Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-t5pn6 Total loading time: 0 Render date: 2024-04-24T23:48:05.335Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  23 February 2017

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
Deep Homology?
Uncanny Similarities of Humans and Flies Uncovered by Evo-Devo
, pp. 135 - 266
Publisher: Cambridge University Press
Print publication year: 2017

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

Abbott, A., and other members of the C. elegans Sequencing Consortium (1998). Genome sequence of the nematode C. elegans: a platform for investigating biology. Science 282, 20122018.Google Scholar
Abbott, L.A., and Lindenmayer, A. (1981). Models for growth of clones in hexagonal cell arrangements: applications in Drosophila wing disc epithelia and plant epidermal tissues. J. Theor. Biol. 90, 495544.Google Scholar
Abdus-Saboor, I., Al Nufal, M.J., Agha, M.V., Ruinart de Brimont, M., Fleischmann, A., and Shykind, B.M. (2016). An expression refinement process ensures singular odorant receptor gene choice. Curr. Biol. 26, 10831090.CrossRefGoogle ScholarPubMed
Abelló, G., and Alsina, B. (2007). Establishment of a proneural field in the inner ear. Int. J. Dev. Biol. 51, 483493.CrossRefGoogle ScholarPubMed
Abmayr, S.M., and Pavlath, G.K. (2012). Myoblast fusion: lessons from flies and mice. Development 139, 641656.CrossRefGoogle ScholarPubMed
Aboitiz, F., and Montiel, J. (2003). One hundred million years of interhemispheric communication: the history of the corpus callosum. Braz. J. Med. Biol. Res. 36, 409420.Google Scholar
Aboobaker, A.A., and Blaxter, M. (2010). The nematode story: Hox gene loss and rapid evolution. In Deutsch, J.S. (ed.), Hox Genes: Studies from the 20th to the 21st Century. Landes Bioscience, Austin, TX, pp. 101110.Google Scholar
Aboobaker, A.A., Tomancak, P., Patel, N., Rubin, G.M., and Lai, E.C. (2005). Drosophila microRNAs exhibit diverse spatial expression patterns during embryonic development. PNAS 102, #50, 1801718022.Google Scholar
Abouheif, E. (1997). Developmental genetics and homology: a hierarchical approach. Trends Ecol. Evol. 12, 405408.CrossRefGoogle ScholarPubMed
Abouheif, E., Akam, M., Dickinson, W.J., Holland, P.W.H., Meyer, A., Patel, N.H., Raff, R.A., Roth, V.L., and Wray, G.A. (1997). Homology and developmental genes. Trends Genet. 13, 432433.CrossRefGoogle ScholarPubMed
Abramov, I., and Gordon, J. (1994). Color appearance: on seeing red – or yellow, or green, or blue. Annu. Rev. Psychol. 45, 451485.CrossRefGoogle ScholarPubMed
Abu-Shaar, M., and Mann, R.S. (1998). Generation of multiple antagonistic domains along the proximodistal axis during Drosophila leg development. Development 125, 38213830.Google Scholar
Abu-Shaar, M., Ryoo, H.D., and Mann, R.S. (1999). Control of the nuclear localization of Extradenticle by competing nuclear import and export signals. Genes Dev. 13, 935945.CrossRefGoogle ScholarPubMed
Abuin, L., Bargeton, B., Ulbrich, M.H., Isacoff, E.Y., Kellenberger, S., and Benton, R. (2011). Functional architecture of olfactory ionotropic glutamate receptors. Neuron 69, 4460.Google Scholar
Acampora, D., Annino, A., Tuorto, F., Puelles, E., Lucchesi, W., Papalia, A., and Simeone, A. (2005). Otx genes in the evolution of the vertebrate brain. Brain Res. Bull. 66, 410420.Google Scholar
Achim, K., and Arendt, D. (2014). Structural evolution of cell types by step-wise assembly of cellular modules. Curr. Opin. Genet. Dev. 27, 102108.CrossRefGoogle ScholarPubMed
Adachi, R., Sasaki, Y., Morita, K., Shirakawa, H., Goto, T., Furuyama, A., and Isono, K. (2012). Behavioral analysis of Drosophila transformants expressing human taste receptor genes in the gustatory receptor neurons. J. Neurogenet. 26, 198205.Google Scholar
Adam, J., Myat, A., Le Roux, I., Eddison, M., Henrique, D., Ish-Horowicz, D., and Lewis, J. (1998). Cell fate choices and the expression of Notch, Delta and Serrate homologues in the chick inner ear: parallels with Drosophila sense-organ development. Development 125, 46454654.Google Scholar
Adams, D.S., Robinson, K.R., Fukumoto, T., Yuan, S., Albertson, R.C., Yelick, P., Kuo, L., McSweeney, M., and Levin, M. (2006). Early, H+-V-ATPase-dependent proton flux is necessary for consistent left-right patterning of non-mammalian vertebrates. Development 133, 16571671.CrossRefGoogle ScholarPubMed
Adler, P.N., and Nathans, J. (2016). The cellular compass. Sci. Am. 314, 6671.CrossRefGoogle ScholarPubMed
Affolter, M., Marty, T., and Vigano, M.A. (1999). Balancing import and export in development. Genes Dev. 13, 913915.Google Scholar
Afzelius, B.A. (1976). A human syndrome caused by immobile cilia. Science 193, 317319.CrossRefGoogle Scholar
Afzelius, B.A. (1995). Situs inversus and ciliary abnormalities. What is the connection? Int. J. Dev. Biol. 39, 839844.Google ScholarPubMed
Afzelius, B.A., and Stenram, U. (2006). Prevalence and genetics of immotile-cilia syndrome and left-handedness. Int. J. Dev. Biol. 50, 571573.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
Aguilar-Hidalgo, D., Domínguez-Cejudo, M.A., Amore, G., Brockmann, A., Lemos, M.C., Córdoba, A., and Casares, F. (2013). A Hh-driven gene network controls specification, pattern and size of the Drosophila simple eyes. Development 140, 8292.Google Scholar
Aguilera, F., McDougall, C., and Degnan, B.M. (2013). Origin, evolution and classification of type-3 copper proteins: lineage-specific gene expansions and losses across the Metazoa. BMC Evol. Biol. 13, Article 96.Google Scholar
Ahn, Y., Mullan, H.E., and Krumlauf, R. (2014). Long-range regulation by shared retinoic acid response elements modulates dynamic expression of posterior Hoxb genes in CNS development. Dev. Biol. 388, 134144.Google Scholar
Ahnelt, P.K., and Kolb, H. (2000). The mammalian photoreceptor: mosaic-adaptive design. Prog. Retin. Eye Res. 19, 711777.Google Scholar
Akam, M. (1987). The molecular basis for metameric pattern in the Drosophila embryo. Development 101, 122.Google Scholar
Akam, M. (1989). Hox and HOM: homologous gene clusters in insects and vertebrates. Cell 57, 347349.Google Scholar
Akam, M. (1998). Hox genes, homeosis and the evolution of segment identity: no need for hopeless monsters. Int. J. Dev. Biol. 42, 445451.Google Scholar
Akam, M. (1998). Hox genes: from master genes to micromanagers. Curr. Biol. 8, R676R678.Google Scholar
Akins, K.A., and Hahn, M. (2014). More than mere colouring: the role of spectral information in human vision. Br. J. Philos. Sci. 65, 125171.Google Scholar
Akiyama-Oda, Y., and Oda, H. (2006). Axis specification in the spider embryo: dpp is required for radial-to-axial symmetry transformation and sog for ventral patterning. Development 133, 23472357.Google Scholar
Al-Qattan, M.M. (2003). Congenital duplication of the palm in a patient with multiple anomalies. J. Hand Surg. Br. 28B, 276279.Google Scholar
Al Qattan, M.M. (2004). On the emerging evidence of a new category of duplication in the human hand: the dorsoventral duplication. Plast. Reconstr. Surg. 114, 12331237.Google Scholar
Ala-Laurila, P., and Rieke, F. (2014). Coincidence detection of single-photon responses in the inner retina at the sensitivity limit of vision. Curr. Biol. 24, 28882898.Google Scholar
Albalat, R. (2009). The retinoic acid machinery in invertebrates: ancestral elements and vertebrate innovations. Mol. Cell. Endocrinol. 313, 2335.CrossRefGoogle ScholarPubMed
Albalat, R. (2012). Evolution of the genetic machinery of the visual cycle: a novelty of the vertebrate eye? Mol. Biol. Evol. 29, 14611469.Google Scholar
Albalat, R., Brunet, F., Laudet, V., and Schubert, M. (2011). Evolution of retinoid and steroid signaling: vertebrate diversification from an amphioxus perspective. Genome Biol. Evol. 3, 9851005.Google Scholar
Albert, J. (2011). Sensory transduction: the “swarm intelligence” of auditory hair bundles. Curr. Biol. 21, R632R634.Google Scholar
Albert, J.T., and Göpfert, M.C. (2015). Hearing in Drosophila. Curr. Opin. Neurobiol. 34, 7985.Google Scholar
Alberts, B., Johnson, A., Lewis, J., Raff, M., Roberts, K., and Walter, P. (2002). Molecular Biology of the Cell, 4th edn. Garland, New York, NY.Google Scholar
Alexander, T., Nolte, C., and Krumlauf, R. (2009). Hox genes and segmentation of the hindbrain and axial skeleton. Annu. Rev. Cell Dev. Biol. 25, 431456.Google Scholar
Alfano, C., and Studer, M. (2013). Neocortical arealization: evolution, mechanisms, and open questions. Dev. Neurobiol. 73, 411447.CrossRefGoogle ScholarPubMed
Alié, A., and Manuel, M. (2010). The backbone of the post-synaptic density originated in a unicellular ancestor of choanoflagellates and metazoans. BMC Evol. Biol. 10, Article 34.CrossRefGoogle Scholar
Allan, D.W., and Thor, S. (2015). Transcriptional selectors, masters, and combinatorial codes: regulatory principles of neural subtype specification. WIREs Dev. Biol. 4, 505528.Google Scholar
Allen, A.E., Storchi, R., Martial, F.P., Petersen, R.S., Montemurro, M.A., Brown, T.M., and Lucas, R.J. (2014). Melanopsin-driven light adaptation in mouse vision. Curr. Biol. 24, 24812490.Google Scholar
Allen, J., and Chilton, J.K. (2009). The specific targeting of guidance receptors within neurons: who directs the directors? Dev. Biol. 327, 411.Google Scholar
Allison, W.T., Barthel, L.K., Skebo, K.M., Takechi, M., Kawamura, S., and Raymond, P.A. (2010). Ontogeny of cone photoreceptor mosaics in zebrafish. J. Comp. Neurol. 518, 41824195.Google Scholar
Alsina, B., Giraldez, F., and Pujades, C. (2009). Patterning and cell fate in ear development. Int. J. Dev. Biol. 53, 15031513.Google Scholar
Álvarez-Fernández, C., Tamirisa, S., Prada, F., Chernomoretz, A., Podhajcer, O., Blanco, E., and Martín-Blanco, E. (2015). Identification and functional analysis of healing regulators in Drosophila. PLoS Genet. 11, #2, e1004965.Google Scholar
Ambegaonkar, A.A., and Irvine, K.D. (2015). Coordination of planar cell polarity pathways through Spiny legs. eLife 4, e09946.CrossRefGoogle ScholarPubMed
Ameisen, J.C. (2002). On the origin, evolution, and nature of programmed cell death: a timeline of four billion years. Cell Death Differ. 9, 367393. [See alsoGoogle ScholarGoogle Scholar
Amemiya, C.T., Miyake, T., and Rast, J.P. (2005). Echinoderms. Curr. Biol. 15, R944R946.Google Scholar
Amemiya, C.T., Prohaska, S.J., Hill-Force, A., Cook, A., Wasserscheid, J., Ferrier, D.E.K., Pascual-Anaya, J., Garcia-Fernàndez, J., Dewar, K., and Stadler, P.F. (2008). The amphioxus Hox cluster: characterization, comparative genomics, and evolution. J. Exp. Zool. B. Mol. Dev. Evol. 310, 465477.Google Scholar
Amodio, V., Tevy, M.F., Traina, C., Ghosh, T.K., and Capovilla, M. (2012). Transactivation in Drosophila of human enhancers by human transcription factors involved in congenital heart diseases. Dev. Dyn. 241, 190199.Google Scholar
Amour, A., Bird, M., Chaudry, L., Deadman, J., Hayes, D., and Kay, C. (2004). General considerations for proteolytic cascades. Biochem. Soc. Trans. 32, 1516.Google Scholar
Amrein, H., and Thorne, N. (2005). Gustatory perception and behavior in Drosophila melanogaster. Curr. Biol. 15, R673R684.Google Scholar
Anderson, E., Peluso, S., Lettice, L.A., and Hill, R.E. (2012). Human limb abnormalities caused by disruption of hedgehog signaling. Trends Genet. 28, 364373.Google Scholar
Anderson, J.M., Horton, P., Kim, E.-H., and Chow, W.S. (2012). Towards elucidation of dynamic structural changes of plant thylakoid architecture. Philos. Trans. R. Soc. Lond. B 367, 35153524.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
Andretic, R. (2015). Neurobiology: what drives flies to sleep? Curr. Biol. 25, R1086R1088.Google Scholar
Andrews, G.L., Tanglao, S., Farmer, W.T., Morin, S., Brotman, S., Berberoglu, M.A., Price, H., Fernandez, G.C., Mastick, G.S., Charron, F., and Kidd, T. (2008). Dscam guides embryonic axons by Netrin-dependent and -independent functions. Development 135, 38393848.Google Scholar
Angelini, D.R., and Kaufman, T.C. (2005). Comparative developmental genetics and the evolution of arthropod body plans. Annu. Rev. Genet. 39, 95119.Google Scholar
Angelini, D.R., and Kaufman, T.C. (2005). Insect appendages and comparative ontogenetics. Dev. Biol. 286, 5777.Google Scholar
Angioy, A.M., Desogus, A., Barbarossa, I.T., Anderson, P., and Hansson, B.S. (2003). Extreme sensitivity in an olfactory system. Chem. Senses 28, 279284.Google Scholar
Angueyra, J.M., Pulido, C., Malagón, G., Nasi, E., and Gomez, M.d.P. (2012). Melanopsin-expressing amphioxus photoreceptors transduce light via a phospholipase C signaling cascade. PLoS ONE 7, #1, e29813.Google Scholar
Annona, G., Holland, N.D., and D'Aniello, S. (2015). Evolution of the notochord. EvoDevo 6, Article 30.Google Scholar
Anonymous, (2001). Little humans with wings? Spectra: The Newsletter of the Carnegie Institution Spring 2001, p. 2004.Google Scholar
Antic, D., Stubbs, J.L., Suyama, K., Kinter, C., Scott, M.P., and Axelrod, J. (2010). Planar cell polarity enables posterior localization of nodal cilia and left-right axis determination during mouse and Xenopus embryogenesis. PLoS ONE 5, #2, e8999.Google Scholar
Apostolopoulou, A.A., Rist, A., and Thum, A.S. (2015). Taste processing in Drosophila larvae. Front. Integr. Neurosci. 9, Article 50.Google Scholar
Appel, T.A. (1987). The Cuvier–Geoffroy Debate. Oxford University Press, New York, NY.Google Scholar
Applebury, M.L., Antoch, M.P., Baxter, L.C., Chun, L.L.Y., Falk, J.D., Farhangfar, F., Kage, K., Krzystolik, M.G., Lyass, L.A., and Robbins, J.T. (2000). The murine cone photoreceptor: a single cone type expresses both S and M opsins with retinal spatial patterning. Neuron 27, 513523.Google Scholar
Araújo, S.J., and Tear, G. (2003). Axon guidance mechanisms and molecules: lessons from invertebrates. Nat. Rev. Neurosci. 4, 910922.CrossRefGoogle ScholarPubMed
Aravind, L., Dixit, V.M., and Koonin, E.V. (2001). Apoptotic molecular machinery: vastly increased complexity in vertebrates revealed by genome comparisons. Science 291, 12791284.CrossRefGoogle ScholarPubMed
Arendt, D. (2003). Evolution of eyes and photoreceptor cell types. Int. J. Dev. Biol. 47, 563571.Google Scholar
Arendt, D. (2005). Genes and homology in nervous system evolution: comparing gene functions, expression patterns, and cell type molecular fingerprints. Theory Biosci. 124, 185197.Google Scholar
Arendt, D. (2008). The evolution of cell types in animals: emerging principles from molecular studies. Nat. Rev. Genet. 9, 868882.CrossRefGoogle ScholarPubMed
Arendt, D., Denes, A.S., Jékely, G., and Tessmar-Raible, K. (2008). The evolution of nervous system centralization. Philos. Trans. R. Soc. Lond. B 363, 15231528.Google Scholar
Arendt, D., Hausen, H., and Purschke, G. (2009). The “division of labour” model of eye evolution. Philos. Trans. R. Soc. Lond. B 364, 28092817.Google Scholar
Arendt, D., and Nübler-Jung, K. (1994). Inversion of dorsoventral axis? Nature 371, 26.Google Scholar
Arendt, D., and Nübler-Jung, K. (1996). Common ground plans in early brain development in mice and flies. BioEssays 18, 255259.Google Scholar
Arendt, D., and Nübler-Jung, K. (1997). Dorsal or ventral: similarities in fate maps and gastrulation patterns in annelids, arthropods and chordates. Mech. Dev. 61, 721.Google Scholar
Arendt, D., and Nübler-Jung, K. (1999). Comparison of early nerve cord development in insects and vertebrates. Development 126, 23092325.Google Scholar
Arendt, D., and Nübler-Jung, K. (1999). Rearranging gastrulation in the name of yolk: evolution of gastrulation in yolk-rich amniote eggs. Mech. Dev. 81, 322.Google Scholar
Arendt, D., Tessmar-Raible, K., Snyman, H., Dorresteijn, A.W., and Wittbrodt, J. (2004). Ciliary photoreceptors with a vertebrate-type opsin in an invertebrate brain. Science 306, 869871.Google Scholar
Arendt, D., Tosches, M.A., and Marlow, H. (2016). From nerve net to nerve ring, nerve cord and brain: evolution of the nervous system. Nat. Rev. Neurosci. 17, 6172.Google Scholar
Arendt, D., and Wittbrodt, J. (2001). Reconstructing the eyes of Urbilateria. Philos. Trans. R. Soc. Lond. B 356, 15451563.Google Scholar
Armitage, S.A.O., Freiburg, R.Y., Kurtz, J., and Bravo, I.G. (2012). The evolution of Dscam genes across the arthropods. BMC Evol. Biol. 12, Article 53. [See alsoGoogle ScholarGoogle Scholar
Armitage, S.A.O., Peuss, R., and Kurtz, J. (2015). Dscam and pancrustacean immune memory: a review of the evidence. Dev. Comp. Immunol. 48, 315323.Google Scholar
Armstrong, E.J., and Bischoff, J. (2004). Heart valve development: endothelial cell signaling and differentiation. Circ. Res. 95, 459470.Google Scholar
Armstrong, J.R., and Ferguson, M.W.J. (1995). Ontogeny of the skin and the transition from scar-free to scarring phenotype during wound healing in the pouch young of a marsupial, Monodelphis domestica. Dev. Biol. 169, 242260.CrossRefGoogle ScholarPubMed
Arnheiter, H. (1998). Eyes viewed from the skin. Nature 391, 632633.Google Scholar
Arnone, M.I., Rizzo, F., Annunciata, R., Cameron, R.A., Peterson, K.J., and Martínez, P. (2006). Genetic organization and embryonic expression of the ParaHox genes in the sea urchin S. purpuratus: insights into the relationship between clustering and colinearity. Dev. Biol. 300, 6373.Google Scholar
Arnosti, D.N., and Kulkarni, M.M. (2005). Transcriptional enhancers: intelligent enhanceosomes or flexible billboards? J. Cell. Biochem. 94, 890898.Google Scholar
Aronova, M.Z. (2009). Structural models of “simple” sense organs by the example of the first metazoa. J. Evol. Biochem. Physiol. 45, 179196.Google Scholar
Aronowicz, J., and Lowe, C.J. (2006). Hox gene expression in the hemichordate Saccoglossus kowalevskii and the evolution of deuterostome nervous systems. Integr. Comp. Biol. 46, 890901.Google Scholar
Arshavsky, V.Y. (2010). Vision: the retinoid cycle in Drosophila. Curr. Biol. 20, R96R98.Google Scholar
Arthur, W. (2002). The emerging conceptual framework of evolutionary developmental biology. Nature 415, 757764.Google Scholar
Arya, R., and White, K. (2015). Cell death in development: signaling pathways and core mechanisms. Semin. Cell Dev. Biol. 39, 1219.Google Scholar
Arzi, A., and Sobel, N. (2011). Olfactory perception as a compass for olfactory neural maps. Trends Cogn. Sci. 15, 537545.Google Scholar
Asadi-Pooya, A.A., Sharan, A., Nei, M., and Sperling, M.R. (2008). Corpus callosostomy. Epilepsy Behav. 13, 271278.Google Scholar
Ashburner, M. (2006). Won for All: How the Drosophila Genome Was Sequenced. Cold Spring Harbor Laboratory Press, Plainview, NY.Google Scholar
Ashkenazi, A., and Salvesen, G. (2014). Regulated cell death: signaling and mechanisms. Annu. Rev. Cell Dev. Biol. 30, 337356.CrossRefGoogle ScholarPubMed
Aso, Y., Hattori, D., Yu, Y., Johnston, R.M., Iyer, N.A., Ngo, T.-T.B., Dionne, H., Abbott, L.F., Axel, R., Tanimoto, H., and Rubin, G.M. (2014). The neuronal architecture of the mushroom body provides a logic for associative learning. eLife 3, e04577.Google Scholar
Aso, Y., Sitaraman, D., Ichinose, T., Kaun, K.R., Vogt, K., Belliart-Guérin, G., Plaçais, P.-Y., Robie, A.A., Yamagata, N., Schnaitmann, C., Rowell, W.J., Johnston, R.M., Ngo, T.-T.B., Chen, N., Korff, W., Nitabach, M.N., Heberlein, U., Preat, T., Branson, K.M., Tanimoto, H., and Rubin, G.M. (2014). Mushroom body output neurons encode valence and guide memory-based action selection in Drosophila. eLife 3, e04580.Google Scholar
Asteriti, S., Grillner, S., and Cangiano, L. (2015). A Cambrian origin for vertebrate rods. eLife 4, e07166.Google Scholar
Atkinson, P.J., Najarro, E.H., Sayyid, Z.N., and Cheng, A.G. (2015). Sensory hair development and regeneration: similarities and differences. Development 142, 15611571.Google Scholar
Audibert, A., Simon, F., and Gho, M. (2005). Cell cycle diversity involves differential regulation of Cyclin E activity in the Drosophila bristle cell lineage. Development 132, 22872297.Google Scholar
Austin, J.R. II, and Staehelin, L.A. (2011). Three-dimensional architecture of grana and stroma thylakoids of higher plants as determined by electron tomography. Plant Physiol. 155, 16011611.Google Scholar
Avasthi, P., and Marshall, W.F. (2012). Stages of ciliogenesis and regulation of ciliary length. Differentiation 83, S30S42.CrossRefGoogle ScholarPubMed
Averof, M. (2002). Arthropod Hox genes: insights on the evolutionary forces that shape gene functions. Curr. Opin. Genet. Dev. 12, 386392.Google Scholar
Aw, S., and Levin, M. (2008). What's left in asymmetry? Dev. Dyn. 237, 34533463.Google Scholar
Aw, S., and Levin, M. (2009). Is left-right asymmetry a form of planar cell polarity? Development 136, 355366.Google Scholar
Axelrod, J.D., and Bergmann, D.C. (2014). Coordinating cell polarity: heading in the right direction? Development 141, 32983302.Google Scholar
Aylsworth, A.S. (2001). Clinical aspects of defects in the determination of laterality. Am. J. Med. Genet. 101, 345355.Google Scholar
Azevedo, F.A.C., Carvalho, L.R.B., Grinberg, L.T., Farfel, J.M., Ferretti, R.E.L., Leite, R.E.P., Filho, W.J., Lent, R., and Herculano-Houzel, S. (2009). Equal numbers of neuronal and nonneuronal cells make the human brain an isometrically scaled-up primate brain. J. Comp. Neurol. 513, 532541.Google Scholar
Babcock, L.E. (1993). The right and the sinister. Nat. Hist. 102, #7, 3239.Google Scholar
Bachmann, A., and Knust, E. (1998). Dissection of cis-regulatory elements of the Drosophila gene Serrate. Dev. Genes Evol. 208, 346351.Google Scholar
Bachmann, A., and Knust, E. (1998). Positive and negative control of Serrate expression during early development of the Drosophila wing. Mech. Dev. 76, 6778.Google Scholar
Backfisch, B., Rajan, V.B.V., Fischer, R.M., Lohs, C., Arboleda, E., Tessmar-Raible, K., and Raible, F. (2013). Stable transgenesis in the marine annelid Platynereis dumerilii sheds new light on photoreceptor evolution. PNAS 110, #1, 193198.Google Scholar
Badano, J.L., and Katsanis, N. (2006). Life without centrioles: cilia in the spotlight. Cell 125, 12281230.Google Scholar
Bae, B.-I., Jayaraman, D., and Walsh, C.A. (2015). Genetic changes shaping the human brain. Dev. Cell 32, 423434.Google Scholar
Baek, J.H., Hatakeyama, J., Sakamoto, S., Ohtsuka, T., and Kageyama, R. (2006). Persistent and high levels of Hes1 expression regulate boundary formation in the developing central nervous system. Development 133, 24672476.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
Baguñà, J., and Garcia-Fernàndez, J. (2003). Evo-devo: the long and winding road. Int. J. Dev. Biol. 47, 705713.Google Scholar
Bailes, H.J., and Lucas, R.J. (2010). Melanopsin and inner retinal photoreception. Cell. Mol. Life Sci. 67, 99111.Google Scholar
Bailly, X., Reichert, H., and Hartenstein, V. (2013). The urbilaterian brain revisited: novel insights into old questions from new flatworm clades. Dev. Genes Evol. 223, 149157.CrossRefGoogle ScholarPubMed
Bainbridge, D. (2008). Beyond the Zonules of Zinn: A Fantastic Journey Through Your Brain. Harvard University Press, Cambridge, MA.Google Scholar
Baird, I.L. (1974). Some aspects of the comparative anatomy and evolution of the inner ear in submammalian vertebrates. Brain Behav. Evol. 10, 1136.Google Scholar
Bakalenko, N.I., Novikova, E.L., Nesterenko, A.Y., and Kulakova, M.A. (2013). Hox gene expression during postlarval development of the polychaete Alitta virens. EvoDevo 4, Article 13.Google Scholar
Baker, C.V.H., O'Neill, P., and McCole, R.B. (2008). Lateral line, otic and epibranchial placodes: developmental and evolutionary links? J. Exp. Zool. B. Mol. Dev. Evol. 310, 370383.CrossRefGoogle ScholarPubMed
Baker, N.E. (2001). Master regulatory genes; telling them what to do. BioEssays 23, 763766.Google Scholar
Baker, N.E., and Firth, L.C. (2011). Retinal determination genes function along with cell-cell signals to regulate Drosophila eye development. BioEssays 33, 538546.Google Scholar
Baker, R.E., Schnell, S., and Maini, P.K. (2009). Waves and patterning in developmental biology: vertebrate segmentation and feather bud formation as case studies. Int. J. Dev. Biol. 53, 783794.Google Scholar
Balaban, E. (2005). Brain switching: studying evolutionary behavioral changes in the context of individual brain development. Int. J. Dev. Biol. 49, 117124.Google Scholar
Balavoine, G. (2014). Segment formation in annelids: patterns, processes and evolution. Int. J. Dev. Biol. 58, 469483.Google Scholar
Balavoine, G., and Adoutte, A. (2003). The segmented Urbilateria: a testable scenario. Integr. Comp. Biol. 43, 137147.Google Scholar
Balavoine, G., de Rosa, R., and Adoutte, A. (2002). Hox clusters and bilaterian phylogeny. Mol. Phylogenet. Evol. 24, 366373.Google Scholar
Baliga, B., and Kumar, S. (2003). Apaf-1/cytochrome c apoptosome: an essential initiator of caspase activation or just a sideshow? Cell Death Differ. 10, 1618.Google Scholar
Baltimore, D. (2001). Our genome unveiled. Nature 409, 814816.Google Scholar
Balzeau, A., Gilissen, E., and Grimaud-Hervé, D. (2102). Shared pattern of endocranial shape asymmetries among great apes, anatomically modern humans, and fossil hominins. PLoS ONE 7, #1, e29581.Google Scholar
Barendregt, M., Harvey, B.M., Rokers, B., and Dumoulin, S.O. (2015). Transformation from a retinal to a cyclopean representation in human visual cortex. Curr. Biol. 25, 19821987.Google Scholar
Bargmann, C.I. (2006). Comparative chemosensation from receptors to ecology. Nature 444, 295301.Google Scholar
Barinaga, M. (1995). Focusing on the eyeless gene. Science 267, 17661767.Google Scholar
Barish, S., and Volkan, P.C. (2015). Mechanisms of olfactory receptor neuron specification in Drosophila. WIREs Dev. Biol. 4, 609621.Google Scholar
Barnes, J. (ed.) (1984). The Complete Works of Aristotle: The Revised Oxford Translation. Princeton University Press, Princeton, NJ.Google Scholar
Barr-Gillespie, P.-G. (2015). Assembly of hair bundles, an amazing problem for cell biology. Mol. Biol. Cell 26, #15, 27272732. [See alsoGoogle ScholarGoogle Scholar
Barretto, R.P.J., Gillis-Smith, S., Chandrashekar, J., Yarmolinsky, D.A., Schnitzer, M.J., Ryba, N.J.P., and Zuker, C.S. (2015). The neural representation of taste quality at the periphery. Nature 517, 373376.Google Scholar
Barrios, N., González-Pérez, E., Hernández, R., and Campuzano, S. (2015). The homeodomain Iroquois proteins control cell cycle progression and regulate the size of developmental fields. PLoS Genet. 11, #8, e1005463.Google Scholar
Bartlett, H., Veenstra, G.J.C., and Weeks, D.L. (2010). Examining the cardiac NK-2 genes in early heart development. Pediatr. Cardiol. 31, 335341.Google Scholar
Bartusiak, M. (2015). Rings, rings, rings: it's Saturn Giganticus! Nat. Hist. 123, #5, 1011.Google Scholar
Barucca, M., Canapa, A., and Bicscotti, M.A. (2016). An overview of Hox genes in Lophotrochozoa: evolution and functionality. J. Dev. Biol. 4, Article 4010012.CrossRefGoogle ScholarPubMed
Basch, M.L., Brown, R.M. II, Jen, H.-I., and Groves, A.K. (2016). Where hearing starts: the development of the mammalian cochlea. J. Anat. 228, 233254.Google Scholar
Basto, R., Lau, J., Vinogradova, T., Gardiol, A., Woods, C.G., Khodjakov, A., and Raff, J.W. (2006). Flies without centrioles. Cell 125, 13751386.Google Scholar
Basu, B., and Brueckner, M. (2008). Cilia: multifunctional organelles at the center of vertebrate left-right asymmetry. Curr. Top. Dev. Biol. 85, 151174.Google Scholar
Bates, C.J. (1995). Vitamin A. Lancet 345, 3135.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
Battelle, B.-A., Kempler, K.E., Saraf, S.R., Marten, C.E., Dugger, D.R. Jr., Speiser, D.I., and Oakley, T.H. (2015). Opsins in Limulus eyes: characterization of three visible lightsensitive opsins unique to and co-expressed in median eye photoreceptors and a peropsin/RGR that is expressed in all eyes. J. Exp. Biol. 218, 466479.Google Scholar
Baughman, K.W., McDougall, C., Cummins, S.F., Hall, M., Degnan, B.M., Satoh, N., and Shoguchi, E. (2014). Genomic organization of Hox and ParaHox clusters in the echinoderm, Acanthaster planci. Genesis 52, 952958.Google Scholar
Bayascas, J.B., Castillo, E., and Saló, E. (1998). Platyhelminthes have a Hox code differently activated during regeneration, with genes closely related to those of spiralian protostomes. Dev. Genes Evol. 208, 467473.Google Scholar
Baylies, M.K., Bate, M., and Ruiz Gomez, M. (1998). Myogenesis: a view from Drosophila. Cell 93, 921927.Google Scholar
Bayramli, X., and Fuss, S.H. (2012). Born to run: patterning the Drosophila olfactory system. Dev. Cell 22, 240241.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
Bear, A., and Monteiro, A. (2013). Both cell-autonomous mechanisms and hormones contribute to sexual development in vertebrates and insects. BioEssays 35, 725732.Google Scholar
Beatus, P., and Lendahl, U. (1998). Notch and neurogenesis. J. Neurosci. Res. 54, 125136.Google Scholar
Beaulé, V., Tremblay, S., Lafleur, L.-P., Tremblay, S., Lassonde, M., Lepage, J.-F., and Théoret, H. (2015). Cortical thickness in adults with agensis of the corpus callosum. Neuropsychologia 77, 359365.Google Scholar
Bechstedt, S., and Howard, J. (2007). Models of hair cell mechanotransduction. Curr. Top. Membr. 59, 399424.Google Scholar
Bechstedt, S., and Howard, J. (2008). Hearing mechanics: a fly in your ear. Curr. Biol. 18, R869R870.Google Scholar
Beckwith, E.J., and Yanovsky, M.J. (2014). Circadian regulation of gene expression: at the crossroads of transcriptional and post-transcriptional regulatory networks. Curr. Opin. Genet. Dev. 27, 3542.Google Scholar
Beisel, K.W., Wang-Lundberg, Y., Maklad, A., and Fritzsch, B. (2005). Development and evolution of the vestibular sensory apparatus of the mammalian ear. J. Vestib. Res. 15, 225241.CrossRefGoogle ScholarPubMed
Bell, A. (2008). The pipe and the pinwheel: is pressure an effective stimulus for the 9 + 0 primary cilium? Cell Biol. Int. 32, 462468.Google Scholar
Bell, A.T.A., and Niven, J.E. (2014). Individual-level, context-dependent handedness in the desert locust. Curr. Biol. 24, R382R383.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
Bell-Pedersen, D., Cassone, V.M., Earnest, D.J., Golden, S.S., Hardin, P.E., Thomas, T.L., and Zoran, M.J. (2005). Circadian rhythms from multiple oscillators: lessons from diverse organisms. Nat. Rev. Genet. 6, 544556.Google Scholar
Bellardita, C., and Kiehn, O. (2015). Phenotypic characterization of speed-associated gait changes in mice reveals modular organization of locomotor networks. Curr. Biol. 25, 14261436.Google Scholar
Belle, M.D.C., and Piggins, H.D. (2012). Circadian time redoxed. Science 337, 805806.Google Scholar
Bellen, H.J., and Yamamoto, S. (2015). Morgan's legacy: fruit flies and the functional annotation of conserved genes. Cell 163, 1214.Google Scholar
Belluscio, L., Lodovichi, C., Feinstein, P., Mombaerts, P., and Katz, L.C. (2002). Odorant receptors instruct functional circuitry in the mouse olfactory bulb. Nature 419, 296300.Google Scholar
Bengtson, S., Cunningham, J.A., Yin, C., and Donoghue, P.C.J. (2012). A merciful death for the “earliest bilaterian,” Vernanimalcula. Evol. Dev. 14, 421427.Google Scholar
Benito-Gutiérrez, È., and Arendt, D. (2009). CNS evolution: new insight from the mud. Curr. Biol. 19, R640R642.CrossRefGoogle ScholarPubMed
Benton, R., Sachse, S., Michnick, S.W., and Vosshall, L.B. (2006). Atypical membrane topology and heteromeric function of Drosophila odorant receptors in vivo. PLoS Biol. 4, #2, e20.Google Scholar
Benton, R., Vannice, K.S., Gomez-Diaz, C., and Vosshall, L.B. (2009). Variant ionotropic glutamate receptors as chemosensory receptors in Drosophila. Cell 136, 149162. [See alsoGoogle ScholarGoogle Scholar
Benzer, S. (1973). Genetic dissection of behavior. Sci. Am. 229, #6, 2437.Google Scholar
Bernards, A., and Hariharan, I.K. (2001). Of flies and men: studying human disease in Drosophila. Curr. Opin. Genet. Dev. 11, 274278.Google Scholar
Berry, J.A., Cervantes-Sandoval, I., Chakraborty, M., and Davis, R.L. (2015). Sleep facilitates memory by blocking dopamine neuron-mediated forgetting. Cell 161, 16561667.Google Scholar
Berson, D.M., Dunn, F.A., and Takao, M. (2002). Phototransduction by retinal ganglion cells that set the circadian clock. Science 295, 10701073.Google Scholar
Berthelsen, J., Kilstrup-Nielsen, C., Blasi, F., Mavilio, F., and Zappavigna, V. (1999). The subcellular localization of PBX1 and EXD proteins depends on nuclear import and export signals and is modulated by association with PREP1 and HTH. Genes Dev. 13, 946953.Google Scholar
Bertrand, N., Castro, D.S., and Guillemot, F. (2002). Proneural genes and the specification of neural cell types. Nat. Rev. Neurosci. 3, 517530.Google Scholar
Bertrand, S., and Escriva, H. (2011). Evolutionary crossroads in developmental biology: amphioxus. Development 138, 48194830.Google Scholar
Bertrand, V., Bisso, P., Poole, R.J., and Hobert, O. (2011). Notch-dependent induction of left/right asymmetry in C. elegans interneurons and motorneurons. Curr. Biol. 21, 12251231.Google Scholar
Bertucci, P., and Arendt, D. (2013). Somatic and visceral nervous systems – an ancient duality. BMC Biol. 11, Article 11.Google Scholar
Beshel, J., and Zhong, Y. (2013). Graded encoding of food odor value in the Drosophila brain. J. Neurosci. 33, #40, 1569315704.Google Scholar
Besson, C., Bernard, F., Corson, F., Roualt, H., Reynaud, E., Keder, A., Mazouni, K., and Schweisguth, F. (2015). Planar cell polarity breaks the symmetry of PAR protein distribution prior to mitosis in Drosophila sensory organ precursor cells. Curr. Biol. 25, 11041110.Google Scholar
Bettex, D.A., Prêtre, R., and Chassot, P.-G. (2014). Is our heart a well-designed pump? The heart along animal evolution. Eur. Heart J. 35, 23222332.Google Scholar
Beurg, M., Xiong, W., Zhao, B., Müller, U., and Fettiplace, R. (2015). Subunit determination of the conductance of hair-cell mechanotransducer channels. PNAS 112, #5, 15891594.Google Scholar
Bharti, K., Gasper, M., Ou, J., Brucato, M., Clore-Gronenborn, K., Pickel, J., and Arnheiter, H. (2012). A regulatory loop involving PAX6, MITF, and WNT signaling controls retinal pigment epithelium development. PLoS Genet. 8, #7, e1002757.Google Scholar
Bhat, K.M. (2005). Slit-Roundabout signaling neutralizes Netrin-Frazzled-mediated attractant cue to specify the lateral positioning of longitudinal axon pathways. Genetics 170, 149159.Google Scholar
Bhat, K.M., and Schedl, P. (1997). Requirement for engrailed and invected genes reveals novel regulatory interactions between engrailed/invected, patched, gooseberry and wingless during Drosophila neurogenesis. Development 124, 16751688.Google Scholar
Bhatia, S., Monahan, J., Ravi, V., Gautier, P., Murdoch, E., Brenner, S., Van Heyningen, V., Venkatesh, B., and Kleinjan, D.A. (2014). A survey of ancient conserved non-coding elements in the PAX6 locus reveals a landscape of interdigitated cis-regulatory archipelagos. Dev. Biol. 387, 214228.Google Scholar
Bier, E. (1997). Anti-neural-inhibition: a conserved mechanism for neural induction. Cell 89, 681684.Google Scholar
Bier, E. (2005). Drosophila, the golden bug, emerges as a tool for human genetics. Nat. Rev. Genet. 6, 923.Google Scholar
Bier, E. (2011). Evolution of development: diversified dorsoventral patterning. Curr. Biol. 21, R591R594.Google Scholar
Bier, E., and Bodmer, R. (2004). Drosophila, an emerging model for cardiac disease. Gene 342, 111.Google Scholar
Bier, E., and De Robertis, E.M. (2015). BMP gradients: a paradigm for morphogen-mediated developmental patterning. Science 348, aaa 5838.Google Scholar
Bilak, A., and Su, T.T. (2009). Regulation of Drosophila melanogaster pro-apoptotic gene hid. Apoptosis 14, 943949.Google Scholar
Billeter, J.-C., and Levine, J.D. (2013). Who is he and what is he to you? Recognition in Drosophila melanogaster. Curr. Opin. Neurobiol. 23, 1723.Google Scholar
Bisazza, A., Rogers, L.J., and Vallortigara, G. (1998). The origins of cerebral asymmetry: a review of evidence of behavioural and brain lateralization in fishes, reptiles and amphibians. Neurosci. Biobehav. Rev. 22, 411426.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
Bishopric, N.H. (2005). Evolution of the heart from bacteria to man. Ann. N. Y. Acad. Sci. 1047, 1329.Google Scholar
Black, B.L., and Olson, E.N. (1998). Transcriptional control of muscle development by Myocyte enhancer factor-2 (Mef2) proteins. Annu. Rev. Cell Dev. Biol. 14, 167196.Google Scholar
Black, D.L., and Zipursky, S.L. (2008). To cross or not to cross: alternately spliced forms of the Robo3 receptor regulate discrete steps in axonal midline crossing. Neuron 58, 297298.Google Scholar
Blair, S.S. (2009). Segmentation in animals. Curr. Biol. 18, R991R995.Google Scholar
Blair, S.S. (2014). Planar cell polarity: the importance of getting it backwards. Curr. Biol. 24, R835R838.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, J., Pandey, R., Wasser, M., and Udolph, G. (2011). Orthodenticle is necessary for survival of a cluster of clonally related dopaminergic neurons in the Drosophila larval and adult brain. Neural Dev. 6, Article 34.Google Scholar
Blaxter, M., and Sunnucks, P. (2011). Velvet worms. Curr. Biol. 21, R238R240.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
Blumer, M.J.F. (1996). Alterations of the eyes during ontogenesis in Aporrhais pespelecani (Mollusca, Caenogastropoda). Zoomorphology 116, 123131.Google Scholar
Bode, H. (2011). Axis formation in hydra. Annu. Rev. Genet. 45, 105117.Google Scholar
Bodmer, R. (1993). The gene tinman is required for specification of the heart and visceral muscles in Drosophila. Development 118, 719729.Google Scholar
Bodmer, R., and Venkatesh, T.V. (1998). Heart development in Drosophila and vertebrates: conservation of molecular mechanisms. Dev. Genet. 22, 181186.Google Scholar
Boekhoff-Falk, G. (2005). Hearing in Drosophila: development of Johnston's organ and emerging parallels to vertebrate ear development. Dev. Dyn. 232, 550558.Google Scholar
Boekhoff-Falk, G., and Eberl, D.F. (2014). The Drosophila auditory system. WIREs Dev. Biol. 3, 179191.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.Google Scholar
Bok, M.J., Porter, M.L., Place, A.R., and Cronin, T.W. (2014). Biological sunscreens tune polychromatic ultraviolet vision in mantis shrimp. Curr. Biol. 24, 16361642.Google Scholar
Bokolia, N.P., and Mishra, M. (2015). Hearing molecules, mechanism and transportation: modeled in Drosophila melanogaster. Dev. Neurobiol. 75, 109130.Google Scholar
Boorman, C.J., and Shimeld, S.M. (2002). The evolution of left-right asymmetry in chordates. BioEssays 24, 10041011.Google Scholar
Boorman, C.J., and Shimeld, S.M. (2002). Pitx homeobox genes in Ciona and amphioxus show left-right asymmetry is a conserved chordate character and define the ascidian adenohypophysis. Evol. Dev. 4, 354365.Google Scholar
Borges, R., Johnson, W.E., O'Brien, S.J., Vasconcelos, V., and Antunes, A. (2012). The role of gene duplication and unconstrained selective pressures in the melanopsin gene family evolution and vertebrate circadian rhythm regulation. PLoS ONE 7, #12, e52413.Google Scholar
Borst, A. (2009). Drosophila's view on insect vision. Curr. Biol. 19, R36R47.Google Scholar
Bouchard, M., de Caprona, D., Busslinger, M., Xu, P., and Fritzsch, B. (2010). Pax2 and Pax8 cooperate in mouse inner ear morphogenesis and innervation. BMC Dev. Biol. 10, Article 89.Google Scholar
Bowmaker, J.K. (2012). Evolution of the vertebrate eye. In Lazareva, O.F., Shimizu, T., and Wasserman, E.A. (eds.), How Animals See the World: Comparative Behavior, Biology, and Evolution of Vision. Oxford University Press, New York, NY, pp. 441472.Google Scholar
Bowmaker, J.K., and Hunt, D.M. (2006). Evolution of vertebrate visual pigments. Curr. Biol. 16, R484R489.Google Scholar
Boyd, J.L., Skove, S.L., Rouanet, J.P., Pilaz, L.-J., Bepler, T., Gordan, R., Wray, G.A., and Silver, D.L. (2015). Human-chimpanzee differences in a FZD8 enhancer alter cell-cycle dynamics in the developing neocortex. Curr. Biol. 25, 772779.Google Scholar
Boyden, E.A. (1977). Development and growth of the airways. In Hodson, W.A. (ed.), Development of the Lung. Marcel Dekker, New York, NY, pp. 335.Google Scholar
Bozza, T., Vassalli, A., Fuss, S.H., Zhang, J.-J., Weiland, B., Pacifico, R., Feinstein, P., and Mombaerts, P. (2009). Mapping of Class I and Class II odorant receptors to glomerular domains by two distinct types of olfactory sensory neurons in mice. Neuron 61, 220233.Google Scholar
Brachmann, C.B., and Cagan, R.L. (2003). Patterning the fly eye: the role of apoptosis. Trends Genet. 19, 9196.Google Scholar
Bradford, D.K., Cole, S.J., and Cooper, H.M. (2009). Netrin-1: diversity in development. Int. J. Biochem. Cell Biol. 41, 487493.Google Scholar
Bradshaw, J.L., and Rogers, L.J. (1993). The Evolution of Lateral Asymmetries, Language, Tool Use, and Intellect. Academic Press, New York, NY.Google Scholar
Brainard, D.H., and Hurlbert, A.C. (2015). Colour vision: understanding #TheDress. Curr. Biol. 25, R551R554.Google Scholar
Brand, T. (2003). Heart development: molecular insights into cardiac specification and early morphogenesis. Dev. Biol. 258, 119.Google Scholar
Brandler, W.M., Morris, A.P., Evans, D.M., Scerri, T.S., Kemp, J.P., Timpson, N.J., St. Pourcain, B., Smith, G.D., Ring, S.M., Stein, J.L., Monaco, A.P., Talcott, J.B., Fisher, S.E., Webber, C., and Paracchini, S. (2013). Common variants in left/right asymmetry genes and pathways are associated with relative hand skill. PLoS Genet. 9, #9, e1003751.Google Scholar
Brauckmann, S. (2012). Karl Ernst von Baer (1792–1876) and evolution. Int. J. Dev. Biol. 56, 653660.Google Scholar
Bray, D. (1998). Signaling complexes: biophysical constraints on intracellular communication. Annu. Rev. Biophys. Biomol. Struct. 27, 5975.Google Scholar
Brazeau, M.D., and Friedman, M. (2015). The origin and early phylogenetic history of jawed vertebrates. Nature 520, 490497.Google Scholar
Breer, H., Fleischer, J., and Strotmann, J. (2006). The sense of smell: multiple olfactory subsystems. Cell. Mol. Life Sci. 63, 14651475.Google Scholar
Bremner, A.J., and van Velzen, J. (2015). Sensorimotor control: retuning the body-world interface. Curr. Biol. 25, R159R161.Google Scholar
Brena, C., Chipman, A.D., Minelli, A., and Akam, M. (2006). Expression of trunk Hox genes in the centipede Strigamia maritima: sense and anti-sense transcripts. Evol. Dev. 8, 252265.Google Scholar
Brennan, P.A., and Zufall, F. (2006). Pheromonal communication in vertebrates. Nature 444, 308315.Google Scholar
Breslin, P.A.S., and Spector, A.C. (2008). Mammalian taste perception. Curr. Biol. 18, R148R155.Google Scholar
Brierly, A.S. (2014). Diel vertical migration. Curr. Biol. 24, R1074R1076.Google Scholar
Briscoe, A.D., and Chittka, L. (2001). The evolution of color vision in insects. Annu. Rev. Entomol. 46, 471510.Google Scholar
Briscoe, J., and Small, S. (2015). Morphogen rules: design principles of gradient-mediated embryo patterning. Development 142, 39964009.Google Scholar
Brites, D., Brena, C., Ebert, D., and Du Pasquier, L. (2013). More than one way to produce protein diversity: duplication and limited alternative splicing of an adhesion molecule gene in basal arthropods. Evolution 67, 29993011.Google Scholar
Brockes, J.P., and Kumar, A. (2005). Appendage regeneration in adult vertebrates and implications for regenerative medicine. Science 310, 19191922. [See alsoGoogle ScholarGoogle Scholar
Broihier, H.T., Kuzin, A., Zhu, Y., Odenwald, W., and Skeath, J.B. (2004). Drosophila homeodomain protein Nkx6 coordinates motoneuron subtype identity and axonogenesis. Development 131, 52335242.Google Scholar
Bronowski, J. (1956). Science and Human Values. Harper & Row, New York, NY.Google Scholar
Brooke, N.M., Garcia-Fernàndez, J., and Holland, P.W.H. (1998). The ParaHox gene cluster is an evolutionary sister of the Hox gene cluster. Nature 392, 920922.Google Scholar
Brookes, M. (2001). Fly: The Unsung Hero of Twentieth-Century Science. HarperCollins, New York, NY.Google Scholar
Brose, K., Bland, K.S., Wang, K.H., Arnott, D., Henzel, W., Goodman, C.S., Tessier-Lavigne, M., and Kidd, T. (1999). Slit proteins bind Robo receptors and have an evolutionarily conserved role in repulsive axon guidance. Cell 99, 795806.Google Scholar
Brown, D.D., Martz, S.N., Binder, O., Goetz, S.C., Price, B.M.J., Smith, J.C., and Conlon, F.L. (2005). Tbx5 and Tbx20 act synergistically to control vertebrate heart morphogenesis. Development 132, 553563.Google Scholar
Brown, N.A., and Lander, A. (1993). On the other hand … Nature 363, 303304.Google Scholar
Brown, N.A., McCarthy, A., and Wolpert, L. (1990). The development of handed asymmetry in aggregation chimeras of situs inversus mutant and wild-type mouse embryo. Development 110, 949954.Google Scholar
Brown, N.A., and Wolpert, L. (1990). The development of handedness in left/right asymmetry. Development 109, 19.Google Scholar
Brown, N.L., Patel, S., Brzezinski, J., and Glaser, T. (2001). Math5 is required for retinal ganglion cell and optic nerve formation. Development 128, 24972508.Google Scholar
Brown, S.D.M., Hardisty-Hughes, R.E., and Mburu, P. (2008). Quiet as a mouse: dissecting the molecular and genetic basis of hearing. Nat. Rev. Genet. 9, 277290.Google Scholar
Brown, T.M., Tsujimura, S.-i., Allen, A.E., Wynne, J., Bedford, R., Vickery, G., Vugler, A., and Lucas, R.J. (2012). Melanopsin-based brightness discrimination in mice and humans. Curr. Biol. 22, 11311141.Google Scholar
Brownstone, R.M., and Wilson, J.M. (2007). Strategies for delineating spinal locomotor rhythm-generating networks and the possible role of Hb9 interneurones in rhythmogenesis. Brain Res. Rev. 57, 6476.Google Scholar
Brunet, I., Di Nardo, A.A., Sonnier, L., Beurdeley, M., and Prochiantz, A. (2007). The topological role of homeoproteins in the developing central nervous system. Trends Neurosci. 30, 260267.Google Scholar
Brunet, T., Lauri, A., and Arendt, D. (2015). Did the notochord evolve from an ancient axial muscle? The axochord hypothesis. BioEssays 37, 836850.Google Scholar
Bryant, D.A., and Frigaard, N.-U. (2006). Prokaryotic photosynthesis and phototropy illuminated. Trends Microbiol. 14, 488496.Google Scholar
Bryant, P.J. (1993). The Polar Coordinate Model goes molecular. Science 259, 471472.Google Scholar
Bryant, S.V., and Gardiner, D.M. (2016). The relationship between growth and pattern formation. Regeneration 3, 103122.Google Scholar
Bryant, S.V., and Iten, L.E. (1976). Supernumerary limbs in amphibians: experimental production in Notophthalmus viridescens and a new interpretation of their formation. Dev. Biol. 50, 212234.Google Scholar
Bryantsev, A.L., and Cripps, R.M. (2009). Cardiac gene regulatory networks in Drosophila. Biochim. Biophys. Acta 1789, 343353.Google Scholar
Brzezinski, J.A., and Reh, T.A. (2015). Photoreceptor cell fate specification in vertebrates. Development 142, 32633273.Google Scholar
Bucher, G., Farzana, L., Brown, S.J., and Klingler, M. (2005). Anterior localization of maternal mRNAs in a short germ insect lacking bicoid. Evol. Dev. 7, 142149.Google Scholar
Buchon, N., Osman, D., David, F.P.A., Fang, H.Y., Boquete, J.-P., Deplancke, B., and Lemaitre, B. (2013). Morphological and molecular characterization of adult midgut compartmentalization in Drosophila. Cell Rep. 3, 17251738.Google Scholar
Buckingham, M., Meilhac, S., and Zaffran, S. (2005). Building the mammalian heart from two sources of myocardial cells. Nat. Rev. Genet. 6, 826835.Google Scholar
Budd, G.E. (2001). Why are arthropods segmented? Evol. Dev. 3, 332342.Google Scholar
Budd, G.E. (2008). The earliest fossil record of the animals and its significance. Philos. Trans. R. Soc. Lond. B 363, 14251434.Google Scholar
Budd, G.E. (2012). Cambrian nervous wrecks. Nature 490, 180181.Google Scholar
Budd, G.E. (2013). At the origin of animals: the revolutionary Cambrian fossil record. Curr. Genomics 14, 344354.Google Scholar
Budd, G.E., and Jackson, I.S.C. (2016). Ecological innovations in the Cambrian and the origins of the crown group phyla. Philos. Trans. R. Soc. Lond. B 371, 20150287.Google Scholar
Buhr, E., and Van Gelder, R.N. (2014). The making of the master clock. eLife 3, e04014.Google Scholar
Bullock, T.H., Orkand, R., and Grinnell, A. (1977). Introduction to Nervous Systems. W.H. Freeman, San Francisco, CA.Google Scholar
Burke, A.C., Nelson, C.E., Morgan, B.A., and Tabin, C. (1995). Hox genes and the evolution of vertebrate axial morphology. Development 121, 333346.Google Scholar
Burke, R.D. (2011). Deuterostome neuroanatomy and the body plan paradox. Evol. Dev. 13, 110115.Google Scholar
Burmester, T., and Hankeln, T. (2007). The respiratory proteins of insects. J. Insect Physiol. 53, 285294.Google Scholar
Burn, S.F., Boot, M.J., de Angelis, C., Doohan, R., Arques, C.G., Torres, M., and Hill, R.E. (2008). The dynamics of spleen morphogenesis. Dev. Biol. 318, 303311.Google Scholar
Buschbeck, E.K., and Friedrich, M. (2008). Evolution of insect eyes: tales of ancient heritage, deconstruction, reconstruction, remodeling, and recycling. Evol. Educ. Outreach 1, 448462.Google Scholar
Buschbeck, E.K., and Hauser, M. (2009). The visual system of male scale insects. Naturwissenschaften 96, 365374.Google Scholar
Butler, S.J., and Tear, G. (2007). Getting axons onto the right path: the role of transcription factors in axon guidance. Development 134, 439448.Google Scholar
Butts, T., Holland, P.W.H., and Ferrier, D.E.K. (2008). The Urbilaterian Super-Hox cluster. Trends Genet. 24, 259262.Google Scholar
Butts, T., Holland, P.W.H., and Ferrier, D.E.K. (2010). Ancient homeobox gene loss and the evolution of chordate brain and pharynx development: deductions from amphioxus gene expression. Proc. R. Soc. Lond. B 277, 33813389.Google Scholar
Byrne, M., Martinez, P., and Morris, V. (2016). Evolution of a pentameral body plan was not linked to translocation of anterior Hox genes: the echinoderm HOX cluster revisited. Evol. Dev. 18, 137143.Google Scholar
Cagan, R. (2009). Principles of Drosophila eye differentiation. Curr. Top. Dev. Biol. 89, 115135.Google Scholar
Cagan, R.L., and Ready, D.F. (1989). The emergence of order in the Drosophila pupal retina. Dev. Biol. 136, 346362.Google Scholar
Cajal, S.R., and Sánchez, D. (1915). Contribución al conocimiento de los centros nerviosos de los insectos. Trab. Lab. Invest. Biol. Univ. Madrid 13, 1167 + 162 plates.Google Scholar
Callaerts, P., Halder, G., and Gehring, W.J. (1997). Pax-6 in development and evolution. Annu. Rev. Neurosci. 20, 483532.Google Scholar
Callander, D.C., Alcorn, M.R., Birsoy, B., and Rothman, J.H. (2014). Natural reversal of left-right gut/gonad asymmetry in C. elegans males is independent of embryonic chirality. Genesis 52, 581587.Google Scholar
Cameron, R.A., Rowen, L., Nesbitt, R., Bloom, S., Rast, J.P., Berney, K., Arenas-Mena, C., Martinez, P., Lucas, S., Richardson, P.M., Davidson, E.H., Peterson, K.J., and Hood, L. (2006). Unusual gene order and organization of the sea urchin Hox cluster. J. Exp. Zool. B. Mol. Dev. Evol. 306, 4558.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
Campo-Paysaa, F., Marlétaz, F., Laudet, V., and Schubert, M. (2008). Retinoic acid signaling in development: tissue-specific functions and evolutionary origins. Genesis 46, 640656.Google Scholar
Cande, J., Prud'homme, B., and Gompel, N. (2013). Smells like evolution: the role of chemoreceptor evolution in behavioral change. Curr. Opin. Neurobiol. 23, 152158.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
Cañestro, C., and Postlethwait, J.H. (2007). Development of a chordate anterior–posterior axis without classical retinoic acid signaling. Dev. Biol. 305, 522538.Google Scholar
Cannon, J.T., Vellutini, B.C., Smith, J. III, Ronquist, F., Jondelius, U., and Hejnol, A. (2016). Xenacoelomorpha is the sister group to Nephrozoa. Nature 530, 8993. [See alsoGoogle ScholarGoogle Scholar
Canto-Soler, M.V., and Adler, R. (2006). Optic cup and lens development requires Pax6 expression in the early optic vesicle during a narrow time window. Dev. Biol. 294, 119132.Google Scholar
Cantore, E. (1977). Scientific Man: The Humanistic Significance of Science. Institute for Scientific Humanism, New York, NY.Google Scholar
Cao, N., Huang, Y., Zheng, J., Spencer, C.I., Zhang, Y., Fu, J.-D., Nie, B., Xie, M., Zhang, M., Wang, H., Ma, T., Xu, T., Shi, G., Srivastava, D., and Ding, S. (2016). Conversion of human fibroblasts into functional cardiomyocytes by small molecules. Science 352, 12161220.Google Scholar
Capdevila, J., Tsukui, T., Rodríguez Estaban, C., Zappavigna, V., and Izpisua Belmonte, J.C. (1999). Control of vertebrate limb outgrowth by the proximal factor Meis2 and distal antagonism of BMPs by Gremlin. Mol. Cell 4, 839849.Google Scholar
Capellini, T.D., Zappavigna, V., and Selleri, L. (2011). Pbx homeodomain proteins: TALEnted regulators of limb patterning and outgrowth. Dev. Dyn. 240, 10631086.Google Scholar
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
Capozzoli, N.J. (1995). Why are vertebrate nervous systems crossed? Med. Hypotheses 45, 471475.Google Scholar
Capozzoli, N.J. (1999). Why do we speak with the left hemisphere? Med. Hypotheses 52, 497503.Google Scholar
Carapuço, M., Nóvoa, A., Bobola, N., and Mallo, M. (2005). Hox genes specify vertebral types in the presomitic mesoderm. Genes Dev. 19, 21162121.Google Scholar
Carlson, B.M. (1994). Human Embryology and Developmental Biology. Mosby, St. Louis, MO.Google Scholar
Carmeliet, P., and Tessier-Lavigne, M. (2005). Common mechanisms of nerve and blood vessel wiring. Nature 436, 193200.Google Scholar
Carroll, S.B. (1990). Zebra patterns in fly embryos: activation of stripes or repression of interstripes? Cell 60, 916.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, NY.Google Scholar
Carroll, S.B. (2005). The origins of form. Nat. Hist. 114, #9, 5863.Google Scholar
Carroll, S.B., DiNardo, S., O'Farrell, P.H., White, R.A.H., and Scott, M.P. (1988). Temporal and spatial relationships between segmentation and homeotic gene expression in Drosophila embryos: distributions of the fushi tarazu, engrailed, Sex combs reduced, Antennapedia, and Ultrabithorax proteins. Genes Dev. 2, 350360.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 edn. Blackwell, Malden, MA.Google Scholar
Carson, H.L. (1983). Chromosomal sequences and interisland colonizations in Hawaiian Drosophila. Genetics 103, 465482.Google Scholar
Casey, B., and Hackett, B.P. (2000). Left-right axis malformations in man and mouse. Curr. Opin. Genet. Dev. 10, 257261.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 Scholar
Catania, K.C. (2011). Natural-born killer. Sci. Am. 304, #4, 8487.Google Scholar
Catela, C., Shin, M.M., and Dasen, J.S. (2015). Assembly and function of spinal circuits for motor control. Annu. Rev. Cell Dev. Biol. 31, 669698.Google Scholar
Cattenoz, P.B., and Giangrande, A. (2015). New insights in the clockwork mechanism regulating lineage specification: lessons from the Drosophila nervous system. Dev. Dyn. 244, 332341.Google Scholar
Caubit, X., Coré, N., Boned, A., Kerridge, S., Djabali, M., and Fasano, L. (2000). Vertebrate orthologues of the Drosophila region-specific patterning gene teashirt. Mech. Dev. 91, 445448.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
Cayouette, M., and Raff, M. (2002). Asymmetric segregation of Numb: a mechanism for neural specification from Drosophila to mammals. Nat. Neurosci. 5, 12651269.Google Scholar
Cehajic-Kapetanovic, J., Eleftheriou, C., Allen, A.E., Milosavljevic, N., Pienaar, A., Bedford, R., Davis, K.E., Bishop, P.N., and Lucas, R.J. (2015). Restoration of vision with ectopic expression of human rod opsin. Curr. Biol. 25, 21112122.Google Scholar
Certel, S.J., and Thor, S. (2004). Specification of Drosophila motoneuron identity by the combinatorial action of POU and LIM-HD factors. Development 131, 54295439.Google Scholar
Chacon-Heszele, M.F., and Chen, P. (2009). Mouse models for dissecting vertebrate planar cell polarity signaling in the inner ear. Brain Res. 1277, 130140.Google Scholar
Challis, R.C., Tian, H., Wang, J., He, J., Jiang, J., Chen, X., Yin, W., Connelly, T., Ma, L., Yu, R., Pluznick, J.L., Storm, D.R., Huang, L., Zhao, K., and Ma, M. (2015). An olfactory cilia pattern in the mammalian nose ensures high sensitivity to odors. Curr. Biol. 25, 25032512.Google Scholar
Chan, Y.-H., and Marshall, W.F. (2012). How cells know the size of their organelles. Science 337, 11861189.Google Scholar
Chan, Y.-M., and Jan, Y.N. (1999). Conservation of neurogenic genes and mechanisms. Curr. Opin. Neurobiol. 9, 582588.Google Scholar
Chandrashekar, J., Hoon, M.A., Ryba, N.J.P., and Zuker, C.S. (2006). The receptors and cells for mammalian taste. Nature 444, 288294.Google Scholar
Chang, C.-P., Neilson, J.R., Bayle, J.H., Gestwicki, J.E., Kuo, A., Stankunas, K., Graef, I.A., and Crabtree, G.R. (2004). A field of myocardial-endocardial NFAT signaling underlies heart valve morphogenesis. Cell 118, 649663.Google Scholar
Chang, D.C., and Reppert, S.M. (2001). The circadian clocks of mice and men. Neuron 29, 555558.Google Scholar
Chang, H., Cahill, H., Smallwood, P.M., Wang, Y., and Nathans, J. (2015). Identification of Astrotactin2 as a genetic modifier that regulates the global orientation of mammalian hair follicles. PLoS Genet. 11, #9, e1005532.Google Scholar
Chang, H.-Y., and Ready, D.F. (2000). Rescue of photoreceptor degeneration in rhodopsin-null Drosophila mutants by activated Rac1. Science 290, 19781980.Google Scholar
Chang, S., Johnston, R.J. Jr., and Hobert, O. (2003). A transcriptional regulatory cascade that controls left/right asymmetry in chemosensory neurons of C. elegans. Genes Dev. 17, 21232137.Google Scholar
Charité, J., de Graaff, W., Consten, D., Reijnen, M.J., Korving, J., and Deschamps, J. (1998). Transducing positional information to the Hox genes: critical interaction of cdx gene products with position-sensitive regulatory elements. Development 125, 43494358.Google Scholar
Charlton-Perkins, M., and Cook, T.A. (2010). Building a fly eye: terminal differentiation events of the retina, corneal lens, and pigmented epithelia. Curr. Top. Dev. Biol. 93, 129173.Google Scholar
Charpentier, M.S., and Conlon, F.L. (2013). Cellular and molecular mechanisms underlying blood vessel lumen formation. BioEssays 36, 251259.Google Scholar
Chatelin, L., Volovitch, M., Joliot, A.H., Perez, F., and Prochiantz, A. (1996). Transcription factor Hoxa-5 is taken up by cells in culture and conveyed to their nuclei. Mech. Dev. 55, 111117.Google Scholar
Chea, H.K., Wright, C.V., and Swalla, B.J. (2005). Nodal signaling and the evolution of deuterostome gastrulation. Dev. Dyn. 234, 269278.Google Scholar
Cheatle Jarvela, A.M., and Pick, L. (2016). Evo-devo: discovery of diverse mechanisms regulating development. Curr. Top. Dev. Biol. 117, 253274.Google Scholar
Chédotal, A. (2011). Further tales of the midline. Curr. Opin. Neurobiol. 21, 6875.Google Scholar
Cheesman, S.E., Layden, M.J., Ohlen, T.V., Doe, C.Q., and Eisen, J.S. (2004). Zebrafish and fly Nkx6 proteins have similar CNS expression patterns and regulate motoneuron formation. Development 131, 52215232.Google Scholar
Chen, C.-K., Woodruff, M.L., Chen, F.S., Shim, H., Cilluffo, M.C., and Fain, G.L. (2010). Replacing the rod with the cone transducin α subunit decreases sensitivity and accelerates response decay. J. Physiol. 588, 32313241.Google Scholar
Chen, H., Lun, Y., Ovchinnikov, D., Kokubo, H., Oberg, K.C., Pepicelli, C.V., Gan, L., Lee, B., and Johnson, R.L. (1998). Limb and kidney defects in Lmx1b mutant mice suggest an involvement of LMX1B in human nail patella syndrome. Nat. Genet. 19, 5155.Google Scholar
Chen, H., Xu, Z., Mei, C., Yu, D., and Small, S. (2012). A system of repressor gradients spatially organizes the boundaries of Bicoid-dependent target genes. Cell 149, 618629.Google Scholar
Chen, J.-Y. (2011). The origins and key innovations of vertebrates and arthropods. Paleoworld 20, 257278.Google Scholar
Chen, S.-K., Badea, T.C., and Hattar, S. (2011). Photoentrainment and pupillary light reflex are mediated by distinct populations of ipRGCs. Nature 476, 9295.Google Scholar
Chen, Z., Gore, B.B., Long, H., Ma, L., and Tessier-Lavigne, M. (2008). Alternative splicing of the Robo3 axon guidance receptor governs the midline switch from attraction to repulsion. Neuron 58, 325332.Google Scholar
Chen, Z., Zhu, J.-y., Fu, Y., Richman, A., and Han, Z. (2016). Wnt4 is required for ostia development in the Drosophila heart. Dev. Biol. 413, 188198.Google Scholar
Cheng, N., Tsunenari, T., and Yau, K.-W. (2009). Intrinsic light response of retinal horizontal cells of teleosts. Nature 460, 899903.Google Scholar
Cherry, S., Jin, E.J., Özel, M.N., Lu, Z., Agi, E., Wang, D., Jung, W.-H., Epstein, D., Meinertzhagen, I.A., Chan, C.-C., and Hiesinger, P.R. (2013). Charcot–Marie–Tooth 2B mutations in rab7 cause dosage-dependent neurodegeneration due to partial loss of function. eLife 2, e01064.Google Scholar
Chesler, A., and Firestein, S. (2008). Current views on odour receptors. Nature 452, 944.Google Scholar
Chiang, A.S., Lin, C.Y., Chuang, C.C., Chang, H.M., and Hsieh, C.-H. (2011). Three-dimensional reconstruction of brain-wide wiring networks in Drosophila at single-cell resolution. Curr. Biol. 21, 111.Google Scholar
Chien, C.-B. (1998). Why does the growth cone cross the road? Neuron 20, 36.Google Scholar
Chien, Y.-H., Keller, R., Kintner, C., and Shook, D.R. (2015). Mechanical strain determines the axis of planar polarity in ciliated epithelia. Curr. Biol. 25, 27742784.Google Scholar
Chilton, J.K. (2006). Molecular mechanisms of axon guidance. Dev. Biol. 292, 1324.Google Scholar
Chintapalli, V.R., Terhzaz, S., Wang, J., Al Bratty, M., Watson, D.G., Herzyk, P., Davies, S.A., and Dow, J.A.T. (2012). Functional correlates of positional and gender-specific renal asymmetry in Drosophila. PLoS ONE 7, #4, e32577.Google Scholar
Chipman, A.D. (2008). Annelids step forward. Evol. Dev. 10, 141142.Google Scholar
Chipman, A.D. (2010). Parallel evolution of segmentation by co-option of ancestral gene regulatory networks. BioEssays 32, 6070.Google Scholar
Chipman, A.D., Ferrier, D.E.K., Brena, C., Qu, J., Hughes, D.S.T., Schröder, R., Torres-Oliva, M., Znassi, N., Jiang, H., Almeida, F.C., Alonso, C.R., Apostolou, Z., Aqrawi, P., Arthur, W., Barna, J.C.J., Blankenburg, K.P., Brites, D., Capella-Gutiérrez, S., Coyle, M., Dearden, P.K., Du Pasquier, L., Duncan, E.J., Ebert, D., Eibner, C., Erikson, G., Evans, P.D., Extavour, C.G., Francisco, L., Gabaldón, T., Gillis, W.J., Goodwin-Horn, E.A., Green, J.E., Griffiths-Jones, S., Grimmelikhuijzen, C.J.P., Gubbala, S., Guigó, R., Han, Y., Hauser, F., Havlak, P., Hayden, L., Helbing, S., Holder, M., Hui, J.H.L., Hunn, J.P., Hunnekuhl, V.S., Jackson, L., Javaid, M., Jhangiani, S.N., Jiggins, F.M., Jones, T.E., Kaiser, T.S., Kalra, D., Kenny, N.J., Korchina, V., Kovar, C.L., Kraus, F.B., Lapraz, F., Lee, S.L., Lv, J., Mandapat, C., Manning, G., Mariotti, M., Mata, R., Mathew, T., Neumann, T., Newsham, I., Ngo, D.N., Ninova, M., Okwuonu, G., Ongeri, F., Palmer, W.J., Patil, S., Patraquim, P., Pham, C., Pu, L.-L., Putnam, N.H., Rabouille, C., Ramos, O.M., Rhodes, A.C., Robertson, H.E., Robertson, H.M., Ronshaugen, M., Rozas, J., Saada, N., Sánchez-Gracia, A., Scherer, S.E., Schurko, A.M., Siggens, K.W., Simmons, D., Stief, A., Stolle, E., Telford, M.J., Tessmar-Raible, K., Thornton, R., van der Zee, M., von Haeseler, A., Williams, J.M., Willis, J.H., Wu, Y., Zou, X., Lawson, D., Muzny, D.M., Worley, K.C., Gibbs, R.A., Akam, M., and Richards, S. (2014). The first myriapod genome sequence reveals conservative arthropod gene content and genome organisation in the centipede Strigamia maritima. PLoS Biol. 12, #11, e1002005.Google Scholar
Chisholm, A., and Tessier-Lavigne, M. (1999). Conservation and divergence of axon guidance mechanisms. Curr. Opin. Neurobiol. 9, 603615.Google Scholar
Chisholm, A.D., and Horvitz, H.R. (1995). Patterning of the Caenorhabditis elegans head region by the Pax-6 family member vab-3. Nature 377, 5255.Google Scholar
Choksi, S.P., Southall, T.D., Bossing, T., Edoff, K., de Wit, E., Fischer, B.E., van Steensel, B., Micklem, G., and Brand, A.H. (2006). Prospero acts as a binary switch between self-renewal and differentiation in Drosophila neural stem cells. Dev. Cell 11, 775789.Google Scholar
Chouhan, N.S., Wolf, R., Helfrich-Förster, C., and Heisenberg, M. (2015). Flies remember the time of day. Curr. Biol. 25, 16191624.Google Scholar
Chourrout, D., Delsuc, F., Chourrout, P., Edvardsen, R.B., Rentzsch, F., Renfer, E., Jensen, M.F., Zhu, B., de Jong, P., Steele, R.E., and Technau, U. (2006). Minimal ProtoHox cluster inferred from bilaterian and cnidarian Hox complements. Nature 442, 684687.Google Scholar
Chow, R.L., Altmann, C.R., Lang, R.A., and Hemmati-Brivanlou, A. (1999). Pax6 induces ectopic eyes in a vertebrate. Development 126, 42134222.Google Scholar
Christiaen, L., Jaszczyszyn, Y., Kerfant, M., Kano, S., Thermes, V., and Joly, J.-S. (2007). Evolutionary modification of mouth position in deuterostomes. Semin. Cell Dev. Biol. 18, 502511.Google Scholar
Christian, J.L. (2012). Morphogen gradients in development: from form to function. WIREs Dev. Biol. 1, 315.Google Scholar
Christodoulou, F., Raible, F., Tomer, R., Simakov, O., Trachana, K., Klaus, S., Snyman, H., Hannon, G.J., Bork, P., and Arendt, D. (2010). Ancient animal microRNAs and the evolution of tissue identity. Nature 463, 10841088.Google Scholar
Christoffels, V.M., Smits, G.J., Kispert, A., and Moorman, A.F.M. (2010). Development of the pacemaker tissues of the heart. Circ. Res. 106, 240254.Google Scholar
Chuang, C.-F., VanHoven, M.K., Fetter, R.D., Verselis, V.K., and Bargmann, C.I. (2007). An innexin-dependent cell network establishes left-right neuronal asymmetry in C. elegans. Cell 129, 787799.Google Scholar
Chuang, J.-Z., Zhao, Y., and Sung, C.-H. (2007). SARA-regulated vesicular targeting underlies formation of the light-sensing organelle in mammalian rods. Cell 130, 535547.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
Clandinin, T.R., and Giocomo, L.M. (2015). Internal compass puts flies in their place. Nature 521, 165166.Google Scholar
Clark, K.L., Yutzey, K.E., and Benson, D.W. (2006). Transcription factors and congenital heart defects. Annu. Rev. Physiol. 68, 97121.Google Scholar
Clarke, P.G.H. (1981). Chance, repetition, and error in the development of normal nervous systems. Perspect. Biol. Med. 25, 219.Google Scholar
Clarke, S.L., VanderMeer, J.E., Wenger, A.M., Schaar, B.T., Ahituv, N., and Bejerano, G. (2012). Human developmental enhancers conserved between deuterostomes and protostomes. PLoS Genet. 8, #8, e1002852.Google Scholar
Clayton, J.D., Kyriacou, C.P., and Reppert, S.M. (2001). Keeping time with the human genome. Nature 409, 829831.Google Scholar
Cloney, R.A. (1982). Ascidian larvae and the events of metamorphosis. Am. Zool. 22, 817826.Google Scholar
Clowney, E.J., LeGros, M.A., Mosley, C.P., Clowney, F.G., Markenskoff-Papadimitriou, E.C., Myllys, M., Barnea, G., Larabell, C.A., and Lomvardas, S. (2012). Nuclear aggregation of olfactory receptor genes governs their monogenic expression. Cell 151, 724737.Google Scholar
Cogan, G.B., Thesen, T., Carlson, C., Doyle, W., Devinsky, O., and Pesaran, B. (2014). Sensory-motor transformations for speech occur bilaterally. Nature 507, 9498.Google Scholar
Cohen, B., Simcox, A.A., and Cohen, S.M. (1993). Allocation of the thoracic imaginal primordia in the Drosophila embryo. Development 117, 597608.Google Scholar
Cohen, E.D., and Morrisey, E.E. (2008). A house with many rooms: how the heart got its chambers with foxn4. Genes Dev. 22, 706710.Google Scholar
Cohen, S.M., Brönner, G., Küttner, F., Jürgens, G., and Jäckle, H. (1989). Distal-less encodes a homoeodomain protein required for limb development in Drosophila. Nature 338, 432434.Google Scholar
Collett, T.S. (2002). Insect vision: controlling actions through optic flow. Curr. Biol. 12, R615R617.Google Scholar
Colley, N.J. (2000). Actin’ up with Rac1. Science 290, 19021903.Google Scholar
Collin, S.P., Knight, M.A., Davies, W.L., Potter, I.C., Hunt, D.M., and Trezise, A.E.O. (2003). Ancient color vision: multiple opsin genes in the ancestral vertebrates. Curr. Biol. 13, R864R865.Google Scholar
Collins, A.G., and Valentine, J.W. (2001). Defining phyla: evolutionary pathways to metazoan body plans. Evol. Dev. 3, 432442.Google Scholar
Collins, M.M., Baumholtz, A.I., Simard, A., Gregory, M., Cyr, D.G., and Ryan, A.K. (2015). Claudin-10 is required for relay of left–right patterning cues from Hensen's node to the lateral plate mesoderm. Dev. Biol. 401, 236248.Google Scholar
Collins, M.M., and Ryan, A.K. (2014). Are there conserved roles for the extracellular matrix, cilia, and junctional complexes in left-right patterning? Genesis 52, 488502.Google Scholar
Collu, G.M., and Mlodzik, M. (2015). Planar polarity: converting a morphogen gradient into cellular polarity. Curr. Biol. 25, R372R374.Google Scholar
Comai, G., and Tajbakhsh, S. (2014). Molecular and cellular regulation of skeletal myogenesis. Curr. Top. Dev. Biol. 110, 173.Google Scholar
Concha, M.L., Russell, C., Regan, J.C., Tawk, M., Sidi, S., Gilmour, D.T., Kapsimali, M., Sumoy, L., Goldstone, K., Amaya, E., Kimelman, D., Nicolson, T., Gründer, S., Gomperts, M., Clarke, J.D.W., and Wilson, S.W. (2003). Local tissue interactions across the dorsal midline of the forebrain establish CNS laterality. Neuron 39, 423438.Google Scholar
Conradt, B. (2009). Genetic control of programmed cell death during animal development. Annu. Rev. Genet. 43, 493523.Google Scholar
Cook, C.E., Chenevert, J., Larsson, T.A., Arendt, D., Houliston, E., and Lénárt, P. (2016). Old knowledge and new technologies allow rapid development of model organisms. Mol. Biol. Cell 27, 882887.Google Scholar
Cook, J.E. (1996). Spatial properties of retinal mosaics: an empirical evaluation of some existing measures. Vis. Neurosci. 13, 1530.Google Scholar
Cook, T. (2003). Cell diversity in the retina: more than meets the eye. BioEssays 25, 921925.Google Scholar
Cook, T., and Desplan, C. (2001). Photoreceptor subtype specification: from flies to humans. Semin. Cell Dev. Biol. 12, 509518.Google Scholar
Cooke, J. (2004). Developmental mechanism and evolutionary origin of vertebrate left/right asymmetries. Biol. Rev. 79, 377407.Google Scholar
Cooke, J. (2004). The evolutionary origins and significance of vertebrate left-right organisation. BioEssays 26, 413421.Google Scholar
Copeland, J.W.R., Nasiadka, A., Dietrich, B.H., and Krause, H.M. (1996). Patterning of the Drosophila embryo by a homeodomain-deleted Ftz polypeptide. Nature 379, 162165.Google Scholar
Copf, T., Schröder, R., and Averof, M. (2004). Ancestral role of caudal genes in axis elongation and segmentation. PNAS 101, 1771117715.Google Scholar
Corballis, M.C. (1991). The Lopsided Ape: Evolution of the Generative Mind. Oxford University Press, New York, NY.Google Scholar
Corballis, M.C. (2014). Left brain, right brain: facts and fantasies. PLoS Biol. 12, #1, e1001767.Google Scholar
Corballis, M.C., and Morgan, M.J. (1978). On the biological basis of human laterality. I. Evidence for a maturational left-right gradient. Behav. Brain Sci. 2, 261269.Google Scholar
Corballis, M.C., and Morgan, M.J. (1978). On the biological basis of human laterality. II. The mechanisms of inheritance. Behav. Brain Sci. 2, 270336.Google Scholar
Cordes, R., Schuster-Gossler, K., Serth, K., and Gossier, A. (2004). Specification of vertebral identity is coupled to Notch signalling and the segmentation clock. Development 131, 12211233.Google Scholar
Córdoba, S., and Estella, C. (2014). The bHLH-PAS transcription factor Dysfusion regulates tarsal joint formation in response to Notch activity during Drosophila leg development. PLoS Genet. 10, #10, e1004621.Google Scholar
Coren, S., and Porac, C. (1977). Fifty centuries of right-handedness: the historical record. Science 198, 631632.Google Scholar
Corless, J.M. (2012). Cone outer segments: a biophysical model of membrane dynamics, shape retention, and lamella formation. Biophys. J. 102, 26972705.Google Scholar
Costa, A., Sanchez-Guardado, L., Juniat, S., Gale, J.E., Daudet, N., and Henrique, D. (2015). Generation of sensory hair cells by genetic programming with a combination of transcription factors. Development 142, 19481959.Google Scholar
Coulter, D.E., Swaykus, E.A., Beran-Koehn, M.A., Goldberg, D., Wieschaus, E., and Schedl, P. (1990). Molecular analysis of odd-skipped, a zinc finger encoding segmentation gene with a novel pair-rule expression pattern. EMBO J. 8, 37953804.Google Scholar
Couso, J.P. (2009). Segmentation, metamerism and the Cambrian explosion. Int. J. Dev. Biol. 53, 13051316.Google Scholar
Coutelis, J.-B., Géminard, C., Spéder, P., Suzanne, M., Petzoldt, A.G., and Noselli, S. (2013). Drosophila left/right asymmetry establishment is controlled by the Hox gene Abdominal-B. Dev. Cell 24, 8997.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
Coutelis, J.B., Petzoldt, A.G., Spéder, P., Suzanne, M., and Noselli, S. (2008). Left-right asymmetry in Drosophila. Semin. Cell Dev. Biol. 19, 252262.Google Scholar
Couto, A., Alenius, M., and Dickson, B.J. (2005). Molecular, anatomical, and functional organization of the Drosophila olfactory system. Curr. Biol. 15, 15351547.Google Scholar
Couturier, L., Vodovar, N., and Schweisguth, F. (2012). Endocytosis by Numb breaks Notch symmetry at cytokinesis. Nat. Cell Biol. 14, 131139.Google Scholar
Cowan, W.M., Fawcett, J.W., O'Leary, D.D.M., and Stanfield, B.B. (1984). Regressive events in neurogenesis. Science 225, 12581265.Google Scholar
Craig, D.A., and Mary-Sasal, N. (2013). A detailed description of Simulium (Meilloniellum) adersi (Pomeroy) from Mayotte, Comoro islands, with comments on bionomics and biogeography (Diptera: Simuliidae). Zootaxa 3641, 129148.Google Scholar
Cripps, R.M., and Olson, E.N. (2002). Control of cardiac development by an evolutionarily conserved transcriptional network. Dev. Biol. 246, 1428.Google Scholar
Crocker, J., Tamori, Y., and Erives, A. (2008). Evolution acts on enhancer organization to fine-tune gradient threshold readouts. PLoS Biol. 6, #11, e263.Google Scholar
Cronin, T.W., Johnsen, S., Marshall, N.J., and Warrant, , , E.J. (2014). Visual Ecology. Princeton University Press, Princeton, NJ.Google Scholar
Cronin, T.W., and Porter, M.L. (2014). The evolution of invertebrate photopigments and photoreceptors. In Hunt, D.M., Hankins, M.W., Collin, S.P., and Marshall, N.J. (eds.), Evolution of Visual and Non-Visual Pigments. Springer, New York, NY, pp. 105135.Google Scholar
Croset, V., Rytz, R., Cummins, S.F., Budd, A., Brawand, D., Kaessmann, H., Gibson, T.J., and Benton, R. (2010). Ancient protostome origin of chemosensory ionotropic glutamate receptors and the evolution of insect taste and olfaction. PLoS Genet. 6, #8, e1001064.Google Scholar
Crouzet, S.M., Busch, N.A., and Ohla, K. (2015). Taste quality decoding parallels taste sensations. Curr. Biol. 25, 890896.Google Scholar
Crow, J.F. (2001). Shannon's brief foray into genetics. Genetics 159, 915917.Google Scholar
Cummins, M., Pueyo, J.I., Greig, S.A., and Couso, J.P. (2003). Comparative analysis of leg and antenna development in wild-type and homeotic Drosophila melanogaster. Dev. Genes Evol. 213, 319327.Google Scholar
Curcio, C.A., Sloan, K.R. Jr., Packer, O., Hendrickson, A.E., and Kalina, R.E. (1987). Distribution of cones in human and monkey retina: individual variability and radial asymmetry. Science 236, 579582.Google Scholar
Currie, K.W., Brown, D.D.R., Zhu, S., Xu, C.J., Voisin, V., Bader, G.D., and Pearson, B.J. (2016). HOX gene complement and expression in the planarian Schmidtea mediterranea. EvoDevo 7, Article 7.Google Scholar
Curto, G.G., Gard, C., and Ribes, V. (2015). Structures and properties of PAX linked regulatory networks architecting and pacing the emergence of neuronal diversity. Semin. Cell Dev. Biol. 44, 7586.Google Scholar
Cveki, A., and Piatigorsky, J. (1996). Lens development and crystallin gene expression: many roles for Pax-6. BioEssays 18, 621630.Google Scholar
da Fonseca, R.N., Lynch, J.A., and Roth, S. (2009). Evolution of axis formation: mRNA localization, regulatory circuits and posterior specification in non-model arthropods. Curr. Opin. Genet. Dev. 19, 404411.Google Scholar
Dacey, D.M., Liao, H.-W., Peterson, B.B., Robinson, F.R., Smith, V.C., Pokorny, J., Yau, K.-W., and Gamlin, P.D. (2005). Melanopsin-expressing ganglion cells in primate retina signal colour and irradiance and project to the LGN. Nature 433, 749754.Google Scholar
Dahmann, C., Oates, A.C., and Brand, M. (2011). Boundary formation and maintenance in tissue development. Nat. Rev. Genet. 12, 4355.Google Scholar
Dallos, P. (2008). Cochlear amplification, outer hair cells and prestin. Curr. Opin. Neurobiol. 18, 370376.Google Scholar
Dallos, P., and Fakler, B. (2002). Prestin, a new type of motor protein. Nat. Rev. Mol. Cell Biol. 3, 104111.Google Scholar
Dallos, P., Wu, X., Cheatham, M.A., Gao, J., Zheng, J., Anderson, C.T., Jia, S., Wang, X., Cheng, W.H.Y., Sengupta, S., He, D.Z.Z., and Zuo, J. (2008). Prestin-based outer hair cell motility is necessary for mammalian cochlear amplification. Neuron 58, 333339.Google Scholar
Dalton, R.P., and Lomvardas, S. (2015). Chemosensory receptor specificity and regulation. Annu. Rev. Neurosci. 38, 331349.Google Scholar
Dambly-Chaudière, C., Sapède, D., Soubiran, F., Decorde, K., Gompel, N., and Ghysen, A. (2003). The lateral line of zebrafish: a model system for the analysis of morphogenesis and neural development in vertebrates. Biol. Cell 95, 579587.Google Scholar
Damen, W.G.M. (2002). fushi tarazu: a Hox gene changes its role. BioEssays 24, 992995.Google Scholar
Damen, W.G.M. (2007). Evolutionary conservation and divergence of the segmentation process in arthropods. Dev. Dyn. 236, 13791391. [See alsoGoogle ScholarGoogle Scholar
Danchin, E.G.J., and Pontarotti, P. (2004). Stastical evidence for a more than 800-million-year-old evolutionarily conserved genomic region in our genome. J. Mol. Evol. 59, 587597.Google Scholar
Daneman, R., and Barres, B.A. (2005). The blood–brain barrier: lessons from moody flies. Cell 123, 912.Google Scholar
Darras, S., and Nishida, H. (2001). The BMP signaling pathway is required together with the FGF pathway for notochord induction in the ascidian embryo. Development 128, 26292638.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
Dasen, J.S., and Jessell, T.M. (2009). Hox networks and the origins of motor neuron diversity. Curr. Top. Dev. Biol. 88, 169200.Google Scholar
Dasen, J.S., Tice, B.C., Brenner-Morton, S., and Jessell, T.M. (2005). A Hox regulatory network establishes motor neuron pool identity and target-muscle connectivity. Cell 123, 477491.Google Scholar
Datta, R.R., Cruickshank, T., and Kumar, J.P. (2011). Differential selection within the Drosophila retinal determination network and evidence for functional divergence between paralog pairs. Evol. Dev. 13, 5871.Google Scholar
Davidson, B.P., and Tam, P.P.L. (2000). The node of the mouse embryo. Curr. Biol. 10, R617R619.Google Scholar
Davidson, E.H. (2001). Genomic Regulatory Systems: Development and Evolution. Academic Press, New York, NY.Google Scholar
Davidson, E.H. (2006). The Regulatory Genome: Gene Regulatory Networks in Development and Evolution. Academic Press, New York, NY.Google Scholar
Davies, W.I.L., Collin, S.P., and Hunt, D.M. (2012). Molecular ecology and adaptation of visual pigments in craniates. Mol. Ecol. 21, 31213158.Google Scholar
Davis, G.K., and Patel, N.H. (1999). The origin and evolution of segmentation. Trends Genet. 9, #12, M68M72.Google Scholar
Davis, G.K., and Patel, N.H. (2002). Short, long, and beyond: molecular and embryological approaches to insect segmentation. Annu. Rev. Entomol. 47, 669699. [See alsoGoogle ScholarGoogle Scholar
Davis, G.K., and Patel, N.H. (2003). Playing by pair-rules? BioEssays 25, 425429.Google Scholar
Davis, R.H. (2004). The age of model organisms. Nat. Rev. Genet. 5, 6976.Google Scholar
Davison, A., McDowell, G.S., Holden, J.M., Johnson, H.F., Koutsovoulos, G.D., Liu, M.M., Hulpiau, P., Van Roy, F., Wade, C.M., Banerjee, R., Yang, F., Chiba, S., Davey, J.W., Jackson, D.J., Levin, M., and Blaxter, M.L. (2016). Formin is associated with left-right asymmetry in the pond snail and the frog. Curr. Biol. 26, 654660.Google Scholar
Dawkins, R. (1998). Unweaving the Rainbow: Science, Delusion and the Appetite for Wonder. Houghton Mifflin, New York, NY.Google Scholar
de Brito Sanchez, G., and Giurfa, M. (2011). A comparative analysis of neural taste processing in animals. Philos. Trans. R. Soc. Lond. B 366, 21712180.Google Scholar
de Bruyne, M., and Warr, C.G. (2005). Molecular and cellular organization of insect chemosensory neurons. BioEssays 28, 2334.Google Scholar
de Celis, J.F., and Barrio, R. (2009). Regulation and function of Spalt proteins during animal development. Int. J. Dev. Biol. 53, 13851398.Google Scholar
de Ibarra, N.H., Vorobyev, M., and Menzel, R. (2014). Mechanisms, functions and ecology of colour vision in the honeybee. J. Comp. Physiol. A 200, 411433.Google Scholar
de Melo, J., Peng, G.-H., Chen, S., and Blackshaw, S. (2011). The Spalt family transcription factor Sall3 regulates the development of cone photoreceptors and retinal horizontal interneurons. Development 138, 23252336.Google Scholar
de Mendoza, A., Sebé-Pedrós, A., Sestak, M.S., Matejcic, M., Torruella, G., Domazet-Loso, T., and Ruiz-Trillo, I. (2013). Transcription factor evolution in eukaryotes and the assembly of the regulatory toolkit in multicellular lineages. PNAS 110, E4858E4866.Google Scholar
de Monasterio, F.M., Shein, S.J., and McCrane, E.P. (1981). Staining of blue-sensitive cones of the macaque retina by a fluorescent dye. Science 231, 12781281.Google Scholar
De Robertis, E.M. (2008). Evo-devo: variations on ancestral themes. Cell 132, 185195.Google Scholar
De Robertis, E.M. (2008). The molecular ancestry of segmentation mechanisms. PNAS 105, #43, 1641116412.Google Scholar
De Robertis, E.M., and Sasai, Y. (1996). A common plan for dorsoventral patterning in Bilateria. Nature 380, 3740.Google Scholar
de Rosa, R., Grenier, J.K., Andreeva, T., Cook, C.E., Adoutte, A., Akam, M., Carroll, S.B., and Balavoine, G. (1999). Hox genes in brachiopods and priapulids and protostome evolution. Nature 399, 772776.Google Scholar
de Rosa, R., Prud'homme, B., and Balavoine, G. (2005). caudal and even-skipped in the annelid Platynereis dumerilii and the ancestry of posterior growth. Evol. Dev. 7, 574587.Google Scholar
de Velasco, B., Erclik, T., Shy, D., Sclafani, J., Lipshitz, H., McInnes, R., and Hartenstein, V. (2007). Specification and development of the pars intercerebralis and pars lateralis, neuroendocrine command centers in the Drosophila brain. Development 302, 309323.Google Scholar
de Visser, J.A.G.M., and Krug, J. (2014). Empirical fitness landscapes and the predictability of evolution. Nat. Rev. Genet. 15, 480490.Google Scholar
Deans, M.R. (2013). A balance of form and function: planar polarity and development of the vestibular maculae. Semin. Cell Dev. Biol. 24, 490498.Google Scholar
Dearden, P., and Akam, M. (1999). Developmental evolution: axial patterning in insects. Curr. Biol. 9, R591R594.Google Scholar
Decker, R.S., Koyama, E., and Pacifici, M. (2014). Genesis and morphogenesis of limb synovial joints and articular cartilage. Matrix Biol. 39, 510.Google Scholar
Degnan, B.M., Vervoort, M., Larroux, C., and Richards, G.S. (2009). Early evolution of metazoan transcription factors. Curr. Opin. Genet. Dev. 19, 591599.Google Scholar
Dekkers, M.P.J., Nikoletopoulou, V., and Barde, Y.-A. (2013). Death of developing neurons: new insights and implications for connectivity. J. Cell Biol. 203, 385393. [See alsoGoogle ScholarGoogle 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
Delidakis, C., and Artavanis-Tsakonas, S. (1992). The Enhancer of split [E(spl)] locus of Drosophila encodes seven independent helix-loop-helix proteins. PNAS 89, 87318735.Google Scholar
Delling, M., Indzhykulian, A.A., Liu, X., Li, Y., Xie, T., Corey, D.P., and Clapham, D.E. (2016). Primary cilia are not calcium-responsive mechanosensors. Nature 531, 656660.Google Scholar
Delmas, P., Hao, J., and Rodat-Despoix, L. (2011). Molecular mechanisms of mechanotransduction in mammalian sensory neurons. Nat. Rev. Neurosci. 12, 139153.Google Scholar
Delsuc, F., Brinkmann, H., Chourrout, D., and Philippe, H. (2006). Tunicates and not cephalochordates are the closest relatives of vertebrates. Nature 439, 965968.Google Scholar
Delventhal, R., and Carlson, J.R. (2016). Bitter taste receptors confer diverse functions to neurons. eLife 5, e11181.Google Scholar
DeMaria, S., Berke, A.P., Van Name, E., Heravian, A., Ferreira, T., and Ngai, J. (2013). Role of a ubiquitously expressed receptor in the vertebrate olfactory system. J. Neurosci. 33, #38, 1523515247.Google Scholar
DeMaria, S., and Ngai, J. (2010). The cell biology of smell. J. Cell Biol. 191, 443452.Google Scholar
Demuth, J.P., and Wade, M.J. (2007). Maternal expression increases the rate of bicoid evolution by relaxing selective constraint. Genetica 129, 3743.Google Scholar
Denes, A.S., Jékely, G., Steinmetz, P.R.H., Raible, F., Snyman, H., Prud'homme, B., Ferrier, D.E.K., Balavoine, G., and Arendt, D. (2007). Molecular architecture of annelid nerve cord supports common origin of nervous system centralization in Bilateria. Cell 129, 277288.Google Scholar
Deng, H., Tan, T., and Yuan, L. (2015). Advances in the molecular genetics of non-syndromic polydactyly. Expert Rev. Mol. Med. 17, e18.CrossRefGoogle ScholarPubMed
Denver, R.J. (2008). Chordate metamorphosis: ancient control by iodothyronines. Curr. Biol. 18, R567R569.Google Scholar
Depetris-Chauvin, A., Galagovsky, D., and Grosjean, Y. (2015). Chemicals and chemoreceptors: ecologically relevant signals driving behavior in Drosophila. Front. Ecol. Evol. 3, Article 41.Google Scholar
Derelle, R., Lopez, P., Le Guyader, H., and Manuel, M. (2007). Homeodomain proteins belong to the ancestral molecular toolkit of eukaryotes. Evol. Dev. 9, 212219.Google Scholar
Deschamps, J. (2007). Ancestral and recently recruited global control of the Hox genes in development. Curr. Opin. Genet. Dev. 17, 422427.Google Scholar
Deschamps, J., and van Nes, J. (2005). Developmental regulation of the Hox genes during axial morphogenesis in the mouse. Development 132, 29312942.Google Scholar
Deutsch, J.S. (2004). Segments and parasegments in arthropods: a functional perspective. BioEssays 26, 11171125.Google Scholar
Deutsch, J.S. (2010). Homeosis and beyond. What is the function of the Hox genes? In Deutsch, J.S. (ed.), Hox Genes: Studies from the 20th to the 21st Century. Landes Bioscience, Austin, TX, pp. 155165.Google Scholar
Devenport, D. (2014). The cell biology of planar cell polarity. J. Cell Biol. 207, 171179.Google Scholar
Devenport, D., and Fuchs, E. (2008). Planar polarization in embryonic epidermis orchestrates global asymmetric morphogenesis of hair follicles. Nature Cell Biol. 10, 12571268.CrossRefGoogle ScholarPubMed
Dewey, E.B., Taylor, D.T., and Johnston, C.A. (2015). Cell fate decision making through oriented cell division. J. Dev. Biol. 3, 129157.Google Scholar
Dhouailly, D., Olivera-Martinez, I., Fliniaux, I., Missier, S., Viallet, J.P., and Thelu, J. (2004). Skin field formation: morphogenetic events. Int. J. Dev. Biol. 48, 8591.Google Scholar
Diaz-Benjumea, F.J., and Cohen, S.M. (1993). Interaction between dorsal and ventral cells in the imaginal disc directs wing development in Drosophila. Cell 75, 741752.CrossRefGoogle ScholarPubMed
Dicke, R.H., Peebles, P.J.E., Roll, P.G., and Wilkinson, D.T. (1965). Cosmic black-body radiation. Astrophys. J. 142, 414419.CrossRefGoogle 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
Dickinson, M.H. (2015). Motor control: how dragonflies catch their prey. Curr. Biol. 25, R232R234.Google Scholar
Dickinson, M.H., Farley, C.T., Full, R.J., Koehl, M.A.R., Kram, R., and Lehman, S. (2000). How animals move: an integrative view. Science 288, 100106.CrossRefGoogle ScholarPubMed
Dickson, B. (2001). Moving on. Science 291, 19101911.Google Scholar
Dickson, B.J., and Gilestro, G.F. (2006). Regulation of commissural axon pathfinding by Slit and its Robo receptors. Annu. Rev. Cell Dev. Biol. 22, 651675.CrossRefGoogle ScholarPubMed
Dimiccoli, M., Girard, B., Berthoz, A., and Bennequin, D. (2013). Striola magica: a functional explanation of otolith geometry. J. Comput. Neurosci. 35, 125154.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. Dyn. 244, 13571374.Google Scholar
Dissel, S., Angadi, V., Kirszenblat, L., Suzuki, Y., Donlea, J., Klose, M., Koch, Z., English, D., Winsky-Sommerer, R., van Swinderen, B., and Shaw, P.J. (2015). Sleep restores behavioral plasticity to Drosophila mutants. Curr. Biol. 25, 12701281.Google Scholar
Dissel, S., Hansen, C.N., Özkaya, Ö., Hemsley, M., Kyriacou, C.P., and Rosato, E. (2014). The logic of circadian organization in Drosophila. Curr. Biol. 24, 22572266.Google Scholar
Do, M.T.H., and Yau, K.-W. (2010). Intrinsically photosensitive retinal ganglion cells. Physiol. Rev. 90, 15471581.CrossRefGoogle ScholarPubMed
Domigan, C.K., Ziyad, S., and Iruela-Arispe, M.L. (2014). Canonical and noncanonical vascular endothelial growth factor pathways: new developments in biology and signal transduction. Arterioscler. Thromb. Vasc. Biol. 35, 3039.Google Scholar
Domínguez, L., González, A., and Moreno, N. (2015). Patterns of hypothalamic regionalization in amphibians and reptiles: common traits revealed by a genoarchitectonic approach. Front. Neuroanat. 9, Article 3.Google Scholar
Domyan, E.T., Kronenberg, Z., Infante, C.R., Vickrey, A.I., Stringham, S.A., Bruders, R., Guernsey, M.W., Park, S., Payne, J., Beckstead, R.B., Kardon, G., Menke, D.B., Yandell, M., and Shapiro, M.D. (2016). Molecular shifts in limb identity underlie development of feathered feet in two domestic avian species. eLife 5, e12115.Google Scholar
Donelson, N.C., and Sanyal, S. (2015). Use of Drosophila in the investigation of sleep disorders. Exp. Neurol. 274, 7279.Google Scholar
Dong, Y., Cirimotich, C.M., Pike, A., Chandra, R., and Dimopoulos, G. (2012). Anopheles NF-kB-regulated splicing factors direct pathogen-specific repertoires of the hypervariable pattern recognition receptor AgDscam. Cell Host Microbe 12, 521530.Google Scholar
Dong, Y., Taylor, H.E., and Dimopoulos, G. (2006). AgDscam, a hypervariable immunoglobulin domain-containing receptor of the Anopheles gambiae innate immune system. PLoS Biol. 4, #7, e229.Google Scholar
Donner, A.L., and Maas, R.L. (2004). Conservation and non-conservation of genetic pathways in eye specification. Int. J. Dev. Biol. 48, 743753.Google Scholar
Döring, C., Gosda, J., Tessmar-Raible, K., Hausen, H., Arendt, D., and Purschke, G. (2013). Evolution of clitellate phaosomes from rhabdomeric photoreceptor cells of polychaetes: a study in the leech Helobdella robusta (Annelida, Sedentaria, Clitellata). Front. Zool. 10, Article 52.Google Scholar
dos Reis, M., Thawornwattana, Y., Angelis, K., Telford, M.J., Donoghue, P.C.J., and Yang, Z. (2015). Uncertainty in the timing of origin of animals and the limits of precision in molecular timescales. Curr. Biol. 25, 29392950.Google Scholar
Douglas, R.H., Partridge, J.C., Dulai, K., Hunt, D., Mullineaux, C.W., Tauber, A.Y., and Hynninen, P.H. (1998). Dragon fish see using chlorophyll. Nature 393, 423424.Google Scholar
Doupé, D.P., and Jones, P.H. (2013). Cycling progenitors maintain epithelia while diverse cell types contribute to repair. BioEssays 35, 443451.Google Scholar
Dowling, J.E. (2012). The Retina: An Approachable Part of the Brain, 2nd edn. Harvard University Press, Cambridge, MA.Google Scholar
Downey, G., and Lende, D.H. (2012). Evolution and the brain. In Lende, D.H. and Downey, G. (eds.), The Encultured Brain: An Introduction to Neuroanthropology. MIT Press, Cambridge, MA., pp. 103137.Google Scholar
Driever, W. (2004). The Bicoid morphogen papers (II): account from Wolfgang Driever. Cell S116, S7S9.Google Scholar
Driever, W., and Nüsslein-Volhard, C. (1988). The bicoid protein determines position in the Drosophila embryo in a concentration-dependent manner. Cell 54, 95104.Google Scholar
Driever, W., and Nüsslein-Volhard, C. (1988). A gradient of bicoid protein in Drosophila embryos. Cell 54, 8393.Google Scholar
Driver, E.C., Sillers, L., Coate, T.M., Rose, M.F., and Kelley, M.W. (2013). The Atoh1-lineage gives rise to hair cells and supporting cells within the mammalian cochlea. Dev. Biol. 376, 8698.Google Scholar
Dror, A.A., and Avraham, K.B. (2010). Hearing impairment: a panoply of genes and functions. Neuron 68, 293308.Google Scholar
Duboc, V., Dufourcq, P., Blader, P., and Roussigné, M. (2015). Asymmetry of the brain: development and implications. Annu. Rev. Genet. 49, 647672.Google Scholar
Duboc, V., and Lepage, T. (2006). A conserved role for the nodal signaling pathway in the establishment of dorso-ventral and left-right axes in deuterostomes. J. Exp. Zool. B. Mol. Dev. Evol. 310, 4153.Google Scholar
Duboc, V., Röttinger, E., Lapraz, F., Besnardeau, L., and Lepage, T. (2005). Left-right asymmetry in the sea urchin embryo is regulated by Nodal signaling on the right side. Dev. Cell 9, 147158.CrossRefGoogle ScholarPubMed
Duboule, D. (2007). The rise and fall of Hox gene clusters. Development 134, 25492560.Google Scholar
Duboule, D., and Dollé, P. (1989). The structural and functional organization of the murine HOX gene family resembles that of Drosophila homeotic genes. EMBO J. 8, 14971505.Google Scholar
Dubowy, C.M., and Cavanaugh, D.J. (2014). Sleep: a neuropeptidergic wake-up call for flies. Curr. Biol. 24, R1092R1094.Google Scholar
Dubrulle, J., and Pourquié, O. (2004). Coupling segmentation to axis formation. Development 131, 57835793.Google Scholar
DuBuc, T.Q., Ryan, J.F., Shinzato, C., Satoh, N., and Martindale, M.Q. (2012). Coral comparative genomics reveal expanded Hox cluster in the cnidarian-bilaterian ancestor. Integr. Comp. Biol. 52, 835841.Google Scholar
Dudley, R., and Yanoviak, S.P. (2011). Animal aloft: the origins of aerial behavior and flight. Integr. Comp. Biol. 51, 926936.CrossRefGoogle ScholarPubMed
Duelli, P. (1978). An insect retina without microvilli in the male scale insect, Eriococcus sp. (Eriococcidae, Homoptera). Cell Tissue Res. 187, 417427.Google Scholar
Dugas-Ford, J., and Ragsdale, C.W. (2015). Levels of homology and the problem of the neocortex. Annu. Rev. Neurosci. 38, 351368.Google Scholar
Dulac, C. (2006). Charting olfactory maps. Science 314, 606607.Google Scholar
Dulai, K.S., von Dornum, M., Mollon, J.D., and Hunt, D.M. (1999). The evolution of trichromatic color vision by opsin gene duplication in New World and Old World primates. Genome Res. 9, 629638.Google Scholar
Duncan, J.S., and Fritzsch, B. (2012). Evolution of sound and balance perception: innovations that aggregate single hair cells into the ear and transform a gravistatic sensor into the organ of Corti. Anat. Rec. 295, 17601774.Google Scholar
Dunn, C.W., Giribet, G., Edgecombe, G.D., and Hejnol, A. (2014). Animal phylogeny and its evolutionary implications. Annu. Rev. Ecol. Evol. Syst. 45, 371395.Google Scholar
Dunn, C.W., and Ryan, J.F. (2015). The evolution of animal genomes. Curr. Opin. Genet. Dev. 35, 2532.Google Scholar
Dupé, V., and Lumsden, A. (2001). Hindbrain patterning involves graded responses to retinoic acid signalling. Development 128, 21992208.Google Scholar
Durston, A.J. (2012). Global posterior prevalence is unique to vertebrates: a dance to the music of time? Dev. Dyn. 241, 17991807.Google Scholar
Durston, A.J., Wacker, S., Bardine, N., and Jansen, H.J. (2012). Time space translation: a Hox mechanism for vertebrate A-P patterning. Curr. Genomics 13, 300307.Google Scholar
Duttke, S.H.C., Doolittle, R.F., Wang, Y.-L., and Kadonga, J.T. (2014). TRF2 and the evolution of the bilateria. Genes Dev. 28, 20712076.Google Scholar
Duverger, O., and Morasso, M.I. (2014). To grow or not to grow: hair morphogenesis and human genetic hair disorders. Semin. Cell Dev. Biol. 25–26, 2223.Google Scholar
Duysens, J., and Van de Crommert, H.W.A.A. (1998). Neural control of locomotion. Part 1: The central pattern generator from cats to humans. Gait Posture 7, 131141.Google Scholar
Dyer, M.A., Livesey, F.J., Cepko, C.L., and Oliver, G. (2003). Prox1 function controls progenitor cell proliferation and horizontal cell genesis in the mammalian retina. Nat. Genet. 34, 5358.Google Scholar
Eakin, R.M. (1965). Evolution of photoreceptors. Cold Spring Harb. Symp. Quant. Biol. 30, 363370.Google Scholar
Eakin, R.M. (1979). Evolutionary significance of photoreceptors: in retrospect. Am. Zool. 19, 647653.Google Scholar
Eakin, R.M., and Brandenberger, J.L. (1980). Unique eye of probable evolutionary significance. Science 211, 11891190.Google Scholar
Eakin, R.M., and Westfall, J.A. (1962). Fine structure of photoreceptors in amphioxus. J. Ultrastruct. Res. 6, 531539.Google Scholar
Eatock, R.A., and Hurley, K.M. (2003). Functional development of hair cells. Curr. Top. Dev. Biol. 57, 389447.Google Scholar
Eatock, R.A., and Songer, J.E. (2011). Vestibular hair cells and afferents: two channels for head motion signals. Annu. Rev. Neurosci. 34, 501534.Google Scholar
Eberl, D.F., and Boekhoff-Falk, G. (2007). Development of Johnston's organ in Drosophila. Int. J. Dev. Biol. 51, 679687.Google Scholar
Eberl, D.F., Hardy, R.W., and Kernan, M.J. (2000). Genetically similar transduction mechanisms for touch and hearing in Drosophila. J. Neurosci. 20, #16, 59815988.Google Scholar
Ecker, J.L., Dumitrescu, O.N., Wong, K.Y., Alam, N.M., Chen, S.-K., LeGates, T., Renna, J.M., Prusky, G.T., Berson, D.M., and Hattar, S. (2010). Melanopsin-expressing retinal ganglion-cell photoreceptors: cellular diversity and role in pattern vision. Neuron 67, 4960.Google Scholar
Eddison, M., Le Roux, I., and Lewis, J. (2000). Notch signaling in the development of the inner ear: lessons from Drosophila. PNAS 97, 1169211699.Google Scholar
Edelman, G.M. (1993). Neural Darwinism: selection and reentrant signaling in higher brain function. Neuron 10, 115125.Google Scholar
Edgar, L.G., Carr, S., Wang, H., and Wood, W.B. (2001). Zygotic expression of the caudal homolog pal-1 is required for posterior patterning in Caenorhabditis elegans embryogenesis. Dev. Biol. 229, 7188.Google Scholar
Edgar, R.S., Green, E.W., Zhao, Y., van Ooijen, G., Olmedo, M., Qin, X., Xu, Y., Pan, M., Valekunja, U.K., Feeney, K.A., Maywood, E.S., Hastings, M.H., Baliga, N.S., Merrow, M., Millar, A.J., Johnson, C.H., Kyriacou, C.P., O'Neill, J.S., and Reddy, A.B. (2012). Peroxiredoxins are conserved markers of circadian rhythms. Nature 485, 459464.Google Scholar
Effertz, T., Scharr, A.L., and Ricci, A.J. (2015). The how and why of identifying the hair cell mechano-electrical transduction channel. Pflugers Arch. 467, 7384.Google Scholar
Effertz, T., Wiek, R., and Göpfert, M.C. (2011). NompC TRP channel is essential for Drosophila sound receptor function. Curr. Biol. 21, 592597.Google Scholar
Egger, B., Chell, J.M., and Brand, A.H. (2008). Insights into neural stem cell biology from flies. Philos. Trans. R. Soc. Lond. B 363, 3956.Google Scholar
Eichmann, A., and Thomas, J.-L. (2013). Molecular parallels between neural and vascular development. Cold Spring Harb. Perspect. Med. 3, a006551.Google Scholar
Elliott, D.A., Solloway, M.J., Wise, N., Biben, C., Costa, M.W., Furtado, M.B., Lange, M., Dunwoodie, S., and Harvey, R.P. (2006). A tyrosine-rich domain within homeodomain transcription factor Nkx2–5 is an essential element in the early cardiac transcriptional regulatory machinery. Development 133, 13111322.Google Scholar
Elofsson, R., and Dahl, E. (1970). The optic neuropiles and chiasmata of Crustacea. Z. Zellforsch. Mikrosk Anat. 107, 343360.Google Scholar
Elsaesser, R., and Paysan, J. (2007). The sense of smell, its signalling pathways, and the dichotomy of cilia and microvilli in olfactory sensory cells. BMC Neurosci. 8 (Suppl. 3), Article S1.Google Scholar
Enard, W. (2015). Human evolution: enhancing the brain. Curr. Biol. 25, R409R430.Google Scholar
Enjin, A., Zaharieva, E.E., Frank, D.D., Mansourian, S., Suh, G.B., Gallio, M., and Stensmyr, M.C. (2016). Humidity sensing in Drosophila. Curr. Biol. 26, 13521358.Google Scholar
Epstein, M., Pillemer, G., Yelin, R., Yisraeli, J.K., and Fainsod, A. (1997). Patterning of the embryo along the anterior-posterior axis: the role of the caudal genes. Development 124, 38053814.Google Scholar
Erclik, T., Hartenstein, V., Lipshitz, H.D., and McInnes, R.R. (2008). Conserved role of the Vsx genes supports a monophyletic origin for bilaterian visual systems. Curr. Biol. 18, 12781287.Google Scholar
Erclik, T., Hartenstein, V., McInnes, R.R., and Lipshitz, H.D. (2009). Eye evolution at high resolution: the neuron as a unit of homology. Dev. Biol. 332, 7079.Google Scholar
Erickson, T., French, C.R., and Waskiewicz, A.J. (2010). Meis1 specifies positional information in the retina and tectum to organize the zebrafish visual system. Neural Dev. 5, Article 22.Google Scholar
Eriksson, B.J., Larson, E.T., Thörnqvist, P.-O., Tait, N.N., and Budd, G.E. (2005). Expression of engrailed in the developing brain and appendages of the onychophoran Euperipatoides kanangrensis (Reid). J. Exp. Zool. B. Mol. Dev. Evol. 304, 220228.Google Scholar
Ernsberger, U. (2015). Can the “neuron theory” be complemented by a universal mechanism for generic neuronal differentiation? Cell Tissue Res. 359, 343384.Google Scholar
Ernst, O.P., Lodowski, D.T., Elstner, M., Hegemann, P., Brown, L.S., and Kandori, H. (2014). Microbial and animal rhodopsins: structures, functions, and molecular mechanisms. Chem. Rev. 114, 126163.Google Scholar
Erskine, L., and Herrera, E. (2007). The retinal ganglion cell axon's journey: insights into molecular mechanims of axon guidance. Dev. Biol. 308, 114.Google Scholar
Erwin, D.H., and Davidson, E.H. (2002). The last common bilaterian ancestor. Development 129, 30213032.Google Scholar
Esteves, F.F., Springhorn, A., Kague, E., Taylor, E., Pyrowolakis, G., Fisher, S., and Bier, E. (2014). BMPs regulate msx gene expression in the dorsal neuroectoderm of Drosophila and vertebrates by distinct mechanisms. PLoS Genet. 10, #9, e1004625.Google Scholar
Estévez-Calvar, N., Romero, A., Figueras, A., and Novoa, B. (2013). Genes of the mitochondrial apoptotic pathway in Mytilus galloprovincialis. PLoS ONE 8, #4, e61502.Google Scholar
Etchberger, J.F., Flowers, E.B., Poole, R.J., Bashllari, E., and Hobert, O. (2009). Cis-regulatory mechanisms of left/right asymmetric neuron-subtype specification in C. elegans. Development 136, 147160.Google Scholar
Evans, C.J., Hartenstein, V., and Banerjee, U. (2003). Thicker than blood: conserved mechanisms in Drosophila and vertebrate hematopoiesis. Dev. Cell 5, 673690.Google Scholar
Evans, C.J., Sinenko, S.A., Mandal, L., Martinez-Agosto, J.A., Hartenstein, V., and Benerjee, U. (2008). Genetic dissection of hematopoiesis using Drosophila as a model system. Adv. Dev. Biol. 18, 259299.Google Scholar
Evans, S.M. (1999). Vertebrate tinman homologues and cardiac differentiation. Semin. Cell Dev. Biol. 10, 7383.Google Scholar
Evans, T.A., and Bashaw, G.J. (2010). Axon guidance at the midline: of mice and flies. Curr. Opin. Neurobiol. 20, 7985.Google Scholar
Evans, T.A., Santiago, C., Arbeille, E., and Bashaw, G.J. (2015). Robo2 acts in trans to inhibit Slit-Robo1 repulsion in pre-crossing commissural axons. eLife 4, e08407.Google Scholar
Ewing, T. (1993). Genetic “master switch” for left-right symmetry found. Science 260, 624625.Google Scholar
Extavour, C.G., and Akam, M. (2003). Mechanisms of germ cell specification across the metazoans: epigenesis and preformation. Development 130, 58695884.Google Scholar
Extavour, C.G.M. (2007). Evolution of the bilaterian germ line: lineage origin and modulation of specification mechanisms. Integr. Comp. Biol. 47, 770785.Google Scholar
Extavour, C.G.M. (2008). Urbisexuality: the evolution of bilaterian germ cell specification and reproductive systems. In Minelli, A. and Fusco, G. (eds.), Evolving Pathways: Key Themes in Evolutionary Developmental Biology. Cambridge University Press, New York, NY, pp. 321342.Google Scholar
Ezan, J., and Montcouquiol, M. (2013). Revisiting planar cell polarity in the inner ear. Semin. Cell Dev. Biol. 24, 499506. [See alsoGoogle ScholarGoogle Scholar
Fabri, M., and Polonara, G. (2013). Functional topography of human corpus callosum: an fMRI mapping study. Neural Plast. 2013, Article 251308.Google Scholar
Fähling, M., Mrowka, R., Steege, A., Kirschner, K.M., Benko, E., Förstera, B., Persson, P.B., Thiele, B.J., Meier, J.C., and Scholz, H. (2009). Translational regulation of the human achaete-scute homologue-1 by fragile X mental retardation protein. J. Biol. Chem. 284, #7, 42554266.Google Scholar
Fain, G.L. (2003). Sensory Transduction. Sinauer, Sunderland, MA.Google Scholar
Fain, G.L. (2016). Phototransduction: making the chromophore to see through the murk. Curr. Biol. 25, R1126R1142.Google Scholar
Fain, G.L., Hardie, R., and Laughlin, S.B. (2010). Phototransduction and the evolution of photoreceptors. Curr. Biol. 20, R114R124.Google Scholar
Fan, J.-Y., Preuss, F., Muskus, M.J., Bjes, E.S., and Price, J.L. (2009). Drosophila and vertebrate Casein Kinase Iδ exhibits evolutionary conservation of circadian function. Genetics 181, 139152.Google Scholar
Fanto, M., and McNeill, H. (2004). Planar polarity from flies to vertebrates. J. Cell Sci. 117, 527533.Google Scholar
Farah, M.J. (2015). An ethics toolbox for neurotechnology. Neuron 86, 3437.Google Scholar
Farfán, C., Shigeno, S., Nödl, M.-T., and de Couet, H.G. (2009). Developmental expression of apterous/Lhx2/9 in the sepiolid squid Euprymna scolopes supports an ancestral role in neural development. Evol. Dev. 11, 354362.Google Scholar
Farnum, C.E., and Wilsman, N.J. (2011). Axonemal positioning and orientation in three-dimensional space for primary cilia: what is known, what is assumed, and what needs clarification. Dev. Dyn. 240, 24052431.Google Scholar
Farrar, N.R., and Spencer, G.E. (2008). Pursuing a “turning point” in growth cone research. Dev. Biol. 318, 102111.Google Scholar
Farris, S.M. (2015). Evolution of brain elaboration. Philos. Trans. R. Soc. Lond. B 370, 20150054.Google Scholar
Faurie, C., and Raymond, M. (2004). Handedness frequency over more than ten thousand years. Proc. R. Soc. Lond. B 271, S43S45.Google Scholar
Fears, S.C., Scheibel, K., Abaryan, Z., Lee, C., Service, S.K., Jorgensen, M.J., Fairbanks, L.A., Cantor, R.M., Freimer, N.B., and Woods, R.P. (2011). Anatomic brain asymmetry in vervet monkeys. PLoS ONE 6, #12, e28243.Google Scholar
Feiler, R., Bjornson, R., Kirschfeld, K., Mismer, D., Rubin, G.M., Smith, D.P., Socolich, M., and Zuker, C.S. (1992). Ectopic expression of ultraviolet-rhodopsins in the blue photoreceptor cells of Drosophila: visual physiology and photochemistry of transgenic animals. J. Neurosci. 12, #10, 38623868.Google Scholar
Fekete, D.M., and Wu, D.K. (2002). Revisiting cell fate specification in the inner ear. Curr. Opin. Neurobiol. 12, 3542.Google Scholar
Ferguson, L., Marlétaz, F., Carter, J.-M., Taylor, W.R., Gibbs, M., Breuker, C.J., and Holland, P.W.H. (2014). Ancient expansion of the Hox cluster in Lepidoptera generated four homeobox genes implicated in extra-embryonic tissue formation. PLoS Genet. 10, #10, e1004698.Google Scholar
Fernald, R.D. (2000). Evolution of eyes. Curr. Opin. Neurobiol. 10, 444450.Google Scholar
Fernald, R.D. (2004). Evolving eyes. Int. J. Dev. Biol. 48, 701705.Google Scholar
Fernald, R.D. (2006). Casting a genetic light on the evolution of eyes. Science 313, 19141918.Google Scholar
Ferreira, L.M.R., and Mostajo-Radji, M.A. (2013). How induced pluripotent stem cells are redefining personalized medicine. Gene 520, 16.Google Scholar
Ferreira, T., Wilson, S.R., Choi, Y.G., Risso, D., Dudoit, S., Speed, T.P., and Ngai, J. (2014). Silencing of odorant receptor genes by G protein βγ signaling ensures the expression of one odorant receptor per olfactory sensory neuron. Neuron 81, 847859.Google Scholar
Ferrier, D.E.K. (2012). Evolutionary crossroads in developmental biology: annelids. Development 139, 25432653.Google Scholar
Ferrier, D.E.K. (2016). The origin of the Hox/ParaHox genes, the Ghost Locus hypothesis and the complexity of the first animal. Brief. Funct. Genomics 15, 19 (doi 10.1093).Google Scholar
Ferrier, D.E.K., and Minguillón, C. (2003). Evolution of the Hox/ParaHox gene clusters. Int. J. Dev. Biol. 47, 605611.Google Scholar
Fettiplace, R., and Kim, K.X. (2014). The physiology of mechanoelectrical transduction channels in hearing. Physiol. Rev. 94, 951986.Google Scholar
Feuda, R., Hamilton, S.C., McInerney, J.O., and Pisani, D. (2012). Metazoan opsin evolution reveals a simple route to animal vision. PNAS 109, #46, 1886818872.Google Scholar
Finlay, B.L. (2008). The developing and evolving retina: using time to organize form. Brain Res. 1192, 516.Google Scholar
Fiore, V.G., Dolan, R.J., Strausfeld, N.J., and Hirth, F. (2015). Evolutionarily conserved mechanisms for the selection and maintenance of behavioural activity. Philos. Trans. R. Soc. Lond. B 370, 20150053.Google Scholar
Firestein, S. (2001). How the olfactory system makes sense of scents. Nature 413, 211218.Google Scholar
Fishman, M.C., and Olson, E.N. (1997). Parsing the heart: genetic modules for organ assembly. Cell 91, 153156.Google Scholar
Fitch, D.H.A., and Sudhaus, W. (2002). One small step for worms, one giant leap for “Bauplan”? Evol. Dev. 4, 243246.Google Scholar
FitzGerald, B.J., Richardson, K., and Wesson, D.W. (2014). Olfactory tubercle stimulation alters odor preference behavior and recruits forebrain reward and motivational centers. Front. Behav. Neurosci. 8, Article 81.Google Scholar
Flanagan, J.G., and Van Vactor, D. (1998). Through the looking glass: axon guidance at the midline choice point. Cell 92, 429432.Google Scholar
Fleming, A., Kishida, M.G., Kimmel, C.B., and Keynes, R.J. (2015). Building the backbone: the development and evolution of vertebral patterning. Development 142, 17331744.Google Scholar
Florio, M., Albert, M., Taverna, E., Namba, T., Brandl, H., Lewitus, E., Haffner, C., Sykes, A., Wong, F.K., Peters, J., Guhr, E., Klemroth, S., Prüfer, K., Kelso, J., Naumann, R., Nüsslein, I., Dahl, A., Lachmann, R., Pääbo, S., and Huttner, W.B. (2015). Human-specific gene ARHGAP11B promotes basal progenitor amplification and neocortex expansion. Science 347, 14651470.Google Scholar
Foltys, H., Krings, T., Meister, I.G., Sparing, R., Boroojerdi, B., Thron, A., and Topper, R. (2003). Motor representation in patients rapidly recovering after stroke: a functional magnetic resonance imaging and transcranial magnetic stimulation study. Clin. Neurophysiol. 114, 24042415.Google Scholar
Fomenou, M.D., Scaal, M., Stockdale, F.E., Christ, B., and Huang, R. (2005). Cells of all somitic compartments are determined with respect to segmental identity. Dev. Dyn. 233, 13861393.Google Scholar
Fontdevila, A. (2011). The Dynamic Genome: A Darwinian Approach. Oxford University Press, New York, NY.Google Scholar
Formosa-Jordan, P., and Ibañes, M. (2014). Competition in Notch signaling with cis enriches cell fate decisions. PLoS ONE 9, #4, e95744.Google Scholar
Foronda, D., Martin, P., and Sánchez-Herrero, E. (2012). Drosophila Hox and sex-determination genes control segment elimination through EGFR and extramacrochaete activity. PLoS Genet. 8, #8, e1002874.Google Scholar
Forouhar, A.S., Liebling, M., Hickerson, A., Nasiraei-Moghaddam, A., Tsai, H.-J., Hove, J.R., Fraser, S.E., Dickenson, M.E., and Gharib, M. (2006). The embryonic vertebrate heart tube is a dynamic suction pump. Science 312, 751753.Google Scholar
Fortey, R. (2012). Horseshoe Crabs and Velvet Worms: The Story of the Animals and Plants That Time Has Left Behind. Knopf, New York, NY.Google Scholar
Fortini, M.E. (2009). Notch signaling: the core pathway and its posttranslational regulation. Dev. Cell 16, 633647.Google Scholar
Fortunato, S.A.V., Adamski, M., Ramos, O.M., Leininger, S., Liu, J., Ferrier, D.E.K., and Adamska, M. (2014). Calcisponges have a ParaHox gene and dynamic expression of dispersed NK homeobox genes. Nature 514, 620623.Google Scholar
Fortunato, S.A.V., Leininger, S., and Adamski, M. (2014). Evolution of the Pax-Six-Eya-Dach network: the calcisponge case study. EvoDevo 5, Article 23.Google Scholar
Foster, K.W. (2009). Eye evolution: two eyes can be better than one. Curr. Biol. 19, R208R210.Google Scholar
Foster, R.G., and Hankins, M.W. (2007). Circadian vision. Curr. Biol. 17, R746R751.Google Scholar
Franchini, L.F., and Pollard, K.S. (2015). Can a few non-coding mutations make a human brain? BioEssays 37, 10541061.Google Scholar
Frankel, N., Davis, G.K., Vargas, D., Wang, S., Payre, F., and Stern, D.L. (2010). Phenotypic robustness conferred by apparently redundant transcriptional enhancers. Nature 466, 490493.Google Scholar
Franz, E.A., Chiaroni-Clarke, R., Woodrow, S., Glendining, K.A., Jasoni, C.L., Robertson, S.P., Gardner, R.J.M., and Markie, D. (2015). Congenital mirror movements: phenotypes associated with DCC and RAD51 mutations. J. Neurol. Sci. 351, 140145.Google Scholar
Franze, K. (2013). The mechanical control of nervous system development. Development 140, 30693077.Google Scholar
Frasnelli, E., Vallortigara, G., and Rogers, L.J. (2012). Left-right asymmetries of behaviour and nervous system in invertebrates. Neurosci. Biobehav. Rev. 36, 12731291.Google Scholar
Frawley, L.E., and Orr-Weaver, T.L. (2015). Polyploidy. Curr. Biol. 25, R353R357.Google Scholar
Freeman, G., and Lundelius, J.W. (1982). The developmental genetics of dextrality and sinistrality in the gastropod Lymnaea peregra. W. Roux's Arch. 191, 6983.Google Scholar
Freeman, M.R., and Doherty, J. (2006). Glial cell biology in Drosophila and vertebrates. Trends Neurosci. 29, 8290.Google Scholar
Freeman, R., Ikuta, T., Wu, M., Koyanagi, R., Kawashima, T., Tagawa, K., Humphreys, T., Fang, G.-C., Fujiyama, A., Saiga, H., Lowe, C., Worley, K., Jenkins, J., Schmutz, J., Kirschner, M., Rokhsar, D., Satoh, N., and Gerhart, J. (2012). Identical genomic organization of two hemichordate Hox clusters. Curr. Biol. 22, 20532058.Google Scholar
Freibaum, B.D., Lu, Y., Lopez-Gonzalez, R., Kim, N.C., Almeida, S., Lee, K.-H., Badders, N., Valentine, M., Miller, B.L., Wong, P.C., Petrucelli, L., Kim, H.J., Gao, F.-B., and Taylor, J.P. (2015). GGGGCC repeat expansion in C9orf72 compromises nucleocytoplasmic transport. Nature 525, 129133.Google Scholar
French, V., Bryant, P.J., and Bryant, S.V. (1976). Pattern regulation in epimorphic fields. Science 193, 969981.Google Scholar
Frentiu, F.D., and Briscoe, A.D. (2008). A butterfly eye's view of birds. BioEssays 30, 11511162.Google Scholar
Frenzel, H., Bohlender, J., Pinsker, K., Wohlleben, B., Tank, J., Lechner, S.G., Schiska, D., Jaijo, T., Rüshendorf, F., Saar, K., Jordan, J., Millán, J.M., Gross, M., and Lewin, G.R. (2012). A genetic basis for mechanosensory traits in humans. PLoS Biol. 10, #5, e1001318.Google Scholar
Friedmann, D., Hoagland, A., Berlin, S., and Isacoff, E.Y. (2015). A spinal opsin controls early neural activity and drives a behavioral light response. Curr. Biol. 25, 6974.Google Scholar
Friedrich, F. (2006). Ancient mechanisms of visual sense organ development based on comparison of the gene networks controlling larval eye, ocellus, and compound eye specification in Drosophila. Arthropod Struct. Dev. 35, 357378.Google Scholar
Friedrich, M. (2015). Evo-devo gene toolkit update: at least seven Pax transcription factor subfamilies in the last common ancestor of bilaterian animals. Evol. Dev. 17, 255257.Google Scholar
Frigon, A. (2012). Central pattern generators of the mammalian spinal cord. Neuroscientist 18, 5669.Google Scholar
Fritsch, M., Wollesen, T., de Oliveira, A.L., and Wanninger, A. (2015). Unexpected co-linearity of Hox gene expression in an aculiferan mollusk. BMC Evol. Biol. 15, Article 151.Google Scholar
Fritzsch, B., and Beisel, K.W. (2001). Evolution and development of the vertebrate ear. Brain Res. Bull. 55, 711721.Google Scholar
Fritzsch, B., and Beisel, K.W. (2003). Molecular conservation and novelties in vertebrate ear development. Curr. Top. Dev. Biol. 57, 144.Google Scholar
Fritzsch, B., and Beisel, K.W. (2004). Keeping sensory cells and evolving neurons to connect them to the brain: molecular conservation and novelties in vertebrate ear development. Brain Behav. Evol. 64, 182197.Google Scholar
Fritzsch, B., Beisel, K.W., Pauley, S., and Soukup, G. (2007). Molecular evolution of the vertebrate mechanosensory cell and ear. Int. J. Dev. Biol. 51, 663678.Google Scholar
Fritzsch, B., Eberl, D.F., and Beisel, K.W. (2010). The role of bHLH genes in ear development and evolution: revisiting a 10-year-old hypothesis. Cell. Mol. Life Sci. 67, 30893099.Google Scholar
Fritzsch, B., Jahan, I., Pan, N., Kersigo, J., Duncan, J., and Kopecky, B. (2011). Dissecting the molecular basis of organ of Corti development: where are we now? Hear. Res. 276, 1626.Google Scholar
Fritzsch, B., Pauley, S., and Beisel, K.W. (2006). Cells, molecules and morphogenesis: the making of the vertebrate ear. Brain Res. 1091, 151171.Google Scholar
Fritzsch, B., and Piatigorsky, J. (2005). Ancestry of photic and mechanic sensation? Science 308, 11131114.Google Scholar
Fritzsch, B., and Straka, H. (2014). Evolution of vertebrate mechanosensory hair cells and inner ears: toward identifying stimuli that select mutation driven altered morphologies. J. Comp. Physiol. A 200, 518.Google Scholar
Fröbius, A.C., Matus, D.Q., and Seaver, E.C. (2008). Genomic organization and expression demonstrate spatial and temporal Hox gene colinearity in the lophotrochozoan Capitella sp. I. PLoS ONE 3, #12, e4004.Google Scholar
Frohns, A., Frohns, F., Naumann, S.C., Layer, P.G., and Löbrich, M. (2014). Inefficient double-strand break repair in murine rod photoreceptors with inverted heterochromatin organization. Curr. Biol. 24, 10801090.Google Scholar
Fu, X., Yan, Y., Xu, P.S., Geerlof-Vidavsky, I., Chong, W., Gross, M.L., and Holy, T.E. (2015). A molecular code for identity in the vomeronasal system. Cell 163, 313323.Google Scholar
Fu, Y., Kefalov, V., Luo, D.-G., Xue, T., and Yau, K.-W. (2008). Quantal noise from human red cone pigment. Nat. Neurosci. 11, 565571.Google Scholar
Fuchs, E. (2007). Scratching the surface of skin development. Nature 445, 834842.Google Scholar
Fuchs, P.A. (2015). How many proteins does it take to gate hair cell mechanotransduction? PNAS 112, #5, 12541255.Google Scholar
Fuchs, Y., and Steller, H. (2011). Programmed cell death in animal development and disease. Cell 147, 742758.Google Scholar
Fuerst, P.G., Bruce, F., Rounds, R.P., Erskine, L., and Burgess, R.W. (2012). Cell autonomy of DSCAM function in retinal development. Dev. Biol. 361, 326337.Google Scholar
Fuerst, P.G., Koizumi, A., Masland, R.H., and Burgess, R.W. (2008). Neurite arborization and mosaic spacing in the mouse retina require DSCAM. Nature 451, 470474.Google Scholar
Fujii, S., Yavuz, A., Slone, J., Jagge, C., Song, X., and Amrein, H. (2015). Drosophila sugar receptors in sweet taste perception, olfaction, and internal nutrient sensing. Curr. Biol. 25, 621627.Google Scholar
Fukushige, T., Brodigan, T.M., Schriefer, L.A., Waterston, R.H., and Krause, M. (2006). Defining the transcriptional redundancy of early bodywall muscle development in C. elegans: evidence for a unified theory of animal muscle development. Genes Dev. 20, 33953406.Google Scholar
Furlong, R.F., and Holland, P.W.H. (2002). Bayesian phylogenetic analysis supports monophyly of Ambulacraria and of cyclostomes. Zool. Sci. 19, 593599.Google Scholar
Fusco, M.A., Wajsenzon, I.J.R., de Carvalho, S.L., da Silva, R.T., Einicker-Lamas, M., Cavalcante, L.A., and Allodi, S. (2014). Vascular endothelial growth factor-like and its receptor in a crustacean optic ganglia: a role in neuronal differentiation? Biochem. Biophys. Res. Comm. 447, 299303.Google Scholar
Fuss, S.H., and Ray, A. (2009). Mechanisms of odorant receptor gene choice in Drosophila and vertebrates. Mol. Cell. Neurosci. 41, 101112.Google Scholar
Gaillard, D., Xu, M., Liu, F., Millar, S.E., and Barlow, L.A. (2015). β-Catenin signaling biases multipotent lingual epithelial progenitors to differentiate and acquire specific taste cell fates. PLoS Genet. 11, #5, e1005208.Google Scholar
Galic, M., and Matis, M. (2015). Polarized trafficking provides spatial cues for planar cell polarization within a tissue. BioEssays 37, 678686.Google Scholar
Galindo, M.I., Fernández-Garza, D., Phillips, R., and Couso, J.P. (2011). Control of Distal-less expression in the Drosophila appendages by functional 3’ enhancers. Dev. Biol. 353, 396410.Google Scholar
Galizia, C.G., and Rössler, W. (2010). Parallel olfactory systems in insects: anatomy and function. Annu. Rev. Entomol. 55, 399420.Google Scholar
Galli-Resta, L. (1998). Patterning the vertebrate retina: the early appearance of retinal mosaics. Semin. Cell Dev. Biol. 9, 279284.Google Scholar
Gallo, G. (2013). Mechanisms underlying the initiation and dynamics of neuronal filopodia: from neurite formation to synaptogenesis. Int. Rev. Cell Mol. Biol. 301, 95156.Google Scholar
Ganmor, E., Segev, R., and Schneidman, E. (2015). A thesaurus for a neural population code. eLife 4, e06134.Google Scholar
Gao, Q., Yuan, B., and Chess, A. (2000). Convergent projections of Drosophila olfactory neurons to specific glomeruli in the antennal lobe. Nat. Neurosci. 3, 780785. [See alsoGoogle ScholarGoogle Scholar
Gao, Y., Lan, Y., Liu, H., and Jiang, R. (2011). The zinc finger transcription factors Osr1 and Osr2 control synovial joint formation. Dev. Biol. 352, 8391.Google Scholar
Gao, Y., Lan, Y., Ovitt, C.E., and Jiang, R. (2009). Functional equivalence of the zinc finger transcription factors Osr1 and Osr2 in mouse development. Dev. Biol. 328, 200209.Google Scholar
Garcia-Fernàndez, J. (2005). The genesis and evolution of homeobox gene clusters. Nat. Rev. Genet. 6, 881892.Google Scholar
Garcia-Fernàndez, J. (2005). Hox, ParaHox, ProtoHox: facts and guesses. Heredity 94, 145152.Google Scholar
Garcia-Fernàndez, J., and Benito-Gutiérrez, É. (2009). It's a long way from amphioxus: descendants of the earliest chordate. BioEssays 31, 665675.Google Scholar
García-Frigola, C., Carreres, M.I., Vegar, C., Mason, C., and Herrera, E. (2008). Zic2 promotes axonal divergence at the optic chiasm midline by EphB1-dependent and -independent mechanisms. Development 135, 18331841.Google Scholar
Gardner, K.H., and Correa, F. (2012). How plants see the invisible. Science 335, 14511452.Google Scholar
Garm, A., and Nilsson, D.-E. (2014). Visual navigation in starfish: first evidence for the use of vision and eyes in starfish. Proc. R. Soc. Lond. B 281, 20132011.Google Scholar
Garm, A., Oskarsson, M., and Nilsson, D.-E. (2011). Box jellyfish use terrestrial visual cues for navigation. Curr. Biol. 21, 798803.Google Scholar
Garrett, A.M., and Burgess, R.W. (2015). Self-awareness in the retina. eLife 4, e10233.Google Scholar
Garrity, P.A. (2010). Feel the light. Nature 468, 900901.Google Scholar
Garry, D.J., and Olson, E.N. (2006). A common progenitor at the heart of development. Cell 127, 11011104.Google Scholar
Garstang, M., and Ferrier, D.E.K. (2013). Time is of the essence for ParaHox homeobox gene clustering. BMC Biol. 11, Article 72.Google Scholar
Garvie, C.W., and Wolberger, C. (2001). Recognition of specific DNA sequences. Mol. Cell 8, 937946.Google Scholar
Gaunt, S.J. (1997). Chick limbs, fly wings and homology at the fringe. Nature 386, 324325.Google Scholar
Gaunt, S.J., Drage, D., and Trubshaw, R.C. (2008). Increased Cdx protein dose effects upon axial patterning in transgenic lines of mice. Development 135, 25112520.Google Scholar
Gaviño, M.A., and Reddien, P.W. (2011). A Bmp/Admp regulatory circuit controls maintenance and regeneration of dorsal-ventral polarity in planarians. Curr. Biol. 21, 294299.Google Scholar
Gee, H. (2008). The amphioxus unleashed. Nature 453, 9991000.Google Scholar
Gehring, W., and Rosbash, M. (2003). The coevolution of blue-light photoreception and circadian rhythms. J. Mol. Evol. 57, S286S289.Google Scholar
Gehring, W.J. (1985). Homeotic genes, the homeo box, and the genetic control of development. Cold Spring Harb. Symp. Quant. Biol. 50, 243251.Google Scholar
Gehring, W.J. (1998). Master Control Genes in Development and Evolution: The Homeobox Story. Yale University Press, New Haven, CT.Google Scholar
Gehring, W.J. (2002). The genetic control of eye development and its implications for the evolution of the various eye-types. Int. J. Dev. Biol. 46, 6573.Google Scholar
Gehring, W.J. (2012). The animal body plan, the prototypic body segment, and eye evolution. Evol. Dev. 14, 3446.Google Scholar
Gehring, W.J. (2014). The evolution of vision. WIREs Dev. Biol. 3, 140.Google Scholar
Gehring, W.J., and Ikeo, K. (1999). Pax6: mastering eye morphogenesis and eye evolution. Trends Genet. 15, 371377.Google Scholar
Géléoc, G.S.G., and Holt, J.R. (2014). Sound strategies for hearing restoration. Science 344, 596.Google Scholar
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
Genikhovich, G., Fried, P., Prünster, M.M., Schinko, J.B., Gilles, A.F., Fredman, D., Meier, K., Iber, D., and Technau, U. (2015). Axis patterning by BMPs: cnidarian network reveals evolutionary constraints. Cell Rep. 10, 16461654.Google Scholar
Geoffroy St.-Hilaire, E. (1822). Considérations générales sur la vertèbre. Mém. Mus. Hist. Nat. 9, 89119 + Planches V–VII (ff).Google Scholar
Gerdes, J.M., Davis, E.E., and Katsanis, N. (2009). The vertebrate primary cilium in development, homeostasis, and disease. Cell 137, 3245.Google Scholar
Gerhart, J., and Kirschner, M. (1997). Cells, Embryos, and Evolution. Blackwell Science, Malden, MA.Google Scholar
Gerhart, J., Lowe, C., and Kirschner, M. (2005). Hemichordates and the origin of chordates. Curr. Opin. Genet. Dev. 15, 461467.Google Scholar
Gerkema, M.P., Davies, W.I.L., Foster, R.G., Menaker, M., and Hut, R.A. (2013). The nocturnal bottleneck and the evolution of activity patterns in mammals. Proc. R. Soc. Lond. B 280, 20130508. [See alsoGoogle ScholarGoogle Scholar
Gerstner, J.R., and Yin, J.C.P. (2010). Circadian rhythms and memory formation. Nat. Rev. Neurosci. 11, 577588.Google Scholar
Ghysen, A. (1990). Origins of segment periodicity. Nature 344, 297298.Google Scholar
Ghysen, A. (2003). The origin and evolution of the nervous system. Int. J. Dev. Biol. 47, 555562.Google Scholar
Gibbons, B. (1986). The intimate sense of smell. Natl. Geogr. 170, #3, 324361.Google Scholar
Gibson, G. (2000). Evolution: Hox genes and the cellared wine principle. Curr. Biol. 10, R452R455.Google Scholar
Giessel, A.J., and Datta, S.R. (2014). Olfactory maps, circuits and computations. Curr. Opin. Neurobiol. 24, 120132.Google Scholar
Gilbert, A.N., and Firestein, S. (2002). Dollars and scents: commercial opportunities in olfaction and taste. Nat. Neurosci. 5 (Suppl.), 10431045.Google Scholar
Gilbert, J.-M. (2002). The evolution of engrailed genes after duplication and speciation events. Dev. Genes Evol. 212, 307318.Google Scholar
Gilbert, S.F., and Bolker, J.A. (2001). Homologies of process and modular elements of embryonic construction. J. Exp. Zool. 291, 112.Google Scholar
Gilbert, S.L., Dobyns, W.B., and Lahn, B.T. (2005). Genetic links between brain development and brain evolution. Nat. Rev. Genet. 6, 581590.Google Scholar
Gilestro, G.F. (2008). Redundant mechanisms for regulation of midline crossing in Drosophila. PLoS ONE 3, #11, e3798.Google Scholar
Giljov, A., Karenina, K., Ingram, J., and Malashichev, Y. (2015). Parallel emergence of true handedness in the evolution of marsupials and placentals. Curr. Biol. 25, 18781884.Google Scholar
Gillespie, P.G., and Walker, R.G. (2001). Molecular basis of mechanosensory transduction. Nature 413, 194202.Google Scholar
Gillies, T.E., and Cabernard, C. (2011). Cell division orientation in animals. Curr. Biol. 21, R599R609.CrossRefGoogle ScholarPubMed
Giorgianni, M., and Patel, N.H. (2005). Conquering land, air and water: the evolution and development of arthropod appendages. In Briggs, D.E.G. (ed.), Evolving Form and Function: Fossils and Development. Yale University Peabody Museum of Natural History, New Haven, CT., pp. 159180.Google Scholar
Gitton, Y., Dahmane, N., Balk, S., Ruiz i Altaba, A., Neidhardt, L., Scholze, M., Herrmann, B.G., Kanlem, P., Benkahla, A., Schrinner, S., Yildirimman, R., Herwig, R., Lehrach, H., and Yaspo, M.-L. (2002). A gene expression map of human chromosome 21 orthologues in the mouse. Nature 420, 586590.Google Scholar
Godwin, J. (2014). The promise of perfect adult tissue repair and regeneration in mammals: learning from regenerative amphibians and fish. BioEssays 36, 861871.Google Scholar
Goetz, J.J., Farris, C., Chowdhury, R., and Trimarchi, J.M. (2014). Making of a retinal cell: insights into retinal cell-fate determination. Int. Rev. Cell Mol. Biol. 308, 273321.Google Scholar
Goldbeter, A., Gonze, D., and Pourquié, O. (2007). Sharp developmental thresholds defined through bistability by antagonistic gradients of retinoic acid and FGF signaling. Dev. Dyn. 236, 14951508.Google Scholar
Goldman, A.L., van der Goes van Naters, W., Lessing, D., Warr, C.G., and Carlson, J.R. (2005). Coexpression of two functional odor receptors in one neuron. Neuron 45, 661666.Google Scholar
Goldsmith, T.H. (1990). Optimization, constraint, and history in the evolution of eyes. Q. Rev. Biol. 65, 281322.Google Scholar
Goldsmith, T.H. (2013). Evolutionary tinkering with visual photoreception. Vis. Neurosci. 30, 2137.Google Scholar
Goltsev, Y., Fuse, N., Frasch, M., Zinzen, R.P., Lanzaro, G., and Levine, M. (2007). Evolution of the dorsal-ventral patterning network in the mosquito, Anopheles gambiae. Development 134, 24152424.Google Scholar
Gómez-Skarmeta, J.L., and Modolell, J. (2002). Iroquois genes: genomic organization and function in vertebrate neural development. Curr. Opin. Genet. Dev. 12, 403408.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
González, J., Ranz, J.M., and Ruiz, A. (2002). Chromosomal elements evolve at different rates in the Drosophila genome. Genetics 161, 11371154.Google Scholar
González-Morales, N., Géminard, C., Lebreton, G., Cerezo, D., Coutelis, J.-B., and Noselli, S. (2015). The atypical cadherin Dachsous controls left-right asymmetry in Drosophila. Dev. Cell 33, 675689.Google Scholar
Goode, D.K., Callaway, H.A., Cerda, G.A., Lewis, K.E., and Elgar, G. (2011). Minor change, major difference: divergent functions of highly conserved cis-regulatory elements subsequent to whole genome duplication events. Development 138, 879884.Google Scholar
Goode, D.K., and Elgar, G. (2009). The PAX258 gene subfamily: a comparative perspective. Dev. Dyn. 238, 29512974.Google Scholar
Goodrich, L.V., and Strutt, D. (2011). Principles of planar polarity in animal development. Development 138, 18771892.Google Scholar
Gooijers, J., and Swinnen, S.P. (2014). Interactions between brain structure and behavior: the corpus callosum and bimanual coordination. Neurosci. Biobehav. Rev. 43, 119.Google Scholar
Göpfert, M.C., and Hennig, R.M. (2016). Hearing in insects. Annu. Rev. Entomol. 61, 257276.Google Scholar
Göpfert, M.C., and Robert, D. (2002). The mechanical basis of Drosophila audition. J. Exp. Biol. 205, 11991208.Google Scholar
Göpfert, M.C., Stocker, H., and Robert, D. (2002). atonal is required for exoskeletal joint formation in the Drosophila auditory system. Dev. Dyn. 225, 106109.Google Scholar
Gorbunov, D., Sturlese, M., Nies, F., Kluge, M., Bellanda, M., Battistutta, R., and Oliver, D. (2014). Molecular architecture and the structural basis for anion interaction in prestin and SLC26 transporters. Nat. Commun. 5, Article 3622.Google Scholar
Gorfinkiel, N., Morata, G., and Guerrero, I. (1997). The homeobox gene Distal-less induces ventral appendage development in Drosophila. Genes Dev. 11, 22592271.Google Scholar
Gosse, N.J., Nevin, L.M., and Baier, H. (2008). Retinotopic order in the absence of axon competition. Nature 452, 892895.Google Scholar
Goto, S., and Hayashi, S. (1997). Cell migration within the embryonic limb primordium of Drosophila as revealed by a novel fluorescence method to visualize mRNA and protein. Dev. Genes Evol. 207, 194198.Google Scholar
Goto, S., and Hayashi, S. (1997). Specification of the embryonic limb primordium by graded activity of Decapentaplegic. Development 124, 125132.Google Scholar
Gottardo, M., Pollarolo, G., Llamazares, S., Reina, J., Riparbelli, M.G., Callaini, G., and Gonzalez, C. (2015). Loss of centrobin enables daughter centrioles to form sensory cilia in Drosophila. Curr. Biol. 25, 23192324.Google Scholar
Gould, S.J. (1977). Ontogeny and Phylogeny. Harvard University Press, Cambridge, MA.Google Scholar
Gould, S.J. (1994). Common pathways of illumination. Nat. Hist. 103, #12, 1020.Google Scholar
Gould, S.J. (1997). As the worm turns. Nat. Hist. 106, #1, 2427, 68–73.Google Scholar
Gould, S.J. (2002). The Structure of Evolutionary Theory. Harvard University Press, Cambridge, MA.Google Scholar
Graham, A., Butts, T., Lumsden, A., and Kiecker, C. (2014). What can vertebrates tell us about segmentation? EvoDevo 5, Article 24.Google Scholar
Graham, A., Papalopulu, N., and Krumlauf, R. (1989). The murine and Drosophila homeobox gene complexes have common features of organization and expression. Cell 57, 367378.Google Scholar
Graham, D.M., Wong, K.Y., Shapiro, P., Frederick, C., Pattabiraman, K., and Berson, D.M. (2008). Melanopsin ganglion cells use a membrane-associated rhabdomeric phototransduction cascade. J. Neurophysiol. 99, 25222532.Google Scholar
Graham, T.G.W., Tabei, S.M.A., Dinner, A.R., and Rebay, I. (2010). Modeling bistable cell-fate choices in the Drosophila eye: qualitative and quantitative perspectives. Development 137, 22652278.Google Scholar
Grande, C., and Patel, N.H. (2009). Nodal signalling is involved in left–right asymmetry in snails. Nature 457, 10071011.Google Scholar
Grati, M., and Kachar, B. (2011). Myosin VIIa and sans localization at stereocilia upper tip-link density implicates these Usher syndrome proteins in mechanotransduction. PNAS 108, #28, 1147611481.Google Scholar
Graur, D., Zheng, Y., Price, N., Azevedo, R.B.R., Zufall, R.A., and Elhaik, E. (2013). On the immortality of television sets: “function” in the human genome according to the evolution-free gospel of ENCODE. Genome Biol. 5, 578590.Google Scholar
Graveley, B.R. (2005). Mutually exclusive splicing of the insect Dscam pre-mRNA directed by competing intronic RNA secondary structures. Cell 123, 6573.Google Scholar
Graveley, B.R., Kaur, A., Gunning, D., Zipursky, S.L., Rowen, L., and Clemens, J.C. (2004). The organization and evolution of the dipteran and hymenopteran Down syndrome cell adhesion molecule (Dscam) genes. RNA 10, 14991506.Google Scholar
Graw, J. (2010). Eye development. Curr. Top. Dev. Biol. 90, 343386.Google Scholar
Graziussi, D.F., Suga, H., Schmid, V., and Gehring, W.J. (2012). The “eyes absent” (eya) gene in the eye-bearing hydrozoan jellyfish Cladonema radiatum: conservation of the retinal determination network. J. Exp. Zool. B. Mol. Dev. Evol. 318, 257267.Google Scholar
Green, J., and Akam, M. (2013). Evolution of the pair rule gene network: insights from a centipede. Dev. Biol. 382, 235245.Google Scholar
Green, M.M. (2002). It really is not a fruit fly. Genetics 162, 13.Google Scholar
Greenberg, L., and Hatini, V. (2009). Essential roles for lines in mediating leg and antennal proximodistal patterning and generating a stable Notch signaling interface at segment borders. Dev. Biol. 330, 93104.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
Greenow, K., and Clarke, A.R. (2012). Controlling the stem cell compartment and regeneration in vivo: the role of pluripotency pathways. Physiol. Rev. 92, 7599.Google Scholar
Greenspan, R.J., and Dierick, H.A. (2004). “Am not I a fly like thee?” From genes in fruit flies to behavior in humans. Hum. Mol. Genet. 13, R267R273.Google Scholar
Greenspan, R.J., and van Swinderen, B. (2004). Cognitive consonance: complex brain functions in the fruit fly and its relatives. Trends Neurosci. 27, 707711.Google Scholar
Greer, J.M., Puetz, J., Thomas, K.R., and Capecchi, M.R. (2000). Maintenance of functional equivalence during paralogous Hox gene evolution. Nature 403, 661665.Google Scholar
Gregor, T., McGregor, A.P., and Wieschaus, E.F. (2008). Shape and function of the Bicoid morphogen gradient in dipteran species with different sized embryos. Dev. Biol. 316, 350358.Google Scholar
Gregory, T.R. (2008). The evolution of complex organs. Evol. Educ. Outreach 1, 358389.Google Scholar
Greven, H. (2007). Comments on the eyes of tardigrades. Arthropod Struct. Dev. 36, 401407.Google Scholar
Griffin, C., Kleinjan, D.A., Doe, B., and van Heyningen, V. (2002). New 3’ elements control Pax6 expression in the developing pretectum, neural retina and olfactory region. Mech. Dev. 112, 89100.Google Scholar
Grigorian, M., and Hartenstein, V. (2013). Hematopoiesis and hematopoietic organs in arthropods. Dev. Genes Evol. 223, 103115.Google Scholar
Grill, S.W. (2010). Forced to be unequal. Science 330, 597598.Google Scholar
Grimes, A.C., and Kirby, M.L. (2009). The outflow tract of the heart in fishes: anatomy, genes and evolution. J. Fish Biol. 74, 9831036.Google Scholar
Grindley, J.C., Davidson, D.R., and Hill, R.E. (1995). The role of Pax-6 in eye and nasal development. Development 121, 14331442.Google Scholar
Gros, J., Feistel, K., Viebahn, C., Blum, M., and Tabin, C. (2009). Cell movements at Hensen's node establish left/right asymmetric gene expression in the chick. Science 324, 941944.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
Grünbaum, B., and Shephard, G.C. (1987). Tilings and Patterns. W.H. Freeman, New York, NY.Google Scholar
Grus, W.E., and Zhang, J. (2006). Origin and evolution of the vertebrate vomeronasal system viewed through system-specific genes. BioEssays 28, 709718.Google Scholar
Grus, W.E., and Zhang, J. (2009). Origin of the genetic components of the vomeronasal system in the common ancestor of all extant vertebrates. Mol. Biol. Evol. 26, 407419.Google Scholar
Gruss, M., Bushell, T.J., Bright, D.P., Lieb, W.R., Mathie, A., and Franks, N.P. (2004). Two-pore-domain K+ channels are a novel target for the anesthetic gases xenon, nitrous oxide, and cyclopropane. Mol. Pharmacol. 65, 443452.Google Scholar
Guarner, A., Manjón, C., Edwards, K., Steller, H., Suzanne, M., and Sánchez-Herrero, E. (2014). The zinc finger homeodomain-2 gene of Drosophila controls Notch targets and regulates apoptosis in the tarsal segments. Dev. Biol. 385, 350365.Google Scholar
Guarnieri, D.J., and Heberlein, U. (2003). Drosophila melanogaster, a genetic model system for alcohol research. Int. Rev. Neurobiol. 54, 199228.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
Gubb, D., Green, C., Huen, D., Coulson, D., Johnson, G., Tree, D., Collier, S., and Roote, J. (1999). The balance between isoforms of the Prickle LIM domain protein is critical for planar polarity in Drosophila imaginal discs. Genes Dev. 13, 23152327.Google Scholar
Guerin, M.B., McKernan, D.P., O'Brien, C.J., and Cotter, T.G. (2006). Retinal ganglion cells: dying to survive. Int. J. Dev. Biol. 50, 665674.Google Scholar
Guertin, P.A. (2009). The mammalian central pattern generator for locomotion. Brain Res. Rev. 62, 4556.Google Scholar
Guichard, C., Harricane, M.-C., Lafitte, J.-J., Godard, P., Zaegel, M., Tack, V., Lalau, G., and Bouvagnet, P. (2001). Axonemal dynein intermediate-chain gene (DNAI1) mutations result in situs inversus and primary ciliary diskinesia (Kartagener syndrome). Am. J. Hum. Genet. 68, 10301035.Google Scholar
Guillemot, F., Lo, L.-C., Johnson, J.E., Auerbach, A., Anderson, D.J., and Joyner, A.L. (1993). Mammalian achaete-scute homolog 1 is required for the early development of olfactory and autonomic neurons. Cell 75, 463476.Google Scholar
Guillery, R.W. (1994). No crossing at the chiasm. Nature 367, 597598.Google Scholar
Gulino, A., Di Marcotullio, L., and Screpanti, I. (2010). The multiple functions of Numb. Exp. Cell Res. 316, 900906.Google Scholar
Güntürkün, O. (2012). Brain asymmetry in vertebrates. In Lazareva, O.F., Shimizu, T., and Wasserman, E.A. (eds.), How Animals See the World: Comparative Behavior, Biology, and Evolution of Vision. Oxford University Press, Oxford, pp. 501519.Google Scholar
Guo, N., Hawkins, C., and Nathans, J. (2004). Frizzled6 controls hair patterning in mice. PNAS 101, #25, 92779281.Google Scholar
Gurtan, A.M., and Sharp, P.A. (2013). The role of miRNAs in regulating gene expression networks. J. Mol. Biol. 425, 35823600.Google Scholar
Guthrie, S. (1999). Axon guidance: starting and stopping with Slit. Curr. Biol. 9, R432R435.Google Scholar
Gutierrez-Mazariegos, J., Schubert, M., and Laudet, V. (2014). Evolution of retinoic acid receptors and retinoic acid signaling. In Asson-Batres, M.A. and Rochette-Egly, C. (eds.), The Biochemistry of Retinoic Acid Receptors I: Structure, Activation, and Function at the Molecular Level. Springer, New York, NY, pp. 5573.Google Scholar
Gutierrez-Mazariegos, J., Theodosiou, M., Campo-Paysaa, F., and Schubert, M. (2011). Vitamin A: a multifunctional tool for development. Semin. Cell Dev. Biol. 22, 603610.Google Scholar
Ha, T.S., and Smith, D.P. (2009). Odorant and pheromone receptors in insects. Front. Cell. Neurosci. 3, Article 10.Google Scholar
Häder, T., La Rosée, A., Ziebold, U., Busch, M., Taubert, H., Jäckle, H., and Rivera-Pomar, R. (1998). Activation of posterior pair-rule stripe expression in response to maternal caudal and zygotic knirps activities. Mech. Dev. 71, 177186.Google Scholar
Haeussler, M., Jaszczyszyn, Y., Christiaen, L., and Joly, J.-S. (2010). A cis-regulatory signature for chordate anterior neuroectodermal genes. PLoS Genet. 6, #4, e1000912.Google Scholar
Hafen, E., Kuroiwa, A., and Gehring, W.J. (1984). Spatial distribution of transcripts from the segmentation gene fushi tarazu during Drosophila embryonic development. Cell 37, 833841.Google Scholar
Haithcock, J., Billington, N., Choi, K., Fordham, J., Sellers, J.R., Stafford, W.F., White, H., and Forgacs, E. (2011). The kinetic mechanism of mouse myosin VIIA. J. Biol. Chem. 286, 88198828.Google Scholar
Haldane, J.B.S. (1928). On being the right size. In Possible Worlds and Other Papers. Harper & Bros., New York, NY, pp. 2028.Google Scholar
Halder, G., Callaerts, P., and Gehring, W.J. (1995). Induction of ectopic eyes by targeted expression of the eyeless gene in Drosophila. Science 267, 17881792.Google Scholar
Hale, R., and Strutt, D. (2015). Conservation of planar polarity pathway function across the animal kingdom. Annu. Rev. Genet. 49, 529551.Google Scholar
Hall, B.K. (1984). Homology: The Hierarchical Basis of Comparative Biology. Academic Press, San Diego, CA.Google Scholar
Hall, B.K. (2012). Parallelism, deep homology, and evo-devo. Evol. Dev. 14, 2933.Google Scholar
Hall, B.K., and Olson, W.M. (eds.) (2003). Keywords and Concepts in Evolutionary Developmental Biology. Harvard University Press, Cambridge, MA.Google Scholar
Hall, J.C. (2003). A neurogeneticist's manifesto. J. Neurogenet. 17, 190.Google Scholar
Hallem, E.A., and Carlson, J.R. (2004). The odor coding system of Drosophila. Trends Genet. 20, 453459.Google Scholar
Hallem, E.A., and Carlson, J.R. (2006). Coding of odors by a receptor repertoire. Cell 125, 143160.Google Scholar
Hallsson, J.H., Haflidadóttir, B.S., Stivers, C., Odenwald, W., Arnheiter, H., Pignoni, F., and Steingrímsson, E. (2004). The basic helix-loop-helix leucine zipper transcription factor Mitf is conserved in Drosophila and functions in eye development. Genetics 167, 233241.Google Scholar
Halpern, M.E., Hobert, O., and Wright, C.V.E. (2014). Left-right asymmetry: advances and enigmas. Genesis 52, 451454.Google Scholar
Hamada, H. (2015). Role of physical forces in embryonic development. Semin. Cell Dev. Biol. 47–48, 8891.Google Scholar
Hamilton, C.R., and Vermeire, B.A. (1988). Complementary hemispheric specialization in monkeys. Science 242, 16911694.Google Scholar
Hamilton, E.E., and Kay, S.A. (2008). Snapshot: circadian clock proteins. Cell 135, 368.Google Scholar
Hammond, K.L., Baxendale, S., McCauley, D., Ingham, P.W., and Whitfield, T.T. (2009). Expression of patched, prdm1 and engrailed in the lamprey somite reveals conserved responses to Hedgehog signaling. Evol. Dev. 11, 2740.Google Scholar
Hammond, K.L., and Whitfield, T.T. (2006). The developing lamprey ear closely resembles the zebrafish otic vesicle: otx1 expression can account for all major patterning differences. Development 133, 13471357.Google Scholar
Hampel, S., Franconville, R., Simpson, J.H., and Seeds, A.M. (2015). A neural command circuit for grooming movement control. eLife 4, e308758.Google Scholar
Hanchate, N.K., Kondoh, K., Lu, Z., Kuang, D., Ye, X., Qui, X., Pachter, L., Trapnell, C., and Buck, L.B. (2015). Single-cell transcriptomics reveals receptor transformations during olfactory neurogenesis. Science 350, 12511255.Google Scholar
Hankins, M.W., and Hughes, S. (2014). Vision: melanopsin as a novel irradiance detector at the heart of vision. Curr. Biol. 24, R1055R1057.Google Scholar
Hannibal, R.L., and Patel, N.H. (2013). What is a segment? EvoDevo 4, Article 35.Google Scholar
Hanson, I.M. (2001). Mammalian homologues of the Drosophila eye specification genes. Semin. Cell Dev. Biol. 12, 475484.Google Scholar
Hansson, B.S., and Stensmyr, M.C. (2011). Evolution of insect olfaction. Neuron 72, 698711.Google Scholar
Hao, I., Green, R.B., Dunaevsky, O., Lengyel, J.A., and Rauskolb, C. (2003). The odd-skipped family of zinc finger genes promotes Drosophila leg segmentation. Dev. Biol. 263, 282295. [See alsoGoogle ScholarGoogle Scholar
Hara, Y., Yamaguchi, M., Akasaka, K., Nakano, H., Nonaka, M., and Amemiya, S. (2006). Expression patterns of Hox genes in larvae of the sea lily Metacrinus rotundus. Dev. Genes Evol. 216, 797809.Google Scholar
Harada, T., Harada, C., and Parada, L.F. (2007). Molecular regulation of visual system development: more than meets the eye. Genes Dev. 21, 367378.Google Scholar
Hardie, R., and Raghu, P. (2001). Visual transduction in Drosophila. Nature 413, 186193.Google Scholar
Hardie, R.C. (1985). Functional organization of the fly retina. In Ottoson, D. (ed.), Progress in Sensory Physiology. Springer-Verlag, Berlin, pp. 179.Google Scholar
Hardie, R.C. (1986). The photoreceptor array of the dipteran retina. Trends Neurosci. 9, 419423.Google Scholar
Hardie, R.C. (2012). Polarization vision: Drosophila enters the arena. Curr. Biol. 22, R12R14.Google Scholar
Hardie, R.C., and Franze, K. (2012). Photomechanical responses in Drosophila photoreceptors. Science 338, 260263.Google Scholar
Hardin, P.E. (2000). From biological clock to biological rhythms. Genome Biol. 1, #4, reviews1023.Google Scholar
Hardin, P.E. (2005). The circadian timekeeping system of Drosophila. Curr. Biol. 15, R714R722.Google Scholar
Hardin, P.E. (2011). Molecular genetic analysis of circadian timekeeping in Drosophila. Adv. Genet. 74, 141173.Google Scholar
Harding, K., and Levine, M. (1988). Gap genes define the limits of Antennapedia and Bithorax gene expression during early development in Drosophila. EMBO J. 7, 205214.Google Scholar
Hardison, R.C. (2008). Globin genes on the move. J. Biol. 7, Article 35.Google Scholar
Hardison, R.C. (2012). Evolution of hemoglobin and its genes. Cold Spring Harb. Perspect. Med. 2, a011627.Google Scholar
Harris, W.A. (1997). Pax-6: where to be conserved is not conservative. PNAS 94, 20982100.Google Scholar
Harrison, R.G. (1917). Transplantation of limbs. PNAS 3, #4, 245251.Google Scholar
Hartenstein, V. (1993). Atlas of Drosophila Development. Cold Spring Harbor Laboratory Press, Plainview, NY.Google Scholar
Hartenstein, V. (2006). The neuroendocrine system of invertebrates: a developmental and evolutionary perspective. J. Endocrinology 190, 555570.Google Scholar
Hartenstein, V., and Mandal, L. (2006). The blood/vascular system in a phylogenetic perspective. BioEssays 28, 12031210.Google Scholar
Hartenstein, V., and Stollewerk, A. (2015). The evolution of early neurogenesis. Dev. Cell 32, 390407.Google Scholar
Hartenstein, V., Takashima, S., and Adams, K.L. (2010). Conserved genetic pathways controlling the development of the diffuse endocrine system in vertebrates and Drosophila. Gen. Comp. Endocrinol. 166, 462469.Google Scholar
Hartenstein, V., Tepass, U., and Gruszynski-Defeo, E. (1994). Embryonic development of the stomatogastric nervous system in Drosophila. J. Comp. Neurol. 350, 367381.Google Scholar
Hartline, D.K. (2011). The evolutionary origins of glia. Glia 59, 12151236.Google Scholar
Hartman, M.A., Finan, D., Sivaramakrishnan, S., and Spudich, J.A. (2011). Principles of unconventional myosin function and tarteting. Annu. Rev. Cell Dev. Biol. 27, 133155.Google Scholar
Harvey, R.P. (1996). NK-2 homeobox genes and heart development. Dev. Biol. 178, 203216.Google Scholar
Harvey, R.P. (2002). Patterning the vertebrate heart. Nat. Rev. Genet. 3, 544556.Google Scholar
Harzsch, S. (2002). The phylogenetic significance of crustacean optic neuropils and chiasmata: a re-examination. J. Comp. Neurol. 453, 1021.Google Scholar
Harzsch, S. (2004). The tritocerebrum of Euarthropoda: a “non-drosophilocentric” perspective. Evol. Dev. 6, 303309.Google Scholar
Harzsch, S., Vilpoux, K., Blackburn, D.C., Platchetzki, D., Brown, N.L., Melzer, R., Kempler, K.E., and Battelle, B.A. (2006). Evolution of arthropod visual systems: development of the eyes and central visual pathways in the horseshoe crab Limulus polyphemus Linnaeus, 1758 (Chelicerata, Xiphosura). Dev. Dyn. 235, 26412655.Google Scholar
Haskel-Ittah, M., Ben-Zvi, D., Branski-Arieli, M., Schejter, E.D., Shilo, B.-Z., and Barkai, N. (2012). Self-organized shuttling: generating sharp dorsoventral polarity in the early Drosophila embryo. Cell 150, 10161028.Google Scholar
Hassan, B.A., and Hiesinger, P.R. (2015). Beyond molecular codes: simple rules to wire complex brains. Cell 163, 285291. [For a paean to the wonders of science in general and neuroscience in particular, seeGoogle ScholarGoogle Scholar
Hattar, S., Liao, H.-W., Takao, M., Berson, D.M., and Yau, K.-W. (2002). Melanopsin-containing retinal ganglion cells: architecture, projections, and intrinsic photosensitivity. Science 295, 10651070.Google Scholar
Hattori, D., Chen, Y., Matthews, B.J., Salwinski, L., Sabatti, C., Grueber, W.B., and Zipursky, S.L. (2009). Robust discrimination between self and non-self neurites requires thousands of Dscam1 isoforms. Nature 461, 644648.Google Scholar
Hattori, D., Demir, E., Kim, H.W., Viragh, E., Zipursky, S.L., and Dickson, B. (2007). Dscam diversity is essential for neuronal wiring and self-recognition. Nature 449, 223227.Google Scholar
Hattori, D., Millard, S.S., Wojtowicz, W.M., and Zipursky, S.L. (2008). Dscam-mediated cell recognition regulates neural circuit formation. Annu. Rev. Cell Dev. Biol. 24, 597620.Google Scholar
Haun, C., Alexander, J., Stainier, D.Y., and Okkema, P.G. (1998). Rescue of Caenorhabditis elegans pharyngeal development by a vertebrate heart specification gene. PNAS 95, 50725075.Google Scholar
Hauser, F.E., van Hazel, I., and Chang, B.S.W. (2014). Spectral tuning in vertebrate short wavelength-sensitive 1 (SWS1) visual pigments: can wavelength sensitivity be inferred from sequence data? J. Exp. Zool. B. Mol. Dev. Evol. 322, 529539.Google Scholar
Hayashi, T., and Murakami, R. (2001). Left-right asymmetry in Drosophila melanogaster gut development. Dev. Growth Differ. 43, 239246.Google Scholar
Haynes, P.R., Christmann, B.L., and Griffith, L.C. (2015). A single pair of neurons links sleep to memory consolidation in Drosophila melanogaster. eLife 4, e03868.Google Scholar
Hayward, A.G. II, Joshi, P., and Skromne, I. (2015). Spatiotemporal analysis of zebrafish hox gene regulation by Cdx4. Dev. Dyn. 244, 15641573.Google Scholar
He, D.Z.Z., Zheng, J., Kalinec, F., Kakehata, S., and Santos-Sacchi, J. (2006). Tuning in to the amazing outer hair cell: membrane wizardry with a twist and shout. J. Membr. Biol. 209, 119134.Google Scholar
He, H., Kise, Y., Izadifar, A., Urwyler, O., Ayaz, D., Parthasarthy, A., Yan, B., Erfurth, M.-L., Dascenco, D., and Schmucker, D. (2014). Cell-intrinsic requirement of Dscam1 isoform diversity for axon collateral formation. Science 344, 11821886.Google Scholar
He, S., Dong, W., Deng, Q., Weng, S., and Sun, W. (2003). Seeing more clearly: recent advances in understanding retinal circuitry. Science 302, 408411.Google Scholar
Heckman, C.A., and Plummer, H.K. III (2013). Filopodia as sensors. Cell. Signal. 25, 22982311.Google Scholar
Heed, T. (2010). Touch perception: how we know where we are touched. Curr. Biol. 20, R604R606.Google Scholar
Heffer, A., Löhr, U., and Pick, L. (2011). ftz evolution: findings, hypotheses and speculations. BioEssays 33, 910918.Google Scholar
Heffer, A., and Pick, L. (2013). Conservation and variation in Hox genes: how insect models pioneered the evo-devo field. Annu. Rev. Entomol. 58, 161179.Google Scholar
Heffer, A., Shultz, J.W., and Pick, L. (2010). Surprising flexibility in a conserved Hox transcription factor over 550 million years of evolution. PNAS 107, 1804018045.Google Scholar
Heimberg, A., and McGlinn, E. (2012). Building a robust A-P axis. Curr. Genomics 13, 278288.Google Scholar
Heine, P., Dohle, E., Bumsted-O'Brien, K., Engelkamp, D., and Schulte, D. (2008). Evidence for an evolutionary conserved role of homothorax/Meis1/2 during vertebrate retina development. Development 135, 805811.Google Scholar
Hejnol, A., and Dunn, C.W. (2016). Animal evolution: are phyla real? Curr. Biol. 26, R424R426.Google Scholar
Hejnol, A., and Lowe, C.J. (2014). Animal evolution: stiff or squishy notochord origins? Curr. Biol. 24, R1131R1133.Google Scholar
Hejnol, A., and Martindale, M.Q. (2009). Coordinated spatial and temporal expression of Hox genes during embryogenesis in the acoel Convolutriloba longifissura. BMC Biol. 7, Article 65.Google Scholar
Hejnol, A., and Rentzsch, F. (2015). Neural nets. Curr. Biol. 25, R782R786.Google Scholar
Hejnol, A., and Scholtz, G. (2004). Clonal analysis of Distal-less and engrailed expression patterns during early morphogenesis of uniramous and biramous crustacean limbs. Dev. Genes Evol. 214, 473485.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. (1991). Bristle patterning in Drosophila. BioEssays 13, 633640.Google Scholar
Held, L.I. Jr. (1992). Models for Embryonic Periodicity. Karger, Basel.Google Scholar
Held, L.I. Jr. (2002). Imaginal Discs: The Genetic and Cellular Logic of Pattern Formation. Cambridge University Press, New York, NY.Google Scholar
Held, L.I. Jr. (2005). Suppressing apoptosis fails to cure “extra-joint syndrome” or to stop sex-comb rotation. Dros. Inf. Serv. 88, 910.Google Scholar
Held, L.I. Jr. (2009). Quirks of Human Anatomy: An Evo-Devo Look at the Human Body. Cambridge University Press, New York, NY.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). The evolutionary geometry of human anatomy: discovering our inner fly. Evol. Anthropol. 19, 227235.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, NY. [See alsoGoogle ScholarGoogle Scholar
Held, L.I. Jr., Duarte, C.M., and Derakhshanian, K. (1986). Extra joints and misoriented bristles on Drosophila legs. In Slavkin, H. (ed.), Progress in Developmental Biology (Part A). Alan R. Liss, New York, NY, pp. 293296.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. Roux's Arch. Dev. Biol. 195, 145157.Google Scholar
Helenius, I.T., and Beitel, G.J. (2008). The first “Slit” is the deepest: the secret to a hollow heart. J. Cell Biol. 182, 221223.Google Scholar
Helfrich-Förster, C. (2004). The circadian clock in the brain: a structural and functional comparison between mammals and insects. J. Comp. Physiol. A 190, 601613.Google Scholar
Hellige, J.B. (1993). Hemispheric Asymmetry: What's Right and What's Left. Harvard University Press, Cambridge, MA.Google Scholar
Helmstädter, M., Lüthy, K., Gödel, M., Simons, M., Ashish, , Nihalani, D., Rensing, S.A., Fischbach, K.-F., and Huber, T.B. (2012). Functional study of mammalian Neph proteins in Drosophila melanogaster. PLoS ONE 7, #7, e40300.Google Scholar
Hemani, Y., and Soller, M. (2012). Mechanisms of Drosophila Dscam mutually exclusive splicing regulation. Biochem. Soc. Trans. 40, 804809.Google Scholar
Hering, L., Henze, M.J., Kohler, M., Kelber, A., Bleidorn, C., Leschke, M., Nickel, B., Meyer, M., Kircher, M., Sunnucks, P., and Mayer, G. (2012). Opsins in Onychophora (velvet worms) suggest a single origin and subsequent diversification of visual pigments in arthropods. Mol. Biol. Evol. 29, 34513458.Google Scholar
Herranz, H., Eichenlaub, T., and Cohen, S.M. (2016). Cancer in Drosophila: imaginal discs as a model for epithelial tumor formation. Curr. Top. Dev. Biol. 116, 181199.Google Scholar
Herrera, E., and Garcia-Frigola, C. (2008). Genetics and development of the optic chiasm. Frontiers Biosci. 13, 16461653.Google Scholar
Herron, J.C., Freeman, S., Hodin, J., Miner, B., and Sidor, C. (2014). Evolutionary Analysis, 5th edn. Pearson, Upper Saddle River, NJ.Google Scholar
Heyn, P., Kalinka, A.T., Tomancak, P., and Neugebauer, K.M. (2014). Introns and gene expression: cellular constraints, transcriptional regulation, and evolutionary consequences. BioEssays 37, 148154.Google Scholar
Hibino, T., Ishii, Y., Levin, M., and Nishino, A. (2006). Ion flow regulates left-right asymmetry in sea urchin development. Dev. Genes Evol. 216, 265276.Google Scholar
Hibino, T., Nishino, A., and Amemiya, S. (2006). Phylogenetic correspondence of the body axes in bliaterians is revealed by the right-sided expression of Pitx genes in echinoderm larvae. Dev. Growth Differ. 48, 587595.Google Scholar
Hilbrant, M., Almudi, I., Leite, D.J., Kuncheria, L., Posnien, N., Nunes, M.D.S., and McGregor, A.P. (2014). Sexual dimorphism and natural variation within and among species in the Drosophila retinal mosaic. BMC Evol. Biol. 14, Article 240.Google Scholar
Hildebrand, J.G., and Shepherd, G.M. (1997). Mechanisms of olfactory discrimination: converging evidence for common principles across phyla. Annu. Rev. Neurosci. 20, 595631.Google Scholar
Himanen, J.-P., Saha, N., and Nikolov, D.B. (2007). Cell-cell signaling via Eph receptors and ephrins. Curr. Opin. Cell Biol. 19, 534542.Google Scholar
Hires, S.A., Pammer, L., Svoboda, K., and Golomb, D. (2013). Tapered whiskers are required for active tactile sensation. eLife 2, e01350.Google Scholar
Hiromi, Y., and Gehring, W.J. (1987). Regulation and function of the Drosophila segmentation gene fushi tarazu. Cell 50, 963974.Google Scholar
Hirth, F. (2010). On the origin and evolution of the tripartite brain. Brain Behav. Evol. 76, 310.Google Scholar
Hirth, F., Kammermeier, L., Frei, E., Walldorf, U., Noll, M., and Reichert, H. (2003). An urbilaterian origin of the tripartite brain: developmental genetic insights from Drosophila. Development 130, 23652373.Google Scholar
Hirth, F., and Reichert, H. (1999). Conserved genetic programs in insect and mammalian brain development. BioEssays 21, 677684.Google Scholar
Hisatomi, O., and Tokunaga, F. (2002). Molecular evolution of proteins involved in vertebrate phototransduction. Comp. Biochem. Physiol. B 133, 509522.Google Scholar
Hoare, D.J., McCrohan, C.R., and Cobb, M. (2008). Precise and fuzzy coding by olfactory sensory neurons. J. Neurosci. 28, #39, 97109722.Google Scholar
Hodin, J. (2000). Plasticity and constraints in development and evolution. J. Exp. Zool. 288, 120.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, #42, 96699679.Google Scholar
Hofmeyer, K., and Treisman, J. (2008). Sensory systems: seeing the world in a new light. Curr. Biol. 18, R919R921.Google Scholar
Hogan, B.L.M. (1995). Upside-down ideas vindicated. Nature 376, 210211.Google Scholar
Holland, L.Z. (2000). Body-plan evolution in the Bilateria: early antero-posterior patterning and the deuterostome-protostome dichotomy. Curr. Opin. Genet. Dev. 10, 434442.Google Scholar
Holland, L.Z. (2005). Non-neural ectoderm is really neural: evolution of developmental patterning mechanisms in the non-neural ectoderm of chordates and the problem of sensory cell homologies. J. Exp. Zool. B. Mol. Dev. Evol. 304, 304323.Google Scholar
Holland, L.Z. (2014). Genomics, evolution and development of amphioxus and tunicates: the Goldilocks Principle. J. Exp. Zool. B. Mol. Dev. Evol. 324, 342352.Google Scholar
Holland, L.Z. (2015). Evolution of basal deuterostome nervous systems. J. Exp. Biol. 218, 637645.Google Scholar
Holland, L.Z. (2015). The origin and evolution of chordate nervous systems. Philos. Trans. R. Soc. Lond. B 370, 20150048.Google Scholar
Holland, L.Z. (2016). Tunicates. Curr. Biol. 26, R141R156.Google Scholar
Holland, L.Z., Carvalho, J.E., Escriva, H., Laudet, V., Schubert, M., Shimeld, S.M., and Yu, J.-K. (2013). Evolution of bilaterian central nervous systems: a single origin? EvoDevo 4, Article 27.Google Scholar
Holland, L.Z., and Holland, N.D. (1996). Expression of AmphiHox-1 and AmphiPax-1 in amphioxus embryos treated with retinoic acid: insights into evolution and patterning of the chordate nerve cord and pharanx. Development 122, 18291838.Google Scholar
Holland, L.Z., and Holland, N.D. (1999). Chordate origins of the vertebrate central nervous system. Curr. Opin. Neurobiol. 9, 596602.Google Scholar
Holland, L.Z., and Holland, N.D. (2007). A revised fate map for amphioxus and the evolution of axial patterning in chordates. Integr. Comp. Biol. 47, 360372.Google Scholar
Holland, L.Z., Kene, M., Williams, N.A., and Holland, N.D. (1997). Sequence and embryonic expression of the amphioxus engrailed gene (AmphiEn): the metameric pattern of transcription resembles that of its segment-polarity homolog in Drosophila. Development 124, 17231732.Google Scholar
Holland, N.D. (2003). Early central nervous system evolution: an era of skin brains? Nat. Rev. Neurosci. 4, 111.Google Scholar
Holland, N.D., Holland, L.Z., and Holland, P.W.H. (2015). Scenarios for the making of vertebrates. Nature 520, 450455.Google Scholar
Holland, P.W.H. (2001). Beyond the Hox: how widespread is homeobox gene clustering? J. Anat. 199, 1323.Google Scholar
Holland, P.W.H. (2013). Evolution of homeobox genes. WIREs Dev. Biol. 2, 3145.Google Scholar
Holland, P.W.H., Booth, H.A.F., and Bruford, E.A. (2007). Classification and nomenclature of all human homeobox genes. BMC Biol. 5, Article 47.Google Scholar
Holley, S.A., Jackson, P.D., Sasai, Y., Lu, B., De Robertis, E.M., Hoffmann, F.M., and Ferguson, E.L. (1995). A conserved system for dorsal-ventral patterning in insects and vertebrates involving sog and chordin. Nature 376, 249253.Google Scholar
Holt, R.D. (2000). Use it or lose it. Nature 407, 689690.Google Scholar
Honda, H., Kodama, R., Takeuchi, T., Yamanaka, H., Watanabe, K., and Eguchi, G. (1984). Cell behaviour in a polygonal cell sheet. J. Embyrol. Exp. Morph. 83 (Suppl.), 313327.Google Scholar
Hong, W., and Luo, L. (2014). Genetic control of wiring specificity in the fly olfactory system. Genetics 196, 1729.Google Scholar
Honigberg, L., and Kenyon, C. (2000). Establishment of left/right asymmetry in neuroblast migration by UNC-40/DCC, UNC-73/Trio and DPY-19 proteins in C. elegans. Development 127, 46554668.Google Scholar
Hornstein, E.P., O'Carroll, D.C., Anderson, J.C., and Laughlin, S.B. (2000). Sexual dimorphism matches photoreceptor performance to behavioural requirements. Proc. R. Soc. Lond. B 267, 21112117.Google Scholar
Hosaka, Y., Saito, T., Sugita, S., Hikata, T., Kobayashi, H., Fukai, A., Taniguchi, Y., Hirata, M., Akiyama, H., Chung, U.-i., and Kawaguchi, H. (2013). Notch signaling in chondrocytes modulates endochondral ossification and osteoarthritis development. PNAS 110, #5, 18751880.Google Scholar
Houle, M., Sylvestre, J.-R., and Lohnes, D. (2003). Retinoic acid regulates a subset of Cdx1 function in vivo. Development 130, 65556567.Google Scholar
Houweling, A.C., Dildrop, R., Peters, T., Mummenhoff, J., Moorman, A.F.M., Rüther, U., and Christoffels, V.M. (2001). Gene and cluster-specific expression of the Iroquois family members during mouse development. Mech. Dev. 107, 169174.Google Scholar
Howard, J., Grill, S.W., and Bois, J.S. (2011). Turing's next steps: the mechanochemical basis of morphogenesis. Nat. Rev. Mol. Cell Biol. 12, 400406.Google Scholar
Hoy, R.R. (2012). Convergent evolution of hearing. Science 338, 894895.Google Scholar
Hoyle, C.H.V. (2011). Evolution of neuronal signalling: transmitters and receptors. Autonomic Neurosci. Basic Clin. 165, 2853.Google Scholar
Hozumi, S., Maeda, R., Taniguchi, K., Kanai, M., Shirakabe, S., Sasamura, T., Spéder, P., Noselli, S., Aigaki, T., Murakami, R., and Matsuno, K. (2006). An unconventional myosin in Drosophila reverses the default handedness in visceral organs. Nature 440, 798802.Google Scholar
Hozumi, S., Maeda, R., Taniguchi-Kanai, M., Okumura, T., Taniguchi, K., Kawakatsu, Y., Nakazawa, N., Hatori, R., and Matsuno, K. (2008). Head region of unconventional Myosin I family members is responsible for the organ-specificity of their roles in left-right polarity in Drosophila. Dev. Dyn. 237, 35283537.Google Scholar
Hsu, Y.-C., Chuang, J.-Z., and Sung, C.-H. (2015). Light regulates the ciliary protein transport and outer segment disc renewal of mammalian photoreceptors. Dev. Cell 32, 731742.Google Scholar
Huang, J., Wang, Y., Raghavan, S., Feng, S., Kiesewetter, K., and Wang, J. (2011). Human down syndrome cell adhesion molecules (DSCAMs) are functionally conserved with Drosophila Dscam[TM1] isoforms in controlling neurodevelopment. Insect Biochem. Mol. Biol. 41, 778787.Google Scholar
Huang, R., Zhi, Q., Schmidt, C., Wilting, J., Brand-Saberi, B., and Christ, B. (2000). Sclerotomal origin of the ribs. Development 127, 527532.Google Scholar
Huberman, A.D. (2009). Mammalian DSCAMs: they won't help you find a partner, but they'll guarantee you some personal space. Neuron 64, 441443.Google Scholar
Hudspeth, A.J. (2008). Making an effort to listen: mechanical amplification in the ear. Neuron 59, 530545.Google Scholar
Hudspeth, A.J. (2014). Integrating the active process of hair cells with cochlear function. Nat. Rev. Neurosci. 15, 600614.Google Scholar
Hueber, S.D., Weiller, G.F., Djordjevic, M.A., and Frickey, T. (2010). Improving Hox protein classification across the major model organisms. PLoS ONE 5, #5, e10820.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, C.L., and Kaufman, T.C. (2002). Exploring the myriapod body plan: expression patterns of the ten Hox genes in a centipede. Development 129, 12251238.Google Scholar
Hughes, C.L., and Kaufman, T.C. (2002). Hox genes and the evolution of the arthropod body plan. Evol. Dev. 4, 459499.Google Scholar
Hughes, C.L., Liu, P.Z., and Kaufman, T.C. (2004). Expression patterns of the rogue Hox genes Hox3/zen and fushi tarazu in the apterygote insect Thermobia domestica. Evol. Dev. 6, 393401.Google Scholar
Hughes, N.C. (2003). Trilobite body patterning and the evolution of arthropod tagmosis. BioEssays 25, 386395.Google Scholar
Hui, J.H.L., McDougall, C., Monteiro, A.S., Holland, P.W.H., Arendt, D., Balavoine, G., and Ferrier, D.E.K. (2012). Extensive chordate and annelid macrosynteny reveals ancestral homeobox gene organization. Mol. Biol. Evol. 29, 157165.Google Scholar
Hummel, K.P., and Chapman, D.B. (1959). Visceral inversion and associated anomalies in the mouse. J. Hered. 50, 913.Google Scholar
Hummel, T., Vasconcelos, M.L., Clemens, J.C., Fishilevich, Y., Vosshall, L.B., and Zipursky, S.L. (2003). Axonal targeting of olfactory receptor neurons in Drosophila is controlled by Dscam. Neuron 37, 221231.Google Scholar
Hummer, T.A., and McClintock, M.K. (2009). Putative human pheromone androstadienone attunes the mind specifically to emotional information. Horm. Behav. 55, 548559.Google Scholar
Hunnekuhl, V.S., and Akam, M. (2014). An anterior medial cell population with an apical-organ-like transcriptional profile that pioneers the central nervous system in the centipede Strigamia maritima. Dev. Biol. 396, 136149.Google Scholar
Hunt, D.M., Carvalho, L.S., Cowing, J.A., and Davies, W.L. (2009). Evolution and spectral tuning of visual pigments in birds and mammals. Philos. Trans. R. Soc. Lond. B 364, 29412955.Google Scholar
Hunt, D.M., and Collin, S.P. (2014). The evolution of photoreceptors and visual photopigments in vertebrates. In Hunt, D.M., Hankins, M.W., Collin, S.P., and Marshall, N.J. (eds.), Evolution of Visual and Non-Visual Pigments. Springer, New York, NY, pp. 163217.Google Scholar
Hunt, D.M., Dulai, K.S., Partridge, J.C., Cottrill, P., and Bowmaker, J.K. (2001). The molecular basis for spectral tuning of rod visual pigments in deep-sea fish. J. Exp. Biol. 204, 33333344.Google Scholar
Hunt, P., and Krumlauf, R. (1992). Hox codes and positional specification in vertebrate embryonic axes. Annu. Rev. Cell Biol. 8, 227256.Google Scholar
Hunter, T. (2012). Why nature chose phosphates to modify proteins. Philos. Trans. R. Soc. Lond. B 367, 25132516.Google Scholar
Hyman, S.E. (2005). Neurotransmitters. Curr. Biol. 15, R154R158.Google Scholar
Ide, H. (2012). Bone pattern formation in mouse limbs after amputation at the forearm level. Dev. Dyn. 241, 435441.Google Scholar
Ifkovits, J.L., Addis, R.C., Epstein, J.A., and Gearhart, J.D. (2014). Inhibition of TGFβ signaling increases direct conversion of fibroblasts to induced cardiomyocytes. PLoS ONE 9, #2, e89678.Google Scholar
Ikuta, T., Yoshida, N., Satoh, N., and Saiga, H. (2004). Ciona intestinalis Hox gene cluster: its dispersed structure and residual colinear expression in development. PNAS 101, #42, 1511815123.Google Scholar
Illergard, K., Ardell, D.H., and Elofsson, A. (2009). Structure is three to ten times more conserved than sequence: a study of structural response in protein cores. Proteins 77, 499508.Google Scholar
Im, S.H., and Galko, M.J. (2012). Pokes, sunburn, and hot sauce: Drosophila as an emerging model for the biology of nociception. Dev. Dyn. 241, 1626.Google Scholar
Imai, T. (2012). Positional information in neural map development: lessons from the olfactory system. Dev. Growth Differ. 54, 358365.Google Scholar
Imai, T. (2014). Construction of functional neuronal circuitry in the olfactory bulb. Semin. Cell Dev. Biol. 35, 180188.Google Scholar
Imai, T., Sakano, H., and Vosshall, L.B. (2010). Topographic mapping: the olfactory system. Cold Spring Harb. Perspect. Biol. 2, a001776.Google Scholar
Inbal, A., Halachmi, N., Dibner, C., Frank, D., and Salzberg, A. (2001). Genetic evidence for the transcriptional-activating function of Homothorax during adult fly development. Development 128, 34053413.Google Scholar
Ingber, D.E., and Levin, M. (2007). What lies at the interface of regenerative medicine and developmental biology? Development 134, 25412547.Google Scholar
Ingham, P.W., and Martinez-Arias, A. (1986). The correct activation of Antennapedia and bithorax complex genes requires the fushi tarazu gene. Nature 324, 592597.Google Scholar
Innan, H., and Kondrashov, F. (2010). The evolution of gene duplications: classifying and distinguishing between models. Nat. Rev. Genet. 11, 97108.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
Irimia, M., Piñeiro, C., Maeso, I., Gómez-Skarmeta, J.L., Casares, F., and Garcia-Fernàndez, J. (2010). Conserved developmental expression of Fezf in chordates and Drosophila and the origin of the Zona Limitans Intrathalamica (ZLI) brain organizer. EvoDevo 1, Article 7.Google Scholar
Irimia, M., Tena, J.J., Alexis, M.S., Fernandez-Miñan, A., Maeso, I., Bogdanovic, O., de la Calle-Mustienes, E., Roy, S.W., Gómez-Skarmeta, J.L., and Fraser, H.B. (2012). Extensive conservation of ancient microsynteny across metazoans due to cis-regulatory constraints. Genome Res. 22, 23562367.Google Scholar
Irish, V.F., Martinez-Arias, A., and Akam, M. (1989). Spatial regulation of the Antennapedia and Ultrabithorax homeotic genes during Drosophila early development. EMBO J. 8, 15271537.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 Wieschaus, E. (1994). fringe, a boundary-specific signaling molecule, mediates interactions between dorsal and ventral cells during Drosophila wing development. Cell 79, 595606.Google Scholar
Irvine, S.Q., and Martindale, M.Q. (2001). Comparative analysis of Hox gene expression in the polychaete Chaetopterus: implications for the evolution of body plan regionalization. Am. Zool. 41, 640651.Google Scholar
Isayama, T., Chen, Y., Kono, M., Fabre, E., Slavsky, M., DeGrip, W.J., Ma, J.-x., Crouch, R.K., and Makino, C.L. (2013). Coexpression of three opsins in cone photoreceptors of the salamander Ambystoma tigrinum. J. Comp. Neurol. 522, 22492265.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
Itskov, P.M., and Ribeiro, C. (2013). The dilemmas of the gourmet fly: the molecular and neuronal mechanisms of feeding and nutrient decision making in Drosophila. Front. Neurosci. 7, Article 12.Google Scholar
Iwai, M., Yokono, M., and Nakano, A. (2014). Visualizing structural dynamics of thylakoid membranes. Sci. Rep. 4, Article 3768.Google Scholar
Iwaki, D.D., and Lengyel, J.A. (2002). A Delta-Notch signaling border regulated by Engrailed/Invected repression specifies boundary cells in the Drosophila hindgut. Mech. Dev. 114, 7184.Google Scholar
Iwasaki, Y., Hosoya, T., Takebayashi, H., Ogawa, Y., Hotta, Y., and Ikenaka, K. (2003). The potential to induce glial differentiation is conserved between Drosophila and mammalian glial cells missing genes. Development 130, 60276035.Google Scholar
Izaddoost, S., Nam, S.-C., Bhat, M.A., Bellen, H.J., and Choi, K.-W. (2002). Drosophila Crumbs is a positional cue in photoreceptor adherens junctions and rhabdomeres. Nature 416, 178182.Google Scholar
Izpisúa Belmonte, J.C. (1999). How the body tells left from right. Sci. Am. 280, #6, 4651.Google Scholar
Jablonski, N.G. (2010). The naked truth. Sci. Am. 302, #2, 4249.Google Scholar
Jack, T., Regulski, M., and McGinnis, W. (1988). Pair-rule segmentation genes regulate the expression of the homeotic selector gene, Deformed. Genes Dev. 2, 635651.Google Scholar
Jackson, D.J., Meyer, N.P., Seaver, E., Pang, K., McDougall, C., Moy, V.N., Gordon, K., Degnan, B.M., Martindale, M.Q., Burke, R.D., and Peterson, K.J. (2010). Developmental expression of COE across the Metazoa supports a conserved role in neuronal cell-type specification and mesodermal development. Dev. Genes Evol. 220, 221234.Google Scholar
Jacobo, A., and Hudspeth, A.J. (2014). Reaction–diffusion model of hair-bundle morphogenesis. PNAS 111, #43, 1544415449.Google Scholar
Jacobs, D.K., Hughes, N.C., Fitz-Gibbon, S.T., and Winchell, C.J. (2005). Terminal addition, the Cambrian radiation and the Phanerozoic evolution of bilaterian form. Evol. Dev. 7, 498514.Google Scholar
Jacobs, D.K., Nakanishi, N., Yuan, D., Camara, A., Nichols, S.A., and Hartenstein, V. (2007). Evolution of sensory structures in basal metazoa. Integr. Comp. Biol. 47, 712723.Google Scholar
Jacobs, G.H. (2012). The evolution of vertebrate color vision. In López-Larrea, C. (ed.), Sensing in Nature. Springer, New York, NY, pp. 156172.Google Scholar
Jacobs, G.H., and Nathans, J. (2009). The evolution of primate color vision. Sci. Am. 300, #4, 5663.Google Scholar
Jacobs, G.H., Williams, G.A., Cahill, H., and Nathans, J. (2007). Emergence of novel color vision in mice engineered to express a human cone photopigment. Science 315, 17231725.Google Scholar
Jaeger, J. (2011). The gap gene network. Cell. Mol. Life Sci. 68, 243274.Google Scholar
Jaeger, J., Manu, , and Reinitz, J. (2012). Drosophila blastoderm patterning. Curr. Opin. Genet. Dev. 22, 533541.Google Scholar
Jafari, S., and Alenius, M. (2015). Cis-regulatory mechanisms for robust olfactory sensory neuron class-restricted odorant receptor gene expression in Drosophila. PLoS Genet. 11, #3, e1005051.Google Scholar
Jafari, S., Alkhori, L., Schleiffer, A., Brochtrup, A., Hummel, T., and Alenius, M. (2012). Combinatorial activation and repression by seven transcription factors specify Drosophila odorant receptor expression. PLoS Biol. 10, #3, e1001280.Google Scholar
Jahan, I., Pan, N., Elliott, K.L., and Fritzsch, B. (2015). The quest for restoring hearing: understanding ear development more completely. BioEssays 37, 10161027.Google Scholar
Jane, S.M., Ting, S.B., and Cunningham, J.M. (2005). Epidermal impermeable barriers in mouse and fly. Curr. Opin. Genet. Dev. 15, 447453.Google Scholar
Jankowski, R. (ed.) (2013). The Evo-Devo Origin of the Nose, Anterior Skull Base and Midface. Spinger, New York, NY.Google Scholar
Janssen, R., and Budd, G.E. (2013). Deciphering the onychophoran “segmentation gene cascade”: gene expression reveals limited involvement of pair rule gene orthologs in segmentation, but a highly conserved segment polarity gene network. Dev. Biol. 382, 224234.Google Scholar
Janssen, R., and Budd, G.E. (2016). Gene expression analysis reveals that Delta/Notch signalling is not involved in onychophoran segmentation. Dev. Genes Evol. 226, 6977.Google Scholar
Janssen, R., and Damen, W.G.M. (2006). The ten Hox genes of the millipede Glomeris marginata. Dev. Genes Evol. 216, 451465.Google Scholar
Janssen, R., Eriksson, B.J., Budd, G.E., Akam, M., and Prpic, N.-M. (2010). Gene expression patterns in an onychophoran reveal that regionalization predates limb segmentation in pan-arthropods. Evol. Dev. 12, 363372.Google Scholar
Janssen, R., Eriksson, B.J., Tait, N.N., and Budd, G.E. (2014). Onychophoran Hox genes and the evolution of arthropod Hox gene expression. Front. Zool. 11, Article 22.Google Scholar
Janssen, R., Jörgensen, M., Prpic, N.-M., and Budd, G.E. (2015). Aspects of dorso-ventral and proximo-distal limb patterning in onychophorans. Evol. Dev. 17, 2133.Google Scholar
Jarman, A.P. (2000). Developmental genetics: vertebrates and insects see eye to eye. Curr. Biol. 10, R857R859.Google Scholar
Jarman, A.P. (2002). Studies of mechanosensation using the fly. Hum. Mol. Genet. 11, #10, 12151218.Google Scholar
Jarman, A.P., and Groves, A.K. (2013). The role of Atonal transcription factors in the development of mechanosensitive cells. Semin. Cell Dev. Biol. 24, 438447.Google Scholar
Jaumouillé, E., Almeida, P.M., Stähli, P., Koch, R., and Nagoshi, E. (2015). Transcriptional regulation via nuclear receptor crosstalk required for the Drosophila circadian clock. Curr. Biol. 25, 15021508.Google Scholar
Jaw, T.J., You, L.-R., Knoepfler, P.S., Yao, L.-C., Pai, C.-Y., Tang, C.-Y., Chang, L.-P., Berthelsen, J., Blasi, F., Kamps, M.P., and Sun, Y.H. (2000). Direct interaction of two homeoproteins, Homothorax and Extradenticle, is essential for EXD nuclear localization and function. Mech. Dev. 91, 279291.Google Scholar
Jaworski, A., Tom, I., Tong, R.K., Gildea, H.K., Koch, A.W., Gonzalez, L.C., and Tessier-Lavigne, M. (2015). Operational redundancy in axon guidance through the multifunctional receptor Robo3 and its ligand NELL2. Science 350, 961965.Google Scholar
Jefferis, G.S.X.E., Potter, C.J., Chan, A.M., Marin, E.C., Rohlfing, T., Maurer, C.R. Jr., and Luo, L. (2007). Comprehensive maps of Drosophila higher olfactory centers: spatially segregated fruit and pheromone representation. Cell 128, 11871203.Google Scholar
Jeffery, G., and Erskine, L. (2005). Variations in the architecture and development of the vertebrate optic chiasm. Prog. Retin. Eye Res. 24, 721753.Google Scholar
Jékely, G. (2013). Global view of the evolution and diversity of metazoan neuropeptide signaling. PNAS 110, #21, 87028707.Google Scholar
Jékely, G., Paps, J., and Nielsen, C. (2015). The phylogenetic position of ctenophores and the origin(s) of nervous systems. EvoDevo 6, Article 1. [See alsoGoogle ScholarGoogle Scholar
Jensen, B., Wang, T., Christoffels, V.M., and Moorman, A.F.M. (2013). Evolution and development of the building plan of the vertebrate heart. Biochim. Biophys. Acta 1833, 783794.Google Scholar
Jessell, T.M. (2000). Neuronal specification in the spinal cord: inductive signals and transcription codes. Nat. Rev. Genet. 1, 2029.Google Scholar
Ji, C., Wu, L., Zhao, W., Wang, S., and Lv, J. (2012). Echinoderms have bilateral tendencies. PLoS ONE 7, #1, e28978.Google Scholar
Jiang, H., and Edgar, B.A. (2012). Intestinal stem cell function in Drosophila and mice. Curr. Opin. Genet. Dev. 22, 354360.Google Scholar
Jiang, X., Shen, S., Cadwell, C.R., Berens, P., Sinz, F., Ecker, A.S., Patel, S., and Tolias, A.S. (2015). Principles of connectivity among morphologically defined cell types in the adult neocortex. Science 350, 1055.Google Scholar
Jiao, Y., Lau, T., Hatzikirou, H., Meyer-Hermann, M., Corbo, J.C., and Torquato, S. (2014). Avian photoreceptor patterns represent a disordered hyperuniform solution to a multiscale packing problem. Phys. Rev. E 89, Article 022721.Google Scholar
Jimenez-Gurl, E., and Pujades, C. (2011). An ancient mechanism of hindbrain patterning has been conserved in vertebrate evolution. Evol. Dev. 13, 3846.Google Scholar
Jockusch, E.L., Williams, T.A., and Nagy, L. (2004). The evolution of patterning of serially homologous appendages in insects. Dev. Genes Evol. 214, 324338.Google Scholar
Joffe, B., Peichl, L., Hendrickson, A., Leonhardt, H., and Solovei, I. (2014). Diurnality and nocturnality in primates: an analysis from the rod photoreceptor nuclei perspective. Evol. Biol. 41, 111.Google Scholar
Johansson, J.A., and Headon, D.J. (2014). Regionalisation of the skin. Semin. Cell Dev. Biol. 25–26, 310.Google Scholar
Johnsen, S. (2012). The Optics of Life: A Biologist's Guide to Light in Nature. Princeton University Press, Princeton, NJ.Google Scholar
Johnson, J.-L.F., and Leroux, M.R. (2010). cAMP and cGMP signaling: sensory systems with prokaryotic roots adopted by eukaryotic cilia. Trends Cell Biol. 20, 435444.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
Johnston, R.J. Jr., and Hobert, O. (2003). A microRNA controlling left/right neuronal asymmetry in Caenorhabditis elegans. Nature 426, 845849.Google Scholar
Johnston, R.J. Jr., Otake, Y., Sood, P., Vogt, N., Behnia, R., Vasiliauskas, D., McDonald, E., Xie, B., Koenig, S., Wolf, R., Cook, T., Gebelein, B., Kussell, E., Nakagoshi, H., and Desplan, C. (2011). Interlocked feedforward loops control cell-type-specific rhodopsin expression in the Drosophila eye. Cell 145, 956968.Google Scholar
Joliot, A., and Prochiantz, A. (2008). Homeoproteins as natural Penetratin cargoes with signaling properties. Adv. Drug Deliv. Rev. 60, 608613.Google Scholar
Joly, W., Mugat, B., and Maschat, F. (2007). Engrailed controls the organization of the ventral nerve cord through frazzled regulation. Dev. Biol. 301, 542554.Google Scholar
Jonasova, K., and Kozmik, Z. (2008). Eye evolution: lens and cornea as an upgrade of animal visual system. Semin. Cell Dev. Biol. 19, 7181.Google Scholar
Jones, C., Qian, D., Kim, S.M., Li, S., Ren, D., Knapp, L., Sprinzak, D., Avraham, K.B., Matsuzaki, F., Chi, F., and Chen, P. (2014). Ankrd6 is a mammalian functional homolog of Drosophila planar cell polarity gene diego and regulates coordinated cellular orientation in the mouse inner ear. Dev. Biol. 395, 6272.Google Scholar
Joseph, R.M., and Carlson, J.R. (2015). Drosophila chemoreceptors: a molecular interface between the chemical world and the brain. Trends Genet. 31, 683695.Google Scholar
Jost, M., Fernández-Zapata, J., Polanco, M.C., Ortiz-Guerrero, J.M., Chen, P.Y.-T., Kang, G., Padmanabhan, S., Elías-Arnanz, M., and Drennan, C.L. (2015). Structural basis for gene regulation by a B12-dependent photoreceptor. Nature 526, 536541.Google Scholar
Ju, Y.-E.S., Lucey, B.P., and Holtzman, D.M. (2014). Sleep and Alzheimer disease pathology: a bidirectional relationship. Nat. Rev. Neurol. 10, 115119.Google Scholar
Jung, H., and Dasen, J.S. (2015). Evolution of patterning systems and circuit elements for locomotion. Dev. Cell 32, 408422.Google Scholar
Kageyama, R., Ohtsuka, T., Shimojo, H., and Imayoshi, I. (2009). Dynamic regulation of Notch signaling in neural progenitor cells. Curr. Opin. Cell Biol. 21, 733740.Google Scholar
Kainz, F., Ewen-Campen, B., Akam, M., and Extavour, C.G. (2011). Notch/Delta signalling is not required for segment generation in the basally branching insect Gryllus bimaculatus. Development 138, 50155026.Google Scholar
Kaiser, M. (2015). Neuroanatomy: Connectome connects fly and mammalian brain networks. Curr. Biol. 25, R409R430.Google Scholar
Kaltenbach, J.A., Falzarano, P.R., and Simpson, T.H. (1994). Postnatal development of the hamster cochlea. II. Growth and differentiation of stereocilia bundles. J. Comp. Neurol. 350, 187198.Google Scholar
Kaltezioti, V., Kouroupi, G., Oikonomaki, M., Mantouvalou, E., Stergiopoulos, A., Charonis, A., Rohrer, H., Matsas, R., and Politis, P.K. (2010). Prox1 regulates the Notch1-mediated inhibition of neurogenesis. PLoS Biol. 8, #12, e1000565.Google Scholar
Kamikouchi, A. (2013). Auditory neuroscience in fruit flies. Neurosci. Res. 76, 113118.Google Scholar
Kamikouchi, A., Inagaki, H.K., Effertz, T., Hendrich, O., Fiala, A., Göpfert, M.C., and Ito, K. (2009). The neural basis of Drosophila gravity-sensing and hearing. Nature 458, 165171.Google Scholar
Kamikouchi, A., Shimada, T., and Ito, K. (2006). Comprehensive classification of the auditory sensory projections in the brain of the fruit fly Drosophila melanogaster. J. Comp. Neurol. 499, 317356.Google Scholar
Kammandel, B., Chowdhury, K., Stoykova, A., Aparicio, S., Brenner, S., and Gruss, P. (1999). Distinct cis-essential modules direct the time-space pattern of the Pax6 gene activity. Dev. Biol. 205, 7997.Google Scholar
Kammermeier, L., and Reichert, H. (2001). Common developmental genetic mechanisms for patterning invertebrate and vertebrate brains. Brain Res. Bull. 55, 675682.Google Scholar
Kandel, E.R. (2012). The molecular biology of memory: cAMP, PKA, CRE, CREB-1, CREB-2, and CPEB. Mol. Brain 5, Article 14.Google Scholar
Kang, J., Hu, J., Karra, R., Dickson, A.L., Tornini, V.A., Nachtrab, G., Gemberling, M., Goldman, J.A., Black, B.L., and Poss, K.D. (2016). Modulation of tissue repair by regeneration enhancer elements. Nature 532, 201206.Google Scholar
Kang, K., Pulver, S.R., Panzano, V.C., Chang, E.C., Griffith, L.C., Theobald, D.L., and Garrity, P.A. (2010). Analysis of Drosophila TRPA1 reveals an ancient origin for human chemical noniception. Nature 464, 597600.Google Scholar
Kantor, D.B., and Kolodkin, A.L. (2003). Curbing the excesses of youth: molecular insights into axonal pruning. Neuron 38, 849852.Google Scholar
Kaprielian, Z., Imondi, R., and Runko, E. (2000). Axon guidance at the midline of the developing CNS. Anat. Rec. 261, 176197.Google Scholar
Kaprielian, Z., Runko, E., and Imondi, R. (2001). Axon guidance at the midline choice point. Dev. Dyn. 221, 154181.Google Scholar
Karadge, U.B., Gosto, M., and Nicotra, M.L. (2015). Allorecognition proteins in an invertebrate exhibit homophilic interactions. Curr. Biol. 25, 28452850.Google Scholar
Karavitaki, K.D., and Corey, D.P. (2010). Sliding adhesion confers coherent motion to hair cell stereocilia and parallel gating to transduction channels. J. Neurosci. 30, 90519063.Google Scholar
Kardon, G., Heanue, T.A., and Tabin, C.J. (2004). The Pax/Six/Eya/Dach network in development and evolution. In Schlosser, G. and Wagner, G.P. (eds.), Modularity in Development and Evolution. University of Chicago Press, Chicago, IL, pp. 5980.Google Scholar
Karten, H.J. (2015). Vertebrate brains and evolutionary connectomics: on the origins of the mammalian “neocortex”. Philos. Trans. R. Soc. Lond. B 370, 20150060.Google Scholar
Kasahara, M. (2007). The 2R hypothesis: an update. Curr. Opin. Immunol. 19, 547552.Google Scholar
Kashalikar, S.J. (1988). An explanation for the development of decussations in the central nervous system. Med. Hypotheses 26, 18.Google Scholar
Kato, H.E., Inoue, K., Abe-Yoshizumi, R., Kato, Y., Ono, H., Konno, M., Hososhima, S., Ishizuka, T., Hoque, M.R., Kunitomo, H., Ito, J., Yoshizawa, S., Yamashita, K., Takemoto, M., Nishizawa, T., Taniguchi, R., Kogure, K., Maturana, A.D., Iino, Y., Yawo, H., Ishitani, R., Kandori, H., and Nureki, O. (2015). Structural basis for Na+ transport mechanism by a light-driven Na+ pump. Nature 521, 4853.Google Scholar
Katta, S., Krieg, M., and Goodman, M.B. (2015). Feeling force: physical and physiological principles enabling sensory mechanotransduction. Annu. Rev. Cell Dev. Biol. 31, 347371.Google Scholar
Katz, B., and Minke, B. (2009). Drosophila photoreceptors and signaling mechanisms. Front. Cell. Neurosci. 3, Article 2.Google Scholar
Katz, M.J., and Grenander, U. (1982). Developmental matching and the numerical matching hypothesis for neuronal cell death. J. Theor. Biol. 98, 501517.Google Scholar
Katz, P.S. (2016). Evolution of central pattern generators and rhythmic behaviours. Philos. Trans. R. Soc. Lond. B 371, 20150057.Google Scholar
Katz, P.S., and Harris-Warrick, R.M. (1999). The evolution of neuronal circuits underlying species-specific behavior. Curr. Opin. Neurobiol. 9, 628633.Google Scholar
Kaufman, L. (2005). One fish, two fish, red fish, blue fish: why are coral reefs so colorful? Natl. Geogr. 207, #5, 86109.Google Scholar
Kaufman, T.C., Lewis, R., and Wakimoto, B. (1980). Cytogenetic analysis of chromosome 3 in Drosophila melanogaster: the homoeotic gene complex in polytene chromosome interval 84A-B. Genetics 94, 115133.Google Scholar
Kaupp, U.B. (2010). Olfactory signalling in vertebrates and insects: differences and commonalities. Nat. Rev. Neurosci. 11, 188200.Google Scholar
Kavaler, J., Fu, W., Duan, H., Noll, M., and Posakony, J.W. (1999). An essential role for the Drosophila Pax2 homolog in the differentiation of adult sensory organs. Development 126, 22612272.Google Scholar
Kavanau, J.L. (2001). Memory failures, dream illusions and mental malfunction. Neuropsychobiology 44, 199211.Google Scholar
Kavlie, R.G., and Albert, J.T. (2013). Chordotonal organs. Curr. Biol. 23, R334R335.Google Scholar
Kavlie, R.G., Fritz, J.L., Nies, F., Göpfert, M.C., Oliver, D., Albert, J.T., and Eberl, D.F. (2015). Prestin is an anion transporter dispensable for mechanical feedback amplification in Drosophila hearing. J. Comp. Physiol. A 201, 5160.Google Scholar
Kawamori, A., Shimaji, K., and Yamaguchi, M. (2012). Dynamics of endoreplication during Drosophila posterior scutellar macrochaete development. PLoS ONE 7, #6, e38714.Google Scholar
Kawasumi, A., Nakamura, T., Iwai, N., Yashiro, K., Saijoh, Y., Belo, J.A., Shiratori, H., and Hamada, H. (2011). Left-right asymmetry in the level of active Nodal protein produced in the node is translated into left-right asymmetry in the lateral plate of mouse embryos. Dev. Biol. 353, 321330.Google Scholar
Kay, J.N., Link, B.A., and Baier, H. (2005). Staggered cell-intrinsic timing of ath5 expression underlies the wave of ganglion cell neurogenesis in the zebrafish retina. Development 132, 25732585.Google Scholar
Kayser, M.S., Mainwaring, B., Yue, Z., and Sehgal, A. (2015). Sleep deprivation suppresses aggression in Drosophila. eLife 4, e07643.Google Scholar
Kazmierczak, P., and Müller, U. (2011). Sensing sound: molecules that orchestrate mechanotransduction by hair cells. Trends Neurosci. 35, 220229.Google Scholar
Kechad, A., Jolicoeur, C., Tufford, A., Mattar, P., Chow, R.W.Y., Harris, W.A., and Cayouette, M. (2012). Numb is required for the production of terminal asymmetric cell divisions in the developing mouse retina. J. Neurosci. 32, #48, 1719717210.Google Scholar
Kefalov, V., Fu, Y., Marsh-Armstrong, N., and Yau, K.-W. (2003). Role of visual pigment properties in rod and cone phototransduction. Nature 425, 526531.Google Scholar
Keil, T.A. (2012). Sensory cilia in arthropods. Arthropod Struct. Dev. 41, 515534.Google Scholar
Kelber, A., and Osorio, D. (2010). From spectral information to animal colour vision: experiments and concepts. Proc. R. Soc. Lond. B 277, 16171625.Google Scholar
Keller, R. (2012). Physical biology returns to morphogenesis. Science 338, 201203.Google Scholar
Kelly, M., and Chen, P. (2007). Shaping the mammalian auditory sensory organ by the planar cell polarity pathway. Int. J. Dev. Biol. 51, 535547.Google Scholar
Kelly, M.C., and Chen, P. (2009). Development of form and function in the mammalian cochlea. Curr. Opin. Neurobiol. 19, 395401.Google Scholar
Kelly, M.W. (2007). Cellular commitment and differentiation in the organ of Corti. Int. J. Dev. Biol. 51, 571583.Google Scholar
Kenaley, C.P., DeVaney, S.C., and Fjeran, T.T. (2013). The complex evolutionary history of seeing red: molecular phylogeny and the evolution of an adaptive visual system in deep-sea dragonfishes (Stomiiformes: Stomiidae). Evolution 68, 9961013.Google Scholar
Kennedy, B., and Malicki, J. (2009). What drives cell morphogenesis: a look inside the vertebrate photoreceptor. Dev. Dyn. 238, 21152138.Google Scholar
Kernan, M.J. (2007). Mechanotransduction and auditory transduction in Drosophila. Pflugers Arch. 454, 703720.Google Scholar
Kerner, P., Ikmi, A., Coen, D., and Vervoort, M. (2009). Evolutionary history of the iroquois/Irx genes in metazoans. BMC Evol. Biol. 9, Article 74.Google Scholar
Kerner, P., Simionato, E., Le Gouar, M., and Vervoort, M. (2009). Orthologs of key vertebrate neural genes are expressed during neurogenesis in the annelid Platynereis dumerilii. Evol. Dev. 11, 513524.Google Scholar
Kessel, M. (1992). Respecification of vertebral identities by retinoic acid. Development 115, 487501.Google Scholar
Kessel, M., and Gruss, P. (1991). Homeotic transformations of murine vertebrae and concomitant alteration of Hox codes induced by retinoic acid. Cell 67, 89104.Google Scholar
Kettle, C., Johnstone, J., Jowett, T., Arthur, H., and Arthur, W. (2003). The pattern of segment formation, as revealed by engrailed expression, in a centipede with a variable number of segments. Evol. Dev. 5, 198207.Google Scholar
Kidd, T., Bland, K.S., and Goodman, C.S. (1999). Slit is the midline repellent for the Robo receptor in Drosophila. Cell 96, 785794.Google Scholar
Kiecker, C., and Lumsden, A. (2005). Compartments and their boundaries in vertebrate brain development. Nat. Rev. Neurosci. 6, 553564.Google Scholar
Kiecker, C., and Lumsden, A. (2012). The role of organizers in patterning the nervous system. Annu. Rev. Neurosci. 35, 347367.Google Scholar
Kiernan, A.E. (2013). Notch signaling during cell fate determination in the inner ear. Semin. Cell Dev. Biol. 24, 470479.Google Scholar
Kim, H., Kim, K., and Yim, J. (2013). Biosynthesis of drosopterins, the red eye pigments of Drosophila melanogaster. IUBMB Life 65, 334340.Google Scholar
Kim, J.H., Jin, P., Duan, R., and Chen, E.H. (2015). Mechanisms of myoblast fusion during muscle development. Curr. Opin. Genet. Dev. 32, 162170.Google Scholar
Kim, S.-G., Ashe, J., Hendrich, K., Ellermann, J.M., Merkle, H., Ugurbil, K., and Georgopoulos, A.P. (1993). Functional magnetic resonance imaging of motor cortex: hemispheric asymmetry and handedness. Science 261, 615617.Google Scholar
Kim, S.Y., Paylor, S.W., Magnuson, T., and Schumacher, A. (2007). Juxtaposed Polycomb complexes co-regulate vertebral identity. Development 133, 49574968.Google Scholar
Kimelman, D., and Martin, B.L. (2011). Anterior-posterior patterning in early development: three strategies. WIREs Dev. Biol. 1, 253266.Google Scholar
Kimmel, C.B. (1996). Was Urbilateria segmented? Trends Genet. 12, 329331.Google Scholar
King, T., and Brown, N.A. (1995). The embryo's one-sided genes. Curr. Biol. 5, 13641366.Google Scholar
King, T., and Brown, N.A. (1999). Embryonic asymmetry: the left side gets all the best genes. Curr. Biol. 9, R18R22.Google Scholar
Kingsley, M.C.S., and Ramsay, M.A. (1988). The spiral in the tusk of the narwhal. Arctic 41, 236238.Google Scholar
Kinsbourne, M. (2013). Somatic twist: a model for the evolution of decussation. Neuropsychology 27, 511515.Google Scholar
Kipryushina, Y.O., Yakovlev, K.V., and Odintsova, N.A. (2015). Vascular endothelial growth factors: a comparison between invertebrates and vertebrates. Cytokine Growth Factor Rev. 26, 687695.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
Kiser, P.D., Golczak, M., and Palczewski, K. (2013). Chemistry of the retinoid (visual) cycle. Chem. Rev. 114, 194232.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
Klaus, A., Saga, Y., Taketo, M.M., Tzahor, E., and Birchmeier, W. (2007). Distinct roles of Wnt/β-catenin and Bmp signaling during early cardiogenesis. PNAS 104, #47, 1853118536.Google Scholar
Klein, R., and Kania, A. (2014). Ephrin signalling in the developing nervous system. Curr. Opin. Neurobiol. 27, 1624.Google Scholar
Klein, T., and Martinez Arias, A. (1998). Interactions among Delta, Serrate and Fringe modulate Notch activity during Drosophila wing development. Development 125, 29512962.Google Scholar
Klezovitch, O., and Vasioukhin, V. (2013). Your gut is right to turn left. Dev. Cell 26, 553554.Google Scholar
Kliethermes, C.L. (2015). Conservation of the ethanol-induced locomotor stimulant response among arthropods. Brain Behav. Evol. 85, 3746.Google Scholar
Klomp, J., Athy, D., Kwan, C.W., Bloch, N.I., Sandmann, T., Lemke, S., and Schmidt-Ott, U. (2015). A cysteine-clamp gene drives embryo polarity in the midge Chironomus. Science 348, 10401042.Google Scholar
Klug, W.S., Cummings, M.R., and Spencer, C. (2006). Concepts in Genetics, 8th edn. Pearson Prentice Hall, Upper Saddle River, NJ.Google Scholar
Kmita, M., and Duboule, D. (2003). Organizing axes in time and space: 25 years of colinear tinkering. Science 301, 331333.Google Scholar
Knaden, M., and Hansson, B.S. (2014). Mapping odor valence in the brain of flies and mice. Curr. Opin. Neurobiol. 24, 3438.Google Scholar
Kobayashi, D., and Takeda, H. (2012). Ciliary motility: the components and cytoplasmic preassembly mechanisms of the axonemal dyneins. Differentiation 83, S23S29.Google Scholar
Koh, T.-W., Gorur-Shandilya, S., Menuz, K., Larter, N.K., Stewart, S., and Carlson, J.R. (2014). The Drosophila IR20a clade of ionotropic receptors are candidate taste and pheromone receptors. Neuron 83, 850865.Google Scholar
Kohl, J., and Jefferis, G.S.X.E. (2011). Neuroanatomy: decoding the fly brain. Curr. Biol. 21, R19R20.Google Scholar
Kohler, R.E. (1994). Lords of the Fly: Drosophila Genetics and the Experimental Life. University of Chicago Press, Chicago, IL.Google Scholar
Koirala, S., and Ko, C.-P. (2004). Pruning an axon piece by piece: a new mode of synapse elimination. Neuron 44, 578580.Google Scholar
Kojima, T. (2004). The mechanism of Drosophila leg development along the proximodistal axis. Dev. Growth Differ. 46, 115129.Google Scholar
Konopova, B., and Akam, M. (2014). The Hox genes Ultrabithorax and abdominal-A specify three different types of abdominal appendage in the springtail Orchesella cincta (Collembola). EvoDevo 5, Article 2.Google Scholar
Konstantinides, N., and Averof, M. (2014). A common cellular basis for muscle regeneration in arthropods and vertebrates. Science 343, 788791.Google Scholar
Koop, D., Holland, N.D., Sémon, M., Alvarez, S., Rodriguez de Lera, A., Laudet, V., Holland, L.Z., and Schubert, M. (2010). Retinoic acid signaling targets Hox genes during the amphioxus gastrula stage: insights into early anterior-posterior patterning of the chordate body plan. Dev. Biol. 338, 98106.Google Scholar
Kopp, A. (2011). Drosophila sex combs as a model of evolutionary innovations. Evol. Dev. 13, 504522.Google Scholar
Kosaki, K., and Casey, B. (1998). Genetics of human left-right axis malformations. Semin. Cell Dev. Biol. 9, 8999.Google Scholar
Koschowitz, M.-C., Fischer, C., and Sander, M. (2014). Beyond the rainbow. Science 346, 416418.Google Scholar
Koshiba-Takeuchi, K., Mori, A.D., Kaynak, B.L., Cebra-Thomas, J., Sukonnik, T., Georges, R.O., Latham, S., Beck, L., Henkelman, R.M., Black, B.L., Olson, E.N., Wade, J., Takeuchi, J.K., Nemer, M., Gilbert, S.F., and Bruneau, B.G. (2009). Reptilian heart development and the molecular basis of cardiac chamber evolution. Nature 461, 9598.Google Scholar
Kotkamp, K., Klingler, M., and Schoppmeier, M. (2010). Apparent role of Tribolium orthodenticle in anteroposterior blastoderm patterning largely reflects novel functions in dorsoventral axis formation and cell survival. Development 137, 18531862.Google Scholar
Kottler, B., and van Swinderen, B. (2014). Taking a new look at how flies learn. eLife 3, e03978.Google Scholar
Koulakov, A., Gelperin, A., and Rinberg, D. (2007). Olfactory coding with all-or-nothing glomeruli. J. Neurophysiol. 98, 31343142.Google Scholar
Kowalczyk, A.P., and Moses, K. (2002). Photoreceptor cells in flies and mammals: crumby homology? Dev. Cell 2, 253254.Google Scholar
Koyama, E., Yasuda, T., Minugh-Purvis, N., Kinumatsu, T., Yallowitz, A.R., Wellik, D.M., and Pacifici, M. (2010). Hox11 genes establish synovial joint organization and phylogenetic characteristics in developing mouse zeugopod skeletal elements. Development 137, 37953800.Google Scholar
Koyanagi, M., Nagata, T., Katoh, K., Yamashita, S., and Tokunaga, F. (2008). Molecular evolution of arthropod color vision deduced from multiple opsin genes of jumping spiders. J. Mol. Evol. 66, 130137.Google Scholar
Kozlov, A.S., Baumgart, J., Risler, T., Versteegh, C.P.C., and Hudspeth, A.J. (2011). Forces between clustered stereocilia minimize friction in the ear on a subnanometre scale. Nature 474, 376379.Google Scholar
Kozmik, Z. (2005). Pax genes in eye development and evolution. Curr. Opin. Genet. Dev. 15, 430438.Google Scholar
Kozmik, Z. (2008). The role of Pax genes in eye evolution. Brain Res. Bull. 75, 335339.Google Scholar
Kozmikova, I., Smolikova, J., Vlcek, C., and Kozmik, Z. (2011). Conservation and diversification of an ancestral chordate gene regulatory network for dorsoventral patterning. PLoS ONE 6, #2, e14650.Google Scholar
Kram, Y.A., Mantey, S., and Corbo, J.C. (2010). Avian cone photoreceptors tile the retina as five independent, self-organizing mosaics. PLoS ONE 5, #2, e8992.Google Scholar
Krapp, H.G. (2009). Ocelli. Curr. Biol. 19, R435R437.Google Scholar
Kreahling, J.M., and Graveley, B.R. (2005). The iStem, a long-range RNA secondary structure element required for efficient exon inclusion in the Drosophila Dscam pre-mRNA. Mol. Cell. Biol. 25, #23, 1025110260.Google Scholar
Krendel, M., and Mooseker, M.S. (2005). Myosins: tails (and heads) of functional diversity. Physiology 20, 239251.Google Scholar
Krishnan, A., and Schiöth, H.B. (2015). The role of G protein-coupled receptors in the early evolution of neurotransmission and the nervous system. J. Exp. Biol. 218, 562571.Google Scholar
Kronhamn, J., Frei, E., Daube, M., Jiao, R., Shi, Y., Noll, M., and Rasmuson-Lestander, Å. (2002). Headless flies produced by mutations in the paralogous Pax6 genes eyeless and twin of eyeless. Development 129, 10151026.Google Scholar
Krueger, D.D., Tuffy, L.P., Papadopoulos, T., and Brose, N. (2012). The role of neurexins and neuroligins in the formation, maturation, and function of vertebrate synapses. Curr. Opin. Neurobiol. 22, 412422.Google Scholar
Kulakova, M., Bakalenko, N., Novikova, E., Cook, C.E., Eliseeva, E., Steinmetz, P.R.H., Kostyuchenko, R.P., Dondua, A., Arendt, D., Akam, M., and Andreeva, T. (2007). Hox gene expression in larval development of the polychaetes Nereis virens and Platynereis dumerilii (Annelida, Lophotrochozoa). Dev. Genes Evol. 217, 3954.Google Scholar
Kumar, A., and Shivashankar, G.V. (2012). Mechanical force alters morphogenetic movements and segmental gene expression patterns during Drosophila embryogenesis. PLoS ONE 7, #3, e33089.Google Scholar
Kumar, J. (2006). The molecular circuitry governing retinal determination. Biochim. Biophys. Acta 1789, 306314.Google Scholar
Kumar, J.P. (2001). Signalling pathways in Drosophila and vertebrate retinal development. Nat. Rev. Genet. 2, 846857.Google Scholar
Kumar, J.P. (2012). Building an ommatidium one cell at a time. Dev. Dyn. 241, 136149.Google Scholar
Kumar, J.P., and Ready, D.F. (1995). Rhodopsin plays an essential structural role in Drosophila photoreceptor development. Development 121, 43594370.Google Scholar
Kunst, M., Hughes, M.E., Raccuglia, D., Felix, M., Li, M., Barnett, G., Duah, J., and Nitabach, M.N. (2014). Calcitonin gene-related peptide neurons mediate sleep-specific circadian output in Drosophila. Curr. Biol. 24, 26522664.Google Scholar
Kuo, D.-H., and Weisblat, D.A. (2011). A new molecular logic for BMP-mediated dorsoventral patterning in the leech Helobdella. Curr. Biol. 21, 12821288.Google Scholar
Kuranaga, E., Matsunuma, T., Kanuka, H., Takemoto, K., Koto, A., Kimura, K.-i., and Miura, M. (2011). Apoptosis controls the speed of looping morphogenesis in Drosophila male terminalia. Development 138, 14931499.Google Scholar
Kuroda, J., Nakamura, M., Yoshida, M., Yamamoto, H., Maeda, T., Taniguchi, K., Nakazawa, N., Hatori, R., Ishio, A., Ozaki, A., Shimaoka, S., Ito, T., Iida, H., Okumura, T., Maeda, R., and Matsuno, K. (2012). Canonical Wnt signaling in the visceral muscle is required for left-right asymmetric development of the Drosophila midgut. Mech. Dev. 128, 625639.Google Scholar
Kurokawa, M., and Kornbluth, S. (2009). Caspases and kinases in a death grip. Cell 138, 838854.Google Scholar
Kyrchanova, O., Mogila, V., Wolle, D., Magbanua, J.P., White, R., Georgiev, P., and Schedl, P. (2015). The boundary paradox in the Bithorax complex. Mech. Dev. 138, 122132.Google Scholar
Labhart, T., and Nilsson, D.-E. (1995). The dorsal eye of the dragonfly Sympetrum: specializations for prey detection against the blue sky. J. Comp. Physiol. A 176, 437453.Google Scholar
Lacalli, T. (2003). Body plans and simple brains. Nature 424, 263264.Google Scholar
Lacalli, T. (2004). Light on ancient photoreceptors. Nature 432, 454455.Google Scholar
Lacalli, T. (2008). Head organization and the head/trunk relationship in protochordates: problems and prospects. Integr. Comp. Biol. 48, 620629.Google Scholar
Lacalli, T. (2010). The emergence of the chordate body plan: some puzzles and problems. Acta Zool. (Stockholm) 91, 410.Google Scholar
Lacalli, T. (2014). Echinoderm conundrums: Hox genes, heterochrony, and an excess of mouths. EvoDevo 5, Article 46.Google Scholar
Lacalli, T.C. (2004). Sensory systems in amphioxus: a window on the ancestral chordate condition. Brain Behav. Evol. 64, 148162.Google Scholar
Lacalli, T.C. (2008). Mucus secretion and transport in amphioxus larvae: organization and ultrastructure of the food trapping system, and implications for head evolution. Acta Zool. (Stockholm) 89, 219230.Google Scholar
Lacin, H., Zhu, Y., Wilson, B.A., and Skeath, J.B. (2014). Transcription factor expression uniquely identifies most postembryonic neuronal lineages in the Drosophila thoracic central nervous system. Development 141, 10111021.Google Scholar
Lacoste, A.M.B., Schoppik, D., Robson, D.N., Haesemeyer, M., Portugues, R., Li, J.M., Randlett, O., Wee, C.L., Engert, F., and Schier, A.F. (2015). A convergent and essential interneuron pathway for Mauthner-cell-mediated escapes. Curr. Biol. 25, 15261534.Google Scholar
Lai, E.C. (2004). Notch signaling: control of cell communication and cell fate. Development 131, 965973.Google Scholar
Lai, E.C., and Orgogozo, V. (2004). A hidden program in Drosophila peripheral neurogenesis revealed: fundamental principles underlying sensory organ diversity. Dev. Biol. 269, 117.Google Scholar
Lall, S., Ludwig, M.Z., and Patel, N.H. (2003). Nanos plays a conserved role in axial patterning outside of the Diptera. Curr. Biol. 13, 224229.Google Scholar
Lall, S., and Patel, N.H. (2001). Conservation and divergence in molecular mechanisms of axis formation. Annu. Rev. Genet. 35, 407437.Google Scholar
Lamb, T.D. (2011). Evolution of the eye. Sci. Am. 305, #1, 6469.Google Scholar
Lamb, T.D. (2013). Evolution of phototransduction, vertebrate photoreceptors and retina. Prog. Retin. Eye Res. 36, 52119.Google Scholar
Lambrechts, D., and Carmeliet, P. (2004). Sculpting heart valves with NFATc and VEGF. Cell 118, 532534.Google Scholar
LaMendola, N.P., and Bever, T.G. (1997). Peripheral and cerebral asymmetries in the rat. Science 278, 483486.Google Scholar
Land, M.F. (2013). Animal vision: rats watch the sky. Curr. Biol. 23, R611R613.Google Scholar
Land, M.F. (2014). Animal vision: starfish can see at last. Curr. Biol. 24, R200R201.Google Scholar
Land, M.F., and Nilsson, D.-E. (2012). Animal Eyes, 2nd edn. Oxford University Press, New York, NY.Google Scholar
Lander, A.D. (2007). Morpheus unbound: reimagining the morphogen gradient. Cell 128, 245256.Google Scholar
Landgraf, M., and Thor, S. (2006). Development of Drosophila motoneurons: specification and morphology. Semin. Cell Dev. Biol. 17, 311.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
Langen, M., Agi, E., Altschuler, D.J., Wu, L.F., Altschuler, S.J., and Hiesinger, P.R. (2015). The developmental rules of neural superposition in Drosophila. Cell 162, 120133.Google Scholar
Langen, M., Koch, M., Yan, J., De Geest, N., Erfurth, M.-L., Pfeiffer, B.D., Schmucker, D., Moreau, Y., and Hassan, B.A. (2013). Mutual inhibition among postmitotic neurons regulates robustness of brain wiring in Drosophila. eLife 2, e00337.Google Scholar
Lanot, R., Zachary, D., Holder, F., and Meister, M. (2001). Postembryonic hematopoiesis in Drosophila. Dev. Biol. 230, 247257.Google Scholar
Lapraz, F., Besnardeau, L., and Lepage, T. (2009). Patterning of the dorsal-ventral axis in echinoderms: insights into the evolution of the BMP-Chordin signaling network. PLoS Biol. 7, #11, e1000248.Google Scholar
Larroux, C., Fahey, B., Degnan, S.M., Adamski, M., Rokhsar, D.S., and Degnan, B.M. (2007). The NK homeobox gene cluster predates the origin of Hox genes. Curr. Biol. 17, 706710.Google Scholar
Larsen, C., Bardet, P.-L., Vincent, J.-P., and Alexandre, C. (2008). Specification and positioning of parasegment grooves in Drosophila. Dev. Biol. 321, 310318.Google Scholar
Larsson, M.C., Domingos, A.I., Jones, W.D., Chiappe, M.E., Amrein, H., and Vosshall, L.B. (2004). Or83b encodes a broadly expressed odorant receptor essential for Drosophila olfaction. Neuron 43, 703714.Google Scholar
Larsson, M.L. (2015). Binocular vision, the optic chiasm, and their associations with vertebrate motor behavior. Front. Ecol. Evol. 3, Article 89.Google Scholar
Laubichler, M.D., and Maienschein, J. (eds.) (2007). From Embryology to Evo-Devo: A History of Developmental Evolution. MIT Press, Cambridge, MA.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
Lauri, A., Brunet, T., Handberg-Thorsager, M., Fischer, A.H.L., Simakov, O., Steinmetz, P.R.H., Tomer, R., Keller, P.J., and Arendt, D. (2014). Development of the annelid axochord: insights into notochord evolution. Science 345, 13651368.Google Scholar
Lavagnino, N.J., Arya, G.H., Korovaichuk, A., and Fanara, J.J. (2013). Genetic architecture of olfactory behavior in Drosophila melanogaster: differences and similarities across development. Behav. Genet. 43, 348359.Google Scholar
Lavagnino, N.J., and Fanara, J.J. (2016). Changes across development influence visible and cryptic natural variation of Drosophila melanogaster olfactory response. Evol. Biol. 43, 96108.Google Scholar
Lawrence, P.A., and Casal, J. (2013). The mechanisms of planar cell polarity, growth and the Hippo pathway: some known unknowns. Dev. Biol. 377, 18. [See alsoGoogle ScholarGoogle Scholar
Layton, W.M. Jr. (1976). Random determination of a developmental process: reversal of normal visceral asymmetry in the mouse. J. Hered. 67, 336338.Google Scholar
Le Douarin, N.M., and Dieterlen-Liévre, F. (2013). How studies on the avian embryo have opened new avenues in the understanding of development: a view about the neural and hematopoietic systems. Dev. Growth Differ. 55, 114.Google Scholar
Leal, W.S. (2013). Odorant reception in insects: roles of receptors, binding proteins, and degrading enzymes. Annu. Rev. Entomol. 58, 373391.Google Scholar
Lebestky, T., Chang, T., Hartenstein, V., and Banerjee, U. (2000). Specification of Drosophila hematopoietic lineage by conserved transcription factors. Science 288, 146149.Google Scholar
LeBon, L., Lee, T.V., Sprinzak, D., Jafar-Nejad, H., and Elowitz, M.B. (2014). Fringe proteins modulate Notch-ligand cis and trans interactions to specify signaling states. eLife 3, e02950.Google Scholar
Leclère, L., and Rentzsch, F. (2014). RGM regulates BMP-mediated secondary axis formation in the sea anemone Nematostella vectensis. Cell Rep. 9, 19211930.Google Scholar
Lee, C., Kim, N., Roy, M., and Graveley, B.R. (2010). Massive expansions of Dscam splicing diversity via staggered homologous recombination during arthropod evolution. RNA 16, 91105.Google Scholar
Lee, H.-G., Kim, Y.-C., Dunning, J.S., and Han, K.-A. (2008). Recurring ethanol exposure induces disinhibited courtship in Drosophila. PLoS ONE, #1, e1391.Google Scholar
Lee, M.S.Y., Jago, J.B., García-Bellido, D.C., Edgecombe, G.D., Gehling, J.G., and Paterson, J.R. (2011). Modern optics in exceptionally preserved eyes of early Cambrian arthropods from Australia. Nature 474, 631634.Google Scholar
Lee, Y., and Rio, D.C. (2015). Mechanisms and regulation of alternative pre-mRNA splicing. Annu. Rev. Biochem. 84, 291323.Google Scholar
Lefebvre, J.L., Kostadinov, D., Chen, W.V., Maniatis, T., and Sanes, J.R. (2012). Protocadherins mediate dendritic self-avoidance in the mammalian nervous system. Nature 488, 517521.Google Scholar
Lefebvre, J.L., Sanes, J.R., and Kay, J.N. (2015). Development of dendritic form and function. Annu. Rev. Cell Dev. Biol. 31, 741777.Google Scholar
Leinwand, S.G., and Chalasani, S.H. (2011). Olfactory networks: from sensation to perception. Curr. Opin. Genet. Dev. 21, 806811.Google Scholar
Leite-Castro, J., Beviano, V., Rodrigues, P.N., and Freitas, R. (2016). HoxA genes and the fin-to-limb transition in vertebrates. J. Dev. Biol. 4, Article 4010010. [See alsoGoogle ScholarGoogle Scholar
Lelli, A., Asai, Y., Forge, A., Holt, J.R., and Géléoc, G.S.G. (2009). Tonotopic gradient in the developmental acquisition of sensory transduction in outer hair cells of the mouse cochlea. J. Neurophysiol. 101, 29612973.Google Scholar
Lem, J., Krasnoperova, N.V., Calvert, P.D., Kosaras, B., Cameron, D.A., Nicolò, M., Makino, C.L., and Sidman, R.L. (1999). Morphological, physiological, and biochemical changes in rhodopsin knockout mice. PNAS 96, 736741.Google Scholar
LeMasurier, M., and Gillespie, P.G. (2005). Hair-cell mechanotransduction and cochlear amplification. Neuron 48, 403415.Google Scholar
Lemke, S., and Schmidt-Ott, U. (2009). Evidence for a composite anterior determinant in the hover fly Episyrphus balteatus (Syrphidae), a cyclorrhaphan fly with an anterodorsal serosa anlage. Development 136, 117127.Google Scholar
Lemke, S., Stauber, M., Shaw, P.J., Rafiqi, A.M., Prell, A., and Schmidt-Ott, U. (2008). bicoid occurrence and Bicoid-dependent hunchback regulation in lower cyclorrhaphan flies. Evol. Dev. 10, 413420.Google Scholar
Lemon, R.N. (2008). Descending pathways in motor control. Annu. Rev. Neurosci. 31, 195218.Google Scholar
Lemons, D., Fritzenwanker, J.H., Gerhart, J., Lowe, C.J., and McGinnis, W. (2010). Co-option of an anteroposterior head axis patterning system for proximodistal patterning of appendages in early bilaterian evolution. Dev. Biol. 344, 358362.Google Scholar
Lemons, D., and McGinnis, W. (2006). Genomic evolution of Hox gene clusters. Science 313, 19181922.Google Scholar
Lengyel, J.A., and Iwaki, D.D. (2002). It takes guts: the Drosophila hindgut as a model system for organogenesis. Dev. Biol. 243, 119.Google Scholar
Leondaritis, G., and Eickholt, B.J. (2015). Short lives with long-lasting effects: filopodia protrusions in neuronal branching morphogenesis. PLoS Biol. 13, #9, e1002241.Google Scholar
Leroi, A.M. (2003). Mutants: On Genetic Variety and the Human Body. Viking Press, New York, NY.Google Scholar
Lesser, M.P., Carleton, K.L., Böttger, S.A., Barry, T.M., and Walker, C.W. (2011). Sea urchin tube feet are photosenory organs that express a rhabdomeric-like opsin and PAX6. Proc. R. Soc. Lond. B 278, 33713379.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
Letzkus, P., Ribi, W.A., Wood, J.T., Zhu, H., Zhang, S.-W., and Srinivasan, M.V. (2006). Lateralization of olfaction in the honeybee Apis mellifera. Curr. Biol. 16, 14711476.Google Scholar
Levin, M. (2005). Left-right asymmetry in embryonic development: a comprehensive review. Mech. Dev. 122, 325.Google Scholar
Levin, M., Johnson, R.L., Stern, C.D., Kuehn, M., and Tabin, C. (1995). A molecular pathway determining left-right asymmetry in chick embryogenesis. Cell 82, 803814.Google Scholar
Levin, M., and Mercola, M. (1998). The compulsion of chirality: toward an understanding of left-right asymmetry. Genes Dev. 12, 763769.Google Scholar
Levin, M., and Palmer, A.R. (2007). Left-right patterning from the inside out: widespread evidence for intracellular control. BioEssays 29, 271287.Google Scholar
Levin, M., Roberts, D.J., Holmes, L.B., and Tabin, C. (1996). Laterality defects in conjoined twins. Nature 384, 321.Google Scholar
Levin, M., Thorlin, T., Robinson, K.R., Nogi, T., and Mercola, M. (2002). Asymmetries in H+/K+-ATPase and cell membrane potentials comprise a very early step in left-right patterning. Cell 111, 7789.Google Scholar
Levine, J.S., and MacNichol, E.F. Jr. (1982). Color vision in fishes. Sci. Am. 246, #2, 140149.Google Scholar
Levine, M. (2010). Transcriptional enhancers in animal development and evolution. Curr. Biol. 20, R754R763.Google Scholar
Levine, M., Cattoglio, C., and Tijan, R. (2014). Looping back to leap forward: transcription enters a new era. Cell 157, 1325.Google Scholar
Levine, S.S., Weiss, A., Erdjument-Bromage, H., Shao, Z., Tempst, P., and Kingston, R.E. (2002). The core of the Polycomb repressive complex is compositionally and functionally conserved in flies and humans. Mol. Cell. Biol. 22, #17, 60706078.Google Scholar
Levy, D.L., and Heald, R. (2012). Mechanisms of intracellular scaling. Annu. Rev. Cell Dev. Biol. 28, 113135.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., 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, M.A., and Steel, K.P. (2012). A cornucopia of candidates for deafness. Cell 150, 879881.Google Scholar
Lewitzky, M., Simister, P.C., and Feller, S.M. (2012). Beyond “furballs” and “dumpling soups”: towards a molecular architecture of signaling complexes and networks. FEBS Lett. 586, 27402750.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, L., and Ginty, D.D. (2014). The structure and organization of lanceolate mechanosensory complexes at mouse hair follicles. eLife 3, e01901.Google Scholar
Li, L., Rutlin, M., Abraira, V.E., Cassidy, C., Kus, L., Gong, S., Jankowski, M.P., Luo, W., Heintz, N., Koerber, R., Woodbury, C.J., and Ginty, D.D. (2011). The functional organization of cutaneous low-threshold mechanosensory neurons. Cell 147, 16151627.Google Scholar
Li, Q., Barish, S., Okuwa, S., Maciejewski, A., Brandt, A.T., Reinhold, D., Jones, C.D., and Volkan, P.C. (2016). A functionally conserved gene regulatory network module governing olfactory neuron diversity. PLoS Genet. 12, #1, e1005780.Google Scholar
Li, Q., and Liberles, S.D. (2015). Aversion and attraction through olfaction. Curr. Biol. 25, R120R129.Google Scholar
Li, S., Sukeena, J.M., Simmons, A.B., Hansen, E.J., Nuhn, R.E., Samuels, I.S., and Fuerst, P.G. (2015). DSCAM promotes refinement in the mouse retina through cell death and restriction of exploring dendrites. J. Neurosci. 35, #14, 56405654.Google Scholar
Liang, H.-L., Xu, M., Chuang, Y.-C., and Rushlow, C. (2012). Response to the BMP gradient requires highly combinatorial inputs from multiple patterning systems in the Drosophila embryo. Development 139, 19561964.Google Scholar
Liang, X., Holy, T.E., and Taghert, P.H. (2016). Synchronous Drosophila circadian pacemakers display nonsynchronous Ca2+ rhythms in vivo. Science 351, 976981.Google Scholar
Liang, Z., and Biggin, M.D. (1998). Eve and ftz regulate a wide array of genes in blastoderm embryos: the selector homeoproteins directly or indirectly regulate most genes in Drosophila. Development 125, 44714482.Google Scholar
Liberles, S.D. (2014). Mammalian pheromones. Annu. Rev. Physiol. 76, 151175.Google Scholar
Liberles, S.D., and Buck, L.B. (2006). A second class of chemosensory receptors in the olfactory epithelium. Nature 442, 645650.Google Scholar
Liberman, M.C., Gao, J., He, D.Z.Z., Wu, X., Jia, S., and Zuo, J. (2002). Prestin is required for electromotility of the outer hair cell and for the cochlear amplifier. Nature 419, 300304.Google Scholar
Lichtneckert, R., and Reichert, H. (2005). Insights into the urbilaterian brain: conserved genetic patterning mechanisms in insect and vertebrate brain development. Heredity 94, 465477.Google Scholar
Lienkamp, S., Ganner, A., and Walz, G. (2012). Inversin, Wnt signaling and primary cilia. Differentiation 83, S49S55.Google Scholar
Lilly, B. (2014). We have contact: endothelial cell–smooth muscle cell interactions. Physiology 29, 234241.Google Scholar
Lillywhite, P.G. (1980). The insect's compound eye. Trends Neurosci. 3, 169173.Google Scholar
Lin, C.H., and Rankin, C.H. (2012). Alcohol addiction: chronic ethanol leads to cognitive dependence in Drosophila. Curr. Biol. 22, R1043R1044.Google Scholar
Lin, C.Y., Chuang, C.C., Hua, T.E., Chen, C.C., Dickson, B.J., Greenspan, R.J., and Chiang, A.S. (2013). A comprehensive wiring diagram of the protocerebral bridge for visual information processing in the Drosophila brain. Cell Rep. 3, 17391753.Google Scholar
Lin, G., Chen, Y., and Slack, J.M.W. (2013). Imparting regenerative capacity to limbs by progenitor cell transplantation. Dev. Cell 24, 4151.Google Scholar
Lin, L., Rao, Y., and Isacson, O. (2005). Netrin-1 and slit-2 regulate and direct neurite growth of ventral midbrain dopaminergic neurons. Mol. Cell. Neurosci. 28, 547555.Google Scholar
Lin, R., Rittenhouse, D., Sweeney, K., Potluri, P., and Wallace, D.C. (2015). TSPO, a mitochondrial outer membrane protein, controls ethanol-related behaviors in Drosophila. PLoS Genet. 11, #8, e1005366.Google Scholar
Lin, S.-Y., and Burdine, R.D. (2005). Brain asymmetry: switching from left to right. Curr. Biol. 15, R343R345.Google Scholar
Lin, Y.Y., and Gubb, D. (2009). Molecular dissection of Drosophila Prickle isoforms distinguishes their essential and overlapping roles in planar cell polarity. Dev. Biol. 325, 386399.Google Scholar
Lindeman, R.E., Gearhart, M.D., Minkina, A., Krentz, A.D., Bardwell, V.J., and Zarkower, D. (2015). Sexual cell-fate reprogramming in the ovary by DMRT1. Curr. Biol. 25, 764771.Google Scholar
Lindemann, B. (2001). Receptors and transduction in taste. Nature 413, 219225.Google Scholar
Lindsey, D.T., Brown, A.M., Brainard, D.H., and Apicella, C.L. (2015). Hunter-gatherer color naming provides new insight into the evolution of color terms. Curr. Biol. 25, 24412446.Google Scholar
Lindsley, D.L., and Zimm, G.G. (1992). The Genome of Drosophila melanogaster. Academic Press, New York, NY.Google Scholar
Ling, F., Dahanukar, A., Weiss, L.A., Kwon, J.Y., and Carlson, J.R. (2014). The molecular and cellular basis of taste coding in the legs of Drosophila. J. Neurosci. 34, 71487164.Google Scholar
Linnaeus, C. (1758–59). Systema naturae per regna tria naturae: secundum classes, ordines, genera, species, cum characteribus, differentiis, synonymis, locis., 10th edn. Laurentius Salvius, Stockholm.Google Scholar
Linneweber, G.A., Winking, M., and Fischbach, K.-F. (2015). The cell adhesion molecules Roughest, Hibris, Kin of Irre and Sticks and Stones are required for long range spacing of the Drosophila wing disc sensory sensilla. PLoS ONE 10, #6, e0128490.Google Scholar
Liu, G., Li, W., Wang, L., Kar, A., Guan, K.-L., Rao, Y., and Wu, J.Y. (2009). DSCAM functions as a netrin receptor in commissural axon pathfinding. PNAS 106, #8, 29512956.Google Scholar
Liu, P.Z., and Kaufman, T.C. (2005). Short and long germ segmentation: unanswered questions in the evolution of a developmental mode. Evol. Dev. 7, 629646. [See alsoGoogle ScholarGoogle Scholar
Liu, Q.-X., Hiramoto, M., Ueda, H., Gojobori, T., Hiromi, Y., and Hirose, S. (2009). Midline governs axon pathfinding by coordinating expression of two major guidance systems. Genes Dev. 23, 11651170.Google Scholar
Liu, Z., Li, G.-H., Huang, J.-F., Murphy, R.W., and Shi, P. (2012). Hearing aid for vertebrates via multiple episodic adaptive events on prestin genes. Mol. Biol. Evol. 29, #9, 21872198.Google Scholar
Liu, Z., Qi, F.-Y., Zhou, X., Ren, H.-Q., and Shi, P. (2014). Parallel sites implicate functional convergence of the hearing gene Prestin among echolocating mammals. Mol. Biol. Evol. 31, #9, 24152424.Google Scholar
Liu, Z., Yang, C.-P., Sugino, K., Fu, C.-C., Liu, L.-Y., Yao, X., Lee, L.P., and Lee, T. (2015). Opposing intrinsic temporal gradients guide neural stem cell production of varied neuronal fates. Science 350, 317320.Google Scholar
Livesey, F.J., and Cepko, C.L. (2001). Vertebrate neural cell-fate determination: lessons from the retina. Nat. Rev. Neurosci. 2, 109118.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
Lodato, S., and Arlotta, P. (2015). Generating neuronal diversity in the mammalian cerebral cortex. Annu. Rev. Cell Dev. Biol. 31, 699720.Google Scholar
Logan, M.A., and Vetter, M.L. (2004). Do-it-yourself tiling: dendritic growth in the absence of homotypic contacts. Neuron 43, 439440.Google Scholar
Lohnes, D. (2003). The Cdx1 homeodomain protein: an integrator of posterior signaling in the mouse. BioEssays 25, 971980.Google Scholar
Löhr, U., and Pick, L. (2005). Cofactor-interaction motifs and the cooption of a homeotic Hox protein into the segmentation pathway of Drosophila melanogaster. Curr. Biol. 15, 643649.Google Scholar
Long, H., Sabatier, C., Ma, L., Plump, A., Yuan, W., Ornitz, D.M., Tamada, A., Murakami, F., Goodman, C.S., and Tessier-Lavigne, M. (2004). Conserved roles for Slit and Robo proteins in midline commissural axon guidance. Neuron 42, 213223.Google Scholar
Longley, R.L. Jr., and Ready, D.F. (1995). Integrins and the development of three-dimensional structure in the Drosophila compound eye. Dev. Biol. 171, 415433.Google Scholar
Longobardi, L., Li, T., Tagliafierro, L., Temple, J.D., Willcockson, H.H., Ye, P., Esposito, A., Xu, F., and Spagnoli, A. (2015). Synovial joints: from development to homeostasis. Curr. Osteoporos. Rep. 13, 4151.Google Scholar
Looby, P., and Loudon, A.S.I. (2005). Gene duplication and complex circadian clocks in mammals. Trends Genet. 12, 4653.Google Scholar
Lopes, C.S., and Casares, F. (2010). hth maintains the pool of eye progenitors and its downregulation by Dpp and Hh couples retinal fate acquisition with cell cycle exit. Dev. Biol. 339, 7888.Google Scholar
Lopez-Rios, J. (2016). The many lives of SHH in limb development and evolution. Semin. Cell Dev. Biol. 49, 116124.Google Scholar
Lorimer, T., Gomez, F., and Stoop, R. (2015). Mammalian cochlea as a physics guided evolution-optimized hearing sensor. Sci. Rep. 5, Article 12492.Google Scholar
Losos, J.B. (2011). Convergence, adaptation, and constraint. Evolution 65, 18271840.Google Scholar
Loudon, A.S.I. (2012). Circadian biology: a 2.5 billion year old clock. Curr. Biol. 22, R570R571.Google Scholar
Lovato, T.L., and Cripps, R.M. (2016). Regulatory networks that direct the development of specialized cell types in the Drosophila heart. J. Cardiovasc. Dev. Dis. 3, Article 18.Google Scholar
Lovato, T.L., Sensibaugh, C.A., Swingle, K.L., Martinez, M.M., and Cripps, R.M. (2015). The Drosophila transcription factors Tinman and Pannier activate and collaborate with Myocyte Enhancer Factor-2 to promote heart cell fate. PLoS ONE 10, #7, e0132965.Google Scholar
Lovejoy, N.R. (2000). Reinterpreting recapitulation: systematics of needlefishes and their allies (Teleostei: Beloniformes). Evolution 54, 13491362.Google Scholar
Lowe, C.J., Clarke, D.N., Medeiros, D.M., Rokhsar, D.S., and Gerhart, J. (2015). The deuterostome context of chordate origins. Nature 520, 456465.Google Scholar
Lowe, C.J., Terasaki, M., Wu, M., Freeman, R.M. Jr., Runft, L., Kwan, K., Haigo, S., Aronowicz, J., Lander, E., Gruber, C., Smith, M., Kirschner, M., and Gerhart, J. (2006). Dorsoventral patterning in hemichordates: insights into early chordate evolution. PLoS Biol. 4, #9, e1291.Google Scholar
Lowe, C.J., Wu, M., Salic, A., Evans, L., Lander, E., Stange-Thomann, N., Gruber, C.E., Gerhart, J., and Kirschner, M. (2003). Anteroposterior patterning in hemichordates and the origins of the chordate nervous system. Cell 113, 853865.Google Scholar
Lowe, L.A., Supp, D.M., Sampath, K., Yokoyama, T., Wright, C.V.E., Potter, S.S., Overbeek, P., and Kuehn, M.R. (1996). Conserved left-right asymmetry of nodal expression and alterations in murine situs inversus. Nature 381, 158161.Google Scholar
Lowery, L.A., and Sive, H. (2004). Strategies of vertebrate neurulation and a re-evaluation of teleost neural tube formation. Mech. Dev. 121, 11891197.Google Scholar
Lowery, L.A., and Sive, H. (2009). Totally tubular: the mystery behind function and origin of the brain ventricular system. BioEssays 31, 446458.Google Scholar
Lu, C.-H., Rincón-Limas, D.E., and Botas, J. (2000). Conserved overlapping and reciprocal expression of msh/Msx1 and apterous/Lhx2 in Drosophila and mice. Mech. Dev. 99, 177181.Google Scholar
Lu, Q., Senthilan, P.R., Effertz, T., Nadrowski, B., and Göpfert, M.C. (2009). Using Drosophila for studying fundamental processes in hearing. Integr. Comp. Biol. 49, 674680.Google Scholar
Lu, X., and Sipe, C.W. (2016). Developmental regulation of planar cell polarity and hair-bundle morphogenesis in auditory hair cells: lessons from human and mouse genetics. WIREs Dev. Biol. 5, 85101.Google Scholar
Lucas, R.J. (2013). Mammalian inner retinal photoreception. Curr. Biol. 23, R125R133.Google Scholar
Luke, G.N., Castro, L.F.C., McLay, K., Bird, C., Coulson, A., and Holland, P.W.H. (2003). Dispersal of NK homeobox gene clusters in amphioxus and humans. PNAS 100, #9, 52925295.Google Scholar
Lumpkin, E.A., Marshall, K.L., and Nelson, A.M. (2010). The cell biology of touch. J. Cell Biol. 191, 237248.Google Scholar
Lundberg, Y.W., Xu, Y., Thiessen, K.D., and Kramer, K.L. (2015). Mechanisms of otoconia and otolith development. Dev. Dyn. 244, 239253.Google Scholar
Luo, D.-G., and Yau, K.-W. (2008). How vision begins: an odyssey. PNAS 105, #29, 98559862.Google Scholar
Luo, D.-G., Yue, W.W.S., Ala-Laurila, P., and Yau, K.-W. (2011). Activation of visual pigments by light and heat. Science 332, 13071312.Google Scholar
Lynch, J., and Desplan, C. (2003). Evolution of development: beyond Bicoid. Curr. Biol. 13, R557R559.Google Scholar
Lynch, J.A., Brent, A.E., Leaf, D.S., Pultz, M.A., and Desplan, C. (2006). Localized maternal orthodenticle patterns anterior and posterior in the long germ wasp Nasonia. Nature 439, 728732.Google Scholar
Lynch, J.A., and Roth, S. (2011). The evolution of dorsal-ventral patterning mechanisms in insects. Genes Dev. 25, 107118.Google Scholar
Lynch, M. (2007). The Origins of Genome Architecture. Sinauer, Sunderland, MA.Google Scholar
Lyons, D.B., Allen, W.E., Goh, T., Tsai, L., Barnea, G., and Lomvardas, S. (2013). An epigenetic trap stabilizes singular olfactory receptor expression. Cell 154, 325336.Google Scholar
Ma, J.-x., Znoiko, S., Othersen, K.L., Ryan, J.C., Das, J., Isayama, T., Kono, M., Oprian, D.D., Corson, D.W., Cornwall, M.C., Cameron, D.A., Harosi, F.I., Makino, C.L., and Crouch, R.K. (2001). A visual pigment expressed in both rod and cone photoreceptors. Neuron 32, 451461.Google Scholar
Ma, X., Edgecombe, G.D., Hou, X., Goral, T., and Strausfeld, N.J. (2015). Preservation pathways of corresponding brains of a Cambrian euarthropod. Curr. Biol. 25, 29692975.Google Scholar
Macdonald, P.M., and Struhl, G. (1986). A molecular gradient in early Drosophila embryos and its role in specifying the body pattern. Nature 324, 537545.Google Scholar
Mackay, T.F.C., and Anholt, R.R.H. (2006). Of flies and man: Drosophila as a model for human complex traits. Annu. Rev. Genomics Hum. Genet. 7, 339367.Google Scholar
Mackin, K.A., Roy, R.A., and Theobald, D.L. (2014). An empirical test of convergent evolution in rhodopsins. Mol. Biol. Evol. 31, 8595.Google Scholar
MacLean, J.A. II, Lorenzetti, D., Hu, Z., Salerno, W.J., Miller, J., and Wilkinson, M.F. (2006). Rhox homeobox gene cluster: recent duplication of three family members. Genesis 44, 122129.Google Scholar
MacNeilage, P.F., Rogers, L.J., and Vallortigara, G. (2009). Origins of the left and right brain. Sci. Am. 301, #1, 6067.Google Scholar
Maderspacher, F. (2016). Snail chirality: the unwinding. Curr. Biol. 26, R215R217.Google Scholar
Maeso, I., Irimia, M., Tena, J.J., Casares, F., and Gómez-Skarmeta, J.L. (2013). Deep conservation of cis-regulatory elements in metazoans. Philos. Trans. R. Soc. Lond. B 368, 20130020.Google Scholar
Maeso, I., Irimia, M., Tena, J.J., González-Pérez, E., Tran, D., Ravi, V., Venkatesh, B., Campuzano, S., Gómez-Skarmeta, J.L., and Garcia-Fernàndez, J. (2012). An ancient genomic regulatory block conserved across bilaterians and its dismantling in tetrapods by retrogene replacement. Genome Res. 22, 642655.Google Scholar
Magariños, M., Contreras, J., Aburto, M.R., and Varela-Nieto, I. (2012). Early development of the vertebrate inner ear. Anat. Rec. 295, 17751790.Google Scholar
Maier, E.C., Saxena, A., Alsina, B., Bronner, M.E., and Whitfield, T.T. (2014). Sensational placodes: neurogenesis in the otic and olfactory systems. Dev. Biol. 389, 5067.Google Scholar
Maier, J.X., Blankenship, M.L., Li, J.X., and Katz, D.B. (2015). A multisensory network for olfactory processing. Curr. Biol. 25, 26422650.Google Scholar
Mainguy, G., In der Rieden, P.M.J., Berezikov, E., Woltering, J.M., Plasterk, R.H.A., and Durston, A.J. (2003). A position-dependent organisation of retinoid response elements is conserved in the vertebrate Hox clusters. Trends Genet. 19, 476479.Google Scholar
Maizel, A., Bensaude, O., Prochiantz, A., and Joliot, A. (1999). A short region of its homeodomain is necessary for Engrailed nuclear export and secretion. Development 126, 31833190.Google Scholar
Makino, C.L., Riley, C.K., Looney, J., Crouch, R.K., and Okada, T. (2010). Binding of more than one retinoid to visual opsins. Biophys. J. 99, 23662373.Google Scholar
Maklad, A., Kamel, S., Wong, E., and Fritzsch, B. (2010). Development and organization of polarity-specific segregation of primary vestibular afferent fibers in mice. Cell Tissue Res. 340, 303321.Google Scholar
Maklad, A., Reed, C., Johnson, N.S., and Fritzsch, B. (2014). Anatomy of the lamprey ear: morphological evidence for occurrence of horizontal semicircular ducts in the labyrinth of Petromyzon marinus. J. Anat. 224, 432446.Google Scholar
Malicki, J. (2004). Cell fate decisions and patterning in the vertebrate retina: the importance of timing, asymmetry, polarity and waves. Curr. Opin. Neurobiol. 14, 1521.Google Scholar
Malik, S. (2014). Polydactyly: phenotypes, genetics and classification. Clin. Genet. 85, 203212.Google Scholar
Mallo, M., and Alonso, C.R. (2013). The regulation of Hox gene expression during animal development. Development 140, 39513963.Google Scholar
Mallo, M., Wellik, D.M., and Deschamps, J. (2010). Hox genes and regional patterning of the vertebrate body plan. Dev. Biol. 344, 715.Google Scholar
Malnic, B., Hirono, J., Sato, T., and Buck, L.B. (1999). Combinatorial receptor codes for odors. Cell 96, 713723.Google Scholar
Mandal, L., Banerjee, U., and Hartenstein, V. (2004). Evidence for a fruit fly hemangioblast and similarities between lymph-gland hematopoiesis in fruit fly and mammal aorta-gonadal-mesonephros mesoderm. Nat. Genet. 36, 10191023.Google Scholar
Manfroid, I., Caubit, X., Kerridge, S., and Fasano, L. (2003). Three putative murine Teashirt orthologs specify trunk structures in Drosophila in the same way as the Drosophila teashirt gene. Development 131, 10651073.Google Scholar
Manjón, C., Sánchez-Herrero, E., and Suzanne, M. (2007). Sharp boundaries of Dpp signalling trigger local cell death required for Drosophila leg morphogenesis. Nature Cell Biol. 9, 5763.Google Scholar
Mann, R.S., and Abu-Shaar, M. (1996). Nuclear import of the homeodomain protein Extradenticle in response to Wg and Dpp signalling. Nature 383, 630633.Google Scholar
Mann, R.S., and Carroll, S.B. (2002). Molecular mechanisms of selector gene function and evolution. Curr. Opin. Genet. Dev. 12, 592600.Google Scholar
Männer, J., Wessel, A., and Yelbuz, T.M. (2010). How does the tubular embryonic heart work? Looking for the physical mechanism generating unidirectional blood flow in the valveless embryonic heart tube. Dev. Dyn. 239, 10351046.Google Scholar
Mannschreck, A., and von Angerer, E. (2011). The scent of roses and beyond: molecular structures, analysis, and practical applications of odorants. J. Chem. Educ. 88, 15011506.Google Scholar
Mano, H., and Fukada, Y. (2007). A median third eye: pineal gland retraces evolution of vertebrate photoreceptive organs. Photochem. Photobiol. 83, 1118.Google Scholar
Manor, U., Disanza, A., Grati, M.H., Andrade, L., Lin, H., Di Fiore, P.P., Scita, G., and Kachar, B. (2011). Regulation of stereocilia length by Myosin XVa and whirlin depends on the actin-regulatory protein Eps8. Curr. Biol. 21, 167172.Google Scholar
Manoussaki, D., Chadwick, R.S., Ketten, D.R., Arruda, J., Dimitriadis, E.K., and O'Malley, J.T. (2008). The influence of cochlear shape on low-frequency hearing. PNAS 105, #16, 61626166.Google Scholar
Manoussaki, D., Dimitriadis, E.K., and Chadwick, R.S. (2006). Cochlea's graded curvature effect on low frequency waves. Phys. Rev. Lett. 96, Article 088701.Google Scholar
Mansfield, J.H., and McGlinn, E. (2012). Evolution, expression, and developmental function of Hox-embedded miRNAs. Curr. Top. Dev. Biol. 99, 3157.Google Scholar
Manuel, M. (2009). Early evolution of symmetry and polarity in metazoan body plans. C. R. Biol. 332, 184209.Google Scholar
Marcotti, W. (2012). Functional assembly of mammalian cochlear hair cells. Exp. Physiol. 97, 438451.Google Scholar
Marcus, G., Marblestone, A., and Dean, T. (2014). The atoms of neural computation. Science 346, 551552.Google Scholar
Marcus, G.F. (2006). Cognitive architecture and descent with modification. Cognition 101, 443465.Google Scholar
Maricich, S.M., and Zoghbi, H.Y. (2006). Getting back to basics. Cell 126, 1115.Google Scholar
Marie, B., and Blagburn, J.M. (2003). Differential roles of Engrailed paralogs in determining sensory axon guidance and synaptic target recognition. J. Neurosci. 23, #21, 78547862.Google Scholar
Marlétaz, F., Holland, L.Z., Laudet, V., and Schubert, M. (2006). Retinoic acid signaling and the evolution of chordates. Int. J. Biol. Sci. 2, 3847.Google Scholar
Marlow, H., and Arendt, D. (2014). Ctenophore genomes and the origin of neurons. Curr. Biol. 24, R757R761.Google Scholar
Marlow, H., Tosches, M.A., Tomer, R., Steinmetz, P.R., Lauri, A., Larsson, T., and Arendt, D. (2014). Larval body patterning and apical organs are conserved in animal evolution. BMC Biol. 12, Article 7.Google Scholar
Marlow, H.Q., Srivastava, M., Matus, D.Q., Rokhsar, D., and Martindale, M.Q. (2009). Anatomy and development of the nervous system of Nematostella vectensis, an anthozoan cnidarian. Dev. Neurobiol. 69, 235254.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
Marquardt, T., Ashery-Padan, R., Andrejewski, N., Scardigli, R., Guillemot, F., and Gruss, P. (2001). Pax6 is required for the multipotent state of retinal progenitor cells. Cell 105, 4355.Google Scholar
Marques-Souza, H., Aranda, M., and Tautz, D. (2008). Delimiting the conserved features of hunchback function for the trunk organization of insects. Development 135, 881888.Google Scholar
Marshall, C.R., Raff, E.C., and Raff, R.A. (1994). Dollo's law and the death and resurrection of genes. PNAS 91, 1228312287.Google Scholar
Marshall, J., and Arikawa, K. (2014). Unconventional colour vision. Curr. Biol. 24, R1150R1154.Google Scholar
Marshall, K.L., and Lumpkin, E.A. (2012). The molecular basis of mechanosensory transduction. In López-Larrea, C. (ed.), Sensing in Nature. Landes Bioscience, Austin, TX, pp. 142155.Google Scholar
Marshall, W.F. (2008). Basal bodies: platforms for building cilia. Curr. Top. Dev. Biol. 85, 122.Google Scholar
Marshall, W.F. (2015). How cells measure length on subcellular scales. Trends Cell Biol. 25, 760768.Google Scholar
Marshall, W.F., and Nonaka, S. (2006). Cilia: tuning in to the cell's antenna. Curr. Biol. 16, R604R614.Google Scholar
Martin, A., and Orgogozo, V. (2013). The loci of repeated evolution: a catalog of genetic hotspots of phenotypic variation. Evolution 67, 12351250.Google Scholar
Martin, A., Serano, J.M., Jarvis, E., Bruce, H.S., Wang, J., Ray, S., Barker, C.A., O'Connell, L.C., and Patel, N.H. (2016). CRISPR/Cas9 mutagenesis reveals versatile roles of Hox genes in crustacean limb specification and evolution. Curr. Biol. 26, 1426.Google Scholar
Martin, B.L., and Kimelman, D. (2009). Wnt signaling and the evolution of embryonic posterior development. Curr. Biol. 19, R215R219.Google Scholar
Martin, J.H. (2005). The corticospinal system: from development to motor control. Neuroscientist 11, 161173.Google Scholar
Martin, J.P., Guo, P., Mu, L., Harley, C.M., and Ritzmann, R.E. (2015). Central-complex control of movement in the freely walking cockroach. Curr. Biol. 25, 27952803.Google Scholar
Martin, J.P., and Hildebrand, J.G. (2010). Innate recognition of pheromone and food odors in moths: a common mechanism in the antennal lobe? Front. Behav. Neurosci. 4, Article 159.Google Scholar
Martindale, M.Q., and Hejnol, A. (2009). A developmental perspective: changes in the position of the blastopore during bilaterian evolution. Dev. Cell 17, 162174.Google Scholar
Martinez, S.R., Gay, M.S., and Zhang, L. (2015). Epigenetic mechanisms in heart development and disease. Drug Discov. Today 20, 799811.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 edn. Benjamin Cummings, San Francisco, CA.Google Scholar
Mashanov, V.S., Zueva, O.R., Heinzeller, T., Aschauer, B., and Dolmatov, I.Y. (2007). Developmental origin of the adult nervous system in a holothurian: an attempt to unravel the enigma of neurogenesis in echinoderms. Evol. Dev. 9, 244256.Google Scholar
Mason, A.C., Oshinsky, M.L., and Hoy, R.R. (2011). Hyperacute directional hearing in a microscale auditory system. Nature 410, 686690.Google Scholar
Masse, N.Y., Turner, G.C., and Jefferis, G.S.X.E. (2009). Olfactory information processing in Drosophila. Curr. Biol. 19, R700R713.Google Scholar
Matis, M., Russler-Germain, D.A., Hu, Q., Tomlin, C.J., and Axelrod, J. (2014). Microtubules provide directional information for core PCP function. eLife 3, e02893.Google Scholar
Matsui, A., Go, Y., and Niimura, Y. (2010). Degeneration of olfactory receptor gene repertoires in primates: no direct link to full trichromatic vision. Mol. Biol. Evol. 27, 11921200.Google Scholar
Matsunami, M., Sumiyama, K., and Saitou, N. (2010). Evolution of conserved non-coding sequences within the vertebrate Hox clusters through the two-round whole genome duplications revealed by phylogenetic footprinting analysis. J. Mol. Evol. 71, 427436.Google Scholar
Matsuo, E., and Kamikouchi, A. (2013). Neuronal encoding of sound, gravity, and wind in the fruit fly. J. Comp. Physiol. A 199, 253262.Google Scholar
Matthews, B.J., Kim, M.E., Flanagan, J.J., Hattori, D., Clemens, J.C., Zipursky, S.L., and Grueber, W.B. (2007). Dendrite self-avoidance is controlled by Dscam. Cell 129, 593604.Google Scholar
Maung, S.M.T.W., and Jenny, A. (2011). Planar cell polarity in Drosophila. Organogenesis 7, 115.Google Scholar
Maurange, C., and Gould, A.P. (2005). Brainy but not too brainy: starting and stopping neuroblast divisions in Drosophila. Trends Neurosci. 28, 3036.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
Mayer, G. (2006). Structure and development of onychophoran eyes: what is the ancestral visual organ in arthropods? Arthropod Struct. Dev. 35, 231245.Google Scholar
Mayer, G., Kato, C., Quast, B., Chisholm, R.H., Landman, K.A., and Quinn, L.M. (2010). Growth patterns in Onychophora (velvet worms): lack of a localised posterior proliferation zone. BMC Evol. Biol. 10, Article 339.Google Scholar
Mayer, G., and Whitington, P.M. (2009). Neural development in Onychophora (velvet worms) suggests a step-wise evolution of segmentation in the nervous system of Panarthropoda. Dev. Biol. 335, 263275.Google Scholar
Maynard Smith, J. (1968). The counting problem. In Waddington, C.H. (ed.), Towards a Theoretical Biology. I. Prolegomena. Aldine, Chicago, IL, pp. 120124.Google Scholar
Mazzoni, E.O., Celik, A., Wernet, M.F., Vasiliauskas, D., Johnston, R.J., Cook, T.A., Pichaud, F., and Desplan, C. (2008). Iroquois Complex genes induce co-expression of rhodopsins in Drosophila. PLoS Biol. 6, #4, e97.Google Scholar
McAlpine, J.F. (1981). Morphology and terminology: adults. In McAlpine, J.F., Peterson, B.V., Shewell, G.E., Teskey, H.J., Vockeroth, J.R., and Wood, D.M. (eds.), Manual of Nearctic Diptera. Agriculture Canada, Ottawa, pp. 963.Google Scholar
McCabe, K.L., Gunther, E.C., and Reh, T.A. (1999). The development of the pattern of retinal ganglion cells in the chick retina: mechanisms that control differentiation. Development 126, 57135724.Google Scholar
McDevitt, D.S., Brahma, S.K., Jeanny, J.-C., and Hicks, D. (1993). Presence and foveal enrichment of rod opsin in the “all cone” retina of the American chameleon. Anat. Rec. 237, 299307.Google Scholar
McDougall, C., Korchagina, N., Tobin, J.L., and Ferrier, D.E.K. (2011). Annelid Distal-less/Dlx duplications reveal varied post-duplication fates. BMC Evol. Biol. 11, Article 241.Google Scholar
McGary, K.L., Park, T.J., Woods, J.O., Cha, H.J., Wallingford, J.B., and Marcotte, E.M. (2010). Systematic discovery of nonobvious human disease models through orthologous phenotypes. PNAS 107, #14, 65446549.Google Scholar
McGhee, G.R. Jr. (2011). Convergent Evolution: Limited Forms Most Beautiful. MIT Press, Cambridge, MA.Google Scholar
McGinnis, W. (1994). A century of homeosis, a decade of homeoboxes. Genetics 137, 607611.Google Scholar
McGinnis, W., and Krumlauf, R. (1992). Homeobox genes and axial patterning. Cell 68, 283302.Google Scholar
McGinnis, W., and Kuziora, M. (1994). The molecular architects of body design. Sci. Am. 270, #2, 5866.Google Scholar
McGinnis, W., Levine, M.S., Hafen, E., Kuroiwa, A., and Gehring, W.J. (1984). A conserved DNA sequence in homoeotic genes of the Drosophila Antennapedia and bithorax complexes. Nature 308, 428433.Google Scholar
McGregor, A.P. (2005). How to get ahead: the origin, evolution and function of bicoid. BioEssays 27, 904913.Google Scholar
McGregor, A.P. (2006). Wasps, beetles and the beginning of the ends. BioEssays 28, 683686.Google Scholar
McGurk, L., Berson, A., and Bonini, N.M. (2015). Drosophila as an in vivo model for human neurodegenerative disease. Genetics 201, 377402.Google Scholar
McHenry, M.J. (2005). The morphology, behavior, and biomechanics of swimming in ascidian larvae. Can. J. Zool. 83, 6274.Google Scholar
McInerney, J.O., and O'Connell, M.J. (2014). Ghost locus appears. Nature 514, 570571.Google Scholar
McKinney, R.M., Vernier, C., and Ben-Shahar, Y. (2015). The neural basis for insect pheromonal communication. Curr. Opin. Insect Sci. 12, 8692.Google Scholar
McManus, C. (2005). Reversed bodies, reversed brains, and (some) reversed behaviors: of zebrafish and men. Dev. Cell 8, 796797.Google Scholar
McManus, I.C., Martin, N., Stubbings, G.F., Chung, E.M.K., and Mitchison, H.M. (2004). Handedness and situs inversus in primary ciliary dyskinesia. Proc. R. Soc. Lond. B 271, 25792582.Google Scholar
McMillen, P., and Holley, S.A. (2015). The tissue mechanics of vertebrate body elongation and segmentation. Curr. Opin. Genet. Dev. 32, 106111.Google Scholar
Meadows, R. (2015). Odors help fruit flies escape parasitoid wasps. PLoS Biol. 13, #12, e1002317.Google Scholar
Medioni, C., Sénatore, S., Salmand, P.-A., Lalevée, N., Perrin, L., and Sémériva, M. (2009). The fabulous destiny of the Drosophila heart. Curr. Opin. Genet. Dev. 19, 518525.Google Scholar
Meganathan, K., Sotiriadou, I., Natarajan, K., Hescheler, J., and Sachinidis, A. (2015). Signaling molecules, transcription growth factors and other regulators revealed from in-vivo and in-vitro models for the regulation of cardiac development. Int. J. Cardiol. 183, 117128.Google Scholar
Mehta, T.K., Ravi, V., Yamasaki, S., Lee, A.P., Lian, M.M., Tay, B.-H., Tohari, S., Yanai, S., Tay, A., Brenner, S., and Venkatesh, B. (2013). Evidence for at least six Hox clusters in the Japanese lamprey (Lethenteron japonicum). PNAS 110, 1604416049.Google Scholar
Meier, T., Chabaud, F., and Reichert, H. (1991). Homologous patterns in the embryonic development of the peripheral nervous system in the grasshopper Schistocerca gregaria and the fly Drosophila melanogaster. Development 112, 241253.Google Scholar
Meijers, R., Puettmann-Holgado, R., Skiniotis, G., Liu, J.-h., Walz, T., Wang, J.-h., and Schmucker, D. (2007). Structural basis of Dscam isoform specificity. Nature 449, 487491.Google Scholar
Meinhardt, H. (2015). Dorsoventral patterning by the Chordin-BMP pathway: a unified model from a pattern-formation perspective for Drosophila, vertebrates, sea urchins and Nematostella. Dev. Biol. 405, 137148.Google Scholar
Meister, M. (2015). On the dimensionality of odor space. eLife 4, e07865.Google Scholar
Melnattur, K., and Shaw, P.J. (2015). Learning and memory: do bees dream? Curr. Biol. 25, R1040R1041.Google Scholar
Menda, G., Shamble, P.S., Nitzany, E.I., Golden, J.R., and Hoy, R.R. (2014). Visual perception in the brain of a jumping spider. Curr. Biol. 24, 25802585.Google Scholar
Menzel, R., Greggers, U., Smith, A., Berger, S., Brandt, R., Brunke, S., Bundrock, G., Hülse, S., Plümpe, T., Schaupp, F., Schüttler, E., Stach, S., Stindt, J., Stollhoff, N., and Watzl, S. (2005). Honey bees navigate according to a map-like spatial memory. PNAS 102, #8, 30403045.Google Scholar
Merabet, S., and Mann, R.S. (2016). To be specific or not: the critical relationship between Hox and TALE proteins. Trends Genet. 32, 334347.Google Scholar
Meraldi, P. (2016). Centrosomes in spindle organization and chromosome segregation: a mechanistic view. Chromosome Res. 24, 1934.Google Scholar
Mercader, N., Leonardo, E., Azpiazu, N., Serrano, A., Morata, G., Martínez-A, C., and Torres, M. (1999). Conserved regulation of proximodistal limb axis development by Meis1/Hth. Nature 402, 425429.Google Scholar
Meredith, R.W., Gatesy, J., Emerling, C.A., York, V.M., and Springer, M.S. (2013). Rod monochromacy and the coevolution of cetacean retinal opsins. PLoS Genet. 9, #4, e1003432.Google Scholar
Merkel, M., Sagner, A., Gruber, F.S., Etournay, R., Blasse, C., Myers, E., Eaton, S., and Jülicher, F. (2014). The balance of Prickle/Spiny-legs isoforms controls the amount of coupling between core and Fat PCP systems. Curr. Biol. 24, 21112123.Google Scholar
Merriam, L.C., and Chess, A. (2007). cis-regulatory elements within the odorant receptor coding region. Cell 131, 844846.Google Scholar
Meunier, H., Vauclair, J., and Fagard, J. (2012). Human infants and baboons show the same pattern of handedness for a communicative gesture. PLoS ONE 7, #3, e33959.Google Scholar
Mhatre, N. (2015). Active amplification in insect ears: mechanics, models and molecules. J. Comp. Physiol. A 201, 1937.Google Scholar
Mieko, C., and Bier, E. (2008). EvoD/Vo: the origins of BMP signalling in the neuroectoderm. Nat. Rev. Genet. 9, 663677.Google Scholar
Mihrshahi, R. (2006). The corpus callosum as an evolutionary innovation. J. Exp. Zool. B. Mol. Dev. Evol. 306, 817.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
Mikeladze-Dvali, T., Wernet, M.F., Pistillo, D., Mazzoni, E.O., Teleman, A.A., Chen, Y.-W., Cohen, S., and Desplan, C. (2005). The growth regulators warts/lats and melted interact in a bistable loop to specify opposite fates in Drosophila R8 photoreceptors. Cell 122, 775787.Google Scholar
Milán, M., Weihe, U., Tiong, S., Bender, W., and Cohen, S.M. (2001). msh specifies dorsal cell fate in the Drosophila wing. Development 128, 32633268.Google Scholar
Milev, N.B., and Reddy, A.B. (2015). Circadian redox oscillations and metabolism. Trends Endocr. Metab. 26, 430437.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
Millard, S.S., Lu, Z., Zipursky, S.L., and Meinertzhagen, I.A. (2010). Drosophila Dscam proteins regulate postsynaptic specificity at multiple-contact synapses. Neuron 67, 761768.Google Scholar
Miller, A. (1941). Position of adult testes in Drosophila melanogaster Meigen. PNAS 27, 3541.Google Scholar
Miller, C.J., and Davidson, L.A. (2013). The interplay between cell signalling and mechanics in developmental processes. Nat. Rev. Genet. 14, 733744.Google Scholar
Miller, G. (2009). On the origin of the nervous system. Science 325, 2426.Google Scholar
Minelli, A. (2003). The origin and evolution of appendages. Int. J. Dev. Biol. 47, 573581.Google Scholar
Minelli, A. (2005). A morphologist's perspective on terminal growth and segmentation. Evol. Dev. 7, 568573.Google Scholar
Minelli, A. (2015). Grand challenges in evolutionary developmental biology. Front. Ecol. Evol. 2, Article 85. [See alsoGoogle ScholarGoogle Scholar
Minelli, A., and Fusco, G. (2004). Evo-devo perspectives on segmentation: model organisms, and beyond. Trends Ecol. Evol. 19, 423429.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
Mishra, M., Oke, A., Lebel, C., McDonald, E.C., Plummer, Z., Cook, T.A., and Zelhof, A.C. (2010). Pph13 and Orthodenticle define a dual regulatory pathway for photoreceptor cell morphogenesis and function. Development 137, 28952904.Google Scholar
Miskolczi-McCallum, C.M., Scavetta, R.J., Svendsen, P.C., Soanes, K.H., and Brook, W.J. (2005). The Drosophila melanogaster T-box genes midline and H15 are conserved regulators of heart development. Dev. Biol. 278, 459472.Google Scholar
Missbach, C., Dweck, H.K.M., Vogel, H., Vilcinskas, A., Stensmyr, M., Hansson, B.S., and Gross-Wilde, E. (2014). Evolution of insect olfactory receptors. eLife 3, e02115.Google Scholar
Missler, M., and Südhof, T.C. (1998). Neurexins: three genes and 1001 products. Trends Genet. 14, 2026.Google Scholar
Mitchell, K.J., Doyle, J.L., Serafini, T., Kennedy, T.E., Tessier-Lavigne, M., Goodman, C.S., and Dickson, B. (1996). Genetic analysis of Netrin genes in Drosophila: netrins guide CNS commissural axons and peripheral motor axons. Neuron 17, 203215.Google Scholar
Mitchell, M.E., Sander, T.L., Klinkner, D.B., and Tomita-Mitchell, A. (2007). The molecular basis of congenital heart disease. Semin. Thorac. Cardiovasc. Surg. 19, 228237.Google Scholar
Mito, T., Nakamura, T., and Noji, S. (2010). Evolution of insect development: to the hemimetabolous paradigm. Curr. Opin. Genet. Dev. 20, 355361.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
Mittmann, B., and Scholtz, G. (2001). Distal-less expression in embryos of Limulus polyphemus (Chelicerata, Xiphosura) and Lepisma saccharina (Insecta, Zygentoma) suggests a role in the development of mechanoreceptors, chemoreceptors, and the CNS. Dev. Genes Evol. 211, 232243.Google Scholar
Miura, S.K., Martins, A., Zhang, K.X., Graveley, B.R., and Zipursky, S.L. (2013). Probabilistic splicing of Dscam1 establishes identity at the level of single neurons. Cell 155, 11661177.Google Scholar
Miyazono, S., Isayama, T., Delori, F.C., and Makino, C.L. (2011). Vitamin A activates rhodopsin and sensitizes it to ultraviolet light. Vis. Neurosci. 28, 485497.Google Scholar
Mizunami, M. (1994). Processing of contrast signals in the insect ocellar system. Zool. Sci. 11, 175190.Google Scholar
Mizutani, C.M., and Bier, E. (2008). EvoD/Vo: the origins of BMP signalling in the neuroectoderm. Nat. Rev. Genet. 9, 663677.Google Scholar
Mizutani, C.M., Meyer, N., Roelink, H., and Bier, E. (2006). Threshold-dependent BMP-mediated repression: a model for a conserved mechanism that patterns the neuroectoderm. PLoS Biol. 4, #10, e1313.Google Scholar
Moczek, A.P., Sears, K.E., Stollewerk, A., Wittkopp, P.J., Diggle, P., Dworkin, I., Ledon-Rettig, C., Matus, D.Q., Roth, S., Abouheif, E., Brown, F.D., Chiu, C.-H., Cohen, S., De Tomaso, A.W., Gilbert, S.F., Hall, B., Love, A.C., Lyons, D.C., Sanger, T.J., Smith, J., Specht, C., Vallejo-Marin, M., and Extavour, C.G. (2015). The significance and scope of evolutionary developmental biology: a vision for the 21st century. Evol. Dev. 17, 198219.Google Scholar
Mogilner, A., and Fogelson, B. (2015). Cytoskeletal chirality: swirling cells tell left from right. Curr. Biol. 25, R501R503.Google Scholar
Möglich, A., Yang, X., Ayers, R.A., and Moffat, K. (2010). Structure and function of plant photoreceptors. Annu. Rev. Plant Biol. 61, 2147.Google Scholar
Mohammad, F., Aryal, S., Ho, J., Stewart, J.C., Norman, N.A., Tan, T.L., Eisaka, A., and Claridge-Chang, A. (2016). Ancient anxiety pathways influence Drosophila defense behaviors. Curr. Biol. 26, 981986.Google Scholar
Molina, M.D., de Crozé, N., Haillot, E., and Lepage, T. (2013). Nodal: master and commander of the dorsal-ventral and left-right axes in the sea urchin embryo. Curr. Opin. Genet. Dev. 23, 445453.Google Scholar
Molina, M.D., Neto, A., Maeso, I., Gómez-Skarmeta, J.L., Saló, E., and Cebrià, F. (2011). Noggin and noggin-like genes control dorsoventral axis regeneration in planarians. Curr. Biol. 21, 300305.Google Scholar
Molina, M.D., Saló, E., and Cebrià, F. (2007). The BMP pathway is essential for re-specification and maintenance of the dorsoventral axis in regenerating and intact planarians. Development 311, 7994.Google Scholar
Mombaerts, P. (2004). Genes and ligands for odorant, vomeronasal and taste receptors. Nat. Rev. Neurosci. 5, 263278.Google Scholar
Monahan, K., and Lomvardas, S. (2015). Monoallelic expression of olfactory receptors. Annu. Rev. Cell Dev. Biol. 31, 721740.Google Scholar
Monahan-Earley, R., Dvorak, A.M., and Aird, W.C. (2013). Evolutionary origins of the blood vascular system and endothelium. J. Thromb. Haemost. 11 (Suppl. 1), 4666.Google Scholar
Monier, B., Tevy, F., Perrin, L., Capovilla, M., and Semeriva, M. (2007). Downstream of Hox genes. Fly 1, 5967.Google Scholar
Montcouquiol, M., Sans, N., Huss, D., Kach, J., Dickman, J.D., Forge, A., Rachel, R.A., Copeland, N.G., Jenkins, N.A., Bogani, D., Murdoch, J., Warchol, M.E., Wenthold, R.J., and Kelley, M.W. (2006). Asymmetric localization of Vangl2 and Fz3 indicate novel mechanisms for planar cell polarity in mammals. J. Neurosci. 26, #19, 52655275.Google Scholar
Montealegre-Z., F., Jonsson, T., Robson-Brown, K.A., Postles, M., and Robert, D. (2012). Convergent evolution between insect and mammalian audition. Science 338, 968971.Google Scholar
Monteiro, A.S., and Ferrier, D.E.K. (2006). Hox genes are not always colinear. Int. J. Biol. Sci. 2, 95103.Google Scholar
Monteiro, A.S., Shierwater, B., Dellaporta, S.L., and Holland, P.W.H. (2006). A low diversity of ANTP class homeobox genes in Placozoa. Evol. Dev. 8, 174182.Google Scholar
Montell, C. (1999). Visual transduction in Drosophila. Annu. Rev. Cell Dev. Biol. 15, 231268.Google Scholar
Montell, C. (2009). A taste of the Drosophila gustatory receptors. Curr. Opin. Neurobiol. 19, 345353.Google Scholar
Mooi, R., and David, B. (2008). Radial symmetry, the anterior/posterior axis, and echinoderm Hox genes. Ann. Rev. Ecol. Evol. Syst. 39, 4362. [See alsoGoogle ScholarGoogle Scholar
Moore, M.S., DeZazzo, J., Luk, A.Y., Tully, T., Singh, C.M., and Heberlein, U. (1998). Ethanol intoxication in Drosophila: genetic and pharmacological evidence for regulation by the cAMP signaling pathway. Cell 93, 9971007.Google Scholar
Moorman, A.F.M., Christoffels, V.M., Anderson, R.H., and van den Hoff, M.J.B. (2007). The heart-forming fields: one or mutiple? Philos. Trans. R. Soc. Lond. B 362, 12571265.Google Scholar
Moran, D.T., Rowley, J.C. III, Jafek, B.W., and Lovell, M.A. (1982). The fine structure of the olfactory mucosa in man. J. Neurocytol. 11, 721746.Google Scholar
Morante, J., and Desplan, C. (2008). The color-vision circuit in the medulla of Drosophila. Curr. Biol. 18, 553565.Google Scholar
Morata, G., Macías, A., Urquía, N., and González-Reyes, A. (1990). Homoeotic genes. Semin. Cell Biol. 1, 219227.Google Scholar
Moreira, I.S. (2014). Structural features of the G-protein/GPCR interactions. Biochim. Biophys. Acta 1840, 1633.Google Scholar
Morell, V. (1991). A hand on the bird: and one on the bush. Science 254, 3334.Google Scholar
Moreno, E., De Mulder, K., Salvenmoser, W., Ladurner, P., and Martínez, P. (2010). Inferring the ancestral function of the posterior Hox gene within the bilateria: controlling the maintenance of reproductive structures, the musculature and the nervous system in the acoel flatworm Isodiametra pulchra. Evol. Dev. 12, 258266.Google Scholar
Moreno, E., Permanyer, J., and Martinez, P. (2011). The origin of patterning systems in Bilateria: insights from the Hox and ParaHox genes in Acoelomorpha. Genomics Proteomics Bioinformatics 9, 6576.Google Scholar
Morey, C., Da Silva, N.R., Perry, P., and Bickmore, W.A. (2007). Nuclear reorganisation and chromatin decondensation are conserved, but distinct, mechanisms linked to Hox gene activation. Development 134, 909919.Google Scholar
Morgan, D., Goodship, J., Essner, J.J., Vogan, K.J., Turnpenny, L., Yost, H.J., Tabin, C.J., and Strachan, T. (2002). The left-right determinant inversin has highly conserved ankyrin repeat and IQ domains and interacts with calmodulin. Hum. Genet. 110, 377384.Google Scholar
Morgan, D., Turnpenny, L., Goodship, J., Dai, W., Majumder, K., Matthews, L., Gardner, A., Schuster, G., Vein, L., Harrison, W., Elder, F.F.B., Penman-Splitt, M., Overbeek, P., and Strachan, T. (1998). Inversin, a novel gene in the vertebrate left-right axis pathway, is partially deleted in the inv mouse. Nat. Genet. 20, 149156.Google Scholar
Morgan, M.J. (1992). On the evolutionary origin of right handedness. Curr. Biol. 2, 1517.Google Scholar
Morgan, R. (2004). Conservation of sequence and function in the Pax6 regulatory elements. Trends Genet. 20, 283287.Google Scholar
Morgan, R. (2006). Engrailed: complexity and economy of a multi-functional transcription factor. FEBS Lett. 580, 25312533.Google Scholar
Mori, K., Nagao, H., and Yoshihara, Y. (1999). The olfactory bulb: coding and processing of odor molecule information. Science 286, 711715.Google Scholar
Mori, T., Mchaourab, H., and Johnson, C.H. (2015). Circadian clocks: unexpected biochemical cogs. Curr. Biol. 25, R827R844.Google Scholar
Morin, L.P. (1994). The circadian visual system. Brain Res. Rev. 67, 102127.Google Scholar
Morokuma, J., Ueno, M., Kawanishi, H., Saiga, H., and Nishida, H. (2002). HrNodal, the ascidian nodal-related gene, is expressed in the left side of the epidermis, and lies upstream of HrPitx. Dev. Genes Evol. 212, 439446.Google Scholar
Moroz, L.L. (2009). On the independent origins of complex brains and neurons. Brain Behav. Evol. 74, 177190.Google Scholar
Moroz, L.L., Kocot, K.M., Citarella, M.R., Dosung, S., Norekian, T.P., Povolotskaya, I.S., Grigorenko, A.P., Dailey, C., Berezikov, E., Buckley, K.M., Ptitsyn, A., Reshetov, D., Mukherjee, K., Moroz, T.P., Bobkova, Y., Yu, F., Kapitonov, V.V., Jurka, J., Bobkov, Y.V., Swore, J.J., Girardo, D.O., Fodor, A., Gusev, F., Sanford, R., Bruders, R., Kittler, E., Mills, C.E., Rast, J.P., Derelle, R., Solovyev, V.V., Kondrashov, F.A., Swalla, B.J., Sweedler, J.V., Rogaev, E.I., Halanych, K.M., and Kohn, A.B. (2014). The ctenophore genome and the evolutionary origins of neural systems. Nature 510, 109114.Google Scholar
Moroz, L.L., and Kohn, A.B. (2016). Independent origins of neurons and synapses: insights from ctenophores. Philos. Trans. R. Soc. Lond. B 371, 20150041.Google Scholar
Morrissey, M.E., Shelton, S., Brockerhoff, S.E., Hurley, J.B., and Kennedy, B.N. (2011). PRE-1, a cis element sufficient to enhance cone- and rod- specific expression in differentiating zebrafish photoreceptors. BMC Dev. Biol. 11, Article 3.Google Scholar
Morshedian, A., and Fain, G.L. (2015). Single-photon sensitivity of lamprey rods with cone-like outer segments. Curr. Biol. 25, 484487.Google Scholar
Mosca, T.J., and Luo, L. (2014). Synaptic organization of the Drosophila antennal lobe and its regulation by the Teneurins. eLife 3, e03726.Google Scholar
Mouchel-Vielh, E., Blin, M., Rigolot, C., and Deutsch, J.S. (2002). Expression of a homologue of the fushi tarazu (ftz) gene in a ciripede crustacean. Evol. Dev. 4, 7685.Google Scholar
Moussian, B., and Uv, A.E. (2005). An ancient control of epithelial barrier formation and wound healing. BioEssays 27, 987990.Google Scholar
Mu, X., Fu, X., Sun, H., Liang, S., Maeda, H., Firishman, L.J., and Klein, W.H. (2005). Ganglion cells are required for normal progenitor-cell proliferation but not cell-fate determination or patterning in the developing mouse retina. Curr. Biol. 15, 525530.Google Scholar
Muchlinski, M.N. (2010). A comparative analysis of vibrissa count and infraorbital foramen area in primates and other mammals. J. Hum. Evol. 58, 447473.Google Scholar
Muerdter, F., and Stark, A. (2014). Hiding in plain sight. Nature 512, 374375.Google Scholar
Mukckli, L., Naumer, M.J., and Singer, W. (2009). Bilateral visual field maps in a patient with only one hemisphere. PNAS 106, #31, 1303413039.Google Scholar
Mukouyama, Y.-s., Shin, D., Britsch, S., Taniguchi, M., and Anderson, D.J. (2002). Sensory nerves determine the pattern of arterial differentiation and blood vessel branching in the skin. Cell 109, 693705.Google Scholar
Müller, G.B. (2003). Homology: the evolution of morphological organization. In Müller, G.B. and Newman, S.A. (eds.), Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology. MIT Press, Cambridge, MA, pp. 5169.Google Scholar
Muller, H.J. (1932). Further studies on the nature and causes of gene mutations. Proc. 6th Int. Congr. Genet. 1, 213255.Google Scholar
Müller, P., Rogers, K.W., Yu, S.R., Brand, M., and Schier, A.F. (2013). Morphogen transport. Development 140, 16211638.Google Scholar
Muneoka, K., Han, M., and Gardiner, D.M. (2008). Regrowing human limbs. Sci. Am. 298, #4, 5663.Google Scholar
Munnamalai, V., and Fekete, D.M. (2013). Wnt signaling during cochlear development. Semin. Cell Dev. Biol. 24, 480489.Google Scholar
Muñoz-Chápuli, R. (2011). Evolution of angiogenesis. Int. J. Dev. Biol. 55, 345351.Google Scholar
Murawala, P., Tanaka, E.M., and Currie, J.D. (2012). Regeneration: the ultimate example of wound healing. Semin. Cell Dev. Biol. 23, 954962.Google Scholar
Murdock, D.J.E., and Donoghue, P.C.J. (2011). Evolutionary origins of animal skeletal biomineralization. Cells Tissues Organs 194, 98102.Google Scholar
Mure, L.S., Cornut, P.-L., Rieux, C., Drouyer, E., Denis, P., Gronfier, C., and Cooper, H.M. (2009). Melanopsin bistability: a fly's eye technology in the human retina. PLoS ONE 4, #6, e5991.Google Scholar
Murray, J.D. (1989). Mathematical Biology. Springer-Verlag, Berlin.Google Scholar
Musiek, E.S., Xiong, D.D., and Holtzman, D.M. (2014). Sleep, circadian rhythms, and the pathogenesis of Alzheimer Disease. Exp. Mol. Med. 47, e148.Google Scholar
Mustafi, D., Engel, A.H., and Palczewski, K. (2009). Structure of cone photoreceptors. Prog. Retin. Eye Res. 28, 289302.Google Scholar
Mustárdy, L., Buttle, K., Steinbach, G., and Garab, G. (2008). The three-dimensional network of thylakoid membranes in plants: quasihelical model of the granum-stroma assembly. Plant Cell 20, 25522557.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
Nadeau, J.H., and Taylor, B.A. (1984). Lengths of chromosomal segments conserved since divergence of man and mouse. PNAS 81, 814818.Google Scholar
Nadrowski, B., Albert, J.T., and Göpfert, M.C. (2008). Transducer-based force generation explains active process in Drosophila hearing. Curr. Biol. 18, 13651372.Google Scholar
Naganathan, S.R., Fürthauer, S., Nishikawa, M., Jülicher, F., and Grill, S.W. (2014). Active torque generation by the actomyosin cell cortex drives left–right symmetry breaking. eLife 3, e04165.Google Scholar
Nakagawa, T., and Vosshall, L.B. (2009). Controversy and consensus: noncanonical signaling mechanisms in the insect olfactory system. Curr. Opin. Neurobiol. 19, 284292.Google Scholar
Nakamura, T., and Hamada, H. (2012). Left-right patterning: conserved and divergent mechanisms. Development 139, 32573262.Google Scholar
Nakanishi, K. (1991). Why 11-cis-retinal? Am. Zool. 31, 479489.Google Scholar
Nakano, S., Stillman, B., and Horvitz, H.R. (2011). Replication-coupled chromatin assembly generates a neuronal bilateral asymmetry in C. elegans. Cell 147, 15251536.Google Scholar
Nam, J., and Nei, M. (2005). Evolutionary change of the numbers of homeobox genes in bilaterian animals. Mol. Biol. Evol. 22, 23862394.Google Scholar
Nam, J.-H., Cotton, J.R., and Grant, W. (2007). A virtual hair cell. I. Addition of gating spring theory into a 3-D bundle mechanical model. Biophys. J. 92, 19181928.Google Scholar
Namigai, E.K.O., Kenny, N.J., and Shimeld, S.M. (2014). Right across the tree of life: the evolution of left-right asymmetry in the Bilateria. Genesis 52, 458470.Google Scholar
Narayanan, C.H., and Hamburger, V. (1971). Motility in chick embryos with substitution of lumbosacral by brachial and brachial by lumbosacral spinal cord segments. J. Exp. Zool. 178, 415432.Google Scholar
Narayanan, P., Chatterton, P., Ikeda, A., Ikeda, S., Corey, D.P., Ervasti, J.M., and Perrin, B.J. (2015). Length regulation of mechanosensitive stereocilia depends on very slow actin dynamics and filament-severing proteins. Nat. Commun. 6, Article 6855.Google Scholar
Narendra, V., Rocha, P.P., An, D., Raviram, R., Skok, J.A., Mazzoni, E.O., and Reinberg, D. (2015). CTCF establishes discrete functional chromatin domains at the Hox clusters during differentiation. Science 347, 10171021.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
Narris, P.M. (2001). In a fly's ear. Nature 410, 644645.Google Scholar
Natale, A., Sims, C., Chiusano, M.L., Amoroso, A., D'Aniello, E., Fucci, L., Krumlauf, R., Branno, M., and Locascio, A. (2011). Evolution of anterior Hox regulatory elements among chordates. BMC Evol. Biol. 11, Article 330.Google Scholar
Nathans, J. (1987). Molecular biology of visual pigments. Annu. Rev. Neurosci. 10, 163194.Google Scholar
Nathans, J. (1999). The evolution and physiology of human color vision: insights from molecular genetic studies of visual pigments. Neuron 24, 299312.Google Scholar
Navis, A., and Bagnat, M. (2015). Developing pressures: fluid forces driving morphogenesis. Curr. Opin. Genet. Dev. 32, 2430.Google Scholar
Nawabi, H., and Castellani, V. (2011). Axonal commissures in the central nervous system: how to cross the midline? Cell. Mol. Life Sci. 68, 25392553.Google Scholar
Nayak, G.D., Ratnayaka, H.S.K., Goodyear, R.J., and Richardson, G.P. (2007). Development of the hair bundle and mechanotransduction. Int. J. Dev. Biol. 51, 597608.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
Negre, B., Ranz, J.M., Casals, F., Cáceres, M., and Ruiz, A. (2003). A new split of the Hox gene complex in Drosophila: relocation and evolution of the gene labial. Mol. Biol. Evol. 20, 20422054.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
Nei, M., Niimura, Y., and Nozawa, M. (2008). The evolution of animal chemosensory receptor gene repertoires: roles of chance and necessity. Nat. Rev. Genet. 9, 951963.Google Scholar
Nelson, B.R., Gumuscu, B., Hartman, B.H., and Reh, T.A. (2006). Notch activity is downregulated just prior to retinal ganglion cell differentiation. Dev. Neurosci. 28, 128141.Google Scholar
Nelson, C. (2004). Selector genes and the genetic control of developmental modules. In Schlosser, G. and Wagner, G.P. (eds.), Modularity in Development and Evolution. University of Chicago Press, Chicago, IL, pp. 1733.Google Scholar
Nemer, M. (2008). Genetic insights into normal and abnormal heart development. Cardiovasc. Pathol. 17, 4854.Google Scholar
Netter, F.H. (2003). Atlas of Human Anatomy, 3rd edn. Icon Learning Systems, Teterboro, NJ.Google Scholar
Neumann, C.J., and Nuesslein-Volhard, C. (2000). Patterning of the zebrafish retina by a wave of Sonic hedgehog activity. Science 289, 21372139.Google Scholar
Neves, J., Demaria, M., Campisi, J., and Jasper, H. (2015). Of flies, mice, and men: evolutionarily conserved tissue damage responses and aging. Dev. Cell 32, 918.Google Scholar
Ng, T.H., Chiang, Y.-A., Yeh, Y.-C., and Wang, H.-C. (2014). Review of Dscam-mediated immunity in shrimp and other arthropods. Dev. Comp. Immunol. 46, 129138.Google Scholar
Nichols, C.D. (2006). Drosophila melanogaster neurobiology, neuropharmacology, and how the fly can inform central nervous system drug discovery. Pharmacol. Ther. 112, 677700.Google Scholar
Nickle, B., and Robinson, P.R. (2007). The opsins of the vertebrate retina: insights from structural, biochemical, and evolutionary studies. Cell. Mol. Life Sci. 64, 29172932.Google Scholar
Nicklen, P. (2007). Arctic ivory: hunting the narwhal. Natl. Geogr. 212, #2, 110129.Google Scholar
Nicol, X., and Gaspar, P. (2014). Routes to cAMP: shaping neuronal connectivity with distinct adenylate cyclases. Eur. J. Neurosci. 39, 17421751.Google Scholar
Nie, J., Mahato, S., Mustill, W., Tipping, C., Bhattacharya, S.S., and Zelhof, A.C. (2012). Cross species analysis of Prominin reveals a conserved cellular role in invertebrate and vertebrate photoreceptor cells. Dev. Biol. 371, 312320.Google Scholar
Niederreither, K., and Dollé, P. (2008). Retinoic acid in development: towards an integrated view. Nat. Rev. Genet. 9, 541553.Google Scholar
Niehrs, C. (2010). On growth and form: a Cartesian coordinate system of Wnt and BMP signaling specifies bilaterian body axes. Development 137, 845857.Google Scholar
Nielsen, C. (2005). Larval and adult brains. Evol. Dev. 7, 483489.Google Scholar
Nielsen, C., and Martinez, P. (2003). Patterns of gene expression: homology or homocracy? Dev. Genes Evol. 213, 149154.Google Scholar
Nilsson, D.-E. (1994). Eyes as optical alarm systems in fan worms and ark clams. Philos. Trans. R. Soc. Lond. B 346, 195212.Google Scholar
Nilsson, D.-E. (1996). Eye ancestry: old genes for new eyes. Curr. Biol. 6, 3942.Google Scholar
Nilsson, D.-E. (2004). Eye evolution: a question of genetic promiscuity. Curr. Opin. Neurobiol. 14, 407414.Google Scholar
Nilsson, D.-E. (2005). Photoreceptor evolution: ancient siblings serve different tasks. Curr. Biol. 15, R94R96.Google Scholar
Nilsson, D.-E. (2009). The evolution of eyes and visually guided behaviour. Philos. Trans. R. Soc. Lond. B 364, 28332847.Google Scholar
Nilsson, D.-E. (2013). Eye evolution and its functional basis. Vis. Neurosci. 30, 520.Google Scholar
Nilsson, D.-E., and Arendt, D. (2008). Eye evolution: the blurry beginning. Curr. Biol. 18, R1096R1098.Google Scholar
Nilsson, D.-E., Gislén, L., Coates, M.M., Skogh, C., and Garm, A. (2005). Advanced optics in a jellyfish eye. Nature 435, 201205.Google Scholar
Niswander, L. (2003). Pattern formation: old models out on a limb. Nat. Rev. Genet. 4, 133143.Google Scholar
Nitta, K.R., Jolma, A., Yin, Y., Morgunova, E., Kivioja, T., Akhtar, J., Hens, K., Toivonen, J., Deplancke, B., Furlong, E.E.M., and Taipale, J. (2015). Conservation of transcription factor binding specificities across 600 million years of bilateria evolution. eLife 4, e04837.Google Scholar
Niwa, N., Akimoto-Kato, A., Niimi, T., Tojo, K., Machida, R., and Hayashi, S. (2010). Evolutionary origin of the insect wing via integration of two developmental modules. Evol. Dev. 12, 168176.Google Scholar
Niwa, N., Hiromi, Y., and Okabe, M. (2004). A conserved developmental program for sensory organ formation in Drosophila melanogaster. Nat. Genet. 36, 293297.Google Scholar
Nödl, M.-T., Fossati, S.M., Domingues, P., Sánchez, F.J., and Zullo, L. (2015). The making of an octopus arm. EvoDevo 6, Article 19.Google Scholar
Nolte, C., and Krumlauf, R. (2007). Expression of Hox genes in the nervous system of vertebrates. In Papageorgiou, S. (ed.), HOX Gene Expression. Landes Bioscience, Austin, TX, pp. 1441.Google Scholar
Nomaksteinsky, M., Kassabov, S., Chettouh, Z., Stoeklé, H.-C., Bonnaud, L., Fortin, G., Kandel, E.R., and Brunet, J.-F. (2013). Ancient origin of somatic and visceral neurons. BMC Biol. 11, Article 53.Google Scholar
Nomaksteinsky, M., Röttinger, E., Dufour, H.D., Chettouh, Z., Lowe, C.J., Martindale, M.Q., and Brunet, J.-F. (2009). Centralization of the deuterostome nervous system predates chordates. Curr. Biol. 19, 12641269.Google Scholar
Nonaka, S., Tanaka, Y., Okada, Y., Takeda, S., Harada, A., Kanai, Y., Kido, M., and Hirokawa, N. (1998). Randomization of left-right asymmetry due to loss of nodal cilia generating leftward flow of extraembryonic fluid in mice lacking KIF3B motor protein. Cell 95, 829837 (Erratum: Cell 1999, 99, 117).Google Scholar
Noonan, J.P. (2009). Regulatory DNAs and the evolution of human development. Curr. Opin. Genet. Dev. 19, 557564.Google Scholar
Noordermeer, D., and Duboule, D. (2013). Chromatin architectures and Hox gene collinearity. Curr. Top. Dev. Biol. 104, 113148.Google Scholar
Noordermeer, D., Leleu, M., Splinter, E., Rougemont, J., De Laat, W., and Duboule, D. (2011). The dynamic architecture of Hox gene clusters. Science 334, 222225.Google Scholar
Nordström, K.J.V., Almén, M.S., Edstam, M.M., Fredriksson, R., and Schiöth, H.B. (2011). Independent HHsearch, Needleman-Wunsch-based, and motif analyses reveal the overall hierarchy for most of the G protein-coupled receptor families. Mol. Biol. Evol. 28, 24712480.Google Scholar
Noro, B., Culi, J., McKay, D.J., Zhang, W., and Mann, R.S. (2006). Distinct functions of homeodomain-containing and homeodomain-less isoforms encoded by homothorax. Genes Dev. 20, 16361650.Google Scholar
Norris, D.P. (2012). Cilia, calcium and the basis of left-right asymmetry. BMC Biol. 10, Article 102.Google Scholar
Norris, D.P., and Jackson, P.K. (2016). Calcium contradictions in cilia. Nature 531, 582583.Google Scholar
Northcutt, R.G. (2012). Evolution of centralized nervous systems: two schools of evolutionary thought. PNAS 109 (Suppl. 1), 1062610633.Google Scholar
Nossa, C.W., Havlak, P., Yue, J.-X., Lv, J., Vincent, K.Y., Brockmann, H.J., and Putnam, N.H. (2014). Joint assembly and genetic mapping of the Atlantic horseshoe crab genome reveals ancient whole genome duplication. GigaScience 3, Article 9.Google Scholar
Novarino, G., Akizu, N., and Gleeson, J.G. (2011). Modeling human disease in humans: the ciliopathies. Cell 147, 7079.Google Scholar
Nübler-Jung, K., and Arendt, D. (1994). Is ventral in insects dorsal in vertebrates? A history of embryological arguments favouring axis inversion in chordate ancestors. Roux's Arch. Dev. Biol. 203, 357366.Google Scholar
Nübler-Jung, K., and Arendt, D. (1996). Enteropneusts and chordate evolution. Curr. Biol. 6, 352353.Google Scholar
Nürnberger, J., Bacallao, R.L., and Phillips, C.L. (2002). Inversin forms a complex with catenins and N-cadherin in polarized epithelial cells. Mol. Biol. Cell 13, 30963106.Google Scholar
Nürnberger, J., Kribben, A., Saez, A.O., Heusch, G., Philipp, T., and Phillips, C.L. (2004). The Invs gene encodes a microtubule-associated protein. J. Am. Soc. Nephrol. 15, 17001710.Google Scholar
Nusse, R. (2001). An ancient cluster of Wnt paralogues. Trends Genet. 17, 443.Google Scholar
Nüsslein-Volhard, C. (1996). Gradients that organize embryo development. Sci. Am. 275, #2, 5461.Google Scholar
Nweeia, M.T., Eichmiller, F.C., Hauschka, P.V., Donahue, G.A., Orr, J.R., Ferguson, S.H., Watt, C.A., Mead, J.G., Potter, C.W., Dietz, R., Giuseppetti, A.A., Black, S.R., Trachtenberg, A.J., and Kuo, W.P. (2014). Sensory ability in the narwhal tooth organ system. Anat. Rec. 297, 599617.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
O'Connell, L.A. (2013). Evolutionary development of neural systems in vertebrates and beyond. J. Neurogenet. 27, 6985.Google Scholar
O'Grady, G.E., and McIver, S.B. (1987). Fine structure of the compound eye of the black fly Simulium vittatum (Diptera: Simuliidae). Can. J. Zool. 65, 14541469.Google Scholar
O'Leary, D.D.M., and Sahara, S. (2008). Genetic regulation of arealization of the neocortex. Curr. Opin. Neurobiol. 18, 90100.Google Scholar
O'Riordan, M.X.D., Bauler, L.D., Scott, F.L., and Duckett, C.S. (2008). Inhibitor of apoptosis proteins in eukaryotic evolution and development: a model of thematic conservation. Dev. Cell 15, 497508.Google Scholar
Oakley, T.H. (2003). The eye as a replicating and diverging, modular developmental unit. Trends Ecol. Evol. 18, 623627.Google Scholar
Ochoa, C., and Rasskin-Gutman, D. (2015). Evo-devo mechanisms underlying the continuum between homology and homoplasy. J. Exp. Zool. B. Mol. Dev. Evol. 324, 91103.Google Scholar
Ochoa-Espinosa, A., Yu, D., Tsirigos, A., Struffi, P., and Small, S. (2009). Anterior-posterior positional information in the absence of a strong Bicoid gradient. PNAS 106, #10, 38233828.Google Scholar
Ochoa-Espinosa, A., Yucel, G., Kaplan, L., Pare, A., Pura, N., Oberstein, A., Papatsenko, D., and Small, S. (2005). The role of binding site cluster strength in Bicoid-dependent patterning in Drosophila. PNAS 102, #14, 49604965.Google Scholar
Odden, J.P., Holbrook, S., and Doe, C.Q. (2002). Drosophila HB9 is expressed in a subset of motoneurons and interneurons, where it regulates gene expression and axon pathfinding. J. Neurosci. 22, #21, 91439149.Google Scholar
Ogura, A., Ikeo, K., and Gojobori, T. (2004). Comparative analysis of gene expression for convergent evolution of camera eye between octopus and human. Genome Res. 14, 15551561.Google Scholar
Ogura, A., Yoshida, M.-a., Moritaki, T., Okuda, Y., Sese, J., Shimizu, K.K., Sousounis, K., and Tsonis, P.A. (2013). Loss of the six3/6 controlling pathways might have resulted in pinhole-eye evolution in Nautilus. Sci. Rep. 3, Article 1432.Google Scholar
Oka, K., Yoshiyama, N., Tojo, K., Machida, R., and Hatakeyama, M. (2010). Characterization of abdominal appendages in the sawfly, Athalia rosae (Hymenoptera), by morphological and gene expression analyses. Dev. Genes Evol. 220, 5359.Google Scholar
Okada, T., Sugihara, M., Bondar, A.-N., Elstner, M., Entel, P., and Buss, V. (2004). The retinal conformation and its environment in rhodopsin in light of a new 2.2 Å crystal structure. J. Mol. Biol. 342, 571583.Google Scholar
Okada, Y., Nonaka, S., Tanaka, Y., Saijoh, Y., Hamada, H., and Hirokawa, N. (1999). Abnormal nodal flow precedes situs inversus in iv and inv mice. Mol. Cell 4, 459468.Google Scholar
Okawa, H., Sampath, A.P., Laughlin, S.B., and Fain, G.L. (2008). ATP consumption by mammalian rod photoreceptors in darkness and light. Curr. Biol. 18, 19171921.Google Scholar
Okkema, P.G., Ha, E., Haun, C., Chen, W., and Fire, A. (1997). The Caenorhabditis elegans NK-2 homeobox gene ceh-22 activates pharyngeal muscle gene expression in combination with pha-1 and is required for normal pharyngeal development. Development 124, 39653973.Google Scholar
Okoruwa, O.E., Weston, M.D., Sanjeevi, D.C., Millemon, A.R., Fritzsch, B., Hallworth, R., and Beisel, K.W. (2008). Evolutionary insights into the unique electromotility motor of mammalian outer hair cells. Evol. Dev. 10, 300315.Google Scholar
Okray, Z., and Hassan, B.A. (2013). Genetic approaches in Drosophila for the study neurodevelopmental disorders. Neuropharmacology 68, 150156.Google Scholar
Okumura, T., Fujiwara, H., Taniguchi, K., Kuroda, J., Nakazawa, N., Nakamura, M., Hatori, R., Ishio, A., Maeda, R., and Matsuno, K. (2010). Left-right asymmetric morphogenesis of the anterior midgut depends on the activation of a non-muscle myosin II in Drosophila. Dev. Biol. 344, 693706.Google Scholar
Okumura, T., Sasamura, T., Inatomi, M., Hozumi, S., Nakamura, M., Hatori, R., Taniguchi, K., Nakazawa, N., Suzuki, E., Maeda, R., Yamakawa, T., and Matsuno, K. (2015). Class I myosins have overlapping and specialized functions in left-right asymmetric development in Drosophila. Genetics 199, 11831199.Google Scholar
Oland, L.A., and Tolbert, L.P. (2011). Roles of glial cells in neural circuit formation: insights from research in insects. Glia 59, 12731295.Google Scholar
Olesnicky, E.C., Brent, A.E., Tonnes, L., Walker, M., Pultz, M.A., Leaf, D., and Desplan, C. (2006). A caudal mRNA gradient controls posterior development in the wasp Nasonia. Development 133, 39733982.Google Scholar
Oliveira, M.B., Liedholm, S.E., Lopez, J.E., Lochte, A.A., Pazio, M., Martin, J.P., Mörch, P.R., Salakka, S., York, J., Yoshimoto, A., and Janssen, R. (2014). Expression of arthropod distal limb-patterning genes in the onychophoran Euperipatoides kanangrensis. Dev. Genes Evol. 224, 8796.Google Scholar
Olivera-Martinez, I., Viallet, J.P., Michon, F., Pearton, D.J., and Dhouailly, D. (2004). The different steps of skin formation in vertebrates. Int. J. Dev. Biol. 48, 107115.Google Scholar
Olson, E.N. (2006). Gene regulatory networks in the evolution and development of the heart. Science 313, 19221927.Google Scholar
Olson, E.N., and Klein, W.H. (1998). Muscle minus MyoD. Dev. Biol. 202, 153156.Google Scholar
Onai, T., Yu, J.-K., Blitz, I.L., Cho, K.W.Y., and Holland, L.Z. (2010). Opposing Nodal/Vg1 and BMP signals mediate axial patterning in embryos of the basal chordate amphioxus. Dev. Biol. 344, 377389.Google Scholar
Onuma, Y., Takahashi, S., Asashima, M., Kurata, S., and Gehring, W.J. (2002). Conservation of Pax 6 function and upstream activation by Notch signaling in eye development of frogs and flies. PNAS 99, #4, 20202025.Google Scholar
Oppenheim, R.W. (1991). Cell death during development of the nervous system. Annu. Rev. Neurosci. 14, 453501.Google Scholar
Oppenheimer, J.M. (1974). Asymmetry revisited. Am. Zool. 14, 867879.Google Scholar
Ordan, E., and Volk, T. (2015). A non-signaling role of Robo2 in tendons is essential for Slit processing and muscle patterning. Development 142, 35123518.Google Scholar
Orii, H., Kato, K., Umesono, Y., Sakurai, T., Agata, K., and Watanabe, K. (1999). The planarian HOM/HOX genes (Plox) expressed along the anteroposterior axis. Dev. Biol. 210, 456468.Google Scholar
Orkin, S.H., and Zon, L.I. (2008). Hematopoiesis: an evolving paradigm for stem cell biology. Cell 132, 631644.Google Scholar
Orly, G., Manor, U., and Gov, N.S. (2015). A biophysical model for the staircase geometry of stereocilia. PLoS ONE 10, #7, e0127926.Google Scholar
Ortega-Hernández, J. (2015). Lobopodians. Curr. Biol. 25, R873R875.Google Scholar
Ortiz, C.O., Etchberger, J.F., Posy, S.L., Frøkjaer-Jensen, C., Lockery, S., Honig, B., and Hobert, O. (2006). Searching for neuronal left/right asymmetry: genomewide analysis of nematode receptor-type guanylyl cyclases. Genetics 173, 131149.Google Scholar
Osborne, P.W., and Ferrier, D.E.K. (2010). Chordate Hox and ParaHox gene clusters differ dramatically in their repetitive element content. Mol. Biol. Evol. 27, 217220.Google Scholar
Osorio, D. (2007). Spam and the evolution of the fly's eye. BioEssays 29, 111115.Google Scholar
Osorio, D., and Bossomaier, T.R.J. (1992). Human cone-pigment spectral sensitivities and the reflectances of natural surfaces. Biol. Cybern. 67, 217222.Google Scholar
Osorio, D., and Vorobyev, M. (2008). A review of the evolution of animal color vision and visual communication signals. Vision Res. 48, 20422051.Google Scholar
Otsuna, H., Shinomiya, K., and Ito, K. (2014). Parallel neural pathways in higher visual centers of the Drosophila brain that mediate wavelength-specific behavior. Front. Neural Circuits 8, Article 8. [See alsoGoogle ScholarGoogle Scholar
Otto, E.A., Schermer, B., Obara, T., O'Toole, J.F., Hiller, K.S., Mueller, A.M., Ruf, R.G., Hoefele, J., Beekmann, F., Landau, D., Foreman, J.W., Goodship, J.A., Strachan, T., Kispert, A., Wolf, M.T., Gagnadoux, M.F., Nivet, H., Antignac, C., Walz, G., Drummond, I.A., Benzing, T., and Hildebrandt, F. (2003). Mutations in INVS encoding inversin cause nephronophthisis type 2, linking renal cystic disease to the function of primary cilia and left-right axis determination. Nat. Genet. 34, 413420.Google Scholar
Ou, G., Stuurman, N., D'Ambrosio, M., and Vale, R.D. (2010). Polarized myosin produces unequal-size daughters during asymmetric cell division. Science 330, 677680.Google Scholar
Ovespian, S.V., and Vesselkin, N.P. (2014). Wiring prior to firing: the evolutionary rise of electrical and chemical modes of synaptic transmission. Rev. Neuroscience 25, 821832.Google Scholar
Özbudak, E.M., and Lewis, J. (2008). Notch signalling synchronizes the zebrafish segmentation clock but is not needed to create somite boundaries. PLoS Genet. 4, #2, e15.Google Scholar
Özüak, O., Buchta, T., Roth, S., and Lynch, J.A. (2014). Dorsoventral polarity of the Nasonia embryo primarily relies on a BMP gradient formed without input from Toll. Curr. Biol. 24, 23932398.Google Scholar
Pace, R.M., Grbic, M., and Nagy, L.M. (2016). Composition and genomic organization of arthropod Hox clusters. EvoDevo 7, Article 11. [See alsoGoogle ScholarGoogle Scholar
Pacifici, M., Koyama, E., and Iwamoto, M. (2005). Mechanisms of synovial joint and articular cartilage formation: recent advances, but many lingering mysteries. Birth Defects Res. C 75, 237248. [See alsoGoogle ScholarGoogle Scholar
Packard, A. (1972). Cephalopods and fish: the limits of convergence. Biol. Rev. 47, 241307.Google Scholar
Page, D.T. (2002). Inductive patterning of the embryonic brain in Drosophila. Development 129, 21212128.Google Scholar
Paixão-Côrtes, V.R., Salzano, F.M., and Bortolini, M.C. (2015). Origins and evolvability of the PAX family. Semin. Cell Dev. Biol. 44, 6474.Google Scholar
Palmer, A.R. (1996). From symmetry to asymmetry: phylogenetic patterns of asymmetry variation in animals and their evolutionary significance. PNAS 93, 1427914286.Google Scholar
Palmer, A.R. (2004). Symmetry breaking and the evolution of development. Science 306, 828833.Google Scholar
Palmer, L.M., Schulz, J.M., Murphy, S.C., Ledergerber, D., Murayama, M., and Larkum, M.E. (2012). The cellular basis of GABAB-mediated interhemispheric inhibition. Science 335, 989993.Google Scholar
Palmer, W.J., and Jiggins, F.M. (2015). Comparative genomics reveals the origins and diversity of arthropod immune systems. Mol. Biol. Evol. 32, #8, 21112129.Google Scholar
Panda, S., Hogenesch, J.B., and Kay, S.A. (2002). Circadian rhythms from flies to humans. Nature 417, 329335.Google Scholar
Panda, S., Nayak, S.K., Campo, B., Walker, J.R., Hogenesch, J.B., and Jegla, T. (2005). Illumination of the melanopsin signaling pathway. Science 307, 600604.Google Scholar
Pandey, U.B., and Nichols, C.D. (2011). Human disease models in Drosophila melanogaster and the role of the fly in therapeutic drug discovery. Pharmacol. Rev. 63, 411436.Google Scholar
Pandur, P. (2005). What does it take to make a heart? Biol. Cell 97, 197210.Google Scholar
Pandur, P., Sirbu, I.O., Kühl, S.J., Philipp, M., and Kühl, M. (2013). Islet1-expressing cardiac progenitor cells: a comparison across species. Dev. Genes Evol. 223, 117129.Google Scholar
Pang, K., and Martindale, M.Q. (2008). Developmental expression of homeobox genes in the ctenophore Mnemiopsis leidyi. Dev. Genes Evol. 218, 307319.Google Scholar
Panganiban, G., Irvine, S.M., Lowe, C., Roehl, H., Corley, L.S., Sherbon, B., Grenier, J.K., Fallon, J.F., Kimble, J., Walker, M., Wray, G.A., Swalla, B.J., Martindale, M.Q., and Carroll, S.B. (1997). The origin and evolution of animal appendages. PNAS 94, 51625166.Google Scholar
Panganiban, G., and Rubenstein, J.L.R. (2002). Developmental functions of the Distal-less/Dlx homeobox genes. Development 129, 43714386.Google Scholar
Pani, A.M., Mullarkey, E.E., Aronowicz, J., Assimacopoulos, S., Grove, E.A., and Lowe, C.J. (2012). Ancient deuterostome origins of vertebrate brain signalling centres. Nature 483, 289294.Google Scholar
Papatsenko, D. (2009). Stripe formation in the early fly embryo: principles, models, and networks. BioEssays 31, 11721180.Google Scholar
Papatsenko, D., Goltsev, Y., and Levine, M. (2009). Organization of developmental enhancers in the Drosophila embryo. Nucleic Acids Res. 37, 56655677.Google Scholar
Papatsenko, D., Nazina, A., and Desplan, C. (2001). A conserved regulatory element present in all Drosophila rhodopsin genes mediates Pax6 functions and participates in the fine-tuning of cell-specific expression. Mech. Dev. 101, 143153.Google Scholar
Papillon, D., Perez, Y., Fasano, L., Le Parco, Y., and Caubit, X. (2005). Restricted expression of a median Hox gene in the central nervous system of chaetognaths. Dev. Genes Evol. 215, 369373.Google Scholar
Paris, M., Escriva, H., Schubert, M., Brunet, F., Brtko, J., Ciesielski, F., Roecklin, D., Vivat-Hannah, V., Jamin, E.L., Cravedi, J.-P., Scanlan, T.S., Renaud, J.-P., Holland, N.D., and Laudet, V. (2008). Amphioxus postembryonic development reveals the homology of chordate metamorphosis. Curr. Biol. 18, 825830.Google Scholar
Park, J.H., Scheerer, P., Hofmann, K.P., and Choe, H.-W. (2008). Crystal structure of the ligand-free G-protein-coupled receptor opsin. Nature 454, 183187.Google Scholar
Park, M.S., Nakagawa, E., Schoenberg, M.R., Benbadis, S.R., and Vale, F.L. (2013). Outcome of corpus callosotomy in adults. Epilepsy Behav. 28, 181184.Google Scholar
Parker, H.J., Bronner, M.E., and Krumlauf, R. (2014). A Hox regulatory network of hindbrain segmentation is conserved to the base of vertebrates. Nature 514, 490493.Google Scholar
Parr, B.A., and McMahon, A.P. (1995). Dorsalizing signal Wnt-7a required for normal polarity of D-V and A-P axes of mouse limb. Nature 374, 350353.Google Scholar
Parr, B.A., Shea, M.J., Vassileva, G., and McMahon, A.P. (1993). Mouse Wnt genes exhibit discrete domains of expression in the early embryonic CNS and limb buds. Development 119, 247261.Google Scholar
Partridge, J.C., and Cuthill, I.C. (2010). Animal behaviour: ultraviolet fish faces. Curr. Biol. 20, R318R320.Google Scholar
Pascual, A., Huang, K.-L., Neveu, J., and Préat, T. (2004). Brain asymmetry and long-term memory. Nature 427, 605606.Google Scholar
Pascual-Anaya, J., Adachi, N., Álvarez, S., Kuratani, S., D'Aniello, S., and Garcia-Fernàndez, J. (2012). Broken colinearity of the amphioxus Hox cluster. EvoDevo 3, Article 28.Google Scholar
Pascual-Anaya, J., Albuixech-Crespo, B., Somorjai, I.M.L., Carmona, R., Oisi, Y., Álvarez, S., Kuratani, S., Muñoz-Chápuli, R., and Garcia-Fernàndez, J. (2013). The evolutionary origins of chordate hematopoesis and vertebrate endothelia. Dev. Biol. 375, 182192.Google Scholar
Pascual-Anaya, J., D'Aniello, S., Kuratani, S., and Garcia-Fernàndez, J. (2013). Evolution of Hox gene clusters in deuterostomes. BMC Dev. Biol. 13, Article 26.Google Scholar
Passamaneck, Y.J., Furchheim, N., Hejnol, A., Martindale, M.Q., and Lüter, C. (2011). Ciliary photoreceptors in the cerebral eyes of a protostome larva. EvoDevo 2, Article 6.Google Scholar
Patel, A.J., Honoré, E., Lesage, F., Fink, M., Romey, G., and Lazdunski, M. (1999). Inhalational anesthetics activate two-pore-domain background K+ channels. Nat. Neurosci. 2, 422426.Google Scholar
Patel, N.H. (1994). Developmental evolution: insights from studies of insect segmentation. Science 266, 581590.Google Scholar
Patel, N.H., Martin-Blanco, E., Coleman, K.G., Poole, S.J., Ellis, M.C., Kornberg, T.B., and Goodman, C.S. (1989). Expression of engrailed proteins in arthropods, annelids, and chordates. Cell 58, 955968.Google Scholar
Patel-Hett, S., and D'Amore, P.A. (2011). Signal transduction in vasculogenesis and developmental angiogenesis. Int. J. Dev. Biol. 55, 353363.Google Scholar
Paterson, J.R., García-Bellido, D.C., Lee, M.S.Y., Brock, G.A., Jago, J.B., and Edgecombe, G.D. (2011). Acute vision in the giant Cambrian predator Anomalocaris and the origin of compound eyes. Nature 480, 237240.Google Scholar
Patraquim, P., Warnefors, M., and Alonso, C.R. (2011). Evolution of Hox post-transcriptional regulation by alternative polyadenylation and microRNA modulation within 12 Drosophila genomes. Mol. Biol. Evol. 28, 24532460.Google Scholar
Patthey, C., Schlosser, G., and Shimeld, S.M. (2014). The evolutionary history of vertebrate cranial placodes. I. Cell type evolution. Dev. Biol. 389, 8297.Google Scholar
Pause, B.M. (2012). Processing of body odor signals by the human brain. Chem. Percept. 5, 5563.Google Scholar
Pavlou, H.J., and Goodwin, S.F. (2013). Courtship behavior in Drosophila melanogaster: towards a “courtship connectome”. Curr. Opin. Neurobiol. 23, 7683.Google Scholar
Pearson, J.C., Lemons, D., and McGinnis, W. (2005). Modulating Hox gene functions during animal body patterning. Nat. Rev. Genet. 6, 893904.Google Scholar
Pechmann, M., McGregor, A.P., Schwager, E.E., Feitosa, N.M., and Damen, W.G.M. (2009). Dynamic gene expression is required for anterior regionalization in a spider. PNAS 106, #5, 14681472.Google Scholar
Pecot, M.Y., Chen, Y., Akin, O., Chen, Z., Tsui, C.Y.K., and Zipursky, S.L. (2014). Sequential axon-derived signals couple target survival and layer specificity in the Drosophila visual system. Neuron 82, 320333.Google Scholar
Peel, A. (2004). The evolution of arthropod segmentation mechanisms. BioEssays 26, 11081116.Google Scholar
Peel, A., and Akam, M. (2003). Evolution of segmentation: rolling back the clock. Curr. Biol. 13, R708R710.Google Scholar
Peel, A.D. (2008). The evolution of developmental gene networks: lessons from comparative studies on holometabolous insects. Philos. Trans. R. Soc. Lond. B 363, 15391547.Google Scholar
Peel, A.D., Chipman, A.D., and Akam, M. (2005). Arthropod segmentation: beyond the Drosophila paradigm. Nat. Rev. Genet. 6, 905916.Google Scholar
Peeters, H., and Devriendt, K. (2006). Human laterality disorders. Eur. J. Med. Genet. 49, 349362.Google Scholar
Peichl, L., Behrmann, G., and Kröger, R.H.H. (2001). For whales and seals the ocean is not blue: a visual pigment loss in marine mammals. Eur. J. Neurosci. 13, 15201528.Google Scholar
Peirson, S.N., Halford, S., and Foster, R.G. (2009). The evolution of irradiance detection: melanopsin and the non-visual opsins. Philos. Trans. R. Soc. Lond. B 364, 28492865.Google Scholar
Pelaz, S., Urquía, N., and Morata, G. (1993). Normal and ectopic domains of the homeotic gene Sex combs reduced of Drosophila. Development 117, 917923.Google Scholar
Pellikka, M., Tanentzapf, G., Pinto, M., Smith, C., McGlade, C.J., Ready, D.F., and Tepass, U. (2002). Crumbs, the Drosophila homologue of human CRB1/RP12, is essential for photoreceptor morphogenesis. Nature 416, 143149.Google Scholar
Peng, J., and Charron, F. (2013). Lateralization of motor control in the human nervous system: genetics of mirror movements. Curr. Opin. Neurobiol. 23, 109118.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. (2013). Opsins: not just for eyes. Science 339, 754755.Google Scholar
Pennisi, E. (2015). Of mice and men. Science 349, 2123.Google Scholar
Penzias, A.A., and Wilson, R.W. (1965). A measurement of excess antenna temperature at 4080 Mc/s. Astrophys. J. 142, 419421.Google Scholar
Pérez-Pomares, J.M., González-Rosa, J.M., and Muñoz-Chápuli, R. (2009). Building the vertebrate heart: an evolutionary approach to cardiac development. Int. J. Dev. Biol. 53, 14271443.Google Scholar
Perry, M.W., Boettiger, A.N., Bothma, J.P., and Levine, M. (2010). Shadow enhancers foster robustness of Drosophila gastrulation. Curr. Biol. 20, 15621567.Google Scholar
Persat, A., and Gitai, Z. (2014). Bacterial evolution: rewiring modules to get in shape. Curr. Biol. 24, R522R524. [See alsoGoogle ScholarGoogle Scholar
Pertea, M., and Salzberg, S.L. (2010). Between a chicken and a grape: estimating the number of human genes. Genome Biol. 11, Article 206.Google Scholar
Perusek, L., and Maeda, T. (2013). Vitamin A derivatives as treatment options for retinal degenerative diseases. Nutrients 5, 26462666.Google Scholar
Peso, M., Elgar, M.A., and Barron, A.B. (2015). Pheromonal control: reconciling physiological mechanism with signalling theory. Biol. Rev. 90, 542559.Google Scholar
Peterkin, T., Gibson, A., Loose, M., and Patient, R. (2005). The roles of GATA-4, -5 and -6 in vertebrate heart development. Semin. Cell Dev. Biol. 16, 8394.Google Scholar
Petralia, R.S., Mattson, M.P., and Yao, P.J. (2014). Aging and longevity in the simplest animals and the quest for immortality. Ageing Res. Rev. 16, 6682.Google Scholar
Petros, T.J., Rebsam, A., and Mason, C.A. (2008). Retinal axon growth at the optic chiasm: to cross or not to cross. Annu. Rev. Neurosci. 31, 295315.Google Scholar
Petrovic, J., Formosa-Jordan, P., Luna-Escalante, J.C., Abelló, G., Ibañes, M., Neves, J., and Giraldez, F. (2014). Ligand-dependent Notch signaling strength orchestrates lateral induction and lateral inhibition in the developing inner ear. Development 141, 23132324.Google Scholar
Petrovic, M., and Schmucker, D. (2015). Axonal wiring in neural development: target-independent mechanisms help to establish precision and complexity. BioEssays 37, 9961004.Google Scholar
Petzoldt, A.G., Coutelis, J.-B., Géminard, C., Spéder, P., Suzanne, M., Cerezo, D., and Noselli, S. (2012). DE-Cadherin regulates unconventional Myosin ID and Myosin IC in Drosophila left-right asymmetry establishment. Development 139, 18741884.Google Scholar
Peyer, B. (1947). An early description of Drosophila. J. Hered. 38, 194199.Google Scholar
Pham, H., Yu, H., and Laski, F.A. (2008). Cofilin/ADF is required for retinal elongation and morphogenesis of the Drosophila rhabdomere. Dev. Biol. 318, 8291.Google Scholar
Piatigorsky, J., and Kozmik, Z. (2004). Cubozoan jellyfish: an Evo/Devo model for eyes and other sensory systems. Int. J. Dev. Biol. 48, 719729.Google Scholar
Pichaud, F. (2014). Transcriptional regulation of tissue organization and cell morphogenesis: the fly retina as a case study. Dev. Biol. 385, 168178.Google Scholar
Pichaud, F., Briscoe, A., and Desplan, C. (1999). Evolution of color vision. Curr. Opin. Neurobiol. 9, 622627.Google Scholar
Pichaud, F., and Casares, F. (2000). homothorax and iroquois-C genes are required for the establishment of territories within the developing eye disc. Mech. Dev. 96, 1525.Google Scholar
Pichaud, F., and Desplan, C. (2002). Pax genes and eye organogenesis. Curr. Opin. Genet. Dev. 12, 430434.Google Scholar
Pichaud, F., Treisman, J., and Desplan, C. (2001). Reinventing a common strategy for patterning the eye. Cell 105, 912.Google Scholar
Pick, L., and Heffer, A. (2012). Hox gene evolution: multiple mechanisms contributing to evolutionary novelties. Ann. N. Y. Acad. Sci. 1256, 1532.Google Scholar
Pickard, G.E., and Sollars, P.J. (2012). Intrinsically photosensitive retinal ganglion cells. Rev. Physiol. Biochem. Pharmacol. 162, 5990.Google Scholar
Pierce, R.J., Wu, W., Hirai, H., Ivens, A., Murphy, L.D., Noël, C., Johnston, D.A., Artiguenave, F., Adams, M., Cornette, J., Viscogliosi, E., Capron, M., and Balavoine, G. (2005). Evidence for a dispersed Hox gene cluster in the platyhelminth parasite Schistosoma mansoni. Mol. Biol. Evol. 22, 24912503.Google Scholar
Piggins, H.D. (2002). Human clock genes. Ann. Med. 34, 394400.Google Scholar
Piñeiro, C., Lopes, C.S., and Casares, F. (2014). A conserved transcriptional network regulates lamina development in the Drosophila visual system. Development 141, 28382847.Google Scholar
Pinheiro, D., and Bellaïche, Y. (2014). Making the most of the midbody remnant: specification of the dorsal-ventral axis. Dev. Cell 28, 219220.Google Scholar
Pinkoviezky, I., and Gov, N.S. (2014). Traffic jams and shocks of molecular motors inside cellular protrusions. Phys. Rev. E 89, Article 052703.Google Scholar
Pinnell, J., Lindeman, P.S., Colavito, S., Lowe, C., and Savage, R.M. (2006). The divergent roles of the segmentation gene hunchback. Integr. Comp. Biol. 46, 519532.Google Scholar
Pirrotta, V., Chan, C.S., McCabe, D., and Qian, S. (1995). Distinct parasegmental and imaginal enhancers and the establishment of the expression pattern of the Ubx gene. Genetics 141, 14391450.Google Scholar
Pitt, J.N., and Kaeberlein, M. (2015). Why is aging conserved and what can we do about it? PLoS Biol. 13, #4, e1002131.Google Scholar
Pizzari, T. (2006). Evolution: the paradox of sperm leviathans. Curr. Biol. 16, R462R464.Google Scholar
Plachetzki, D.C., Degnan, B.M., and Oakley, T.H. (2007). The origins of novel protein interactions during animal opsin evolution. PLoS ONE 2, #10, e1054.Google Scholar
Plachetzki, D.C., Fong, C.R., and Oakley, T.H. (2010). The evolution of phototransduction from an ancestral cyclic nucleotide gated pathway. Proc. R. Soc. Lond. B 277, 19631969.Google Scholar
Plachetzki, D.C., and Oakley, T.H. (2007). Key transitions during the evolution of animal phototransduction: novelty, “tree-thinking,” co-option, and co-duplication. Integr. Comp. Biol. 47, 759769.Google Scholar
Plachetzki, D.C., Serb, J.M., and Oakley, T.H. (2005). New insights into the evolutionary history of photoreceptor cells. Trends Ecol. Evol. 20, 465467.Google Scholar
Plavicki, J., Mader, S., Pueschel, E., Peebles, P., and Boekhoff-Falk, G. (2012). Homeobox gene distal-less is required for neuronal differentiation and neurite outgrowth in the Drosophila olfactory system. PNAS 109, #5, 15781583.Google Scholar
Plavicki, J.S., Squirrell, J.M., Eliceiri, K.W., and Boekhoff-Falk, G. (2016). Expression of the Drosophila homeobox gene, Distal-less, supports an ancestral role in neural development. Dev. Dyn. 245, 8795.Google Scholar
Pohl, C., and Bao, Z. (2010). Chiral forces organize left-right patterning in C. elegans by uncoupling midline and anteroposterior axis. Dev. Cell 19, 402412.Google Scholar
Pollock, J.A., and Benzer, S. (1988). Transcript localization of four opsin genes in the three visual organs of Drosophila: RH2 is ocellus specific. Nature 333, 779782.Google Scholar
Ponce, C.R., and Born, R.T. (2008). Stereopsis. Curr. Biol. 18, R845R850.Google Scholar
Ponnambalam, S., and Alberghina, M. (2011). Evolution of the VEGF-regulated vascular network from a neural guidance system. Mol. Neurobiol. 43, 192206.Google Scholar
Poodry, C.A. (1980). Epidermis: morphology and development. In Ashburner, M. and Wright, T.R.F. (eds.), The Genetics and Biology of Drosophila, Vol. 2d. Academic Press, New York, NY, pp. 443497.Google Scholar
Popovici, C., Isnardon, D., Birnbaum, D., and Roubin, R. (2002). Caenorhabditis elegans receptors related to mammalian vascular endothelial growth factor receptors are expressed in neural cells. Neurosci. Lett. 329, 116120.Google Scholar
Porcher, A., and Dostatni, N. (2010). The Bicoid morphogen system. Curr. Biol. 20, R249R254.Google Scholar
Porter, M.L., Blasic, J.R., Bok, M.J., Cameron, E.G., Pringle, T., Cronin, T.W., and Robinson, P.R. (2012). Shedding new light on opsin evolution. Proc. R. Soc. Lond. B 279, 314.Google Scholar
Portugues, R., Severi, K.E., Wyart, C., and Ahrens, M.B. (2013). Optogenetics in a transparent animal: circuit function in the larval zebrafish. Curr. Opin. Neurobiol. 23, 119126.Google Scholar
Poss, K.D. (2010). Advances in understanding tissue regenerative capacity and mechanisms in animals. Nat. Rev. Genet. 11, 710722.Google Scholar
Poulain, F.E., and Yost, H.J. (2015). Heparan sulfate proteoglycans: a sugar code for vertebrate development? Development 142, 34563467.Google Scholar
Pourquié, O. (2003). Vertebrate somitogenesis: a novel paradigm for animal segmentation? Int. J. Dev. Biol. 47, 597603.Google Scholar
Powell, L.M., and Jarman, A.P. (2008). Context dependence of proneural bHLH proteins. Curr. Opin. Genet. Dev. 18, 411417.Google Scholar
Prasad, B.C., and Reed, R.R. (1999). Chemosensation: molecular mechanisms in worms and mammals. Trends Genet. 15, 150153.Google Scholar
Prasov, L., and Glaser, T. (2012). Pushing the envelope of retinal ganglion cell genesis: context dependent function of Math5 (Atoh7). Dev. Biol. 368, 214230.Google Scholar
Pregitzer, P., Greschista, M., Breer, H., and Krieger, J. (2014). The sensory neurone membrane protein SNMP1 contributes to the sensitivity of a pheromone detection system. Insect Mol. Biol. 23, 733742.Google Scholar
Pribil, M., Labs, M., and Leister, D. (2014). Structure and dynamics of thylakoids in land plants. J. Exp. Botany 65, 19551972.Google Scholar
Price, D.C., Egizi, A., and Fonseca, D.M. (2015). Characterization of the doublesex gene within the Culex pipiens complex suggests regulatory plasticity at the base of the mosquito sex determination cascade. BMC Evol. Biol. 15, Article 108.Google Scholar
Prieto-Godino, L.L., Diegelmann, S., and Bate, M. (2012). Embryonic origin of olfactory circuitry in Drosophila: contact and activity-mediated interactions pattern connectivity in the antennal lobe. PLoS Biol. 10, #10, e1001400.Google Scholar
Prince, V.E. (2002). The Hox paradox: more complex(es) than imagined. Dev. Biol. 249, 115.Google Scholar
Prochiantz, A., and Joliot, A. (2003). Can transcription factors function as cell-cell signalling molecules? Nat. Rev. Mol. Cell Biol. 4, 814819.Google Scholar
Prosser, H.M., Rzadzinska, A.K., Steel, K.P., and Bradley, A. (2008). Mosaic complementation demonstrates a regulatory role for myosin VIIa in actin dynamics of stereocilia. Mol. Cell. Biol. 28, #5, 17021712.Google Scholar
Provencio, I. (2011). The hidden organ in our eyes. Sci. Am. 304, #5, 5459.Google Scholar
Provencio, I., Rollag, M.D., and Castrucci, A.M. (2002). Photoreceptive net in the mammalian retina. Nature 415, 493.Google Scholar
Prpic, N.-M. (2008). Arthropod appendages: a prime example for the evolution of morphological diversity and innovation. In Minelli, A. and Fusco, G. (eds.), Evolving Pathways: Key Themes in Evolutionary Developmental Biology. Cambridge University Press, New York, NY, pp. 381398.Google Scholar
Prpic, N.-M., and Damen, W.G.M. (2009). Notch-mediated segmentation of the appendages is a molecular phylotypic trait of the arthropods. Dev. Biol. 326, 262271.Google Scholar
Prud'homme, B., de Rosa, R., Arendt, D., Julien, J.-F., Pajaziti, R., Dorresteijn, A.W.C., Adoutte, A., Wittbrodt, J., and Balavoine, G. (2003). Arthropod-like expression patterns of engrailed and wingless in the annelid Platynereis dumerilii suggest a role in segment formation. Curr. Biol. 13, 18761881.Google Scholar
Pueyo, J.I., and Couso, J.P. (2005). Parallels between the proximal-distal development of vertebrate and arthropod appendages: homology without an ancestor? Curr. Opin. Genet. Dev. 15, 439446.Google Scholar
Pugh, E.N. (2001). Rods are rods and cones cones, and (never) the twain shall meet. Neuron 32, 375380.Google Scholar
Pujol, R., Pickett, S.B., Nguyen, T.B., and Stone, J.S. (2014). Large basolateral processes on type II hair cells are novel processing units in mammalian vestibular organs. J. Comp. Neurol. 522, 31413159.Google Scholar
Punzo, C., Kurata, S., and Gehring, W.J. (2001). The eyeless homeodomain is dispensable for eye development in Drosophila. Genes Dev. 15, 17161723.Google Scholar
Purcell, P., Oliver, G., Mardon, G., Donner, A.L., and Maas, R.L. (2005). Pax6-dependence of Six3, Eya1 and Dach1 expression during lens and nasal placode induction. Gene Expr. Patterns 6, 110118.Google Scholar
Purschke, G., Arendt, D., Hausen, H., and Müller, M.C.M. (2006). Photoreceptor cells and eyes in Annelida. Arthropod Struct. Dev. 35, 211230.Google Scholar
Putnam, N.H., Butts, T., Ferrier, D.E.K., Furlong, R.F., Hellsten, U., Kawashima, T., Robinson-Rechavi, M., Shoguchi, E., Terry, A., Yu, J.-K., Benito-Gutiérrez, È., Dubchak, I., Garcia-Fernàndez, J., Gibson-Brown, J.J., Grigoriev, I.V., Horton, A.C., de Jong, P.J., Jurka, J., Kapitonov, V.V., Kohara, Y., Kuroki, Y., Lindquist, E., Lucas, S., Osoegawa, K., Pennacchio, L.A., Salamov, A.A., Satou, Y., Sauka-Spengler, T., Schmutz, J., Shin-I, T., Toyoda, A., Bronner-Fraser, M., Fujiyama, A., Holland, L.Z., Holland, P.W.H., Satoh, N., and Rokhsar, D.S. (2008). The amphioxus genome and the evolution of the chordate karyotype. Nature 453, 10641071.Google Scholar
Pyrpassopoulos, S., Feeser, E.A., Mazerik, J.N., Tyska, M.J., and Ostap, E.M. (2012). Membrane-bound Myo1c powers asymmetric motility of actin filaments. Curr. Biol. 22, 16881692.Google Scholar
Qian, L., Liu, J., and Bodmer, R. (2005). Neuromancer Tbx20-related genes (H15/midline) promote cell fate specification and morphogenesis of the Drosophila heart. Dev. Biol. 279, 509524.Google Scholar
Qian, L., Wythe, J.D., Liu, J., Cartry, J., Vogler, G., Mohapatra, B., Otway, R.T., Huang, Y., King, I.N., Maillet, M., Zheng, Y., Crawley, T., Taghli-Lamallem, O., Semsarian, C., Dunwoodie, S., Winlaw, D., Harvey, R.P., Fatkin, D., Towbin, J.A., Molkentin, J.D., Srivastava, D., Ocorr, K., Bruneau, B.G., and Bodmer, R. (2011). Tinman/Nkx2–5 acts via miR-1 and upstream of Cdc42 to regulate heart function across species. J. Cell Biol. 193, 11811196.Google Scholar
Quan, X.-J., and Hassan, B.A. (2005). From skin to nerve: flies, vertebrates and the first helix. Cell. Mol. Life Sci. 62, 20362049.Google Scholar
Quiquand, M., Yanze, N., Schmich, J., Schmid, V., Galliot, B., and Piraino, S. (2009). More constraint on ParaHox than Hox gene families in early metazoan evolution. Dev. Biol. 328, 173187.Google Scholar
Quiring, R., Walldorf, U., Kloter, U., and Gehring, W.J. (1994). Homology of the eyeless gene of Drosophila to the Small eye gene in mice and Aniridia in humans. Science 265, 785789.Google Scholar
Rabe, N., Gezelius, H., Vallstedt, A., Memic, F., and Kullander, K. (2009). Netrin-1-dependent spinal interneuron subtypes are required for the formation of left-right alternating locomotor circuitry. J. Neurosci. 29, #50, 1564215649.Google Scholar
Rader, A.J., Anderson, G., Isin, B., Khorana, H.G., Bahar, I., and Klein-Seetharaman, J. (2004). Identification of core amino acids stabilizing rhodopsin. PNAS 101, #19, 72467251.Google Scholar
Raff, R.A. (1996). The Shape of Life: Genes, Development, and the Evolution of Animal Form. University of Chicago Press, Chicago, IL.Google Scholar
Raff, R.A., and Kaufman, T.C. (1983). Embryos, Genes, and Evolution: The Developmental-Genetic Basis of Evolutionary Change. Macmillan, New York, NY.Google Scholar
Raft, S., and Groves, A.K. (2015). Segregating neural and mechanosensory fates in the developing ear: patterning, signaling, and transcriptional control. Cell Tissue Res. 359, 315332.Google Scholar
Raines, A.M., Magella, B., Adam, M., and Potter, S.S. (2015). Key pathways regulated by HoxA9,10,11/HoxD9,10,11 during limb development. BMC Dev. Biol. 15, Article 28.Google Scholar
Raj, B., and Blencowe, B.J. (2015). Alternative splicing in the mammalian nervous system: recent insights into mechanisms and functional roles. Neuron 87, 1427.Google Scholar
Rajagopalan, S., Vivancos, V., Nicolas, E., and Dickson, B.J. (2000). Selecting a longitudinal pathway: Robo receptors specify the lateral position of axons in the Drosophila CNS. Cell 103, 10331045.Google Scholar
Rakic, P. (2009). Evolution of the neocortex: a perspective from developmental biology. Nat. Rev. Neurosci. 10, 724735.Google Scholar
Ramaekers, A., Magnenat, E., Marin, E.C., Gendre, N., Jefferis, G.S.X.E., Luo, L., and Stocker, R.F. (2005). Glomerular maps without cellular redundancy at successive levels of the Drosophila larval olfactory circuit. Curr. Biol. 15, 982992.Google Scholar
Ramanathan, D.S., Gulati, T., and Ganguly, K. (2015). Sleep-dependent reactivation of ensembles in motor cortex promotes skill consolidation. PLoS Biol. 13, #9, e1002263.Google Scholar
Ramdya, P., and Benton, R. (2010). Evolving olfactory systems on the fly. Trends Genet. 26, 307316.Google Scholar
Ramel, M.-C., and Hill, C.S. (2013). The ventral to dorsal BMP activity gradient in the early zebrafish embryo is determined by graded expression of BMP ligands. Dev. Biol. 378, 170182.Google Scholar
Ramirez, M.D., and Oakley, T.H. (2015). Eye-independent, light-activated chromatophore expansion (LACE) and expression of phototransduction genes in the skin of Octopus bimaculoides. J. Exp. Biol. 218, 15131520.Google Scholar
Ramos, C., and Robert, B. (2005). msh/Msx gene family in neural development. Trends Genet. 21, 624632.Google Scholar
Ramos, O.M., Barker, D., and Ferrier, D.E.K. (2012). Ghost loci imply Hox and ParaHox existence in the last common ancestor of animals. Curr. Biol. 22, 19511956.Google Scholar
Ranade, S.S., Yang-Zhou, D., Kong, S.W., McDonald, E.C., Cook, T.A., and Pignoni, F. (2008). Analysis of the Otd-dependent transcriptome supports the evolutionary conservation of CRX/OTX/OTD functions in flies and vertebrates. Dev. Biol. 315, 521534.Google Scholar
Ranganayakulu, G., Elliott, D.A., Harvey, R.P., and Olson, E.N. (1998). Divergent roles for NK-2 class homeobox genes in cardiogenesis in flies and mice. Development 125, 30373048.Google Scholar
Rao-Mirotznik, R., Harkins, A.B., Buchsbaum, G., and Sterling, P. (1995). Mammalian rod terminal: architecture of a binary synapse. Neuron 14, 561569.Google Scholar
Rasband, K., Hardy, M., and Chien, C.-B. (2003). Generating X: formation of the optic chiasm. Neuron 39, 885888.Google Scholar
Rauskolb, C. (2001). The establishment of segmentation in the Drosophila leg. Development 128, 45114521.Google Scholar
Ravni, A., Qu, Y., Goffinet, A.M., and Tissir, F. (2009). Planar cell polarity cadherin Celsr1 regulates skin hair patterning in the mouse. J. Invest. Dermatol. 129, 25072509.Google Scholar
Ray, A., van der Goes van Naters, W., Shiraiwa, T., and Carlson, J.R. (2007). Mechanisms of odor receptor gene choice in Drosophila. Neuron 53, 353369.Google Scholar
Raya, Á., and Izpisúa Belmonte, J.C. (2006). Left-right asymmetry in the vertebrate embryo: from early information to higher-level integration. Nat. Rev. Genet. 7, 283293.Google Scholar
Raybaud, C. (2010). The corpus callosum, the other great forebrain commissures, and the septum pellucidum: anatomy, development, and malformation. Neuroradiology 52, 447477.Google Scholar
Raymond, P.A., and Barthel, L.K. (2004). A moving wave patterns the cone photoreceptor mosaic array in the zebrafish retina. Int. J. Dev. Biol. 48, 935945.Google Scholar
Raymond, P.A., Colvin, S.M., Jabeen, Z., Nagashima, M., Barthel, L.K., Hadidjojo, J., Popova, L., Pejaver, V.R., and Lubensky, D.K. (2014). Patterning the cone mosaic array in zebrafish retina requires specification of ultraviolet-sensitive cones. PLoS ONE 9, #1, e85325.Google Scholar
Ready, D.F., Hanson, T.E., and Benzer, S. (1976). Development of the Drosophila retina, a neurocrystalline lattice. Dev. Biol. 53, 217240.Google Scholar
Rebay, I., Silver, S.J., and Tootle, T.L. (2005). New vision from Eyes absent: transcription factors as enzymes. Trends Genet. 21, 163171.Google Scholar
Rebeiz, M., Castro, B., Liu, F., Yue, F., and Posakony, J.W. (2012). Ancestral and conserved cis-regulatory architectures in developmental control genes. Dev. Biol. 362, 282294.Google Scholar
Rebeiz, M., Stone, T., and Posakony, J.W. (2005). An ancient transcriptional regulatory linkage. Dev. Biol. 281, 299308.Google Scholar
Reddien, P.W., Bermange, A.L., Kicza, A.M., and Sánchez Alvarado, A. (2007). BMP signaling regulates the dorsal planarian midline and is needed for asymmetric regeneration. Development 134, 40434051.Google Scholar
Reddy, P.C., Unni, M.K., Gungi, A., Agarwal, P., and Galande, S. (2015). Evolution of Hox-like genes in Cnidaria: study of Hydra Hox repertoire reveals tailor-made Hox-code for Cnidarians. Mech. Dev. 138, 8796.Google Scholar
Reed, R.R. (2004). After the Holy Grail: establishing a molecular basis for mammalian olfaction. Cell 116, 329336.Google Scholar
Reese, B.E. (2011). Development of the retina and optic pathway. Vision Res. 51, 613632.Google Scholar
Reese, B.E., and Keeley, P.W. (2015). Design principles and developmental mechanisms underlying retinal mosaics. Biol. Rev. 90, 854876.Google Scholar
Reese, B.E., and Tan, S.-S. (1998). Clonal boundary analysis in the developing retina using X-inactivation transgenic mosaic mice. Semin. Cell Dev. Biol. 9, 285292.Google Scholar
Reeves, R.R., and Mitchell, E. (1981). The whale behind the tusk. Nat. Hist. 90, #8, 5057.Google Scholar
Rehorn, K.-P., Thelen, H., Michelson, A.M., and Reuter, R. (1996). A molecular aspect of hematopoiesis and endoderm development common to vertebrates and Drosophila. Development 122, 40234031.Google Scholar
Reichert, H. (2009). Evolutionary conservation of mechanisms for neural regionalization, proliferation and interconnection in brain development. Biol. Lett. 5, 112116.Google Scholar
Reilly, S.K., Yin, J., Ayoub, A.E., Emera, D., Leng, J., Cotney, J., Sarro, R., Rakic, P., and Noonan, J.P. (2015). Evolutionary changes in promoter and enhancer activity during human corticogenesis. Science 347, 11551159.Google Scholar
Reim, I., and Frasch, M. (2010). Genetic and genomic dissection of cardiogenesis in the Drosophila model. Pediatr. Cardiol. 31, 325334.Google Scholar
Reim, I., Mohler, J.P., and Frasch, M. (2005). Tbx20-related genes, mid and H15, are required for tinman expression, proper patterning, and normal differentiation of cardioblasts in Drosophila. Mech. Dev. 122, 10561069.Google Scholar
Reingruber, J., Holcman, D., and Fain, G.L. (2015). How rods respond to single photons: key adaptations of a G-protein cascade that enable vision at the physical limit of perception. BioEssays 37, 12431252.Google Scholar
Reiter, L.T., Potocki, L., Chien, S., Gribskov, M., and Bier, E. (2001). A systematic analysis of human disease-associated gene sequences in Drosophila melanogaster. Genome Res. 11, 11141125.Google Scholar
Repérant, J., Lemire, M., Miceli, D., and Peyrichoux, J. (1976). A radioautographic study of the visual system in fresh water teleosts following intraocular injection of tritiated fucose and proline. Brain Res. 118, 123131.Google Scholar
Reppert, S.M., and Weaver, D.R. (2002). Coordination of circadian timing in mammals. Nature 418, 935941.Google Scholar
Reuter, M., Mäntylä, K., and Gustafsson, M.K.S. (1998). Organization of the orthogon: main and minor nerve cords. Hydrobiologia 383, 175182.Google Scholar
Rezával, C., Fabre, C.C.G., and Goodwin, S.F. (2011). Invertebrate neuroethology: food play and sex. Curr. Biol. 21, R960R962.Google Scholar
Rezsohazy, R., Saurin, A.J., Maurel-Zaffran, C., and Graba, Y. (2015). Cellular and molecular insights into Hox protein action. Development 142, 12121227.Google Scholar
Rhinn, M., and Dollé, P. (2012). Retinoic acid signalling during development. Development 139, 843858.Google Scholar
Riabinina, O., Dai, M., Duke, T., and Albert, J.T. (2011). Active process mediates species-specific tuning of Drosophila ears. Curr. Biol. 21, 658664.Google Scholar
Riccomagno, M.M., and Kolodkin, A.L. (2015). Sculpting neural circuits by axon and dendrite pruning. Annu. Rev. Cell Dev. Biol. 31, 779805.Google Scholar
Richards, T.A., and Gomes, S.L. (2015). How to build a microbial eye. Nature 523, 166167.Google Scholar
Richardson, M.K. (2009). The Hox complex: an interview with Denis Duboule. Int. J. Dev. Biol. 53, 717723.Google Scholar
Richmond, D.L., and Oates, A.C. (2012). The segmentation clock: inherited trait or universal design principle? Curr. Opin. Genet. Dev. 22, 600606.Google Scholar
Rida, P.C.G., and Chen, P. (2009). Line up and listen: planar cell polarity regulation in the mammalian inner ear. Semin. Cell Dev. Biol. 20, 978985.Google Scholar
Riddle, R.D., Ensini, M., Nelson, C., Tsuchida, T., Jessell, T.M., and Tabin, C. (1995). Induction of the LIM homeobox gene Lmx1 by WNT7a establishes dorsoventral pattern in the vertebrate limb. Cell 83, 631640.Google Scholar
Ridley, M. (2004). Evolution, 3rd edn. Blackwell, Malden, MA.Google Scholar
Rieckhof, G.E., Casares, F., Ryoo, H.D., Abu-Shaar, M., and Mann, R.S. (1997). Nuclear translocation of Extradenticle requires homothorax, which encodes an Extradenticle-related homeodomain protein. Cell 91, 171183.Google Scholar
Rieder, L.E., and Larschan, E.N. (2014). Wisdom from the fly. Trends Genet. 30, 479481.Google Scholar
Rieke, F., and Rudd, M.E. (2009). The challenges natural images pose for visual adaptation. Neuron 64, 605616.Google Scholar
Riley, B.B., Chiang, M.-Y., Farmer, L., and Heck, R. (1999). The deltaA gene of zebrafish mediates lateral inhibition of hair cells in the inner ear and is regulated by pax2.1. Development 126, 56695678.Google Scholar
Rincón-Limas, D.E., Lu, C.-H., Canal, I., Calleja, M., Rodríguez-Esteban, C., Izpisúa-Belmonte, J.C., and Botas, J. (1999). Conservation of the expression and function of apterous orthologs in Drosophila and mammals. PNAS 96, 21652170.Google Scholar
Rister, J., and Desplan, C. (2011). The retinal mosaics of opsin expression in invertebrates and vertebrates. Dev. Neurobiol. 71, 12121226. [See alsoGoogle ScholarGoogle Scholar
Rister, J., Desplan, C., and Vasiliauskas, D. (2013). Establishing and maintaining gene expression patterns: insights from sensory receptor patterning. Development 140, 493503.Google Scholar
Rister, J., Razzaq, A., Boodram, P., Desai, N., Tsanis, C., Chen, H., Jukam, D., and Desplan, C. (2015). Single-base pair differences in a shared motif determine differential Rhodopsin expression. Science 350, 12581261.Google Scholar
Rivera, A.S., and Weisblat, D.A. (2009). And Lophotrochozoa makes three: Notch/Hes signaling in annelid segmentation. Dev. Genes Evol. 219, 3743.Google Scholar
Robert, D., and Hoy, R.R. (2007). Auditory systems in insects. In North, G. and Greenspan, R.J. (eds.), Invertebrate Neurobiology. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY, pp. 155184.Google Scholar
Robert, J.S. (2001). Interpreting the homeobox: metaphors of gene action and activation in development and evolution. Evol. Dev. 3, 287295.Google Scholar
Robinson, B.G., Khurana, S., Kuperman, A., and Atkinson, N.S. (2012). Neural adaptation leads to cognitive ethanol dependence. Curr. Biol. 22, 23382341.Google Scholar
Robinson, B.G., Khurana, S., Pohl, J.B., Li, W.-k., Ghezzi, A., Cady, A.M., Najjar, K., Hatch, M.M., Shah, R.R., Bhat, A., Hariri, O., Haroun, K.B., Young, M.C., Fife, K., Hooten, J., Tran, T., Goan, D., Desai, F., Husain, F., Godinez, R.M., Sun, J.C., Corpuz, J., Moran, J., Zhong, A.C., Chen, W.Y., and Atkinson, N.S. (2011). A low concentration of ethanol impairs learning but not motor and sensory behavior in Drosophila larvae. PLoS ONE 7, #5, e37394.Google Scholar
Robledo, R.F., Rajan, L., Li, X., and Lufkin, T. (2002). The Dlx5 and Dlx6 homeobox genes are essential for craniofacial, axial, and appendicular skeletal development. Genes Dev. 16, 10891101.Google Scholar
Rochlin, K., Yu, S., Roy, S., and Baylies, M.K. (2010). Myoblast fusion: when it takes more to make one. Dev. Biol. 341, 6683.Google Scholar
Rodriguez, I., Greer, C.A., Mok, M.Y., and Mombaerts, P. (2000). A putative pheromone receptor gene expressed in human olfactory mucosa. Nat. Genet. 26, 1819.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
Rodríguez-Trelles, F., Tarrío, R., and Ayala, F.J. (2005). Is ectopic expression caused by deregulatory mutations or due to gene-regulation leaks with evolutionary potential? BioEssays 27, 592601.Google Scholar
Rogers, G.E. (2004). Hair follicle differentiation and regulation. Int. J. Dev. Biol. 48, 163170.Google Scholar
Rogers, K.W., and Schier, A.F. (2011). Morphogen gradients: from generation to interpretation. Annu. Rev. Cell Dev. Biol. 27, 377407.Google Scholar
Rogers, L.J. (2014). Asymmetry of brain and behavior in animals: its development, function, and human relevance. Genesis 52, 555571.Google Scholar
Rogers, L.J., Vallortigara, G., and Andrew, R.J. (2013). Divided Brains: The Biology and Behaviour of Brain Asymmetries. Cambridge University Press, Cambridge.Google Scholar
Rogulja-Ortmann, A., Lüer, K., Seibert, J., Rickert, C., and Technau, G.M. (2007). Programmed cell death in the embryonic central nervous system of Drosophila melanogaster. Development 134, 105116.Google Scholar
Rokas, A. (2008). The origins of multicellularity and the early history of the genetic toolkit for animal development. Annu. Rev. Genet. 42, 235251.Google Scholar
Root, C.M., Denny, C.A., Hen, R., and Axel, R. (2014). The participation of cortical amygdala in innate odour-driven behaviour. Nature 515, 269273.Google Scholar
Rosato, E., and Kyriacou, C.P. (2011). The role of natural selection in circadian behaviour: a molecular-genetic approach. Essays Biochem. 49, 7185.Google Scholar
Rosbash, M. (2009). The implications of multiple circadian clock origins. PLoS Biol. 7, #3, e1000062.Google Scholar
Rosenbaum, D.M., Rasmussen, S.G.F., and Kobilka, B.K. (2009). The structure and function of G-protein-coupled receptors. Nature 459, 356363.Google Scholar
Rosenberg, M.I., Lynch, J.A., and Desplan, C. (2009). Heads and tails: evolution of antero-posterior patterning in insects. Biochim. Biophys. Acta 1789, 333342.Google Scholar
Rosenstein, J.M., Krum, J.M., and Ruhrberg, C. (2010). VEGF in the nervous system. Organogenesis 6, 107114.Google Scholar
Rosenwasser, A.M. (2009). Functional neuroanatomy of sleep and circadian rhythms. Brain Res. Rev. 61, 281306.Google Scholar
Rosselló, R.A., Chen, C.-C., Dai, R., Howard, J.T., Hochgeschwender, U., and Jarvis, E.D. (2013). Mammalian genes induce partially reprogrammed pluripotent stem cells in non-mammalian vertebrate and invertebrate species. eLife 2, e00036.Google Scholar
Roth, G. (2015). Convergent evolution of complex brains and high intelligence. Philos. Trans. R. Soc. Lond. B 370, 20150049.Google Scholar
Roth, S., and Lynch, J. (2012). Does the Bicoid gradient matter? Cell 149, 511512.Google Scholar
Roth, S., and Panfilio, K.A. (2012). Making waves for segments. Science 336, 306307.Google Scholar
Roth, S., Stein, D., and Nüsslein-Volhard, C. (1989). A gradient of nuclear localization of the dorsal protein determines dorsoventral pattern in the Drosophila embryo. Cell 59, 11891202.Google Scholar
Röttinger, E., and Lowe, C.J. (2012). Evolutionary crossroads in developmental biology: hemichordates. Development 139, 24632475.Google Scholar
Roush, S., and Slack, F.J. (2008). The let-7 family of microRNAs. Trends Cell Biol. 18, 505516.Google Scholar
Rovira, M., Saavedra, P., Casal, J., and Lawrence, P.A. (2015). Regions within a single epidermal cell of Drosophila can be planar polarised independently. eLife 4, e06303.Google Scholar
Roy, S.W. (2006). Intron-rich ancestors. Trends Genet. 22, 468471.Google Scholar
Royo, J.L., Maeso, I., Irimia, M., Gao, F., Peter, I.S., Lopes, C.S., D'Aniello, S., Casares, F., Davidson, E.H., Garcia-Fernández, J., and Gómez-Skarmeta, J.L. (2011). Transphyletic conservation of developmental regulatory state in animal evolution. PNAS 108, #34, 1418614191.Google Scholar
Ru, H., Chambers, M.G., Fu, T.-M., Tong, A.B., Liao, M., and Wu, H. (2015). Molecular mechanism of V(D)J recombination from synaptic RAG1-RAG2 complex structures. Cell 163, 11381152.Google Scholar
Rubin, G.M. (2001). The draft sequences: comparing species. Nature 409, 820821.Google Scholar
Rubinstein, R., Thu, C.A., Goodman, K.M., Wolcott, H.N., Bahna, F., Mannepalli, S., Ahlsen, G., Chevee, M., Halim, A., Clausen, H., Maniatis, T., Shapiro, L., and Honig, B. (2015). Molecular logic of neuronal self-recognition through protocadherin domain interactions. Cell 163, 629642.Google Scholar
Rudrapatna, V.A., Cagan, R.L., and Das, T.K. (2012). Drosophila cancer models. Dev. Dyn. 241, 107118.Google Scholar
Rushlow, C.A., Han, K., Manley, J.L., and Levine, M. (1989). The graded distribution of the dorsal morphogen is initiated by selective nuclear transport in Drosophila. Cell 59, 11651177.Google Scholar
Rusten, T.E., Cantera, R., Kafatos, F.C., and Barrio, R. (2002). The role of TGFβ signaling in the formation of the dorsal nervous system is conserved between Drosophila and chordates. Development 129, 35753584.Google Scholar
Ruzickova, J., Piatigorsky, J., and Kozmik, Z. (2009). Eye-specific expression of an ancestral jellyfish PaxB gene interferes with Pax6 function despite its conserved Pax6/Pax2 characteristics. Int. J. Dev. Biol. 53, 469482.Google Scholar
Ryan, J.F., and Baxevanis, A.D. (2007). Hox, Wnt, and the evolution of the primary body axis: insights from the early-divergent phyla. Biol. Direct 2, Article 37.Google Scholar
Ryan, J.F., and Chiodin, M. (2015). Where is my mind? How sponges and placozoans may have lost neural cell types. Philos. Trans. R. Soc. Lond. B 370, 20150059.Google Scholar
Ryan, J.F., Mazza, M.E., Pang, K., Matus, D.Q., Baxevanis, A.D., Martindale, M.Q., and Finnerty, J.R. (2007). Pre-bilaterian origins of the Hox cluster and the Hox code: evidence from the sea anemone, Nematostella vectensis. PLoS ONE 2, #1, e153.Google Scholar
Ryan, T.J., and Grant, S.G.N. (2009). The origin and evolution of synapses. Nat. Rev. Neurosci. 10, 701712.Google Scholar
Ryoo, H.D., Marty, T., Casares, F., Affolter, M., and Mann, R.S. (1999). Regulation of Hox target genes by a DNA bound Homothorax/Hox/Extradenticle complex. Development 126, 51375148.Google Scholar
Rytz, R., Croset, V., and Benton, R. (2013). Ionotropic receptors (IRs): chemosensory ionotropic glutamate receptors in Drosophila and beyond. Insect Biochem. Mol. Biol. 43, 888897.Google Scholar
Ryu, J.-R., Najand, N., and Brook, W.J. (2011). Tinman is a direct activator of midline in the Drosophila dorsal vessel. Dev. Dyn. 240, 8695.Google Scholar
Saavedra, P., Vincent, J.-P., Palacios, I.M., Lawrence, P.A., and Casal, J. (2014). Plasticity of both planar cell polarity and cell identity during the development of Drosophila. eLife 3, e01569.Google Scholar
Saburi, S., Hester, I., Goodrich, L., and McNeill, H. (2012). Functional interactions between Fat family cadherins in tissue morphogenesis and planar polarity. Development 139, 18061820.Google Scholar
Sadaf, S., Reddy, O.V., Sane, S.P., and Hasan, G. (2015). Neural control of wing coordination in flies. Curr. Biol. 25, 8086.Google Scholar
Saga, Y. (2012). The mechanism of somite formation in mice. Curr. Opin. Genet. Dev. 22, 331338.Google Scholar
Saha, D., and Raman, B. (2015). Relating early olfactory processing with behavior: a perspective. Curr. Opin. Insect Sci. 12, 5463.Google Scholar
Sakamaki, K., Imai, K., Tomii, K., and Miller, D.J. (2015). Evolutionary analyses of caspase-8 and its paralogs: deep origins of the apoptotic signaling pathways. BioEssays 37, 767776.Google Scholar
Sakmar, T.P. (2012). Redder than red. Science 338, 12991300.Google Scholar
Sakurai, T., Namiki, S., and Kanzaki, R. (2014). Molecular and neural mechanisms of sex pheromone reception and processing in the silkmoth Bombyx mori. Front. Physiol. 5, Article 125.Google Scholar
Salcedo, E., Farrell, D.M., Zheng, L., Phistry, M., Bagg, E.E., and Britt, S.G. (2009). The green-absorbing Drosophila Rh6 visual pigment contains a blue-shifting amino acid substitution that is conserved in vertebrates. J. Biol. Chem. 284, 57175722.Google Scholar
Salcedo, E., Huber, A., Henrich, S., Chadwell, L.V., Chou, W.-H., Paulsen, R., and Britt, S.G. (1999). Blue- and green-absorbing visual pigments of Drosophila: ectopic expression and physiological characterization of the R8 photoreceptor cell-specific Rh5 and Rh6 rhodopsins. J. Neurosci. 19, #24, 1071610726.Google Scholar
Sallé, J., Campbell, S.D., Gho, M., and Audibert, A. (2012). CycA is involved in the control of endoreplication dynamics in the Drosophila bristle lineage. Development 139, 547557.CrossRefGoogle ScholarPubMed
Salzberg, A., and Bellen, H.J. (1996). Invertebrate versus vertebrate neurogenesis: variations on the same theme? Dev. Genet. 18, 110.Google Scholar
Samadi, L., and Steiner, G. (2010). Expression of Hox genes during the larval development of the snail, Gibbula varia (L.): further evidence of non-colinearity in molluscs. Dev. Genes Evol. 220, 161172.CrossRefGoogle ScholarPubMed
Sánchez, L., and Guerrero, I. (2001). The development of the Drosophila genital disc. BioEssays 23, 698707.Google Scholar
Sánchez-Gracia, A., Vieira, F.G., and Rozas, J. (2009). Molecular evolution of the major chemosensory gene families in insects. Heredity 103, 208216.Google Scholar
Sánchez-Herrero, E., Vernós, I., Marco, R., and Morata, G. (1985). Genetic organization of Drosophila bithorax complex. Nature 313, 108113.CrossRefGoogle ScholarPubMed
Sander, M., Neubüser, A., Ee, H.C., Martin, G.R., and German, M.S. (1997). Genetic analysis reveals that PAX6 is required for normal transcription of pancreatic hormone genes and islet development. Genes Dev. 11, 16621673.Google Scholar
Sanes, J.R., and Zipursky, S.L. (2010). Design principles of insect and vertebrate visual systems. Neuron 66, 1536.Google Scholar
Santiago, C., Labrador, J.-P., and Bashaw, G.J. (2014). The homeodomain transcription factor Hb9 controls axon guidance in Drosophila through the regulation of Robo receptors. Cell Rep. 7, 153165.Google Scholar
Sarin, S., O'Meara, M., Flowers, E.B., Antonio, C., Poole, R.J., Didiano, D., Johnston, R.J. Jr., Chang, S., Narula, S., and Hobert, O. (2007). Genetic screens for Caenorhabditis elegans mutants defective in left/right asymmetric neuronal fate specification. Genetics 176, 21092130.Google Scholar
Sarnat, H.B., and Netsky, M.G. (1981). Evolution of the Nervous System. Oxford University Press, New York, NY.Google Scholar
Sarnat, H.B., and Netsky, M.G. (2002). When does a ganglion become a brain? Evolutionary origin of the central nervous system. Semin. Pediatr. Neurol. 9, 240253.Google Scholar
Sarrazin, A.F., Peel, A.D., and Averof, M. (2012). A segmentation clock with two-segment periodicity in insects. Science 336, 338341.Google Scholar
Sasai, Y. (2001). Roles of Sox factors in neural determination: conserved signaling in evolution? Int. J. Dev. Biol. 45, 321326.Google Scholar
Sassi, N., Laadhar, L., Driss, M., Kallel-Sellami, M., Sellami, S., and Makni, S. (2011). The role of the Notch pathway in healthy and osteoarthritic articular cartilage: from experimental models to ex vivo studies. Arthritis Res. Ther. 13, Article 208.Google Scholar
Sato, K., Pellegrino, M., Nakagawa, T., Nakagawa, T., Vosshall, L.B., and Touhara, K. (2008). Insect olfactory receptors are heteromeric ligand-gated ion channels. Nature 452, 10021006.Google Scholar
Sato, M., Suzuki, T., and Nakai, Y. (2013). Waves of differentiation in the fly visual system. Dev. Biol. 380, 111.Google Scholar
Satoh, A.K., Li, B.X., Xia, H., and Ready, D.F. (2008). Calcium-activated Myosin V closes the Drosophila pupil. Curr. Biol. 18, 951955.Google Scholar
Satoh, N., Tagawa, K., and Takahashi, H. (2012). How was the notochord born? Evol. Dev. 14, 5675.Google Scholar
Satou, Y., and Satoh, N. (2006). Gene regulatory networks for the development and evolution of the chordate heart. Genes Dev. 20, 26342638.Google Scholar
Saudemont, A., Dray, N., Hudry, B., Le Gouar, M., Vervoort, M., and Balavoine, G. (2008). Complementary striped expression patterns of NK homeobox genes during segment formation in the annelid Platynereis. Dev. Biol. 317, 430443.Google Scholar
Savory, J.G.A., Pilon, N., Grainger, S., Sylvestre, J.-R., Béland, M., Houle, M., Oh, K., and Lohnes, D. (2009). Cdx1 and Cdx2 are functionally equivalent in vertebral patterning. Dev. Biol. 330, 114122.Google Scholar
Sawaya, M.R., Wojtowicz, W.M., Andre, I., Qian, B., Wu, W., Baker, D., Eisenberg, D., and Zipursky, S.L. (2008). A double S shape provides the structural basis for the extraordinary binding specificity of Dscam isoforms. Cell 134, 10071018.Google Scholar
Scheerer, P., Park, J.H., Hildebrand, P.W., Kim, Y.J., Krauss, N., Choe, H.-W., Hofmann, K.P., and Ernst, O.P. (2008). Crystal structure of opsin in its G-protein-interacting conformation. Nature 455, 497502.CrossRefGoogle ScholarPubMed
Schetelig, M.F., Schmid, B.G.M., Zimowska, G., and Wimmer, E.A. (2008). Plasticity in mRNA expression and localization of orthodenticle within higher Diptera. Evol. Dev. 10, 700704.Google Scholar
Schier, A.F. (2009). Nodal morphogens. Cold Spring Harb. Perspect. Biol. 1, a003459.Google Scholar
Schilling, T.F., and Knight, R.D. (2001). Origins of anteroposterior patterning and Hox gene regulation during chordate evolution. Philos. Trans. R. Soc. Lond. B 356, 15991613.Google Scholar
Schilling, T.F., Nie, Q., and Lander, A.D. (2012). Dynamics and precision in retinoic acid morphogen gradients. Curr. Opin. Genet. Dev. 22, 562569.Google Scholar
Schinko, J.B., Kreuzer, N., Offen, N., Posnien, N., Wimmer, E.A., and Bucher, G. (2008). Divergent functions of orthodenticle, empty spiracles and buttonhead in early head patterning of the beetle Tribolium castaneum (Coleoptera). Dev. Biol. 317, 600613.Google Scholar
Schippers, K.J., and Nichols, S.A. (2014). Deep, dark secrets of melatonin in animal evolution. Cell 159, 910.Google Scholar
Schlosser, G., Patthey, C., and Shimeld, S.M. (2014). The evolutionary history of vertebrate cranial placodes. II. Evolution of ectodermal patterning. Dev. Biol. 389, 98119.CrossRefGoogle ScholarPubMed
Schmidt, J., Francois, V., Bier, E., and Kimelman, D. (1995). Drosophila short gastrulation induces an ectopic axis in Xenopus: evidence for conserved mechanisms of dorsal-ventral patterning. Development 121, 43194328.Google Scholar
Schmidt, M.H. (2014). The energy allocation function of sleep: a unifying theory of sleep, torpor, and continuous wakefulness. Neurosci. Biobehav. Rev. 47, 122153.Google Scholar
Schmidt, T.M., Chen, S.-K., and Hattar, S. (2011). Intrinsically photosensitive retinal ganglion cells: many subtypes, diverse functions. Trends Neurosci. 34, 572580.Google Scholar
Schmidt-Rhaesa, A. (2007). The Evolution of Organ Systems. Oxford University Press, New York, NY.Google Scholar
Schmidt-Ullrich, R., and Paus, R. (2005). Molecular principles of hair follicle induction and morphogenesis. BioEssays 27, 247261.Google Scholar
Schmucker, D., and Chen, B. (2009). Dscam and DSCAM: complex genes in simple animals, complex animals yet simple genes. Genes Dev. 23, 147156.Google Scholar
Schmucker, D., Clemens, J.C., Shu, H., Worby, C.A., Xiao, J., Muda, M., Dixon, J.E., and Zipursky, S.L. (2000). Drosophila Dscam is an axon guidance receptor exhibiting extraordinary molecular diversity. Cell 101, 671684.Google Scholar
Schnaitmann, C., Garbers, C., Wachtler, T., and Tanimoto, H. (2013). Color discrimination with broadband photoreceptors. Curr. Biol. 23, 23752382.Google Scholar
Schneider, M.D. (2016). Heartbreak hotel: a convergence in cardiac regeneration. Development 143, 14351441.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
Schoenwolf, G.C., Bleyl, S.B., Brauer, P.R., and Francis-West, P.H. (2009). Larsen's Human Embryology, 4th edn. Churchill Livingstone, Philadelphia, PA.Google Scholar
Scholtz, G. (2002). The Articulata hypothesis: or what is a segment? Org. Divers. Evol. 2, 197215.Google Scholar
Schön, P., Tsuchiya, K., Lenoir, D., Mochizuki, T., Guichard, C., Takai, S., Maiti, A.K., Nihei, H., Weil, J., Yokoyama, T., and Bouvagnet, P. (2002). Identification, genomic organization, chromosomal mapping and mutation analysis of the human INV gene, the ortholog of a murine gene implicated in left-right axis development and biliary atresia. Hum. Genet. 110, 157165.Google Scholar
Schonegg, S., Hyman, A.A., and Wood, W.B. (2014). Timing and mechanism of the initial cue establishing handed left-right asymmetry in Caenorhabditis elegans embryos. Genesis 52, 572580.Google Scholar
Schoppmeier, M., Fischer, S., Schmitt-Engel, C., Löhr, U., and Klingler, M. (2009). An ancient anterior patterning system promotes caudal repression and head formation in Ecdysozoa. Curr. Biol. 19, 18111815.Google Scholar
Schreiner, D., Nguyen, T.-M., Russo, G., Heber, S., Patrignani, A., Ahrné, E., and Scheiffele, P. (2014). Targeted combinatorial alternative splicing generates brain region-specific repertoires of neurexins. Neuron 84, 386398.Google Scholar
Schröder, R. (2003). The genes orthodenticle and hunchback substitute for bicoid in the beetle Tribolium. Nature 422, 621625.Google Scholar
Schroeder, J.A., Jackson, L.F., Lee, D.C., and Camenisch, T.D. (2003). Form and function of developing heart valves: coordination by extracellular matrix and growth factor signaling. J. Mol. Med. 81, 392403.Google Scholar
Schroeder, M.D., Greer, C., and Gaul, U. (2011). How to make stripes: deciphering the transition from non-periodic to periodic patterns in Drosophila segmentation. Development 138, 30673078.Google Scholar
Schubert, M., Escriva, H., Xavier-Neto, J., and Laudet, V. (2006). Amphioxus and tunicates as evolutionary model systems. Trends Ecol. Evol. 21, 269277.Google Scholar
Schubert, M., Yu, J.-K., Holland, N.D., Escriva, H., Laudet, V., and Holland, L.Z. (2005). Retinoic acid signaling acts via Hox1 to establish the posterior limit of the pharynx in the chordate amphioxus. Development 132, 6173.Google Scholar
Schughart, K., Kappen, C., and Ruddle, F.H. (1988). Mammalian homeobox-containing genes: genome organization, structure, expression and evolution. Br. J. Cancer 58 (Suppl. IX), 913.Google Scholar
Schulte, D., and Bumsted-O'Brien, K.M. (2008). Molecular mechanisms of vertebrate retina development: implications for ganglion cell and photoreceptor patterning. Brain Res. 1192, 151164.Google Scholar
Schulze, J., and Schierenberg, E. (2009). Embryogenesis of Romanomermis culicivorax: an alternative way to construct a nematode. Dev. Biol. 334, 1021.Google Scholar
Schuster, S., Strauss, R., and Götz, K.G. (2002). Virtual-reality techniques resolve the visual cues used by fruit flies to evaluate object distances. Curr. Biol. 12, 15911594.Google Scholar
Schwab, I.R. (2012). Evolution's Witness: How Eyes Evolved. Oxford University Press, New York, NY.Google Scholar
Schwabe, T., Borycz, J.A., Meinertzhagen, I.A., and Clandinin, T.R. (2014). Differential adhesion determines the organization of synaptic fascicles in the Drosophila visual system. Curr. Biol. 24, 13041313.Google Scholar
Schwander, M., Kachar, B., and Müller, U. (2010). The cell biology of hearing. J. Cell Biol. 190, 920.Google Scholar
Schwartz, T.W., and Hubbell, W.L. (2008). A moving story of receptors. Nature 455, 473474.Google Scholar
Schwartz, W.J. (2004). Sunrise and sunset in fly brains. Nature 431, 751752.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
Schweickert, A., Walentek, P., Thumberger, T., and Danilchik, M. (2012). Linking early determinants and cilia-driven leftward flow in left-right axis specificiation of Xenopus laevis: a theoretical approach. Differentiation 83, S67S77.Google Scholar
Schweitzer, R., Zelzer, E., and Volk, T. (2010). Connecting muscles to tendons: tendons and musculoskeletal development in flies and vertebrates. Development 137, 28072817.Google Scholar
Scotland, R.W. (2011). What is parallelism? Evol. Dev. 13, 214227.Google Scholar
Scott, C.A., and Dahanukar, A. (2014). Sensory coding of olfaction and taste. In Dubnau, J. (ed.), Behavioral Genetics of the Fly (Drosophila melanogaster). Cambridge University Press, New York, NY, pp. 4965.Google Scholar
Scott, M.P. (1994). Intimations of a creature. Cell 79, 11211124.Google Scholar
Scott, M.P. (2000). Development: the natural history of genes. Cell 100, 2740.Google Scholar
Scott, M.P. (2016). Homeodomains, hedgehogs, and happiness. Curr. Top. Dev. Biol. 117, 331337.Google Scholar
Scott, M.P., and Carroll, S.B. (1987). The segmentation and homeotic gene network in early Drosophila development. Cell 51, 689698.Google Scholar
Scott, M.P., and Weiner, A.J. (1984). Structural relationships among genes that control development: sequence homology between the Antennapedia, Ultrabithorax, and fushi tarazu loci of Drosophila. PNAS 81, 41154119.Google Scholar
Seaver, E.C. (2003). Segmentation: mono- or polyphyletic? Int. J. Dev. Biol. 47, 583595.Google Scholar
Sebé-Pedrós, A., Burkhardt, P., Sánchez-Pons, N., Fairclough, S.R., Lang, B.F., King, N., and Ruiz-Trillo, I. (2013). Insights into the origin of metazoan filopodia and microvilli. Mol. Biol. Evol. 30, 20132023.Google Scholar
Seeds, A.M., Ravbar, P., Chung, P., Hampel, S., Midgley, F.M. Jr., Mensh, B.D., and Simpson, J.H. (2014). A suppression hierarchy among competing motor programs drives sequential grooming in Drosophila. eLife 3, e02951.Google Scholar
Seeger, M., Tear, G., Ferres-Marco, D., and Goodman, C.S. (1993). Mutations affecting growth cone guidance in Drosophila: genes necessary for guidance toward or away from the midline. Neuron 10, 409426.Google Scholar
Seelig, J.D., and Jayaraman, V. (2015). Neural dynamics for landmark orientation and angular path integration. Nature 521, 186191.Google Scholar
Sehadova, H., Glaser, F.T., Gentile, C., Simoni, A., Giesecke, A., and Albert, J.T. (2009). Temperature entrainment of Drosophila's circadian clock involves the gene nocte and signaling from peripheral sensory tissues to the brain. Neuron 64, 251266.Google Scholar
Seibert, J., Volland, D., and Urbach, R. (2009). Ems and Nkx6 are central regulators in dorsoventral patterning of the Drosophila brain. Development 136, 39373947.Google Scholar
Seifert, A.W., Kiama, S.G., Seifert, M.G., Goheen, J.R., Palmer, T.M., and Maden, M. (2012). Skin shedding and tissue regeneration in African spiny mice (Acomys). Nature 489, 561565.Google Scholar
Seifert, A.W., Monaghan, J.R., Voss, S.R., and Maden, M. (2012). Skin regeneration in adult axolotls: a blueprint for scar-free healing in vertebrates. PLoS ONE 7, #4, e32875.Google Scholar
Seifert, J.R.K., and Mlodzik, M. (2007). Frizzled/PCP signalling: a conserved mechanism regulating cell polarity and directed motility. Nat. Rev. Genet. 8, 126138.Google Scholar
Seipel, K., Eberhardt, M., Müller, P., Pescia, E., Yanze, N., and Schmid, V. (2004). Homologs of vascular endothelial growth factor and receptor, VEGF and VEGFR, in the jellyfish Podocoryne carnea. Dev. Dyn. 231, 303312.Google Scholar
Sekharan, S., and Morokuma, K. (2011). Why 11-cis-retinal? Why not 7-cis-, 9-cis-, or 13-cis-retinal in the eye? J. Am. Chem. Soc. 133, 1905219055.Google Scholar
Selverston, A.I. (2010). Invertebrate central pattern generator circuits. Philos. Trans. R. Soc. Lond. B 365, 23292345.Google Scholar
Semmelhack, J.L., and Wang, J.W. (2009). Select Drosophila glomeruli mediate innate olfactory attraction and aversion. Nature 459, 218223.CrossRefGoogle ScholarPubMed
Sen, S., Reichert, H., and VijayRaghavan, K. (2013). Conserved roles of ems/Emx and otd/Otx genes in olfactory and visual system development in Drosophila and mouse. Open Biol. 3, Article 120177.Google Scholar
Senoo, H., Sesaki, H., and Iijima, M. (2016). A GPCR handles bacterial sensing in chemotaxis and phagocytosis. Dev. Cell 36, 354356.Google Scholar
Senthilan, P.R., Piepenbrock, D., Ovezmyradov, G., Nadrowski, B., Bechstedt, S., Pauls, S., Winkler, M., Möbius, W., Howard, J., and Göpfert, M.C. (2012). Drosophila auditory organ genes and genetic hearing defects. Cell 150, 10421054.Google Scholar
Seo, H.-C., Edvardsen, R.B., Maeland, A.D., Bjordal, M., Jensen, M.F., Hansen, A., Flaat, M., Weissenbach, J., Lehrach, H., Wincker, P., Reinhardt, R., and Chourrout, D. (2004). Hox cluster disintegration with persistent anteroposterior order of expression in Oikopleura dioica. Nature 431, 6771.Google Scholar
Serizawa, S., Miyamichi, K., Nakatani, H., Suzuki, M., Saito, M., Yoshihara, Y., and Sakano, H. (2003). Negative feedback regulation ensures the one receptor-one olfactory neuron rule in mouse. Science 302, 20882094.Google Scholar
Serizawa, S., Miyamichi, K., Takeuchi, H., Yamagishi, Y., Suzuki, M., and Sakano, H. (2006). A neuronal identity code for the odorant receptor-specific and activity-dependent axon sorting. Cell 127, 10571069.Google Scholar
Shapiro, L. (2007). Self-recognition at the atomic level: understanding the astonishing molecular diversity of homophilic Dscams. Neuron 56, 1013.CrossRefGoogle ScholarPubMed
Sharma, P.P., Santiago, M.A., González-Santillán, E., Monod, L., and Wheeler, W.C. (2015). Evidence of duplicated Hox genes in the most recent common ancestor of extant scorpions. Evol. Dev. 17, 347355.Google Scholar
Shaw, P.J., and Franken, P. (2003). Perchance to dream: solving the mystery of sleep through genetic analysis. J. Neurobiol. 54, 179202.Google Scholar
Sheeba, C.J., Andrade, R.P., and Palmeirim, I. (2016). Getting a handle on embryo limb development: molecular interactions driving limb outgrowth and patterning. Semin. Cell Dev. Biol. 49, 92101.Google Scholar
Shen, H.H. (2014). Inner workings: discovering the split mind. PNAS 111, #51, 18097.Google Scholar
Shen, M.M. (2007). Nodal signaling: developmental roles and regulation. Development 134, 10231034.Google Scholar
Shen, W.L., Kwon, Y., Adegbola, A.A., Luo, J., Chess, A., and Montell, C. (2011). Function of rhodopsin in temperature discrimination in Drosophila. Science 331, 13331336.Google Scholar
Shepherd, G.M. (2006). Smell images and the flavour system in the human brain. Nature 444, 316321.Google Scholar
Sheth, R., Grégoire, D., Dumouchel, A., Scotti, M., Pham, J.M.T., Nemec, S., Bastida, M.F., Ros, M.A., and Kmita, M. (2013). Decoupling the function of Hox and Shh in developing limb reveals multiple inputs of Hox genes on limb growth. Development 140, 21302138.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
Shibazaki, Y., Shimizu, M., and Kuroda, R. (2004). Body handedness is directed by genetically determined cytoskeletal dynamics in the early embryo. Curr. Biol. 14, 14621467.Google Scholar
Shichida, Y., and Matsuyama, T. (2009). Evolution of opsins and phototransduction. Philos. Trans. R. Soc. Lond. B 364, 28812895.Google Scholar
Shimeld, S.M., and Levin, M. (2006). Evidence for the regulation of left-right asymmetry in Ciona intestinalis by ion flux. Dev. Dyn. 235, 15431553.Google Scholar
Shimizu, H., and Fujisawa, T. (2003). Peduncle of Hydra and the heart of higher organisms share a common ancestral origin. Genesis 36, 182186.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
Shimozono, S., Iimura, T., Kitaguchi, T., Higashijima, S.-I., and Miyawaki, A. (2013). Visualization of an endogenous retinoic acid gradient across embryonic development. Nature 496, 363366.Google Scholar
Shinbrot, T., and Young, W. (2008). Why decussate? Topological constraints on 3D wiring. Anat. Rec. 291, 12781292.Google Scholar
Shinohara, K., Chen, D., Nishida, T., Misaki, K., Yonemura, S., and Hamada, H. (2015). Absence of radial spokes in mouse node cilia is required for rotational movement but confers ultrastructural instability as a trade-off. Dev. Cell 35, 236246.Google Scholar
Shinomiya, K., Takemura, S.-Y., Rivlin, P.K., Plaza, S.M., Scheffer, L.K., and Meinertzhagen, I.A. (2015). A common evolutionary origin for the ON- and OFF-edge motion detection pathways of the Drosophila visual system. Front. Neural Circuits 9, Article 33.Google Scholar
Shippy, T.D., Ronshaugen, M., Cande, J., He, J.P., Beeman, R.W., Levine, M., Brown, S.J., and Denell, R. (2008). Analysis of the Tribolium homeotic complex: insights into mechanisms constraining insect Hox clusters. Dev. Genes Evol. 218, 127139.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
Shiraiwa, T. (2008). Multimodal chemosensory integration through the maxillary palp in Drosophila. PLoS ONE 3, #5, e2191.Google Scholar
Shirasaki, R., and Pfaff, S.L. (2002). Transcriptional codes and the control of neuronal identity. Annu. Rev. Neurosci. 25, 251281.Google Scholar
Shlyueva, D., Stampfel, G., and Stark, A. (2014). Transcriptional enhancers: from properties to genome-wide predictions. Nat. Rev. Genet. 15, 272286.Google Scholar
Shohat-Ophir, G., Kaun, K.R., Azanchi, R., and Heberlein, U. (2012). Sexual deprivation increases ethanol intake in Drosophila. Science 335, 13511355.CrossRefGoogle ScholarPubMed
Sholtis, S.J., and Noonan, J.P. (2010). Gene regulation and the origins of human biological uniqueness. Trends Genet. 26, 110118.Google Scholar
Shubin, N., Tabin, C., and Carroll, S. (1997). Fossils, genes and the evolution of animal limbs. Nature 388, 639648.Google Scholar
Shubin, N., Tabin, C., and Carroll, S. (2009). Deep homology and the origins of evolutionary novelty. Nature 457, 818823.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
Siegert, S., Cabuy, E., Scherf, B.G., Kohler, H., Panda, S., Le, Y.-Z., Fehling, H.J., Gaidatzis, D., Stadler, M.B., and Roska, B. (2012). Transcriptional code and disease map for adult retinal cell types. Nat. Neurosci. 15, 487495.Google Scholar
Sienknecht, U.J. (2015). Current concepts of hair cell differentiation and planar cell polarity in inner ear sensory organs. Cell Tissue Res. 361, 2532.Google Scholar
Sienknecht, U.J., Anderson, B.K., Parodi, R.M., Fantetti, K.N., and Fekete, D.M. (2011). Non-cell-autonomous planar cell polarity propagation in the auditory sensory epithelium of vertebrates. Dev. Biol. 352, 2739.Google Scholar
Sillar, K.T. (2009). Escape behaviour: reciprocal inhibition ensures effective escape trajectory. Curr. Biol. 19, R697R699.Google Scholar
Silver, S.J., and Rebay, I. (2005). Signaling circuitries in development: insights from the retinal determination gene network. Development 132, 313.Google Scholar
Silverman, H.B., and Dunbar, M.J. (1980). Aggressive tusk use by the narwhal (Monodon monoceros L.). Nature 284, 5758.Google Scholar
Simakov, O., Marletaz, F., Cho, S.-J., Edsinger-Gonzales, E., Havlak, P., Hellsten, U., Kuo, D.-H., Larsson, T., Lv, J., Arendt, D., Savage, R., Osoegawa, K., de Jong, P., Grimwood, J., Chapman, J.A., Shapiro, H., Aerts, A., Otillar, R.P., Terry, A.Y., Boore, J.L., Grigoriev, I.V., Lindberg, D.R., Seaver, E.C., Weisblat, D.A., Putnam, N.H., and Rokhsar, D.S. (2013). Insights into bilaterian evolution from three spiralian genomes. Nature 493, 526531.Google Scholar
Simeone, A., Puelles, E., and Acampora, D. (2002). The Otx family. Curr. Opin. Genet. Dev. 12, 409415.Google Scholar
Simionato, E., Kerner, P., Dray, N., Le Gouar, M., Ledent, V., Arendt, D., and Vervoort, M. (2008). atonal- and achaete-scute-related genes in the annelid Platynereis dumerilii: insights into the evolution of neural basic-Helix-Loop-Helix genes. BMC Evol. Biol. 8, Article 170.Google Scholar
Simmonds, A.J., Brook, W.J., Cohen, S.M., and Bell, J.B. (1995). Distinguishable functions for engrailed and invected in anterior-posterior patterning in the Drosophila wing. Nature 376, 424427.Google Scholar
Simões-Costa, M.S., Vasconcelos, M., Sampaio, A.C., Cravo, R.M., Linhares, V.L., Hochgreb, T., Yan, C.Y.I., Davidson, B., and Xavier-Neto, J. (2005). The evolutionary origin of cardiac chambers. Dev. Biol. 277, 115.Google Scholar
Simon, A., and Tanaka, E.M. (2013). Limb regeneration. WIREs Dev. Biol. 2, 291300.Google Scholar
Simonnet, F., Célérier, M.-L., and Quéinnec, E. (2006). Orthodenticle and empty spiracles genes are expressed in a segmental pattern in chelicerates. Dev. Genes Evol. 216, 467480.Google Scholar
Simpson, J.H., Bland, K.S., Fetter, R.D., and Goodman, C.S. (2000). Short-range and long-range guidance by Slit and its Robo receptors: a combinatorial code of Robo receptors controls lateral position. Cell 103, 10191032.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
Singh, A., Tare, M., Puli, O.R., and Kango-Singh, M. (2011). A glimpse into dorso-ventral patterning of the Drosophila eye. Dev. Dyn. 241, 6984.Google Scholar
Singh, D., and Pohl, C. (2014). Coupling of rotational cortical flow, asymmetric midbody positioning, and spindle rotation mediates dorsoventral axis formaiton in C. elegans. Dev. Cell 28, 253267.Google Scholar
Singh, N.P., and Mishra, R.K. (2008). A double-edged sword to force posterior dominance of Hox genes. BioEssays 30, 10581061.CrossRefGoogle ScholarPubMed
Singh, S., Stellrecht, C.M., Tang, H.K., and Saunders, G.F. (2000). Modulation of PAX6 homeodomain function by the paired domain. J. Biol. Chem. 275, #23, 1730617313.Google Scholar
Singla, V., and Reiter, J.F. (2006). The primary cilium as the cell's antenna: signaling at a sensory organelle. Science 313, 629633.Google Scholar
Siniscalchi, M., Lusito, R., Vallortigara, G., and Quaranta, A. (2013). Seeing left- or right-asymmetric tail wagging produces different emotional responses in dogs. Curr. Biol. 23, 22792282.Google Scholar
Sitaraman, D., Aso, Y., Jin, X., Chen, N., Felix, M., Rubin, G.M., and Nitabach, M.N. (2015). Propagation of homeostatic sleep signals by segregated synaptic microcircuits of the Drosophila mushroom body. Curr. Biol. 25, 29152927.Google Scholar
Sivanantharajah, L., and Percival-Smith, A. (2015). Differential pleiotropy and HOX functional organization. Dev. Biol. 398, 110.Google Scholar
Skeldon, K.D., Reid, L.M., McInally, V., Dougan, B., and Fulton, C. (1998). Physics of the Theremin. Am. J. Phys. 66, 945955.Google Scholar
Skoyles, J.R. (2006). Human balance, the evolution of bipedalism and dysequilibrium syndrome. Med. Hypotheses 66, 10601068.Google Scholar
Slack, J.M.W., Holland, P.W.H., and Graham, C.F. (1993). The zootype and the phylotypic stage. Nature 361, 490492.Google Scholar
Smallwood, P.M., Ölveczky, B.P., Williams, G.L., Jacobs, G.H., Reese, B.E., Meister, M., and Nathans, J. (2003). Genetically engineered mice with an additional class of cone receptors: implications for the evolution of color vision. PNAS 100, #20, 1170611711.Google Scholar
Smallwood, P.M., Wang, Y., and Nathans, J. (2002). Role of a locus control region in the mutually exclusive expression of human red and green cone pigment genes. PNAS 99, #2, 10081011.Google Scholar
Smear, M., Resulaj, A., Zhang, J., Bozza, T., and Rinberg, D. (2013). Multiple perceptible signals from a single olfactory glomerulus. Nat. Neurosci. 16, 16871691.CrossRefGoogle ScholarPubMed
Smetacek, V. (1992). Mirror-script and left-handedness. Nature 355, 118119.Google Scholar
Smith, A.B. (2008). Deuterostomes in a twist: the origins of a radical new body plan. Evol. Dev. 10, 493503.Google Scholar
Smith, A.T. (2015). Binocular vision: joining up the eyes. Curr. Biol. 25, R661R663.Google Scholar
Smith, F.W., Boothby, T.C., Giovannini, I., Rebecchi, L., Jockusch, E.L., and Goldstein, B. (2016). The compact body plan of tardigrades evolved by the loss of a large body region. Curr. Biol. 26, 224229.Google Scholar
Smyth, V.A., Di Lorenzo, D., and Kennedy, B.N. (2008). A novel, evolutionarily conserved enhancer of cone photoreceptor-specific expression. J. Biol. Chem. 283, #16, 1088110891.Google Scholar
Soba, P., Zhu, S., Emoto, K., Younger, S., Yang, S.-J., Yu, H.-H., Lee, T., Jan, L.Y., and Jan, Y.-N. (2007). Drosophila sensory neurons require Dscam for dendritic self-avoidance and proper dendritic field organization. Neuron 54, 403416.Google Scholar
Sohaskey, M.L., Yu, J., Diaz, M.A., Plaas, A.H., and Harland, R.M. (2008). JAWS coordinates chondrogenesis and synovial joint positioning. Development 135, 22152220.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. (2005). Conserved patterns of cell movements during vertebrate gastrulation. Curr. Biol. 15, R213R228.Google Scholar
Solomon, S.G., and Lennie, P. (2007). The machinery of colour vision. Nat. Rev. Neuro. 8, 276286.Google Scholar
Solovei, I., Kreysing, M., Lanctôt, C., Kösem, S., Peichi, L., Cremer, T., Guck, J., and Joffe, B. (2009). Nuclear architecture of rod photoreceptor cells adapts to vision in mammalian evolution. Cell 137, 356368.Google Scholar
Somel, M., Liu, X., and Khaitovich, P. (2013). Human brain evolution: transcripts, metabolites and their regulators. Nat. Rev. Neurosci. 14, 112127.Google Scholar
Sommer, A., and Vyas, K.S. (2012). A global clinical view on vitamin A and carotenoids. Am. J. Clin. Nutr. 96 (Suppl.), 1204S1206S.Google Scholar
Sommer, R., and Tautz, D. (1991). Segmentation gene expression in the housefly Musca domestica. Development 113, 419430.Google Scholar
Song, E., de Bivort, B., Dan, C., and Kunes, S. (2012). Determinants of the Drosophila odorant receptor pattern. Dev. Cell 22, 363376.Google Scholar
Song, H., Hu, J., Chen, W., Elliott, G., Andre, P., Gao, B., and Yang, Y. (2010). Planar cell polarity breaks bilateral symmetry by controlling ciliary positioning. Nature 466, 378382.Google Scholar
Song, K., Nam, Y.-J., Luo, X., Qi, X., Tan, W., Huang, G.N., Acharya, A., Smith, C.L., Tallquist, M.D., Neilson, E.G., Hill, J.A., Bassel-Duby, R., and Olson, E.N. (2012). Heart repair by reprogramming non-myocytes with cardiac transcription factors. Nature 485, 599604.Google Scholar
Song, Z., Postma, M., Billings, S.A., Coca, D., Hardie, R.C., and Juusola, M. (2012). Stochastic, adaptive sampling of information by microvilli in fly photoreceptors. Curr. Biol. 22, 13711380.Google Scholar
Sopko, R., Lin, Y.B., Makhijani, K., Alexander, B., Perrimon, N., and Brückner, K. (2015). A systems-level interrogation identifies regulators of Drosophila blood cell number and survival. PLoS Genet. 11, #3, e1005056.Google Scholar
Sopko, R., and McNeill, H. (2009). The skinny on Fat: an enormous cadherin that regulates cell adhesion, tissue growth, and planar cell polarity. Curr. Opin. Cell Biol. 21, 717723.Google Scholar
Sorrentino, R.P., Gajewski, K.M., and Schulz, R.A. (2005). GATA factors in Drosophila heart and blood cell development. Semin. Cell Dev. Biol. 16, 107116.Google Scholar
Soshnikova, N., Dewaele, R., Janvier, P., Krumlauf, R., and Duboule, D. (2013). Duplications of hox gene clusters and the emergence of vertebrates. Dev. Biol. 378, 194199.Google Scholar
Sotomayor, M., Weihofen, W.A., Gaudet, R., and Corey, D.P. (2012). Structure of a force-conveying cadherin bond essential for inner-ear mechanotransduction. Nature 492, 128132.Google Scholar
Soukup, V., and Kozmik, Z. (2016). Zoology: a new mouth for amphioxus. Curr. Biol. 26, R367R368.Google Scholar
Soukup, V., Yong, L.W., Lu, T.-M., Huang, S.-W., Kozmik, Z., and Yu, J.-K. (2015). The Nodal signaling pathway controls left-right asymmetric development in amphioxus. EvoDevo 6, Article 5.Google Scholar
Soustelle, L., Trousse, F., Jacques, C., Ceron, J., Cochard, P., Soula, C., and Giangrande, A. (2007). Neurogenic role of Gcm transcription factors is conserved in chicken spinal cord. Development 134, 625634.Google Scholar
Southwell, D.G., Paredes, M.F., Galvao, R.P., Jones, D.L., Froemke, R.C., Sebe, J.Y., Alfaro-Cervello, C., Tang, Y., Garcia-Verdugo, J.M., Rubenstein, J.L., Baraban, S.C., and Alvarez-Buylla, A. (2012). Intrinsically determined cell death of developing cortical interneurons. Nature 491, 109113.Google Scholar
Spanagel, R. (2009). Alcoholism: a systems approach from molecular physiology to addictive behavior. Physiol. Rev. 89, 649705.Google Scholar
Spéder, P., Ádám, G., and Noselli, S. (2006). Type ID unconventional myosin controls left-right asymmetry in Drosophila. Nature 440, 803807.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
Spehr, M., and Munger, S.D. (2009). Olfactory receptors: G protein-coupled receptors and beyond. J. Neurochem. 109, 15701583.Google Scholar
Spence, C. (2013). Multisensory flavour perception. Curr. Biol. 23, R365R370.Google Scholar
Sperry, R.W. (1963). Chemoaffinity in the orderly growth of nerve fiber patterns and connections. PNAS 50, 703710.Google Scholar
Spitz, F., Herkenne, C., Morris, M.A., and Duboule, D. (2005). Inversion-induced disruption of the Hoxd cluster leads to the partition of regulatory landscapes. Nature Genet. 37, 889893.Google Scholar
Spitzweck, B., Brankatschk, M., and Dickson, B.J. (2010). Distinct protein domains and expresion patterns confer divergent axon guidance functions for Drosophila Robo receptors. Cell 140, 409420.Google Scholar
Spletter, M.L., and Luo, L. (2009). A new family of odorant receptors in Drosophila. Cell 136, 2325.Google Scholar
Spoon, C., Moravec, W.J., Rowe, M.H., Grant, J.W., and Peterson, E.H. (2011). Steady-state stiffness of utricular hair cells depends on macular location and hair bundle structure. J. Neurophysiol. 106, 29502963.Google Scholar
Spoon, J.M. (2001). Situs inversus totalis. Neonatal Netw. 20, 5963.Google Scholar
Sprecher, S.G., and Desplan, C. (2008). Switch of rhodopsin expression in terminally differentiated Drosophila sensory neurons. Nature 454, 533537.Google Scholar
Sprecher, S.G., and Reichert, H. (2003). The urbilaterian brain: developmental insights into the evolutionary origin of the brain in insects and vertebrates. Arthropod Struct. Dev. 32, 141156.Google Scholar
Sprecher, S.G., Reichert, H., and Hartenstein, V. (2007). Gene expression patterns in primary neuronal clusters of the Drosophila embryonic brain. Gene Expr. Patterns 7, 584595.Google Scholar
Sproul, D., Gilbert, N., and Bickmore, W.A. (2005). The role of chromatin structure in regulating the expression of clustered genes. Nat. Rev. Genet. 6, 775781.Google Scholar
Spudich, J.L., Sineshchekov, O.A., and Govorunova, E.G. (2014). Mechanism divergence in microbial rhodopsins. Biochim. Biophys. Acta 1837, 546552.Google Scholar
Spudich, J.L., Yang, C.-S., Jung, K.-H., and Spudich, E.N. (2000). Retinylidene proteins: structures and functions from archaea to humans. Annu. Rev. Cell Dev. Biol. 16, 365392.Google Scholar
Srinivasan, M., Zhang, S., and Reinhard, J. (2006). Small brains, smart minds: vision, perception, navigation, and “cognition” in insects. In Warrant, E. and Nilsson, D.-E. (eds.), Invertebrate Vision. Cambridge University Press, New York, NY, pp. 462493.Google Scholar
Srivastava, D. (2006). Making or breaking the heart: from lineage determination to morphogenesis. Cell 126, 10371048.Google Scholar
Srour, M., Rivière, J.-B., Pham, J.M.T., Dubé, M.-P., Girard, S., Morin, S., Dion, P.A., Asselin, G., Rochefort, D., Hince, P., Diab, S., Sharafaddinzadeh, N., Chouinard, S., Théoret, H., Charron, F., and Rouleau, G.A. (2010). Mutations in DCC cause congenital mirror movements. Science 328, 592593.Google Scholar
Stanewsky, R. (2003). Genetic analysis of the circadian system in Drosophila melanogaster and mammals. J. Neurobiol. 54, 111147.Google Scholar
Stark, W.S., Wagner, R.H., and Gillespie, C.M. (1994). Ultraviolet sensitivity of three cone types in the aphakic observer determined by chromatic adaptation. Vision Res. 34, 14571459.Google Scholar
Stauber, M., Jäckle, H., and Schmidt-Ott, U. (1999). The anterior determinant bicoid of Drosophila is a derived Hox class 3 gene. PNAS 96, 37863789.Google Scholar
Stauber, M., Prell, A., and Schmidt-Ott, U. (2002). A single Hox3 gene with composite bicoid and zerknüllt expression characteristics in non-Cyclorrhaphan flies. PNAS 99, #1, 274279.Google Scholar
Stavenga, D.G. (2002). Colour in the eyes of insects. J. Comp. Physiol. A 188, 337348.Google Scholar
Stavenga, D.G., and Hardie, R.C. (2011). Metarhodopsin control by arrestin, light-filtering screening pigments, and visual pigment turnover in invertebrate microvillar photoreceptors. J. Comp. Physiol. A 197, 227241.Google Scholar
Steele, R.E., David, C.N., and Technau, U. (2010). A genomic view of 500 million years of cnidarian evolution. Trends Genet. 27, 713.Google Scholar
Stein, E., and Tessier-Lavigne, M. (2001). Hierarchical organization of guidance receptors: silencing of Netrin attraction by Slit through a Robo/DCC receptor complex. Science 291, 19281938.Google Scholar
Steinmetz, P.R.H., Kostyuchenko, R.P., Fischer, A., and Arendt, D. (2011). The segmental pattern of otx, gbx, and Hox genes in the annelid Platynereis dumerilii. Evol. Dev. 13, 7279.Google Scholar
Steinmetz, P.R.H., Urbach, R., Posnien, N., Eriksson, J., Kostyuchenko, R.P., Brena, C., Guy, K., Akam, M., Bucher, G., and Arendt, D. (2010). Six3 demarcates the anterior-most developing brain region in bilaterian animals. EvoDevo 1, Article 14.Google Scholar
Stengl, M., and Funk, N.W. (2013). The role of the coreceptor Orco in insect olfactory transduction. J. Comp. Physiol. A 199, 897909.Google Scholar
Stennard, F.A., and Harvey, R.P. (2005). T-box transcription factors and their roles in regulatory hierarchies in the developing heart. Development 132, 48974910.Google Scholar
Stephen, L.A., Johnson, E.J., Davis, G.M., McTeir, L., Pinkham, J., Jaberi, N., and Davey, M.G. (2014). The chicken left right organizer has nonmotile cilia which are lost in a stage-dependent manner in the talpid3 ciliopathy. Genesis 52, 600613.Google Scholar
Stephenson-Jones, M., Samuelsson, E., Ericsson, J., Robertson, B., and Grillner, S. (2011). Evolutionary conservation of the basal ganglia as a common vertebrate mechanism for action selection. Curr. Biol. 21, 10811091.Google Scholar
Stergiopoulos, A., Elkouris, M., and Politis, P.K. (2015). Prospero-related homeobox 1 (Prox1) at the crossroads of diverse pathways during adult neural fate specification. Front. Cell. Neurosci. 8, Article 454.Google Scholar
Stern, C. (1941). The growth of testes in Drosophila. I. The relation between vas deferens and testis within various species. J. Exp. Zool. 87, 113158.Google Scholar
Stern, C., and Tokunaga, C. (1968). Autonomous pleiotropy in Drosophila. PNAS 60, 12521259.Google Scholar
Stern, C.D. (1990). Two distinct mechanisms for segmentation? Semin. Dev. Biol. 1, 109116.Google Scholar
Stern, C.D., and Piatkowska, A.M. (2015). Multiple roles of timing in somite formation. Semin. Cell Dev. Biol. 42, 134139.Google Scholar
Stern, D.L. (2013). The genetic causes of convergent evolution. Nat. Rev. Genet. 14, 751764.Google Scholar
Stern, K., and McClintock, M.K. (1998). Regulation of ovulation by human pheromones. Nature 392, 177179.Google Scholar
Steward, R. (1989). Relocalization of the dorsal protein from the cytoplasm to the nucleus correlates with its function. Cell 59, 11791188.Google Scholar
Stocker, R.F. (2001). Drosophila as a focus in olfactory research: mapping of olfactory sensilla by fine structure, odor specificity, odorant receptor expression, and central connectivity. Microsc. Res. Tech. 55, 284296.Google Scholar
Stocum, D.L., and Cameron, J.A. (2011). Looking proximally and distally: 100 years of limb regeneration and beyond. Dev. Dyn. 240, 943968.Google Scholar
Stoick-Cooper, C.L., Moon, R.T., and Weidinger, G. (2007). Advances in signaling in vertebrate regeneration as a prelude to regenerative medicine. Genes Dev. 21, 12921315.Google Scholar
Stokes, M.D., and Holland, N. (1998). The lancelet. Am. Sci. 86, #6, 552560.Google Scholar
Stollewerk, A. (2008). Evolution of neurogenesis in arthropods. In Minelli, A. and Fusco, G. (eds.), Evolving Pathways: Key Themes in Evolutionary Developmental Biology. Cambridge University Press, New York, NY, pp. 359380.Google Scholar
Stollewerk, A., and Simpson, P. (2005). Evolution of early development of the nervous system: a comparison between arthropods. BioEssays 27, 874883.Google Scholar
Störtkuhl, K.F., and Fiala, A. (2011). The smell of blue light: a new approach toward understanding an olfactory neuronal network. Front. Neurosci. 5, Article 72.Google Scholar
Strausfeld, N.J. (2005). The evolution of crustacean and insect optic lobes and the origins of chiasmata. Arthropod Struct. Dev. 34, 235256.Google Scholar
Strausfeld, N.J. (2009). Brain organization and the origin of insects: an assessment. Proc. R. Soc. Lond. B 276, 19291937.Google Scholar
Strausfeld, N.J., and Hildebrand, J.G. (1999). Olfactory systems: common design, uncommon origins? Curr. Opin. Neurobiol. 9, 634639.Google Scholar
Strausfeld, N.J., and Hirth, F. (2013). Deep homology of arthropod central complex and vertebrate basal ganglia. Science 340, 157161. [See alsoGoogle ScholarGoogle Scholar
Strausfeld, N.J., and Hirth, F. (2013). Homology versus convergence in resolving transphyletic correspondences of brain organization. Brain Behav. Evol. 82, 215219.Google Scholar
Streelman, J.T. (2014). Advancing evolutionary developmental biology. In Advances in Evolutionary Developmental Biology. Wiley-Blackwell, New York, NY, pp. 203217.Google Scholar
Strigini, M. (2005). Mechanisms of morphogen transport. J. Neurobiol. 64, 324333.Google Scholar
Strilic, B., Kucera, T., and Lammert, E. (2010). Formation of cardiovascular tubes in invertebrates and vertebrates. Cell. Mol. Life Sci. 67, 32093218.Google Scholar
Struhl, G., Struhl, K., and Macdonald, P.M. (1989). The gradient morphogen bicoid is a concentration-dependent transcriptional activator. Cell 57, 12591273.Google Scholar
Strutt, H., Warrington, S.J., and Strutt, D. (2011). Dynamics of core planar polarity protein turnover and stable assembly into discrete membrane subdomains. Dev. Cell 20, 511525.Google Scholar
Sturtevant, A.H. (1965). A History of Genetics. Harper & Row, New York, NY.Google Scholar
Sturtevant, A.H. (1970). Studies on the bristle pattern of Drosophila. Dev. Biol. 21, 4861.Google Scholar
Su, C.-Y., Menuz, K., and Carlson, J.R. (2009). Olfactory perception: receptors, cells, and circuits. Cell 139, 4559.Google Scholar
Suárez, R., García-González, D., and de Castro, F. (2012). Mutual influences between the main olfactory and vomeronasal systems in development and evolution. Front. Neuroanat. 6, Article 50.Google Scholar
Suga, H., Schmid, V., and Gehring, W.J. (2008). Evolution and functional diversity of jellyfish opsins. Curr. Biol. 18, 5155.Google Scholar
Sun, T., Patoine, C., Abu-Khalil, A., Visvader, J., Sum, E., Cherry, T.J., Orkin, S.H., Geschwind, D.H., and Walsh, C.A. (2005). Early asymmetry of gene transcription in embryonic human left and right cerebral cortex. Science 308, 17941798.Google Scholar
Sun, T., and Walsh, C.A. (2006). Molecular approaches to brain asymmetry and handedness. Nat. Rev. Neurosci. 7, 655662.Google Scholar
Sun, Y., Kanekar, S.L., Vetter, M.L., Gorski, S., Jan, Y.-N., Glaser, T., and Brown, N.L. (2003). Conserved and divergent functions of Drosophila atonal, amphibian, and mammalian Ath5 genes. Evol. Dev. 5, 532541.Google Scholar
Sung, C.-H., and Chuang, J.-Z. (2010). The cell biology of vision. J. Cell Biol. 190, 953963.Google Scholar
Supp, D.M., Witte, D.P., Potter, S.S., and Brueckner, M. (1997). Mutation of an axonemal dynein affects left-right asymmetry in inversus viscerum mice. Nature 389, 963966.Google Scholar
Sutcliffe, B., Forero, M.G., Zhu, B., Robinson, I.M., and Hidalgo, A. (2013). Neuron-type specific functions of DNT1, DNT2 and Spz at the Drosophila neuromuscular junction. PLoS ONE 8, #10, e75902.Google Scholar
Suzanne, M., Petzoldt, A.G., Spéder, P., Coutelis, J.-B., Steller, H., and Noselli, S. (2010). Coupling of apoptosis and L/R patterning controls stepwise organ looping. Curr. Biol. 20, 17731778.Google Scholar
Suzuki, D.G., Murakami, Y., Escriva, H., and Wada, H. (2015). A comparative examination of neural circuit and brain patterning between the lamprey and amphioxus reveals the evolutionary origin of the vertebrate visual center. J. Comp. Neurol. 523, 251261.Google Scholar
Suzuki, T. (2013). How is digit identity determined during limb development? Dev. Growth Differ. 55, 130138.Google Scholar
Suzuki, Y., and Palopoli, M.F. (2001). Evolution of insect abdominal appendages: are prolegs homologous or convergent traits? Dev. Genes Evol. 211, 486492.Google Scholar
Sweeney, E., Fryer, A., Mountford, R., Green, A., and McIntosh, I. (2003). Nail patella syndrome: a review of the phenotype aided by developmental biology. J. Med. Genet. 40, 153162.Google Scholar
Tabin, C., and Wolpert, L. (2007). Rethinking the proximodistal axis of the vertebrate limb in the molecular era. Genes Dev. 21, 14331442.Google Scholar
Tabin, C.J., Carroll, S.B., and Panganiban, G. (1999). Out on a limb: parallels in vertebrate and invertebrate limb patterning and the origin of appendages. Am. Zool. 39, 650663.Google Scholar
Tabin, C.J., and Vogan, K.J. (2003). A two-cilia model for vertebrate left-right axis specification. Genes Dev. 17, 16.Google Scholar
Tabuchi, K., and Südhof, T.C. (2002). Structure and evolution of neurexin genes: insight into the mechanism of alternative splicing. Genomics 79, 849859.Google Scholar
Tahayato, A., Sonneville, R., Pichaud, F., Wernet, M.F., Papatsenko, D., Beaufils, P., Cook, T., and Desplan, C. (2003). Otd/Crx, a dual regulator for the specification of ommatidia subtypes in the Drosophila retina. Dev. Cell 5, 391402.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, M., and Osumi, N. (2002). Pax6 regulates specification of ventral neurone subtypes in the hindbrain by establishing progenitor domains. Development 129, 13271338.Google Scholar
Takashima, S., Gold, D., and Hartenstein, V. (2013). Stem cells and lineages of the intestine: a developmental and evolutionary perspective. Dev. Genes Evol. 223, 85102.Google Scholar
Takashima, S., Yoshimori, H., Yamasaki, N., Matsuno, K., and Murakami, R. (2002). Cell-fate choice and boundary formation by combined action of Notch and engrailed in the Drosophila hindgut. Dev. Genes Evol. 212, 534541.Google Scholar
Takeda, S., and Narita, K. (2012). Structure and function of vertebrate cilia, towards a new taxonomy. Differentiation 83, S4S11.Google Scholar
Talpalar, A.E., Bouvier, J., Borgius, L., Fortin, G., Pierani, A., and Kiehn, O. (2013). Dual-mode operation of neuronal networks involved in left-right alternation. Nature 500, 8588.Google Scholar
Tam, P.P.L., Loebel, D.A.F., and Tanaka, S.S. (2006). Building the mouse gastrula: signals, asymmetry and lineages. Curr. Opin. Genet. Dev. 16, 419425.Google Scholar
Tamakoshi, T., Itakura, T., Chandra, A., Uezato, T., Yang, Z., Xue, X.-D., Wang, B., Hackett, B.P., Yokoyama, T., and Miura, N. (2006). Roles of the Foxj1 and Inv genes in the left-right determination of internal organs in mice. Biochem. Biophys. Res. Comm. 339, 932938.Google Scholar
Tan, X., Pecka, J.L., Tang, J., Okoruwa, O.E., Zhang, Q., Beisel, K.W., and He, D.Z.Z. (2011). From zebrafish to mammal: functional evolution of prestin, the motor protein of cochlear outer hair cells. J. Neurophysiol. 105, 3644.Google Scholar
Tanaka, E.M. (2003). Regeneration: if they can do it, why can't we? Cell 113, 559562. [See alsoGoogle ScholarGoogle Scholar
Tanaka, E.M. (2012). Skin, heal thyself. Nature 489, 508510.Google Scholar
Tanaka, E.M., and Reddien, P.W. (2011). The cellular basis for animal regeneration. Dev. Cell 21, 172185.Google Scholar
Tanaka, M. (2016). Fins into limbs: autopod acquisition and anterior elements reduction by modifying gene networks involving 5Hox, Gli3, and Shh. Dev. Biol. 413, 17.Google Scholar
Tanaka, M., Kasahara, H., Bartunkova, S., Schinke, M., Komuro, I., Inagaki, H., Lee, Y., Lyons, G.E., and Izumo, S. (1998). Vertebrate homologs of tinman and bagpipe: roles of the homeobox genes in cardiovascular development. Dev. Genet. 22, 239249.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
Tang, M., Yuan, W., Bodmer, R., Wu, X., and Ocorr, K. (2013). The role of pygopus in the differentiation of intracardiac valves in Drosophila. Genesis 52, 1928.Google Scholar
Tang, W.J., Fernandez, J.G., Sohn, J.J., and Amemiya, C.T. (2015). Chitin is endogenously produced in vertebrates. Curr. Biol. 25, 897900.Google Scholar
Taniguchi, K., Hozumi, S., Maeda, R., Ooike, M., Sasamura, T., Aigaki, T., and Matsuno, K. (2007). D-JNK signaling in visceral muscle cells controls the laterality of the Drosophila gut. Dev. Biol. 311, 251263.Google Scholar
Tao, Y., and Schulz, R.A. (2007). Heart development in Drosophila. Semin. Cell Dev. Biol. 18, 315.Google Scholar
Tarazona, O.A., Slota, L.A., Lopez, D.H., Zhang, G., and Cohn, M.J. (2016). The genetic program for cartilage development has deep homology within Bilateria. Nature 533, 8689. [See alsoGoogle ScholarGoogle Scholar
Tarchini, B., Jolicoeur, C., and Cayouette, M. (2013). A molecular blueprint at the apical surface establishes planar asymmetry in cochlear hair cells. Dev. Cell 27, 88102.Google Scholar
Taschner, M., Bhogaraju, S., and Lorentzen, E. (2012). Architecture and function of IFT complex proteins in ciliogenesis. Differentiation 83, S12S22.Google Scholar
Tateya, T., Sakamoto, S., Imayoshi, I., and Kageyama, R. (2015). In vivo overactivation of the Notch signaling pathway in the developing cochlear epithelium. Hear. Res. 327, 209217.Google Scholar
Tautz, D. (2004). Segmentation. Dev. Cell 7, 301312.Google Scholar
Tautz, D., and Sommer, R.J. (1995). Evolution of segmentation genes in insects. Trends Genet. 11, 2327.Google Scholar
Taverna, E., Götz, M., and Huttner, W.B. (2015). The cell biology of neurogenesis: toward an understanding of the development and evolution of the neocortex. Annu. Rev. Cell Dev. Biol. 30, 465502.Google Scholar
Taylor, J.S., and Raes, J. (2004). Duplication and divergence: the evolution of new genes and old ideas. Annu. Rev. Genet. 38, 615643.Google Scholar
Technau, G.M., Berger, C., and Urbach, R. (2006). Generation of cell diversity and segmental pattern in the embryonic central nervous system of Drosophila. Dev. Dyn. 235, 861869.Google Scholar
Tee, Y.H., Shemesh, T., Thiagarajan, V., Hariadi, R.F., Anderson, K.L., Page, C., Volkmann, N., Hanein, D., Sivaramakrishnan, S., Kozlov, M.M., and Bershadsky, A.D. (2015). Cellular chirality arising from the self-organization of the actin cytoskeleton. Nat. Cell Biol. 17, #4, 445457.Google Scholar
Teixeira, C.S.S., Cerqueira, N.M.F.S.A., and Ferreira, A.C.S. (2016). Unravelling the olfactory sense: from the gene to odor perception. Chem. Senses 41, 105121.Google Scholar
Telford, M.J. (2000). Evidence for the derivation of the Drosophila fushi tarazu gene from a Hox gene orthologous to lophotrochozoan Lox5. Curr. Biol. 10, 349352.Google Scholar
Telford, M.J. (2007). A single origin of the central nervous system? Cell 129, 237239.Google Scholar
Telford, M.J., and Budd, G.E. (2003). The place of phylogeny and cladistics in Evo-Devo research. Int. J. Dev. Biol. 47, 479490.Google Scholar
Telford, M.J., Budd, G.E., and Philippe, H. (2015). Phylogenomic insights into animal evolution. Curr. Biol. 25, R876R887.Google Scholar
Temple, S.E. (2011). Why different regions of the retina have different spectral sensitivities: a review of mechanisms and functional significance of intraretinal variability in spectral sensitivity in vertebrates. Vis. Neurosci. 28, 281293.Google Scholar
Terakita, A. (2005). The opsins. Genome Biol. 6, Article 213.Google Scholar
Terrell, D., Xie, B., Workman, M., Mahato, S., Zelhof, A.C., Gebelein, B., and Cook, T. (2012). OTX2 and CRX rescue overlapping and photoreceptor-specific functions in the Drosophila eye. Dev. Dyn. 241, 215228.Google Scholar
Tessmar-Raible, K., Raible, F., Christodoulou, F., Guy, K., Rembold, M., Hausen, H., and Arendt, D. (2007). Conserved sensory-neurosecretory cell types in annelid and fish forebrain: insights into hypothalamus evolution. Cell 129, 13891400.Google Scholar
Tettamanti, G., Cattaneo, A.G., Gornati, R., de Eguileor, M., Bernardini, G., and Binelli, G. (2010). Phylogenesis of brain-derived neurotrophic factor (BDNF) in vertebrates. Gene 450, 8593.Google Scholar
Tettamanti, G., Grimaldi, A., Valvassori, R., Rinaldi, R., and de Eguileor, M. (2003). Vascular endothelial growth factor is involved in neoangiogenesis in Hirudo medicinalis (Annelida, Hirudinea). Cytokine 22, 168179.Google Scholar
Thanawala, S.U., Rister, J., Goldberg, G.W., Zuskov, A., Olesnicky, E.C., Flowers, J.M., Jukam, D., Purugganan, M.D., Gavis, E.R., Desplan, C., and Johnston, R.J. Jr. (2013). Regional modulation of a stochastically expressed factor determines photoreceptor subtypes in the Drosophila retina. Dev. Cell 25, 93105.Google Scholar
Tharadra, S.K., Medina, A., and Ray, A. (2013). Advantage of the highly restricted odorant receptor expression pattern in chemosensory neurons of Drosophila. PLoS ONE 8, #6, e66173.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
Thoen, H.H., How, M.J., Chiou, T.-H., and Marshall, J. (2014). A different form of color vision in mantis shrimp. Science 343, 411413.Google Scholar
Thomas, A.L., Davis, S.M., and Dierick, H.A. (2015). Of fighting flies, mice, and men: are some of the molecular and neuronal mechanisms of aggression universal in the animal kingdom? PLoS Genet. 11, #8, e1005416.Google Scholar
Thompson, B.J. (2013). Cell polarity: models and mechanisms from yeast, worms, and flies. Development 140, 1321.Google Scholar
Thompson, D., Regev, A., and Roy, S. (2015). Comparative analysis of gene regulatory networks: from network reconstruction to evolution. Annu. Rev. Cell Dev. Biol. 31, 399428.Google Scholar
Thompson, H., Shaw, M.K., Dawe, H.R., and Shimeld, S.M. (2012). The formation and positioning of cilia in Ciona intestinalis embryos in relation to the generation and evolution of chordate left-right asymmetry. Dev. Biol. 364, 214223.Google Scholar
Thor, S. (2013). Stem cells in multiple time zones. Nature 498, 441443.Google Scholar
Thor, S., and Thomas, J.B. (2002). Motor neuron specification in worms, flies and mice: conserved and “lost” mechanisms. Curr. Opin. Genet. Dev. 12, 558564.Google Scholar
Tickle, C., and Barker, H. (2013). The Sonic hedgehog gradient in the developing limb. WIREs Dev. Biol. 2, 275290.Google Scholar
Tilney, L.G., Connelly, P., Smith, S., and Guild, G.M. (1996). F-Actin bundles in Drosophila bristles are assembled from modules composed of short filaments. J. Cell Biol. 135, 12911308.Google Scholar
Tilney, L.G., Cotanche, D.A., and Tilney, M.S. (1992). Actin filaments, stereocilia and hair cells of the bird cochlea. VI. How the number and arrangement of stereocilia are determined. Development 116, 213226.Google Scholar
Tilney, L.G., and DeRosier, D.J. (2005). How to make a curved Drosophila bristle using straight actin bundles. PNAS 102, #52, 1878518792.Google Scholar
Tilney, L.G., Tilney, M.S., and Cotanche, D.A. (1988). Actin filaments, stereocilia, and hair cells of the bird cochlea. V. How the staircase pattern of stereociliary lengths is generated. J. Cell Biol. 106, 355365.Google Scholar
Tilney, L.G., Tilney, M.S., and DeRosier, D.J. (1992). Actin filaments, stereocilia, and hair cells: how cells count and measure. Annu. Rev. Cell Biol. 8, 257274.Google Scholar
Tilney, L.G., Tilney, M.S., and Guild, G.M. (1995). F-Actin bundles in Drosophila bristles. I. Two filament cross-links are involved in bundling. J. Cell Biol. 130, 629638.Google Scholar
Tingler, M., Ott, T., Tözser, J., Kurz, S., Getwan, M., Tisler, M., Schweickert, A., and Blum, M. (2014). Symmetry breakage in the frog Xenopus: role of Rab11 and the ventral-right blastomere. Genesis 52, 588599.Google Scholar
Tiozzo, S., Christiaen, L., Deyts, C., Manni, L., Joly, J.-S., and Burighel, P. (2005). Embryonic versus blastogenetic development in the compound ascidian Botryllus schlosseri: insights from Pitx expression patterns. Dev. Dyn. 232, 468478.Google Scholar
Tissir, F., and Goffinet, A.M. (2010). Planar cell polarity signaling in neural development. Curr. Opin. Neurobiol. 20, 572577.Google Scholar
Tittel, J.N., and Steller, H. (2000). A comparison of programmed cell death between species. Genome Biol. 1, #3, 16.Google Scholar
Todi, S.V., Franke, J.D., Kiehart, D.P., and Eberl, D.F. (2005). Myosin VIIA defects, which underlie the Usher 1B syndrome in humans, lead to deafness in Drosophila. Curr. Biol. 15, 862868. [See alsoGoogle ScholarGoogle Scholar
Todi, S.V., Sivan-Loukianova, E., Jacobs, J.S., Kiehart, D.P., and Eberl, D.F. (2008). Myosin VIIA, important for human auditory function, is necessary for Drosophila auditory organ development. PLoS ONE 3, #5, e2115.Google Scholar
Tomarev, S.I., Callaerts, P., Kos, L., Zinovieva, R., Halder, G., Gehring, W., and Piatigorsky, J. (1997). Squid Pax-6 and eye development. PNAS 94, 24212426.Google Scholar
Tomer, R., Denes, A.S., Tessmar-Raible, K., and Arendt, D. (2010). Profiling by image registration reveals common origin of annelid mushroom bodies and vertebrate pallium. Cell 142, 800809.Google Scholar
Torgersen, J. (1947). Transposition of viscera, bronchiectasis and nasal polyps. Acta Radiol. 28, 1724.Google Scholar
Torres, M. (2016). Limb regrowth takes two. Nature 533, 328330.Google Scholar
Torres, M., Gómez-Pardo, E., and Gruss, P. (1996). Pax2 contributes to inner ear patterning and optic nerve trajectory. Development 122, 33813391.Google Scholar
Tosches, M.A., and Arendt, D. (2013). The bilaterian forebrain: an evolutionary chimaera. Curr. Opin. Neurobiol. 23, 10801089.Google Scholar
Touhara, K., and Vosshall, L.B. (2009). Sensing odorants and pheromones with chemosensory receptors. Annu. Rev. Physiol. 71, 307332.Google Scholar
Towbin, B.D., Meister, P., and Gasser, S.M. (2009). The nuclear envelope: a scaffold for silencing? Curr. Opin. Genet. Dev. 19, 180186.Google Scholar
Travis, J. (1995). The ghost of Geoffroy Saint-Hilaire: frog and fly genes revive the ridiculed idea that vertebrates resemble upside-down insects. Sci. News 148, 216218.Google Scholar
Treisman, J.E. (2004). Coming to our senses. BioEssays 26, 825828.Google Scholar
Treisman, J.E. (2004). How to make an eye. Development 131, 38233827.Google Scholar
Tromelin, A. (2016). Odour perception: a review of an intricate signalling pathway. Flavour Fragr. J. 31, 107119.Google Scholar
Troost, T., and Klein, T. (2012). Sequential Notch signalling at the boundary of Fringe expressing and non-expressing cells. PLoS ONE 7, #11, e49007.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
Trotier, D. (2011). Vomeronasal organ and human pheromones. Eur. Ann. Otorhinolaryngol. Head Neck Dis. 128, 184190.Google Scholar
Trujillo-Cenóz, O. (1985). The eye: development, structure and neural connections. In Kerkut, G.A. and Gilbert, L.I. (eds.), Comprehensive Insect Physiology, Biochemistry, and Pharmacology. Pergamon Press, Oxford, pp. 171223.Google Scholar
Tsachaki, M., and Sprecher, S.G. (2012). Genetic and developmental mechanisms underlying the formation of the Drosophila compound eye. Dev. Dyn. 241, 4056.Google Scholar
Tschopp, P., and Duboule, D. (2011). A genetic approach to the transcriptional regulation of Hox gene clusters. Annu. Rev. Genet. 45, 145166.Google Scholar
Tschopp, P., Tarchini, B., Spitz, F., Zakany, J., and Duboule, D. (2009). Uncoupling time and space in the collinear regulation of Hox genes. PLoS Genet. 5, #3, e1000398.Google Scholar
Tsigankov, D., and Koulakov, A.A. (2010). Sperry versus Hebb: topographic mapping in Isl2/EphA3 mutant mice. BMC Neurosci. 11, Article 155.Google Scholar
Tuthill, J.C., and Wilson, R.I. (2016). Parallel transformation of tactile signals in central circuits of Drosophila. Cell 164, 10461059.Google Scholar
Tweedt, S.M., and Erwin, D.H. (2015). Origin of metazoan developmental toolkits and their expression in the fossil record. In Ruiz-Trillo, I. and Nedelcu, A.M. (eds.), Evolutionary Transitions to Multicellular Life: Principles and Mechanisms. Springer, Netherlands, pp. 4777.Google Scholar
Uemura, T., and Shimada, Y. (2003). Breaking cellular symmetry along planar axes in Drosophila and vertebrates. J. Biochem. 134, 625630.Google Scholar
Ugur, B., Chen, K., and Bellen, H.J. (2016). Drosophila tools and assays for the study of human diseases. Dis. Model. Mech. 9, 235244.Google Scholar
Ullrich, B., Ushkaryov, Y.A., and Südhof, T.C. (1995). Cartography of neurexins: more than 1000 isoforms generated by alternative splicing and expressed in distinct subsets of neurons. Neuron 14, 497507.Google Scholar
Ullrich-Lüter, E.M., Dupont, S., Arboleda, E., Hausen, H., and Arnone, M.I. (2011). Unique system of photoreceptors in sea urchin tube feet. PNAS 108, #20, 83678372.Google Scholar
Umesono, Y., Watanabe, K., and Agata, K. (1999). Distinct structural domains in the planarian brain defined by the expression of evolutionarily conserved homeobox genes. Dev. Genes Evol. 209, 3139.Google Scholar
Underwood, E. (2015). Plugged pores may underlie some ALS, dementia cases. Science 349, 911912.Google Scholar
Urata, M., Tsuchimoto, J., Yasui, K., and Yamaguchi, M. (2009). The Hox8 of the hemichordate Balanoglossus misakiensis. Dev. Genes Evol. 219, 377382.Google Scholar
Urbach, R. (2007). A procephalic territory in Drosophila exhibiting similarities and dissimilarities compared to the vertebrate midbrain/hindbrain boundary region. Neural Dev. 2, Article 23.Google Scholar
Urbach, R., Volland, D., Seibert, J., and Technau, G.M. (2006). Segment-specific requirements for dorsoventral patterning genes during early brain development in Drosophila. Development 133, 43154330.Google Scholar
Usukura, J., and Obata, S. (1995). Morphogenesis of photoreceptor outer segments in retinal development. Prog. Retin. Eye Res. 15, 113125.Google Scholar
van Alphen, B., and van Swinderen, B. (2013). Drosophila strategies to study psychiatric disorders. Brain Res. Bull. 92, 111.Google Scholar
van Amerongen, R., and Nusse, R. (2009). Towards an integrated view of Wnt signaling in development. Development 136, 32053214.Google Scholar
Van Battum, E.Y., Brignani, S., and Pasterkamp, R.J. (2015). Axon guidance proteins in neurological disorders. Lancet 14, 532546.Google Scholar
Van de Peer, Y., Maere, S., and Meyer, A. (2009). The evolutionary significance of ancient genome duplications. Nat. Rev. Genet. 10, 725732.Google Scholar
van der Knaap, L.J., and van der Ham, J.M. (2011). How does the corpus callosum mediate interhemispheric transfer? A review. Behav. Brain Res. 223, 211221.Google Scholar
van der Linde, D., Konings, E.E.M., Slager, M.A., Witsenburg, M., Helbing, W.A., Takkenberg, J.J.M., and Roos-Hesselink, J.W. (2011). Birth prevalence of congenital heart disease worldwide: a systematic review and meta-analysis. J. Am. Coll. Cardiol. 58, #21, 22412247.Google Scholar
Van Essen, D.C. (1997). A tension-based theory of morphogenesis and compact wiring in the central nervous system. Nature 385, 313318.Google Scholar
van Hateren, J.H., Hardie, R.C., Rudolph, A., Laughlin, S.B., and Stavenga, D.G. (1989). The bright zone, a specialized dorsal eye region in the male blowfly Chrysomyia megacephala. J. Comp. Physiol. A 164, 297308. [See alsoGoogle ScholarGoogle Scholar
van Holde, K.E., Miller, K.I., and Decker, H. (2001). Hemocyanins and invertebrate evolution. J. Biol. Chem. 276, #19, 1556315566.Google Scholar
van Staaden, M.J., and Römer, H. (1998). Evolutionary transition from stretch to hearing organs in ancient grasshoppers. Nature 394, 773776.Google Scholar
Vandenberg, L.N., and Levin, M. (2010). Consistent left-right asymmetry cannot be established by late organizers in Xenopus unless the late organizer is a conjoined twin. Development 137, 10951105.Google Scholar
Vandenberg, L.N., and Levin, M. (2010). Far from solved: a perspective on what we know about early mechanisms of left-right asymmetry. Dev. Dyn. 239, 31313146.Google Scholar
Vandenberg, L.N., and Levin, M. (2013). A unified model for left-right asymmetry? Comparison and synthesis of molecular models of embryonic laterality. Dev. Biol. 379, 115.Google Scholar
Vasiliauskas, D., Mazzoni, E.O., Sprecher, S.G., Brodetskiy, K., Johnston, R.J. Jr., Lidder, P., Vogt, N., Celik, A., and Desplan, C. (2011). Feedback from rhodopsin controls rhodopsin exclusion in Drosophila photoreceptors. Nature 479, 108112.Google Scholar
Vasudevan, D., and Ryoo, H.D. (2015). Regulation of cell death by IAPs and their antagonists. Curr. Top. Dev. Biol. 114, 185208.Google Scholar
Vavouri, T., and Lehner, B. (2009). Conserved noncoding elements and the evolution of animal body plans. BioEssays 31, 727735.Google Scholar
Videnovic, A., Lazar, A.S., Barker, R.A., and Overeem, S. (2014). “The clocks that time us”: circadian rhythms in neurodegenerative disorders. Nat. Rev. Neurol. 10, 683693.Google Scholar
Villar, D., Flicek, P., and Odom, D.T. (2014). Evolution of transcription factor binding in metazoans: mechanisms and functional implications. Nat. Rev. Genet. 15, 221233.Google Scholar
Visel, A., Minovitsky, S., Dubchak, I., and Pennacchio, L.A. (2007). VISTA enhancer browser: a database of tissue-specific human enhancers. Nucleic Acids Res. 35, D88D92.Google Scholar
Vitruvius, (1960). The Ten Books on Architecture. Dover Publications, New York, NY. M.H. Morgan, translator.Google Scholar
Vogel, A., Rodriguez, C., Warnken, W., and Izpisua Belmonte, J.C. (1995). Dorsal cell fate specified by chick Lmx1 during vertebrate limb development. Nature 378, 716720.Google Scholar
Vogt, K., Schnaitmann, C., Dylla, K.V., Knapek, S., Aso, Y., Rubin, G.M., and Tanimoto, H. (2014). Shared mushroom body circuits underlie visual and olfactory memories in Drosophila. eLife 3, e02395.Google Scholar
Vogt, N., and Desplan, C. (2014). Vision. In Dubnau, J. (ed.), Behavioral Genetics of the Fly (Drosophila melanogaster). Cambridge University Press, New York, NY, pp. 3748.Google Scholar
Vogt, T.F., and Duboule, D. (1999). Antagonists go out on a limb. Cell 99, 563566.Google Scholar
Von Allmen, G., Hogga, I., Spierer, A., Karch, F., Bender, W., Gyurkovics, H., and Lewis, E. (1996). Splits in the fruitfly Hox gene complexes. Nature 380, 116.Google Scholar
von Frisch, K. (1967). The Dance Language and Orientation of Bees. Harvard University Press, Cambridge, MA.Google Scholar
von Lintig, J. (2012). Metabolism of carotenoids and retinoids related to vision. J. Biol. Chem. 287, #3, 16271634.Google Scholar
von Schantz, M., Jenkins, A., and Archer, S.N. (2006). Evolutionary history of the vertebrate period genes. J. Mol. Evol. 62, 701707.Google Scholar
Vopalensky, P., and Kozmik, Z. (2009). Eye evolution: common use and independent recruitment of genetic components. Philos. Trans. R. Soc. Lond. B 364, 28192832.Google Scholar
Vopalensky, P., Pergner, J., Liegertova, M., Benito-Gutierrez, E., Arendt, D., and Kozmik, Z. (2013). Molecular analysis of the amphioxus frontal eye unravels the evolutionary origin of the retina and pigment cells of the vertebrate eye. PNAS 109, #38, 1538315388.Google Scholar
Vora, S., and Phillips, B.T. (2015). Centrosome-associated degradation limits b-catenin inheritance by daughter cells after asymmetric division. Curr. Biol. 25, 10051016.Google Scholar
Voss, A.K., Collin, C., Dixon, M.P., and Thomas, T. (2009). Moz and retinoic acid coordinately regulate H3K9 acetylation, Hox gene expression, and segment identity. Dev. Cell 17, 674686.Google Scholar
Vosshall, L.B., and Stensmyr, M.C. (2005). Wake up and smell the pheromones. Neuron 45, 179187.Google Scholar
Vosshall, L.B., and Stocker, R.F. (2007). Molecular architecture of smell and taste in Drosophila. Annu. Rev. Neurosci. 30, 505533.Google Scholar
Vosshall, L.B., Wong, A.M., and Axel, R. (2000). An olfactory sensory map in the fly brain. Cell 102, 147159.Google Scholar
Vulliemoz, S., Raineteau, O., and Jabaudon, D. (2005). Reaching beyond the midline: why are human brains cross wired? Lancet Neurol. 4, 8799.Google Scholar
Wada, H., Escriva, H., Zhang, S., and Laudet, V. (2006). Conserved RARE localization in amphioxus Hox clusters and implications for Hox code evolution in the vertebrate neural crest. Dev. Dyn. 235, 15221531.Google Scholar
Wada, H., Garcia-Fernàndez, J., and Holland, P.W.H. (1999). Colinear and segmental expression of amphioxus Hox genes. Dev. Biol. 213, 131141.Google Scholar
Wadsworth, W.G. (2005). Axon pruning: C. elegans makes the cut. Curr. Biol. 15, R796R798.Google Scholar
Wagner, A. (2011). Genotype networks shed light on evolutionary constraints. Trends Ecol. Evol. 26, 577584.Google Scholar
Wagner, E.F., and Petruzzelli, M. (2015). A waste of insulin interference. Nature 521, 430431.Google Scholar
Wagner, G.P. (2007). The developmental genetics of homology. Nat. Rev. Genet. 8, 473479.Google Scholar
Wagner, G.P. (2014). Homology, Genes, and Evolutionary Innovation. Princeton University Press, Princeton, NJ.Google Scholar
Wagner, G.P. (2016). What is “homology thinking” and what is it for? J. Exp. Zool. B. Mol. Dev. Evol. 326, 38.Google Scholar
Wagner, R.A., Tabibiazar, R., Liao, A., and Quertermous, T. (2005). Genome-wide expression dynamics during mouse embryonic development reveal similarities to Drosophila development. Dev. Biol. 288, 595611.Google Scholar
Wake, D.B., Wake, M.H., and Specht, C.D. (2011). Homoplasy: from detecting pattern to determining process and mechanism of evolution. Science 331, 10321035.Google Scholar
Wallace, M.T. (2015). Multisensory perception: the building of flavor representations. Curr. Biol. 25, R980R1001.Google Scholar
Wallace, V.A. (2008). Proliferative and cell fate effects of Hedgehog signaling in the vertebrate retina. Brain Res. 1192, 6175.Google Scholar
Wallingford, J.B., and Mitchell, B. (2011). Strange as it may seem: the many links between Wnt signaling, planar cell polarity, and cilia. Genes Dev. 25, 201213.Google Scholar
Wan, G., Corfas, G., and Stone, J.S. (2013). Inner ear supporting cells: rethinking the silent majority. Semin. Cell Dev. Biol. 24, 448459.Google Scholar
Wang, G.-Z., Marini, S., Ma, X., Yang, Q., Zhang, X., and Zhu, Y. (2014). Improvement of Dscam homophilic binding affinity throughout Drosophila evolution. BMC Evol. Biol. 14, Article 186.Google Scholar
Wang, J., Zugates, C.T., Liang, I.H., Lee, C.-H.J., and Lee, T. (2002). Drosophila Dscam is required for divergent segregation of sister branches and suppresses ectopic bifurcation of axons. Neuron 33, 559571.Google Scholar
Wang, R., Chen, C.-C., Hara, E., Rivas, M.V., Roulhac, P.L., Howard, J.T., Chakraborty, M., Audet, J.-N., and Jarvis, E.D. (2014). Convergent differential regulation of SLIT-ROBO axon guidance genes in the brains of vocal learners. J. Comp. Neurol. 523, 892906.Google Scholar
Wang, S., and Samakovlis, C. (2012). Grainy head and its target genes in epithelial morphogenesis and wound healing. Curr. Top. Dev. Biol. 98, 3563.Google Scholar
Wang, W., Nossoni, Z., Berbasova, T., Watson, C.T., Yapici, I., Lee, K.S.S., Vasileiou, C., Geiger, J.H., and Borhan, B. (2012). Tuning the electronic absorption of protein-embedded all-trans-retinal. Science 338, 13401343.Google Scholar
Wang, X., Wang, T., Jiao, Y., von Lintig, J., and Montell, C. (2010). Requirement for an enzymatic visual cycle in Drosophila. Curr. Biol. 20, 93102.Google Scholar
Wang, Y., Gao, Y., Imsland, F., Gu, X., Feng, C., Liu, R., Song, C., Tixier-Boichard, M., Gourichon, D., Li, Q., Chen, K., Li, H., Andersson, L., Hu, X., and Li, N. (2012). The Crest phenotype in chicken is associated with ectopic expression of Hoxc8 in cranial skin. PLoS ONE 7, #4, e34012.Google Scholar
Wang, Z., Nudelman, A., and Storm, D.R. (2007). Are pheromones detected through the main olfactory epithelium? Mol. Neurobiol. 35, 317323.Google Scholar
Wang, Z., Singhvi, A., Kong, P., and Scott, K. (2004). Taste representations in the Drosophila brain. Cell 117, 981991.Google Scholar
Wangler, M.F., Yamamoto, S., and Bellen, H.J. (2015). Fruit flies in biomedical research. Genetics 199, 639653.Google Scholar
Wanninger, A., Kristof, A., and Brinkmann, N. (2009). Sipunculans and segmentation. Commun. Integr. Biol. 2, 5659.Google Scholar
Warchol, M.E., and Montcouquiol, M. (2010). Maintained expression of the planar cell polarity molecule Vangl2 and reformation of hair cell orientation in the regenerating inner ear. J. Assoc. Res. Otolaryngol. 11, 395406.Google Scholar
Ward, E.J., Zhou, X., Riddiford, L.M., Berg, C.A., and Ruohola-Baker, H. (2006). Border of Notch activity establishes a boundary between the two dorsal appendage tube cell types. Dev. Biol. 297, 461470.Google Scholar
Wardill, T.J., List, O., Li, X., Dongre, S., McCulloch, M., Ting, C.-Y., O'Kane, C.J., Tang, S., Lee, C.-H., Hardie, R.C., and Juusola, M. (2012). Multiple spectral inputs improve motion discrimination in the Drosophila visual system. Science 336, 925931.Google Scholar
Ware, M., Dupé, V., and Schubert, F.R. (2015). Evolutionary conservation of the early axon scaffold in the vertebrate brain. Dev. Dyn. 244, 12021214.Google Scholar
Warner, J.F., and McClay, D.R. (2014). Left-right asymmetry in the sea urchin. Genesis 52, 481487.Google Scholar
Warrant, E.J. (2015). Photoreceptor evolution: ancient “cones” turn out to be rods. Curr. Biol. 25, R148R151.Google Scholar
Warrant, E.J., and Johnsen, S. (2013). Vision and the light environment. Curr. Biol. 23, R990R994.Google Scholar
Warrant, E.J., and McIntyre, P.D. (1993). Arthropod eye design and the physical limits to spatial resolving power. Prog. Neurobiol. 40, 413461.Google Scholar
Wartlick, O., Jülicher, F., and Gonzalez Gaitan, M. (2014). Growth control by a moving morphogen gradient during Drosophila eye development. Development 141, 18841893.Google Scholar
Washington, I., Zhou, J., Jockusch, S., Turro, N.J., Nakanishi, K., and Sparrow, J.R. (2007). Chlorophyll derivatives as visual pigments for super vision in the red. Photochem. Photobiol. Sci. 6, 775779.Google Scholar
Washington, N.L., Haendel, M.A., Mungall, C.J., Ashburner, M., Westerfield, M., and Lewis, S.E. (2009). Linking human diseases to animal models using ontology-based phenotype annotation. PLoS Biol. 7, #11, e1000247.Google Scholar
Wässle, H. (1999). A patchwork of cones. Nature 397, 473475.Google Scholar
Watanabe, D., Saijoh, Y., Nonaka, S., Sasaki, G., Ikawa, Y., Yokoyama, T., and Hamada, H. (2003). The left-right determinant Inversin is a component of node monocilia and other 9+0 cilia. Development 130, 17251734.Google Scholar
Watanabe, H., Schmidt, H.A., Kuhn, A., Höger, S.K., Kocagöz, Y., Laumann-Lipp, N., Özbek, S., and Holstein, T.W. (2014). Nodal signalling determines biradial symmetry in Hydra. Nature 515, 112115.Google Scholar
Watanabe, S., Umeki, N., Ikebe, R., and Ikebe, M. (2008). Impacts of Usher syndrome type 1B mutations on human myosin VIIa motor function. Biochemistry 47, 95059513.Google Scholar
Watson, F.L., Püttmann-Holgado, R., Thomas, F., Lamar, D.L., Hughes, M., Kondo, M., Rebel, V.I., and Schmucker, D. (2005). Extensive diversity of Ig-superfamily proteins in the immune system of insects. Science 309, 18741878.Google Scholar
Wawersik, S., and Maas, R.L. (2000). Vertebrate eye development as modeled in Drosophila. Hum. Mol. Genet. 9, 917925.Google Scholar
Weaver, J. (2012). Striking similarities in fly and vertebrate olfactory network formation. PLoS Biol. 10, #10, e1001401.Google Scholar
Weavers, H., Prieto-Sánchez, S., Grawe, F., Garcia-López, A., Artero, R., Wilsch-Braüninger, M., Ruiz-Gómez, M., Skaer, H., and Denholm, B. (2009). The insect nephrocyte is a podocyte-like cell with a filtration slit diaphragm. Nature 457, 322326.Google Scholar
Weber, T., Göpfert, M.C., Winter, H., Zimmermann, U., Kohler, H., Meier, A., Hendrich, O., Rohbock, K., Robert, D., and Knipper, M. (2003). Expression of prestin-homologous solute carrier (SLC26) in auditory organs of nonmammalian vertebrates and insects. PNAS 100, #13, 76907695.Google Scholar
Weddell, T.D., Mellado-Lagarde, M., Lukashkina, V.A., Lukashkin, A.N., Zuo, J., and Russell, I.J. (2011). Prestin links extrinsic tuning to neural excitation in the mammalian cochlea. Curr. Biol. 21, R682R683.Google Scholar
Wehner, R. (2005). Brainless eyes. Nature 435, 157159.Google Scholar
Weiner, J. (1999). Time, Love, Memory. Random House, New York, NY.Google Scholar
Weir, P.T., and Dickinson, M.H. (2012). Flying Drosophila orient to sky polarization. Curr. Biol. 22, 2127.Google Scholar
Weirauch, M.T., and Hughes, T.R. (2011). A catalogue of eukaryotic transcription factor types, their evolutionary origin, and species distribution. In Hughes, T.R. (ed.), A Handbook of Transcription Factors. Springer, New York, NY.Google Scholar
Weisblat, D.A., and Kuo, D.-H. (2014). Developmental biology of the leech Helobdella. Int. J. Dev. Biol. 58, 429443.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
Welniarz, Q., Dusart, I., Gallea, C., and Roze, E. (2015). One hand clapping: lateralization of motor control. Front. Neuroanat. 9, Article 75.Google Scholar
Welsh, I.C., Thomsen, M., Gludish, D.W., Alfonso-Parra, C., Bai, Y., Martin, J.F., and Kurpios, N.A. (2013). Integration of left-right Pitx2 transcription and Wnt signaling drives asymmetric gut morphogenesis via Daam2. Dev. Cell 26, 629644.Google Scholar
Werner, H.B. (2013). Do we have to reconsider the evolutionary emergence of myelin? Front. Cell. Neurosci. 7, Article 217.Google Scholar
Werner, M.E., Ward, H.H., Phillips, C.L., Miller, C., Gattone, V.H., and Bacallao, R.L. (2013). Inversin modulates the cortical actin network during mitosis. Am. J. Physiol. Cell Physiol. 305, C36C47.Google Scholar
Wernet, M.F., Celik, A., Mikeladze-Dvali, T., and Desplan, C. (2007). Generation of uniform fly retinas. Curr. Biol. 17, R1002R1003.Google Scholar
Wernet, M.F., and Desplan, C. (2004). Building a retinal mosaic: cell-fate decision in the fly eye. Trends Cell Biol. 14, 576584.Google Scholar
Wernet, M.F., Huberman, A.D., and Desplan, C. (2014). So many pieces, one puzzle: cell type specification and visual circuitry in flies and mice. Genes Dev. 28, 25652584.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
Wernet, M.F., Meier, K.M., Baumann-Klausener, F., Dorfman, R., Weihe, U., Labhart, T., and Desplan, C. (2014). Genetic dissection of photoreceptor subtype specification by the Drosophila melanogaster zinc finger proteins Elbow and No ocelli. PLoS Genet. 10, #3, e1004210.Google Scholar
Wernet, M.F., Perry, M.W., and Desplan, C. (2015). The evolutionary diversity of insect retinal mosaics: common design principles and emerging molecular logic. Trends Genet. 31, 316328.Google Scholar
Wernet, M.F., Velez, M.M., Clark, D.A., Baumann-Klausener, F., Brown, J.R., Klovstad, M., Labhart, T., and Clandinin, T.R. (2012). Genetic dissection reveals two separate retinal substrates for polarization vision in Drosophila. Curr. Biol. 22, 1220.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
Wheeler, S.R., Stagg, S.B., and Crews, S.T. (2008). Multiple Notch signaling events control Drosophila CNS midline neurogenesis, gliogenesis and neuronal identity. Development 135, 30713079.Google Scholar
White, L.E., Lucas, G., Richards, A., and Purves, D. (1994). Cerebral asymmetry and handedness. Nature 368, 197198.Google Scholar
White, R.A.H., and Lehmann, R. (1986). A gap gene, hunchback, regulates the spatial expression of Ultrabithorax. Cell 47, 311321.Google Scholar
Whitehouse, D. (2009). Renaissance Genius: Galileo Galilei and His Legacy to Modern Science. Sterling, New York, NY.Google Scholar
Whitesides, G.M., and Grzybowski, B. (2002). Self-assembly at all scales. Science 295, 24182421.Google Scholar
Whitfield, T.T. (2015). Development of the inner ear. Curr. Opin. Genet. Dev. 32, 112118.Google Scholar
Whitfield, T.T., and Monk, N. (2014). Julian Hart Lewis, F.R.S. (1946–2014). Dev. Cell 29, 507509.Google Scholar
Wibowo, I., Pinto-Teixeira, F., Satou, C., Higashijima, S.-i., and López-Schier, H. (2011). Compartmentalized Notch signaling sustains epithelial mirror symmetry. Development 138, 11431152.Google Scholar
Wicher, D. (2012). Functional and evolutionary aspects of chemoreceptors. Front. Cell. Neurosci. 6, Article 48.Google Scholar
Wicher, D., Schäfer, R., Bauernfeind, R., Stensmyr, M.C., Heller, R., Heinemann, S.H., and Hansson, B.S. (2008). Drosophila odorant receptors are both ligand-gated and cyclic-nucleotide-activated cation channels. Nature 452, 10071011.Google Scholar
Wieschaus, E., and Nüsslein-Volhard, C. (2014). Walter Gehring (1939–2014). Curr. Biol. 24, R632R634.Google Scholar
Wikler, K.C., and Rakic, P. (1991). Relation of an array of early-differentiating cones to the photoreceptor mosaic in the primate retina. Nature 351, 397400.Google Scholar
Wilkins, A.S. (1989). Organizing the Drosophila posterior pattern: why has the fruit fly made life so complicated for itself? BioEssays 11, 6769.Google Scholar
Wilkins, A.S. (2002). The Evolution of Developmental Pathways. Sinauer, Sunderland, MA.Google Scholar
Wilkins, A.S. (2014). “The genetic tool-kit”: the life history of an important metaphor. In Advances in Evolutionary Developmental Biology. Wiley-Blackwell, New York, NY, pp. 114.Google Scholar
Williams, D.S. (1991). Actin filaments and photoreceptor membrane turnover. BioEssays 13, 171178.Google Scholar
Williams, R.J. (1971). You Are Extraordinary. Pyramid Books, New York, NY.Google Scholar
Williams, T.A., Nulsen, C., and Nagy, L.M. (2002). A complex role for Distal-less in crustacean appendage development. Dev. Biol. 241, 302312.Google Scholar
Wilmer, P. (2003). Convergence and homoplasy in the evolution of organismal form. In Müller, G.B. and Newman, S.A. (eds.), Origination of Organismal Form: Beyond the Gene in Developmental and Evolutionary Biology. MIT Press, Cambridge, MA., pp. 3349.Google Scholar
Wilson, M.J., and Dearden, P.K. (2011). Diversity in insect axis formation: two orthodenticle genes and hunchback act in anterior patterning and influence dorsoventral organization in the honeybee (Apis mellifera). Development 138, 34973507.Google Scholar
Wilson, R.I. (2013). Early olfactory processing in Drosophila: mechanisms and principles. Annu. Rev. Neurosci. 36, 217241.Google Scholar
Wilson, R.I., and Mainen, Z.F. (2006). Early events in olfactory processing. Annu. Rev. Neurosci. 29, 163201.Google Scholar
Wilson, S.I., Shafer, B., Lee, K.J., and Dodd, J. (2008). A molecular program for contralateral trajectory: Rig-1 control by LIM homeodomain transcription factors. Neuron 59, 413424.Google Scholar
Wilson-Sanders, S.E. (2011). Invertebrate models for biomedical research, testing, and education. ILAR J. 52, 126152.Google Scholar
Wimmer, E.A., Carleton, A., Harjes, P., Turner, T., and Desplan, C. (2000). bicoid-independent formation of thoracic segments in Drosophila. Science 287, 24762479.Google Scholar
Winchell, C.J., and Jacobs, D.K. (2013). Expression of the Lhx genes apterous and lim1 in an errant polychaete: implications for bilaterian appendage evolution, neural development, and muscle diversification. EvoDevo 4, Article 4.Google Scholar
Winchell, C.J., Valencia, J.E., and Jacobs, D.K. (2010). Expression of Distal-less, dachshund, and optomotor blind in Neanthes arenaceodentata (Annelida, Nereididae) does not support homology of appendage-forming mechanisms across the Bilateria. Dev. Genes Evol. 220, 275295.Google Scholar
Winston, R., and Wilson, D.E. (2006). Human. Dorling Kindersley, New York, NY.Google Scholar
Winterbottom, E.F., Illes, J.C., Faas, L., and Isaacs, H.V. (2010). Conserved and novel roles for the Gsh2 transcription factor in primary neurogenesis. Development 137, 26232631.Google Scholar
Wodarz, A., and Huttner, W.B. (2003). Asymmetric cell division during neurogenesis in Drosophila and vertebrates. Mech. Dev. 120, 12971309.Google Scholar
Wojtowicz, W.M., Wu, W., Andre, I., Qian, B., Baker, D., and Zipursky, S.L. (2007). A vast repertoire of Dscam binding specificities arises from modular interactions of variable Ig domains. Cell 130, 11341145.Google Scholar
Wolf, F.W., and Heberlein, U. (2003). Invertebrate models of drug abuse. J. Neurobiol. 54, 161178.Google Scholar
Wolff, G.H., and Strausfeld, N.J. (2015). Genealogical correspondence of mushroom bodies across invertebrate phyla. Curr. Biol. 25, 3844.Google Scholar
Wolff, G.H., and Strausfeld, N.J. (2016). Genealogical correspondence of a forebrain centre implies an executive brain in the protostome–deuterostome bilaterian ancestor. Philos. Trans. R. Soc. Lond. B 371, 20150055. [See alsoGoogle ScholarGoogle Scholar
Wollesen, T., Monje, S.V.R., Todt, C., Degnan, B.M., and Wanninger, A. (2015). Ancestral role of Pax2/5/8 in molluscan brain and multimodal sensory system development. BMC Evol. Biol. 15, Article 231.Google Scholar
Wolpert, L. (1968). The French Flag Problem: a contribution to the discussion on pattern development and regulation. In Waddington, C.H. (ed.), Towards a Theoretical Biology. I. Prolegomena. Aldine, Chicago, IL, pp. 125133.Google Scholar
Wolpert, L. (1969). Positional information and the spatial pattern of cellular differentiation. J. Theor. Biol. 25, 147.Google Scholar
Wolpert, L. (2014). Revisiting the F-shaped molecule: is its identity solved? Genesis 52, 455457.Google Scholar
Wootton, R.J. (1976). The fossil record and insect flight. In Rainey, R.C. (ed.), Insect Flight. Wiley, New York, NY, pp. 235254.Google Scholar
Wotton, K.R., Weierud, F.K., Juárez-Morales, J.L., Alvares, L.E., Dietrich, S., and Lewis, K.E. (2009). Conservation of gene linkage in dispersed vertebrate NK homeobox clusters. Dev. Genes Evol. 219, 481496.Google Scholar
Wray, G.A. (1998). Promoter logic. Science 279, 18711872.Google Scholar
Wray, G.A. (2001). Resolving the Hox paradox. Science 292, 22562257.Google Scholar
Wray, G.A., and Abouheif, E. (1998). When is homology not homology? Curr. Opin. Genet. Dev. 8, 675680.Google Scholar
Wright, G.A. (2015). Olfaction: smells like fly food. Curr. Biol. 25, R144R146.Google Scholar
Wu, J., and Cohen, S.M. (2000). Proximal distal axis formation in the Drosophila leg: distinct functions of Teashirt and Homothorax in the proximal leg. Mech. Dev. 94, 4756.Google Scholar
Wu, J., and Mlodzik, M. (2009). A quest for the mechanism regulating global planar cell polarity of tissues. Trends Cell Biol. 19, 295305.Google Scholar
Wu, J.Y., and Rao, Y. (1999). Fringe: defining borders by regulating the Notch pathway. Curr. Opin. Neurobiol. 9, 537543.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
Wu, X. (2010). Wg signaling in Drosophila heart development as a pioneering model. J. Genet. Genomics 37, 593603.Google Scholar
Wurst, W., and Bally-Cuif, L. (2001). Neural plate patterning: upstream and downstream of the isthmic organizer. Nat. Rev. Neurosci. 2, 99108.Google Scholar
Wyart, C., Webster, W.W., Chen, J.H., Wilson, S.R., McClary, A., Khan, R.M., and Sobel, N. (2007). Smelling a single component of male sweat alters levels of cortisol in women. J. Neurosci. 27, #6, 12611265.Google Scholar
Wyatt, T.D. (2015). The search for human pheromones: the lost decades and the necessity of returning to first principles. Proc. R. Soc. Lond. B 282, 20142994.Google Scholar
Xavier-Neto, J., Castro, R.A., Sampaio, A.C., Azambuja, A.P., Castillo, H.A., Cravo, R.M., and Simões-Costa, M.S. (2007). Parallel avenues in the evolution of hearts and pumping organs. Cell. Mol. Life Sci. 64, 719734.Google Scholar
Xiang, Y., Yuan, Q., Vogt, N., Looger, L.L., Jan, L.Y., and Jan, Y.N. (2010). Light-avoidance-mediating photoreceptors tile the Drosophila larval body wall. Nature 468, 921926.Google Scholar
Xie, B., Charlton-Perkins, M., McDonald, E., Gebelein, B., and Cook, T. (2007). Senseless functions as a molecular switch for color photoreceptor differentiation in Drosophila. Development 134, 42434253.Google Scholar
Xu, P.-X., Zhang, X., Heaney, S., Yoon, A., Michelson, A.M., and Maas, R.L. (1999). Regulation of Pax6 expression is conserved between mice and flies. Development 126, 383395.Google Scholar
Xue, T., Do, M.T.H., Riccio, A., Jiang, Z., Hsieh, J., Wang, H.C., Merbs, S.L., Welsbie, D.S., Yoshioka, T., Weissgerber, P., Stolz, S., Flockerzi, V., Freichel, M., Simon, M.I., Clapham, D.E., and Yau, K.-W. (2011). Melanopsin signalling in mammalian iris and retina. Nature 479, 6773.Google Scholar
Yack, J.E. (2004). The structure and function of auditory chordotonal organs in insects. Microsc. Res. Tech. 63, 315337.Google Scholar
Yager, D., and Hoy, R.R. (1986). The cyclopean ear: a new sense for the praying mantis. Science 231, 727729.Google Scholar
Yager, D., and May, M. (1993). Coming in on a wing and an ear. Nat. Hist. 102, #1, 2833.Google Scholar
Yamagata, M., and Sanes, J.R. (2008). Dscam and Sidekick proteins direct lamina-specific synaptic connections in vertebrate retina. Nature 451, 465469.Google Scholar
Yamaguchi, S., Desplan, C., and Heisenberg, M. (2010). Contribution of photoreceptor subtypes to spectral wavelength preference in Drosophila. PNAS 107, #12, 56345639.Google Scholar
Yamaguchi, S., Wolf, R., Desplan, C., and Heisenberg, M. (2008). Motion vision is independent of color in Drosophila. PNAS 105, #12, 49104915.Google Scholar
Yamaguchi, Y., and Miura, M. (2015). Programmed cell death and caspase functions during neural development. Curr. Top. Dev. Biol. 114, 159184.Google Scholar
Yamaguchi, Y., and Miura, M. (2015). Programmed cell death in neurodevelopment. Dev. Cell 32, 478490.Google Scholar
Yamamoto, S., Jaiswal, M., Charng, W.-L., Gambin, T., Karaca, E., Mirzaa, G., Wiszniewski, W., Sandoval, H., Haelterman, N.A., Xiong, B., Zhang, K., Bayat, V., David, G., Li, T., Chen, K., Gala, U., Harel, T., Pehlivan, D., Penney, S., Vissers, L.E.L.M., de Ligt, J., Jhangiani, S.N., Xie, Y., Tsang, S.H., Parman, Y., Sivaci, M., Battaloglu, E., Muzny, D., Wan, Y.-W., Liu, Z., Lin-Moore, A.T., Clark, R.D., Curry, C.J., Link, N., Schulze, K.L., Boerwinkle, E., Dobyns, W.B., Allikmets, R., Gibbs, R.A., Chen, R., Lupski, J.R., Wangler, M.F., and Bellen, H.J. (2014). A Drosophila genetic resource of mutants to study mechanisms underlying human genetic diseases. Cell 159, 200214.Google Scholar
Yan, B. (2010). Numb: from flies to humans. Brain Dev. 32, 293298.Google Scholar
Yang, H.H., and Clandinin, T.R. (2014). What can fruit flies teach us about karate? eLife 3, e04040.Google Scholar
Yang, J., Ortega-Hernández, J., Butterfield, N.J., Liu, Y., Boyan, G.S., Hou, J.-b., Lan, T., and Zhang, X.-g. (2016). Fuxianhuiid ventral nerve cord and early nervous system evolution in Panarthropoda. PNAS 113, #11, 29882993.Google Scholar
Yang, L., Li, R., Kaneko, T., Takle, K., Morikawa, R.K., Essex, L., Wang, X., Zhou, J., Emoto, K., Xiang, Y., and Ye, B. (2014). Trim9 regulates activity-dependent fine-scale topography in Drosophila. Curr. Biol. 24, 10241030.Google Scholar
Yang, M., and Meyer-Rochow, V.B. (2004). Fine-structural details of the photoreceptor membranes in the ocellus of the scale-insect parasite Centrodora sp. (Hymenoptera; Aphenelidae): a case of gene transfer between host and parasite? Biocell 28, 151154.Google Scholar
Yang, Y., Kovács, M., Sakamoto, T., Zhang, F., Kiehart, D.P., and Sellers, J.R. (2006). Dimerized Drosophila myosin VIIa: a processive motor. PNAS 103, #15, 57465751.Google Scholar
Yang, Y., and Mlodzik, M. (2015). Wnt-Frizzled/Planar cell polarity signaling: cellular orientation by facing the wind (Wnt). Annu. Rev. Cell Dev. Biol. 31, 623646.Google Scholar
Yang, Z., Bertolucci, F., Wolf, R., and Heisenberg, M. (2013). Flies cope with uncontrollable stress by learned helplessness. Curr. Biol. 23, 799803.Google Scholar
Yao, Z., and Shafer, O.T. (2014). The Drosophila circadian clock is a variably coupled network of multiple peptidergic units. Science 343, 15161520.Google Scholar
Yarmolinsky, D.A., Zuker, C.S., and Ryba, N.J.P. (2009). Common sense about taste: from mammals to insects. Cell 139, 234244.Google Scholar
Yau, K.-W., and Hardie, R.C. (2009). Phototransduction motifs and variations. Cell 139, 246264.Google Scholar
Yekta, S., Tabin, C.J., and Bartel, D.P. (2008). MicroRNAs in the Hox network: an apparent link to posterior prevalence. Nat. Rev. Genet. 9, 789796.Google Scholar
Yeo, R.A., and Gangestad, S.W. (1994). Developmental origins of variation in human hand preference. In Markow, T.A. (ed.), Developmental Instability: Its Origins and Evolutionary Implications. Kluwer, London, pp. 283298.Google Scholar
Yi, C.H., and Yuan, J. (2009). The Jekyll and Hyde functions of caspases. Dev. Cell 16, 2134.Google Scholar
Yi, H., and Norell, M.A. (2015). The burrowing origin of modern snakes. Sci. Adv. 1, e1500743.Google Scholar
Yogev, S., and Shen, K. (2014). Cellular and molecular mechanisms of synaptic specificity. Annu. Rev. Cell Dev. Biol. 30, 417437.Google Scholar
Yokoyama, S., and Radlwimmer, F.B. (2001). The molecular genetics and evolution of red and green color vision in vertebrates. Genetics 158, 16971710.Google Scholar
Yokoyama, S., Xing, J., Liu, Y., Faggionato, D., Altun, A., and Starmer, W.T. (2014). Epistatic adaptive evolution of human color vision. PLoS Genet. 10, #12, e1004884.Google Scholar
Yokoyama, T., Copeland, N.G., Jenkins, N.A., Montgomery, C.A., Elder, F.F.B., and Overbeek, P.A. (1993). Reversal of left-right asymmetry: a situs inversus mutation. Science 260, 679682.Google Scholar
Yong, E. (2015). Seeing the light. Natl. Geogr. 229, #2, 3057.Google Scholar
Yoshiba, S., and Hamada, H. (2014). Roles of cilia, fluid flow, and Ca2+ signaling in breaking of left-right symmetry. Trends Genet. 30, 1017.Google Scholar
Yoshida, M.-a., Shigeno, S., Tsuneki, K., and Furuya, H. (2010). Squid vascular endothelial growth factor receptor: a shared molecular signature in the convergent evolution of closed circulatory systems. Evol. Dev. 12, 2533.Google Scholar
Yoshida, M.-a., Yura, K., and Ogura, A. (2014). Cephalopod eye evolution was modulated by the acquisition of Pax-6 splicing variants. Sci. Rep. 4, Article 4256.Google Scholar
Yost, H.J. (1999). Diverse initiaiton in a conserved left-right pathway? Curr. Opin. Genet. Dev. 9, 422426.Google Scholar
Yost, H.J. (2003). Left-right asymmetry: nodal cilia make and catch a wave. Curr. Biol. 13, R808R809.Google Scholar
Young, M.W. (2000). The tick-tock of the biological clock. Sci. Am. 282, #3, 6471.Google Scholar
Young, M.W., and Kay, S.A. (2001). Time zones: a comparative genetics of circadian clocks. Nat. Rev. Genet. 2, 702715.Google Scholar
Yu, J.-K., Holland, L.Z., and Holland, N.D. (2002). An amphioxus nodal gene (AmphiNodal) with early symmetrical expression in the organizer and mesoderm and later asymmetrical expression associated with left-right axis formation. Evol. Dev. 4, 418425.Google Scholar
Yu, J.-K., Satou, Y., Holland, N.D., Shin-I, T., Kohara, Y., Satoh, N., Bronner-Fraser, M., and Holland, L.Z. (2007). Axial patterning in cephalochordates and the evolution of the organizer. Nature 445, 613617.Google Scholar
Yu, W., and Hardin, P.E. (2006). Circadian oscillators of Drosophila and mammals. J. Cell Sci. 119, 47934795.Google Scholar
Yuan, S., Yu, X., Asara, J.M., Heuser, J.E., Ludtke, S.J., and Akey, C.W. (2011). The holo-apoptosome: activation of procaspase-9 and interactions with caspase-3. Structure 19, 10841096.Google Scholar
Yuan, S., Yu, X., Topf, M., Dorstyn, L., Kumar, S., Ludtke, S.J., and Akey, C.W. (2011). Structure of the Drosophila apoptosome at 6.9 Å resolution. Structure 19, 128140.Google Scholar
Zaccardi, G., Kelber, A., Sison-Mangus, M.P., and Briscoe, A.D. (2006). Color discrimination in the red range with only one long-wavelength sensitive opsin. J. Exp. Biol. 209, 19441955.Google Scholar
Zaffran, S., El Robrini, N., and Bertrand, N. (2014). Retinoids and cardiac development. J. Dev. Biol. 2, 5071.Google Scholar
Zaffran, S., Reim, I., Qian, L., Lo, P.C., Bodmer, R., and Frasch, M. (2006). Cardioblast-intrinsic Tinman activity controls proper diversification and differentiation of myocardial cells in Drosophila. Development 133, 40734083.Google Scholar
Zagozewski, J.L., Zhang, Q., Pinto, V.I., Wigle, J.T., and Eisenstat, D.D. (2014). The role of homeobox genes in retinal development and disease. Dev. Biol. 393, 195208.Google Scholar
Zak, M., Klis, S.F.L., and Grolman, W. (2015). The Wnt and Notch signalling pathways in the developing cochlea: formation of hair cells and induction of regenerative potential. Int. J. Dev. Neurosci. 47, 247258.Google Scholar
Zakaria, S., Mao, Y., Kuta, A., de Sousa, C.F., Gaufo, G.O., McNeill, H., Hindges, R., Guthrie, S., Irvine, K.D., and Francis-West, P.H. (2014). Regulation of neuronal migration by Dchs1-Fat4 planar cell polarity. Curr. Biol. 24, 16201627.Google Scholar
Zakin, L., and De Robertis, E.M. (2010). Extracellular regulation of BMP signaling. Curr. Biol. 20, R89R92.Google Scholar
Zamora, S., Rahman, I.A., and Smith, A.B. (2012). Plated Cambrian bilaterians reveal the earliest stages of echinoderm evolution. PLoS ONE 7, #6, e38296.Google Scholar
Zampini, V., Rüttiger, L., Johnson, S.L., Franz, C., Furness, D.N., Waldhaus, J., Xiong, H., Hackney, C.M., Holley, M.C., Offenhauser, N., Di Fiore, P.P., Knipper, M., Masetto, S., and Marcotti, W. (2011). Eps8 regulates hair bundle length and functional maturation of mammalian auditory hair cells. PLoS Biol. 9, #4, e1001048.Google Scholar
Zarin, A.A., Asadzadeh, J., Hokamp, K., McCartney, D., Yang, L., Bashaw, G.J., and Labrador, J.-P. (2014). A transcription factor network coordinates attraction, repulsion, and adhesion combinatorially to control motor axon pathway selection. Neuron 81, 12971311.Google Scholar
Zarin, A.A., Asadzadeh, J., and Labrador, J.-P. (2014). Transcriptional regulation of guidance at the midline and in motor circuits. Cell. Mol. Life Sci. 71, 419432.Google Scholar
Zdobnov, E.M., and Bork, P. (2007). Quantification of insect genome divergence. Trends Genet. 23, 1620.Google Scholar
Zeitouni, B., Sénatore, S., Séverac, D., Aknin, C., Sémériva, M., and Perrin, L. (2007). Signalling pathways involved in adult heart formation revealed by gene expression profiling in Drosophila. PLoS Genet. 3, #10, e174.Google Scholar
Zhang, J. (2003). Evolution by gene duplication: an update. Trends Ecol. Evol. 18, 292298.Google Scholar
Zhang, J., and Webb, D.M. (2003). Evolutionary deterioration of the vomeronasal pheromone transduction pathway in catarrhine primates. PNAS 100, #14, 83378341.Google Scholar
Zhang, W., Cheng, L.E., Kittelmann, M., Li, J., Petkovic, M., Cheng, T., Jin, P., Guo, Z., Göpfert, M.C., Jan, L.Y., and Jan, Y.N. (2015). Ankyrin repeats convey force to gate the NOMPC mechanotransduction channel. Cell 162, 13911402.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
Zhang, Y., Yang, Y., Trujillo, C., Zhong, W., and Leung, Y.F. (2012). The expression of irx7 in the inner nuclear layer of zebrafish retina is essential for a proper retinal development and lamination. PLoS ONE 7, #4, e36145.Google Scholar
Zhao, L., Svingen, T., Ng, E.T., and Koopman, P. (2015). Female-to-male sex reversal in mice caused by transgenic overexpression of Dmrt1. Development 142, 10831088.Google Scholar
Zheng, J., Shen, W., He, D.Z.Z., Long, K.B., Madison, L.D., and Dallos, P. (2000). Prestin is the motor protein of cochlear outer hair cells. Nature 405, 149155.Google Scholar
Zheng, L., Zheng, J., Whitlon, D.S., García-Añoveros, J., and Bartles, J.R. (2010). Targeting of the hair cell proteins cadherin 23, harmonin, myosin XVa, espin, and prestin in an epithelial cell model. J. Neurosci. 30, #21, 71877201.Google Scholar
Zhong, Y.-f., and Holland, P.W.H. (2011). The dynamics of vertebrate homeobox gene evolution: gain and loss of genes in mouse and human lineages. BMC Evol. Biol. 11, Article 169.Google Scholar
Zhou, F., and Roy, S. (2015). Snapshot: motile cilia. Cell 162, 224.Google Scholar
Zhu, B., Pennack, J.A., McQuilton, P., Forero, M.G., Mizuguchi, K., Sutcliffe, B., Gu, C.-J., Fenton, J.C., and Hidalgo, A. (2008). Drosophila neurotrophins reveal a common mechanism for nervous system formation. PLoS Biol. 6, #11, e284.Google Scholar
Zhukovsky, E.A., and Oprian, D.D. (1989). Effect of carboxylic acid side chains on the absorption maximum of visual pigments. Science 246, 928930.Google Scholar
Zhukovsky, E.A., Robinson, P.R., and Oprian, D.D. (1991). Transducin activation by rhodopsin without a covalent bond to the 11-cis-retinal chromophore. Science 251, 558560.Google Scholar
Ziauddin, J., and Schneider, D.S. (2012). Where does innate immunity stop and adaptive immunity begin? Cell Host Microbe 12, 394395.Google Scholar
Ziegler, A.B., Berthelot-Grosjean, M., and Grosjean, Y. (2013). The smell of love in Drosophila. Front. Physiol. 4, Article 72.Google Scholar
Zimmer, C. (2001). Evolution: The Triumph of an Idea. HarperCollins, New York, NY.Google Scholar
Zimmer, C. (2013). Genes are us. And them. Natl. Geogr. 224, #1, 102103.Google Scholar
Zimmerman, A., Bai, L., and Ginty, D.D. (2014). The gentle touch receptors of mammalian skin. Science 346, 950954.Google Scholar
Zinn, K. (2007). Dscam and neuronal uniqueness. Cell 129, 455456.Google Scholar
Zinn, K., and Sun, Q. (1999). Slit branches out: a secreted protein mediates both attractive and repulsive axon guidance. Cell 97, 14.Google Scholar
Zinzen, R.P., Cande, J., Ronshaugen, M., Papatsenko, D., and Levine, M. (2006). Evolution of the ventral midline in insect embryos. Dev. Cell 11, 895902.Google Scholar
Zipursky, S.L., and Sanes, J.R. (2010). Chemoaffinity revisited: Dscams, protocadherins, and neural circuit assembly. Cell 143, 343353.Google Scholar
Zollinger, D.R., Baalman, K.L., and Rasband, M.N. (2015). The ins and outs of polarized axonal domains. Annu. Rev. Cell Dev. Biol. 31, 647667.Google Scholar
Zordan, M.A., and Sandrelli, F. (2015). Circadian clock dysfunction and psychiatric disease: could fruit flies have a say? Front. Neurol. 6, Article 80.Google Scholar
Zou, Z., and Buck, L.B. (2006). Combinatorial effects of odorant mixes in olfactory cortex. Science 311, 14771481.Google Scholar
Zuber, M.E., Gestri, G., Viczian, A.S., Barsacchi, G., and Harris, W.A. (2003). Specification of the vertebrate eye by a network of eye field transcription factors. Development 130, 51555167.Google Scholar
Zuker, C.S. (1994). On the evolution of eyes: would you like it simple or compound? Science 265, 742743.Google Scholar
Zumdahl, S.S. (1986). Chemistry. D.C. Heath, Lexington, MA.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: Deep Homology?
  • Online publication: 23 February 2017
  • Chapter DOI: https://doi.org/10.1017/9781316550175.012
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: Deep Homology?
  • Online publication: 23 February 2017
  • Chapter DOI: https://doi.org/10.1017/9781316550175.012
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: Deep Homology?
  • Online publication: 23 February 2017
  • Chapter DOI: https://doi.org/10.1017/9781316550175.012
Available formats
×