression of Hoxc8 in cranial skin. PLoS ONE 7 #4, e34012.CrossRefGoogle ScholarPubMed
Wang, Z., Dong, D., Ru, B., Young, R.L., Han, N., Guo, T., and Zhang, S. (2010). Digital gene expression tag profiling of bat digits provides robust candidates contributing to wing formation. BMC Genomics 11, Article 619 (12 pp.).CrossRefGoogle ScholarPubMed
Ward, A.B. and Brainerd, E.L. (2007). Evolution of axial patterning in elongate fishes. Biol. J. Linnean Soc. 90, 97–116.CrossRefGoogle Scholar
Ward, C.V. (1997). Functional anatomy and phyletic implications of the hominoid trunk and hindlimb. In Function, Phylogeny, and Fossils: Miocene Hominoid Evolution and Adaptations (Begun, D.R., Ward, C.V., and Rose, M.D., eds.). Plenum Press, New York, NY, pp. 101–130.CrossRefGoogle Scholar
Ward, M. (1997). Everest 1951: the footprints attributed to the Yeti – myth and reality. Wilderness Environ. Med. 8, 29–32.CrossRefGoogle Scholar
Ward, P.D. (1991). On Methuselah’s Trail: Living Fossils and the Great Extinctions. W. H. Freeman, New York, NY.Google Scholar
Ward, P.S. (2007). Phylogeny, classification, and species-level taxonomy of ants (Hymenoptera: Formicidae). Zootaxa 1668, 549–563.Google Scholar
Ward, S.J. (1998). Numbers of teats and pre- and post-natal litter sizes in small diprotodont marsupials. J. Mammalogy 79, 999–1008.CrossRefGoogle Scholar
Warner, J.F., Lyons, D.C., and McClay, D.R. (2012). Left-right asymmetry in the sea urchin embryo: BMP and the asymmetrical origins of the adult. PLoS Biol. 10 #10, e1001404.CrossRefGoogle ScholarPubMed
Warrant, E.J. (2007). Visual ecology: hiding in the dark. Curr. Biol. 17, R209–R211.CrossRefGoogle Scholar
Warren, I. and Smith, H. (2007). Stalk-eyed flies (Diopsidae): modelling the evolution and development of an exaggerated sexual trait. BioEssays 29, 300–307.CrossRefGoogle ScholarPubMed
Warren, R. and Carroll, S. (1995). Homeotic genes and diversification of the insect body plan. Curr. Opin. Gen. Dev. 5, 459–465.CrossRefGoogle ScholarPubMed
Warren, R.W., Nagy, L., Selegue, J., Gates, J., and Carroll, S. (1994). Evolution of homeotic gene regulation and function in flies and butterflies. Nature 372, 458–461.CrossRefGoogle ScholarPubMed
Warren, W.C., Hillier, L.W., Graves, J.A.M., Birney, E., Ponting, C.P., Grützner, F., Belov, K., Miller, W., Clarke, L., Chinwalla, A.T., Yang, S.-P., Heger, A., Locke, D.P., Miethke, P., Waters, P.D., Veyrunes, F., Fulton, L., Fulton, B., Graves, T., Wallis, J., Puente, X.S., López-Otín, C., Ordóñez, G.R., Eichler, E.E., Chen, L., Cheng, Z., Deakin, J.E., Alsop, A., Thompson, K., Kirby, P., Papenfuss, A.T., Wakefield, M.J., Olender, T., Lancet, D., Huttley, G.A., Smit, A.F.A., Pask, A., Temple-Smith, P., Batzer, M.A., Walker, J.A., Konkel, M.K., Harris, R.S., Whittington, C.M., Wong, E.S.W., Gemmell, N.J., Buschiazzo, E., Jentzsch, I.M.V., Merkel, A., Schmitz, J., Zemann, A., Churakov, G., Kriegs, J.O., Brosius, J., Murchison, E.P., Sachidanandam, R., Smith, C., Hannon, G.J., Tsend-Ayush, E., McMillan, D., Attenborough, R., Rens, W., Ferguson-Smith, M., Lefèvre, C.M., Sharp, J.A., Nicholas, K.R., Ray, D.A., Kube, M., Reinhardt, R., Pringle, T.H., Taylor, J., Jones, R.C., Nixon, B., Dacheux, J.-L., Niwa, H., Sekita, Y., Huang, X., Stark, A., Kheradpour, P., Kellis, M., Flicek, P., Chen, Y., Webber, C., Hardison, R., Nelson, J., Hallsworth-Pepin, K., Delehaunty, K., Markovic, C., Minx, P., Feng, Y., Kremitzki, C., Mitreva, M., Glasscock, J., Wylie, T., Wohldmann, P., Thiru, P., Nhan, M.N., Pohl, C.S., Smith, S.M., Hou, S., Nefedov, M., de Jong, P.J., Renfree, M.B., Mardis, E.R., and Wilson, R.K. (2008). Genome analysis of the platypus reveals unique signatures of evolution. Nature 453, 175–183.CrossRefGoogle Scholar
Warrick, D., Hedrick, T., Fernández, M.J., Tobalkse, B., and Biewener, A. (2012). Hummingbird flight. Curr. Biol. 22, R472–R477.CrossRefGoogle ScholarPubMed
Washio, Y., Aritaki, M., Fujinami, Y., Shimizu, D., Yokoi, H., and Suzuki, T. (2013). Ocular-side lateralization of adult-type chromatophore precursors: development of pigment asymmetry in metamorphosing flounder larvae. J. Exp. Zool. (Mol. Dev. Evol.) 320B, 151–165.CrossRefGoogle Scholar
Wasik, B.R., Rose, D.J., and Moczek, A.P. (2010). Beetle horns are regulated by the Hox gene, Sex combs reduced, in a species- and sex-specific manner. Evol. Dev. 12, 353–362.CrossRefGoogle Scholar
Watanabe, M., Iwashita, M., Ishii, M., Kurachi, Y., Kawakami, A., Kondo, S., and Okada, N. (2006). Spot pattern of leopard Danio is caused by mutation in the zebrafish connexin41.8 gene. EMBO Reports 7, 893–897.CrossRefGoogle ScholarPubMed
Watanabe, M. and Kondo, S. (2012). Changing clothes easily: connexin41.8regulates skin pattern variation. Pigment Cell Melanoma Res. 25, 326–330.CrossRefGoogle ScholarPubMed
Watanabe, M., Watanabe, D., and Kondo, S. (2012). Polyamine sensitivity of gap junctions is required for skin pattern formation in zebrafish. Sci. Reports 2, Article 473 (5 pp.).Google ScholarPubMed
Weatherbee, S.D., Behringer, R.R., Rasweiler, J.J., IV, and Niswander, L.A. (2006). Interdigital webbing retention in bat wings illustrates genetic changes underlying amniote limb diversification. PNAS 103 #41, 15103–15107.CrossRefGoogle ScholarPubMed
Weatherbee, S.D. and Carroll, S.B. (1999). Selector genes and limb identity in arthropods and vertebrates. Cell 97, 283–286.CrossRefGoogle ScholarPubMed
Weatherbee, S.D., Halder, G., Kim, J., Hudson, A., and Carroll, S. (1998). Ultrabithorax regulates genes at several levels of the wing-patterning hierarchy to shape the development of the Drosophila haltere. Genes Dev. 12, 1474–1482.CrossRefGoogle ScholarPubMed
Weatherbee, S.D., Nijhout, H.F., Grunert, L.W., Halder, G., Galant, R., Selegue, J., and Carroll, S. (1999). Ultrabithorax function in butterfly wings and the evolution of insect wing patterns. Curr. Biol. 9, 109–115.CrossRefGoogle ScholarPubMed
Weaver, J. (2012). Striking similarities in fly and vertebrate olfactory network formation. PLoS Biol. 10 #10, e1001401.CrossRefGoogle ScholarPubMed
Weavers, H., Prieto-Sánchez, S., Grawe, F., Garcia-López, A., Artero, R., Wilsch-Braüninger, M., Ruiz-Gómez, M., Skaer, H., and Denholm, B. (2009). The insect nephrocyte is a podocyte-like cell with a filtration slit diaphragm. Nature 457, 322–326.CrossRefGoogle ScholarPubMed
Weber, B.H. and Depew, D.J. (2003). Evolution and Learning: The Baldwin Effect Reconsidered. MIT Press, Cambridge, MA.Google Scholar
Weidauer, T., Pauluis, O., and Schumacher, J. (2010). Cloud patterns and mixing properties in shallow moist Rayleigh-Bénard convection. New J. Physics 12, 105002. [See also Schaefer, V.J., and Day, J.A. (1981). A Field Guide to the Atmosphere. The Peterson Field Guide Series, Vol. 26. Houghton Mifflin, Boston, MA.]CrossRefGoogle Scholar
Weil, A. (2003). Teeth as tools. Nature 422, 128.CrossRefGoogle ScholarPubMed
Weisbecker, V. (2011). Monotreme ossification sequences and the riddle of mammalian skeletal development. Evolution 65, 1323–1335.CrossRefGoogle ScholarPubMed
Weiss, K. (2002). Good vibrations: the silent symphony of life. Evol. Anthrop. 11, 176–182.CrossRefGoogle Scholar
Weiss, K. (2004). Doin’ what comes naturally. Evol. Anthrop. 13, 47–52.CrossRefGoogle Scholar
Weiss, K. and Sholtis, S. (2003). Dinner at Baby’s: Werewolves, dinosaur jaws, hen’s teeth, and horse toes. Evol. Anthrop. 12, 247–251.CrossRefGoogle Scholar
Weiss, K.M. (2005). The phenogenetic logic of life. Nature Rev. Genet. 6, 36–45.CrossRefGoogle ScholarPubMed
Weiss, K.M. and Buchanan, A.V. (2009). The Mermaid’s Tale: Four Billion Years of Cooperation in the Making of Living Things. Harvard University Press, Cambridge, MA.Google Scholar
Weiss, K.M. and Fullerton, S.M. (2000). Phenogenetic drift and the evolution of genotype-phenotype relationships. Theor. Pop. Biol. 57, 187–195.CrossRefGoogle ScholarPubMed
Weiss, K.M., Stock, D.W., and Zhao, Z. (1998). Dynamic interactions and the evolutionary genetics of dental patterning. Crit. Rev. Oral Biol. Med. 9, 369–398.CrossRefGoogle ScholarPubMed
Weiss, P. (1969). The living system: determinism stratified. In Beyond Reductionism: New Perspectives in the Life Sciences (Koestler, A. and Smythies, J.R., eds.). Macmillan, New York, NY, pp. 3–55.Google Scholar
Wellik, D.M. and Capecchi, M.R. (2003). Hox10 and Hox11 genes are required to globally pattern the mammalian skeleton. Science 301, 363–367.CrossRefGoogle ScholarPubMed
Welsh, I.C. and O’Brien, T.P. (2009). Signaling integration in the rugae growth zone directs sequential SHH signaling center formation during the rostral outgrowth of the palate. Dev. Biol. 336, 53–67.CrossRefGoogle ScholarPubMed
Werdelin, L. and Olsson, L. (1997). How the leopard got its spots: a phylogenetic view of the evolution of felid coat patterns. Biol. J. Linnean Soc. 62, 383–400.CrossRefGoogle Scholar
Werner, T., Koshikawa, S., Williams, T.M., and Carroll, S.B. (2010). Generation of a novel wing colour pattern by the Wingless morphogen. Nature 464, 1143–1148.CrossRefGoogle ScholarPubMed
West, P.M. and Packer, C. (2002). Sexual selection, temperature, and the lion’s mane. Science 297, 1339–1343.CrossRefGoogle ScholarPubMed
West-Eberhard, M.J. (1998). Evolution in the light of developmental and cell biology, and vice versa. PNAS 95 #15, 8417–8419.CrossRefGoogle ScholarPubMed
West-Eberhard, M.J. (2003). Developmental Plasticity and Evolution. Oxford University Press, New York, NY.Google Scholar
West-Eberhard, M.J. (2005). Developmental plasticity and the origin of species differences. PNAS 102 (Suppl. 1), 6543–6549.CrossRefGoogle ScholarPubMed
Wharton, C.H. (1969). The cottonmouth moccasin on Sea Horse Key, Florida. Bull. Fla. State Mus. Biol. Sci. 14, 227–272.Google Scholar
White, M. (2012). Paradise found. Natl. Geogr. 222 #6, 70–89.Google Scholar
White, M. (2013). When push comes to shove. Natl. Geogr. 223 #4.Google Scholar
White, P.J.T., Heidemann, M., Loh, M., and Smith, J.J. (2013). Integrative cases for teaching evolution. Evo. Edu. Outreach 6, Article 17 (7 pp.).CrossRefGoogle Scholar
White, R.A.H. and Akam, M.E. (1985). Contrabithorax mutations cause inappropriate expression of Ultrabithorax products in Drosophila. Nature 318, 567–569.CrossRefGoogle Scholar
White, R.A.H. and Wilcox, M. (1985). Distribution of Ultrabithorax proteins in Drosophila. EMBO J. 4, 2035–2043.Google ScholarPubMed
White, R.A.H. and Wilcox, M. (1985). Regulation of the distribution of Ultrabithorax proteins in Drosophila. Nature 318, 563–567.CrossRefGoogle Scholar
Whiting, M.F., Bradler, S., and Maxwell, T. (2003). Loss and recovery of wings in stick insects. Nature 421, 264–267.CrossRefGoogle ScholarPubMed
Whiting, M.F. and Wheeler, W.C. (1994). Insect homeotic transformation. Nature 368, 696.CrossRefGoogle Scholar
Whitlatch, T. (2010). Animals Real and Imagined: Fantasy of What Is and What Might Be. Design Studio Press, Culver City, CA.Google Scholar
Whyte, W.A., Orlando, D.A., Hnisz, D., Abraham, B.J., Lin, C.Y., Kagey, M.H., Rahl, P.B., Lee, T.I., and Young, R.A. (2013). Master transcription factors and Mediator establish super-enhancers at key cell identity genes. Cell 153, 307–319.CrossRefGoogle ScholarPubMed
Wibowo, I., Pinto-Teixeira, F., Satou, C., Higashijima, S.-i., and López-Schier, H. (2011). Compartmentalized Notch signaling sustains epithelial mirror symmetry. Development 138, 1143–1152.CrossRefGoogle ScholarPubMed
Widelitz, R.B., Baker, R.E., Plikus, M., Lin, C.-M., Maini, P.K., Paus, R., and Chuong, C.M. (2006). Distinct mechanisms underlie pattern formation in the skin and skin appendages. Birth Defects Res. (Part C) 78, 280–291.CrossRefGoogle ScholarPubMed
Wiegmann, B.M., Yeates, D.K., Thorne, J.L., and Kishino, H. (2003). Time flies, a new molecular time-scale for Brachyceran fly evolution without a clock. Syst. Biol. 52, 745–756.CrossRefGoogle ScholarPubMed
Wiehs, D., Fish, F.E., and Nicastro, A.J. (2007). Mechanics of remora removal by dolphin spinning. Marine Mamm. Sci. 23, 707–714.CrossRefGoogle Scholar
Wiens, J.J. (2001). Shape shifters: Time after time, lizards have dropped their legs in favor of a snakelike body form. Nat. Hist. 110 #8, 70–75.Google Scholar
Wiens, J.J. (2001). Widespread loss of sexually selected traits: how the peacock lost its spots. Trends Ecol. Evol. 16, 517–523.CrossRefGoogle Scholar
Wiens, J.J. (2004). Development and evolution of body form and limb reduction in squamates: a response to Sanger and Gibson-Brown. Evolution 58, 2107–2108.CrossRefGoogle Scholar
Wiens, J.J. (2009). Estimating rates and patterns of morphological evolution from phylogenies: lessons in limb lability from Australian Lerista lizards. J. Biol. 8, e19.CrossRefGoogle ScholarPubMed
Wiens, J.J. (2011). Re-evolution of lost mandibular teeth in frogs after more than 200 million years, and re-evaluating Dollo’s Law. Evolution 65, 1283–1296.CrossRefGoogle ScholarPubMed
Wiens, J.J. and Brandley, M.C. (2009). The evolution of limblessness. In Grzimek’s Animal Life Encyclopedia (Internet edn.). Gale Cengage, Farmington Hills, MI.Google Scholar
Wiens, J.J., Brandley, M.C., and Reeder, T.W. (2006). Why does a trait evolve multiple times within a clade? Repeated evolution of snakelike body form in squamate reptiles. Evolution 60, 123–141.Google ScholarPubMed
Wiens, J.J. and Hoverman, J.T. (2008). Digit reduction, body size, and paedomorphosis in salamanders. Evol. Dev. 10, 449–463.CrossRefGoogle ScholarPubMed
Wiens, J.J., Hutter, C.R., Mulcahy, D.G., Noonan, B.P., Townsend, T.M., Sites, J.W., Jr., and Reeder, T.W. (2012). Resolving the phylogeny of lizards and snakes (Squamata) with extensive sampling of genes and species. Biol. Lett. 8, 1043–1046.CrossRefGoogle ScholarPubMed
Wiens, J.J. and Slingluff, J.L. (2001). How lizards turn into snakes: a phylogenetic analysis of body-form evolution in anguid lizards. Evolution 55, 2303–2318.CrossRefGoogle ScholarPubMed
Wigglesworth, V.B. (1940). Local and general factors in the development of “pattern” in Rhodnius prolixus (Hemiptera). J. Exp. Zool. 17, 180–200.Google Scholar
Wigglesworth, V.B. (1973). Evolution of insect wings and flight. Nature 246, 127–129.CrossRefGoogle Scholar
Wigglesworth, V.B. (1976). The evolution of insect flight. In Insect Flight. (Rainey, R.C., ed.). Symposium of the Royal Entomological Society of London, Vol. 7. Wiley, New York, NY, pp. 255–269.Google Scholar
Wikramanayake, A.H., Hong, M., Lee, P.N., Pang, K., Byrum, C.A., Bince, J.M., Xu, R., and Martindale, M.Q. (2003). An ancient role for nuclear β-catenin in the evolution of axial polarity and germ layer segregation. Nature 426, 446–450.CrossRefGoogle ScholarPubMed
Wilby, O.K. and Ede, D.A. (1975). A model generating the pattern of cartilage skeletal elements in the embryonic chick limb. J. Theor. Biol. 52, 199–217.CrossRefGoogle ScholarPubMed
Wilczynskia, B. and Furlong, E.E.M. (2010). Challenges for modeling global gene regulatory networks during development: Insights from Drosophila. Dev. Biol. 340, 161–169.CrossRefGoogle Scholar
Wilder, E.L. and Perrimon, N. (1995). Dual functions of wingless in the Drosophila leg imaginal disc. Development 121, 477–488.Google ScholarPubMed
Wilkie, A.L., Jordan, S.A., and Jackson, I.J. (2002). Neural crest progenitors of the melanocyte lineage: coat colour patterns revisited. Development 129, 3349–3357.Google ScholarPubMed
Wilkins, A.S. (1989). Organizing the Drosophila posterior pattern: why has the fruit fly made life so complicated for itself?BioEssays 11, 67–69.CrossRefGoogle Scholar
Wilkins, A.S. (1993). Genetic Analysis of Animal Development, 2nd edn. Wiley-Liss, New York, NY.Google Scholar
Wilkins, A.S. (1997). Canalization: a molecular genetic perspective. BioEssays 19, 257–262.CrossRefGoogle ScholarPubMed
Wilkins, A.S. (2007). Between “design” and “bricolage”: Genetic networks, levels of selection, and adaptive evolution. PNAS 104 (Suppl. 1), 8590–8596.CrossRefGoogle ScholarPubMed
Wilkinson, D.M. and Ruxton, G.D. (2012). Understanding selection for long necks in different taxa. Biol. Rev. 87, 616–630.CrossRefGoogle ScholarPubMed
Wilkinson, G.S. (1994). Female choice response to artificial selection on an exaggerated male trait in a stalk-eyed fly. Proc. R. Soc. Lond. B 255, 1–6.CrossRefGoogle Scholar
Wilkinson, G.S. and Johns, P.M. (2005). Sexual selection and the evolution of mating systems in flies. In The Evolutionary Biology of Flies (Yeates, D.K. and Wiegmann, B.M., eds.). Columbia University Press, New York, NY, pp. 312–339.Google Scholar
Wilkinson, G.S., Johns, P.M., Metheny, J.D., and Baker, R.H. (2013). Sex-biased gene expression during head development in a sexually dimorphic stalk-eyed fly. PLoS ONE 8 #3, e59826.CrossRefGoogle Scholar
Wilkinson, M.T. (2007). Sailing the skies: the improbable aeronautical success of the pterosaurs. J. Exp. Biol. 210, 1663–1671.CrossRefGoogle ScholarPubMed
Willey, A. (1911). Convergence in Evolution. John Murray, London.CrossRefGoogle Scholar
Williams, G.C. (1992). Natural Selection: Domains, Levels, and Challenges. Oxford Series in Ecology and Evolution, Vol. 4. Oxford University Press, New York, NY.Google Scholar
Williston, S.W. (1914). Water Reptiles of the Past and Present. University of Chicago Press, Chicago, IL.CrossRefGoogle Scholar
Willmore, K.E. (2010). Development influences evolution. Am. Sci. 98, 220–227.CrossRefGoogle Scholar
Willmore, K.E. (2012). The body plan concept and its centrality in evo-devo. Evo. Edu. Outreach 5, 219–230.CrossRefGoogle Scholar
Willnow, T.E., Christ, A., and Hammes, A. (2012). Endocytic receptor-mediated control of morphogen signaling. Development 139, 4311–4319.CrossRefGoogle ScholarPubMed
Wilson, A.B. and Orr, J.W. (2011). The evolutionary origins of Syngnathidae: pipefishes and seahorses. J. Fish Biol. 78, 1603–1623.CrossRefGoogle ScholarPubMed
Wilson, D.E. and Mittermeier, R.A., eds. Handbook of the Mammals of the World. Vol. 1 (Carnivores). Lynx Edicions, Barcelona.
Wilson, D.E., Mittermeier, R.A., Ruff, S., and Martinez-Vilalta, A., eds. Handbook of the Mammals of the World. Vol. 2 (Hoofed Mammals). Lynx Edicions, Barcelona.
Wilting, A., Buckley-Beason, V.A., Feldhaar, H., Gadau, J., O’Brien, S.J., and Linsenmair, K.E. (2007). Clouded leopard phylogeny revisited: support for species recognition and population division between Borneo and Sumatra. Front. Zool. 4, Article 15 (10 pp.).CrossRefGoogle ScholarPubMed
Winchell, C.J. and Jacobs, D.K. (2013). Expression of the Lhx genes apterous and lim1 in an errant polychaete: implications for bilaterian appendage evolution, neural development, and muscle diversification. EvoDevo 4, Article 4 (19 pp.).CrossRefGoogle Scholar
Winchell, C.J., Valencia, J.E., and Jacobs, D.K. (2010). Expression of Distal-less, dachshund, and optomotor blind in Neanthes arenaceodentata (Annelida, Nereididae) does not support homology of appendage-forming mechanisms across the Bilateria. Dev. Genes Evol. 220, 275–295.CrossRefGoogle Scholar
Winfree, A.T. (1980). The Geometry of Biological Time. Springer-Verlag, Berlin.CrossRefGoogle Scholar
Winfree, A.T. (1984). The prehistory of the Belousov-Zhabotinsky oscillator. J. Chem. Ed. 61, 661–663.CrossRefGoogle Scholar
Winfree, A.T., Winfree, E.M., and Seifert, H. (1985). Organizing centers in a cellular excitable medium. Physica 17D, 109–115.Google Scholar
Wings, O. and Sander, P.M. (2007). No gastric mill in sauropod dinosaurs: new evidence from analysis of gastrolith mass and function in ostriches. Proc. R. Soc. Lond. B 274, 635–640.CrossRefGoogle ScholarPubMed
Winter, S. (1998). The elusive quetzal. Natl. Geogr. 193 #6, 34–45.Google Scholar
Witkop, C.J., Jr., Quevedo, W.C., Jr., Fitzpatrick, T.B., and King, R.A. (1989). Albinism. In The Metabolic Basis of Inherited Disease, Vol. 2, 6th edn. (Scriver, C.R., Beaudet, A.L., Sly, W.S., and Valle, D., eds.). McGraw-Hill, New York, NY, pp. 2905–2947.Google Scholar
Wittkopp, P.J. and Beldade, P. (2009). Development and evolution of insect pigmentation: genetic mechanisms and the potential consequences of pleiotropy. Semin. Cell Dev. Biol. 20, 65–71.CrossRefGoogle ScholarPubMed
Wittkopp, P.J. and Kalay, G. (2012). Cis-regulatory elements: molecular mechanisms and evolutionary processes underlying divergence. Nature Rev. Genet. 13, 59–69.CrossRefGoogle Scholar
Wolfe, J.M., Oliver, J.C., and Monteiro, A. (2010). Evolutionary reduction of the first thoracic limb in butterflies. J. Insect Sci. 11, Article 66 (9 pp.).Google Scholar
Wolfram, S. (1984). Cellular automata as models of complexity. Nature 311, 419–424.CrossRefGoogle Scholar
Wolpert, L. (1968). The French Flag problem: a contribution to the discussion on pattern development and regulation. In Towards a Theoretical Biology. I. Prolegomena (Waddington, C.H., ed.). Aldine, Chicago, IL, pp. 125–133.Google Scholar
Wolpert, L. (1969). Positional information and the spatial pattern of cellular differentiation. J. Theor. Biol. 25, 1–47.CrossRefGoogle ScholarPubMed
Wolpert, L. (1989). Positional information revisited. Development 1989 Suppl., 3–12.Google Scholar
Wolpert, L. (1990). Signals in limb development: STOP, GO, STAY and POSITION. J. Cell Sci. Suppl. 13, 199–208.CrossRefGoogle Scholar
Wolpert, L. (1996). One hundred years of positional information. Trends Genet. 12, 359–364.CrossRefGoogle ScholarPubMed
Wolpert, L., Tickle, C., Jessell, T., Lawrence, P., Meyerowitz, E., Robertson, E., and Smith, J. (2011). Principles of Development, 4th edn. Oxford University Press, New York, NY.Google Scholar
Woltering, J.M. (2012). From lizard to snake: behind the evolution of an extreme body plan. Curr. Genomics 13, 289–299.CrossRefGoogle ScholarPubMed
Woltering, J.M., Vonk, F.J., Müller, H., Bardine, N., Tuduce, I.L., de Bakker, M.A.G., Knöchel, W., Sirbu, I.O., Durston, A.J., and Richardson, M.K. (2009). Axial patterning in snakes and caecilians: Evidence for an alternative interpretation of the Hox code. Dev. Biol. 332, 82–89.CrossRefGoogle ScholarPubMed
Wong, B.B.M. and Rosenthal, G.G. (2006). Female disdain for swords in the swordtail fish. Am. Nat. 167, 136–140.CrossRefGoogle ScholarPubMed
Wong, K. (2002). The mammals that conquered the seas. Sci. Am. 286 #5, 70–79.CrossRefGoogle ScholarPubMed
Wong, K. (2004). Becoming behemoth. Sci. Am. 290 #2, 23–24.CrossRefGoogle ScholarPubMed
Wootton, R.J. (1976). The fossil record and insect flight. In Insect Flight. (Rainey, R.C., ed.). Symposium of the Royal Entomological Society of London, Vol. 7. Wiley, New York, NY, pp. 235–254.Google Scholar
Wootton, R.J. (1992). Functional morphology of insect wings. Annu. Rev. Entomol. 37, 113–140.CrossRefGoogle Scholar
Wootton, R.J. (1999). Invertebrate paraxial locomotory appendages: design, deformation and control. J. Exp. Biol. 202, 3333–3345.Google ScholarPubMed
Wootton, R.J. and Kukalová-Peck, J. (2000). Flight adaptations in Palaeozoic Palaeoptera (Insecta). Biol. Rev. 75, 129–167.CrossRefGoogle Scholar
Wootton, R.J., Kukalová-Peck, J., Newman, D.J.S., and Muzon, J. (1998). Smart engineering in the mid-Carboniferous: how well could Palaeozoic dragonflies fly? Science 282, 749–751.CrossRefGoogle ScholarPubMed
Wourms, M.K. and Wasserman, F.E. (1985). Butterfly wing markings are more advantageous during handling than during the initial strike of an avian predator. Evolution 39, 845–851.CrossRefGoogle ScholarPubMed
Wray, G.A. and Abouheif, E. (1998). When is homology not homology? Curr. Opin. Gen. Dev. 8, 675–680.CrossRefGoogle Scholar
Wu, C.-I. (1996). Now blows the east wind. Nature 380, 105–107.CrossRefGoogle ScholarPubMed
Wu, D., Freund, J.B., Fraser, S.E., and Vermot, J. (2011). Mechanical basis of otolith formation during teleost inner ear development. Dev. Cell 20, 271–278.CrossRefGoogle Scholar
Wu, M.Y. and Hill, C.S. (2009). TGF-β superfamily signaling in embryonic development and homeostasis. Dev. Cell 16, 329–343.CrossRefGoogle ScholarPubMed
Wu, P., Hou, L., Plikus, M., Hughes, M., Scehnet, J., Suksaweang, S., Widelitz, R.B., Jiang, T.-X., and Chuong, C.-M. (2004). Evo-Devo of amniote integuments and appendages. Int. J. Dev. Biol. 48, 249–270.CrossRefGoogle ScholarPubMed
Wu, P., Jiang, T.-X., Shen, J.-Y., Widelitz, R.B., and Chuong, C.-M. (2006). Morphoregulation of avian beaks: comparative mapping of growth zone activities and morphological evolution. Dev. Dynamics 235, 1400–1412.CrossRefGoogle ScholarPubMed
Wu, P., Jiang, T.-X., Suksaweang, S., Widelitz, R.B., and Chuong, C.-M. (2004). Molecular shaping of the beak. Science 305, 1465–1466.CrossRefGoogle ScholarPubMed
Wu, X., Jung, G., and Hammer, J.A., III (2000). Functions of unconvential myosins. Curr. Opin. Cell Biol. 12, 42–51.CrossRefGoogle Scholar
Wueringer, B.E., Squire, L., Jr., Kajiura, S.M., Hart, N.S., and Collin, S.P. (2012). The function of the sawfish’s saw. Curr. Biol. 22, R150–R151.CrossRefGoogle ScholarPubMed
Wund, M.A. (2012). Assessing the impacts of phenotypic plasticity on evolution. Integr. Comp. Biol. 52, 5–15.CrossRefGoogle ScholarPubMed
Wyss, A.R. (1988). Evidence from flipper structure for a single origin of pinnipeds. Nature 334, 427–428.CrossRefGoogle Scholar
Xenia, M., Ilagan, G., and Kopan, R. (2007). Snapshot: Notch signaling pathway. Cell 128, 1246.Google Scholar
Xiang, H., Li, M.W., Guo, J.H., Jiang, J.H., and Huang, Y.P. (2011). Influence of RNAi knockdown for E-complex genes on the silkworm proleg development. Arch. Insect Biochem. Physiol. 76, 1–11.CrossRefGoogle ScholarPubMed
Xiong, F., Tentner, A.R., Huang, P., Gelas, A., Mosaliganti, K.R., Souhait, L., Rannou, N., Swinburne, I.A., Obholzer, N.D., Cowgill, P.D., Schier, A.F., and Megason, S.G. (2013). Specified neural progenitors sort to form sharp domains after noisy Shh signaling. Cell 153, 550–561.CrossRefGoogle ScholarPubMed
Xu, X. (2013). Modular genetic control of innate behaviors. BioEssays 35, 421–424.CrossRefGoogle ScholarPubMed
Xu, X., Dong, G.-X., Hu, X.-S., Miao, L., Zhang, X.-L., Zhang, D.-L., Yang, H.-D., Zhang, T.-Y., Zou, Z.-T., Zhang, T.-T., Zhuang, Y., Bhak, J., Cho, Y.S., Dai, W.-T., Jiang, T.-J., Xie, C., Li, R., and Luo, S.-J. (2013). The genetic basis of white tigers. Curr. Biol. 23, 1031–1035.CrossRefGoogle ScholarPubMed
Xu, X., Sullivan, C., Pittman, M., Choiniere, J.N., Hone, D., Upchurch, P., Tan, Q., Xiao, D., Tan, L., and Han, F. (2011). A monodactyl nonavian dinosaur and the complex evolution of the alvarezsauroid hand. PNAS 108 #6, 2338–2342.CrossRefGoogle ScholarPubMed
Yamada, G., Suzuki, K., Haraguchi, R., Miyagawa, S., Satoh, Y., Kamimura, M., Nakagata, N., Kataoka, H., Kuroiwa, A., and Chen, Y. (2006). Molecular genetic cascades for external genitalia formation: an emerging organogenesis program. Dev. Dynamics 235, 1738–1752.CrossRefGoogle ScholarPubMed
Yamaguchi, M., Yoshimoto, E., and Kondo, S. (2007). Pattern regulation in the stripe of zebrafish suggests an underlying dynamic and autonomous mechanism. PNAS 104 #12, 4790–4793.CrossRefGoogle ScholarPubMed
Yamamoto, S., Charng, W.-L., Rana, N.A., Kakuda, S., Jaiswal, M., Bayat, V., Xiong, B., Zhang, K., Sandoval, H., David, G., Wang, H., Haltiwanger, R.S., and Bellen, H.J. (2012). A mutation in EGF repeat-8 of Notch discriminates between Serrate/Jagged and Delta family ligands. Science 338, 1229–1232.CrossRefGoogle ScholarPubMed
Yamanoue, Y., Setiamarga, D.H.E., and Matsuura, K. (2010). Pelvic fins in teleosts: structure, function and evolution. J. Fish Biol. 77, 1173–1208.CrossRefGoogle ScholarPubMed
Yáñez-Cuna, J.O., Kvon, E.Z., and Stark, A. (2013). Deciphering the transcriptional cis-regulatory code. Trends Genet. 29, 11–22.CrossRefGoogle ScholarPubMed
Yang, A.S. (2001). Modularity, evolvability, and adaptive radiations: a comparison of the hemi- and holometabolous insects. Evol. Dev. 3, 59–72.CrossRefGoogle ScholarPubMed
Yang, Z., Bertolucci, F., Wolf, R., and Heisenberg, M. (2013). Flies cope with uncontrollable stress by learned helplessness. Curr. Biol. 23, 799–803.CrossRefGoogle ScholarPubMed
Yano, T. and Tamura, K. (2013). The making of differences between fins and limbs. J. Anat. 222, 100–113.CrossRefGoogle ScholarPubMed
Yarmolinsky, D.A., Zuker, C.S., and Ryba, N.J.P. (2009). Common sense about taste: from mammals to insects. Cell 139, 234–244.CrossRefGoogle ScholarPubMed
Yau, K.-W. and Hardie, R.C. (2009). Phototransduction motifs and variations. Cell 139, 246–264.CrossRefGoogle ScholarPubMed
Yekta, S., Tabin, C.J., and Bartel, D.P. (2008). MicroRNAs in the Hox network: an apparent link to posterior prevalence. Nature Rev. Genet. 9, 789–796.CrossRefGoogle ScholarPubMed
Yoshida, A. and Aoki, K. (1989). Scale arrangement pattern in a lepidopteran wing. 1. Periodic cellular pattern in the pupal wing of Pieris rapae. Dev. Growth Differ. 31, 601–609.CrossRefGoogle Scholar
Young, B.A. and Kardong, K.V. (2010). The functional morphology of hooding in cobras. J. Exp. Biol. 213, 1521–1528.CrossRefGoogle ScholarPubMed
Young, K.V., Brodie, E.D., Jr., and Brodie, E.D., III (2004). How the horned lizard got its horns. Science 304, 65.CrossRefGoogle ScholarPubMed
Young, N.M. and Hallgrímsson, B. (2005). Serial homology and the evolution of mammalian limb covariation structure. Evolution 59, 2691–2704.CrossRefGoogle ScholarPubMed
Young, R.L. and Wagner, G.P. (2011). Why ontogenetic homology criteria can be misleading: lessons from digit identity transformations. J. Exp. Zool. (Mol. Dev. Evol.) 316B, 165–170. [See also Xu, X., and Makem, S. (2013). Tracing the evolution of avian wing digits. Curr. Biol. 23, R538–R544. Another recent paper that is relevant here is by de Bakker, M.A.G., et al. (2013). Digit loss in archosaur evolution and the interplay between selection and constraints. Nature 500, 445–448.]CrossRefGoogle ScholarPubMed
Young, T., Rowland, J.E., van de Ven, C., Bialecka, M., Novoa, A., Carapuco, M., van Nes, J., de Graaff, W., Duluc, I., Freund, J.-N., Beck, F., Mallo, M., and Deschamps, J. (2009). Cdx and Hox genes differentially regulate posterior axial growth in mammalian embryos. Dev. Cell 17, 516–526.CrossRefGoogle ScholarPubMed
Yu, C.Q., Schwab, I.R., and Dubielzig, R.R. (2009). Feeding the vertebrate retina from the Cambrian to the Tertiary. J. Zool. 278, 259–269.CrossRefGoogle Scholar
Yu, L., Jin, W., Zhang, X., Wang, D., Zheng, J.-s., Yang, G., Xu, S.-x., Cho, S., and Zhang, Y.-p. (2011). Evidence for positive selection on the leptin gene in cetacea and pinnipedia. PLoS ONE 6 #10, e26579.CrossRefGoogle ScholarPubMed
Yue, Z., Jiang, T.-X., Widelitz, R.B., and Chuong, C.-M. (2006). Wnt3a gradient converts radial to bilateral feather symmetry via topological arrangement of epithelia. PNAS 103 #4, 951–955.CrossRefGoogle ScholarPubMed
Zaher, H. and Rieppel, O. (1999). The phylogenetic relationships of Pachyrhachis problematicus, and the evolution of limblessness in snakes (Lepidosauria, Squamata). C. R. Acad. Sci. Paris, Sciences de la terre et des planètes 329, 831–837.Google Scholar
Zahradnicek, O., Horacek, I., and Tucker, A.S. (2008). Viperous fangs: Development and evolution of the venom canal. Mechs. Dev. 125, 786–796.CrossRefGoogle ScholarPubMed
Zakany, J. and Duboule, D. (2007). The role of Hox genes during vertebrate limb development. Curr. Opin. Gen. Dev. 17, 359–366.CrossRefGoogle ScholarPubMed
Zakin, L. and De Robertis, E.M. (2010). Extracellular regulation of BMP signaling. Curr. Biol. 20, R89–R92.CrossRefGoogle ScholarPubMed
Zamora, S., Rahman, I.A., and Smith, A.B. (2012). Plated Cambrian bilaterians reveal the earliest stages of echinoderm evolution. PLoS ONE 7 #6, e38296. [See also Smith, A.B. and Zamora, S. (2013). Cambrian spiral-plated echinoderms from Gondwana reveal the earliest pentaradial body plan. Proc. R. Soc. Lond. B 280. . N.B.: Believe it or not, some sea stars have 50 arms or more. See Janosik, A.M., et al. (2008) Life history of the Antarctic sea star Labidiaster annulatus (Asteroidea: Labidiasteridae) revealed by DNA barcoding. Antarctic Sci. 20, 563–564.]CrossRefGoogle ScholarPubMed
Zbikowski, R. (2002). Red admiral agility. Nature 420, 615–618.CrossRefGoogle ScholarPubMed
Zecca, M., Basler, K., and Struhl, G. (1996). Direct and long-range action of a Wingless morphogen gradient. Cell 87, 833–844.CrossRefGoogle ScholarPubMed
Zeeman, E.C. (1974). Primary and secondary waves in developmental biology. In Lectures on Mathematics in the Life Sciences, Vol. 7. American Mathematical Society, Providence, RI, pp. 69–161.Google Scholar
Zeil, J., Nalbach, G., and Nalbach, H.-O. (1986). Eyes, eye stalks and the visual world of semi-terrestrial crabs. J. Comp. Physiol. A 159, 801–811.CrossRefGoogle Scholar
Zelditch, M.L. (2003). Space, time, and repatterning. In Keywords and Concepts in Evolutionary Developmental Biology (Hall, B.K. and Olson, W.M., eds.). Harvard University Press, Cambridge, MA, pp. 341–348.Google Scholar
Zelditch, M.L. and Fink, W.L. (1996). Heterochrony and heterotopy: innovation and stability in the evolution of form. Paleobiology 22, 241–254.CrossRefGoogle Scholar
Zelenitsky, D.K., Therrien, F., Erickson, G.M., DeBuhr, C.L., Kobayashi, Y., Eberth, D.A., and Hadfield, F. (2012). Feathered non-avian dinosaurs from North America provide insight into wing origins. Science 338, 510–514.CrossRefGoogle ScholarPubMed
Zhai, Z., Han, N., Papagiannouli, F., Hamacher-Brady, A., Brady, N., Sorge, S., Bezdan, D., and Lohmann, I. (2012). Antagonistic regulation of apoptosis and differentiation by the Cut transcription factor represents a tumor-suppressing mechanism in Drosophila. PLoS Genet. 8 #3, e1002582.CrossRefGoogle ScholarPubMed
Zhang, J., Tian, Y., Wang, L., and He, C. (2010). Functional evolutionary developmental biology (evo-devo) of morphological novelties in plants. J. Syst. Evol. 48, 94–101.CrossRefGoogle Scholar
Zhang, X.-g. and Pratt, B.R. (2012). The first stalk-eyed phosphatocopine crustacean from the Lower Cambrian of China. Curr. Biol. 22, 2149–2154.CrossRefGoogle ScholarPubMed
Zheng, X., Zhou, Z., Wang, X., Zhang, F., Zhang, X., Wang, Y., Wei, G., Wang, S., and Xu, X. (2013). Hind wings in basal birds and the evolution of leg feathers. Science 339, 1309–1312.CrossRefGoogle ScholarPubMed
Zheng, Z., Khoo, A., Fambrough, D., Jr., Garza, L., and Booker, R. (1999). Homeotic gene expression in the wild-type and a homeotic mutant of the moth Manduca sexta. Dev. Genes Evol. 209, 460–472.CrossRefGoogle Scholar
Zhong, Y.-f. and Holland, P.W.H. (2011). The dynamics of vertebrate homeobox gene evolution: gain and loss of genes in mouse and human lineages. BMC Evol. Biol. 11, Article 169 (13 pp.).Google ScholarPubMed
Zhou, Q., Yu, L., Shen, X., Li, Y., Xu, W., Yi, Y., and Zhang, Z. (2009). Homology of dipteran bristles and lepidopteran scales: requirement for the Bombyx mori achaete-scute homologue ASH2. Genetics 183, 619–627.CrossRefGoogle ScholarPubMed
Zhou, S., Lo, W.-C., Suhalim, J.L., Digman, M.A., Gratton, E., Nie, Q., and Lander, A.D. (2012). Free extracellular diffusion creates the Dpp morphogen gradient of the Drosophila wing disc. Curr. Biol. 22, 668–675.CrossRefGoogle ScholarPubMed
Zhou, Z., Clark, J., Zhang, F., and Wings, O. (2004). Gastroliths in Yanornis: an indication of the earliest radical diet-switching and gizzard plasticity in the lineage leading to living birds? Naturwissenschaften 91, 571–574.CrossRefGoogle ScholarPubMed
Zhou, Z. and Zhang, F. (2006). A beaked basal ornithurine bird (Aves, Ornithurae) from the Lower Cretaceous of China. Zool. Scripta 35, 363–373.CrossRefGoogle Scholar
Zhu, A.J. and Scott, M.P. (2004). Incredible journey: how do developmental signals travel through tissue? Genes Dev. 18, 2985–2997.CrossRefGoogle ScholarPubMed
Zhu, B., Pennack, J.A., McQuilton, P., Forero, M.G., Mizuguchi, K., Sutcliffe, B., Gu, C.-J., Fenton, J.C., and Hidalgo, A. (2008). Drosophila neurotrophins reveal a common mechanism for nervous system formation. PLoS Biol. 6 #11, 2476–2495 (e284).CrossRefGoogle ScholarPubMed
Zhu, L., Wilken, J., Phillips, N.B., Narendra, U., Chan, G., Stratton, S.M., Kent, S.B., and Weiss, M.A. (2000). Sexual dimorphism in diverse metazoans is regulated by a novel class of intertwined zinc fingers. Genes Dev. 14, 1750–1764.Google ScholarPubMed
Zill, S.N. and Seyfarth, E.-A. (1996). Exoskeletal sensors for walking. Sci. Am. 275 #1, 86–90.CrossRefGoogle Scholar
Zimmer, C. (2002). The rise and fall of the nasal empire. Nat. Hist. 111 #5, 32–35.Google Scholar
Zimmer, C. (2005). Dinosaurs: Why do we have so many questions about the most successful animals that ever lived? Discover 26 #4, 32–39.Google Scholar
Zimmer, C. (2008). The evolution of extraordinary eyes: the cases of flatfishes and stalk-eyed flies. Evo. Edu. Outreach 1, 487–492.CrossRefGoogle Scholar
Zimmer, C. (2011). The long curious extravagant evolution of feathers. Natl. Geogr. 219 #2, 32–57.Google Scholar
Zimmer, C. (2012). The common hand. Natl. Geogr. 221 #5, 98–105.Google Scholar
Zuber, M.E., Gestri, G., Viczian, A.S., Barsacchi, G., and Harris, W.A. (2003). Specification of the vertebrate eye by a network of eye field transcription factors. Development 130, 5155–5167.CrossRefGoogle ScholarPubMed
Zuzarte-Luis, V. and Hurle, J.M. (2005). Programmed cell death in the embryonic vertebrate limb. Semin. Cell Dev. Biol. 16, 261–269.CrossRefGoogle ScholarPubMed
Zwarts, L., Magwire, M.M., Carbone, M.A., Versteven, M., Herteleer, L., Anholt, R.R.H., Callaerts, P., and Mackay, T.F.C. (2011). Complex genetic architecture of Drosophila aggressive behavior. PNAS 108 #41, 17070–17075.CrossRefGoogle ScholarPubMed
Zylinski, S. and Johnsen, S. (2011). Mesopelagic cephalopods switch between transparency and pigmentation to optimize camouflage in the deep. Curr. Biol. 21, 1937–1941.CrossRefGoogle ScholarPubMed

Accessibility standard: Unknown

Accessibility compliance for the PDF of this book is currently unknown and may be updated in the future.

Save book to Kindle

To save this book to your Kindle, first ensure no-reply@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
  • Lewis I. Held, Jr, Texas Tech University
  • Book: How the Snake Lost its Legs
  • Online publication: 05 June 2014
  • Chapter DOI: https://doi.org/10.1017/CBO9781139343497.010
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
  • Lewis I. Held, Jr, Texas Tech University
  • Book: How the Snake Lost its Legs
  • Online publication: 05 June 2014
  • Chapter DOI: https://doi.org/10.1017/CBO9781139343497.010
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
  • Lewis I. Held, Jr, Texas Tech University
  • Book: How the Snake Lost its Legs
  • Online publication: 05 June 2014
  • Chapter DOI: https://doi.org/10.1017/CBO9781139343497.010
Available formats
×