Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-25T15:53:03.889Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  26 February 2021

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

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Animal Anomalies
What Abnormal Anatomies Reveal about Normal Development
, pp. 187 - 260
Publisher: Cambridge University Press
Print publication year: 2021

Access options

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

References

Abdelilah, S., Solnica-Krezel, L., Stainier, D.Y.R., and Driever, W. (1994). Implications for dorsoventral axis determination from the zebrafish mutation janus. Nature 370, 468471.CrossRefGoogle ScholarPubMed
Abe, M. and Kuroda, R. (2019). The development of CRISPR for a mollusc establishes the formin Lsdia1 as the long-sought gene for snail dextral/sinistral coiling. Development 146, dev175976.Google Scholar
Aboitiz, F. and Montiel, J.F. (2019). Morphological evolution of the vertebrate forebrain: from mechanical to cellular processes. Evol. Dev. 21, 330341.Google Scholar
Abu-Shaar, M. and Mann, R.S. (1998). Generation of multiple antagonistic domains along the proximodistal axis during Drosophila leg development. Development 125, 38213830.Google Scholar
Abzhanov, A. (2017). The old and new faces of morphology: the legacy of D’Arcy Thompson’s “theory of transformations” and “laws of growth”. Development 144, 42844297.CrossRefGoogle ScholarPubMed
Acurio, A.E., Rhebergen, F.T., Paulus, S., Courtier-Orgogozo, V., and Lang, M. (2019). Repeated evolution of asymmetric genitalia and right-sided mating behavior in the Drosophila nannoptera species group. BMC Evol. Biol. 19, 109.Google Scholar
Adhikari, K., Fontanil, T., Cal, S., Mendoza-Revilla, J., Fuentes-Guajardo, M., Chacón-Duque, J.C., Al-Saadi, F., Johansson, J.A., Quinto-Sánchez, M., Acuña-Alonzo, V., Jaramillo, C., Arias, W., Lozano, R.B., Pérez, G.M., Gómez-Valdés, J., Villamil-Ramírez, H., Hunemeier, T., Ramallo, V., Silva de Cerqueira, C.C., Hurtado, M., Villegas, V., Granja, V., Gallo, C., Poletti, G., Schuler-Faccini, L., Salzano, F.M., Cátira Bortolinia, M., Canizales-Quinteros, S., Rothhammer, F., Bedoya, G., González-José, R., Headon, D., López-Otín, C., Tobin, D.J., Balding, D., and Ruiz-Linares, A. (2016). A genome-wide association scan in admixed Latin Americans identifies loci influencing facial and scalp hair features. Nat. Commun. 7, 10815.Google Scholar
Agi, E., Langen, M., Altschuler, S.J., Wu, L.F., Zimmermann, T., and Hiesinger, P.R. (2014). The evolution and development of neural superposition. J. Neurogenet. 28, 216232.Google Scholar
Aiello, D. and Lasagna, E. (2018). The myostatin gene: an overview of mechanisms of action and its relevance to livestock animals. Anim. Genet. 49, 505519.CrossRefGoogle ScholarPubMed
Akam, M. (1998). Hox genes: from master genes to micromanagers. Curr. Biol. 8, R676R678.Google Scholar
Akey, J.M., Ruhe, A.L., Akey, D.T., Wong, A.K., Connelly, C.F., Madeoy, J., Nicholas, T.J., and Neff, M.W. (2010). Tracking footprints of artificial selection in the dog genome. PNAS 107, 11601165.CrossRefGoogle ScholarPubMed
Alberch, P. (1982). Developmental constraints in evolutionary processes. In Evolution and Development, Bonner, J.T., editor. Springer-Verlag, Berlin, pp. 313332.CrossRefGoogle Scholar
Alberch, P. (1985). Developmental constraints: why St. Bernards often have an extra digit and poodles never do. Am. Nat. 126, 430433.CrossRefGoogle Scholar
Alberch, P. (1986). Possible dogs. Nat. Hist. 95(12), 48.Google Scholar
Alberch, P. (1989). The logic of monsters: evidence for internal constraint in development and evolution. Geobios 22(Suppl. 2), 2157.Google Scholar
Alberch, P., Gould, S.J., Oster, G.F., and Wake, D.B. (1979). Size and shape in ontogeny and phylogeny. Paleobiology 5, 296317.Google Scholar
Alberts, B., Johnson, A., Lewis, J., Raff, M., Roberts, K., and Walter, P. (2002). Molecular Biology of the Cell, 4th ed. Garland, New York.Google Scholar
Alexander, R.M. (1995). Big flies have bigger cells. Nature 375, 20.Google Scholar
Alibardi, L. (2018). Appendage regeneration in amphibians and some reptiles derived from specific evolutionary histories. J. Exp. Zool. B Mol. Dev. Evol. 330, 396405.Google Scholar
Alibardi, L. (2018). Limb regeneration in humans: dream or reality? Ann. Anat. 217, 16.Google Scholar
Allchin, D. (2019). How the tiger changed its stripes. Am. Biol. Teacher 81, 599604.CrossRefGoogle Scholar
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. B 278, 13731380.CrossRefGoogle ScholarPubMed
Alpert, B.O. (2013). The meaning of the dots on the horses of Pech Merle. Arts 2, 476490.CrossRefGoogle Scholar
Alsina, B. and Whitfield, T.T. (2017). Sculpting the labyrinth: morphogenesis of the developing inner ear. Semin. Cell Dev. Biol. 65, 4759.Google Scholar
Ambegaonkar, A.A. and Irvine, K.D. (2015). Coordination of planar cell polarity pathways through Spiny legs. eLife 4, e09946.Google Scholar
Ambrosi, D., Ben Amar, M., Cyron, C.J., DeSimone, A., Goriely, A., Humphrey, J.D., and Kuhl, E. (2019). Growth and remodelling of living tissues: perspectives, challenges and opportunities. J. R. Soc. Interface 16, 20190233.CrossRefGoogle ScholarPubMed
Anderson, D. and Brenner, S. (2008). Seymour Benzer (1921–2007): restless spirit, and pioneer in molecular genetics. Nature 451, 139.Google Scholar
Andl, T., Reddy, S.T., Gaddapara, T., and Millar, S.E. (2002). WNT signals are required for the initiation of hair follicle development. Dev. Cell 2, 643653.Google Scholar
Andrew, D.J. and Ewald, A.J. (2010). Morphogenesis of epithelial tubes: insights into tube formation, elongation, and elaboration. Dev. Biol. 341, 3455.Google Scholar
Angelini, D.R. and Kaufman, T.C. (2005). Insect appendages and comparative ontogenetics. Dev. Biol. 286, 5777.CrossRefGoogle ScholarPubMed
Anon, . (1957). The Science of Fingerprints. Federal Bureau of Investigation, Washington, DC.Google Scholar
Arendt, J. (2007). Ecological correlates of body size in relation to cell size and cell number: patterns in flies, fish, fruits and foliage. Biol. Rev. 82, 241256.CrossRefGoogle ScholarPubMed
Argyriou, T., Clauss, M., Maxwell, E.E., Furrer, H., and Sánchez-Villagra, M.R. (2016). Exceptional preservation reveals gastrointestinal anatomy and evolution in early actinopterygian fishes. Sci. Rep. 6, 18758.Google Scholar
Arnone, M.I. and Davidson, E.H. (1997). The hardwiring of development: organization and function of genomic regulatory systems. Development 124, 18511864.CrossRefGoogle ScholarPubMed
Arnosti, D.N. and Kulkarni, M.M. (2005). Transcriptional enhancers: intelligent enhanceosomes or flexible billboards? J. Cell. Biochem. 94, 890898.Google Scholar
Artavanis-Tsakonas, S. and Muskavitch, M.A.T. (2010). Notch: the past, the present, and the future. Curr. Top. Dev. Biol. 92, 129.Google Scholar
Artavanis-Tsakonas, S., Rand, M.D., and Lake, R.J. (1999). Notch signaling: cell fate control and signal integration in development. Science 284, 770776.Google Scholar
Arthur, W. (2006). D’Arcy Thompson and the theory of transformations. Nat. Rev. Genet. 7, 401406.Google Scholar
Atallah, J. and Larsen, E. (2009). Genotype–phenotype mapping: developmental biology confronts the toolkit paradox. Int. Rev. Cell Mol. Biol. 278, 119148.Google Scholar
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, 191204.Google Scholar
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, 476483.Google Scholar
Athanasiadis, A.P., Tzannatos, C., Mikos, T., Zafrakas, M., and Bontis, J.N. (2005). A unique case of conjoined triplets. Am. J. Obstet. Gynecol. 192, 20842087.CrossRefGoogle ScholarPubMed
Ather, S., Proudlock, F.A., Welton, T., Morgan, P.S., Sheth, V., Gottlob, I., and Dineen, R.A. (2018). Aberrant visual pathway development in albinism: from retina to cortex. Hum. Brain Mapp. 40, 777788.CrossRefGoogle ScholarPubMed
Audette, D.S., Scheiblin, D.A., and Duncan, M.K. (2017). The molecular mechanisms underlying lens fiber elongation. Exp. Eye Res. 156, 4149.Google Scholar
Auerbach, C. (1936). The development of the legs, wings, and halteres in wild type and some mutant strains of Drosophila melanogaster. Trans. R. Soc. Edinb. 58, 787815.Google Scholar
Averbach, B. and Chein, O. (1999). Problem Solving Through Recreational Mathematics. Dover, New York.Google Scholar
Aw, S. and Levin, M. (2008). What’s left in asymmetry? Dev. Dynamics 237, 34533463.CrossRefGoogle ScholarPubMed
Aw, W.Y. and Devenport, D. (2016). Planar cell polarity: global inputs establishing cellular asymmetry. Curr. Opin. Cell Biol. 44, 110116.CrossRefGoogle ScholarPubMed
Axelrod, J. (2008). Bad hair days for mouse PCP mutants. Nat. Cell Biol. 10, 12511252.Google Scholar
Ayala-Carmago, A., Ekas, L.A., Flaherty, M.S., Baeg, G.-H., and Bach, E.A. (2007). The JAK/STAT pathway regulates proximo-distal patterning in Drosophila. Dev. Dynamics 236, 27212730.Google Scholar
Ayukawa, T., Akiyama, M., Mummery-Widmer, J.L., Stoeger, T., Sasaki, J., Knoblich, J.A., Senoo, H., Sasaki, T., and Yamazaki, M. (2014). Dachsous-dependent asymmetric localization of Spiny-legs determines planar cell polarity orientation in Drosophila. Cell Rep. 8, 610621.CrossRefGoogle ScholarPubMed
Babler, W.J. (1991). Embryologic development of epidermal ridges and their configurations. Birth Defects Orig. Artic. Ser. 27, 95112.Google Scholar
Baer, M.M., Chanut-Delalande, H., and Affolter, M. (2009). Cellular and molecular mechanisms underlying the formation of biological tubes. Curr. Top. Dev. Biol. 89, 137162.Google Scholar
Baker, N.E. (2011). Proximodistal patterning in the Drosophila leg: models and mutations. Genetics 187, 10031010.Google Scholar
Baker, N.E. and Brown, N.L. (2018). All in the family: proneural bHLH genes and neuronal diversity. Development 145, 159426.Google Scholar
Baker, R.E., Schnell, S., and Maini, P.K. (2006). A clock and wavefront mechanism for somite formation. Dev. Biol. 293, 116126.Google Scholar
Ball, P. (1999). The Self-Made Tapestry: Pattern Formation in Nature. Oxford University Press, New York.Google Scholar
Ball, P. (2015). Forging patterns and making waves from biology to geology: a commentary on Turing (1952) “The chemical basis of morphogenesis”. Philos. Trans. R. Soc. Lond. B 370, 20140218.Google Scholar
Bando, T., Mito, T., Hamada, Y., Ishimaru, Y., Noji, S., and Ohuchi, H. (2018). Molecular mechanisms of limb regeneration: insights from regenerating legs of the cricket Gryllus bimaculatus. Int. J. Dev. Biol. 62, 559569.Google Scholar
Bando, T., Mito, T., Maeda, Y., Nakamura, T., Ito, F., Watanabe, T., Ohuchi, H., and Noji, S. (2009). Regulation of leg size and shape by the Dachsous/Fat signalling pathway during regeneration. Development 136, 22352245.Google Scholar
Bangru, S. and Kalsotra, A. (2020). Cellular and molecular basis of liver regeneration. Semin. Cell Dev. Biol. 100, 7487.Google Scholar
Bao, R., Dia, S.E., Issa, H.A., Alhusein, D., and Friedrich, M. (2018). Comparative evidence of an exceptional impact of gene duplication on the developmental evolution of Drosophila and the higher Diptera. Front. Ecol. Evol. 6, 63.Google Scholar
Barad, O., Hornstein, E., and Barkai, N. (2011). Robust selection of sensory organ precursors by the Notch–Delta pathway. Curr. Opin. Cell Biol. 23, 663667.Google Scholar
Bard, J. (2011). A systems biology representation of developmental anatomy. J. Anat. 218, 591599.Google Scholar
Bard, J.B.L. (1977). A unity underlying the different zebra striping patterns. J. Zool. Lond. 183, 527539.Google Scholar
Bard, J.B.L. (1981). A model for generating aspects of zebra and other mammalian coat patterns. J. Theor. Biol. 93, 363385.Google Scholar
Bard, J.B.L. (2008). Waddington’s legacy to developmental and theoretical biology. Biol. Theory 3, 188197.Google Scholar
Bard, J.B.L. (2018). Tinkering and the origins of heritable anatomical variation in vertebrates. Biology 7, 7010020.Google Scholar
Barham, G. and Clarke, N.M.P. (2008). Genetic regulation of embryological limb development with relation to congenital limb deformity in humans. J. Child. Orthop. 2, 19.Google Scholar
Barmina, O. and Kopp, A. (2007). Sex-specific expression of a HOX gene associated with rapid morphological evolution. Dev. Biol. 311, 277286.Google Scholar
Barnes, J., ed. (1984). The Complete Works of Aristotle: The Revised Oxford Translation, Vol. 1. Princeton University Press, Princeton, NJ.Google Scholar
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, 11671181.CrossRefGoogle ScholarPubMed
Barriga, E.H. and Mayor, R. (2015). Embryonic cell–cell adhesion: a key player in collective neural crest migration. Curr. Top. Dev. Biol. 112, 301323.CrossRefGoogle ScholarPubMed
Bartos, L., Bubenik, G.A., and Kuzmova, E. (2012). Endocrine relationships between rank-related behavior and antler growth in deer. Front. Biosci. E4, 11111126.CrossRefGoogle Scholar
Bassnett, S. and Costello, M.J. (2017). The cause and consequence of fiber cell compaction in the vertebrate lens. Exp. Eye Res. 156, 5057.Google Scholar
Bassnett, S., Shi, Y., and Vrensen, G.F.J.M. (2011). Biological glass: structural determinants of eye lens transparency. Philos. Trans. R. Soc. Lond. B 366, 12501264.Google Scholar
Bastida, M.F., Pérez-Gómez, R., Trofka, A., Zhu, J., Rada-Iglesias, A., Sheth, R., Stadler, H.S., Mackem, S., and Ros, M.A. (2020). The formation of the thumb requires direct modulation of Gli3 transcription by Hoxa13. PNAS 117, 10901096.Google Scholar
Bate, M. and Martinez Arias, A. (1991). The embryonic origin of imaginal discs in Drosophila. Development 112, 755761.Google Scholar
Bateson, G. (1971). A re-examination of “Bateson’s rule”. J. Genet. 60, 230240.Google Scholar
Bateson, W. (1894). Materials for the Study of Variation Treated with Especial Regard to Discontinuity in the Origin of Species. Macmillan, London.Google Scholar
Bateson, W. (1909). Mendel’s Principles of Heredity. Cambridge University Press, Cambridge.Google Scholar
Baudouin-Gonzalez, L., Santos, M.A., Tempesta, C., Sucena, E., Roch, F., and Tanaka, K. (2017). Diverse cis-regulatory mechanisms contribute to expression evolution of tandem gene duplicates. Mol. Biol. Evol. 34, 31323147.Google Scholar
Baumeister, F.A.M., Egger, J., Schildhauer, M.T., and Stengel-Rutkowski, S. (1993). Ambras syndrome: delineation of a unique hypertrichosis universalis congenita and association with a balanced pericentric inversion (8) (p11.2; q22). Clin. Genet. 44, 121128.Google Scholar
Bazin-Lopez, N., Valdivia, L.E., Wilson, S.W., and Gestri, G. (2015). Watching eyes take shape. Curr. Opin. Genet. Dev. 32, 7379.Google Scholar
Bazopoulou-Kyrkanidou, E. (2001). Chimeric creatures in Greek mythology and reflections in science. Am. J. Med. Genet. 100, 6680.Google Scholar
Beadle, G.W. (1970). Alfred Henry Sturtevant (1891–1970). Am. Philos. Soc. Yearbook 1970, 166171.Google Scholar
Bear, J.E. and Haugh, J.M. (2014). Directed migration of mesenchymal cells: where signaling and the cytoskeleton meet. Curr. Opin. Cell Biol. 30, 7482.Google Scholar
Bechtel, H.B. (1995). Reptile and Amphibian Variants: Colors, Patterns, and Scales. Krieger, Malabar, FL.Google Scholar
Beebe, D.C., Vasiliev, O., Guo, J., Shui, Y.-B., and Bassnett, S. (2001). Changes in adhesion complexes define stages in the differentiation of lens fiber cells. Invest. Ophthalmol. Vis. Sci. 42, 727734.Google ScholarPubMed
Beermann, F., Orlow, S.J., and Lamoreux, M.L. (2004). The Tyr (albino) locus of the laboratory mouse. Mamm. Genome 15, 749758.Google Scholar
Begley, S. (1982). How life begins. Newsweek (Jan. 11, 1982), 38–43.Google Scholar
Beira, J.V. and Paro, R. (2016). The legacy of Drosophila imaginal discs. Chromosoma 125, 573592.Google Scholar
Bell, M.L., Earl, J.B., and Britt, S.G. (2007). Two types of Drosophila R7 photoreceptor cells are arranged randomly: a model for stochastic cell-fate determination. J. Comp. Neurol. 502, 7585.Google Scholar
Bellaïche, Y., Gho, M., Kaltschmidt, J., Brand, A., and Schweisguth, F. (2001). Frizzled regulates localization of cell-fate determinants and mitotic spindle rotation during asymmetric cell division. Nat. Cell Biol. 3, 5057.Google Scholar
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. Brief. Funct. Genomics 9, 193207.Google Scholar
Beloussov, L.V., Opitz, J.M., and Gilbert, S.F. (1997). Life of Alexander G. Gurwitsch and his relevant contribution to the theory of morphogenetic fields. Int. J. Dev. Biol. 41, 771779.Google Scholar
Bénazéraf, B. and Pourquie, O. (2013). Formation and segmentation of the vertebrate body axis. Annu. Rev. Cell Dev. Biol. 29, 126.Google Scholar
Benkel, B.F., Rouvinen-Watt, K., Farid, H., and Anistoroaei, R. (2009). Molecular characterization of the Himalayan mink. Mamm. Genome 20, 256259.Google Scholar
Benzer, S. (1971). From the gene to behavior. JAMA 218, 10151022.Google Scholar
Benzer, S. (1973). Genetic dissection of behavior. Sci. Am. 229(6), 2437.Google Scholar
Bercovitch, F.B. (2019). Giraffe taxonomy, geographic distribution and conservation. Afr. J. Ecol. 58, 150158.Google Scholar
Bergman, J. (2002). Darwin’s ape-men and the exploitation of deformed humans. Technical Journal 16, 116122.Google Scholar
Bergmann, P., Richter, S., Glöckner, N., and Betz, O. (2018). Morphology of hindwing veins in the shield bug Graphosoma italicum (Heteroptera: Pentatomidae). Arthropod Struct. Dev. 47, 375390.Google Scholar
Bernard, B.A. (2017). The hair follicle enigma. Exp. Dermatol. 26, 472477.Google Scholar
Bernays, M.E. and Smith, R. (1999). Convergent strabismus in a white Bengal tiger. Aust. Vet. J. 77, 152155.Google Scholar
Berton, P. (1977). The Dionne Years: A Thirties Melodrama. W. W. Norton, New York.Google 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, 221227.Google Scholar
Bhalla, U.S. and Iyengar, R. (1999). Emergent properties of networks of biological signaling pathways. Science 283, 381387.Google Scholar
Biesecker, L.G. (2011). Polydactyly: how many disorders and how many genes? 2010 update. Dev. Dynamics 240, 931942.Google Scholar
Biesecker, L.G. and Spinner, N.B. (2013). A genomic view of mosaicism and human disease. Nat. Rev. Genet. 14, 307320.CrossRefGoogle ScholarPubMed
Bigas, A. and Espinosa, L. (2016). Notch signaling in cell-cell communication pathways. Curr. Stem Cell Rep. 2, 349355.CrossRefGoogle Scholar
Binns, W., James, L.F., Shupe, J.L., and Everett, G. (1963). A congenital cyclopian-type malformation in lambs induced by maternal ingestion of a range plant, Veratrum californicum. Am. J. Vet. Res. 24, 11641175.Google Scholar
Biosa, G., Bastianoni, S., and Rustici, M. (2006). Chemical waves. Chem. Eur. J. 12, 34303437.Google Scholar
Bishop, S.A., Klein, T., Martinez Arias, A., and Couso, J.P. (1999). Composite signalling from Serrate and Delta establishes leg segments in Drosophila through Notch. Development 126, 29933003.Google Scholar
Bizzarri, M., Giuliani, A., Minini, M., Monti, N., and Cucina, A. (2020). Constraints shape cell function and morphology by canalizing the developmental path along the Waddington’s landscape. BioEssays 42, 1900108.Google Scholar
Black, S.D. and Gerhart, J.C. (1986). High frequency twinning in Xenopus eggs centrifuged before first cleavage. Dev. Biol. 116, 228240.Google Scholar
Blair, S.S. (2004). Developmental biology: Notching the hindbrain. Curr. Biol. 14, R570R572.Google Scholar
Blair, S.S., Brower, D.L., Thomas, J.B., and Zavortink, M. (1994). The role of apterous in the control of dorsoventral compartmentalization and PS integrin gene expression in the developing wing of Drosophila. Development 120, 18051815.Google Scholar
Blanco, J., Girard, F., Kamachi, Y., Kondoh, H., and Gehring, W. (2005). Functional analysis of the chicken d1-crystallin enhancer activity in Drosophila reveals remarkable evolutionary conservation between chicken and fly. Development 132, 18951905.Google 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, 175187.Google Scholar
Blaustein, A.R. and Johnson, P.T.J. (2003). Explaining frog deformities. Sci. Am. 288(2), 6065.Google Scholar
Blaustein, A.R. and Johnson, P.T.J. (2003). The complexity of deformed amphibians. Front. Ecol. Environ. 1, 8794.Google Scholar
Blum, M. and Ott, T. (2018). Animal left–right asymmetry. Curr. Biol. 28, R301R304.Google Scholar
Blum, M. and Ott, T. (2019). Mechanical strain, novel genes and evolutionary insights: news from the frog left–right organizer. Curr. Biol. 56, 814.Google Scholar
Blum, M., Feistel, K., Thumberger, T., and Schweickert, A. (2014). The evolution and conservation of left–right patterning mechanisms. Development 141, 16031613.Google Scholar
Blum, M., Schweickert, A., Vick, P., Wright, C.V.E., and Danilchik, M.V. (2014). Symmetry breakage in the vertebrate embryo: when does it happen and how does it work? Dev. Biol. 393, 109123.Google Scholar
Blumberg, M.S. (2009). Freaks of Nature: What Anomalies Tell Us about Development. Oxford University Press, New York.Google Scholar
Boareto, M. (2020). Patterning via local cell–cell interactions in developing systems. Dev. Biol. 460, 7785.Google Scholar
Boer, L.L., Schepens-Franke, A.N., and Oostra, R.J. (2019). Two is a crowd: on the enigmatic etiopathogenesis of conjoined twinning. Clin. Anat. 32, 722741.CrossRefGoogle ScholarPubMed
Bohn, H. (1965). Analyse der Regenerationfähigkeit der Insektenextremität durch Amputations- und Transplantationsversuche an Larven der Afrikanischen Schabe (Leucophaea maderae Fabr.). II. Achsendetermination. Roux Arch. Entw.-Mech. 156, 449503.Google Scholar
Böhni, R., Riesgo-Escovar, J., Oldham, S., Brogiolo, W., Stocker, H., Andruss, B.F., Beckingham, K., and Hafen, E. (1999). Autonomous control of cell and organ size by CHICO, a Drosophila homolog of vertebrate IRS1–4. Cell 97, 865875.Google Scholar
Bohring, A., Stamm, T., Spaich, C., Haase, C., Spree, K., Hehr, U., Hoffmann, M., Ledig, S., Sel, S., Wieacker, P., and Röpke, A. (2009). WNT10A mutations are a frequent cause of a broad spectrum of ectodermal dysplasias with sex-biased manifestation pattern in heterozygotes. Am. J. Hum. Genet. 85, 97105.CrossRefGoogle ScholarPubMed
Bökel, C. and Brand, M. (2014). Endocytosis and signaling during development. Cold Spring Harb. Perspect. Biol. 6, a017020.Google Scholar
Boklage, C.E. (2006). Embryogenesis of chimeras, twins and anterior midline asymmetries. Hum. Reprod. 21, 579591.Google Scholar
Bolk, L. (1926). Das Problem der Menschwerdung. Gustav Fischer, Jena.Google Scholar
Borok, M.J., Tran, D.A., Ho, M.C.W., and Drewell, R.A. (2010). Dissecting the regulatory switches of development: lessons from enhancer evolution in Drosophila. Development 137, 513.Google Scholar
Bosch, M., Bishop, S.-A., Baguña, J., and Couso, J.-P. (2010). Leg regeneration in Drosophila abridges the normal developmental program. Int. J. Dev. Biol. 54, 12411250.Google Scholar
Botelho, J.F., Smith-Paredes, D., Soto-Acuña, S., Núñez-León, D., Palma, V., and Vargas, A.O. (2016). Greater growth of proximal metatarsals in bird embryos and the evolution of hallux position in the grasping foot. J. Exp. Zool. B Mol. Dev. Evol. 328, 106118.Google Scholar
Botstein, D. and Maurer, R. (1982). Genetic approaches to the analysis of microbial development. Annu. Rev. Genet. 16, 6183.Google Scholar
Bower, M.A., McGivney, B.A., Campana, M.G., Gu, J., Andersson, L.S., Barrett, E., Davis, C.R., Mikko, S., Stock, F., Voronkova, V., Bradley, D.G., Fahey, A.G., Lindgren, G., MacHugh, D.E., Sulimova, G., and Hill, E.W. (2012). The genetic origin and history of speed in the Thoroughbred racehorse. Nat. Commun. 3, 643.Google Scholar
Bownes, M. and Seiler, M. (1977). Developmental effects of exposing Drosophila embryos to ether vapour. J. Exp. Zool. 199, 923.Google Scholar
Boyce, D. (1992). A sight to behold: Deidre finds a toad with eyes in its mouth. The Hamilton Spectator (Sept. 3, 1992).Google Scholar
Bozorgmehr, J.E.H. (2014). The role of self-organization in developmental evolution. Theory Biosci. 133, 145163.CrossRefGoogle ScholarPubMed
Brakefield, P.M. (1999). Butterfly wings: the evolution of development of colour patterns. BioEssays 21, 391401.Google Scholar
Brakefield, P.M., French, V., and Zwaan, B.J. (2003). Development and the genetics of evolutionary change within insect species. Annu. Rev. Ecol. Evol. Syst. 34, 633660.Google Scholar
Bray, S. (1998). Notch signalling in Drosophila: three ways to use a pathway. Semin. Cell Dev. Biol. 9, 591597.Google Scholar
Bredov, D. and Volodyaev, I. (2018). Increasing complexity: mechanical guidance and feedback loops as a basis for self-organization in morphogenesis. BioSystems 173, 133156.Google Scholar
Brewer, A.A. (2009). Visual maps: to merge or not to merge. Curr. Biol. 19, R945R947.Google Scholar
Bridges, C.B. and Brehme, K.S. (1944). The Mutants of Drosophila melanogaster. Carnegie Institution of Washington, Washington, DC.Google Scholar
Brigham, P.A., Cappas, A., and Uno, H. (1988). The stumptailed macaque as a model for androgenetic alopecia: effects of topical minoxidil analyzed by use of the folliculogram. Clin. Dermatol. 6(4), 177187.Google Scholar
Briscoe, J. and Kicheva, A. (2017). The physics of development 100 years after D’Arcy Thompson’s “On Growth and Form”. Mech. Dev. 145, 2631.Google Scholar
Brison, N., Debeer, P., and Tylzanowski, P. (2013). Joining the fingers: a HOXD13 story. Dev. Dynamics 243, 3748.Google Scholar
Britton, N.F. (1986). Reaction–Diffusion Equations and Their Applications to Biology. Academic Press, New York.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, 213221.CrossRefGoogle ScholarPubMed
Brommage, R., Powell, D.R., and Vogel, P. (2019). Predicting human disease mutations and identifying drug targets from mouse gene knockout phenotyping campaigns. Dis. Model. Mech. 12, dmm038224.Google Scholar
Bronner, M.E. and LeDouarin, N.M. (2012). Development and evolution of the neural crest: an overview. Dev. Biol. 366, 29.Google Scholar
Browd, S.R., Goodrich, J.T., and Walker, M.L. (2008). Craniopagus twins. J. Neurosurg. Pediatr. 1, 120.Google Scholar
Brower, J.S., Wootton-Gorges, S.L., Costouros, J.G., Boakes, J., and Greenspan, A. (2003). Congenital diplopodia. Pediatr. Radiol. 33, 797799.Google Scholar
Brown, D.M., Brenneman, R.A., Koepfli, K.-P., Pollinger, J.P., Milá, B., Georgiadis, N.J., Louis, E.E. Jr., Grether, G.F., Jacobs, D.K., and Wayne, R.K. (2007). Extensive population genetic structure in the giraffe. BMC Biol. 5, 57.CrossRefGoogle ScholarPubMed
Brückner, K., Perez, L., Clausen, H., and Cohen, S. (2000). Glycosyltransferase activity of Fringe modulates Notch–Delta interactions. Nature 406, 411415.Google Scholar
Brunet, T., Larson, B.T., Linden, T.A., Vermeij, M.J.A., McDonald, K., and King, N. (2019). Light-regulated collective contractility in a multicellular choanoflagellate. Science 366, 326334.Google Scholar
Bryant, P.J. (1971). Regeneration and duplication following operations in situ on the imaginal discs of Drosophila melanogaster. Dev. Biol. 26, 637651.Google Scholar
Bryant, S.V., French, V., and Bryant, P.J. (1981). Distal regeneration and symmetry. Science 212, 9931002.Google Scholar
Budday, S., Nay, R., de Rooij, R., Steinmann, P., Wyrobek, T., Ovaert, T.C., and Kuhl, E. Mechanical properties of gray and white matter brain tissue by indentation. J. Mech. Behav. Biomed. Mater. 46, 318–330.Google Scholar
Buffry, A.D., Mendes, C.C., and McGregor, A.P. (2016). The functionality and evolution of eukaryotic transcriptional enhancers. Adv. Genet. 96, 143206.Google Scholar
Bulger, M. and Groudine, M. (2010). Enhancers: the abundance and function of regulatory sequences beyond promoters. Dev. Biol. 339, 250257.Google Scholar
Bullough, W.S. (1962). The control of mitotic activity in adult mammalian tissues. Biol. Rev. 37, 307342.Google Scholar
Burger, B., Fuchs, D., Sprecher, E., and Itin, P. (2011). The immigration delay disease: adermatoglyphia – inherited absence of epidermal ridges. J. Am. Acad. Dermatol. 64, 974980.Google Scholar
Butler, M.T. and Wallingford, J.B. (2017). Planar cell polarity in development and disease. Nat. Rev. Mol. Cell Biol. 18, 375388.Google Scholar
Cadieu, E., Neff, M.W., Quignon, P., Walsh, K., Chase, K., Parker, H.G., VonHoldt, B.M., Rhue, A., Boyko, A., Byers, A., Wong, A., Mosher, D.S., Elkahloun, A.G., Spady, T.C., André, C., Lark, K.G., Cargill, M., Bustamante, C.D., Wayne, R.K., and Ostrander, E.A. (2009). Coat variation in the domestic dog is governed by variants in three genes. Science 326, 150153.CrossRefGoogle ScholarPubMed
Cai, J., Townsend, J.P., Dodson, T.C., Heiney, P.A., and Sweeney, A.M. (2017). Eye patches: protein assembly of index-gradient squid lenses. Science 357, 564569.Google Scholar
Campbell, G. (2002). Distalization of the Drosophila leg by graded EGF-receptor activity. Nature 418, 781785.Google Scholar
Campbell, G. and Tomlinson, A. (1995). Initiation of the proximodistal axis in insect legs. Development 121, 619628.Google Scholar
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, 44834493.Google Scholar
Campbell, G., Weaver, T., and Tomlinson, A. (1993). Axis specification in the developing Drosophila appendage: the role of wingless, decapentaplegic, and the homeobox gene aristaless. Cell 74, 11131123.Google Scholar
Campos-Ortega, J.A. (1998). The genetics of the Drosophila achaete-scute gene complex: a historical appraisal. Int. J. Dev. Biol. 42, 291297.Google Scholar
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, 8394.Google Scholar
Capdevila, M.P. and García-Bellido, A. (1978). Phenocopies of bithorax mutants. W. Roux Arch. Dev. Biol. 185, 105126.CrossRefGoogle Scholar
Capek, D. and Müller, P. (2019). Positional information and tissue scaling during development and regeneration. Development 146, dev177709.CrossRefGoogle ScholarPubMed
Capilla, A., Johnson, R., Daniels, M., Benavente, M., Bray, S.J., and Galindo, M.I. (2012). Planar cell polarity controls directional Notch signaling in the Drosophila leg. Development 139, 25842593.Google Scholar
Caro, T. (2009). Contrasting coloration in terrestrial mammals. Philos. Trans. R. Soc. Lond. B 364, 537548.Google Scholar
Caro, T. and Mallarino, R. (2020). Coloration in mammals. Trends Ecol. Evol. 35, 357366.CrossRefGoogle ScholarPubMed
Carpenter, A.C., Smith, A.N., Wagner, H., Cohen-Tayar, Y., Rao, S., Wallace, V., Ashery-Padan, R., and Lang, R.A. (2015). Wnt ligands from the embryonic surface ectoderm regulate “bimetallic strip” optic cup morphogenesis in mouse. Development 142, 972982.Google Scholar
Carroll, L. and Gardner, M. (1960). The Annotated Alice: Alice’s Adventures in Wonderland & Through the Looking Glass. Meridian, New York.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., editor. University of Chicago Press, Chicago, IL, pp. 185224.Google Scholar
Carroll, S.B. (2000). Endless forms: the evolution of gene regulation and morphological diversity. Cell 101, 577580.Google Scholar
Carroll, S.B. (2005). Endless Forms Most Beautiful: The New Science of Evo Devo and the Making of the Animal Kingdom. Norton, New York.Google Scholar
Carroll, S.B., Grenier, J.K., and Weatherbee, S.D. (2005). From DNA to Diversity: Molecular Genetics and the Evolution of Animal Design, 2nd ed. Blackwell, Malden, MA.Google Scholar
Carson, H.L. and Kaneshiro, K.Y. (1976). Drosophila of Hawaii: systematics and ecological genetics. Annu. Rev. Ecol. Syst. 7, 311345.Google Scholar
Casares, F. and Mann, R.S. (2001). The ground state of the ventral appendage in Drosophila. Science 293, 14771480.Google Scholar
Casas, E. and Kehrli, M.E. Jr. (2016). A review of selected genes with known effects on performance and health of cattle. Front. Vet. Sci. 3, 113.Google Scholar
Cassina, M., Cagnoli, G.A., Zuccarello, D., Di Gianantonio, E., and Clementi, M. (2017). Human teratogens and genetic phenocopies. Understanding pathogenesis through human genes mutation. Eur. J. Med. Genet. 60, 2231.Google Scholar
Castelli-Gair, J. (1998). Implications of the spatial and temporal regulation of Hox genes on development and evolution. Int. J. Dev. Biol. 42, 437444.Google ScholarPubMed
Castelli-Gair Hombría, J. and Lovegrove, B. (2003). Beyond homeosis: HOX function in morphogenesis and organogenesis. Differentiation 71, 461476.Google Scholar
Cavodeassi, F. and Houart, C. (2011). Brain regionalization: of signaling centers and boundaries. Dev. Neurobiol. 72, 218233.Google Scholar
Cavodeassi, F., del Corral, R.D., Campuzano, S., and Domínguez, M. (1999). Compartments and organising boundaries in the Drosophila eye: the role of the homeodomain Iroquois proteins. Development 126, 49334942.Google Scholar
Chan, C.J., Heisenberg, C.-P., and Hiiragi, T. (2017). Coordination of morphogenesis and cell-fate specification in development. Curr. Biol. 27, R1024R1035.Google Scholar
Chang, H.Y. (2009). Anatomic demarcation of cells: genes to patterns. Science 326, 12061207.Google Scholar
Chang, S. (2017). How squid build their graded-index spherical lenses. Physics Today 70, 2628.Google Scholar
Chang, X., Li, D., Tian, L., Liu, Y., March, M., Wang, T., Hou, C., Pellegrino, R., Levy, R., Jen, M., Soccio, R., Sleiman, P., Hakonarson, H., and Castelo-Soccio, L. (2018). Heterozygous deletion impacting SMARCAD1 in the original kindred with absent dermatoglyphs and associated features (Baird, 1964). J. Pediatr. 194, 248252.Google Scholar
Charlton-Perkins, M., Brown, N.L., and Cook, T.A. (2011). The lens in focus: a comparison of lens development in Drosophila and vertebrates. Mol. Genet. Genomics 286, 189213.CrossRefGoogle ScholarPubMed
Chauhan, B., Plageman, T., Lou, M., and Lang, R. (2015). Epithelial morphogenesis: the mouse eye as a model system. Curr. Top. Dev. Biol. 111, 375399.Google Scholar
Chauhan, B.K., Lou, M., Zheng, Y., and Lang, R.A. (2011). Balanced Rac1 and RhoA activities regulate cell shape and drive invagination morphogenesis in epithelia. PNAS 108, 1828918294.Google Scholar
Chen, J. and Chuong, C.-M. (2011). Patterning skin by planar cell polarity: the multi-talented hair designer. Exp. Dermatol. 21, 8185.Google Scholar
Chen, J., Jacox, L.A., Saldanha, F., and Sive, H. (2017). Mouth development. Wiley Interdiscip. Rev. Dev. Biol. 6, e275.Google Scholar
Chew, K.Y., Yu, H., Pask, A.J., Shaw, G., and Renfree, M.B. (2012). HOXA13 and HOXD13 expression during development of the syndactylous digits in the marsupial Macropus eugenii. BMC Dev. Biol. 12, 2.Google Scholar
Chiang, C., Litingtung, Y., Lee, E., Young, K.E., Corden, J.L., Westphal, H., and Beachy, P.A. (1996). Cyclopia and defective axial patterning in mice lacking Sonic hedgehog gene function. Nature 383, 407413.Google Scholar
Choe, C.P. and Crump, J.G. (2015). Dynamic epithelia of the developing vertebrate face. Curr. Opin. Genet. Dev. 32, 6672.Google Scholar
Chouard, T. (2010). Revenge of the hopeful monster. Nature 463, 864867.Google Scholar
Ciechanska, E., Dansereau, D.A., Svendsen, P.C., Heslip, T.R., and Brook, W.J. (2007). dAP–2 and defective proventriculus regulate Serrate and Delta expression in the tarsus of Drosophila melanogaster. Genome 50, 693705.Google Scholar
Cieslak, J., Borowska, A., Wodas, L., and Mackowski, M. (2018). Interbreed distribution of the myostatin (MSTN) gene 5′-flanking variants and their relationship with horse biometric traits. J. Equine Vet. Sci. 60, 8389.Google Scholar
Cieslak, M., Reissmann, M., Hofreiter, M., and Ludwig, A. (2011). Colours of domestication. Biol. Rev. 86, 885899.Google Scholar
Clark, D.A., Mitra, P.P., and Wang, S.S.-H. (2001). Scalable architecture in mammalian brains. Nature 411, 189193.Google Scholar
Clark, E. (2017). Dynamic patterning by the Drosophila pair-rule network reconciles long-germ and short-germ segmentation. PLoS Biol. 15(9), e2002439.Google Scholar
Clark, W.C. and Russell, M.A. (1977). The correlation of lysosomal activity and adult phenotype in a cell-lethal mutant of Drosophila. Dev. Biol. 57, 160173.Google Scholar
Cline, T.W. (1993). The Drosophila sex determination signal: how do flies count to two? Trends Genet. 9, 385390.Google Scholar
Cloutier, R., Clement, A.M., Lee, M.S.Y., Noël, R., Béchard, I., Roy, V., and Long, J.A. (2020). Elpistostege and the origin of the vertebrate hand. Nature 579, 549554.Google Scholar
Cobourne, M.T. and Sharpe, P.T. (2003). Tooth and jaw: molecular mechanisms of patterning in the first branchial arch. Arch. Oral Biol. 48, 114.Google Scholar
Cock, A.G. and Forsdyke, D.R. (2008). Treasure Your Exceptions: The Science and Life of William Bateson. Springer, New York.Google Scholar
Coen, E. (1999). The Art of Genes: How Organisms Make Themselves. Oxford University Press, New York.Google Scholar
Cohen, M.M. Jr. (2006). Holoprosencephaly: clinical, anatomic, and molecular dimensions. Birth Defects Res. A Clin. Mol. Teratol. 76, 658673.Google Scholar
Cohen, S.M. (1993). Imaginal disc development. In The Development of Drosophila melanogaster, Bate, M. and Martinez Arias, A., editors. Cold Spring Harbor Laboratory Press, Plainview, NY, pp. 747841.Google Scholar
Cohen, S.M. (2003). Long-range signalling by touch. Nature 426, 503504.Google Scholar
Coile, D.C. (2005). Encyclopedia of Dog Breeds. Barron’s Educational Series, Hauppauge, NY.Google Scholar
Colas, J.-F. and Schoenwolf, G.C. (2001). Towards a cellular and molecular understanding of neurulation. Dev. Dynamics 221, 117145.Google Scholar
Coletti, S.M., Ide, C.F., Blankenau, A.J., and Meyer, R.L. (1990). Ocular dominance stripe formation by regenerated isogenic double temporal retina in Xenopus laevis. J. Neurobiol. 21, 276282.Google Scholar
Collins, T.N., Mao, Y., Li, H., Bouaziz, M., Hong, A., Feng, G.-S., Wang, F., Quilliam, L.A., Chen, L., Park, T., Curran, T., and Zhang, X. (2018). Crk proteins transduce FGF signaling to promote lens fiber cell elongation. eLife 7, e32586.Google Scholar
Condic, M.L., Fristrom, D., and Fristrom, J.W. (1991). Apical cell shape changes during Drosophila imaginal leg disc elongation: a novel morphogenetic mechanism. Development 111, 2333.Google Scholar
Conlin, L.K., Thiel, B.D., Bonnemann, C.G., Medne, L., Ernst, L.M., Zackai, E.H., Deardorff, M.A., Krantz, I.D., Hakonarson, H., and Spinner, N.B. (2010). Mechanisms of mosaicism, chimerism and uniparental disomy identified by single nucleotide polymorphism array analysis. Hum. Mol. Genet. 19, 12631275.Google Scholar
Constantine-Paton, M. and Law, M.I. (1978). Eye-specific termination bands in tecta of three-eyed frogs. Science 202, 639641.Google 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, 185217.Google 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, 455476.Google Scholar
Cooper, K.L. (2019). Developmental and evolutionary allometry of the mammalian limb skeleton. Integr. Comp. Biol. 59, 13561368.Google Scholar
Cooper, S.B. and Van Leeuwen, J., eds. (2013). Alan Turing: His Work and Impact. Elsevier, New York.Google Scholar
Corallo, D., Trapani, V., and Bonaldo, P. (2015). The notochord: structure and functions. Cell. Mol. Life Sci. 72, 29893008.Google Scholar
Cordero, R.J.B. and Casadevall, A. (2020). Melanin. Curr. Biol. 30, R142R143.Google Scholar
Cordingley, J.E., Sundaresan, S.R., Fischhoff, I.R., Shapiro, B., Ruskey, J., and Rubenstein, D.I. (2009). Is the endangered Grevy’s zebra threatened by hybridization? Anim. Conserv. 12, 505513.Google Scholar
Córdoba, S. and Estella, C. (2018). The transcription factor Dysfusion promotes fold and joint morphogenesis through regulation of Rho1. PLoS Genet. 14(8), e1007584.Google Scholar
Corona, M., Libbrecht, R., and Wheeler, D.E. (2016). Molecular mechanisms of phenotypic plasticity in social insects. Curr. Opin. Insect Sci. 13, 5560.Google Scholar
Corson, F., Couturier, L., Roualt, H., Mazouni, K., and Schweisguth, F. (2017). Self-organized Notch dynamics generate stereotyped sensory organ patterns in Drosophila. Science 356, 501.Google Scholar
Cortes, C., Francou, A., De Bono, C., and Kelly, R.G. (2018). Epithelial properties of the second heart field. Circ. Res. 122, 142154.Google Scholar
Coulombre, J.L. and Coulombre, A.J. (1963). Lens development: fiber elongation and lens orientation. Science 142, 14891490.Google Scholar
Courgeon, M. and Desplan, C. (2019). Coordination between stochastic and deterministic specification in the Drosophila visual system. Science 366, 325.Google Scholar
Couso, J.P., Bishop, S.A., and Martinez Arias, A. (1994). The wingless signalling pathway and the patterning of the wing margin in Drosophila. Development 120, 621636.Google Scholar
Coutelis, J.-B., González-Morales, N., Géminard, C., and Noselli, S. (2014). Diversity and convergence in the mechanisms establishing L/R asymmetry in metazoa. EMBO Rep. 15, 926937.Google Scholar
Cozzitorto, C. and Spagnoli, F.M. (2019). Pancreas organogenesis: the interplay between surrounding microenvironment(s) and epithelium-intrinsic factors. Curr. Top. Dev. Biol. 132, 221256.Google Scholar
Cranford, T.W., Amundin, M., and Norris, K.S. (1996). Functional morphology and homology in the odontocete nasal complex: implications for sound generation. J. Morphol. 228, 223285.Google Scholar
Creel, D., Garber, S.R., King, R.A., and Witkop, C.J. Jr. (1980). Auditory brainstem anomalies in human albinos. Science 209, 12531255.Google Scholar
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, 141151.Google Scholar
Crews, D. (2003). Sex determination: where environment and genetics meet. Evol. Dev. 5, 5055.Google Scholar
Crispo, E. (2007). The Baldwin effect and genetic assimilation: revisiting two mechanisms of evolutionary change mediated by phenotypic plasticity. Evolution 61, 24692479.Google Scholar
Crow, J.F. and Bender, W. (2004). Edward B. Lewis, 1918–2004. Genetics 168, 17731783.Google Scholar
Cubas, P., de Celis, J.-F., Campuzano, S., and Modolell, J. (1991). Proneural clusters of achaete-scute expression and the generation of sensory organs in the Drosophila imaginal wing disc. Genes Dev. 5, 9961008.Google Scholar
Cubeñas-Potts, C. and Corces, V.G. (2015). Architectural proteins, transcription, and the three-dimensional organization of the genome. FEBS Lett. 589, 29232930.Google Scholar
Cummins, H. and Midlo, C. (1943). Finger Prints, Palms and Soles: An Introduction to Dermatoglyphics. Dover, New York.Google Scholar
Currie, A. (2013). Convergence as evidence. Br. J. Philos. Sci. 64, 763786.Google Scholar
Curtis, A.S.G. (1960). Cortical grafting in Xenopus laevis. J. Embryol. Exp. Morphol. 8, 163173.Google Scholar
Curtis, A.S.G. (1962). Morphogenetic interactions before gastrulation in the amphibian, Xenopus laevis: the cortical field. J. Embryol. Exp. Morphol. 10, 410422.Google Scholar
Curtiss, J., Halder, G., and Mlodzik, M. (2002). Selector and signalling molecules cooperate in organ patterning. Nat. Cell Biol. 4, E48E51.Google Scholar
Cvekl, A. and Ashery-Padan, R. (2014). The cellular and molecular mechanisms of vertebrate lens development. Development 141, 44324447.CrossRefGoogle ScholarPubMed
Cvekl, A. and Zhang, X. (2017). Signaling and gene regulatory networks in mammalian lens development. Trends Genet. 33, 677702.Google Scholar
D’Souza, B., Meloty-Kapella, C., and Weinmaster, G. (2010). Canonical and non-canonical Notch ligands. Curr. Top. Dev. Biol. 92, 73129.Google Scholar
Dall’Olio, S., Fontanesi, L., Costa, L.N., Tassinari, M., Minieri, L., and Falaschini, A. (2010). Analysis of horse myostatin gene and identification of single nucleotide polymorphisms in breeds of different morphological types. J. Biomed. Biotech. 2010, 542945.Google Scholar
Dall’Olio, S., Wang, Y., Sartori, C., Fontanesi, L., and Mantovani, R. (2014). Association of myostatin (MSTN) gene polymorphisms with morphological traits in the Italian Heavy Draft Horse breed. Livestock Sci. 160, 2936.Google Scholar
Darbellay, F. and Duboule, D. (2016). Topological domains, metagenes, and the emergence of pleiotropic regulations at Hox loci. Curr. Top. Dev. Biol. 116, 299314.Google Scholar
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. (1872). The Expression of the Emotions in Man and Animals. John Murray, London.Google Scholar
Dasgupta, A. and Amack, J.D. (2016). Cilia in left–right patterning. Philos. Trans. R. Soc. Lond. B 371, 20150410.Google Scholar
DasGupta, R. and Fuchs, E. (1999). Multiple roles for activated LEF/TCF transcription complexes during hair follicle development and differentiation. Development 126, 45574568.Google Scholar
Davidson, L.A. (2012). Epithelial machines that shape the embryo. Trends Cell Biol. 22, 8287.Google Scholar
Davidson, L.A. (2017). Mechanical design in embryos: mechanical signalling, robustness and developmental defects. Philos. Trans. R. Soc. Lond. B 372, 20150516.Google Scholar
Davies-Thompson, J., Scheel, M., Lanyon, L.J., and Barton, J.J.S. (2013). Functional organisation of visual pathways in a patient with no optic chiasm. Neuropsychologia 51, 12601272.Google Scholar
Davis, A.P., Witte, D.P., Hsieh-Li, H.M., Potter, S.S., and Capecchi, M.R. (1995). Absence of radius and ulna in mice lacking hoxa-11 and hoxd-11. Nature 375, 791795.Google 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.Google Scholar
Dawkins, R. (1996). Climbing Mount Improbable. Norton, New York.Google Scholar
Day, S.J. and Lawrence, P.A. (2000). Measuring dimensions: the regulation of size and shape. Development 127, 29772987.Google Scholar
de Beer, G. (1958). Embryos and Ancestors, 3rd ed. Clarendon Press, Oxford.Google Scholar
de Celis, J.F., García-Bellido, A., and Bray, S.J. (1996). Activation and function of Notch at the dorsal–ventral boundary of the wing imaginal disc. Development 122, 359369.Google Scholar
de Celis, J.F., Tyler, D.M., de Celis, J., and Bray, S.J. (1998). Notch signalling mediates segmentation of the Drosophila leg. Development 125, 46174626.Google 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, 555559.Google Scholar
de Juan Romero, C. and Borrell, V. (2017). Genetic maps and patterns of cerebral cortex folding. Curr. Opin. Cell Biol. 49, 3137.Google Scholar
De Pascalis, C. and Etienne-Manneville, S. (2017). Single and collective cell migration: the mechanics of adhesions. Mol. Biol. Cell 28, 18331846.Google Scholar
De Robertis, E.M. (2009). Spemann’s organizer and the self-regulation of embryonic fields. Mech. Dev. 126, 925941.Google Scholar
De Robertis, E.M., Morita, E.A., and Cho, K.W.Y. (1991). Gradient fields and homeobox genes. Development 112, 669678.Google Scholar
De Robertis, E.M., Moriyama, Y., and Colozza, G. (2017). Generation of animal form by the Chordin/Tolloid/BMP gradient: 100 years after D’Arcy Thompson. Dev. Growth Differ. 59, 580592.Google Scholar
Deane-Coe, P.E., Chu, E.T., Slavney, A., Boyko, A.R., and Sams, A.J. (2018). Direct-to-consumer DNA testing of 6,000 dogs reveals 98.6-kb duplication associated with blue eyes and heterochromia in Siberian Huskies. PLoS Genet. 14(10), e1007648Google Scholar
Degabriele, R. (1980). The physiology of the koala. Sci. Am. 243(1), 110117.Google Scholar
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, 19751982.Google Scholar
Delgado, I. and Torres, M. (2016). Gradients, waves and timers: an overview of limb patterning models. Semin. Cell Dev. Biol. 49, 109115.Google Scholar
Delgado, I. and Torres, M. (2017). Coordination of limb development by crosstalk among axial patterning pathways. Dev. Biol. 429, 382386.Google Scholar
Delpretti, S., Zakany, J., and Duboule, D. (2012). A function for all posterior Hoxd genes during digit development? Dev. Dynamics 241, 792802.Google Scholar
Deng-Lobnig, M. and Martin, A.C. (2020). Divergent and combinatorial mechanical strategies that promote epithelial folding during morphogenesis. Curr. Opin. Genet. Dev. 63, 2429.Google Scholar
Depew, M.J., Lufkin, T., and Rubenstein, J.L.R. (2002). Specification of jaw subdivisions by Dlx genes. Science 298, 381385.Google Scholar
Deschamps, J. (2008). Tailored Hox gene transcription and the making of the thumb. Genes Dev. 22, 293296.Google Scholar
Deutsch, J.S., ed. (2010). Hox Genes: Studies from the 20th to the 21st Century. Advances in Experimental Medicine and Biology. Landes Bioscience, New York.Google Scholar
Devenport, D. (2016). Tissue morphodynamics: translating planar polarity cues into polarized cell behaviors. Semin. Cell Dev. Biol. 55, 99110.Google Scholar
Diaz de la Loza, M.C. and Thompson, B.J. (2017). Forces shaping the Drosophila wing. Mech. Dev. 144, 2332.Google Scholar
Diaz de la Loza, M.C., Loker, R., Mann, R.S., and Thompson, B.J. (2020). Control of tissue morphogenesis by the HOX gene Ultrabithorax. Development 147, dev184564.Google Scholar
Diaz de la Loza, M.C., Ray, R.P., Ganguly, P.S., Alt, S., Davis, J.R., Hoppe, A., Tapon, N., Salbreux, G., and Thompson, B.J. (2018). Apical and basal remodeling control epithelial morphogenesis. Dev. Cell 46, 2339.Google Scholar
Dickerson, B.H., de Souza, A.M., Huda, A., and Dickinson, M.H. (2019). Flies regulate wing motion via active control of a dual-function gyroscope. Curr. Biol. 29, 35173524.Google Scholar
Dickinson, M.H. (1999). Haltere-mediated equilibrium reflexes of the fruit fly, Drosophila melanogaster. Philos. Trans. R. Soc. Lond. B 354, 903916.Google Scholar
Dieters, J., Kowalczyk, W., and Seidl, T. (2016). Simultaneous optimisation of earwig hindwings for flight and folding. Biol. Open 5, 638644.Google Scholar
Dietrich, M.R. (2000). From hopeful monsters to homeotic effects: Richard Goldschmidt’s integration of development, evolution, and genetics. Am. Zool. 40, 738747.Google Scholar
Dietrich, M.R. (2003). Richard Goldschmidt: hopeful monsters and other “heresies”. Nat. Rev. Genet. 4, 6874.Google Scholar
Diogo, R. (2017). Evolution Driven by Organismal Behavior: A Unifying View of Life, Function, Form, Mismatches, and Trends. Springer Nature, Cham, Switzerland.Google Scholar
Diogo, R., Guinard, G., and Diaz, R.E. Jr. (2017). Dinosaurs, chameleons, humans, and evo-devo path: linking Étienne Geoffroy’s teratology, Waddington’s homeorhesis, Alberch’s logic of “monsters,” and Goldschmidt hopeful “monsters”. J. Exp. Zool. B Mol. Dev. Evol. 328, 207229.Google Scholar
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, 196214.Google Scholar
Diogo, R., Smith, C.M., and Ziermann, J.M. (2015). Evolutionary developmental pathology and anthropology: a new field linking development, comparative anatomy, human evolution, morphological variations and defects, and medicine. Dev. Dynamics 244, 13571374.Google Scholar
Dittrich-Reed, D.R. and Fitzpatrick, B.M. (2013). Transgressive hybrids as hopeful monsters. Evol. Biol. 40, 310315.Google Scholar
Docampo, M.J., Zanna, G., Fondevila, D., Cabrera, J., López-Iglesias, C., Carvalho, A., Cerrato, S., Ferrer, L., and Bassols, A. (2011). Increased HAS2-driven hyaluronic acid synthesis in shar-pei dogs with hereditary cutaneous hyaluronosis (mucinosis). Vet. Dermatol. 22, 535545.Google Scholar
Doe, C.Q. and Spana, E.P. (1995). A collection of cortical crescents: asymmetric protein localization in CNS precursor cells. Neuron 15, 991995.Google Scholar
Doherty, D., Feger, G., Younger-Shepherd, S., Jan, L.Y., and Jan, Y.N. (1996). Delta is a ventral to dorsal signal complementary to Serrate, another Notch ligand, in Drosophila wing formation. Genes Dev. 10, 421434.Google Scholar
Doiguchi, M., Nakagawa, T., Imamura, Y., Yoneda, M., Higashi, M., Kubota, K., Yamashita, S., Asahara, H., Iida, M., Fujii, S., Ikura, T., Liu, Z., Nandu, T., Kraus, W.L., Ueda, H., and Ito, T. (2016). SMARCAD1 is an ATP-dependent stimulator of nucleosomal H2A acetylation via CBP, resulting in transcriptional regulation. Sci. Rep. 6, 20179.Google Scholar
Domingos, P.M., Jenny, A., Combie, K.F., del Alamo, D., Mlodzik, M., Steller, H., and Mollereau, B. (2019). Regulation of Numb during planar cell polarity establishment in the Drosophila eye. Mech. Dev. 160, 103583.Google Scholar
Donahue, C.J., Glasser, M.F., Preuss, T.M., Rilling, J.K., and Van Essen, D.C. (2018). Quantitative assessment of prefrontal cortex in humans relative to nonhuman primates. PNAS 115, E5183E5192.Google Scholar
Donahue, C.J., Glasser, M.F., Preuss, T.M., Rilling, J.K., and Van Essen, D.C. (2019). Reply to Barton and Montgomery: a case for preferential prefrontal cortical expansion. PNAS 116, 56.Google Scholar
Dongen, S.V. (2006). Fluctuating asymmetry and developmental instability in evolutionary biology: past, present and future. J. Evol. Biol. 19, 17271743.Google Scholar
Donnelly, D.E. and Morrison, P.J. (2014). Hereditary gigantism: the biblical giant Goliath and his brothers. Ulster Med. J. 83, 8688.Google Scholar
Dougoud, M., Mazza, C., Schwaller, B., and Pecze, L. (2019). Extending the mathematical palette for developmental pattern formation: piebaldism. Bull. Math. Biol. 81, 14611478.Google Scholar
Dover, G. (2000). How genomic and developmental dynamics affect evolutionary processes. BioEssays 22, 11531159.Google Scholar
Drimmer, F. (1973). Very Special People. Amjon Publishers, New York.Google Scholar
Driscoll, C.A., Clutton-Brock, J., Kitchener, A.C., and O’Brien, S.J. (2009). The taming of the cat. Sci. Am. 300(6), 6875.Google Scholar
Drögemüller, C., Karlsson, E.K., Hytönen, M.K., Perloski, M., Dolf, G., Sainio, K., Lohi, H., Lindblad-Toh, K., and Leeb, T. (2008). A mutation in hairless dogs implicates FOXI3 in ectodermal development. Science 321, 1462.Google Scholar
Duan, D., Xia, S., Rekik, I., Wu, Z., Wang, L., Lin, W., Gilmore, J.H., Shen, D., and Li, G. (2020). Individual identification and individual variability analysis based on cortical folding features in developing infant singletons and twins. Hum. Brain Mapp. 41, 19852003.Google Scholar
Duboule, D. (2007). The rise and fall of Hox gene clusters. Development 134, 25492560.Google Scholar
Duncan, I. and Montgomery, G. (2002). E. B. Lewis and the bithorax complex: part I. Genetics 160, 12651272.Google Scholar
Duncan, I. and Montgomery, G. (2002). E. B. Lewis and the bithorax complex: part II. From cis–trans test to the genetic control of development. Genetics 161, 110.Google Scholar
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., editors. Elsevier Academic Press, New York, pp. 131158.Google Scholar
Dybus, A., Proskura, W.S., Sadkowski, S., and Pawlina, E. (2013). A single nucleotide polymorphism in exon 3 of the myostatin gene in different breeds of domestic pigeon (Columba livia var. domestica). Vet. Med. (Praha) 58, 3238.Google Scholar
Ebisuya, M. and Briscoe, J. (2018). What does time mean in development? Development 145, dev164368.Google Scholar
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. Nat. Genet. 44, 348352.Google Scholar
Ede, D.A. (1972). Cell behaviour and embryonic development. Int. J. Neurosci. 3, 165174.Google Scholar
Eder, D., Aegerter, C., and Basler, K. (2017). Forces controlling organ growth and size. Mech. Dev. 114, 5361.Google Scholar
Edgar, B.A. (2006). How flies get their size: genetics meets physiology. Nat. Rev. Genet. 7, 907916.Google Scholar
Edgar, B.A. and Orr-Weaver, T.L. (2001). Endoreplication cell cycles: more for less. Cell 105, 297306.Google Scholar
Edwards, J.S. (1994). In memoriam. Sir Vincent Brian Wigglesworth (1899–1994). Dev. Biol. 166, 361362.Google Scholar
Edwards, J.S. (1998). Sir Vincent Wigglesworth and the coming of age of insect development. Int. J. Dev. Biol. 42, 471473.Google Scholar
Efstratiadis, A. (1998). Genetics of mouse growth. Int. J. Dev. Biol. 42, 955976.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, 267275.Google Scholar
Elgjo, K. and Reichelt, K.L. (2004). Chalones: from aqueous extracts to oligopeptides. Cell Cycle 3, 12081211.Google Scholar
Elliott, K.L., Houston, D.W., and Fritzsch, B. (2015). Sensory afferent segregation in three-eared frogs resemble the dominance columns observed in three-eyed frogs. Sci. Rep. 5, 8338.Google Scholar
Elsdale, T. and Wasoff, F. (1976). Fibroblast cultures and dermatoglyphics: the topology of two planar patterns. W. Roux Arch. Dev. Biol. 180, 121147.Google Scholar
Emerson, S.B. (1985). Jumping and leaping. In Functional Vertebrate Morphology, Hildebrand, M., Bramble, D.M., Liem, K.F., and Wake, D.B., editors. Harvard University Press, Cambridge, MA, pp. 5872.Google Scholar
Emlen, D.J. (2008). The evolution of animal weapons. Annu. Rev. Ecol. Evol. Syst. 39, 387413.Google Scholar
Emlen, D.J. (2014). Animal Weapons: The Evolution of Battle. Henry Holt, New York.Google 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, 860864.Google Scholar
Enard, D., Depaulis, F., and Crollius, H.R. (2010). Human and non-human primate genomes share hotspots of positive selection. PLoS Genet. 6(2), e1000840.Google Scholar
Eom, D.S., Bain, E.J., Patterson, L.B., Grout, M.E., and Parichy, D.M. (2015). Long-distance communication by specialized cellular projections during pigment pattern development and evolution. eLife 4, e12401.Google Scholar
Erickson, J.R. and Echeverri, K. (2018). Learning from regeneration research organisms: the circuitous road to scar free wound healing. Dev. Biol. 433, 144154.Google Scholar
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, 173198.Google Scholar
Estrellas, K.M., Chung, L., Cheu, L.A., Sadtler, K., Majumdar, S., Mula, J., Wolf, M.T., Elisseeff, J.H., and Wagner, K.R. (2018). Biological scaffold-mediated delivery of myostatin inhibitor promotes a regenerative immune response in an animal model of Duchenne muscular dystrophy. J. Biol. Chem. 293, 1559415605.Google Scholar
Etienne-Manneville, S. (2011). Control of polarized cell morphology and motility by adherens junctions. Semin. Cell Dev. Biol. 22, 850857.Google Scholar
Etienne-Manneville, S. (2014). Neighborly relations during collective migration. Curr. Opin. Cell Biol. 30, 5159.Google Scholar
Falk, D., Lepore, F.E., and Noe, A. (2013). The cerebral cortex of Albert Einstein: a description and preliminary analysis of unpublished photographs. Brain 136, 13041327.Google Scholar
Fankhauser, G. (1945). The effects of changes in chromosome number on amphibian development. Q. Rev. Biol. 20, 2078.Google Scholar
Fantauzzo, K.A., Tadin-Strapps, M., You, Y., Mentzer, S.E., Baumeister, F.A.M., Cianfarani, S., Van Maldergem, L., Warburton, D., Sundberg, J.P., and Christiano, A.M. (2008). A position effect on TRPS1 is associated with Ambras syndrome in humans and the Koala phenotype in mice. Hum. Mol. Genet. 17, 35393551.Google Scholar
Fehilly, C.B., Willadsen, S.M., and Tucker, E.M. (1984). Interspecific chimaerism between sheep and goat. Nature 307, 634636.Google Scholar
Feigin, C.Y. and Mallarino, R. (2018). Setting the bar: analyzing the genomes of rock pigeons demonstrates that genetic variation comes in many forms and can have unexpected origins. eLife 7, e39068.Google Scholar
Fernandes, J., Celniker, S.E., Lewis, E.B., and VijayRaghavan, K. (1994). Muscle development in the four-winged Drosophila and the role of the Ultrabithorax gene. Curr. Biol. 4, 957964.Google Scholar
Fernández, V., Llinares-Benadero, C., and Borrell, V. (2016). Cerebral cortex expansion and folding: what have we learned? EMBO J. 35, 10211044.Google Scholar
Ferree, P.L., Deneke, V.E., and Di Talia, S. (2016). Measuring time during early embryonic development. Semin. Cell Dev. Biol. 55, 8088.Google Scholar
Ferreira, R.R., Fukui, H., Chow, R., Vilfan, A., and Vermot, J. (2019). The cilium as a force sensor: myth versus reality. J. Cell Sci. 132, jcs213496.Google Scholar
Figuera, L.E., Pandolfo, M., Dunne, P.W., Cantú, J.M., and Patel, P.I. (1995). Mapping of the congenital generalized hypertrichosis locus to chromosome Xq24–q27.1. Nat. Genet. 10, 202207.Google Scholar
Findlay, G.H. and Harris, W.F. (1977). The topology of hair streams and whorls in man, with an observation on their relationship to epidermal ridge patterns. Am. J. Phys. Anthrop. 46, 427438.Google Scholar
Finet, C., Decaras, A., Armisén, D., and Khila, A. (2018). The achaete-scute complex contains a single gene that controls bristle development in the semi-aquatic bugs. Proc. R. Soc. B 285, 20182387.Google Scholar
Finlay, B.L. and Huang, K. (2020). Developmental duration as an organizer of the evolving mammalian brain: scaling, adaptations, and exceptions. Evol. Dev. 22, 181195.Google Scholar
Finlay, B.L., Darlington, R.B., and Nicastro, N. (2001). Developmental structure in brain evolution. Behav. Brain Sci. 24, 263308.Google Scholar
Fisher, A. and Caudy, M. (1998). The function of hairy-related bHLH repressor proteins in cell fate decisions. BioEssays 20, 298306.Google Scholar
Fitch, C.L., Girton, J.R., and Girton, L. (1992). The suppressor of forked locus in Drosophila melanogaster: genetic and molecular analyses. Genetica 85, 185203.Google Scholar
Fondon, J.W. III and Garner, H.R. (2004). Molecular origins of rapid and continuous morphological evolution. PNAS 101, 1805818063.Google Scholar
Fouilloux, C., Ringler, E., and Rojas, B. (2019). Cannibalism. Curr. Biol. 29, R1295R1297.Google Scholar
Francavilla, A., Ove, P., Polimeno, L., Coetzee, M., Makowka, L., Barone, M., Vanthiel, D.H., and Starzl, T.E. (1988). Regulation of liver size and regeneration: importance in liver-transplantation. Transplant. Proc. 20, 494497.Google Scholar
François, L., Fegraeus, K.J., Eriksson, S., Andersson, L.S., Tesfayonas, Y.G., Viluma, A., Imsland, F., Buys, N., Mikko, S., Lindgren, G., and Velie, B.D. (2016). Conformation traits and gaits in the Icelandic horse are associated with genetic variants in myostatin (MSTN). J. Hered. 107, 431437.Google Scholar
Frank, S.A. (2014). Somatic mosaicism and disease. Curr. Biol. 24, R577R581.Google Scholar
Frantsevich, L. (2016). A Houdini’s trick in a fly: leg unfolding with the aid of transient hinges in an extricating Calliphora vicina (Diptera: Calliphoridae). Arthropod Struct. Dev. 45, 213.Google Scholar
French, V., Bryant, P.J., and Bryant, S.V. (1976). Pattern regulation in epimorphic fields. Science 193, 969981.Google Scholar
Freytes, D.O., Wan, L.Q., and Vunjak-Novakovic, G. (2009). Geometry and force control of cell function. J. Cell. Biochem. 108, 10471058.Google Scholar
Fristrom, D. (1988). The cellular basis of epithelial morphogenesis: a review. Tissue Cell 20, 645690.Google Scholar
Fristrom, D. and Fristrom, J.W. (1993). The metamorphic development of the adult epidermis. In The Development of Drosophila melanogaster, Bate, M. and Martinez Arias, A., editors. Cold Spring Harbor Laboratory Press, Plainview, NY, pp. 843897.Google Scholar
Fristrom, D.K., Fekete, E., and Fristrom, J.W. (1981). Imaginal disc development in a non-pupariating lethal mutant in Drosophila melanogaster. W. Roux Arch. Dev. Biol. 190, 1121.Google Scholar
Fromental-Ramain, C., Warot, X., Lakkaraju, S., Favier, B., Haack, H., Birling, C., Dierich, A., Dollé, P., and Chambon, P. (1996). Specific and redundant functions of the paralogous Hoxa-9 and Hoxd-9 genes in forelimb and axial skeleton patterning. Development 122, 461472.Google Scholar
Fuchs, E. (2007). Scratching the surface of skin development. Nature 445, 834842.Google Scholar
Fujisawa, Y., Kosakamoto, H., Chihara, T., and Miura, M. (2019). Non-apoptotic function of Drosophila caspase activation in epithelial thorax closure and wound healing. Development 146, dev169037.Google Scholar
Furman, D.P. and Bukharina, T.A. (2009). The gene network determining development of Drosophila melanogaster mechanoreceptors. Comp. Biol. Chem. 33, 231234.Google Scholar
Furman, D.P. and Bukharina, T.A. (2018). The bristle pattern development in Drosophila melanogaster: the prepattern and achaete-scute genes. Vavilov J. Genet. Breed. 22, 10461054.Google Scholar
Fusco, G. and Minelli, A. (2010). Phenotypic plasticity in development and evolution: facts and concepts. Philos. Trans. R. Soc. Lond. B 365, 547556.Google Scholar
Galant, R. and Carroll, S.B. (2002). Evolution of a transcriptional repression domain in an insect Hox protein. Nature 415, 910913.Google Scholar
Galindo, M.I., Bishop, S.A., and Couso, J.P. (2005). Dynamic EGFR-Ras signalling in Drosophila leg development. Dev. Dynamics 233, 14961508.Google Scholar
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, 256259.Google Scholar
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, 637646.Google Scholar
Galloni, M., Gyurkovics, H., Schedl, P., and Karch, F. (1993). The bluetail transposon: evidence for independent cis-regulatory domains and domain boundaries in the bithorax complex. EMBO J. 12, 10871097.Google Scholar
Gandolfi, B., Outerbridge, C.A., Beresford, L.G., Myers, J.A., Pimentel, M., Alhaddad, H., Grahn, J.C., Grahn, R.A., and Lyons, L.A. (2010). The naked truth: Sphynx and Devon Rex cat breed mutations in KRT71. Mamm. Genome 21, 509515.Google Scholar
Gao, B., Song, H., Bishop, K., Elliot, G., Garrett, L., English, M.A., Andre, P., Robinson, J., Sood, R., Minami, Y., Economides, A.N., and Yang, Y. (2011). Wnt signaling gradients establish planar cell polarity by inducing Vangl2 phosphorylation through Ror2. Dev. Cell 20, 163176.Google Scholar
Garcia, K.E., Kroenke, C.D., and Bayly, P.V. (2018). Mechanics of cortical folding: stress, growth and stability. Philos. Trans. R. Soc. Lond. B 373, 20170321.Google Scholar
García-Bellido, A. (1975). Genetic control of wing disc development in Drosophila. In Cell Patterning, Porter, R. and Rivers, J., editors. Elsevier, Amsterdam, pp. 161182.Google 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, 631639.Google Scholar
García-Bellido, A. and Merriam, J.R. (1969). Cell lineage of the imaginal discs in Drosophila gynandromorphs. J. Exp. Zool. 170, 6176.Google Scholar
García-Bellido, A., Lawrence, P.A., and Morata, G. (1979). Compartments in animal development. Sci. Am. 241(1), 102110.Google Scholar
Garcia-Cruz, D., Figuera, L.E., and Cantu, J.M. (2002). Inherited hypertrichoses. Clin. Genet. 61, 321329.Google Scholar
Gardner, E.W., Miller, H.M., and Lowney, E.D. (1979). Folded skin associated with underlying nevus lipomatosus. Arch. Derm. 115, 978979.Google Scholar
Gardner, S. (2015). #ThrowBackThursday: the toad and I. TheSpec.com (Jan. 15, 2015).Google Scholar
Garzón-Alvarado, D.A. and Ramirez Martinez, A.M. (2011). A biochemical hypothesis on the formation of fingerprints using a Turing patterns approach. Theor. Biol. Med. Model. 8, 24.Google Scholar
Gawne, R., McKenna, K.Z., and Nijhout, H.F. (2018). Unmodern synthesis: developmental hierarchies and the origin of phenotypes. BioEssays 40, 1600265.Google Scholar
Gayon, J. (2000). History of the concept of allometry. Am. Zool. 40, 748758.Google Scholar
Gebo, D.L. (1987). Functional anatomy of the tarsier foot. Am. J. Phys. Anthrop. 73, 931.Google Scholar
Gee, H. (2013). The Accidental Species: Misunderstandings of Human Evolution. University of Chicago Press, Chicago, IL.Google Scholar
Gehring, W.J. (2012). The animal body plan, the prototypic body segment, and eye evolution. Evol. Dev. 14, 3446.CrossRefGoogle ScholarPubMed
Géminard, C., González-Morales, N., Coutelis, J.B., and Noselli, S. (2014). The Myosin ID pathway and left–right asymmetry in Drosophila. Genesis 52, 471480.Google Scholar
Gerhart, J. (1999). Signaling pathways in development (1998 Warkany lecture). Teratology 60, 226239.Google Scholar
Gerhart, J. (2001). Evolution of the organizer and the chordate body plan. Int. J. Dev. Biol. 45, 133153.Google Scholar
Gerhart, J. (2002). Changing the axis changes the perspective. Dev. Dynamics 225, 380383.Google Scholar
Gerhart, J. (2010). Enzymes, embryos, and ancestors. Annu. Rev. Cell Dev. Biol. 26, 120.Google 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), 85828589.Google Scholar
Gerhart, J., Ubbels, G., Black, S., Hara, K., and Kirschner, M. (1981). A reinvestigation of the role of the grey crescent in axis formation in Xenopus laevis. Nature 292, 511516.Google Scholar
Gerhart, J.C. (1987). Determinants of early amphibian development. Am. Zool. 27, 593605.Google Scholar
Germani, F., Bergantinos, C., and Johnston, L.A. (2018). Mosaic analysis in Drosophila. Genetics 208, 473490.Google Scholar
Geyer, P.K. and Corces, V.G. (1987). Separate regulatory elements are responsible for the complex pattern of tissue-specific and developmental transcription of the yellow locus in Drosophila melanogaster. Genes Dev. 1, 9961004.Google Scholar
Gho, M., Bellaïche, Y., and Schweisguth, F. (1999). Revisiting the Drosophila microchaete lineage: a novel intrinsically asymmetric cell division generates a glial cell. Development 126, 35733584.Google Scholar
Ghysen, A. (2009). Ontogeny of an adventurous mind: the origin of Antonio García-Bellido’s contributions to developmental genetics. Int. J. Dev. Biol. 53, 12771290.Google Scholar
Ghysen, A. and Dambly-Chaudière, C. (1988). From DNA to form: the achaete-scute complex. Genes Dev. 2, 495501.Google Scholar
Ghysen, A. and Dambly-Chaudière, C. (1989). Genesis of the Drosophila peripheral nervous system. Trends Genet. 5, 251255.Google Scholar
Gibert, J.-M. and Simpson, P. (2003). Evolution of cis-regulation of the proneural genes. Int. J. Dev. Biol. 47, 643651.Google Scholar
Gibson, G. and Hogness, D.S. (1996). Effect of polymorphism in the Drosophila regulatory gene Ultrabithorax on homeotic stability. Science 271, 200203.Google Scholar
Gibson, M.C. (2019). Commentary on “Regeneration, duplication and transdetermination in fragments of the leg disc of Drosophila melanogaster”: Schubiger, G. (1971). Dev. Biol. 449, 6382.Google Scholar
Giebel, B. and Wodarz, A. (2012). Notch signaling: Numb makes the difference. Curr. Biol. 22, R133R135.Google Scholar
Giebel, L.B., Tripathi, R.K., King, R.A., and Spritz, R.A. (1991). A tyrosinase gene missense mutation in temperature-sensitive Type I oculocutaneous albinism: a human homologue to the Siamese cat and the Himalayan mouse. J. Clin. Invest. 87, 11191122.Google Scholar
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. 163183.Google Scholar
Gilbert, S.F. (2001). Ecological developmental biology: developmental biology meets the real world. Dev. Biol. 233, 112.Google Scholar
Gilbert, S.F. (2014). Developmental Biology, 10th ed. Sinauer, Sunderland, MA.Google Scholar
Gilbert, S.F. (2016). Developmental plasticity and developmental symbiosis: the return of eco-devo. Curr. Top. Dev. Biol. 116, 415433.Google Scholar
Gilbert, S.F. and Barresi, M.J.F. (2019). Developmental Biology, 11th ed. Sinauer, Sunderland, MA.Google Scholar
Gilgenkrantz, H. and de l’Hortet, A.C. (2018). Understanding liver regeneration: from mechanisms to regenerative medicine. Am J. Pathol. 188, 13161327.Google Scholar
Gillham, N.W. (2001). Evolution by jumps: Francis Galton and William Bateson and the mechanism of evolutionary change. Genetics 159, 13831392.Google Scholar
Gilmour, D., Rembold, M., and Leptin, M. (2017). From morphogen to morphogenesis and back. Nature 541, 311320.Google Scholar
Girton, J.R. (1981). Pattern triplications produced by a cell-lethal mutation in Drosophila. Dev. Biol. 84, 164172.Google Scholar
Girton, J.R. (1982). Genetically induced abnormalities in Drosophila: two or three patterns? Am. Zool. 22, 6577.Google Scholar
Girton, J.R. (1983). Morphological and somatic clonal analyses of pattern triplications. Dev. Biol. 99, 202209.Google Scholar
Girton, J.R. and Berns, M.W. (1982). Pattern abnormalities induced in Drosophila imaginal discs by an ultraviolet laser microbeam. Dev. Biol. 91, 7377.Google Scholar
Girton, J.R. and Bryant, P.J. (1980). The use of cell lethal mutations in the study of Drosophila development. Dev. Biol. 77, 233243.Google Scholar
Girton, J.R. and Kumor, A.L. (1985). The role of cell death in the induction of pattern abnormalities in a cell-lethal mutation of Drosophila. Dev. Genet. 5, 93102.Google Scholar
Girton, J.R. and Russell, M.A. (1980). A clonal analysis of pattern duplication in a temperature-sensitive cell-lethal mutant of Drosophila melanogaster. Dev. Biol. 77, 121.Google Scholar
Girton, J.R. and Russell, M.A. (1981). An analysis of compartmentalization in pattern duplications induced by a cell-lethal mutation in Drosophila. Dev. Biol. 85, 5564.Google Scholar
Gloor, H. (1947). Phänokopie-Versuche mit Äther an Drosophila. Rev. Suisse Zool. 54, 637712.Google Scholar
Gnatzy, W., Grünert, U., and Bender, M. (1987). Campaniform sensilla of Calliphora vicina (Insecta, Diptera). I. Typography. Zoomorphology 106, 312319.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), 4649.Google Scholar
Goldschmidt, R.B. (1952). Homoeotic mutants and evolution. Acta Biotheor. 10, 87104.Google Scholar
Goldstein, B. and Freeman, G. (1997). Axis specification in animal development. BioEssays 19, 105116.Google Scholar
Golovnin, A., Gause, M., Georgieva, S., Gracheva, E., and Georgiev, P. (1999). The su(Hw) insulator can disrupt enhancer–promoter interactions when located more than 20 kilobases away from the Drosophila achaete-scute complex. Mol. Cell. Biol. 19, 34433456.Google Scholar
Gómez, J.A., Ceacero, F., Landete-Castillejos, T., Gaspar-Lopez, E., García, A.J., and Gallego, L. (2012). Factors affecting antler investment in Iberian red deer. Anim. Prod. Sci. 52, 867873.Google Scholar
Gómez-Skarmeta, J.L., Campuzano, S., and Modolell, J. (2003). Half a century of neural prepatterning: the story of a few bristles and many genes. Nat. Rev. Neurosci. 4, 587598.Google 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, 18691882.Google Scholar
Gönczy, P. (2008). Mechanisms of asymmetric cell division: flies and worms pave the way. Nat. Rev. Mol. Cell Biol. 9, 355366.Google Scholar
González-Forero, M. and Gardner, A. (2018). Inference of ecological and social drivers of human brain-size evolution. Nature 557, 554557.Google Scholar
González-Méndez, L., Gradilla, A.-C., and Guerreiro, I. (2019). The cytoneme connection: direct long-distance signal transfer during development. Development 146, dev174607.Google Scholar
Goodman, B.A. and Johnson, P.T.J. (2011). Disease and the extended phenotype: parasites control host performance and survival through induced changes in body plan. PLoS ONE 6(5), e20193.Google Scholar
Goodrich, L.V. and Strutt, D. (2011). Principles of planar polarity in animal development. Development 138, 18771892.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., editor. University of Alberta Press, Edmonton, pp. 185200.Google Scholar
Gotoh, H., Hust, J.A., Miura, T., Niimi, T., Emlen, D.J., and Lavine, L.C. (2015). The Fat/Hippo signaling pathway links within-disc morphogen patterning to whole-animal signals during phenotypically plastic growth in insects. Dev. Dynamics 244, 10391045.Google Scholar
Gou, J., Stotsky, J.A., and Othmer, H.G. (2020). Growth control in the Drosophila wing disk. Wiley Interdiscip. Rev. Syst. Biol. Med. 2020, e1478.Google Scholar
Gould, G.M. and Pyle, W.L. (1896). Anomalies and Curiosities of Medicine. Julian Press, New York.Google Scholar
Gould, S.J. (1966). Allometry and size in ontogeny and phylogeny. Biol. Rev. 41, 587640.Google Scholar
Gould, S.J. (1971). D’Arcy Thompson and the science of form. New Lit. Hist. 2, 229258.Google Scholar
Gould, S.J. (1974). The origin and function of “bizarre” structures: antler size and skull size in the “Irish Elk,” Megaloceros giganteus. Evolution 28, 191220.Google 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), 2230.Google Scholar
Gould, S.J. (1980). The Panda’s Thumb: More Reflections in Natural History. W. W. Norton, New York.Google Scholar
Gould, S.J. (1981). What, if anything, is a zebra? Nat. Hist. 90(7), 612.Google Scholar
Gould, S.J. (1981). What color is a zebra? Nat. Hist. 90(8), 1622.Google Scholar
Gould, S.J. (1982). Living with connections: are Siamese twins one person or two? Nat. Hist. 91(11), 1822.Google Scholar
Gould, S.J. (1986). The egg-a-day barrier. Nat. Hist. 95(7), 1624.Google Scholar
Gould, S.J. (1990). Wonderful Life: The Burgess Shale and the Nature of History. Norton, New York.Google Scholar
Gould, S.J. (1994). Cabinet museums revisited. Nat. Hist. 103(1), 1220.Google Scholar
Govind, C.K. (1989). Asymmetry in lobster claws. Am. Sci. 77, 468474.Google Scholar
Graff, J.M. (1997). Embryonic patterning: to BMP or not to BMP, that is the question. Cell 89, 171174.Google Scholar
Grall, E. and Tschopp, P. (2019). A sense of place, many times over: pattern formation and evolution of repetitive morphological structures. Dev. Dynamics 249, 313327.Google Scholar
Grantham, M.E., Shingleton, A.W., Dudley, E., and Brisson, J.A. (2020). Expression profiling of winged‐ and wingless‐destined pea aphid embryos implicates insulin/insulin growth factor signaling in morph differences. Evol. Dev. 22, 257268.Google Scholar
Graván, C.P. and Lahoz-Beltra, R. (2004). Evolving morphogenetic fields in the zebra skin pattern based on Turing’s morphogen hypothesis. Int. J. Appl. Math. Comput. Sci. 14, 351361.Google Scholar
Gray, G.W. (1948). The great ravelled knot. Sci. Am. 179(10), 2639.Google Scholar
Green, H. and Thomas, J. (1978). Pattern formation by cultured human epidermal cells: development of curved ridges resembling dermatoglyphics. Science 200, 13851388.Google Scholar
Green, J.B.A. and Sharpe, J. (2015). Positional information and reaction–diffusion: two big ideas in developmental biology combine. Development 142, 12031211.Google Scholar
Green, M.C. (1961). Himalayan, a new allele of albino in the mouse. J. Hered. 52, 7375.Google Scholar
Greenberg, L. and Hatini, V. (2011). Systematic expression and loss-of-function analysis defines spatially restricted requirements for Drosophila RhoGEFs and RhoGAPs in leg morphogenesis. Mech. Dev. 128, 517.Google Scholar
Greenwald, I. (2012). Notch and the awesome power of genetics. Genetics 191, 655669.Google Scholar
Greiling, T.M.S. and Clark, J.I. (2012). New insights into the mechanism of lens development using zebra fish. Int. Rev. Cell Mol. Biol. 296, 161.Google Scholar
Grimaldi, D.A. (1987). Amber fossil Drosophilidae (Diptera), with particular reference to the Hispaniolan taxa. Am. Mus. Novitates 2880, 123.Google Scholar
Grimes, D.T. (2019). Making and breaking symmetry in development, growth and disease. Development 146, dev170985.Google Scholar
Grimes, D.T. and Burdine, R.D. (2017). Left–right patterning: breaking symmetry to asymmetric morphogenesis. Trends Genet. 33, 616628.Google Scholar
Grobet, L., Martin, L.J.R., Poncelet, D., Pirottin, D., Brouwers, B., Riquet, J., Schoeberlein, A., Dunner, S., Ménissier, F., Massabanda, J., Fries, R., Hanset, R., and Georges, M. (1997). A deletion in the bovine myostatin gene causes the double-muscled phenotype in cattle. Nat. Genet. 17, 7174.Google Scholar
Grochowska, E., Borys, B., Lisiak, D., and Mroczkowski, S. (2019). Genotypic and allelic effects of the myostatin gene (MSTN) on carcass, meat quality, and biometric traits in Colored Polish Merino sheep. Meat Sci. 151, 417.Google Scholar
Gross, J.B., Kerney, R., Hanken, J., and Tabin, C.J. (2011). Molecular anatomy of the developing limb in the coquí frog, Eleutherodactylus coqui. Evol. Dev. 13, 415426.Google Scholar
Groves, A.K. and Fekete, D.M. (2012). Shaping sound in space: the regulation of inner ear patterning. Development 139, 245257.Google Scholar
Gu, L., Mo, E., Yang, Z., Zhu, X., Fang, Z., Sun, B., Wang, C., Bao, J., and Sung, C. (2007). Expression and localization of insulin-like growth factor-I in four parts of the red deer antler. Growth Factors 25, 264279.Google Scholar
Gubb, D. and García-Bellido, A. (1982). A genetic analysis of the determination of cuticular polarity during development in Drosophila melanogaster. J. Embryol. Exp. Morphol. 68, 3757.Google Scholar
Guerreiro, I. and Duboule, D. (2014). Snakes: hatching a model system for Evo-Devo? Int. J. Dev. Biol. 58, 727732.Google Scholar
Guinard, G. (2015). Introduction to evolutionary teratology, with an application to the forelimbs of Tyrannosauridae and Carnotaurinae (Dinosauria: Theropoda). Evol. Biol. 42, 2041.Google Scholar
Gumbiner, B.M. and Kim, N.-G. (2014). The Hippo–YAP signaling pathway and contact inhibition of growth. J. Cell Sci. 127, 709717.Google Scholar
Gunhaga, L. (2011). The lens: a classical model of embryonic induction providing new insights into cell determination in early development. Philos. Trans. R. Soc. Lond. B 366, 11931203.Google Scholar
Haas, B.J. and Whited, J.L. (2017). Advances in decoding axolotl limb regeneration. Trends Genet. 33, 553565.Google Scholar
Hadorn, E. (1961). Developmental Genetics and Lethal Factors. Methuen, London [translated from 1955 German original, Thieme Verlag, Stuttgart, by U. Mittwoch].Google Scholar
Hadorn, E. (1978). Transdetermination. In The Genetics and Biology of Drosophila, Ashburner, M. and Wright, T.R.F., editors. Academic Press, New York, pp. 555617.Google Scholar
Halder, G. and Johnson, R.L. (2011). Hippo signaling: growth control and beyond. Development 138, 922.Google Scholar
Hall, B.K. (2008). EvoDevo concepts in the work of Waddington. Biol. Theory 3, 198203.Google Scholar
Hall, B.K. (2018). Germ layers, the neural crest and emergent organization in development and evolution. Genesis 56, e23103.Google Scholar
Hall, J.C., Gelbart, W.M., and Kankel, D.R. (1976). Mosaic systems. In The Genetics and Biology of Drosophila, Ashburner, M. and Novitski, E., editors. Academic Press, New York, pp. 265314.Google Scholar
Hallgrímsson, B., Green, R.M., Katz, D.C., Fish, J.L., Bernier, F.P., Roseman, C.C., Young, N.M., Cheverud, J.M., and Marcucio, R.S. (2019). The developmental-genetics of canalization. Semin. Cell Dev. Biol. 88, 6779.Google Scholar
Hallgrímsson, B., Jamniczky, H., Young, N.M., Rolian, C., Parsons, T.E., Boughner, J.C., and Marcucio, R.S. (2009). Deciphering the palimpsest: studying the relationship between morphological integration and phenotypic covariation. Evol. Biol. 36, 355376.Google Scholar
Hamant, O. (2017). Mechano-devo. Mech. Dev. 145, 29.Google Scholar
Hamburger, V. (1988). The Heritage of Experimental Embryology: Hans Spemann and the Organizer. Oxford University Press, New York.Google Scholar
Hamburger, V. (2001). Induction of embryonic primordia by implantation of organizers from a different species. [English translation of 1924 German paper by Hans Spemann and Hilde Mangold.] Int. J. Dev. Biol. 45, 1338.Google Scholar
Hamelin, A., Conchou, F., Fusellier, M., Duchenij, B., Vieira, I., Filhol, E., de Citres, C.D., Tiret, L., Gache, V., and Abitbol, M. (2020). Genetic heterogeneity of polydactyly in Maine Coon cats. J. Feline Med. Surg. 1098612X20905061 [published online, 18 Feb 2020].Google 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, 125.Google Scholar
Hannah-Alava, A. (1960). Genetic mosaics. Sci. Am. 202(5), 118130.Google Scholar
Hannezo, E. and Heisenberg, C.-P. (2019). Mechanochemical feedback loops in development and disease. Cell 178, 1225.Google Scholar
Hanset, R. and Michaux, C. (1985). On the genetic determinism of muscular hypertrophy in the Belgian White and Blue cattle breed. II. Population data. Génét. Sél. Evol. 17, 369386.Google Scholar
Happle, R. (2015). The categories of cutaneous mosaicism: a proposed classification. Am. J. Med. Genet. A 170A, 452459.Google Scholar
Hardie, R.C. (1985). Functional organization of the fly retina. In Progress in Sensory Physiology, Ottoson, D., editor. Springer-Verlag, Berlin, pp. 179.Google Scholar
Harfe, B.D., Scherz, P.J., Nissim, S., Tian, H., McMahon, A.P., and Tabin, C.J. (2004). Evidence for an expansion-based temporal Shh gradient in specifying vertebrate digit identities. Cell 118, 517528.Google Scholar
Hariharan, I.K. (2015). Organ size control: lessons from Drosophila. Dev. Cell 34, 255265.Google Scholar
Hariharan, I.K. and Bilder, D. (2006). Regulation of imaginal disc growth by tumor-suppressor genes in Drosophila. Annu. Rev. Genet. 40, 335361.Google Scholar
Hariharan, I.K. and Serras, F. (2017). Imaginal disc regeneration takes flight. Curr. Opin. Cell Biol. 48, 1016.Google Scholar
Harris, M.L., Chora, L., Bishop, C.A., and Bogart, J.P. (2000). Species- and age-related differences in susceptibility to pesticide exposure for two amphibians, Rana pipiens and Bufo americanus. Bull. Environ. Contam. Toxicol. 64, 263270.Google Scholar
Harris, R.E., Setiawan, L., Saul, J., and Hariharan, I.K. (2016). Localized epigenetic silencing of a damage-activated WNT enhancer limits regeneration in mature Drosophila imaginal discs. eLife 5, e11588.Google Scholar
Harrison, R.G. (1918). Experiments on the development of the fore limb of Amblystoma, a self-differentiating equipotential system. J. Exp. Zool. 25, 413461.Google Scholar
Harrison, R.G. (1921). On relations of symmetry in transplanted limbs. J. Exp. Zool. 32, 1136.Google Scholar
Hartfelder, K., Guidugli-Lazzarini, K.R., Cervoni, M.S., Santos, D.E., and Humann, F.C. (2015). Old threads make new tapestry: rewiring of signalling pathways underlies caste phenotypic plasticity in the honey bee, Apis mellifera L. Adv. Insect Physiol. 48, 136.Google Scholar
Hartwell, L.H. (1967). Macromolecule synthesis in temperature-sensitive mutants of yeast. J. Bacteriol. 93, 16621670.Google Scholar
Harvey, P.H. and Krebs, J.R. (1990). Comparing brains. Science 249, 140146.Google Scholar
Harzsch, S., Benton, J., and Beltz, B.S. (2000). An unusual case of a mutant lobster embryo with double brain and double ventral nerve cord. Arthropod Struct. Dev. 29, 9599.Google 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. Mech. Dev. 111, 1724.Google Scholar
Hassan, B.A. and Hiesinger, P.R. (2015). Beyond molecular codes: simple rules to wire complex brains. Cell 163, 285291.Google Scholar
Hassanpour, M. and Joss, J. (2009). Anatomy and histology of the spiral valve intestine in juvenile Australian lungfish, Neoceratodus forsteri. Open Zool. J. 2, 6285.Google Scholar
Hatchwell, E. and Dennis, N. (1996). Mirror hands and feet: a further case of Laurin–Sandrow syndrome. J. Med. Genet. 33, 426428.Google Scholar
Hattori, A., Sugime, Y., Sasa, C., Miyakaya, H., Ishikawa, Y., Miyazaki, S., Okada, Y., Cornette, R., Lavine, L.C., Emlen, D.J., Koshikawa, S., and Miura, T. (2013). Soldier morphogenesis in the damp-wood termite is regulated by the insulin signaling pathway. J. Exp. Zool. B Mol. Dev. Evol. 320, 295306.Google Scholar
Haupaix, N. and Manceau, M. (2020). The embryonic origin of periodic color patterns. Dev. Biol. 460, 7076.Google Scholar
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.Google Scholar
Haynie, J.L. and Bryant, P.J. (1986). Development of the eye-antenna imaginal disc and morphogenesis of the adult head in Drosophila melanogaster. J. Exp. Zool. 237, 293308.Google Scholar
He, X.J., Zhou, L.B., Pan, Q.Z., Barron, A.B., Yan, W.Y., and Zeng, Z.J. (2017). Making a queen: an epigenetic analysis of the robustness of the honeybee (Apis mellifera) queen developmental pathway. Mol. Ecol. 26, 15981607.Google Scholar
Heer, N.C. and Martin, A.C. (2017). Tension, contraction and tissue morphogenesis. Development 144, 42494260.Google Scholar
Heimeier, R.A., Das, B., Buchholz, D.R., Fiorentino, M., and Shi, Y.-B. (2010). Studies on Xenopus laevis intestine reveal biological pathways underlying vertebrate gut adaptation from embryo to adult. Genome Biol. 11, R55.Google Scholar
Heingard, M., Turetzek, N., Prpic, N.-M., and Janssen, R. (2019). FoxB, a new and highly conserved key factor in arthropod dorsal–ventral (DV) limb patterning. EvoDevo 10, 28.Google Scholar
Heintzman, N.D. and Ren, B. (2009). Finding distal regulatory elements in the human genome. Curr. Opin. Genet. Dev. 19, 541549.Google Scholar
Heisenberg, C.-P. and Bellaïche, Y. (2013). Forces in tissue morphogenesis and patterning. Cell 153, 948962.Google Scholar
Hejtmancik, J.F., Riazuddin, S.A., McGreal, R., Liu, W., Cvekl, A., and Shiels, A. (2015). Lens biology and biochemistry. Prog. Mol. Biol. Transl. Sci. 134, 169201.Google Scholar
Held, L.I. Jr. (1977). Analysis of bristle-pattern formation in Drosophila. PhD thesis, Department of Molecular Biology, University of California, Berkeley, CA.Google Scholar
Held, L.I. Jr. (1979). Pattern as a function of cell number and cell size on the second-leg basitarsus of Drosophila. W. Roux Arch. Dev. Biol. 187, 105127.Google Scholar
Held, L.I. Jr. (1979). A high-resolution morphogenetic map of the second-leg basitarsus in Drosophila melanogaster. W. Roux Arch. Dev. Biol. 187, 129150.Google Scholar
Held, L.I. Jr. (1990). Sensitive periods for abnormal patterning on a leg segment in Drosophila melanogaster. W. Roux Arch. Dev. Biol. 199, 3147.Google Scholar
Held, L.I. Jr. (1990). Arrangement of bristles as a function of bristle number on a leg segment in Drosophila melanogaster. W. Roux Arch. Dev. Biol. 199, 4862.Google Scholar
Held, L.I. Jr. (1991). Bristle patterning in Drosophila. BioEssays 13, 633640.Google Scholar
Held, L.I. Jr. (1992). Models for Embryonic Periodicity. Monographs in Developmental Biology, Vol. 24. Karger, Basel.Google Scholar
Held, L.I. Jr. (1995). Axes, boundaries and coordinates: the ABCs of fly leg development. BioEssays 17, 721732.Google Scholar
Held, L.I. Jr. (2002). Imaginal Discs: The Genetic and Cellular Logic of Pattern Formation. Cambridge University Press, New York.Google Scholar
Held, L.I. Jr. (2002). Bristles induce bracts via the EGFR pathway on Drosophila legs. Mech. Dev. 117, 225234.Google Scholar
Held, L.I. Jr. (2009). Quirks of Human Anatomy: An Evo-Devo Look at the Human Body. Cambridge University Press, New York.Google Scholar
Held, L.I. Jr. (2010). The evo-devo puzzle of human hair patterning. Evol. Biol. 37, 113122.Google Scholar
Held, L.I. Jr. (2010). How does Scr cause first legs to deviate from second legs? Dros. Info. Serv. 93, 132146.Google Scholar
Held, L.I. Jr. (2014). How the Snake Lost Its Legs: Curious Tales from the Frontier of Evo-Devo. Cambridge University Press, New York.Google Scholar
Held, L.I. Jr. (2017). Deep Homology? Uncanny Similarities of Humans and Flies. Cambridge University Press, New York.Google Scholar
Held, L.I. Jr. and Bryant, P.J. (1984). Cell interactions controlling the formation of bristle patterns in Drosophila. In Pattern Formation: A Primer in Developmental Biology, Malacinski, G.M. and Bryant, S.V., editors. Macmillan, New York, pp. 291322.Google Scholar
Held, L.I. Jr. and Heup, M. (1996). Genetic mosaic analysis of decapentaplegic and wingless gene function in the Drosophila leg. Dev. Genes Evol. 206, 180194.Google Scholar
Held, L.I. Jr. and Sessions, S.K. (2019). Reflections on Bateson’s rule: solving an old riddle about why extra legs are mirror-symmetric. J. Exp. Zool. B Mol. Dev. Evol. 332, 219237.Google Scholar
Held, L.I. Jr., Davis, A.L., and Aybar, R.S. (2017). Instigating an “identity crisis” to investigate how a Hox gene acts on fly legs. Dros. Info. Serv. 100, 7589.Google Scholar
Held, L.I. Jr., Duarte, C.M., and Derakhshanian, K. (1986). Extra tarsal joints and abnormal cuticular polarities in various mutants of Drosophila melanogaster. W. Roux Arch. Dev. Biol. 195, 145157.Google Scholar
Held, L.I. Jr., Grimson, M.J., and Du, Z. (2004). Proving an old prediction: the sex comb rotates at 16 to 24 hours after pupariation. Dros. Info. Serv. 87, 7678.Google Scholar
Held, L.I. Jr., McNeme, S.C., and Hernandez, D. (2018). Induction of ectopic transverse rows by Ubx on fly legs. Dros. Info. Serv. 101, 2532.Google Scholar
Heller, E. and Fuchs, E. (2015). Tissue patterning and cellular mechanics. J. Cell Biol. 211, 219231.Google Scholar
Henderson, D.J., Long, D.A., and Dean, C.H. (2018). Planar cell polarity in organ formation. Curr. Opin. Cell Biol. 55, 96103.Google Scholar
Hendrikse, J.L., Parsons, T.E., and Hallgrímsson, B. (2007). Evolvability as the proper focus of evolutionary developmental biology. Evol. Dev. 9, 393401.Google Scholar
Henke, K. and Maas, H. (1946). Über sensible Perioden der allgemeinen Körpergliederung von Drosophila. Nachr. Akad. Wiss. Göttingen. Math.-Phys. Kl. IIb, Biol.-Physiol.-Chem. Abt. 1, 34.Google Scholar
Henneberg, M., Lambert, K.M., and Leigh, C.M. (1998). Fingerprinting a chimpanzee and a koala: animal dermatoglyphics can resemble human ones. In Conference of the Australian and New Zealand International Symposium on the Forensic Sciences, 1996. Sydney.Google Scholar
Henrique, D. and Schweisguth, F. (2019). Mechanisms of Notch signaling: a simple logic deployed in time and space. Development 146, dev172148.Google Scholar
Henry, J.J. and Hamilton, P.W. (2018). Diverse evolutionary origins and mechanisms of lens regeneration. Mol. Biol. Evol. 35, 15631575.Google Scholar
Herndon, J.G., Tigges, J., Anderson, D.C., Klumpp, S.A., and McClure, H.M. (1999). Brain weight throughout the life span of the chimpanzee. J. Comp. Neurol. 409, 567572.Google Scholar
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, 717727.Google Scholar
Hershkovitz, P. (1977). Living New World Monkeys (Platyrrhini) With an Introduction to Primates. Vol. 1. University of Chicago Press, Chicago, IL.Google Scholar
Heyning, J.E. and Lento, G.M. (2002). The evolution of marine mammals. In Marine Mammal Biology: An Evolutionary Approach, Hoelzel, A.R., editor. Blackwell Science, Malden, MA, pp. 3872.Google Scholar
Hill, E.W., McGivney, B.A., Rooney, M.F., Katz, L.M., Parnell, A., and MacHugh, D.E. (2019). The contribution of myostatin (MSTN) and additional modifying genetic loci to race distance aptitude in Thoroughbred horses racing in different geographic regions. Equine Vet. J. 51, 625633.Google Scholar
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, 737743.Google Scholar
Hintz, M., Bartholmes, C., Nutt, P., Ziermann, J., Hameister, S., Neuffer, B., and Theissen, G. (2007). Catching a “hopeful monster”: shepherd’s purse (Capsella bursa-pastoris) as a model system to study the evolution of flower development. J. Exp. Botany 57, 35313542.Google Scholar
Hiscock, T.W. and Megason, S.G. (2015). Mathematically guided approaches to distinguish models of periodic patterning. Development 142, 409419.Google Scholar
Hiscock, T.W. and Megason, S.G. (2015). Orientation of Turing-like patterns by morphogen gradients and tissue anisotropies. Cell Syst. 1, 408416.Google Scholar
Hiscock, T.W., Tschopp, P., and Tabin, C.J. (2017). On the formation of digits and joints during limb development. Dev. Cell 41, 459465.Google Scholar
Hjorth, M., Pourteymour, S., Görgens, S.W., Langleite, T.M., Lee, S., Holen, T., Gulseth, H.L., Birkeland, K.I., Jensen, J., Drevon, C.A., and Norheim, F. (2016). Myostatin in relation to physical activity and dysglycaemia and its effect on energy metabolism in human skeletal muscle cells. Acta Physiol. 217, 4560.Google Scholar
Ho, E.C.Y., Malagón, J.N., Ahuja, A., Singh, R., and Larsen, E. (2018). Rotation of sex combs in Drosophila melanogaster requires precise and coordinated spatio-temporal dynamics from forces generated by epithelial cells. PLoS Comput. Biol. 14(10), e1006455.Google Scholar
Ho, M.-W., Bolton, E., and Saunders, P.T. (1983). Bithorax phenocopy and pattern formation. I. Spatiotemporal characteristics of the phenocopy response. Exp. Cell Biol. 51, 282290.Google Scholar
Ho, M.-W., Saunders, P.T., and Bolton, E. (1983). Bithorax phenocopy and pattern formation. II. A model of prepattern formation. Exp. Cell Biol. 51, 291299.Google Scholar
Hodges, A. (1983). Alan Turing: The Enigma. Simon & Schuster, New York.Google Scholar
Hoekstra, H.E. (2006). Genetics, development and evolution of adaptive pigmentation in vertebrates. Heredity 97, 222234.Google Scholar
Hofer, H., Carroll, J., Neitz, J., Neitz, M., and Williams, D.R. (2005). Organization of the human trichromatic cone mosaic. J. Neurosci. 25, 96699679.Google Scholar
Hoffman, B.D. and Yap, A.S. (2015). Towards a dynamic understanding of cadherin-based mechanobiology. Trends Cell Biol. 25, 803814.Google Scholar
Hoffmann, M.B. and Dumoulin, S.O. (2015). Congenital visual pathway abnormalities: a window onto cortical stability and plasticity. Trends Neurosci. 38, 5565.Google Scholar
Holland, N.D. (2003). Early central nervous system evolution: an era of skin brains? Nat. Rev. Neurosci. 4, 111.Google Scholar
Holland, P.W.H., Marlétaz, F., Maeso, I., Dunwell, T.L., and Paps, J. (2016). New genes from old: asymmetric divergence of gene duplicates and the evolution of development. Philos. Trans. R. Soc. Lond. B 372, 20150480.Google Scholar
Holloway, R.L. (2001). Does allometry mask important brain structure residuals relevant to species-specific behavioral evolution? Behav. Brain Sci. 24, 286287.Google Scholar
Holtfreter, J. and Hamburger, V. (1955). Amphibians. In Analysis of Development, Willier, B.H., Weiss, P.A., and Hamburger, V., editors. Hafner, New York, pp. 230296.Google Scholar
Holtzer, H. (1978). Cell lineages, stem cells and the “quantal” cell cycle concept. In Stem Cells and Tissue Homeostasis, Lord, B.I., Potten, C.S., and Cole, R.J., editors. Cambridge University Press, Cambridge, pp. 127.Google Scholar
Hombria, J.C.-G. and Sotillos, S. (2020). Evo-devo: when four became two plus two. Curr. Biol. 30, R655R657.Google Scholar
Honeycutt, R.L. (2010). Unraveling the mysteries of dog evolution. BMC Biol. 8, 20.Google Scholar
Hoopes, B.C., Rimbault, M., Liebers, D., Ostrander, E.A., and Sutter, N.B. (2012). The insulin-like growth factor 1 receptor (IGF1R) contributes to reduced size in dogs. Mamm. Genome 23, 780790.Google Scholar
Horton, J.C. and Adams, D.L. (2005). The cortical column: a structure without a function. Philos. Trans. R. Soc. Lond. B 360, 837862.Google Scholar
Hosseini, H.S. and Taber, L.A. (2018). How mechanical forces shape the developing eye. Prog. Biophys. Mol. Biol. 137, 2536.Google Scholar
Hotta, Y. and Benzer, S. (1970). Genetic dissection of the Drosophila nervous system by means of mosaics. PNAS 67, 11561163.Google Scholar
Hotta, Y. and Benzer, S. (1972). Mapping of behaviour in Drosophila mosaics. Nature 240, 527535.Google Scholar
Houle, D., Jones, L.T., Fortune, R., and Sztepanacz, J.L. (2019). Why does allometry evolve so slowly? Integr. Comp. Biol. 59, 14291440.Google Scholar
Houston, D.W. (2012). Cortical rotation and messenger RNA localization in Xenopus axis formation. Wiley Interdiscip. Rev. Dev. Biol. 1, 371388.Google Scholar
Huang, T., Zhang, M., Yan, G., Huang, X., Chen, H., Zhou, L., Deng, W., Zhang, Z., Qiu, H., Ai, H., and Huang, L. (2019). Genome-wide association and evolutionary analyses reveal the formation of swine facial wrinkles in Chinese Erhualian pigs. Aging 11, 46724687.Google Scholar
Huelsken, J., Vogel, R., Erdmann, B., Cotsarelis, G., and Birchmeier, W. (2001). β-Catenin controls hair follicle morphogenesis and stem cell differentiation in the skin. Cell 105, 533545.Google Scholar
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, 388407.Google Scholar
Hunter, G.L., Hadjivasiliou, Z., Bonin, H., He, L., Perrimon, N., Charras, G., and Baum, B. (2016). Coordinated control of Notch/Delta signalling and cell cycle progression drives lateral inhibition-mediated tissue patterning. Development 143, 23052310.Google Scholar
Hunter, M.V. and Fernandez-Gonzalez, R. (2017). Coordinating cell movements in vivo: junctional and cytoskeletal dynamics lead the way. Curr. Opin. Cell Biol. 48, 5462.Google Scholar
Huxley, J.S. (1932). Problems of Relative Growth, 2nd ed. Methuen, London.Google Scholar
Huxley, J.S. and Teissier, G. (1936). Terminology of relative growth. Nature 137, 780781.Google Scholar
Ichikawa, M. and Bui, K.H. (2018). Microtubule inner proteins: a meshwork of luminal proteins stabilizing the doublet microtubule. BioEssays 40, 1700209.Google Scholar
Iljin, N.A. and Iljin, V.N. (1930). Temperature effects on the color of the Siamese cat. J. Hered. 21, 309318.Google Scholar
Im, K. and Grant, P.E. (2019). Sulcal pits and patterns in developing human brains. NeuroImage 185, 881890.Google Scholar
Im, K., Pienaar, R., Lee, J.-M., Seong, J.-K., Choi, Y.Y., Lee, K.H., and Grant, P.E. (2011). Quantitative comparison and analysis of sulcal patterns using sulcal graph matching: a twin study. NeuroImage 57, 10771086.Google Scholar
Imes, D.L., Geary, L.A., Grahn, R.A., and Lyons, L.A. (2006). Albinism in the domestic cat (Felis catus) is associated with a tyrosinase (TYR) mutation. Anim. Genet. 37, 175178.Google Scholar
Ingham, P.W. and McMahon, A.P. (2001). Hedgehog signaling in animal development: paradigms and principles. Genes Dev. 15, 30593087.Google Scholar
Irimia, M., Maeso, I., Roy, S.W., and Fraser, H.B. (2013). Ancient cis-regulatory constraints and the evolution of genome architecture. Trends Genet. 29, 521528.Google Scholar
Irvine, K.D. (1999). Fringe, Notch, and making developmental boundaries. Curr. Opin. Genet. Dev. 9, 434441.Google Scholar
Irvine, K.D. and Rauskolb, C. (2001). Boundaries in development: formation and function. Annu. Rev. Cell Dev. Biol. 17, 189214.Google Scholar
Irvine, K.D. and Shraiman, B.I. (2017). Mechanical control of growth: ideas, facts and challenges. Development 144, 42384248.Google Scholar
Ishibashi, S., Saldanha, F.Y.L., and Amaya, E. (2017). Xenopus as a model organism for biomedical research. In Basic Science Methods for Clinical Researchers, Jalali, M., Saldanha, F., and Jalali, M., editors. Academic Press, New York, pp. 263290.Google Scholar
Iten, L.E. and Bryant, S.V. (1975). The interaction between the blastema and stump in the establishment of the anterior–posterior and proximal–distal organization of the limb regenerate. Dev. Biol. 44, 119147.Google Scholar
Ito, M., Yang, Z., Andl, T., Cui, C., Kim, N., Millar, S.E., and Cotsarelis, G. (2007). Wnt-dependent de novo hair follicle regeneration in adult mouse skin after wounding. Nature 447, 316320.Google Scholar
Jablonski, N.G. (2006). Skin: A Natural History. University of California Press, Berkeley, CA.Google Scholar
Jablonski, N.G. (2010). The naked truth. Sci. Am. 302(2), 4249.Google Scholar
Jablonski, N.G. and Chaplain, G. (2017). The colours of humanity: the evolution of pigmentation in the human lineage. Proc. R. Soc. B 372, 20160349.Google Scholar
Jackson, I.J. (2013). How the leopard gets its spots: a transmembrane peptidase specifies feline pigmentation patterns. Pigment Cell Melanoma Res. 26, 438439.Google Scholar
Jacobs, M.D. (2009). Multiscale systems integration in the eye. Wiley Interdiscip. Rev. Syst. Biol. Med. 1, 1527.Google Scholar
Jaeger, J. and Verd, B. (2020). Dynamic positional information: patterning mechanism versus precision in gradient-driven systems. Curr. Top. Dev. Biol. 137, 219246.Google Scholar
Jahner, J.P., Lucas, L.K., Wilson, J.S., and Forister, M.L. (2015). Morphological outcomes of gynandromorphism in Lycaeides butterflies (Lepidoptera: Lycaenidae). J. Insect Sci. 15, 38.Google Scholar
Jahoda, C.A.B. (1998). Cellular and developmental aspects of androgenetic alopecia. Exp. Dermatol. 7, 235248.Google Scholar
Jansen, A.G., Mous, S.E., White, T., Posthuma, D., and Polderman, T.J.C. (2015). What twin studies tell us about the heritability of brain development, morphology, and function: a review. Neuropsychol. Rev. 25, 2746.Google Scholar
Janzen, F.J. and Paukstis, G.L. (1991). Environmental sex determination in reptiles: ecology, evolution, and experimental design. Q. Rev. Biol. 66, 149179.Google Scholar
Jaraud, A., Bossé, P., de Citres, C.D., Tiret, L., Gache, V., and Abitbol, M. (2020). Feline chimerism revealed by DNA profiling. Anim. Genet. 51, 631633.Google Scholar
Jarvik, J. and Botstein, D. (1973). A genetic method for determining the order of events in a biological pathway. PNAS 70, 20462050.Google Scholar
Jattiot, R., Fara, E., Brayard, A., Urdy, S., and Goudemand, N. (2019). Learning from beautiful monsters: phylogenetic and morphogenetic implications of left–right asymmetry in ammonoid shells. BMC Evol. Biol. 19, 210.Google Scholar
Jeong, D., Li, Y., Choi, Y., Yoo, M., Kang, D., Park, J., Choi, J., and Kim, J. (2017). Numerical simulation of the zebra pattern formation on a three-dimensional model. Physica A 475, 106116.Google Scholar
Jernvall, J. and Salazar-Ciudad, I. (2007). The economy of tinkering mammalian teeth. In Tinkering: The Microevolution of Development, Bock, G. and Goode, J., editors. Wiley, Chichester, pp. 207224.Google Scholar
Ji, S., Liu, Q., Zhang, S., Chen, Q., Wang, C., Zhang, W., Xiao, C., Li, Y., Nian, C., Li, J., Li, J., Geng, J., Hong, L., Xie, C., He, Y., Chen, X., Li, X., Yin, Z.-Y., You, H., Lin, K.-H., Wu, Q., Yu, C., Johnson, R.L., Wang, L., Chen, L., Wang, F., and Zhou, D. (2018). FGF15 activates Hippo signaling to suppress bile acid metabolism and liver tumorigenesis. Dev. Cell 48, 460474.Google Scholar
Jiang, J. and Struhl, G. (1996). Complementary and mutually exclusive activities of Decapentaplegic and Wingless organize axial patterning during Drosophila leg development. Cell 86, 401409.Google Scholar
Jidigam, V.K., Srinivasan, R.C., Patthey, C., and Gunhaga, L. (2015). Apical constriction and epithelial invagination are regulated by BMP activity. Biol. Open 4, 17821791.Google Scholar
Jin, H., Fisher, M., and Grainger, R.M. (2012). Defining progressive stages in the commitment process leading to embryonic lens formation. Genesis 50, 728740.Google Scholar
Johnson, M.R., Barsh, G.S., and Mallarino, R. (2018). Periodic patterns in Rodentia: development and evolution. Exp. Dermatol. 28, 509513.Google Scholar
Johnson, N.A.N., Wang, Y., Zeng, Z., Wang, G.-D., Yao, Q., and Chen, K.-P. (2019). Phylogenetic analysis and classification of insect achaete-scute complex genes. J. Asia-Pacific Entomol. 22, 398403.Google Scholar
Johnson, P.T.J. and Sutherland, D.R. (2003). Amphibian deformities and Ribeiroia infection: an emerging helminthiasis. Trends Parasitol. 19, 332335.Google Scholar
Johnson, P.T.J., Lunde, K.B., Ritchie, E.G., and Launer, A.E. (1999). The effect of trematode infection on amphibian limb development and survivorship. Science 284, 802804.Google Scholar
Johnson, P.T.J., Lunde, K.B., Zelmer, D.A., and Werner, J.K. (2003). Limb deformities as an emerging parasitic disease in amphibians: evidence from museum specimens and resurvey data. Conserv. Biol. 17, 17241737.Google Scholar
Johnson, P.T.J., Preu, E.R., Sutherland, D.R., Romansic, J.M., Han, B., and Blaustein, A.R. (2006). Adding infection to injury: synergistic effects of predation and parasitism on amphibian malformations. Ecology 87, 22272235.Google Scholar
Johnston, M. (2020). Model organisms: nature’s gift to disease research. Genetics 214, 233234.Google Scholar
Johnston, R.J. Jr. and Desplan, C. (2010). Stochastic mechanisms of cell fate specification that yield random or robust outcomes. Annu. Rev. Cell Dev. Biol. 26, 689719.Google Scholar
Jordan, W. III, Rieder, L.E., and Larschan, E. (2019). Diverse genome topologies characterize dosage compensation across species. Trends Genet. 35, 308315.Google 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, 6476.Google Scholar
Joshi, S.D. and Davidson, L.A. (2012). Epithelial machines of morphogenesis and their potential application in organ assembly and tissue engineering. Biomech. Model. Mechanobiol. 11, 11091121.Google Scholar
Jülicher, F. and Eaton, S. (2017). Emergence of tissue shape changes from collective cell behaviours. Semin. Cell Dev. Biol. 67, 103112.Google Scholar
Jürgens, H., Peitgen, H.-O., and Saupe, D. (1990). The language of fractals. Sci. Am. 263(2), 6067.Google Scholar
Kaas, J.H. (2000). Organizing principles of sensory representations. In Evolutionary Developmental Biology of the Cerebral Cortex, Bock, G.R. and Cardew, G., editors. J. Wiley & Sons, New York, pp. 188205.Google Scholar
Kaas, J.H. (2005). Serendipity and the Siamese cat: the discovery that genes for coat and eye pigment affect the brain. ILAR J. 46, 357363.Google Scholar
Kaelin, C.B. and Barsh, G. (2013). Genetics of pigmentation in dogs and cats. Annu. Rev. Anim. Biosci. 1, 125156.Google Scholar
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). Specifying and sustaining pigmentation patterns in domestic and wild cats. Science 337, 15361541.Google Scholar
Kageura, H. (1997). Activation of dorsal development by contact between the cortical dorsal determinant and the equatorial core cytoplasm in eggs of Xenopus laevis. Development 124, 15431551.Google Scholar
Kalay, G., Lachowiec, J., Rosas, U., Dome, M.R., and Wittkopp, P.J. (2019). Redundant and cryptic enhancer activities of the Drosophila yellow gene. Genetics 212, 343360.Google Scholar
Kalcheim, C. (2016). Epithelial–mesenchymal transitions during neural crest and somite development. J. Clinical Med. 5, 5010001.Google Scholar
Kalcheim, C. (2018). Neural crest emigration: from start to stop. Genesis 56, e23090.Google Scholar
Kandachar, V. and Roegiers, F. (2012). Endocytosis and control of Notch signaling. Curr. Opin. Cell Biol. 24, 534540.Google Scholar
Kaneshiro, K.Y. (1988). Speciation in the Hawaiian Drosophila: sexual selection appears to play an important role. BioScience 38, 258263.Google Scholar
Kango-Singh, M. and Singh, A. (2009). Regulation of organ size: insights from the Drosophila Hippo signaling pathway. Dev. Dynamics 238, 16271637.Google Scholar
Kango-Singh, M., Nolo, R., Tao, C., Verstreken, P., Hiesinger, P.R., Bellen, H.J., and Halder, G. (2002). Shar-pei mediates cell proliferation arrest during imaginal disc growth in Drosophila. Development 129, 57195730.Google Scholar
Kapanidou, M., Lee, S., and Bolanos-Garcia, V.M. (2015). BubR1 kinase: protection against aneuploidy and premature aging. Trends Mol. Med. 21, 364372.Google Scholar
Kapoor, P. and Gonsalves, W.I. (2020). Of lions, shar-pei, and doughnuts: a tale retold. Blood 135, 10741076.Google Scholar
Karlsson, E.K., Baranowska, I., Wade, C.M., Salmon Hillbertz, N.H.C., Zody, M.C., Anderson, N., Biagi, T.M., Patterson, N., Pielberg, G.R., Kulbokas, E.J. III, Comstock, K.E., Keller, E.T., Mesirov, J.P., von Euler, H., Kämpe, O., Hedhammar, A., Lander, E.S., Andersson, G., Andersson, L., and Lindblad-Toh, K. (2007). Efficient mapping of mendelian traits in dogs through genome-wide association. Nat. Genet. 39, 13211328.Google Scholar
Karzbrun, E., Kshirsagar, A., Cohen, S.R., Hanna, J.H., and Reiner, O. (2018). Human brain organoids on a chip reveal the physics of folding. Nat. Phys. 14, 515522.Google Scholar
Katanaev, V.L., Egger-Adam, D., and Tomlinson, A. (2018). Antagonistic PCP signaling pathways in the developing Drosophila eye. Sci. Rep. 8, 5741.Google Scholar
Katz, L.C. and Crowley, J.C. (2002). Development of cortical circuits: lessons from ocular dominance columns. Nat. Rev. Neurosci. 3, 3442.Google Scholar
Kauffman, S.A. (1983). Developmental constraints: internal factors in evolution. In Development and Evolution, Goodwin, B.C., Holder, N., and Wylie, C.C., editors. Cambridge University Press, Cambridge, pp. 195225.Google Scholar
Kaufman, L. (1925). An experimental study on the partial albinism in Himalayan rabbits. Biol. Generalis Vienna 1, 721.Google Scholar
Kaufman, M.H. (2004). The embryology of conjoined twins. Childs Nerv. Syst. 20, 508525.Google Scholar
Kaufman, T.C., Seeger, M.A., and Olsen, G. (1990). Molecular and genetic organization of the Antennapedia gene complex of Drosophila melanogaster. Adv. Genet. 27, 309362.Google Scholar
Kawahira, N., Ohtsuka, D., Kida, N., Hironaka, K.-i., and Morishita, Y. (2020). Quantitative analysis of 3D tissue deformation reveals key cellular mechanism associated with initial heart looping. Cell Rep. 30, 38893903.Google Scholar
Kawamori, A. and Yamaguchi, M. (2011). DREF is critical for Drosophila bristle development by regulating endoreplication in shaft cells. Cell Struct. Funct. 36, 103119.Google Scholar
Kawasaki, S., Makuuchi, M., Ishizone, S., Matsunami, H., Terada, M., and Kawarazaki, H. (1992). Liver regeneration in recipients and donors after transplantation. Lancet 339, 580581.Google Scholar
Keeler, R.F. and Binns, W. (1968). Teratogenic compounds of Veratrum californicum (Durand). V. Comparison of cyclopian effects of steroidal alkaloids from the plant and structurally related compounds from other sources. Teratology 1, 510.Google Scholar
Kemp, P.R., Griffiths, M., and Polkey, M.I. (2019). Muscle wasting in the presence of disease, why is it so variable? Biol. Rev. 94, 10381055.Google Scholar
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, 407419.Google Scholar
Keyte, A.L. and Smith, K.K. (2014). Heterochrony and developmental timing mechanisms: changing ontogenies in evolution. Semin. Cell Dev. Biol. 34, 99107.Google Scholar
Kidd, W. (1920). Initiative in Evolution. H. F. & G. Witherby, London.Google Scholar
Kidson, S. and Fabian, B. (1979). Pigment synthesis in the Himalayan mouse. J. Exp. Zool. 210, 145152.Google Scholar
Kiecker, C. and Lumsden, A. (2012). The role of organizers in patterning the nervous system. Annu. Rev. Neurosci. 35, 347367.Google Scholar
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, 133138.Google Scholar
Kim, J., Williams, F.J., Dreger, D.L., Plassais, J., Davis, B.W., Parker, H.G., and Ostrander, E.A. (2018). Genetic selection of athletic success in sport-hunting dogs. PNAS 115, E7212E7221.Google Scholar
Kim, M.J., Oh, H.J., Kim, G.A., Park, J.E., Park, E.J., Jang, G., Ra, J.C., Kang, S.K., and Lee, B.C. (2012). Lessons learned from cloning dogs. Reprod. Dom. Anim. 47(Suppl. 4), 115119.Google Scholar
King, R.A., Townsend, D., Oetting, W., Summers, C.G., Olds, D.P., White, J.G., and Spritz, R.A. (1991). Temperature-sensitive tyrosinase associated with peripheral pigmentation in oculocutaneous albinism. J. Clin. Invest. 87, 10461053.Google Scholar
Kingsley, M.C.S. and Ramsay, M.A. (1988). The spiral in the tusk of the narwhal. Arctic 41, 236238.Google Scholar
Kirikoshi, H., Sekihara, H., and Katoh, M. (2001). WNT10A and WNT6, clustered in human chromosome 2q35 region with head-to-tail manner, are strongly coexpressed in SW480 cells. Biochem. Biophys. Res. Comm. 283, 798805.Google Scholar
Kirschner, M.W. and Gerhart, J.C. (2005). The Plausibility of Life: Resolving Darwin’s Dilemma. Yale University Press, New Haven, CT.Google Scholar
Kiskowski, M., Glimm, T., Moreno, N., Gamble, T., and Chiari, Y. (2019). Isolating and quantifying the role of developmental noise in generating phenotypic variation. PLoS Comput. Biol. 15(4), e1006943.Google Scholar
Kivell, T.L., Lemelin, P., Richmond, B.G., and Schmitt, D., eds. (2016). The Evolution of the Primate Hand. Springer, New York.Google Scholar
Klaassen, Z., Shoja, M.M., Tubbs, R.S., and Loukas, M. (2011). Supernumerary and absent limbs and digits of the lower limb: a review of the literature. Clin. Anat. 24, 570575.Google Scholar
Klar, A.J.S. (2003). Human handedness and scalp hair-whorl direction develop from a common genetic mechanism. Genetics 165, 269276.Google Scholar
Klar, A.J.S. (2005). A 1927 study supports a current genetic model for inheritance of human scalp hair-whorl orientation and hand-use preference traits. Genetics 170, 20272030.Google Scholar
Klein, T. and Martinez Arias, A. (1999). The Vestigial gene product provides a molecular context for the interpretation of signals during the development of the wing in Drosophila. Development 126, 913925.Google Scholar
Kley, N.J. and Kearney, M. (2007). Adaptations for digging and burrowing. In Fins into Limbs: Evolution, Development, and Transformation, Hall, B.K., editor. University of Chicago Press, Chicago, IL, pp. 284309.Google Scholar
Klimczewska, K., Kasperczuk, A., and Suwinska, A. (2018). The regulative nature of mammalian embryos. Curr. Top. Dev. Biol. 128, 105149.Google Scholar
Klingenberg, C.P. (1998). Heterochrony and allometry: the analysis of evolutionary change in ontogeny. Biol. Rev. 73, 79123.Google Scholar
Kmita, M. and Duboule, D. (2003). Organizing axes in time and space: 25 years of colinear tinkering. Science 301, 331333.Google Scholar
Kmita, M., Fraudeau, N., Hérault, Y., and Duboule, D. (2002). Serial deletions and duplications suggest a mechanism for the collinearity of Hoxd genes in limbs. Nature 420, 145150.Google Scholar
Knebel, D., Rillich, J., Ayali, A., Pflüger, H.-J., and Rigosi, E. (2018). Ex vivo recordings reveal desert locust forelimb control is asymmetric. Curr. Biol. 28, R1283R1295.Google Scholar
Knezevic, V., De Santo, R., Schughart, K., Huffstadt, U., Chiang, C., Mahon, K.A., and Mackem, S. (1997). Hoxd-12 differentially affects preaxial and postaxial chondrogenic branches in the limb and regulates Sonic hedgehog in a positive feedback loop. Development 124, 45234536.Google Scholar
Koca, Y., Housden, B.E., Gault, W.J., Bray, S.J., and Mlodzik, M. (2019). Notch signaling coordinates ommatidial rotation in the Drosophila eye via transcriptional regulation of the EGF-Receptor ligand Argos. Sci. Rep. 9, 18628.Google Scholar
Koch, S.L., Tridico, S.R., Bernard, B.A., Shriver, M.D., and Jablonski, N.G. (2019). The biology of human hair: a multidisciplinary review. Am. J. Hum. Biol. 2019, e23316.Google Scholar
Koenig, K.M., Sun, P., Meyer, E., and Gross, J.M. (2016). Eye development and photoreceptor differentiation in the cephalopod Doryteuthis pealeii. Development 143, 31683181.Google Scholar
Koga, A., Hisakawa, C., and Yoshizawa, M. (2020). Baboon bearing resemblance in pigmentation pattern to Siamese cat carries a missense mutation in the tyrosinase gene. Genome 63, 275279.Google Scholar
Kohler, R.E. (1994). Lords of the Fly: Drosophila Genetics and the Experimental Life. University of Chicago Press, Chicago, IL.Google Scholar
Kojima, T. (2004). The mechanism of Drosophila leg development along the proximodistal axis. Dev. Growth Differ. 46, 115129.Google Scholar
Kojima, T. (2017). Developmental mechanism of the tarsus in insect legs. Curr. Opin. Insect Sci. 19, 3642.Google Scholar
Kondo, S. and Asai, R. (1995). A reaction–diffusion wave on the skin of the marine angelfish Pomacanthus. Nature 376, 765768.Google Scholar
Kondo, S. and Miura, T. (2010). Reaction–diffusion model as a framework for understanding biological pattern formation. Science 329, 16161620.Google Scholar
Kondo, S. and Shirota, H. (2009). Theoretical analysis of mechanisms that generate the pigmentation pattern of animals. Semin. Cell Dev. Biol. 20, 8289.Google Scholar
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, 851856.Google Scholar
Kondo, T. and Hayashi, S. (2015). Mechanisms of cell height changes that mediate epithelial invagination. Dev. Growth Differ. 57, 313323.Google Scholar
Kong, X.-Z., Mathias, S.R., Guadalupe, T., Group, E.L.W., Glahn, D.C., Franke, B., Crivello, F., Tzourio-Mazoyer, N., Fisher, S.E., Thompson, P.M., and Francks, C. (2018). Mapping cortical brain asymmetry in 17,141 healthy individuals worldwide via the ENIGMA Consortium. PNAS 115, ES154ES163.Google Scholar
Kopp, A. (2009). Metamodels and phylogenetic replication: a systematic approach to the evolution of developmental pathways. Evolution 63, 27712789.Google Scholar
Kopp, A. (2011). Drosophila sex combs as a model of evolutionary innovations. Evol. Dev. 13, 504522.Google Scholar
Kopp, A. (2012). Dmrt genes in the development and evolution of sexual dimorphism. Trends Genet. 28, 175184.Google Scholar
Kopp, A. and True, J.R. (2002). Evolution of male sexual characters in the Oriental Drosophila melanogaster species group. Evol. Dev. 4, 278291.Google Scholar
Koshikawa, S. (2015). Enhancer modularity and the evolution of new traits. Fly 9, 155159.Google Scholar
Krapp, H.G. (2009). Ocelli. Curr. Biol. 19, R435R437.Google Scholar
Kruggel, F. and Solodkin, A. (2019). Determinants of structural segregation and patterning in the human cortex. NeuroImage 196, 248260.Google Scholar
Krumlauf, R. (2016). Hox genes and the hindbrain: a study in segments. Curr. Top. Dev. Biol. 116, 581596.Google Scholar
Kücken, M. (2007). Models for fingerprint pattern formation. Forensic Sci. Int. 171, 8596.Google Scholar
Kücken, M. and Newell, A.C. (2005). Fingerprint formation. J. Theor. Biol. 235, 7183.Google Scholar
Kumar, J.P. (2012). Building an ommatidium one cell at a time. Dev. Dynamics 241, 136149.Google Scholar
Kunhardt, P.B. Jr., Kunhardt, P.B. III, and Kunhardt, P.W. (1995). P. T. Barnum: America’s Greatest Showman. Knopf, New York.Google Scholar
Kuratani, S. (2012). Evolution of the vertebrate jaw from developmental perspectives. Evol. Dev. 14, 7692.Google Scholar
Kutsarova, E., Munz, M., and Ruthazer, E.S. (2017). Rules for shaping neural connections in the developing brain. Front. Neural Circuits 10, 11.Google Scholar
Kwak, J.Y. and Kwon, K.-S. (2019). Pharmacological interventions for treatment of sarcopenia: current status of drug development for sarcopenia. Ann. Geriatr. Med. Res. 23, 98104.Google Scholar
Kwon, B.S., Halaban, R., and Chintamaneni, C. (1989). Molecular basis of mouse Himalayan mutation. Biochem. Biophys. Res. Comm. 161, 252260.Google Scholar
Ladher, R.K. (2017). Changing shape and shaping change: inducing the inner ear. Dev. Biol. 65, 3946.Google Scholar
Ladoux, B., Mège, R.-M., and Trepat, X. (2016). Front–rear polarization by mechanical cues: from single cells to tissues. Trends Cell Biol. 26, 420433.Google Scholar
Lafuente, E. and Beldade, P. (2019). Genomics of developmental plasticity in animals. Front. Genet. 10, 720.Google Scholar
Lamas, J.A., Rueda-Ruzafa, L., and Herrera-Pérez, S. (2019). Ion channels and thermosensitivity: TRP, TREK, or both? Int. J. Mol. Sci. 20, 2371.Google Scholar
Lancaster, M.A. (2018). Crinkle-cut brain organoids. Cell Stem Cell 22, 616618.Google Scholar
Lander, A.D. (2011). Pattern, growth, and control. Cell 144, 955969.Google Scholar
Landge, A.N., Jordan, B.M., Diego, X., and Müller, P. (2020). Pattern formation mechanisms of self-organizing reaction–diffusion systems. Dev. Biol. 460, 211.Google Scholar
Lange, A. and Müller, G.B. (2017). Polydactyly in development, inheritance, and evolution. Q. Rev. Biol. 92, 138.Google Scholar
Lange, A., Nemeschkal, H.L., and Müller, G.B. (2014). Biased polyphenism in polydactylous cats carrying a single point mutation: the Hemingway model for digit novelty. Evol. Biol. 41, 262275.Google Scholar
Larsen, E.W. (1997). Evolution of development: the shuffling of ancient modules by ubiquitous bureaucracies. In Physical Theory in Biology: Foundations and Explorations, Lumsden, C.J., Brandts, W.A., and Trainor, L.E.H., editors. World Scientific, Singapore, pp. 431441.Google Scholar
Larue, L., de Vuyst, F., and Delmas, V. (2013). Modeling melanoblast development. Cell. Mol. Life Sci. 70, 10671079.Google Scholar
Laubichler, M.D. and Hall, B.K. (2008). Conrad Hal Waddington: forefather of theoretical EvoDevo. Biol. Theory 3, 185187.Google Scholar
Laufer, E., Dahn, R., Orozco, O.E., Yeo, C.-Y., Pisenti, J., Henrique, D., Abbott, U.K., Fallon, J.F., and Tabin, C. (1997). Expression of Radical fringe in limb-bud ectoderm regulates apical ectodermal ridge formation. Nature 386, 366373.Google Scholar
Laurent-Gengoux, P., Petit, V., and Larue, L. (2019). Modeling and analysis of melanoblast motion. J. Math. Biol. 79, 21112132.Google Scholar
Lavine, L., Gotoh, H., Brent, C.S., Dworkin, I., and Emlen, D.J. (2015). Exaggerated trait growth in insects. Annu. Rev. Entomol. 60, 453472.Google Scholar
Lawrence, P. (2020). Practice makes perfect: Sir Michael J. Berridge. Curr. Biol. 20, R377.Google Scholar
Lawrence, P.A. (2004). A Wigglesworth classic: how cells make patterns. J. Exp. Biol. 207, 192193.Google Scholar
Lawrence, P.A. (2019). Sydney Brenner: a master of science and wit. Development 146, dev179879.Google Scholar
Lawrence, P.A. and Casal, J. (2018). Planar cell polarity: two genetic systems use one mechanism to read gradients. Development 145, dev168229.Google Scholar
Lawrence, P.A. and Locke, M. (1997). A man for our season. Nature 386, 757758.Google Scholar
Lawrence, P.A., Struhl, G., and Casal, J. (2008). Do the protocadherins Fat and Dachsous link up to determine both planar cell polarity and the dimensions of organs? Nat. Cell Biol. 10, 13791382.Google Scholar
Lawrence, P.A., Struhl, G., and Morata, G. (1979). Bristle patterns and compartment boundaries in the tarsi of Drosophila. J. Embryol. Exp. Morphol. 51, 195208.Google Scholar
Lawton, A.K., Engstrom, T., Rohrbach, D., Omura, M., Turnbull, D.H., Mamou, J., Zhang, T., Schwarz, J.M., and Joyner, A.L. (2019). Cerebellar folding is initiated by mechanical constraints on a fluid-like layer without a cellular pre-pattern. eLife 8, e45019.Google Scholar
Le Garrec, J.-F. and Kerszberg, M. (2008). Modeling polarity buildup and cell fate decision in the fly eye: insight into the connection between the PCP and Notch pathways. Dev. Genes Evol. 218, 413426.Google Scholar
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., editors. Clarendon Press, Oxford, pp. 122.Google Scholar
Lebreton, G., Géminard, C., Lapraz, F., Pyrpassopoulos, S., Cerezo, D., Spéder, P., Ostap, E.M., and Noselli, S. (2018). Molecular to organismal chirality is induced by the conserved myosin 1D. Science 362, 949952.Google Scholar
Lee, D.E., Cavener, D.R., and Bond, M.L. (2018). Seeing spots: quantifying mother–offspring similarity and assessing fitness consequences of coat pattern traits in a wild population of giraffes (Giraffa camelopardalis). PeerJ 6, e5690.Google Scholar
Lee, L.A. and Orr-Weaver, T.L. (2003). Regulation of cell cycles in Drosophila development: intrinsic and extrinsic cues. Annu. Rev. Genet. 37, 545578.Google Scholar
Lee, S.-J. (2007). Quadrupling muscle mass in mice by targeting TGF-β signaling pathways. PLoS ONE 2(8), e789.Google Scholar
Lee, S.-J. and McPherron, A.C. (2001). Regulation of myostatin activity and muscle growth. PNAS 98, 93069311.Google Scholar
Leevers, S.J., Weinkove, D., MacDougall, L.K., Hafen, E., and Waterfield, M.D. (1996). The Drosophila phosphoinositide 3-kinase Dp110 promotes cell growth. EMBO J. 15, 65846594.Google Scholar
Lejeune, E., Dortdivanlioglu, B., Kuhl, E., and Linder, C. (2019). Understanding the mechanical link between oriented cell division and cerebellar morphogenesis. Soft Matter 15, 22042215.Google Scholar
Lemons, D. and McGinnis, W. (2006). Genomic evolution of Hox gene clusters. Science 313, 19181922.Google Scholar
Lenhoff, H.M., Wang, P.P., Greenberg, F., and Bellugi, U. (1997). Williams Syndrome and the brain. Sci. Am. 277(6), 6873.Google Scholar
Lerner, A.B. and Fitzpatrick, T.B. (1950). Biochemistry of melanin formation. Physiol. Rev. 30, 91126.Google Scholar
Leroi, A.M. (2003). Mutants: On Genetic Variety and the Human Body. Viking Press, New York.Google Scholar
Lettice, L.A., Williamson, I., Devenney, P.S., Kilanowski, F., Dorin, J., and Hill, R.E. (2014). Development of five digits is controlled by a bipartite long-range cis-regulator. Development 141, 17151725.Google Scholar
Leventhal, A.G., Vitek, D.J., and Creel, D.J. (1985). Abnormal visual pathways in normally pigmented cats that are heterozygous for albinism. Science 229, 13951397.Google Scholar
Levin, M. (1999). Twinning and embryonic left–right asymmetry. Laterality 4, 197208.Google Scholar
Levin, M. (2012). Morphogenetic fields in embryogenesis, regeneration, and cancer: non-local control of complex patterning. BioSystems 109, 243261.Google Scholar
Levin, M., Klar, A.J.S., and Ramsdell, A.F. (2016). Introduction to provocative questions in left–right asymmetry. Philos. Trans. R. Soc. Lond. B 371, 20150399.Google Scholar
Levin, M., Roberts, D.J., Holmes, L.B., and Tabin, C. (1996). Laterality defects in conjoined twins. Nature 384, 321.Google Scholar
Levis, N.A. and Pfennig, D.W. (2020). Plasticity‐led evolution: a survey of developmental mechanisms and empirical tests. Evol. Dev. 22, 7187.Google Scholar
Lewis, E.B. (1976). Alfred Henry Sturtevant. In Dictionary of Scientific Biography. C. Scribner’s Sons, New York, pp. 133138.Google Scholar
Lewis, E.B. (1978). A gene complex controlling segmentation in Drosophila. Nature 276, 565570.Google Scholar
Lewis, E.B. (1994). Homeosis: the first 100 years. Trends Genet. 10, 341343.Google Scholar
Lewis, E.B. (1998). The bithorax complex: the first fifty years. Int. J. Dev. Biol. 42, 403415.Google Scholar
Lewis, E.B., Pfeiffer, B.D., Mathog, D.R., and Celniker, S.E. (2003). Evolution of the homeobox complex in the Diptera. Curr. Biol. 13, R587R588.Google Scholar
Lewis, J. (1982). Continuity and discontinuity in pattern formation. In Developmental Order: Its Origin and Regulation, Subtelny, S. and Green, P.B., editors. Liss, New York, pp. 511531.Google Scholar
Li, A., Xue, J., and Peterson, E.H. (2008). Architecture of the mouse utricle: macular organization and hair bundle heights. J. Neurophysiol. 99, 718733.Google Scholar
Li, C., Littlejohn, R.P., and Suttie, J.M. (1999). Effects of insulin-like growth factor 1 and testosterone on the proliferation of antlerogenic cells in vitro. J. Exp. Zool. 284, 8290.Google Scholar
Li, C., Zhao, H., Liu, Z., and McMahon, C. (2014). Deer antler: a novel model for studying organ regeneration in mammals. Int. J. Biochem. Cell Biol. 56, 111122.Google Scholar
Li, H., Mao, Y., Bouaziz, M., Yu, H., Qu, X., Wang, F., Feng, G.-S., Shawber, C., and Zhang, X. (2019). Lens differentiation is controlled by the balance between PDGF and FGF signaling. PLoS Biol. 17(2), e3000133.Google Scholar
Li, P. and Elowitz, M.B. (2019). Communication codes in developmental signaling pathways. Development 146, dev170977.Google Scholar
Li, X., Guo, C., and Li, L. (2019). Functional morphology and structural characteristics of the hind wings of the bamboo weevil Cyrtotrachelus buqueti (Coleoptera, Curculionidae). Anim. Cells Syst. 23, 143153.Google Scholar
Lichtwark, G.A. and Kelly, L.A. (2020). Ahead of the curve in the evolution of human feet. Nature 579, 3132.Google Scholar
Lim, Y.H., Moscato, Z., and Choate, K.A. (2017). Mosaicism in cutaneous disorders. Annu. Rev. Genet. 51, 123141.Google Scholar
Lin, A.Y. and Wang, L.H. (2018). Molecular therapies for muscular dystrophies. Curr. Treat. Options Neurol. 20, 27.Google Scholar
Lin, J.Y. and Fisher, D.E. (2007). Melanocyte biology and skin pigmentation. Nature 445, 843850.Google Scholar
Lincoln, G.A., Clarke, I.J., Hut, R.A., and Hazelrigg, D.G. (2006). Characterizing a mammalian circannual pacemaker. Science 314, 19411944.Google Scholar
Lindsley, D.L. and Grell, E.H. (1968). Genetic Variations of Drosophila melanogaster. Carnegie Institution of Washington, Washington, DC.Google Scholar
Lipshitz, H.D., ed. (2004). Genes, Development and Cancer: The Life and Work of Edward B. Lewis. Springer, New York.Google Scholar
Lipshitz, H.D. (2005). From fruit flies to fallout: Ed Lewis and his science. Dev. Dynamics 232, 529546.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” Megaloceros giganteus. Nature 438, 850853.Google Scholar
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, 979983.Google Scholar
Liu, F., Chen, Y., Zhu, G., Hysi, P.G., Wu, S., Adhikari, K., Breslin, K., Pospiech, E., Hamer, M.A., Peng, F., Muralidharan, C., Acuna-Alonzo, V., Canizales-Quinteros, S., Bedoya, G., Gallo, C., Poletti, G., Rothhammer, F., Bortolini, M.C., Gonzalez-Jose, R., Zeng, C., Xu, S., Jin, L., Uitterlinden, A.G., Ikram, M.A., van Duijn, C.M., Nijsten, T., Walsh, S., Branicki, W., Wang, S., Ruiz-Linares, A., Spector, T.D., Martin, N.G., Medland, S.E., and Kayser, M. (2018). Meta-analysis of genome-wide association studies identifies 8 novel loci involved in shape variation of human head hair. Hum. Mol. Genet. 27, 559575.Google Scholar
Liu, F., Hamer, M.A., Deelen, J., Lall, J.S., Jacobs, L., van Heemst, D., Murray, P.G., Wollstein, A., de Craen, A.J.M., Uh, H.-W., Zeng, C., Hofman, A., Uitterlinden, A.G., Houwing-Duistermaat, J.J., Pardo, L.M., Beekman, M., Slagboom, P.E., Nijsten, T., Kayser, M., and Gunn, D.A. (2016). The MC1R gene and youthful looks. Curr. Biol. 26, 12131220.Google Scholar
Liu, R.T., Liaw, S.S., and Maini, P.K. (2006). Two-stage Turing model for generating pigment patterns on the leopard and the jaguar. Phys. Rev. E 74, 011914.Google Scholar
Liu, W., Selever, J., Lu, M.-F., and Martin, J.F. (2003). Genetic dissection of Pitx2 in craniofacial development uncovers new functions in branchial arch morphogenesis, late aspects of tooth morphogenesis and cell migration. Development 130, 63756385.Google Scholar
Liu, X. and Smagghe, G. (2019). Roles of the insulin signaling pathway in insect development and organ growth. Peptides 122, 169923.Google Scholar
Llinares-Benadero, C. and Borrell, V. (2019). Deconstructing cortical folding: genetic, cellular and mechanical determinants. Nat. Rev. Neurosci. 20, 161176.Google Scholar
Lo Celso, C., Prowse, D.M., and Watt, F.M. (2004). Transient activation of β-catenin signalling in adult mouse epidermis is sufficient to induce new hair follicles but continuous activation is required to maintain hair follicle tumours. Development 131, 17871799.Google Scholar
Loison, L. (2013). Georges Teissier (1900–1972) and the Modern Synthesis in France. Genetics 195, 295302.Google Scholar
Long, H.K., Prescott, S.L., and Wysocka, J. (2016). Ever-changing landscapes: transcriptional enhancers in development and evolution. Cell 167, 11701187.Google Scholar
Long, J.A. and Cloutier, R. (2020). The unexpected origin of fingers. Sci. Am. 322(6), 4653.Google Scholar
Long, K.R. and Huttner, W.B. (2019). How the extracellular matrix shapes neural development. Open Biol. 9, 180216.Google Scholar
Long, K.R., Newland, B., Florio, M., Kalebic, N., Langen, B., Kolterer, A., Wimberger, P., and Huttner, W.B. (2018). Extracellular matrix components HAPLN1, lumican, and collagen I cause hyaluronic acid-dependent folding of the developing human neocortex. Neuron 99, 702719.Google Scholar
Lui, J.C. and Baron, J. (2011). Mechanisms limiting body growth in mammals. Endocr. Rev. 32, 422440.Google Scholar
Lunde, K.B. and Johnson, P.T.J. (2012). A practical guide for the study of malformed amphibians and their causes. J. Herpetol. 46, 429441.Google Scholar
Luo, S.-J., Liu, Y.-C., and Xu, X. (2019). Tigers of the world: genomics and conservation. Annu. Rev. Anim. Biosci. 7, 521548.Google Scholar
Lyon, M.F. (1999). X-chromosome inactivation. Curr. Biol. 9, R235R237.Google Scholar
Lyons, L.A., Imes, D.L., Rah, H.C., and Grahn, R.A. (2005). Tyrosinase mutations associated with Siamese and Burmese patterns in the domestic cat (Felis catus). Anim. Genet. 36, 119126.Google Scholar
Maartens, A. (2017). On Growth and Form in context: an interview with Matthew Jarron. Development 144, 41994202.Google Scholar
Maartens, A.P. and Brown, N.H. (2015). Anchors and signals: the diverse roles of integrins in development. Curr. Top. Dev. Biol. 112, 233272.Google Scholar
Maas, A.-H. (1948). Über die Auslösbarkeit von Temperatur-Modifikationen während der Embryonal-Entwicklung von Drosophila melanogaster Meigen. W. Roux Arch. Entw.-Mech. Org. 143, 515572.Google Scholar
Macabenta, F. and Stathopoulos, A. (2019). Sticking to a plan: adhesion and signaling control spatial organization of cells within migrating collectives. Curr. Opin. Genet. Dev. 57, 3946.Google 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
MacKenzie, T.C., Crombleholme, T.M., Johnson, M.P., Schnaufer, L., Flake, A.W., Hedrick, H.L., Howell, L.J., and Adzick, N.S. (2002). The natural history of prenatally diagnosed conjoined twins. J. Pediatr. Surg. 37, 303309.Google Scholar
Madhavan, M.M. and Schneiderman, H.A. (1977). Histological analysis of the dynamics of growth of imaginal discs and histoblast nests during the larval development of Drosophila melanogaster. W. Roux Arch. Dev. Biol. 183, 269305.Google Scholar
Maeda, R.K. and Karch, F. (2006). The ABC of the BX-C: the bithorax complex explained. Development 133, dev02323.Google Scholar
Maeda, R.K. and Karch, F. (2015). The open for business model of the bithorax complex in Drosophila. Chromosoma 124, 293307.Google Scholar
Magklara, A. and Lomvardas, S. (2013). Stochastic gene expression in mammals: lessons from olfaction. Trends Cell Biol. 23, 449456.Google Scholar
Malacinski, G.M., ed. (1990). Cytoplasmic Organization Systems. McGraw-Hill, New York.Google Scholar
Malagón, J.N., Ahuja, A., Sivapatham, G., Hung, J., Lee, J., Muñoz-Gómez, S.A., Atallah, J., Singh, R.S., and Larsen, E. (2014). Evolution of Drosophila sex comb length illustrates the inextricable interplay between selection and variation. PNAS 111, E4103E4109.Google Scholar
Mallarino, R., Henegar, C., Mirasierra, M., Manceau, M., Schradin, C., Vallejo, M., Beronja, S., Barsh, G.S., and Hoekstra, H.E. (2016). Developmental mechanisms of stripe patterns in rodents. Nature 539, 518523.Google Scholar
Mallo, M. (2018). Reassessing the role of Hox genes during vertebrate development and evolution. Trends Genet. 34, 209217.Google Scholar
Malmström, T. and Kröger, R.H.H. (2006). Pupil shapes and lens optics in the eyes of terrestrial vertebrates. J. Exp. Biol. 209, 1825.Google Scholar
Manmadhan, S. and Ehmer, U. (2019). Hippo signaling in the liver: a long and ever-expanding story. Front. Cell Dev. Biol. 7, 33.Google Scholar
Mann, R.S. and Morata, G. (2000). The developmental and molecular biology of genes that subdivide the body of Drosophila. Annu. Rev. Cell Dev. Biol. 16, 243271.Google Scholar
Manocha, S., Farokhnia, N., Khosropanah, S., Bertol, J.W., Junior, J.S., and Fakhouri, W.D. (2019). Systematic review of hormonal and genetic factors involved in the nonsyndromic disorders of the lower jaw. Dev. Dynamics 248, 162172.Google Scholar
Marcellini, S., Gibert, J.-M., and Simpson, P. (2005). achaete, but not scute, is dispensable for the peripheral nervous system of Drosophila. Dev. Biol. 285, 545553.Google Scholar
Marcon, L. and Sharpe, J. (2012). Turing patterns in development: what about the horse part? Curr. Opin. Genet. Dev. 22, 578584.Google Scholar
Marmor, M.F., Choi, S.S., Zawadzki, R.J., and Werner, J.S. (2008). Visual insignificance of the foveal pit. Arch. Ophthalmol. 126, 907913.Google Scholar
Marsh, J.L. and Theisen, H. (1999). Regeneration in insects. Semin. Cell Dev. Biol. 10, 365375.Google Scholar
Martin, A.C. and Goldstein, B. (2014). Apical constriction: themes and variations on a cellular mechanism driving morphogenesis. Development 141, 19871998.Google Scholar
Martín, M., Organista, M.F., and de Celis, J.F. (2016). Structure of developmental gene regulatory networks from the perspective of cell fate-determining genes. Transcription 7, 3237.Google Scholar
Martinez Arias, A. and Steventon, B. (2018). On the nature and function of organizers. Development 145, dev159525.Google Scholar
Martinez Arias, A., Nichols, J., and Schröter, C. (2013). A molecular basis for developmental plasticity in early mammalian embryos. Development 140, 34993510.Google Scholar
Martini, F.H., Ober, W.C., Garrison, C.W., Welch, K., Hutchings, R.T., and Ireland, K. (2004). Fundamentals of Anatomy and Physiology, 6th ed. Benjamin Cummings, San Francisco, CA.Google Scholar
Mason, C. and Guillery, R. (2019). Conversations with Ray Guillery on albinism: linking Siamese cat visual pathway connectivity to mouse retinal development. Eur. J. Neurosci. 49, 913927.Google Scholar
Massey, J.H., Chung, D., Siwanowicz, I., Stern, D.L., and Wittkopp, P.J. (2019). The yellow gene influences Drosophila male mating success through sex comb melanization. eLife 8, e49388.Google Scholar
Matis, M. (2020). The mechanical role of microtubules in tissue remodeling. BioEssays 42, 1900244.Google Scholar
Matis, M. and Axelrod, J. (2013). Regulation of PCP by the Fat signaling pathway. Genes Dev. 27, 22072220.Google Scholar
Matsuda, K., Gotoh, H., Tajika, Y., Sushida, T., Aonuma, H., Niimi, T., Akiyama, M., Inoue, Y., and Kondo, S. (2017). Complex furrows in a 2D epithelial sheet code the 3D structure of a beetle horn. Sci. Rep. 7, 13939.Google Scholar
Matt, G. and Umen, J. (2016). Volvox: a simple algal model for embryogenesis, morphogenesis and cellular differentiation. Dev. Biol. 419, 99113.Google Scholar
May-Simera, H. and Kelley, M.W. (2012). Planar cell polarity in the inner ear. Curr. Top. Dev. Biol. 101, 111140.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, 265287.Google Scholar
Mayor, R. and Theveneau, E. (2013). The neural crest. Development 140, 22472251.Google Scholar
McAvoy, J.W., Dawes, L.J., Sugiyama, Y., and Lovicu, F.J. (2017). Intrinsic and extrinsic regulatory mechanisms are required to form and maintain a lens of the correct size and shape. Exp. Eye Res. 156, 3440.Google Scholar
McClure, K.D. and Schubiger, G. (2005). Developmental analysis and squamous morphogenesis of the peripodial epithelium in Drosophila imaginal discs. Development 132, 50335042.Google Scholar
McElreath, R. (2018). Sizing up human brain evolution. Nature 557, 496497.Google Scholar
McGhee, G.R. Jr. (2011). Convergent Evolution: Limited Forms Most Beautiful. Vienna Series in Theoretical Biology, Müller, G.B., Wagner, G.P., and Callebaut, W., editors. MIT Press, Cambridge, MA.Google Scholar
McHugo, G.P., Dover, M.J., and MacHugh, D.E. (2019). Unlocking the origins and biology of domestic animals using ancient DNA and paleogenomics. BMC Biol. 17, 98.Google Scholar
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, 6171.Google Scholar
McKenna, D.D. and Farrell, B.D. (2010). 9-Genes reinforce the phylogeny of holometabola and yield alternate views on the phylogenetic placement of Strepsiptera. PLoS ONE 5(7), e11887.Google Scholar
Mcketton, L., Kelly, K.R., and Schneider, K.A. (2014). Abnormal lateral geniculate nucleus and optic chiasm in human albinism. J. Comp. Neurol. 522, 26802687.Google Scholar
McLean, W.H.I. (2008). Combing the genome for the root cause of baldness. Nat. Genet. 11, 12701271.Google Scholar
McNeill, H. (2010). Planar cell polarity: keeping hairs straight is not so simple. Cold Spring Harb. Perspect. Biol. 2, a003376.Google Scholar
McPherron, A.C. and Lee, S.-J. (1997). Double muscling in cattle due to mutations in the myostatin gene. PNAS 94, 1245712461.Google Scholar
McPherron, A.C. and Lee, S.-J. (2002). Suppression of body fat accumulation in myostatin-deficient mice. J. Clin. Invest. 109, 595601.Google Scholar
McPherron, A.C., Lawler, A.M., and Lee, S.-J. (1997). Regulation of skeletal muscle mass in mice by a new TGF-β superfamily member. Nature 387, 8390.Google Scholar
Meinhardt, H. (2009). Models for the generation and interpretation of gradients. Cold Spring Harb. Perspect. Biol. 1, a001362.Google Scholar
Meinhardt, H. and Gierer, A. (1974). Applications of a theory of biological pattern formation based on lateral inhibition. J. Cell Sci. 15, 321346.Google Scholar
Meinhardt, H. and Gierer, A. (2000). Pattern formation by local self-activation and lateral inhibition. BioEssays 22, 753760.Google Scholar
Men, W., Falk, D., Sun, T., Chen, W., Li, J., Yin, D., Zang, L., and Fan, M. (2014). The corpus callosum of Albert Einstein’s brain: another clue to his high intelligence? Brain 137, 18.Google Scholar
Mermoud, J.E., Rowbotham, S.P., and Varga-Weisz, P.D. (2011). Keeping chromatin quiet: how nucleosome remodeling restores heterochromatin after replication. Cell Cycle 10, 40174025.Google Scholar
Merrell, A.J. and Stanger, B.Z. (2019). A feedback loop controlling organ size. Dev. Cell 48, 425426.Google Scholar
Mian, A., Gabra, N.I., Sharma, T., Topale, N., Gielecki, J., Tubbs, R.S., and Loukas, M. (2017). Conjoined twins: from conception to separation, a review. Clin. Anat. 30, 385396.Google Scholar
Michalopoulos, G.K. and DeFrances, M.C. (1997). Liver regeneration. Science 276, 6066.Google Scholar
Mickelson, J.R. and Valberg, S.J. (2015). The genetics of skeletal muscle disorders in horses. Annu. Rev. Anim. Biosci. 3, 197217.Google Scholar
Miesfeld, J.B. and Brown, N.L. (2019). Eye organogenesis: a hierarchical view of ocular development. Curr. Top. Dev. Biol. 132, 351393.Google Scholar
Migeon, B.R. (2016). An overview of X inactivation based on species differences. Semin. Cell Dev. Biol. 56, 111116.Google Scholar
Mikeladze-Dvali, T., Desplan, C., and Pistillo, D. (2005). Flipping coins in the fly retina. Curr. Top. Dev. Biol. 69, 115 (+ color plates).Google Scholar
Millar, S.E., Willert, K., Salinas, P.C., Roelink, H., Nusse, R., Sussman, D.J., and Barsh, G.S. (1999). WNT signaling in the control of hair growth and structure. Dev. Biol. 207, 133149.Google Scholar
Miller, G.S. Jr. (1931). Human Hair and Primate Patterning. Smithsonian Miscellaneous Collections Vol. 85 No. 10. Smithsonian Institution, Washington, DC.Google Scholar
Miller, S.W. and Posakony, J.W. (2018). Lateral inhibition: two modes of nonautonomous negative autoregulation by neuralized. PLoS Genet. 14(7), e1007528.Google Scholar
Miller, S.W., Rebeiz, M., Atanasov, J.E., and Posakony, J.W. (2014). Neural precursor-specific expression of multiple Drosophila genes is driven by dual enhancer modules with overlapping function. PNAS 111, 1719417199.Google Scholar
Mills, L.S., Zimova, M., Oyler, J., Running, S., Abatzoglou, J.T., and Lukacs, P.M. (2013). Camouflage mismatch in seasonal coat color due to decreased snow duration. PNAS 110, 73607365.Google Scholar
Milner, M.J., Bleasby, A.J., and Kelly, S.L. (1984). The role of the peripodial membrane of leg and wing imaginal discs of Drosophila melanogaster during evagination and differentiation in vitro. W. Roux Arch. Dev. Biol. 193, 180186.Google Scholar
Minor, K.M., Patterson, E.E., Keating, M.K., Gross, S.D., Ekenstedt, K.J., Taylor, S.M., and Mickelson, J.R. (2011). Presence and impact of the exercise-induced collapse associated DNM1 mutation in Labrador retrievers and other breeds. Vet. J. 189, 214219.Google Scholar
Mirth, C. and Akam, M. (2002). Joint development in the Drosophila leg: cell movements and cell populations. Dev. Biol. 246, 391406.Google Scholar
Mirth, C.K. and Shingleton, A.W. (2019). Coordinating development: how do animals integrate plastic and robust developmental processes? Front. Cell Dev. Biol. 7, 8.Google 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, 38233833.Google Scholar
Mitogawa, K., Makanae, A., and Satoh, A. (2018). Hyperinnervation improves Xenopus laevis limb regeneration. Dev. Biol. 433, 276286.Google Scholar
Mittwoch, U. (2000). Genetics of sex determination: exceptions that prove the rule. Mol. Genet. Metab. 71, 405410.Google Scholar
Miyashita, T. and Diogo, R. (2016). Evolution of serial patterns in vertebrate pharyngeal apparatus and paired appendages via assimilation of dissimilar units. Front. Ecol. Evol. 4, 71.Google Scholar
Moczek, A.P. (2010). Phenotypic plasticity and diversity in insects. Philos. Trans. R. Soc. Lond. B 365, 593603.Google Scholar
Modolell, J. and Campuzano, S. (1998). The achaete-scute complex as an integrating device. Int. J. Dev. Biol. 42, 275282.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, 356367.Google Scholar
Moiseff, A. (1989). Binaural disparity cues available to the barn owl for sound localization. J. Comp. Physiol. A 164, 629636.Google Scholar
Monier, B. and Suzanne, M. (2015). The morphogenetic role of apoptosis. Curr. Top. Dev. Biol. 114, 335362.Google Scholar
Monk, P.B. and Othmer, H.G. (1989). Relay, oscillations and wave propagation in a model of Dictyostelium discoideum. In Lectures on Mathematics in the Life Sciences, Vol. 21. American Mathematical Society, Providence, RI, pp. 87122.Google Scholar
Monsoro-Burq, A.H. and Levin, M. (2018). Avian models and the study of invariant asymmetry: how the chicken and the egg taught us to tell right from left. Int. J. Dev. Biol. 62, 6377.Google Scholar
Montagu, M.F.A. (1962). Time, morphology, and neoteny in the evolution of man. In Culture and the Evolution of Man, Montagu, M.F.A., editor. Oxford University Press, New York, pp. 324342.Google Scholar
Montavon, T. and Soshnikova, N. (2014). Hox gene regulation and timing in embryogenesis. Semin. Cell Dev. Biol. 34, 7684.Google Scholar
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, 346359.Google Scholar
Montgomery, S.H., Mundy, N.I., and Barton, R.A. (2016). Brain evolution and development: adaptation, allometry and constraint. Proc. R. Soc. B 283, 20160433.Google Scholar
Moog, U., Felbor, U., Has, C., and Zim, B. (2020). Disorders caused by genetic mosaicism. Dtsch. Arztebl. Int. 117, 119125.Google Scholar
Moore, R. and Alexandre, P. (2020). Delta–Notch signaling: the long and the short of a neuron’s influence on progenitor fates. J. Dev. Biol. 8, 8.Google Scholar
Moran, C., Gillies, C.B., and Nicholas, F.W. (1984). Fertile male tortoiseshell cats: mosaicism due to gene instability? J. Hered. 75, 397402.Google Scholar
Morata, G., Shlevkov, E., and Pérez-Garijo, A. (2011). Mitogenic signaling from apoptotic cells in Drosophila. Dev. Growth Differ. 53, 168176.Google Scholar
Moreno-Marmol, T., Cavodeassi, F., and Bovolenta, P. (2018). Setting eyes on the retinal pigment epithelium. Front. Cell Dev. Biol. 6, 145.Google Scholar
Morgan, T.H. and Bridges, C.B. (1919). The origin of gynandromorphs. In Contributions to the Genetics of Drosophila melanogaster. Carnegie Institution of Washington, Washington, DC, pp. 1122.Google Scholar
Moriyama, Y. and De Robertis, E.M. (2018). Embryonic regeneration by relocalization of the Spemann organizer during twinning in Xenopus. PNAS 115, E4815E4822.Google Scholar
Morris, D. (1967). The Naked Ape: A Zoologist’s Study of the Human Animal. Random House, New York.Google Scholar
Mosher, D.S., Quignon, P., Bustamante, C.D., Sutter, N.B., Mellersh, C.S., Parker, H.G., and Ostrander, E.A. (2007). A mutation in the myostatin gene increases muscle mass and enhances racing performance in heterozygote dogs. PLoS Genet. 3(5), e79.Google Scholar
Moulton, D.E., Goriely, A., and Chirat, R. (2018). How seashells take shape. Sci. Am. 318(4), 6875.Google Scholar
Moya, I.M. and Halder, G. (2019). Hippo–YAP/TAZ signalling in organ regeneration and regenerative medicine. Nat. Rev. Mol. Cell Biol. 20, 211226.Google Scholar
Muckli, L., Naumer, M.J., and Singer, W. (2009). Bilateral visual field maps in a patient with only one hemisphere. PNAS 106, 1303413039.Google 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.Google Scholar
Muller, G.H. (1990). Skin diseases of the Chinese Shar-Pei. Adv. Clin. Dermatol. 20, 16551670.Google Scholar
Murisier, F., Guichard, S., and Beermann, F. (2007). Distinct distal regulatory elements control tyrosinase expression in melanocytes and the retinal pigment epithelium. Dev. Biol. 303, 838847.Google Scholar
Murray, J.D. (1981). On pattern formation mechanisms for lepidopteran wing patterns and mammalian coat markings. Philos. Trans. Roy. Soc. Lond. B 295, 473496.Google Scholar
Murray, J.D. (1981). A pre-pattern formation mechanism for animal coat markings. J. Theor. Biol. 88, 161199.Google Scholar
Murray, J.D. (1988). How the leopard gets its spots. Sci. Am. 258(3), 8087.Google Scholar
Murray, J.D. (1989). Mathematical Biology. Springer-Verlag, Berlin.Google Scholar
Murray, J.D. (1990). Turing’s theory of morphogenesis: its influence on modelling biological pattern and form. Bull. Math. Biol. 52, 119152.Google Scholar
Murray, J.D. (2012). Vignettes from the field of mathematical biology: the application of mathematics to biology and medicine. Interface Focus 2, 397406.Google Scholar
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. R. Soc. B 239, 279293.Google Scholar
Mutzel, V., Okamoto, I., Dunkel, I., Saitou, M., Giorgetti, L., Heard, E., and Schulz, E.G. (2019). A symmetric toggle switch explains the onset of random X inactivation in different mammals. Nat. Struct. Mol. Biol. 26, 350360.Google Scholar
Nacu, E., Gromberg, E., Oliveira, C.R., Drechsel, D., and Tanaka, E.M. (2016). FGF8 and SHH substitute for anterior–posterior tissue interactions to induce limb regeneration. Nature 533, 407410.Google Scholar
Naef, A. (1926). Über die Urformen der Anthropomorphen und die Stammesgeschichte des Menschenschädels. Naturwiss. 14, 445452.Google Scholar
Närhi, K., Järvinen, E., Birchmeier, W., Taketo, M.M., Mikkola, M.L., and Thesleff, I. (2008). Sustained epithelial β-catenin activity induces precocious hair development but disrupts hair follicle down-growth and hair shaft formation. Development 135, 10191028.Google Scholar
Nascone, N. and Mercola, M. (1997). Organizer induction determines left–right asymmetry in Xenopus. Dev. Biol. 189, 6878.Google Scholar
Nasoori, A. (2020). Formation, structure, and function of extra-skeletal bones in mammals. Biol. Rev. Camb. Philos. Soc. 95, 9861019.Google 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, 450462.Google Scholar
Negre, B. and Ruiz, A. (2007). HOM-C evolution in Drosophila: is there a need for Hox gene clustering? Trends Genet. 23, 5559.Google Scholar
Negre, B. and Simpson, P. (2009). Evolution of the achaete-scute complex in insects: convergent duplication of proneural genes. Trends Genet. 25, 147152.Google Scholar
Negre, B., Casillas, S., Suzanne, M., Sánchez-Herrero, E., Akam, M., Nefedov, M., Barbadilla, A., de Jong, P., and Ruiz, A. (2005). Conservation of regulatory sequences and gene expression patterns in the disintegrating Drosophila Hox complex. Genome Res. 15, 692700.Google 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, 14491466.Google Scholar
Nelson, C.M. (2016). On buckling morphogenesis. J. Biomech. Eng. 138, 021005.Google Scholar
Nemec, S., Luxey, M., Jain, D., Sung, A.H., Pastinen, T., and Drouin, J. (2017). Pitx1 directly modulates the core limb development program to implement hindlimb identity. Development 144, 33253335.Google Scholar
Nerurkar, N.L., Mahadevan, L., and Tabin, C.J. (2017). BMP signaling controls buckling forces to modulate looping morphogenesis of the gut. PNAS 114, 22772282.Google Scholar
Neto-Silva, R.M., Wells, B.S., and Johnston, L.A. (2009). Mechanisms of growth and homeostasis in the Drosophila wing. Annu. Rev. Cell Dev. Biol. 25, 197220.Google Scholar
Neubauer, S., Gunz, P., Scott, N.A., Hublin, J.-J., and Mitteroecker, P. (2020). Evolution of brain lateralization: a shared hominid pattern of endocranial asymmetry is much more variable in humans than in great apes. Sci. Adv. 6, eaax9935.Google Scholar
Neumann, C.J. and Cohen, S.M. (1998). Boundary formation in Drosophila wing: Notch activity attenuated by the POU protein Nubbin. Science 281, 409413.Google Scholar
Neveu, M.M., Holder, G.E., Ragge, N.K., Sloper, J.J., Collin, J.R.O., and Jeffery, G. (2006). Early midline interactions are important in mouse optic chiasm formation but are not critical in man: a significant distinction between man and mouse. Eur. J. Neurosci. 23, 30343042.Google Scholar
Ng, C.S. and Kopp, A. (2008). Sex combs are important for male mating success in Drosophila melanogaster. Behav. Genet. 38, 195201.Google Scholar
Nguyen, H.Q., Lee, S.D., and Wu, C.-T. (2019). Paircounting. Trends Genet. 36, 787789.Google Scholar
Nicholas, F.W. and Hobbs, M. (2013). Mutation discovery for Mendelian traits in non-laboratory animals: a review of achievements up to 2012. Anim. Genet. 45, 157170.Google Scholar
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, 13091313.Google Scholar
Nijhout, H.F. (1999). Control mechanisms of polyphenic development in insects. BioScience 49, 181192.Google Scholar
Nijhout, H.F. (2001). Elements of butterfly wing patterns. J. Exp. Zool. 291, 213225.Google Scholar
Nijhout, H.F. (2003). The control of growth. Development 130, 58635867.Google Scholar
Nijhout, H.F. (2003). Development and evolution of adaptive polyphenisms. Evol. Dev. 5, 918.Google Scholar
Nijhout, H.F. (2010). Molecular and physiological basis of colour pattern formation. Adv. Insect Physiol. 38, 219265.Google Scholar
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, 4352.Google Scholar
Nijhout, H.F. and McKenna, K.Z. (2019). Allometry, scaling, and ontogeny of form: an introduction to the symposium. Integr. Comp. Biol. 59, 12751280.Google Scholar
Nissen, S.B., Perera, M., Gonzalez, J.M., Morgani, S.M., Jensen, M.H., Sneppen, K., Brickman, J.M., and Trusina, A. (2017). Four simple rules that are sufficient to generate the mammalian blastocyst. PLoS Biol. 15, e2000737.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, 22692279.Google Scholar
Nogare, D.D. and Chitnis, A.B. (2017). Self-organizing spots get under your skin. PLoS Biol. 15(12), e2004412.Google Scholar
Nonaka, S., Yoshiba, S., Watanabe, D., Ikeuchi, S., Goto, T., Marshall, W.F., and Hamada, H. (2005). De novo formation of left–right asymmetry by posterior tilt of nodal cilia. PLoS Biol. 3(8), e268.Google Scholar
Noonan, J.P. (2009). Regulatory DNAs and the evolution of human development. Curr. Opin. Genet. Dev. 19, 557564.Google Scholar
Nousbeck, J., Burger, B., Fuchs-Telem, D., Pavlovsky, M., Fenig, S., Sarig, O., Itin, P., and Sprecher, E. (2011). A mutation in a skin-specific isoform of SMARCAD1 causes autosomal-dominant adermatoglyphia. Am. J. Hum. Genet. 89, 302307.Google Scholar
Nüsslein-Volhard, C. (2019). Animal Beauty: On the Evolution of Biological Aesthetics. MIT Press, Cambridge, MA.Google Scholar
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, 10061016.Google Scholar
Nyholt, D.R., Gillespie, N.A., Heath, A.C., and Martin, N.G. (2003). Genetic basis of male pattern baldness. J. Invest. Dermatol. 121, 15611564.Google Scholar
O’Brien, S.J. (2004). Cats. Curr. Biol. 14, R988R989.Google Scholar
O’Connell, J.E.A. (1976). Craniopagus twins: surgical anatomy and embryology and their implications. J. Neurol., Neurosurg., and Psychiatry 39, 122.Google Scholar
O’Grady, P. and DeSalle, R. (2018). Hawaiian Drosophila as an evolutionary model clade: days of future past. BioEssays 40, 1700246.Google Scholar
Oda, H., Iwasaki-Yokozawa, S., Usui, T., and Akiyama-Oda, Y. (2020). Experimental duplication of bilaterian body axes in spider embryos: Holm’s organizer and self-regulation of embryonic fields. Dev. Genes Evol. 230, 4963.Google Scholar
Oetting, W.S. (2000). The tyrosinase gene and oculocutaneous albinism Type 1 (OCA1): a model for understanding the molecular biology of melanin formation. Pigment Cell Res. 13, 320325.Google Scholar
Okada, T.S. (2004). From embryonic induction to cell lineages: revisiting old problems for modern study. Int. J. Dev. Biol. 48, 739742.Google Scholar
Olman, C.A., Bao, P., Engel, S.A., Grant, A.N., Purington, C., Qiu, C., Schallmo, M.-P., and Tjan, B.S. (2018). Hemifield columns co-opt ocular dominance column structure in human achiasma. NeuroImage 164, 5966.Google Scholar
Olson, D.J., Oh, D., and Houston, D.W. (2015). The dynamics of plus end polarization and microtubule assembly during Xenopus cortical rotation. Dev. Biol. 401, 249263.Google Scholar
Olson, M.E. (2012). The developmental renaissance in adaptationism. Trends Ecol. Evol. 27, 278287.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.Google Scholar
Oppenheim, R.W. (1991). Cell death during development of the nervous system. Annu. Rev. Neurosci. 14, 453501.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, 920.Google Scholar
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, 433476.Google Scholar
Ostrander, E.A., Wayne, R.K., Freedman, A.H., and Davis, B.W. (2017). Demographic history, selection and functional diversity of the canine genome. Nat. Rev. Genet. 18, 705720.Google Scholar
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), 382.Google Scholar
Outters, P., Jaeger, S., Zaarour, N., and Ferrier, P. (2015). Long-range control of V(D)J recombination & allelic exclusion: modeling views. Adv. Immunol. 128, 363413.Google Scholar
Pagliara, V., Nasso, R., Ascione, A., Masullo, M., and Arcone, R. (2019). Myostatin and plasticity of skeletal muscle tissue. J. Hum. Sport Exercise 14(5proc), S1931S1937.Google Scholar
Palmer, A.R. (2005). Antisymmetry. In Variation: A Central Concept in Biology, Hallgrímsson, B. and Hall, B.K., editors. Elsevier Academic Press, New York, pp. 359397.Google Scholar
Palmer, A.R. and Strobeck, C. (1986). Fluctuating asymmetry: measurement, analysis, patterns. Annu. Rev. Ecol. Syst. 17, 391421.Google Scholar
Pan, Y., Liu, Z., Shen, J., and Kopan, R. (2005). Notch1 and 2 cooperate in limb ectoderm to receive an early Jagged2 signal regulating interdigital apoptosis. Dev. Biol. 286, 472482.Google Scholar
Pan, Y., Tsai, C.-J., Ma, B., and Nussinov, R. (2010). Mechanisms of transcription factor selectivity. Trends Genet. 26, 7583.Google Scholar
Pandya, P., Orgaz, J.L., and Sanz-Moreno, V. (2017). Actomyosin contractility and collective migration: may the force be with you. Curr. Opin. Cell Biol. 48, 8796.Google Scholar
Papert, S. (1980). Mindstorms: Children, Computers, and Powerful Ideas. Basic Books, New York.Google Scholar
Parchure, A., Vyas, N., and Mayor, S. (2018). Wnt and Hedgehog: secretion of lipid-modified morphogens. Trends Cell Biol. 28, 157170.Google Scholar
Park, K., Kang, J., Subedi, K.P., Ha, J.-H., and Park, C.S. (2008). Canine polydactyl mutations with heterogeneous origin in the conserved intronic sequence of LMBR1. Genetics 179, 21632172.Google Scholar
Parker, H.G., Harris, A., Dreger, D.L., Davis, B.W., and Ostrander, E.A. (2017). The bald and the beautiful: hairlessness in domestic dog breeds. Philos. Trans. R. Soc. Lond. B 372, 20150488.Google Scholar
Parker, H.G., Shearin, A.L., and Ostrander, E.A. (2010). Man’s best friend becomes biology’s Best in Show: genome analyses in the domestic dog. Annu. Rev. Genet. 44, 309336.Google Scholar
Parker, H.G., VonHoldt, B.M., Quignon, P., Margulies, E.H., Shao, S., Mosher, D.S., Spady, T.C., Elkahloun, A., Cargill, M., Jones, P.G., Maslen, C.L., Acland, G.M., Sutter, N.B., Kuroki, K., Bustamante, C.D., Wayne, R.K., and Ostrander, E.A. (2009). An expressed Fgf4 retrogene is associated with breed-defining chondrodysplasia in domestic dogs. Science 325, 995998.Google Scholar
Parks, A.L., Klueg, K.M., Stout, J.R., and Muskavitch, M.A.T. (2000). Ligand endocytosis drives receptor dissociation and activation in the Notch pathway. Development 127, 13731385.Google Scholar
Parks, H.B. (1936). Cleavage patterns in Drosophila and mosaic formation. Ann. Ent. Soc. Am. 29, 350392.Google Scholar
Pass, G. (2018). Beyond aerodynamics: the critical roles of the circulatory and tracheal systems in maintaining insect wing functionality. Arthropod Struct. Dev. 47, 391407.Google Scholar
Pastor-Pareja, J.C., Grawe, F., Martín-Blanco, E., and García-Bellido, A. (2004). Invasive cell behavior during Drosophila imaginal disc eversion is mediated by the JNK signaling cascade. Dev. Cell 7, 387399.Google Scholar
Patel, S.H., Camargo, F.D., and Yimlamai, D. (2017). Hippo signaling in the liver regulates organ size, cell fate, and carcinogenesis. Gastroenterology 152, 533545.Google Scholar
Patwari, P. and Lee, R.T. (2008). Mechanical control of tissue morphogenesis. Circ. Res. 103, 234243.Google Scholar
Pauciullo, A., Knorr, C., Perucatti, A., Iannuzzi, A., Iannuzzi, L., and Erhardt, G. (2016). Characterization of a very rare case of living ewe–buck hybrid using classical and molecular cytogenetics. Sci. Rep. 6, 34781.Google Scholar
Pavan, W.J. and Sturm, R.A. (2019). The genetics of human and hair pigmentation. Annu. Rev. Genom. Hum. Genet. 20, 4172.Google 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, 28552860.Google Scholar
Pearl, E.J., Li, J., and Green, J.B.A. (2017). Cellular systems for epithelial invagination. Philos. Trans. R. Soc. Lond. B 372, 20150526.Google Scholar
Pearson, H. (2007). The roots of accomplishment. Nature 446, 2021.Google Scholar
Pecze, L. (2018). A solution to the problem of proper segment positioning in the course of digit formation. BioSystems 173, 266272.Google Scholar
Pedersen, A.S., Berg, L.C., Almstrup, K., and Thomsen, P.D. (2014). A tortoiseshell male cat: chromosome analysis and histologic examination of the testis. Cytogenet. Genome Res. 142, 107111.Google Scholar
Pener, M.P. and Simpson, S.J. (2009). Locust phase polyphenism: an update. Adv. Insect Physiol. 36, 1272.Google Scholar
Peng, Y. and Axelrod, J.D. (2012). Asymmetric protein localization in planar cell polarity: mechanisms, puzzles, and challenges. Curr. Top. Dev. Biol. 101, 3353.Google Scholar
Pennisi, E. (2018). Buying time. Science 362, 988991.Google Scholar
Penzo-Méndez, A.I. and Stanger, B.Z. (2015). Organ-size regulation in mammals. Cold Spring Harb. Perspect. Biol. 7, a019240.Google Scholar
Pereira, G.L., de Matteis, R., Regitano, L.C.A., Chardulo, L.A.L., and Curi, R.A. (2016). MSTN, CKM, and DMRT3 gene variants in different lines of Quarter Horses. J. Equine Vet. Sci. 39, 3337.Google Scholar
Pérez-Gómez, R., Haro, E., Fernández-Guerrero, M., Bastida, M.F., and Ros, M.A. (2018). Role of Hox genes in regulating digit patterning. Int. J. Dev. Biol. 62, 797805.Google Scholar
Perrimon, N., Pitsouli, C., and Shilo, B.-Z. (2012). Signaling mechanisms controlling cell fate and embryonic patterning. Cold Spring Harb. Perspect. Biol. 4, a005975.Google Scholar
Peterson, T. and Müller, G.B. (2016). Phenotypic novelty in EvoDevo: the distinction between continuous and discontinuous variation and its importance in evolutionary theory. Evol. Biol. 43, 314335.Google Scholar
Petit, F., Sears, K.E., and Ahituv, N. (2017). Limb development: a paradigm of gene regulation. Nat. Rev. Genet. 18, 245258.Google Scholar
Petrij, F., van Veen, K., Mettler, M., and Brückmann, V. (2001). A second acromelanistic allelomorph at the albino locus of the Mongolian gerbil (Meriones unguiculatus). J. Hered. 92, 7478.Google Scholar
Pfennig, D.W. (1992). Polyphenism in spadefoot toad tadpoles as a locally adjusted evolutionary stable strategy. Evolution 46, 14081420.Google Scholar
Pfennig, D.W. (1999). Cannibalistic tadpoles that pose the greatest threat to kin are most likely to discriminate kin. Proc. R. Soc. Lond. B 266, 5762.Google Scholar
Pfennig, D.W. and Collins, J.P. (1993). Kinship affects morphogenesis in cannibalistic salamanders. Nature 362, 836838.Google Scholar
Pfister, K., Shook, D.R., Chang, C., Keller, R., and Skoglund, P. (2016). Molecular model for force production and transmission during vertebrate gastrulation. Development 143, 715727.Google Scholar
Phillips, R.G., Warner, N.L., and Whittle, J.R.S. (1999). Wingless signaling leads to an asymmetric response to Decapentaplegic-dependent signaling during sense organ patterning on the notum of Drosophila melanogaster. Dev. Biol. 207, 150162.Google Scholar
Plonka, P.M., Passeron, T., Brenner, M., Tobin, D.J., Shibahara, S., Thomas, A., Slominski, A., Kadekaro, A.L., Hershkovitz, D., Peters, E., Nordlund, J.J., Abdel-Malek, Z., Takeda, K., Paus, R., Ortonne, J.P., Hearing, V.J., and Schalleruter, K.U. (2009). What are melanocytes really doing all day long…? Exp. Dermatol. 18, 799819.Google Scholar
Plouffe, S.W., Hong, A.W., and Guan, K.-L. (2015). Disease implications of the Hippo/YAP pathway. Trends Mol. Med. 21, 212222.Google Scholar
Poodry, C.A. (1975). A temporal pattern in the development of sensory bristles in Drosophila. W. Roux Arch. Dev. Biol. 178, 203213.Google Scholar
Poodry, C.A. and Schneiderman, H.A. (1970). The ultrastructure of the developing leg of Drosophila melanogaster. W. Roux Arch. Dev. Biol. 166, 144.Google Scholar
Poodry, C.A., Hall, L., and Suzuki, D.T. (1973). Developmental properties of shibirets: a pleiotropic mutation affecting larval and adult locomotion and development. Dev. Biol. 32, 373386.Google Scholar
Portmann, A. (1967). Animal Forms and Patterns: A Study of the Appearance of Animals. Schocken Books, New York.Google Scholar
Posakony, J.W. (1994). Nature versus nurture: asymmetric cell divisions in Drosophila bristle development. Cell 76, 415418.Google Scholar
Price, J. and Allen, S. (2004). Exploring the mechanisms regulating regeneration of deer antlers. Philos. Trans. R. Soc. Lond. B 359, 809822.Google Scholar
Price, J.S., Oyajobi, B.O., Oreffo, R.O., and Russell, R.G. (1994). Cells cultured from the growing tip of red deer antler express alkaline phosphatase and proliferate in response to insulin-like growth factor-I. J. Endocrinol. 143, R9R16.Google Scholar
Prieur, D.S. and Rebsam, A. (2017). Retinal axon guidance at the midline: chiasmatic misrouting and consequences. Dev. Neurobiol. 77, 844860.Google Scholar
Pringle, J.W.S. (1948). The gyroscopic mechanism of the halteres of diptera. Philos. Trans. Roy. Soc. Lond. B 233, 347384.Google Scholar
Projecto-Garcia, J., Biddle, J.F., and Ragsdale, E.J. (2017). Decoding the architecture and origins of mechanisms for developmental polyphenism. Curr. Opin. Genet. Dev. 47, 18.Google Scholar
Protas, M.E. and Patel, N.H. (2008). Evolution of color patterns. Annu. Rev. Cell Dev. Biol. 24, 425446.Google Scholar
Pruvost, M., Bellone, R., Benecke, N., Sandoval-Castellanos, E., Cieslak, M., Kuznetsova, T., Morales-Muñiz, A., O’Connor, T., Reissmann, M., Hofreiter, M., and Ludwig, A. (2011). Genotypes of predomestic horses match phenotypes painted in Paleolithic works of cave art. PNAS 108, 1862618630.Google Scholar
Purcell, R. (1997). Special Cases: Natural Anomalies and Historical Monsters. Chronicle Books, San Francisco, CA.Google Scholar
Purves, D., Riddle, D.R., and LaMantia, A.-S. (1992). Iterated patterns of brain circuitry (or how the cortex gets its spots). Trends Neurosci. 15, 362368.Google Scholar
Quammen, D. (2014). People of the horse. Nat. Geogr. 225(3), 104126.Google Scholar
Rabah, S., Salati, S., and Wani, S. (2008). Mirror hand deformity: a rare congenital anomaly of the upper limb. Internet J. Surg. 21(1), 15.Google Scholar
Ramain, P., Khechumian, R., Khechumian, K., Arbogast, N., Ackermann, C., and Heitzler, P. (2000). Interactions between Chip and the Achaete/Scute-Daughterless heterodimers are required for Pannier-driven proneural patterning. Mol. Cell 6, 781790.Google Scholar
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. Pediatr. 136, 6268.Google Scholar
Randall, V.A. (2007). Hormonal regulation of hair follicles exhibits a biological paradox. Semin. Cell Dev. Biol. 18, 274285.Google Scholar
Raser, J.M. and O’Shea, E.K. (2005). Noise in gene expression: origins, consequences, and control. Science 309, 20102013.Google Scholar
Raspopovic, J., Marcon, L., Russo, L., and Sharpe, J. (2014). Digit patterning is controlled by a Bmp-Sox9-Wnt Turing network modulated by morphogen gradients. Science 345, 566570.Google Scholar
Rasskin-Gutman, D. and De Renzi, M., eds. (2009). Pere Alberch: The Creative Trajectory of an Evo-Devo Biologist. Universitat de València, València.Google Scholar
Rauskolb, C. (2001). The establishment of segmentation in the Drosophila leg. Development 128, 45114521.Google Scholar
Rauskolb, C. and Irvine, K.D. (1999). Notch-mediated segmentation and growth control of the Drosophila leg. Dev. Biol. 210, 339350.Google Scholar
Rebeiz, M., Miller, S.W., and Posakony, J.W. (2011). Notch regulates numb: integration of conditional and autonomous cell fate specification. Development 138, 215225.Google Scholar
Red-Horse, K. and Siekmann, A.F. (2019). Veins and arteries build hierarchical branching patterns differently: bottom-up versus top-down. BioEssays 41, e1800198.Google Scholar
Reddy, G.V. and Rodrigues, V. (1999). A glial cell arises from an additional division within the mechanosensory lineage during development of the microchaete on the Drosophila notum. Development 126, 46174622.Google Scholar
Reh, T.A. and Constantine-Paton, M. (1985). Eye-specific segregation requires neural activity in three-eyes Rana pipiens. J. Neurosci. 5, 11321143.Google Scholar
Reik, E.F. (1976). Four-winged diptera from the upper Permian of Australia. Proc. Linn. Soc. New South Wales 101(4), 250255.Google Scholar
Reilly, P.R. (2008). The Strongest Boy in the World: How Genetic Information Is Reshaping Our Lives. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.Google Scholar
Reiss, J.O., Burke, A.C., Archer, C., De Renzi, M., Dopazo, H., Etxeberría, A., Gale, E.A., Hinchliffe, J.R., de la Rosa Garcia, L.N., Rose, C.S., Rasskin-Gutman, D., and Müller, G.B. (2009). Pere Alberch: originator of EvoDevo. Biol. Theory 3, 351356.Google Scholar
Reiter, F., Wienerroither, S., and Stark, A. (2017). Combinatorial function of transcription factors and cofactors. Curr. Opin. Genet. Dev. 43, 7381.Google Scholar
Reno, P.L., Kjosness, K.M., and Hines, J.E. (2016). The role of Hox in pisiform and calcaneus growth plate formation and the nature of the zeugopod/autopod boundary. J. Exp. Zool. B Mol. Dev. Evol. 326, 303321.Google Scholar
Rensberger, B. (1998). Life Itself: Exploring the Realm of the Living Cell. Oxford University Press, New York.Google Scholar
Ressurreição, M., Warrington, S., and Strutt, D. (2018). Rapid disruption of Dishevelled activity uncovers an intercellular role in maintenance of Prickle in core planar polarity protein complexes. Cell Rep. 25, 14151424.Google Scholar
Restrepo, S., Zartman, J.J., and Basler, K. (2014). Coordination of patterning and growth by the morphogen DPP. Curr. Biol. 24, R245R255.Google Scholar
Rhyu, M.S., Jan, L.Y., and Jan, Y.N. (1994). Asymmetric distribution of numb protein during division of the sensory organ precursor cell confers distinct fates to daughter cells. Cell 76, 477491.Google Scholar
Rice, D.S., Goldowitz, D., Williams, R.W., Hamre, K., Johnson, P.T., Tan, S.-S., and Reese, B.E. (1999). Extrinsic modulation of reginal ganglion cell projections: analysis of the albino mutation in pigmentation mosaic mice. Dev. Biol. 216, 4156.Google Scholar
Rice, G.R., Barmina, O., Luecke, D., Hu, K., Arbeitman, M., and Kopp, A. (2019). Modular tissue-specific regulation of doublesex underpins sexually dimorphic development in Drosophila. Development 146, dev178285.Google Scholar
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., editors. Johns Hopkins University Press, Baltimore, MD, pp. 154170.Google Scholar
Richardson, J. and Simpson, P. (2006). A conserved trans-regulatory landscape for scute expression on the notum of cyclorraphous Diptera. Dev. Genes Evol. 216, 2938.Google 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. B Mol. Dev. Evol. 312, 639664.Google Scholar
Richelle, J. and Ghysen, A. (1979). Determination of sensory bristles and pattern formation in Drosophila. I. A model. Dev. Biol. 70, 418437.Google Scholar
Rieppel, O. (2001). Turtles as hopeful monsters. BioEssays 23, 987991.Google Scholar
Rimbault, M., Beale, H.C., Schoenebeck, J.J., Hoopes, B.C., Allen, J.J., Kilroy-Glynn, P., Wayne, R.K., Sutter, N.B., and Ostrander, E.A. (2013). Derived variants at six genes explain nearly half of size reduction in dog breeds. Genome Res. 23, 19851995.Google Scholar
Robinson, R. (1959). Genetic studies of the Syrian hamster. II. Partial albinism. Heredity 13, 165177.Google Scholar
Rodríguez-Carballo, E., Lopez-Delisle, L., Zhan, Y., Fabre, P.J., Beccari, L., El-Idrissi, I., Huynh, T.H.N., Ozadam, H., Dekker, J., and Duboule, D. (2017). The HoxD cluster is a dynamic and resilient TAD boundary controlling the segregation of antagonistic regulatory landscapes. Genes Dev. 31, 22642281.Google Scholar
Rodriguez-Estaban, C., Schwabe, J.W.R., de la Peña, J., Foys, B., Eshelman, B., and Izpisua Belmonte, J.C. (1997). Radical fringe positions the apical ectodermal ridge at the dorsoventral boundary of the vertebrate limb. Nature 386, 360366.Google Scholar
Rogers, G.E. (2004). Hair follicle differentiation and regulation. Int. J. Dev. Biol. 48, 163170.Google Scholar
Rolian, C. (2014). Genes, development, and evolvability in primate evolution. Evol. Anthrop. 23, 93104.Google Scholar
Rolian, C. (2020). Endochondral ossification and the evolution of limb proportions. Wiley Interdiscip. Rev. Dev. Biol. 2020, e373.Google Scholar
Rollo, C.D. (1995). Phenotypes: Their Epigenetics, Ecology and Evolution. Chapman & Hall, New York.Google Scholar
Romani, S., Campuzano, S., Macagno, E.R., and Modolell, J. (1989). Expression of achaete and scute genes in Drosophila imaginal discs and their function in sensory organ development. Genes Dev. 3, 9971007.Google Scholar
Romero, D.M., Bahi-Buisson, N., and Francis, F. (2018). Genetics and mechanisms leading to human cortical malformations. Semin. Cell Dev. Biol. 76, 3375.Google Scholar
Rongioletti, F., Merlo, G.R., Cinotti, E., Fausti, V., Cozzani, E., Cribier, B., Metze, D., Calonje, E., Kanitakis, J., Kempf, W., Stefanato, C.M., Marinho, E., and Parodi, A. (2013). Scleromyxedema: a multicenter study of characteristics, comorbidities, course, and therapy in 30 patients. J. Am. Acad. Dermatol. 69, 6672.Google Scholar
Rooney, M.F., Hill, E.W., Kelly, V.P., and Porter, R.K. (2018). The “speed gene” effect of myostatin arises in Thoroughbred horses due to a promoter proximal SINE insertion. PLoS ONE 13(10), e0205664.Google Scholar
Rørth, P. (2012). Fellow travellers: emergent properties of collective cell migration. EMBO Rep. 13(11), 984991.Google Scholar
Roselló-Díez, A., Arques, C.G., Delgado, I., Giovinazzo, G., and Torres, M. (2014). Diffusible signals and epigenetic timing cooperate in late proximo-distal limb patterning. Development 141, 15341543.Google Scholar
Rosenberger, A.L. and Preuschoft, H. (2012). Evolutionary morphology, cranial biomechanics and the origins of tarsiers and anthropoids. Palaeobiodivers. Palaeoenviron. 92, 507525.Google Scholar
Ross, C.M. (1969). Generalized folded skin with an underlying lipomatous nevus: “the Michelin Tire Baby”. Arch. Derm. 100, 320323.Google Scholar
Ross, C.M. (1972). Generalized folded skin with an underlying lipomatous nevus: the Michelin Tire Baby. Arch. Derm. 106, 766.Google Scholar
Roth, G. and Dicke, U. (2019). Origin and evolution of human cognition. Progr. Brain Res. 250, 285316.Google Scholar
Roy, S., Shashidhara, L.S., and VijayRaghavan, K. (1997). Muscles in the Drosophila second thoracic segment are patterned independently of autonomous homeotic gene function. Curr. Biol. 7, 222227.Google Scholar
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, 400411.Google Scholar
Rozowski, M. and Akam, M. (2002). Hox gene control of segment-specific bristle patterns in Drosophila. Genes Dev. 16, 11501162.Google Scholar
Rudel, D. and Sommer, R.J. (2003). The evolution of developmental mechanisms. Dev. Biol. 264, 1537.Google Scholar
Ruiz-Losada, M., Blom-Dahl, D., Córdoba, S., and Estella, C. (2018). Specification and patterning of Drosophila appendages. J. Dev. Biol. 6, jdb6030017.Google Scholar
Russell, L.B., ed. (1978). Genetic Mosaics and Chimeras in Mammals. Plenum, New York.Google Scholar
Russell, M.A. (1974). Pattern formation in the imaginal discs of a temperature-sensitive cell-lethal mutant of Drosophila melanogaster. Dev. Biol. 40, 2439.Google Scholar
Russell, M.A., Girton, J.R., and Morgan, K. (1977). Pattern formation in a ts-cell-lethal mutant of Drosophila: the range of phenotypes induced by larval heat treatments. W. Roux Arch. Dev. Biol. 183, 4159.Google Scholar
Rusting, R.L. (2001). Hair: why it grows, why it stops. Sci. Am. 284(6), 7079.Google Scholar
Ruvinsky, I. and Gibson-Brown, J.J. (2000). Genetic and developmental bases of serial homology in vertebrate limb evolution. Development 127, 52335244.Google Scholar
Sabarís, G., Laiker, I., Noon, E.P.-B., and Frankel, N. (2019). Actors with multiple roles: pleiotropic enhancers and the paradigm of enhancer modularity. Trends Genet. 35, 423433.Google Scholar
Saha, M. (1991). Spemann seen through a lens. In A Conceptual History of Modern Embryology, Gilbert, S.F., editor. Plenum, New York, pp. 91108.Google Scholar
Saito, K., Nomura, S., Yamamoto, S., Niiyama, R., and Okabe, Y. (2017). Investigation of hindwing folding in ladybird beetles by artificial elytron transplantation and microcomputed tomography. PNAS 114, 56245628.Google Scholar
Saito, K., Yamamoto, S., Maruyama, M., and Okabe, Y. (2014). Asymmetric hindwing foldings in rove beetles. PNAS 111, 1634916352.Google Scholar
Saiz-Lopez, P., Chinnaiya, K., Campa, V.M., Delgado, I., Ros, M.A., and Towers, M. (2015). An intrinsic timer specifies distal structures of the vertebrate limb. Nat. Commun. 6, 8108.Google Scholar
Sakai, M. (2008). Cell-autonomous and inductive processes among three embryonic domains control dorsal–ventral and anterior–posterior development of Xenopus laevis. Dev. Growth Differ. 50, 4962.Google Scholar
Salazar-Ciudad, I., Jernvall, J., and Newman, S.A. (2003). Mechanisms of pattern formation in development and evolution. Development 130, 20272037.Google Scholar
Salser, S.J. and Kenyon, C. (1996). A C. elegans Hox gene switches on, off, on and off again to regulate proliferation, differentiation and morphogenesis. Development 122, 16511661.Google Scholar
Samlaska, C.P., James, W.D., and Sperling, L.C. (1989). Scalp whorls. J. Am. Acad. Dermatol. 21, 553556.Google Scholar
Sánchez-Villagra, M.R. and Menke, P.R. (2005). The mole’s thumb: evolution of the hand skeleton in talpids (Mammalia). Zoology 108, 312.Google Scholar
Sander, K. and Faessler, P.E. (2001). Introducing the Spemann–Mangold organizer: experiments and insights that generated a key concept in developmental biology. Int. J. Dev. Biol. 45, 111.Google Scholar
Sanicola, M., Sekelsky, J., Elson, S., and Gelbart, W.M. (1995). Drawing a stripe in Drosophila imaginal disks: negative regulation of decapentaplegic and patched expression by engrailed. Genetics 139, 745756.Google Scholar
Santamaría, P. (1979). Heat shock induced phenocopies of dominant mutants of the Bithorax Complex in Drosophila melanogaster. Mol. Gen. Genet. 172, 161163.Google Scholar
Santana, S.E., Alfaro, J.L., and Alfaro, M.E. (2012). Adaptive evolution of facial colour patterns in Neotropical primates. Proc. R. Soc. B 279, 22042211.Google Scholar
Sarin, K.Y. and Artandi, S.E. (2007). Aging, graying and loss of melanocyte stem cells. Stem Cell Rev. 3, 212217.Google Scholar
Sarnat, H.B. and Netsky, M.G. (1981). Evolution of the Nervous System. Oxford University Press, New York.Google Scholar
Sawada, R., Aramaki, T., and Kondo, S. (2018). Flexibility of pigment cell behavior permits the robustness of skin pattern formation. Genes Cells 23, 537545.Google Scholar
Saxena, A., Towers, M., and Cooper, K.L. (2017). The origins, scaling and loss of tetrapod digits. Philos. Trans. R. Soc. Lond. B 372, 20150482.Google Scholar
Scheibert, J., Leurent, S., Prevost, A., and Debrégeas, G. (2009). The role of fingerprints in the coding of tactile information probed with a biomimetic sensor. Science 323, 15031506.Google Scholar
Schmidt-Küntzel, A., Eizirik, E., O’Brien, S.J., and Menotti-Raymond, M. (2005). Tyrosinase and Tyrosinase Related Protein 1 alleles specify domestic cat coat color phenotypes of the albino and brown loci. J. Hered. 96, 289301.Google Scholar
Schmidt-Küntzel, A., Nelson, G., David, V.A., Schäffer, A.A., Eizirik, E., Roelke, M.E., Kehler, J.S., Hannah, S.S., O’Brien, S.J., and Menotti-Raymond, M. (2009). A domestic cat X chromosome linkage map and the sex-linked orange locus: mapping of orange, multiple origins and epistasis over nonagouti. Genetics 181, 14151425.Google Scholar
Schmidt-Ullrich, R. and Paus, R. (2005). Molecular principles of hair follicle induction and morphogenesis. BioEssays 27, 247261.Google Scholar
Schneider, I., Kreis, J., Schweickert, A., Blum, M., and Vick, P. (2019). A dual function of FGF signaling in Xenopus left–right axis formation. Development 146, dev173575.Google Scholar
Schneider, M.R., Schmidt-Ullrich, R., and Paus, R. (2009). The hair follicle as a dynamic miniorgan. Curr. Biol. 19, R132R142.Google Scholar
Schöck, F. and Perrimon, N. (2002). Molecular mechanisms of epithelial morphogenesis. Annu. Rev. Cell Dev. Biol. 18, 463493.Google Scholar
Schoenebeck, J.J. and Ostrander, E.A. (2014). Insights into morphology and disease from the dog genome project. Annu. Rev. Cell Dev. Biol. 30, 535560.Google Scholar
Schoenebeck, J.J., Hutchinson, S.A., Byers, A., Beale, H.C., Carrington, B., Faden, D.L., Rimbault, M., Decker, B., Kidd, J.M., Sood, R., Boyko, A.R., Fondon, J.W. III, Wayne, R.K., Bustamante, C.D., Ciruna, B., and Ostrander, E.A. (2012). Variation of BMP3 contributes to dog breed skull diversity. PLoS Genet. 8(8), e1002849.Google Scholar
Schreiber, A.M., Cai, L., and Brown, D.D. (2005). Remodeling of the intestine during metamorphosis of Xenopus laevis. PNAS 102, 37203725.Google Scholar
Schroeder, T.B.H., Houghtaling, J., Wilts, B.D., and Mayer, M. (2018). It’s not a bug, it’s a feature: functional materials in insects. Adv. Mater. 30, 1705322.Google Scholar
Schubiger, G., Schubiger, M., and Sustar, A. (2012). The three leg imaginal discs of Drosophila: “vive la différence”. Dev. Biol. 369, 7690.Google Scholar
Schuelke, M., Wagner, K.R., Stolz, L.E., Hübner, C., Riebel, T., Kömen, W., Braun, T., Tobin, J.F., and Lee, S.-J. (2004). Myostatin mutation associated with gross muscle hypertrophy in a child. N. Engl. J. Med. 350, 26822688.Google Scholar
Schwarzer, W. and Spitz, F. (2014). The architecture of gene expression: integrating dispersed cis-regulatory modules into coherent regulatory domains. Curr. Opin. Genet. Dev. 27, 7482.Google Scholar
Schweisguth, F. and Corson, F. (2019). Self-organization in pattern formation. Dev. Cell 49, 659677.Google Scholar
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, 14651475.Google Scholar
Sengpiel, F. (2008). Binocular vision: only half a brain needed. Curr. Biol. 18, R1054R1056.Google Scholar
Sessions, S.K. and Ruth, S.B. (1990). Explanation for naturally occurring supernumerary limbs in amphibians. J. Exp. Zool. 254, 3847.Google Scholar
Sessions, S.K., Franssen, R.A., and Horner, V.L. (1999). Morphological clues from multilegged frogs: are retinoids to blame? Science 284, 800802.Google Scholar
Shao, X., Ding, Z., Zhao, M., Liu, K., Sun, H., Chen, J., Liu, X., Zhang, Y., Hong, Y., Li, H., and Li, H. (2017). Mammalian Numb protein antagonizes Notch by controlling postendocytic trafficking of the Notch ligand Delta-like 4. J. Biol. Chem. 292, 2062820643.Google Scholar
Sharma, M., Castro-Piedras, I., Simmons, G.E. Jr., and Pruitt, K. (2018). Dishevelled: a masterful conductor of complex Wnt signals. Cell. Signal. 47, 5264.Google 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, 491502.Google Scholar
Shellenbarger, D.L. and Mohler, J.D. (1978). Temperature-sensitive periods and autonomy of pleiotropic effects of l(1)Nts1, a conditional Notch lethal in Drosophila. Dev. Biol. 62, 432446.Google Scholar
Shelton, P.M.J., Truby, P.R., and Shelton, R.G.J. (1981). Naturally occurring abnormalities (Bruchdreifachbildungen) in the chelae of three species of Crustacea (Decapoda) and a possible explanation. J. Embryol. Exp. Morphol. 63, 285304.Google Scholar
Sherwood, C.C. and Gómez-Robles, A. (2017). Brain plasticity and human evolution. Annu. Rev. Anthrop. 46, 399419.Google Scholar
Sherwood, C.C., Bauernfeind, R., Bianchi, S., Raghanti, M.A., and Hof, P.R. (2012). Human brain evolution writ large and small. Progr. Brain Res. 195, 237–254.Google Scholar
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, 14761480.Google Scholar
Shi, Y., Barton, K., De Maria, A., Petrash, J.M., Shiels, A., and Bassnett, S. (2009). The stratified syncytium of the vertebrate lens. J. Cell Sci. 122, 16071615.Google Scholar
Shi, Y., De Maria, A., Lubura, S., Sikic, H., and Bassnett, S. (2015). The penny pusher: a cellular model of lens growth. Invest. Ophthalmol. Vis. Sci. 56, 799809.Google Scholar
Shimizu-Nishikawa, K., Takahashi, J., and Nishikawa, A. (2003). Intercalary and supernumerary regeneration in the limbs of the frog, Xenopus laevis. Dev. Dynamics 227, 563572.Google Scholar
Shimomura, Y., Agalliu, D., Vonica, A., Luria, V., Wajid, M., Baumer, A., Belli, S., Petukhova, L., Schinzel, A., Brivanlou, A.H., Barres, B.A., and Christiano, A.M. (2010). APCDD1 is a novel Wnt inhibitor mutated in hereditary hypotrichosis simplex. Nature 464, 10431047.Google Scholar
Shindo, A. (2018). Models of convergent extension during morphogenesis. Wiley Interdiscip. Rev. Dev. Biol. 7, e293.Google Scholar
Shindo, A., Inoue, Y., Kinoshita, M., and Wallingford, J.B. (2019). PCP-dependent transcellular regulation of actomyosin oscillation facilitates convergent extension of vertebrate tissue. Dev. Biol. 446, 159167.Google Scholar
Shingleton, A.W. and Frankino, W.A. (2013). New perspectives on the evolution of exaggerated traits. BioEssays 35, 100107.Google Scholar
Shingleton, A.W. and Frankino, W.A. (2018). The (ongoing) problem of relative growth. Curr. Opin. Insect Sci. 25, 919.Google Scholar
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, 536548.Google Scholar
Shiota, K., Yamada, S., Komada, M., and Ishibashi, M. (2007). Embryogenesis of holoprosencephaly. Am. J. Med. Genet. A 143A, 30793087.Google Scholar
Shirai, T., Yorimitsu, T., Kiritooshi, N., Matsuzaki, F., and Nakagoshi, H. (2007). Notch signaling relieves the joint-suppressive activity of Defective proventriculus in the Drosophila leg. Dev. Biol. 312, 147156.Google Scholar
Shoji, H. and Iwasa, Y. (2005). Labyrinthine versus stright-striped patterns generated by two-dimensional Turing systems. J. Theor. Biol. 237, 104116.Google Scholar
Shoji, H., Iwasa, Y., and Kondo, S. (2003). Stripes, spots, or reversed spots in two-dimensional Turing systems. J. Theor. Biol. 224, 339350.Google Scholar
Sholtis, S.J. and Noonan, J.P. (2010). Gene regulation and the origins of human biological uniqueness. Trends Genet. 26, 110118.Google Scholar
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, 374381.Google Scholar
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. Mech. Dev. 124, 4358.Google Scholar
Sick, S., Reinker, S., Timmer, J., and Schlake, T. (2006). WNT and DKK determine hair follicle spacing through a reaction–diffusion mechanism. Science 314, 14471450.Google Scholar
Siebel, C. and Lendahl, U. (2017). Notch signaling in development, tissue homeostasis, and disease. Physiol. Rev. 97, 12351294.Google Scholar
Simpson, P. (1997). Notch signalling in development: on equivalence groups and asymmetric developmental potential. Curr. Op. Genet. Dev. 7, 537542.Google Scholar
Simpson, P. (2007). The stars and stripes of animal bodies: evolution of regulatory elements mediating pigment and bristle patterns in Drosophila. Trends Genet. 23, 350358.Google Scholar
Simpson, P. and Marcellini, S. (2006). The origin and evolution of stereotyped patterns of macrochaetes on the nota of cyclorraphous Diptera. Heredity 97, 148156.Google Scholar
Sinclair, R. (1998). Male pattern androgenetic alopecia. BMJ 317, 865869.Google Scholar
Singh, A., Gupta, R., Zaidi, S.H.H., and Singh, A. (2016). Dermatoglyphics: a brief review. Int. J. Adv. Integr. Med. Sci. 1, 111115.Google Scholar
Singh, A., Tare, M., Puli, O.R., and Kango-Singh, M. (2011). A glimpse into dorso-ventral patterning of the Drosophila eye. Dev. Dynamics 241, 6984.Google Scholar
Sjöqvist, M. and Andersson, E.R. (2019). Do as I say, Not(ch) as I do: lateral control of cell fate. Dev. Biol. 447, 5870.Google Scholar
Skaer, N. and Simpson, P. (2000). Genetic analysis of bristle loss in hybrids between Drosophila melanogaster and D. simulans provides evidence for divergence of cis-regulatory sequences in the achaete-scute gene complex. Dev. Biol. 221, 148167.Google Scholar
Skaer, N., Pistillo, D., and Simpson, P. (2002). Transcriptional heterochrony of scute and changes in bristle pattern between two closely related species of blowfly. Dev. Biol. 252, 3145.Google Scholar
Skeath, J.B. and Carroll, S.B. (1991). Regulation of achaete-scute gene expression and sensory organ pattern formation in the Drosophila wing. Genes Dev. 5, 984995.Google Scholar
Skoglund, P. and Keller, R. (2010). Integration of planar cell polarity and ECM signaling in elongation of the vertebrate body plan. Curr. Opin. Cell Biol. 22, 589596.Google Scholar
Skulachev, V.P., Holtze, S., Vyssokikh, M.Y., Bakeeva, L.E., Skulachev, M.V., Markov, A.V., Hildebrandt, T.B., and Sandovnichii, V.A. (2017). Neoteny, prolongation of youth: from naked mole rats to “naked apes” (humans). Physiol. Rev. 97, 699720.Google Scholar
Smaers, J.B., Gómez-Robles, A., Parks, A.N., and Sherwood, C.C. (2017). Exceptional evolutionary expansion of prefrontal cortex in great apes and humans. Curr. Biol. 27, 714720.Google Scholar
Smaers, J.B., Mongle, C.S., Safi, K., and Dechmann, D.K.N. (2019). Allometry, evolution and development of neocortex size in mammals. Progr. Brain Res. 250, 83107.Google Scholar
Smith, D.J., Montenegro-Johnson, T.D., and Lopes, S.S. (2019). Symmetry-breaking cilia-driven flow in embryogenesis. Ann. Rev. Fluid Mech. 51, 105128.Google Scholar
Smith, K.K. (2003). Time’s arrow: heterochrony and the evolution of development. Int. J. Dev. Biol. 47, 613621.Google Scholar
Smith-Bolton, R.K., Worley, M.I., Kanda, H., and Hariharan, I.K. (2009). Regenerative growth in Drosophila imaginal discs is regulated by Wingless and Myc. Dev. Cell 16, 797809.Google Scholar
Soder, A.I., Hoare, S.F., Muire, S., Balmain, A., Parkinson, E.K., and Keith, W.N. (1997). Mapping of the gene for the mouse telomerase RNA component, Terc, to chromosome 3 by fluorescence in situ hybridization and mouse chromosome painting. Genomics 41, 293294.Google Scholar
Soler, C., Daczewska, M., Da Ponte, J.P., Dastugue, B., and Jagla, K. (2004). Coordinated development of muscles and tendons of the Drosophila leg. Development 131, 60416051.Google Scholar
Solnica-Krezel, L. and Sepich, D.S. (2012). Gastrulation: making and shaping germ layers. Annu. Rev. Cell Dev. Biol. 28, 687717.Google Scholar
Sommer, R.J. (2020). Phenotypic plasticity: from theory and genetics to current and future challenges. Genetics 215, 113.Google Scholar
Souder, W. (2000). A Plague of Frogs: Unraveling an Environmental Mystery. University of Minnesota Press, Minneapolis, MN.Google Scholar
Soukup, V., Horácek, I., and Cerny, R. (2013). Development and evolution of the vertebrate primary mouth. J. Anat. 222, 7999.Google Scholar
Sousa, A.M.M., Meyer, K.A., Santpere, G., Gulden, F.O., and Sestan, N. (2017). Evolution of the human nervous system function, structure, and development. Cell 170, 226247.Google Scholar
Spéder, P. and Noselli, S. (2007). Left–right asymmetry: class I myosins show the direction. Curr. Opin. Cell Biol. 19, 8287.Google Scholar
Spemann, H. (1938). Embryonic Development and Induction. Yale University Press, New Haven, CT.Google Scholar
Spemann, H. and Falkenberg, H. (1919). Über asymmetrische Entwicklung und Situs inversus viscerum bei Zwillingen und Doppelbildungen. Archiv. für Entwicklungsmechanik 45, 371422.Google Scholar
Spencer, R. (2000). Craniopagus conjoined twins: typical, parasitic, and intracranial fetus-in-fetu. Neurosurg. Quart. 10, 6079.Google Scholar
Spencer, R. (2000). Theoretical and analytical embryology of conjoined twins. Part I. Embryogenesis. Clin. Anat. 13, 3653.Google Scholar
Spencer, R. (2000). Theoretical and analytical embryology of conjoined twins. Part II. Adjustments to union. Clin. Anat. 13, 97120.Google Scholar
Spencer, R. (2003). Conjoined Twins: Developmental Malformations and Clinical Implications. Johns Hopkins University Press, Baltimore, MD.Google Scholar
Sponenberg, D.P., Carr, G., Simak, E., and Schwink, K. (1990). The inheritance of the leopard complex of spotting patterns in horses. J. Hered. 81, 323331.Google Scholar
St. Johnston, D. (2015). The renaissance of developmental biology. PLoS Biol. 13, e1002149.Google Scholar
Stahl, A.L., Charlton-Perkins, M., Buschbeck, E.K., and Cook, T.A. (2017). The cuticular nature of corneal lenses in Drosophila melanogaster. Dev. Genes Evol. 227, 271278.Google Scholar
Stanford, P.K. (2005). August Weismann’s theory of the germ-plasm and the problem of unconceived alternatives. Hist. Philos. Life Sci. 27, 163199.Google Scholar
Steed, L. (2020). Kitty scary: wrinkly sphynx cat has stolen hearts thanks to his unique appearance. The Sun (9 March 2020).Google Scholar
Steinberg, M.S. (2007). Differential adhesion in morphogenesis: a modern view. Curr. Opin. Genet. Dev. 17, 281286.Google Scholar
Steiner, E. (1976). Establishment of compartments in the developing leg imaginal discs of Drosophila melanogaster. W. Roux Arch. Dev. Biol. 180, 930.Google Scholar
Stent, G.S. (1978). Paradoxes of Progress. W. H. Freeman, San Francisco, CA.Google Scholar
Stern, C. (1954). Genes and developmental patterns. Caryologia (suppl: Proc. 9th Int. Congr. Genet. Part I), 355–369.Google Scholar
Stern, C. (1954). Two or three bristles. Am. Sci. 42, 213247.Google Scholar
Stern, C. (1968). Genetic Mosaics and Other Essays. Harvard University Press, Cambridge, MA.Google Scholar
Stern, C. (1969). Richard Benedict Goldschmidt: April 12, 1878 – April 24, 1958. Persp. Biol. Med. 12, 178203.Google Scholar
Stern, C.D. (2019). The ’omics revolution: how an obsession with compiling lists is threatening the ancient art of experimental design. BioEssays 41, 1900168.Google Scholar
Stern, D. (2006). Morphing into shape. Science 313, 5051.Google Scholar
Sternberg, P.W. (2019). Ablating the fixed lineage conjecture: commentary on Kimble 1981. Dev. Biol. 446, 116.Google Scholar
Stevens, J.L., Edgerton, V.R., and Mitton, S. (1971). Gross anatomy of the hindlimb skeletal system of the Galago senegalensis. Primates 12(3–4), 313321.Google Scholar
Stevenson, R.D., Hill, M.F., and Bryant, P.J. (1995). Organ and cell allometry in Hawaiian Drosophila: how to make a big fly. Proc. Roy. Soc. Lond. B 259, 105110.Google Scholar
Stinckens, A., Luyten, T., Bijttebier, J., Van den Maagdenberg, K., Dieltines, D., Janssens, S., De Smet, S., Georges, M., and Buys, N. (2008). Characterization of the complete porcine MSTN gene and expression levels in pig breeds differing in muscularity. Anim. Genet. 39, 586596.Google Scholar
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, 423433.Google Scholar
Stockard, C.R. (1941). The Genetic and Endocrinic Basis for Differences in Form and Behavior. American Anatomical Memoirs, No. 19. Wistar Institute of Anatomy and Biology, Philadelphia, PA.Google Scholar
Stocum, D.L. (2000). Frog limb deformities: an “eco-devo” riddle wrapped in multiple hypotheses surrounded by insufficient data. Teratology 62, 147150.Google Scholar
Stollewerk, A., Schoppmeier, M., and Damen, W.G.M. (2003). Involvement of Notch and Delta genes in spider segmentation. Nature 423, 863865.Google Scholar
Stone, J.L. and Goodrich, J.T. (2006). The craniopagus malformation: classification and implications for surgical separation. Brain 129, 10841095.Google Scholar
Stopper, G.F., Hecker, L., Franssen, R.A., and Sessions, S.K. (2002). How trematodes cause limb deformities in amphibians. J. Exp. Zool. 294, 252263.Google Scholar
Striedter, G., Srinivasan, S., and Monuki, E.S. (2015). Cortical folding: when, where, how, and why? Annu. Rev. Neurosci. 38, 291307.Google Scholar
Struhl, G. (1984). Splitting the bithorax complex of Drosophila. Nature 308, 454457.Google Scholar
Struhl, G. and Basler, K. (1993). Organizing activity of wingless protein in Drosophila. Cell 72, 527540.Google Scholar
Strutt, D. (2002). The asymmetric subcellular localisation of components of the planar polarity pathway. Semin. Cell Dev. Biol. 13, 225231.Google Scholar
Strutt, D. (2009). Gradients and the specification of planar polarity in the insect cuticle. Cold Spring Harb. Perspect. Biol. 1, a000489.Google Scholar
Strutt, D., Johnson, R., Cooper, K., and Bray, S. (2002). Asymmetric localization of Frizzled and the determination of Notch-dependent cell fate in the Drosophila eye. Curr. Biol. 12, 813824.Google Scholar
Sturtevant, A.H. (1913). The Himalayan rabbit case, with some considerations on multiple allelomorphs. Am. Nat. 47, 234238.Google Scholar
Sturtevant, A.H. (1929). The claret mutant type of Drosophila simulans: a study of chromosome elimination and of cell-lineage. Z. wiss. Zool. 135, 323356.Google Scholar
Sturtevant, A.H. (1970). Studies on the bristle pattern of Drosophila. Dev. Biol. 21, 4861.Google Scholar
Sturtevant, A.H. (2001). Reminiscences of T. H. Morgan. Genetics 159, 15.Google Scholar
Suchy, F. and Nakauchi, H. (2017). Lessons from interspecies mammalian chimeras. Annu. Rev. Cell Dev. Biol. 33, 203217.Google Scholar
Suchy, F. and Nakauchi, H. (2018). Interspecies chimeras. Curr. Opin. Genet. Dev. 52, 3641.Google Scholar
Sui, L., Alt, S., Weigert, M., Dye, N., Eaton, S., Jug, F., Myers, E.W., Jülicher, F., Salbreux, G., and Dahmann, C. (2018). Differential lateral and basal tension drive folding of Drosophila wing discs through two distinct mechanisms. Nat. Commun. 9, 4620.Google Scholar
Suijkerbuijk, S.J.E., van Osch, M.H.J., Bos, F.L., Hanks, S., Rahman, N., and Kops, G.J.P.L. (2010). Molecular causes for BUBR1 dysfunction in the human cancer predisposition syndrome mosaic variegated aneuploidy. Cancer Res. 70, 48914900.Google Scholar
Summerbell, D., Lewis, J.H., and Wolpert, L. (1973). Positional information in chick limb morphogenesis. Nature 244, 492–496.Google Scholar
Sun, B., Tu, J., Liang, Q., Cheng, X., Fan, X., Li, Y., Wallbank, R.W.R., and Yang, M. (2019). Expression of mammalian ASH1 and ASH4 in Drosophila reveals opposing functional roles in neurogenesis. Gene 688, 132139.Google Scholar
Sun, G. and Irvine, K.D. (2014). Control of growth during regeneration. Curr. Top. Dev. Biol. 108, 95120.Google Scholar
Sun, J., Ling, M., Wu, W., Bhushan, B., and Tong, J. (2014). The hydraulic mechanism of the unfolding of hind wings in Dorcus titanus platymelus (Order: Coleoptera). Int. J. Mol. Sci. 15, 60096018.Google Scholar
Sun, M., Li, N., Dong, W., Chen, Z., Liu, Q., Xu, Y., He, G., Shi, Y., Li, X., Hao, J., Luo, Y., Shang, D., Lv, D., Ma, F., Zhang, D., Hua, R., Lu, C., Wen, Y., Cao, L., Irvine, A.D., McLean, W.H.I., Dong, Q., Wang, M.-R., Yu, J., He, L., Lo, W.H.Y., and Zhang, X. (2009). Copy-number mutations on chromosome 17q24.2–q24.3 in congenital generalized hypertrichosis terminalis with or without gingival hyperplasia. Am. J. Hum. Genet. 84, 807813.Google Scholar
Sun, T. and Hevner, R.F. (2014). Growth and folding of the mammalian cerebral cortex: from molecules to malformations. Nat. Rev. Neurosci. 15, 217232.Google Scholar
Sundaram, M.V. and Cohen, J.D. (2017). Time to make the doughnuts: building and shaping seamless tubes. Semin. Cell Dev. Biol. 67, 123131.Google Scholar
Sutter, N.B., Bustamante, C.D., Chase, K., Gray, M.M., Zhao, K., Zhu, L., Padhukasahasram, B., Karlins, E., Davis, S., Jones, P.G., Quignon, P., Johnson, G.S., Parker, H.G., Fretwell, N., Mosher, D.S., Lawler, D.F., Satyaraj, E., Nordborg, M., Lark, K.G., Wayne, R.K., and Ostrander, E.A. (2007). A single IGF1 allele is a major determinant of small size in dogs. Science 316, 112115.Google Scholar
Suttie, J.M., Gluckman, P.D., Butler, J.H., Fennessy, P.F., Corson, I.D., and Laas, F.J. (1985). Insulin-like growth factor 1 (IGF-1) antler-stimulating hormone? Endocrinology 116, 846848.Google Scholar
Suzanne, M. (2016). Molecular and cellular mechanisms involved in leg joint morphogenesis. Semin. Cell Dev. Biol. 55, 131138.Google Scholar
Suzuki, D.G. (2016). Two-headed mutants of the lamprey, a basal vertebrate. Zoological Lett. 2, 22.Google Scholar
Suzuki, D.T. (1970). Temperature-sensitive mutations in Drosophila melanogaster. Science 170, 695706.Google Scholar
Swett, F.H. (1926). On the production of double limbs in amphibians. J. Exp. Zool. 44, 419473.Google Scholar
Symmons, O., Pan, L., Remeseiro, S., Aktas, T., Klein, F., Huber, W., and Spitz, F. (2016). The Shh topological domain facilitates the action of remote enhancers by reducing the effects of genomic distances. Dev. Cell 39, 529543.Google Scholar
Szabad, J., Schüpbach, T., and Wieschaus, E. (1979). Cell lineage and development in the larval epidermis of Drosophila melanogaster. Dev. Biol. 73, 256271.Google Scholar
Szabo, K.T. (1989). Congenital Malformations in Laboratory and Farm Animals. Academic Press, New York.Google Scholar
Szebenyi, A.L. (1969). Cleaning behaviour in Drosophila melanogaster. Anim. Behav. 17, 641651.Google Scholar
Szenker-Ravi, E., Altunoglu, U., Leushacke, M., Bosso-Lefèvre, C., Khatoo, M., Tran, H.T., Naert, T., Noelanders, R., Hajamohideen, A., Beneteau, C., de Sousa, S.B., Karaman, B., Latypova, X., Başaran, S., Yücel, E.B., Tan, T.T., Vlaminck, L., Nayak, S.S., Shukla, A., Girisha, K.M., Le Caignec, C., Soshnikova, N., Uyguner, Z.O., Vleminckx, K., Barker, N., Kayserili, H., and Reversade, B. (2018). RSPO2 inhibition of RNF43 and ZNRF3 governs limb development independently of LGR4/5/6. Nature 557, 564569.Google Scholar
Tabata, T., Schwartz, C., Gustavson, E., Ali, Z., and Kornberg, T.B. (1995). Creating a Drosophila wing de novo, the role of engrailed, and the compartment border hypothesis. Development 121, 33593369.Google Scholar
Taber, L.A. (2014). Morphometrics: transforming tubes into organs. Curr. Opin. Genet. Dev. 27, 713.Google Scholar
Tabin, C. and Laufer, E. (1993). Hox genes and serial homology. Nature 361, 692693.Google Scholar
Tabin, C.J. (1992). Why we have (only) five fingers per hand: Hox genes and the evolution of paired limbs. Development 116, 289296.Google Scholar
Tadin-Strapps, M., Salas-Alanis, J.C., Moreno, L., Warburton, D., Martinez-Mir, A., and Christiano, A.M. (2003). Congenital universal hypertrichosis with deafness and dental anomalies inherited as an X-linked trait. Clin. Genet. 63, 418422.Google Scholar
Tajiri, R. (2017). Cuticle itself as a central and dynamic player in shaping cuticle. Curr. Opin. Insect Sci. 19, 3035.Google Scholar
Tajiri, R., Misaki, K., Yonemura, S., and Hayashi, S. (2010). Dynamic shape changes of ECM-producing cells drive morphogenesis of ball-and-socket joints in the fly leg. Development 137, 20552063.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, 46214626.Google Scholar
Takahashi, H., Abe, M., and Kuroda, R. (2019). GSK3β controls the timing and pattern of the fifth spiral cleavage at the 2–4 cell stage in Lymnaea stagnalis. Dev. Genes Evol. 229, 7381.Google Scholar
Takechi, M., Adachi, N., Hirai, T., Kuratani, S., and Kuraku, S. (2013). The Dlx genes as clues to vertebrate genomics and craniofacial evolution. Semin. Cell Dev. Biol. 24, 110118.Google Scholar
Tallinen, T., Chung, J.Y., Biggins, J.S., and Mahadevan, L. (2014). Gyrification from constrained cortical expansion. PNAS 111, 1266712672.Google Scholar
Tallinen, T., Chung, J.Y., Rousseau, F., Girard, N., Lefèvre, J., and Mahadevan, L. (2016). On the growth and form of cortical convolutions. Nat. Phys. 12, 588593.Google Scholar
Tanaka, E.M. (2016). The molecular and cellular choreography of appendage regeneration. Cell 165, 15981608.Google Scholar
Tanaka, K., Barmina, O., and Kopp, A. (2009). Distinct developmental mechanisms underlie the evolutionary diversification of Drosophila sex combs. PNAS 106, 47644769.Google Scholar
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.Google Scholar
Tanaka, Y., Okada, Y., and Hirokawa, N. (2005). FGF-induced vesicular release of Sonic hedgehog and retinoic acid in leftward nodal flow is critical for left–right determination. Nature 435, 172177.Google Scholar
Tao, X., Chen, X., and Tian, J. (2012). Fingerprint recognition with identical twin fingerprints. PLoS ONE 7(4), e35704.Google Scholar
Tapaltsyan, V., Charles, C., Hu, J., Mindell, D., Ahituv, N., Wilson, G.M., Black, B.L., Viriot, L., and Klein, O.D. (2016). Identification of novel Fgf enhancers and their role in dental evolution. Evol. Dev. 18, 3140.Google Scholar
Tapon, N., Harvey, K.F., Bell, D.W., Wahrer, D.C.R., Schiripo, T.A., Haber, D.A., and Hariharan, I.K. (2002). salvador promotes both cell cycle exit and apoptosis in Drosophila and is mutated in human cancer cell lines. Cell 110, 467478.Google Scholar
Tate, E. (1992). Professor marvels at mutant toad. Eyes inside its mouth upside down to boot. The Hamilton Spectator Metro Section (Sept. 4, 1992), B6.Google Scholar
Taub, R. (2004). Liver regeneration: from myth to mechanism. Nat. Rev. Mol. Cell Biol. 5, 836847.Google Scholar
Tautz, D. (1996). Selector genes, polymorphisms, and evolution. Science 271, 160161.Google Scholar
Taylor, G.K. (2001). Mechanics and aerodynamics of insect flight control. Biol. Rev. 76, 449471.Google Scholar
Taylor, J. and Adler, P.N. (2008). Cell rearrangement and cell division during the tissue level morphogenesis of evaginating Drosophila imaginal discs. Dev. Biol. 313, 739751.Google Scholar
te Welscher, P., Zuniga, A., Kuijper, S., Drenth, T., Goedemans, H.J., Meijlink, F., and Zeller, R. (2002). Progression of vertebrate limb development through SHH-mediated counteraction of Gli3. Science 298, 827830.Google Scholar
Theisen, H., Syed, A., Nguyen, B.T., Lukacsovich, T., Purcell, J., Srivastava, G.P., Iron, D., Gaudenz, K., Nie, Q., Wan, F.Y.M., Waterman, M.L., and Marsh, J.L. (2007). Wingless directly represses DPP morphogen expression via an Armadillo/TCF/Brinker complex. PLoS ONE 2(1), e142.Google Scholar
Theissen, G. (2006). The proper place of hopeful monsters in evolutionary biology. Theory Biosci. 124, 349369.Google Scholar
Theveneau, E. and Mayor, R. (2012). Neural crest delamination and migration: from epithelium-to-mesenchyme transition to collective cell migration. Dev. Biol. 366, 3454.Google Scholar
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, 11401151.Google Scholar
Thomas, C. and Strutt, D. (2012). The roles of the cadherins Fat and Dachsous in planar polarity specification in Drosophila. Dev. Dynamics 241, 2739.Google Scholar
Thompson, D.B. (2019). Diet-induced plasticity of linear static allometry is not so simple for grasshoppers: genotype–environment interaction in ontogeny is masked by convergent growth. Integr. Comp. Biol. 59, 13821398.Google Scholar
Thompson, D.W. (1917). On Growth and Form. Cambridge University Press, Cambridge.Google Scholar
Thompson, D.W. (1942). On Growth and Form, 2nd ed. Cambridge University Press, Cambridge.Google Scholar
Tickle, C. and Towers, M. (2017). Sonic Hedgehog signaling in limb development. Front. Cell Dev. Biol. 5, 14.Google Scholar
Tingler, M., Kurz, S., Maerker, M., Ott, T., Fuhl, F., Schweickert, A., LeBlanc-Straceski, J.M., Noselli, S., and Blum, M. (2018). A conserved role of the unconventional Myosin 1d in laterality determination. Curr. Biol. 28, 810816.Google Scholar
Tisler, M., Schweickert, A., and Blum, M. (2017). Xenopus, an ideal model organism to study laterality in conjoined twins. Genesis 55, e22993.Google Scholar
Tisler, M., Thumberger, T., Schneider, I., Schweickert, A., and Blum, M. (2017). Leftward flow determines laterality in conjoined twins. Curr. Biol. 27, 543548.Google Scholar
Tobias, J.A., Montgomerie, R., and Lyon, B.E. (2012). The evolution of female ornaments and weaponry: social selection, sexual selection and ecological competition. Philos. Trans. R. Soc. Lond. B 367, 22742293.Google Scholar
Toda, S., Blauch, L.R., Tang, S.K.Y., Morsut, L., and Lim, W.A. (2018). Programming self-organizing multicellular structures with synthetic cell-cell signaling. Science 361, 156162.Google Scholar
Tokita, M. (2015). How the pterosaur got its wings. Biol. Rev. 90, 11631178.Google Scholar
Tokunaga, C. (1962). Cell lineage and differentiation on the male foreleg of Drosophila melanogaster. Dev. Biol. 4, 489516.Google Scholar
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., editor. Springer-Verlag, Berlin, pp. 157204.Google Scholar
Tokunaga, C. (1982). Curt Stern, 1902–1981, in memoriam. Jpn. J. Genet. 57, 459466.Google Scholar
Tomaszewski, R. and Bulandra, A. (2015). Ulnar dimelia: diagnosis and management of a rare congenital anomaly of the upper limb. J. Orthop. 12, S121S124.Google Scholar
Tomita, K., Moriyoshi, K., Nakanishi, S., Guillemot, F., and Kageyama, R. (2000). Mammalian achaete-scute and atonal homologs regulate neuronal versus glial fate determination in the central nervous system. EMBO J. 19, 54605472.Google Scholar
Tomoyasu, Y. (2017). Ultrabithorax and the evolution of insect forewing/hindwing differentiation. Curr. Opin. Insect Sci. 19, 815.Google Scholar
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, 20572065.Google Scholar
Tomoyasu, Y., Nakamura, M., and Ueno, N. (1998). Role of Dpp signalling in prepattern formation of the dorsocentral mechanosensory organ in Drosophila melanogaster. Development 125, 42154224.Google Scholar
Tomoyasu, Y., Wheeler, S.R., and Denell, R.E. (2005). Ultrabithorax is required for membranous wing identity in the beetle Tribolium casteneum. Nature 433, 643647.Google Scholar
Tornini, V.A. and Poss, K.D. (2014). Keeping at arm’s length during regeneration. Dev. Cell 29, 139145.Google Scholar
Toro, R. (2012). On the possible shapes of the human brain. Evol. Biol. 39, 600612.Google Scholar
Toussaint, S., Llamosi, A., Morino, L., and Youlatos, D. (2020). The central role of small vertical substrates for the origin of grasping in early primates. Curr. Biol. 30, 16001613.Google Scholar
Tozluoglu, M., Duda, M., Kirkland, N.J., Barrientos, R., Burden, J.J., Muñoz, J.J., and Mao, Y. (2019). Planar differential growth rates initiate precise fold positions in complex epithelia. Dev. Cell 51, 299312.Google Scholar
Triggs-Raine, B. and Natowicz, M.R. (2015). Biology of hyaluronan: insights from genetic disorders of hyaluronan metabolism. World J. Biol. Chem. 6, 110120.Google Scholar
Tripathi, A., Swaroop, S., and Varadarajan, R. (2019). Molecular determinants of temperature-sensitive phenotypes. Biochemistry 58, 17381750.Google Scholar
Troost, T., Schneider, M., and Klein, T. (2015). A re-examination of the selection of the sensory organ precursor of the bristle sensilla of Drosophila melanogaster. PLoS Genet. 11(1), e1004911.Google Scholar
Trueb, L. (1973). Bones, frogs, and evolution. In Evolutionary Biology of the Anurans: Contemporary Research on Major Problems, Vial, J.L., editor. University of Missouri Press, Columbia, MO, pp. 65132.Google Scholar
Turing, A.M. (1952). The chemical basis of morphogenesis. Philos. Trans. Roy. Soc. Lond. B 237, 3772.Google Scholar
Udan, R.S., Kango-Singh, M., Nolo, R., Tao, C., and Halder, G. (2003). Hippo promotes proliferation arrest and apoptosis in the Salvador/Warts pathway. Nat. Cell Biol. 5, 914920.Google Scholar
Ulshafer, R.J. and Clavert, A. (1979). The use of avian double monsters in studies on induction of the nervous system. J. Embryol. Exp. Morphol. 53, 237243.Google Scholar
Umair, M., Ahmad, F., Bilal, M., Ahmad, W., and Alfadhel, M. (2018). Clinical genetics of polydactyly: an updated review. Front. Genet. 9, 447.Google Scholar
Umulis, D.M. and Othmer, H.G. (2013). Mechanisms of scaling in pattern formation. Development 140, 48304843.Google Scholar
Urdy, S., Goudemand, N., and Pantalacci, S. (2016). Looking beyond the genes: the interplay between signaling pathways and mechanics in the shaping and diversification of epithelial tissues. Curr. Top. Dev. Biol. 119, 227290.Google Scholar
Usui, K. and Tokita, M. (2018). Creating diversity in mammalian facial morphology: a review of potential developmental mechanisms. EvoDevo 9, 15.Google Scholar
Usui, K., Goldstone, C., Gibert, J.-M., and Simpson, P. (2008). Redundant mechanisms mediate bristle patterning on the Drosophila thorax. PNAS 105, 2011220117.Google Scholar
Valentin, M.N., Solomon, B.D., Richard, G., Ferreira, C.R., and Kirkorian, A.Y. (2018). Basan gets a new fingerprint: mutations in the skin-specific isoform of SMARCAD1 cause ectodermal dysplasia syndromes with adermatoglyphia. Am. J. Med. Genet. A 176A, 24512455.Google Scholar
Valizadeh, A., Majidinia, M., Samadi-Kafil, H., Yousefi, M., and Yousefi, B. (2019). The roles of signaling pathways in liver repair and regeneration. J. Cell Physiol. 234, 1496614974.Google Scholar
van Amerongen, R. and Nusse, R. (2009). Towards an integrated view of Wnt signaling in development. Development 136, 32053214.Google Scholar
van Arensbergen, J., van Steensel, B., and Bussemaker, H.J. (2014). In search of the determinants of enhancer–promoter interaction specificity. Trends Cell Biol. 24, 695702.Google Scholar
Van Raamsdonk, C.D., Barsh, G.S., Wakamatsu, K., and Ito, S. (2009). Independent regulation of hair and skin color by two G protein-coupled pathways. Pigment Cell Melanoma Res. 22, 819826.Google Scholar
Van Raamsdonk, C.D., Fitch, K.R., Fuchs, H., de Angelis, M.H., and Barsh, G.S. (2004). Effects of G-protein mutations on skin color. Nat. Genet. 36, 961968.Google Scholar
Van Valen, L. (1974). A natural model for the origin of some higher taxa. J. Herpetol. 8, 109121.Google Scholar
Vandamme, N. and Berx, G. (2019). From neural crest cells to melanocytes: cellular plasticity during development and beyond. Cell. Mol. Life Sci. 76, 19191934.Google Scholar
Vandenberg, L.N., Adams, D.S., and Levin, M. (2012). Normalized shape and location of perturbed craniofacial structures in the Xenopus tadpole reveal an innate ability to achieve correct morphology. Dev. Dynamics 241, 863878.Google Scholar
Vandervorst, P. and Ghysen, A. (1980). Genetic control of sensory connections in Drosophila. Nature 286, 6567.Google Scholar
Vargesson, N. (2020). Positional information: a concept underpinning our understanding of developmental biology. Dev. Dynamics 249, 298312.Google Scholar
Venkatasubban, R. (2013). West Texas one-eyed goat appears in “Ripley’s Believe It or Not!”. Midland Reporter-Telegram (Friday, Oct. 18, 2013).Google Scholar
Venot, Q., Blanc, T., Rabia, S.H., Berteloot, L., Ladraa, S., Duong, J.-P., Blanc, E., Johnson, S.C., Hoguin, C., Boccara, O., Sarnacki, S., Boddaert, N., Pannier, S., Martinez, F., Magassa, S., Yamaguchi, J., Knebelmann, B., Merville, P., Grenier, N., Joly, D., Cormier-Daire, V., Michot, C., Bole-Feysot, C., Picard, A., Soupre, V., Lyonnet, S., Sadoine, J., Slimani, L., Chaussain, C., Laroche-Raynaud, C., Guibaud, L., Broissand, C., Amiel, J., Legendre, C., Terzi, F., and Canaud, G. (2018). Targeted therapy in patients with PIK3CA-related overgrowth syndrome. Nature 558, 540546.Google Scholar
Verbrugge, S.A.J., Schönfelder, M., Becker, L., Nezhad, F.Y., de Angelis, M.H., and Wackerhage, H. (2018). Genes whose gain or loss-of-function increases skeletal muscle mass in mice: a systematic literature review. Front. Physiol. 9, 553.Google Scholar
Vergara, M.N., Tsissios, G., and Rio-Tsonis, K.D. (2018). Lens regeneration: a historical perspective. Int. J. Dev. Biol. 62, 351361.Google Scholar
Verheyden, J.M. and Sun, X. (2008). An Fgf/Gremlin inhibitory feedback loop triggers termination of limb bud outgrowth. Nature 454, 638641.Google Scholar
Verma, P.K. and El-Harouni, A.A. (2015). Review of literature: genes related to postaxial polydactyly. Front. Pediatr. 3, 8.Google Scholar
Vicente, C. (2016). An interview with Peter Lawrence. Development 143, 183185.Google Scholar
Vidal, V.P.I., Chaboissier, M.-C., Lützkendorf, S., Cotsarelis, G., Mill, P., Hui, C.-C., Ortone, N., Ortone, J.-P., and Schedl, A. (2005). Sox9 is essential for outer root sheath differentiation and the formation of the hair stem cell compartment. Curr. Biol. 15, 13401351.Google Scholar
Villares, R. and Cabrera, C.V. (1987). The achaete-scute gene complex of D. melanogaster: conserved domains in a subset of genes required for neurogenesis and their homology to myc. Cell 50, 415424.Google Scholar
Vinicius, L. (2005). Human encephalization and developmental timing. J. Human Evol. 49, 762776.Google Scholar
Vogel, G. (2012). Turing pattern fingered for digit formation. Science 338, 1406.Google Scholar
Vogel, G. (2013). How do organs know when they have reached the right size? Science 340, 11561157.Google Scholar
Vogt, T.F. and Duboule, D. (1999). Antagonists go out on a limb. Cell 99, 563566.Google Scholar
Vollmer, J., Casares, F., and Iber, D. (2017). Growth and size control during development. Open Biol. 7, 170190.Google Scholar
von Kalm, L., Fristrom, D., and Fristrom, J. (1995). The making of a fly leg: a model for epithelial morphogenesis. BioEssays 17, 693702.Google Scholar
von Luschan, F. (1907). Ein Haarmensch. Zeitschrift für Ethnologie 39, 425429.Google Scholar
vonHoldt, B.M., Shuldiner, E., Koch, I.J., Kartzinel, R.Y., Hogan, A., Brubaker, L., Wanser, S., Stahler, D., Wynne, C.D.L., Ostrander, E.A., Sinsheimer, J.S., and Udell, M.A.R. (2017). Structural variants in genes associated with human Williams-Beuren syndrome underlie stereotypical hypersociability in domestic dogs. Sci. Adv. 3, e1700398.Google Scholar
Waddington, C.H. (1940). Organizers and Genes. Cambridge University Press, Cambridge.Google Scholar
Waddington, C.H. (1956). Genetic assimilation of the bithorax phenotype. Evolution 10, 113.Google Scholar
Waddington, C.H. (1957). The Strategy of the Genes: A Discussion of Some Aspects of Theoretical Biology. George Allen & Unwin, London.Google Scholar
Walentek, P., Beyer, T., Thumberger, T., Schweickert, A., and Blum, M. (2012). ATP4a Is required for Wnt-dependent Foxj1 expression and leftward flow in Xenopus left–right development. Cell Rep. 1, 516527.Google Scholar
Wallingford, J.B. (2019). We are all developmental biologists. Dev. Cell 50, 132137.Google Scholar
Walsh, C.A. (2019). Rainer W. Guillery and the genetic analysis of brain development. Eur. J. Neurosci. 49, 900908.Google Scholar
Wang, C., Rüther, U., and Wang, B. (2007). The Shh-independent activator function of the full-length Gli3 protein and its role in vertebrate limb digit patterning. Dev. Biol. 305, 460469.Google Scholar
Wang, D., Berg, D., Ba, H., Sun, H., Wang, Z., and Li, C. (2019). Deer antler stem cells are a novel type of cells that sustain full regeneration of a mammalian organ – deer antler. Cell Death Dis. 10, 443.Google Scholar
Wang, Y., Dong, L., and Evans, S.E. (2016). Polydactyly and other limb abnormalities in the Jurassic salamander Chunerpeton from China. Palaeobio. Palaeoenv. 96, 4959.Google Scholar
Wang, Y., Wang, G.-D., He, Q.-L., Luo, Z.-P., Yang, L., Yao, Q., and Chen, K.-P. (2020). Phylogenetic analysis of achaete-scute complex genes in metazoans. Mol. Genet. Genomics 295, 591606.Google Scholar
Wanninger, A. and Wollesen, T. (2019). The evolution of molluscs. Biol. Rev. 94, 102115.Google Scholar
Warman, P.H. and Ennos, A.R. (2009). Fingerprints are unlikely to increase the friction of primate fingerpads. J. Exp. Biol. 212, 20162022.Google Scholar
Warren, I. and Smith, H. (2007). Stalk-eyed flies (Diopsidae): modelling the evolution and development of an exaggerated sexual trait. BioEssays 29, 300307.Google Scholar
Warren, I.A., Gotoh, H., Dworkin, I.M., Emlen, D.J., and Lavine, L.C. (2013). A general mechanism for conditional expression of exaggerated sexually-selected traits. BioEssays 35, 889899.Google Scholar
Warrington, S.J., Strutt, H., Fisher, K.H., and Strutt, D. (2017). A dual function for Prickle in regulating frizzled stability during feedback-dependent amplification of planar polarity. Curr. Biol. 27, 27842797.Google Scholar
Wartlick, O., Mumcu, P., Jülicher, F., and Gonzalez-Gaitan, M. (2011). Understanding morphogenetic growth control: lessons from flies. Nat. Rev. Mol. Cell Biol. 12, 594604.Google Scholar
Watanabe, M. and Kondo, S. (2015). Is pigment patterning in fish skin determined by the Turing mechanism? Trends Genet. 31, 8896.Google Scholar
Watt, K.I., Harvey, K.F., and Gregorevic, P. (2017). Regulation of tissue growth by the mammalian Hippo signaling pathway. Front. Physiol. 8, 942.Google Scholar
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, 14741482.Google Scholar
Weaver, C. and Kimelman, D. (2004). Move it or lose it: axis specification in Xenopus. Development 131, 34913499.Google Scholar
Webb, A.A. and Cullen, C.L. (2010). Coat color and coat color pattern-related neurologic and neuro-ophthalmic diseases. CVJ 51, 653657.Google Scholar
Weber, C., Zhou, Y., Lee, J.G., Looger, L.L., Qian, G., Ge, C., and Capel, B. (2020). Temperature-dependent sex determination is mediated by pSTAT3 repression of Kdm6b. Science 368, 303306.Google Scholar
Weber, M.A. and Sebire, N.J. (2010). Genetics and developmental pathology of twinning. Semin. Fetal Neonatal Med. 15, 313318.Google Scholar
Weiner, J. (1999). Time, Love, Memory. Random House, New York.Google Scholar
Weirauch, M.T. and Hughes, T.R. (2010). Conserved expression without conserved regulatory sequence: the more things change, the more they stay the same. Trends Genet. 26, 6674.Google Scholar
Weiss, P. (1939). Principles of Development. Henry Holt & Co., New York.Google Scholar
Wellik, D.M. and Capecchi, M.R. (2003). Hox10 and Hox11 genes are required to globally pattern the mammalian skeleton. Science 301, 363367.Google Scholar
Welte, M.A., Duncan, I., and Lindquist, S. (1995). The basis for a heat-induced developmental defect: defining crucial lesions. Genes Dev. 9, 22402250.Google Scholar
Welton, T., Ather, S., Proudlock, F.A., Gottlob, I., and Dineen, R.A. (2017). Altered whole-brain connectivity in albinism. Hum. Brain Mapp. 38, 740752.Google Scholar
Wendelin, D.S., Pope, D.N., and Mallory, S.B. (2003). Hypertrichosis. J. Am. Acad. Dermatol. 48, 161179.Google Scholar
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, 383400.Google Scholar
Wernet, M.F. and Desplan, C. (2015). Brain wiring in the fourth dimension. Cell 162, 2022.Google Scholar
Wernet, M.F., Mazzoni, E.O., Celik, A., Duncan, D.M., Duncan, I., and Desplan, C. (2006). Stochastic spineless expression creates the retinal mosaic for color vision. Nature 440, 174180.Google Scholar
Werts, A.D. and Goldstein, B. (2011). How signaling between cells can orient a mitotic spindle. Semin. Cell Dev. Biol. 22, 842849.Google Scholar
West-Eberhard, M.J. (2003). Developmental Plasticity and Evolution. Oxford University Press, New York.Google Scholar
Westgate, G.E., Ginger, R.S., and Green, M.R. (2017). The biology and genetics of curly hair. Exp. Dermatol. 26, 483490.Google Scholar
Westhusin, M. (1997). From mighty mice to mighty cows. Nat. Genet. 17, 45.Google Scholar
Wexler, J.R., Plachetzki, D.C., and Kopp, A. (2014). Pan-metazoan phylogeny of the DMRT gene family: a framework for functional studies. Dev. Genes Evol. 224, 175181.Google Scholar
White, R.A.H. and Akam, M.E. (1985). Contrabithorax mutations cause inappropriate expression of Ultrabithorax products in Drosophila. Nature 318, 567569.Google Scholar
Whiting, M.F. and Wheeler, W.C. (1994). Insect homeotic transformation. Nature 368, 696.Google Scholar
Wianny, F., Kennedy, H., and Dehay, C. (2018). Bridging the gap between mechanics and genetics in cortical folding: ECM as a major driving force. Neuron 99, 625627.Google Scholar
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. C Embryo Today 78, 280291.Google Scholar
Widmann, T.J. and Dahmann, C. (2009). Dpp signaling promotes the cuboidal-to-columnar shape transition of Drosophila wing disc epithelia by regulating Rho1. J. Cell Sci. 122, 13621373.Google Scholar
Widmann, T.J. and Dahmann, C. (2009). Wingless signaling and the control of cell shape in wing imaginal discs. Dev. Biol. 334, 161173.Google Scholar
Wieschaus, E. (1978). Cell lineage relationships in the Drosophila embryo. In Genetic Mosaics and Cell Differentiation, Gehring, W.J., editor. Springer-Verlag, Berlin, pp. 97118.Google Scholar
Wieschaus, E. and Gehring, W. (1976). Gynandromorph analysis of the thoracic disc primordia in Drosophila melanogaster. W. Roux Arch. Dev. Biol. 180, 3146.Google Scholar
Wieschaus, E. and Nüsslein-Volhard, C. (2014). Walter Gehring (1939–2014). Curr. Biol. 24, R632R634.Google Scholar
Wieschaus, E. and Nüsslein-Volhard, C. (2016). The Heidelberg screen for pattern mutants of Drosophila: a personal account. Annu. Rev. Cell Dev. Biol. 32, 146.Google Scholar
Wigglesworth, V.B. (1940). Local and general factors in the development of “pattern” in Rhodnius prolixus (Hemiptera). J. Exp. Biol. 17, 180200.Google Scholar
Wigglesworth, V.B. (1988). The control of pattern as seen in the integument of an insect. BioEssays 9, 2327.Google Scholar
Wilkins, A.S. (2020). A striking example of developmental bias in an evolutionary process: the “domestication syndrome”. Evol. Dev. 22, 143153.Google Scholar
Wilkins, A.S., Wrangham, R., and Fitch, W.T. (2014). The “domestication syndrome” in mammals: a unified explanation based on neural crest cell behavior and genetics. Genetics 197, 795808.Google Scholar
Williams, M.L.K. and Solnica-Krezel, L. (2017). Regulation of gastrulation movements by emergent cell and tissue interactions. Curr. Opin. Cell Biol. 48, 3339.Google Scholar
Williams, R.W., Hogan, D., and Garraghty, P.E. (1994). Target recognition and visual maps in the thalamus of achiasmatic dogs. Nature 367, 637639.Google Scholar
Winchester, G. (1996). The Morgan lineage. Curr. Biol. 6, 100101.Google Scholar
Winchester, G. (2004). Edward B. Lewis 1918–2004. Curr. Biol. 14, R740R742.Google Scholar
Winter, R.M. (1996). Analyzing human developmental abnormalities. BioEssays 18, 965971.Google Scholar
Wolff, G.L. (1955). The effects of environmental temperature on coat color in diverse genotypes of the guinea pig. Genetics 40, 90106.Google Scholar
Wolpert, L. (1969). Positional information and the spatial pattern of cellular differentiation. J. Theor. Biol. 25, 147.Google Scholar
Wolpert, L. (1971). Positional information and pattern formation. Curr. Top. Dev. Biol. 6, 183224.Google Scholar
Wolpert, L. (1981). Positional information and pattern formation. Philos. Trans. Roy. Soc. Lond. B 295, 441450.Google Scholar
Wolpert, L. (1989). Positional information and prepattern in the development of pattern. In Cell to Cell Signalling: From Experiments to Theoretical Models, Goldbeter, A., editor. Academic Press, New York, pp. 133143.Google Scholar
Wolpert, L. (1996). One hundred years of positional information. Trends Genet. 12, 359364.Google Scholar
Wolpert, L., Tickle, C., Martinez Arias, A., Lawrence, P., Lumsden, A., Robertson, E., Meyerowitz, E., and Smith, J. (2015). Principles of Development, 5th ed. Oxford University Press, New York.Google Scholar
Woltering, J.M. and Duboule, D. (2010). The origin of digits: expression patterns versus regulatory mechanisms. Cell 18, 526532.Google Scholar
Wong, R., Geyer, S., Weninger, W., Guimberteau, J.-C., and Wong, J.K. (2016). The dynamic anatomy and patterning of skin. Exp. Dermatol. 25, 9298.Google Scholar
Worley, M.I., Alexander, L.A., and Hariharan, I.K. (2018). CtBP impedes JNK- and Upd/STAT-driven cell fate misspecifications in regenerating Drosophila imaginal discs. eLife 7, e30391.Google Scholar
Wright, S. (1915). The albino series of allelomorphs in guinea-pigs. Am. Nat. 49, 140148.Google Scholar
Wu, J., Greely, H.T., Jaenisch, R., Nakauchi, H., Rossant, J., and Izpisua Belmonte, J.C. (2016). Stem cells and interspecies chimaeras. Nature 540, 5159.Google Scholar
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, 249270.Google Scholar
Wutke, S., Benecke, N., Sandoval-Castellanos, E., Döhle, H.-J., Friederich, S., Gonzalez, J., Hallsson, J.H., Hofreiter, M., Löugas, L., Magnell, O., Morales-Muniz, A., Orlando, L., Pálsdóttir, A.H., Reissmann, M., Ruttkay, M., Trinks, A., and Ludwig, A. (2016). Spotted phenotypes in horses lost attractiveness in the Middle Ages. Sci. Rep. 6, 38548.Google Scholar
Wyatt, T.P.J., Fouchard, J., Lisica, A., Khalilgharibi, N., Baum, B., Recho, P., Kabla, A.J., and Charras, G.T. (2020). Actomyosin controls planarity and folding of epithelia in response to compression. Nat. Mater. 19, 109117.Google Scholar
Xu, H.-J., Xue, J., Lu, B., Zhang, X.-C., Zhuo, J.-C., He, S.-F., Ma, X.-F., Jiang, Y.-Q., Fan, H.-W., Xu, J.-Y., Ye, Y.-X., Pan, P.-L., Li, Q., Bao, Y.-Y., Nijhout, H.F., and Zhang, C.-X. (2015). Two insulin receptors determine alternative wing morphs in planthoppers. Nature 519, 464467.Google Scholar
Yamamoto, D., Jallon, J.-M., and Komatsu, A. (1997). Genetic dissection of sexual behavior in Drosophila melanogaster. Annu. Rev. Entomol. 42, 551585.Google Scholar
Yan, B. (2010). Numb: from flies to humans. Brain Dev. 32, 293298.Google Scholar
Yang, L.Q., Zhang, K., Wu, Q.Y., Li, J., Lai, S.J., Song, T.Z., and Zhang, M. (2019). Identification of two novel single nucleotide polymorphism sites in the myostatin (mstn) gene and their association with carcass traits in meat-type rabbits (Oryctolagus cuniculus). World Rabbit Sci. 27, 249256.Google Scholar
Yang, Z.-Q., Zhang, H.-L., Duan, C.-C., Geng, S., Wang, K., Yu, H.-F., Yue, Z.-P., and Guo, B. (2017). IGF1 regulates RUNX1 expression via IRS1/2: implications for antler chondrocyte differentiation. Cell Cycle 16, 522532.Google Scholar
Ye, X.C., Pegado, V., Patel, M.S., and Wasserman, W.W. (2014). Strabismus genetics across a spectrum of eye misalignment disorders. Clin. Genet. 86, 103111.Google Scholar
Yip, R.K.H., Chan, D., and Cheah, K.S.E. (2019). Mechanistic insights into skeletal development gained from genetic disorders. Curr. Top. Dev. Biol. 133, 343385.Google Scholar
Yokoyama, H., Yonei-Tamura, S., Endo, T., Izpisúa Belmonte, J.C., Tamura, K., and Ide, H. (2000). Mesenchyme with fgf–10 expression is responsible for regenerative capacity in Xenopus limb buds. Dev. Biol. 219, 1829.Google Scholar
Yuan, S. and Brueckner, M. (2018). Left–right asymmetry: myosin 1D at the center. Curr. Biol. 28, R567R569.Google Scholar
Zakany, J. and Duboule, D. (2007). The role of Hox genes during vertebrate limb development. Curr. Opin. Genet. Dev. 17, 359366.Google Scholar
Zalokar, M., Erk, I., and Santamaria, P. (1980). Distribution of ring-X chromosomes in the blastoderm of gynandromorphic D. melanogaster. Cell 19, 133141.Google Scholar
Zanella, M., Vitriolo, A., Andirko, A., Martins, P.T., Sturm, S., O’Rourke, T., Laugsch, M., Malerba, N., Skaros, A., Trattaro, S., Germain, P.-L., Mihailovic, M., Merla, G., Rada-Iglesias, A., Boeckx, C., and Testa, G. (2019). Dosage analysis of the 7q11.23 Williams region identifies BAZ1B as a major human gene patterning the modern human face and underlying self-domestication. Sci. Adv. 5, eaaw7908.Google Scholar
Zanna, G., Docampo, M.J., Fondevila, D., Bardagi, M., Bassols, A., and Ferrer, L. (2009). Hereditary cutaneous mucinosis in shar pei dogs is associated with increased hyaluronan synthase-2 mRNA transcription by cultured dermal fibroblasts. Vet. Dermatol. 20, 377382.Google Scholar
Zanna, G., Fondevila, D., Bardagi, M., Docampo, M.J., Bassols, A., and Ferrer, L. (2008). Cutaneous mucinosis in shar-pei dogs is due to hyaluronic acid deposition and is associated with high levels of hyaluronic acid in serum. Vet. Dermatol. 19, 314318.Google Scholar
Zartman, J.J. and Shvartsman, S.Y. (2010). Unit operations of tissue development: epithelial folding. Annu. Rev. Chem. Biomol. Eng. 1, 231246.Google Scholar
Zhang, C., Wang, F., Gao, Z., Zhang, P., Gao, J., and Wu, X. (2020). Regulation of Hippo signaling by mechanical signals and the cytoskeleton. DNA Cell Biol. 39, 159166.Google Scholar
Zhang, H., Tang, C., Li, X., and Kong, A.W.K. (2014). A study of similarity between genetically identical body vein patterns. In IEEE Symposium on Computational Intelligence in Biometrics and Identity Management (CIBIM). Orlando, FL, 151159.Google Scholar
Zhang, N., Guo, L., and Simpson, J.H. (2020). Spatial comparisons of mechanosensory information govern the grooming sequence in Drosophila. Curr. Biol. 30, 9881001.Google Scholar
Zhang, Y., Andl, T., Yang, S.H., Teta, M., Liu, F., Seykora, J.T., Tobias, J.W., Piccolo, S., Schmidt-Ullrich, R., Nagy, A., Taketo, M.M., Dlugosz, A.A., and Millar, S.E. (2008). Activation of β-catenin signaling programs embryonic epidermis to hair follicle fate. Development 135, 21612172.Google Scholar
Zhang, Y., Tomann, P., Andl, T., Gallant, N.M., Huelsken, J., Jerchow, B., Birchmeier, W., Paus, R., Piccolo, S., Mikkola, M.L., Morrisey, E.E., Overbeek, P.A., Scheidereit, C., Millar, S.E., and Schmidt-Ullrich, R. (2009). Reciprocal requirements for EDA/EDAR/NF-κB and Wnt/β-catenin signaling pathways in hair follicle induction. Dev. Cell 17, 4961.Google Scholar
Zheng, Y. and Pan, D. (2019). The Hippo signaling pathway in development and disease. Dev. Cell 50, 264282.Google Scholar
Zhou, J.X. and Huang, S. (2010). Understanding gene circuits at cell-fate branch points for rational cell reprogramming. Trends Genet. 27, 5562.Google Scholar
Zhou, P., Byrne, C., Jacobs, J., and Fuchs, E. (1995). Lymphoid enhancer factor 1 directs hair follicle patterning and epithelial cell fate. Genes Dev. 9, 570583.Google Scholar
Zhu, J. and Mackem, S. (2017). John Saunders’ ZPA, Sonic hedgehog and digit identity: how does it really all work? Dev. Biol. 429, 391400.Google Scholar
Zilles, K., Palomero-Gallagher, N., and Amunts, K. (2013). Development of cortical folding during evolution and ontogeny. Trends Neurosci. 36, 275284.Google Scholar
Zimova, M., Hackländer, K., Good, J.M., Melo-Ferreira, J., Alves, P.C., and Mills, L.S. (2018). Function and underlying mechanisms of seasonal colour moulting in mammals and birds: what keeps them changing in a warming world? Biol. Rev. 93, 14781498.Google Scholar
Zuniga, A. and Zeller, R. (2014). In Turing’s hands: the making of digits. Science 345, 516517.Google Scholar
Zylberberg, J. and Strowbridge, B.W. (2017). Mechanisms of persistent activity in cortical circuits: possible neural substrates for working memory. Annu. Rev. Neurosci. 40, 603627.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

  • References
  • Lewis I. Held, Jr, Texas Tech University
  • Book: Animal Anomalies
  • Online publication: 26 February 2021
  • Chapter DOI: https://doi.org/10.1017/9781108876612.020
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: Animal Anomalies
  • Online publication: 26 February 2021
  • Chapter DOI: https://doi.org/10.1017/9781108876612.020
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: Animal Anomalies
  • Online publication: 26 February 2021
  • Chapter DOI: https://doi.org/10.1017/9781108876612.020
Available formats
×