Skip to main content Accessibility help
×
Hostname: page-component-76dd75c94c-sgvz2 Total loading time: 0 Render date: 2024-04-30T09:47:50.339Z Has data issue: false hasContentIssue false

Section 3: - Evolution

Published online by Cambridge University Press:  12 August 2022

Michael M. Halassa
Affiliation:
Massachusetts Institute of Technology
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
The Thalamus , pp. 91 - 138
Publisher: Cambridge University Press
Print publication year: 2022

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

References

Aboitiz, F., Morales, D., & Montiel, J. (2003) The evolutionary origin of the mammalian neocortex: towards an integrated developmental and functional approach. Behav Brain Sci 26:535586.Google Scholar
Agrillo, C., & Bisazza, A. (2017) Understanding the origin of number sense: a review of fish studies. Phil Trans Roy Soc B 373:20160511. doi:10.1098/rstb.2016.0511.Google Scholar
Agrillo, A., Petrazzini, M.E., & Dadda, M. (2013) Illusory patterns are fishy for fish, too. Front Neur Circ 7. doi:10.3389/fncir.2013.00137.Google Scholar
Atoji, Y., Sarkar, S., & Wild, J.M. (2016) Proposed homologue of the dorsomedial subdivision and V-shaped layer of the avian hippocampus to Ammon’s horn and dentate gyrus, respectively. Hipp 26:16081617.CrossRefGoogle Scholar
Baars, B.J. (2003) Working memory requires conscious processes, not vice versa. In: Osaka, N. (Ed.), Neural Basis of Consciousness. John Benjamins Publishing Co., Amsterdam, pp. 1126.CrossRefGoogle Scholar
Bachy, I., Vernier, P., & Rétaux, S. (2001) The LIM-homeodomain gene family in the developing Xenopus brain: conservation and divergences with the mouse related to the evolution of the forebrain. J Neurosci 21:76207629.CrossRefGoogle Scholar
Baldwin, M.K.L., Balaram, P., and Kaas, J.H. (2017) The evolution and functions of nuclei of the visual pulvinar in primates. J Comp Neurol 525:32073226.CrossRefGoogle ScholarPubMed
Belekhova, M.G., Kratskin, I.L., Repérant, J., Pierre, J., Vesselkin, N.P., Kenigfest, N.B., Tumanova, N.L., & Chkheidze, D.D. (1991) Localization of GABA-immunoreactive elements in the thalamus of the tortoise Emys orbicularis. Zh Evol Biokhim Fiziol 27:676685.Google Scholar
Belgard, T.G., Montiel, J.F., Wang, W.Z., García-Moreno, F., Margulies, E.H., Ponting, C.P., & Molnár, Z. (2013) Adult pallial transcriptomes surprise in not reflecting predicted homologies across diverse chicken and mouse pallial sectors. Proc Nat Acad Sci 110:1315013155.Google Scholar
Bickford, M.E. (2016) Thalamic circuit diversity: modulation of the driver/modulator framework. Front Neural Circ 9, Art. 86, doi:10.3389/fncir.2015.00086.Google ScholarPubMed
Bielle, F., Marcos-Mondejar, P., Keita, M., Mailhes, C., Verney, C., Nguyen Ba-Chavet, K., Tessier-Lavigne, M., Lopez-Bendito, G., & Garel, S. (2011) Slit2 activity in the migration of guidepost neurons shapes thalamic projections during development and evolution. Neuron 69:10851098.Google Scholar
Bingman, V.P., Rodríguez, F., & Salas, C. (2017) The hippocampus of nonmammalian vertebrates. In: Kaas, J.H., & Striedter, G. (Eds.), Evolution of Nervous Systems. Vol. 1: The Evolution of the Nervous Systems in Nonmammalian Vertebrates, 2nd ed. Elsevier, Amsterdam, pp. 479489.Google Scholar
Braford, M.R., Jr. (1995) Comparative aspects of forebrain organization in the ray-finned fishes: touchstone or not? Brain Behav Evol 46:259274.Google Scholar
Briscoe, S.D., Albertin, C.B., Rowell, J.J., & Ragsdale, C.W. (2018) Neocortical association cell types in the forebrain of birds and alligators. Curr Biol 28:686696.CrossRefGoogle ScholarPubMed
Broglio, C., Rodriguez, F., Gomez, A., Arias, J.L., & Salas, C. (2010) Selective involvement of goldfish lateral pallium in spatial memory. Behav Brain Res 210:191201.Google Scholar
Brown, M., Keynes, R., & Lumsden, A. (2001) The Developing Brain. Oxford University Press, Oxford, UK.Google Scholar
Bruce, L.L., & Butler, A.B. (1984) Telencephalic connections in lizards. I. Projections to cortex. J Comp Neurol 229:585601.Google Scholar
Bruce, L.L., & Neary, T. (1995) The limbic system of tetrapods: a comparative analysis of cortical and amygdalar populations. Brain Behav Evol 46:224234.Google Scholar
Bulchand, S., Grove, E.A., Porter, F.D., & Tole, S. (2001) LIM-homeodomain gene Lhx2 regulates the formation of the cortical hem. Mech Dev 100:165175.CrossRefGoogle ScholarPubMed
Butler, A.B. (1994a) The evolution of the dorsal pallium in the telencephalon of amniotes: cladistic analysis and a new hypothesis. Brain Res Rev 19:66101.Google Scholar
Butler, A.B. (1994b) The evolution of the dorsal thalamus of jawed vertebrates, including mammals: cladistic analysis and a new hypothesis. Brain Res Rev 19:2965.Google Scholar
Butler, A.B. (1995) The dorsal thalamus of jawed vertebrates: a comparative viewpoint. Brain Behav Evol 46:209223.Google Scholar
Butler, A.B. (2007) The dual elaboration hypothesis of the evolution of the dorsal thalamus. In Krubitzer, L.A., & Kaas, J.H. (Eds.), Evolution of Nervous System in Mammals. Elsevier, Amsterdam, pp. 517523.Google Scholar
Butler, A.B. (2008a) Evolution of brains, cognition, and consciousness. Brain Res Bull 75:442449.Google Scholar
Butler, A.B. (2008b) Evolution of the dorsal thalamus: a morphological-functional review. Thalamus Relate Sys 4:3558.Google Scholar
Butler, A. B. (2009) Evolution and the concept of homology. In: Binder, M.D., Hirokawa, N., & Windhorst, U. (Eds.), Encyclopedic Reference of Neuroscience.Springer, New York, pp. 12081212.Google Scholar
Butler, A.B., & Hodos, W. (2005) Comparative Vertebrate Neuroanatomy, 2nd ed. John Wiley & Sons, Hoboken, NJ.Google Scholar
Butler, A.B., Reiner, A., & Karten, H.J. (2011) Evolution of the amniote pallium and the origins of mammalian neocortex. Ann NY Acad Sci 1225:1427.Google Scholar
Butler, A.B., & Saidel, W.M. (2000) Defining sameness: historical, biological, and generative homology. BioEssays 22:846853.Google Scholar
Caballero-Bleda, M. (1988) Región alar del diencéfalo y mesencéfalo en el conejo : quimioarquitectonía de AChE y NADH-diaforasa como contribución a su neuroanatomíca comparada. PhD thesis, Universidad de Murcia.Google Scholar
Campbell, C.B.G., & Hodos, W. (1970) The concept of homology and the evolution of the nervous system. Brain Behav Evol 3:353367.Google Scholar
Caretta, D., Sbriccoli, A., Santarelli, M., Pinto, F., Granato, A., & Minciacchi, D. (1996) Crossed thalamo-cortical and cortico-thalamic projections in adult mice. Neurosci Lett 204:6972.CrossRefGoogle Scholar
Chen, C.C., Winkler, C.M., Pfenning, A.R., & Jarvis, E.D. (2013) Molecular profiling of the developing avian telencephalon: regional timing and brain subdivision continuities. J Comp Neurol 521:36663701.Google Scholar
Clayton, N.S., Bussey, T.J., & Dickinson, A. (2003a) Can animals recall the past and plan for the future? Nat Rev Neurosci 4:685691.Google Scholar
Clayton, N.S., Bussey, T.J., & Dickinson, A. (2003b) Prometheus to Proust: the case for behavioural criteria for “mental time travel.” Trends Cogn Sci 7:436437.Google Scholar
Clayton, N.S., & Dickinson, A. (1998) Episodic-like memory during cache recovery by scrub jays. Nature 395:272274.Google Scholar
Clayton, N.S., & Dickinson, A. (1999) Scrub jays (Aphelocoma coerulescens) remember the relative time of caching as well as the location and content of their caches. J Comp Psychol 113:403416.Google Scholar
Clayton, N.S., Griffiths, D.P., Emery, N.J., & Dickinson, A. (2001) Elements of episodic-like memory in animals. Phil Trans Roy Soc Lond B 356:14831491.Google Scholar
Colombe, J.B., Sylvester, J., Block, J., & Ulinski, P.S. (2004) Subpial and stellate cells: two populations of interneurons in turtle visual cortex. J Comp Neurol 471:333351.Google Scholar
Cordery, P., & Molnár, Z. (1999) Embryonic development of connections in turtle pallium. J Comp Neurol 413:2654.Google Scholar
Crespo, C., Porteros, A., Arévalo, R., Briñón, J.G., Aijón, J., & Alonso, J. R. (1999) Distribution of parvalbumin immunoreactivity in the brain of the tench (Tinca tinca L., 1758). J Comp Neurol 413:549571.Google Scholar
Csillag, A., & Montagnese, C.M. (2005) Thalamotelencephalic organization in birds. Brain Res Bull 66:303310.Google Scholar
Dadda, M., Agrillo, C., Bisazza, A., & Brown, C. (2015) Laterality enhances numerical skills in the guppy, Poecilia reticulata. Front Behav Neurosci 9:285. doi:10.3389/fnbeh.2015.00285.CrossRefGoogle ScholarPubMed
Dadda, M., Piffer, L., Agrillo, C., & Bisazza, A. (2009). Spontaneous number representation in mosquitofish. Cogn 122:343348.Google Scholar
Darmaillacq, A.S., Dickel, L., Rahmani, N., & Shashar, N. (2011) Do reef fish, Variola louit and Scarus niger, perform amodal competition? Evidence from a field study. J Comp Psychol 125:273277.Google Scholar
Dávila, J.C., Andreu, M.J., Real, A., Puelles, L., & Guirado, S. (2002) Mesencephalic and diencephalic afferent connections to the thalamic nucleus rotundus in the lizard, Psammodromus algirus. Europ J Neurosci 16:267282.Google Scholar
Dávila, J.C., Guirado, S., & Puelles, L. (2000) Expression of calcium-binding proteins in the diencephalon of the lizard Psammodromus algirus. J Comp Neurol 427:6792.Google Scholar
de Beer, G. (1971) Homology, an unsolved problem. In: Head, J.J., & Lowenstein, O.E. (Eds.), Oxford Biology Readers, No. 11. Oxford University Press, London.Google Scholar
Delius, J.D., & Bennetto, K. (1972) Cutaneous sensory projections to the avian forebrain. Brain Res 37:205222.Google Scholar
Derman, C.R., & Barbas, H. (1994) Contralateral thalamic projections predominantly reach transitional cortices in rhesus monkey. J Comp Neurol 344:508531.Google Scholar
Desfilis, E., Abellán, A., Sentandreu, V., & Medina, L. (2017) Expression of regulatory genes in the embryonic brain of a lizard and implications for understanding pallial organization and evolution. J Comp Neurol 526:166202.Google Scholar
Desfilis, E., Font, E., Belekhova, M., & Kenigfest, N. (2002) Afferent and efferent projections of the dorsal anterior thalamic nuclei in the lizard Podarcis hispanica (Sauria, Lacertidae). Brain Res Bull 57:477–450.Google Scholar
Desfilis, E., Font, E., & Garcia-Verdugo, J.M. (1998) Trigeminal projections to the dorsal thalamus in a lacertid lizard, Podarcis hispanica. Brain Behav Evol 52:99110.Google Scholar
Díaz, C., Yanes, C., Trujillo, C.M., & Puelles, L. (1994) The lacertidian reticular thalamic nucleus projects topographically upon the dorsal thalamus: experimental study in Gallotia galloti. J Comp Neurol 343:193208.Google Scholar
Diekamp, B., Gagliardo, A., & Güntürkün, O. (2002a) Nonspatial and subdivision-specific working memory deficits after selective lesions of the avian prefrontal cortex. J Neurosci 22:95739580.Google Scholar
Diekamp, B., Kalt, T., & Güntürkün, O. (2002b) Working memory neurons in pigeons. J Neurosci 22:15.Google Scholar
Dinopoulos, A. (1994) Reciprocal connections of the motor neocortical area with the contralateral thalamus in the hedgehog (Erinaceus europaeus) brain. Europ J Neurosci 6:374380.Google Scholar
Dirian, L., Galant, S., Coolen, M., Chen, W., Bedu, S., Houart, C., Bally-Cuif, L., & Foucher, I. (2014) Spatial regionalization and heterochrony in the formation of adult pallial neural stem cells. Dev Cell 30:123136.Google Scholar
Ditz, H.M., & Nieder, A. (2016) Sensory and working memory representations of small and large numerosities in the crow forebrain. J Neurosci 36:1204412052.CrossRefGoogle Scholar
Doron, N.N., & LeDoux, J.E. (2000) Cells in the posterior thalamus project to both amygdala and temporal cortex: a quantitative retrograde double-labeling study in the rat. J Comp Neurol 425:257274.Google Scholar
Dugas-Ford, J., & Ragsdale, C.W. (2015) Levels of homology and the problem of neocortex. Annu Rev Neurosci 38:351368.Google Scholar
Dugas-Ford, J., Rowell, J.J., & Ragsdale, C.W. (2012) Cell-type homologies and the origins of neocortex. Proc Nat’l Acad Sci 109:1697416979.Google Scholar
Durán, E., Ocaña, F.M., Broglio, C., Rodríguez, F., & Salas, C. (2010) Lateral but not medial telencephalic pallium ablation impairs the use of goldfish spatial allocentric strategies in a “hole-board” task. Behav Brain Res 214:480487.CrossRefGoogle ScholarPubMed
Ebbesson, S.O.E. (1980) The parcellation theory and its relation to interspecific variability in brain organization, evolutionary and ontogenetic development and neuronal plasticity. Cell Tiss Res 213:179212, doi:10.1007/BF00234781.CrossRefGoogle ScholarPubMed
Ebbesson, S.O.E. (1984) An update of the parcellation theory. Behav Brain Sci 7:350366, doi:10.1017/S0140525X00018628.Google Scholar
Ebbesson, S.O.E. (2020) How the parcellation theory of comparative forebrain specialization emerged from the Division of Neuropsychiatry at the Walter Reed Army Institute of Research. J Hist Neurosci 30:2455, doi:10.1080/0964704X.2020.1763759.Google Scholar
Emery, N.J., & Clayton, N.S. (2016) An avian perspective on stimulating other minds Learn Behav 44:203204.Google Scholar
Endepols, H., Roden, K., Luksch, H., Dicke, U., & Walkowiak, W. (2004) Dorsal striatopallidal system in anurans. J Comp Neurol 468:299310.Google Scholar
Evans, S.E. (2000) Amniote evolution. In: Bock, G.R., & Cardew, G. (Eds.), Evolutionary Developmental Biology of the Cerebral Cortex (Novartis Foundation Symposium 228). John Wiley & Sons, Ltd., Chichester, pp. 109113.Google Scholar
Fernández, M., Ahumada-Galleguillos, P., Marían, G., & Mpodozis, J. (2019) Intratelencephalic projections of the avian visual dorsal ventricular ridge: laminarly segregated, reciprocally and topographically organized. J Comp Neurol 413:2654.Google Scholar
Folgueira, M., Bayley, P., Navratilova, P., Becker, T.S., Wilson, S.W., & Clarke, J.D.W. (2012) Morphogenesis underlying the development of the everted teleost telencephalon. Neural Dev 7:32.Google Scholar
García-Moreno, F., & Molnár, Z. (2020) Variations of telencephalic development that paved the way for neocortical evolution. Progr Neurobiol 194:101865, doi:10.1016/j.pneurobio.2020.101865.CrossRefGoogle ScholarPubMed
Gilbert, S.R. (1994) Developmental Biology, 4th ed. Sinauer Associates, Inc., Sunderland MA.Google Scholar
Gómez-Laplaza, L.M., & Gerlai, R. (2020) Food density and preferred quantity: discrimination of small and large numbers in angelfish (Pterophyllum scalare). Sci Rep 9:15305. doi:10.1007/s10071-0C20-01355-6.Google Scholar
González, G., Puelles, L., & Medina, L. (2002). Organization of the mouse dorsal thalamus based on topology, calretinin immunostaining, and gene expression. Brain Res Bull 57:439442.Google Scholar
González, M.J., Yáñez, J., and Anadón, R. (1999) Afferent and efferent connections of the torus semicircularis in the sea lamprey: an experimental study. Brain Res 826:8394.Google Scholar
Griffiths, D., Dickinson, A., & Clayton, N. (1999) Episodic memory: what can animals remember about their past? Trends Cogn Sci 3:7480.Google Scholar
Grosenick, L., Clement, T.S., & Fernald, R.D. (2007) Fish can infer social rank by observation alone. Nature 445:429432.Google Scholar
Gruber, R., Schiesti, M., Boeckle, M., Frohnwieser, A., Miller, R., Gray, R.D., Clayton, N.S., & Taylor, A.H. (2019) New Caledonian crows use mental representations to solve metatool problems. Curr Biol 29:686692.Google Scholar
Guillén, M. (1991) Estructura del epitálamo y complejo superior del talámo dorsal en aves: studio embryológico. Posibles homologies con mamíferos. PhD thesis, Universidad de Murcia.Google Scholar
Güntürkün, O. (2020) The surprising power of the avian mind. Sci Am 322:4855.Google Scholar
Halassa, M.M., & Sherman, S.M. (2019) Thalamocortical circuit motifs: a general framework. Neuron 103:762770.CrossRefGoogle ScholarPubMed
Hall, B.K. (1992) Evolutionary Developmental Biology. Chapman & Hall, London.Google Scholar
Hall, J.A., Foster, R.E., Ebner, F.F., & Hall, W.C. (1977) Visual cortex in a reptile, the turtle (Pseudemys scripta and Chrysemys picta). Brain Res 130:197216.Google Scholar
Harting, J.K., Hall, W.C., & Diamond, I.T. (1972) Evolution of the pulvinar. Brain Behav Evol 6:424452.Google Scholar
Harting, J.K., Updyke, B.V., & van Lieshout, D.P. (2001) Striatal projections from the cat visual thalamus. Europ J Neurosci 14:893896.Google Scholar
Heinrich, B. (1999) The Mind of a Raven. Harper Collins Publishers, New York.Google Scholar
Henderson, J., Hurly, T.A., Bateson, M., & Healy, S.D. (2006) Timing in free-living rufous hummingbirds, Selasphorus rufus. Curr Biol 16:512515.Google Scholar
Herculano-Houzel, S. (2020) Birds do have a brain cortex—and think. Sci 369:15671568.Google Scholar
Herkenham, M. (1978) The connections of the nucleus reuniens thalami: evidence for a direct thalamo-hippocampal pathway in the rat. J Comp Neurol 177:589610.Google Scholar
Herrick, J.L., & Keifer, J. (1997) A hypothalamic projection to the turtle red nucleus: An anterograde and retrograde tracing study. Exp Brain Res 116:556560.Google Scholar
Hodos, W., & Campbell, C.B.G. (1969) Scala naturae: why there is no theory in comparative psychology. Psychol Rev 76:337350.Google Scholar
Hofmann, M.H., & Northcutt, R.G. (2012) Forebrain organization in elasmobranchs. Brain Behav Evol 80:142151.Google Scholar
Hollis, D.M., & Boyd, S.K. (2005) Distribution of GABA-like immunoreactive cell bodies in the brains of two amphibians, Rana catesbeiana and Xenopus laevis. Brain Behav Evol 65:127142.Google Scholar
Hoogland, P.V. (1977) Efferent connections of the striatum in Tupinambis nigropunctatus. J Morphol 152:229246.Google Scholar
Hoogland, P.V., & Vermeulen-Van der Zee, (1989) Efferent connections of the dorsal cortex of the lizard Gekko gecko studied with Phaseolus vulgaris-leucoagglutinin. J Comp Neurol 285:289303.Google Scholar
Hotta, T., Ueno, K., Hataji, Y., Kuroshima, H., Fujita, K., & Kohda, M. (2020) Transitive inference in cleaner wrasses (Labroides dimidiatus). PLoSONE 15(8): e0237817. doi:10.1371/journal.pone.0237817.Google Scholar
Hu, J., & Wang, S.-R. (2001) Firing patterns and morphological features of neurons in the pigeon nucleus rotundus. Brain Behav Evol 57:343348.Google Scholar
Husband, S., & Shimizu, T. (1999) Efferent projections of the ectostriatum in the pigeon (Columba livia). J Comp Neurol 406:329345.Google Scholar
Iwaniuk, A.N. (2017) The evolution of cognitive brains in non-mammals. In: Watanabe, S., Hofman, M.A., & Shimizu, T. (Eds.), Evolution of the Brain, Cognition, and Emotion in Vertebrates. Springer Japan KK, Tokyo, pp. 101124.Google Scholar
Jarvis, E.D., Yu, J., Rivas, M.V., Horita, H., Feenders, G., Whitney, O., Jarvis, S.C., Jarvis, E.R., Kubikova, L., Puck, A.E.P., Siang-Bakshi, C., Martin, S., McElroy, M., Hara, E., Howard, J., Pfenning, A., Mouritsen, H., Chen, C.-C., & Wada, K. (2013) Global view of the functional molecular organization of the avian cerebrum: mirror images and functional columns. J Comp Neurol 521:36143665.Google Scholar
Jiao, Y., Medina, L., Veenman, L.C., Toledo, C., Puelles, L., & Reiner, A. (2000) Identification of the anterior nucleus of the ansa lenticularis in birds as the homologue of the mammalian subthalamic nucleus. J Neurosci 20:69987010.Google Scholar
Jones, E.G. (1985) The Thalamus. Plenum Press, New York.Google Scholar
Jones, E.G. (2007) The Thalamus, 2nd ed. Cambridge University Press, Cambridge, UK.Google Scholar
Kaas, J.H. (2017a) Evolution of the visual cortex in primates. In: Kaas, J.H., & Krubitzer, L. (Eds.), Evolution of Nervous Systems. Vol. 3: The Nervous Systems of Non-human Primates, 2nd ed. Elsevier, Amsterdam, pp. 187201.Google Scholar
Kaas, J.H. (2017b) The organization of neocortex in early mammals. In: Kaas, J.H., & Herculano-Houzel, S. (Eds.), Evolution of Nervous Systems. Vol. 2: The Nervous Systems of Early Mammals and Their Evolution, 2nd ed.Elsevier, Amsterdam, pp. 87101.Google Scholar
Karten, H.J. (1968) The ascending auditory pathways in the pigeon (Columba livia): II. Telencephalic projections of the nucleus ovoidalis thalami. Brain Res 11:134153.Google Scholar
Karten, H.J. (1969) The organization of the avian telencephalon and some speculations on the phylogeny of the amniote telencephalon. Ann NY Acad Sci 167:164179.Google Scholar
Karten, H.J., Cox, K., & Mpodozis, J. (1997) Two distinct populations of tectal neurons have unique connections within the retinotectorotundal pathway of the pigeon (Columba livia). J Comp Neurol 387: 449465.Google Scholar
Karten, H.J., & Hodos, W. (1970) Telencephalic projections of the nucleus rotundus in the pigeon (Columba livia). J Comp Neurol 140:3552.CrossRefGoogle ScholarPubMed
Karten, H.J., Hodos, W., Nauta, W.H.J., & Revzin, A. (1973) Neural connections of the “visual Wulst” of the avian telencephalon: experimental studies in the pigeon (Columba livia) and owl (Speotyto cunicularia). J Comp Neurol 150:253278.Google Scholar
Karten, H.J., Konishi, M., & Pettigrew, J.D. (1978) Somatosensory representation in the anterior Wulst of the owl Speotyto cunicularis. Soc Neurosci Abstr 4:554.Google Scholar
Keifer, J., & Lustig, D.G. (2000) Comparison of cortically and subcortically controlled motor systems. II. Distribution of anterogradely labeled terminal boutons on intracellularly filled rubrospinal neurons. J Comp Neurol 416:101111.Google Scholar
Kemp, T.S. (2005) The Origin and Evolution of Mammals. Oxford University Press, Oxford, UK.Google Scholar
Kenigfest, N., Belekhova, M., Repérant, J., Rio, J.-P., Ward, R., & Vesselkin, N. (2005) The turtle thalamic anterior entopeduncular nucleus shares connectional and neurochemical characteristics with the mammalian thalamic reticular nucleus. J Chem Neuroanat 30:129143.Google Scholar
Kenigfest, N., Repérant, J., Belekhova, M., Rio, J.-P., & Ward, R. (2007) Evolution of the visual tectogeniculate and pretectogeniculate pathways in the brain of amniote vertebrates. In: Kaas, J.H., & Bullock, T.H. (Eds.), Evolution of Nervous Systems: A Comprehensive Reference. Vol. 2, Nonmammalian Vertebrates. Elsevier, Amsterdam, pp. 459467.Google Scholar
Kirsch, J.A., Güntürkün, O., & Rose, J. (2008) Insight without cortex: lessons from the avian brain. Consc Cogn 17:475483.Google Scholar
Kohda, M., Hotta, T., Takeyama, T., Awata, S., Tanaka, H., Asai, J., & Jordan, A.L. (2019) If a fish can pass the mark test, what are the implications for consciousness and self-awareness testing in animals? PLoS Biol 17: e3000021. doi:10.1371/journal.pbio.3000021.Google Scholar
Krubitzer, L., & Seelke, A.M.H. (2012) Cortical evolution in mammals: the bane and beauty of phenotypic variability. Proc Nat Acad Sci USA 109, Suppl. 1:1064710654.Google Scholar
Krusche, P., Uller, C., & Dicke, U. (2010) Quantity discrimination in salamanders. J Exp Biol 213:18221828.CrossRefGoogle ScholarPubMed
Kudo, M., Glendenning, K.K., Frost, S.B., & Masterton, R.B. (1986) Origin of mammalian thalamocortical projections. I. Telencephalic projections of the medial geniculate body in the opossum (Didelphis virginiana). J Comp Neurol 245:176197.Google Scholar
Laberge, F., Mühlenbrock-Lenter, S., Dicke, U., & Roth, G. (2008) Thalamo-telencephalic pathways in the fire-bellied toad Bombina orientalis. J Comp Neurol 508:806823.Google Scholar
Lichtneckert, R., & Reichert, H. (2005) Insights into the urbilaterian brain: conserved genetic patterning mechanisms in insect and vertebrate brain development. Heredity 94:465477.Google Scholar
Liem, K.F., Bemis, W.E., Walker, W.F., Jr., & Grande, L. (2001) Functional Anatomy of the Vertebrates: An Evolutionary Perspective. Harcourt College Publishers, Fort Worth.Google Scholar
Llinás, R., & Paré, D. (1991) Coherent oscillations in specific and nonspecific thalamocortical networks and their role in cognition. In Steriade, M., Jones, E.G., & McCormick, D.A. (Eds.), Thalamus. Vol. II: Experimental and Clinical Aspects. Elsevier, Amsterdam, pp. 501516.Google Scholar
Llinás, R., Ribary, U., Contreras, D., & Pedroarena, C. (1998) The neuronal basis for consciousness. Phil Trans Roy Soc Lond 353:1841–1849.Google Scholar
Llinás, R.R., & Steriade, M. (2006) Bursting of thalamic neurons and states of vigilance. J Neurophysiol 95:32973308.Google Scholar
Llinás, R.R., Urbano, F.J., Leznik, E., Ramírez, R.R., & van Marle, H.J.F. (2005) Rhythmic and dysrhythmic thalamocortical dynamics: GABA systems and the edge effect. Trends Neurosci 28:325333.Google Scholar
Lohman, A.H.M., & Van Woerden-Verkley, (1978) Ascending connections to the forebrain in the tegu lizard. J Comp Neurol 182:555594.Google Scholar
López, J.M., Morona, R., Moreno, N., & González, A. (2017) The organization of the central nervous system of lungfishes: an immunohistochemical approach. In: Kaas, J.H., & Striedter, G. (Eds.), Evolution of Nervous Systems. Vol. 1: The Evolution of the Nervous Systems in Nonmammalian Vertebrates, 2nd ed. Elsevier, Amsterdam, pp. 121139.Google Scholar
López-Bendito, G., Flames, N., Ma, L., Fouquet, C., Di Medlio, A., Tessier-Lavigne, M., & Marín, O. (2007) Robo1 and Robo2 cooperate to control the guidance of major axonal tracts in the mammalian forebrain. J Neurosci 27:33953407.Google Scholar
López-Bendito, G., & Molnár, Z. (2003) Thalamocortical development: how are we going to get there? Nat Rev Neurosci 4:276289.Google Scholar
Luksch, H., Cox, K., & Karten, H.J. (1998) Bottlebrush dendritic endings and large dendritic fields: motion-detecting neurons in the tectofugal pathway. J Comp Neurol 396:399414.Google Scholar
Lynn, A.M., Schneider, D.A., & Bruce, L.L. (2015) Development of the avian dorsal thalamus: patterns and gradients of neurogenesis. Brain Behav Evol 86:94109.Google Scholar
Macchi, G., Bentivoglio, M., Minciacchi, D., & Molinari, M. (1996) Trends in the anatomical organization and functional significance of the mammalian thalamus. Ital J Neurol Sci 17:105129.Google Scholar
Mancilla, J.G., & Ulinski, P.S. (2001) Role of GABA(A)-mediated inhibition in controlling the responses of regular spiking cells in turtle visual cortex. Vis Neurosci 18:924.Google Scholar
Manrod, J.D., Hartdegen, R., & Burghardt, G.M. (2008) Rapid solving of a problem apparatus by juvenile black-throated monitor lizards (Varanus albigularis albigularis). Anim Cogn 11:267273.CrossRefGoogle ScholarPubMed
Martínez-de-la-Torre, M. (1985) Estructura del mesencéfalo y diencéfalo en aves y reptiles: aportaciones a una síntesis en la búsqueda de homologías. PhD thesis, Universidad de Murcia.Google Scholar
Martínez-de-la-Torre, M., Pombol, M.A., & Puelles, L., (2011) Distal-less-like protein distribution in the larval lamprey forebrain. Neurosci 178:270284.Google Scholar
Martínez-García, F., Martínez-Marcos, A., & Lanuza, E. (2002) The pallial amygdala of amniote vertebrates: evolution of the concept, evolution of the structure. Brain Res Bull 57:463–4Google Scholar
Martínez-García, F., Novejarque, A., & Lanuza, E. (2007) Evolution of the amygdala in vertebrates. In: Kaas, J.H., & Bullock, T.H. (Eds.), Evolution of Nervous Systems: A Comprehensive Reference. Vol. 2, Nonmammalian Vertebrates. Elsevier, Amsterdam, pp. 255334.Google Scholar
Medina, L., & Abellán, A. (2009) Development and evolution of the pallium. Sem Cell Devel Biol 20:698711.Google Scholar
Medina, L., Abellán, A., Vicario, A., Castro-Robles, B., & Desfilis, E. (2017) The amygdala. In: Kaas, J.H., & Striedter, G. (Eds.), Evolution of Nervous Systems. Vol. 1: The Evolution of the Nervous Systems in Nonmammalian Vertebrates, 2nd ed.Elsevier, Amsterdam, pp. 427478.Google Scholar
Medina, L., Veenman, C.L., & Reiner, A. (1997). Evidence for a possible avian dorsal thalamic region comparable to the mammalian ventral anterior, ventral lateral, and oral ventroposterolateral nuclei. J Comp Neurol 384:86108.Google Scholar
Médina, M., Repérant, J., Dufour, S., Ward, R., Le Belle, N., & Miceli, D. (1994) The distribution of GABA-immunoreactive neurons in the brain of the silver eel (Anguilla anguilla L.). Anat Embryol 189:2539.Google Scholar
Meléndez-Ferro, M., Pérez-Costas, E., Villar-Cheda, , Abalo, X.M., Rodriguez-Muñoz, R., Rodicio, M.C., & Anadón, R. (2002) Ontogeny of γ-aminobutyric acid-immunoreactive neuronal populations in the forebrain and midbrain of the sea lamprey. J Comp Neurol 446:360376.Google Scholar
Molnár, Z., Garel, S., López-Bendito, G., Maness, P., & Price, D.J. (2012) Mechanisms controlling the guidance of thalamocortical axons through the embryonic forebrain. Europ J Neurosci 35:15731585.CrossRefGoogle ScholarPubMed
Montagnese, C.M., Mezey, S.E., & Csillag, A. (2003) Efferent connections of the dorsomedial thalamic nuclei of the domestic chick (Gallus domesticus). J Comp Neurol 459:301326.Google Scholar
Montiel, J.F., & Aboitiz, F. (2015) Pallial patterning and the origin of the isocortex. Front Neurosci 9. doi: 10.3389/fnins.2015.00377.Google Scholar
Moreno, N., Bachy, I., Rétaux, S., & González, A. (2004) LIM-homeodomain genes as developmental and adult genetic markers of Xenopus forebrain functional subdivisions. J Comp Neurol 472:5272.Google Scholar
Moreno, N., Morona, R., López, J.M., and González, A. (2017) The diencephalon and hypothalamus of nonmammalian vertebrates: evolutionary and developmental traits. In: Kaas, J.H., & Striedter, G. (Eds.), Evolution of Nervous Systems. Vol. 1: The Evolution of the Nervous Systems in Nonmammalian Vertebrates, 2nd ed.Elsevier, Amsterdam, pp. 409426.Google Scholar
Morrow, J., Mosher, C., & Gothard, K. (2019) Multisensory neurons in the primate amygdala. J Neurosci 39:36633675.Google Scholar
Mpodozis, J., Cox, K., Shimizu, T., Bischoff, H.J., Woodson, W., & Karten, H.J. (1996) GABAergic inputs to the nucleus rotundus (pulvinar inferior) of the pigeon (Columba livia). J Comp Neurol 374:204222.Google Scholar
Nakagawa, Y., & O’Leary, D.D.M. (2001) Combinatorial expression patterns of LIM-homeodomain and other regulatory genes parcellate developing thalamus. J Neurosci 21:27112725.Google Scholar
Nelson, J.S. (1994) Fishes of the World, 3rd ed., John Wiley & Sons, New York.Google Scholar
Nieder, A., Wagener, L., & Rinnert, P. (2020) A neural correlate of sensory consciousness in a corvid bird. Sci 369:16261629.Google Scholar
*Nieuwenhuys, R., ten Donkelaar, H.J., & Nicholson, C. (1998) The Central Nervous System of Vertebrates, Vol. 1. Springer-Verlag, Berlin.Google Scholar
*Nieuwenhuys, R., ten Donkelaar, H.J., & Nicholson, C. (1998) The Central Nervous System of Vertebrates. Springer-Verlag, Berlin.CrossRefGoogle Scholar
Nomura, T., & Hirata, T. (2017) The neocortical homologues in nonmammalian amniotes: bridging the hierarchical concepts of homology through comparative neurogenesis. In: Kaas, J.H., & Herculano-Houzel, S. (Eds.), Evolution of Nervous Systems. Vol. 2: The Nervous Systems of Early Mammals and Their Evolution, 2nd ed.Elsevier, Amsterdam, pp. 195204.Google Scholar
Northcutt, R.G. (1977) Retinofugal projections in the lepidosirenid lungfishes. J Comp Neurol 174:553574.Google Scholar
Northcutt, R.G. (1980) Retinal projections in the Australian lungfish. Brain Res 185:8590.Google Scholar
Northcutt, R.G. (1984) Evolution of the vertebrate central nervous system: patterns and processes. Amer Zool 24:701716.Google Scholar
Northcutt, R.G. (1985) Central nervous system phylogeny: evaluation of hypotheses. Fortsch Zool 30:497505.Google Scholar
Northcutt, R.G. (1989) Brain variation and phylogenetic trends in elasmobranch fishes. J Exp Zool Suppl 2:83100.Google Scholar
Northcutt, R.G. (2003) The use and abuse of developmental data. Behav Brain Sci 26:565566.Google Scholar
Northcutt, R.G. (2006) Connections of the lateral and medial divisions of the goldfish telencephalic pallium. J Comp Neurol 494:903943.Google Scholar
Northcutt, R.G. (2009) Telencephalic organization in the spotted African lungfish, Protopterus dolloi: a new cytological model. Brain Behav Evol 73:5980.Google Scholar
Northcutt, R.G. (2013) Variations in reptilian brains and cognition. Brain Behav Evol 82:4554.Google Scholar
Northcutt, R.G., & Wicht, H. (1997) Afferent and efferent connections of the lateral and medial pallia of the silver lamprey. Brain Behav Evol 49:119.Google Scholar
Novejarque, A., Lanuza, E., & Martínez-Garcia, F. (2004) Amygdalostriatal projections in reptiles: a tract-tracing study in the lizard Podarcis hispanica. J Comp Neurol 479:287308.Google Scholar
Okano, H., & Temple, S. (2009) Cell types to order: temporal specification of CNS stem cells. Curr Opin Neurobiol 19:112119.Google Scholar
Olkowicz, S., Kocourek, M., Lučan, R.K., Porteš, M., Fitch, W.T., Herculano-Houzel, S., & Němec, P. (2016) Birds have primate-like numbers of neurons in the forebrain. Proc Nat Acad Sci USA 113:72557260.Google Scholar
Panchen, A.L. (1999) Homology—history of a concept. In: Bock, G.R., & Cardew, G. (Eds.), Homology: Novartis Foundation Symposium 222. John Wiley & Sons, Chichester.Google Scholar
Pasko, L., (2010) Tool-like behavior in the sixbar wrasse, Thalassoma hardwicke (Bennett, 1830). Zool Biol 29:767773.Google Scholar
Patton, B.W., & Braithwaite, V.A. (2015) Changing tides: ecological and historical perspectives on fish cognition. WIREs Cogn Sci 6:159176. doi:10.1002/wcs.1337.Google Scholar
Paz-y-Miño, G., Bond, A.B., Kamil, A.C., & Balda, R.P. (2004) Pinyon jays use transitive inference to predict social dominance. Nature 430:778781.Google Scholar
Peake, T.M., Terry, A.M.R., McGregor, P.K., & Dabelsteen, T. (2001) Male great tits eavesdrop on simulated male-to-male vocal interactions. Proc Roy Soc Lond B 268:11831187.Google Scholar
Peake, T.M., Terry, A.M.R., McGregor, P.K., & Dabelsteen, T. (2002) Do great tits assess rivals by combining direct experience with information gathered by eavesdropping? Proc Roy Acad Lond B 269:19251929.Google Scholar
Pepperberg, I.M. (1999) The Alex Studies: Cognitive and Communicative Abilities of Grey Parrots. Harvard University Press, Cambridge, MA.Google Scholar
Pepperberg, I.M. (2006) Grey parrot (Psittacus erithacus) numerical abilities: addition and further experiments on a zero-like concept. J Comp Psychol 120:111.Google Scholar
Pepperberg, I.M. (2013) Abstract concepts: data from a grey parrot. Behav Proc 93:8290.Google Scholar
Pepperberg, I.M. (2020) The comparative psychology of intelligence: some thirty years later. Front Psychol 11. doi:10.3389/fpsyg.2020.00973.Google Scholar
Pepperberg, I.M., & Nakayama, K. (2016) Robust representation of shape in a grey parrot (Psittacus erithacus). Cogn 153:146160.Google Scholar
Polenova, O.A., & Vesselkin, N.P. (1993) Olfactory and nonolfactory projections in the river lamprey (Lampetra fluviatilis) telencephalon. J Hirnforsch 34:261279.Google Scholar
Pombal, M.A., & Megías, M. (2017) The nervous system of jawless vertebrates. In: Kaas, J.H., & Striedter, G. (Eds.), Evolution of Nervous Systems. Vol. 1: The Evolution of the Nervous Systems in Nonmammalian Vertebrates, 2nd ed. Elsevier, Amsterdam, pp. 3757.Google Scholar
Pombal, M.A., Megías, M., Bardet, S.M., & Puelles, L. (2009) New and old thoughts on the segmental organization of the forebrain in lampreys. Brain Behav Evol 74:719.Google Scholar
Portavella, M., Torres, B., & Salas, C. (2004) Avoidance response in goldfish: emotional and temporal involvement of medial and lateral telencephalic pallium. J Neurosci 24:23352342.Google Scholar
Powers, A.S., & Reiner, A. (1980) A stereotaxic atlas of the forebrain and midbrain of the Eastern painted turtle (Chrysemys picta picta). J Hirnforsch 21:125159.Google Scholar
Preuss, T.M., & Goldman-Rakic, P.S. (1987) Crossed corticothalamic and thalamocortical connections of macaque prefrontal cortex. J Comp Neurol 257:269281.Google Scholar
Pritz, M.B. (1995) The thalamus of reptiles and mammals: similarities and differences. Brain Behav Evol 46:197208.Google Scholar
Pritz, M.B. (2015) Crocodilian forebrain: evolution and development. Integrative Comp Biol 55:949–961.Google Scholar
Pritz, M.B., & Stritzel, M.E. (1988) Thalamic nuclei that project to reptilian telencephalon lack GABA and GAD immunoreactive neurons and puncta. Brain Res 457:154159.Google Scholar
Pritz, M.B., & Stritzel, M.E. (1990) A different type of vertebrate thalamic organization. Brain Res 525:330334.Google Scholar
Pritz, M.B., & Stritzel, M.E. (1994) Glutamic acid decarboxylase immunoreactivity in some dorsal thalamic nuclei in Crocodilia. Neurosci Lett 165:109112.Google Scholar
Puelles, L., (2011) Pallio-pallial tangential migrations and growth signaling: new scenario for cortical evolution? Brain Behav Evol 78:108127.Google Scholar
Puelles, L., Amat, J.A., & Martínez-de-la-Torre, M. (1987) Segment-related, mosaic neurogenetic pattern in the forebrain and mesencephalon of early chick embryos: I. Topography of AChE-positive neuroblasts up to stage HH18. J Comp Neurol 266: 247268.Google Scholar
Puelles, L., Ayad, A., Alonso, A., Sandoval, J.E., Martínez-de-la-Torre, M., Medina, L., & Ferran, J.L. (2016) Selective early expression of the orphan nuclear receptor Nr4a2 identifies the claustrum homolog in the avian mesopallium: impact on sauropsidian/mammalian pallium comparisons. J Comp Neurol 524:665703.Google Scholar
Puelles, L., Kuwana, E., Puelles, E., Bulfone, A., Shimamura, K., Keleher, J., Smiga, S., & Rubenstein, J.L. (2000) Pallial and subpallial derivatives of the embryonic chick and mouse telencephalon, traced by the expression of genes Dlx-2, Emx-1, Nkx-2.1, Pax-6, and TBR-1. J Comp Neurol 424:409438.Google Scholar
Puelles, L., & Medina, L. (2002) Field homology as a way to reconcile genetic and developmental variability with adult homology. Brain Res Bull 57:243255.Google Scholar
Puelles, L., Medina, L., Borello, U., Legaz, I., Teissier, A., Pierani, A., & Rubenstein, J.L.R. (2015) Radial derivatives of the mouse ventral pallium traced with Dbx1-LacZ reporters. J Chem Neuroanat 75:219.Google Scholar
Puelles, L., Milán, F.J., & Martínez-de-la-Torre, M. (1996) A segmental map of architectonic subdivisions in the diencephalon of the frog Rana perezi: acetylcholinesterase-histochemical observations. Brain Behav Evol 47:279310.Google Scholar
Puelles, L., Sandoval, J.E., Ayad, A., del Corral, R., Alonso, A., Ferran, J.L., & Martínez-de-la-Torre, (2017) The pallium in reptiles and birds in the light of the updated tetrapartite pallium model. In: Kaas, J.H., & Striedter, G. (Eds.), Evolution of Nervous Systems. Vol. 1: The Evolution of the Nervous Systems in Nonmammalian Vertebrates, 2nd ed.Elsevier, Amsterdam, pp. 519555.CrossRefGoogle Scholar
Raff, R.A. (1996) The Shape of Life: Genes, Development, and the Evolution of Animal Form. University of Chicago Press, Chicago.CrossRefGoogle Scholar
Raghanti, M.A., Munger, E.L., Wicinski, B., Butti, C., & Hof, P.R. (2017) Comparative structure of the cerebral cortex in large mammals. In: Kaas, J.H., & Herculano-Houzel, S. (Eds.), Evolution of Nervous Systems. Vol. 2: The Nervous Systems of Early Mammals and Their Evolution, 2nd ed. Elsevier, Amsterdam, pp. 267289.Google Scholar
Redies, C., Ast, M., Nakagawa, S., Takeichi, M., Martínez-de-la-Torre, M., & Puelles, L. (2000) Morphogenetic fate of diencephalic prosomeres and their subdivisions revealed by mapping cadherin expression. J Comp Neurol 421:481514.Google Scholar
Reiner, A. (1993) Neurotransmitter organization and connections of turtle cortex: implications for the evolution of mammalian isocortex. Comp Biochem Physiol A 104:735748.Google Scholar
Reiner, A. (2002) Functional circuitry of the avian basal ganglia: implications for basal ganglia organization in stem amniotes. Brain Res Bull 57:513528.CrossRefGoogle ScholarPubMed
Reiner, A. (2012) You are who you talk with—a commentary on Dugas-Ford et al. PNAS, 2012. Brain Behav Evol 81:146149 doi:10.1159/000348281.Google Scholar
Reiner, A., Medina, L., & Veenman, L.C. (1998) Structural and functional evolution of the basal ganglia in vertebrates. Brain Res Rev 28:235285.Google Scholar
Reiner, A., Perkel, D.J., Bruce, L.L., Butler, A.B., Csillag, A., Kuenzel, W., Medina, L., Paxinos, G., Shimizu, T., Striedter, G., Wild, M., Ball, G.F., Durand, S., Güntürkün, O., Lee, D.W., Mello, C.V., Powers, A., White, S.A., Hough, G., Kubikova, L., Smulders, T.V., Wada, K., Dugas-Ford, J., Husband, S., Yamamoto, K., Yu, J., Siang, C., & Jarvis, E.D. (2004) Avian Brain Nomenclature Forum. Revised nomenclature for avian telencephalon and some related brainstem nuclei.J Comp Neurol 473:377414.CrossRefGoogle ScholarPubMed
Reiner, A., Yamamoto, K., & Karten, H.J. (2005) Organization and evolution of the avian forebrain. Anat Rec 287A:10801102.Google Scholar
Reisz, R.R. (1997) The origin and early evolutionary history of amniotes. Trends Ecol Evol 12:218222.Google Scholar
Rikhye, R.V., Wimmer, R.D., & Halassa, M.M. (2018) Toward an integrative theory of thalamic function. Annu Rev Neurosci 41:163183. doi:10.1146/annurev-neuro-080317-062144.Google Scholar
Rink, E., & Wullimann, M.F. (2004) Connections of the ventral telencephalon (subpallium) in the zebrafish (Danio rerio). Brain Res 1011:206220.Google Scholar
Rinnert, P., Kirschhock, M.E., & Nieder, A. (2019) Neuronal correlates of spatial working memory in the endbrain of crows. Curr Biol 29:26162624.Google Scholar
Rio, J.-P., Repérant, J., Ward, R., Miceli, D., & Medina, M. (1992) Evidence of GABA-immunopositive neurons in the dorsal part of the lateral geniculate nucleus of reptiles: morphological correlates with interneurons. Neurosci 47:395407.Google Scholar
Rodríguez, F., López, J.C., Vargas, J.P., Broglio, C., Gómez, Y., & Salas, C. (2002) Spatial memory and hippocampal pallium through vertebrate evolution: insights from reptiles and teleost fish. Brain Res Bull 57:499503.Google Scholar
Rodríguez-Moldes, I., Santos-Durán, G.N., Pose-Méndez, S., Quintana-Urzainqui, I., & Candal, E. (2017) The brains of cartilaginous fishes. In: Kaas, J.H., & Striedter, G. (Eds.), Evolution of Nervous Systems. Vol. 1: The Evolution of the Nervous Systems in Nonmammalian Vertebrates, 2nd ed.Elsevier, Amsterdam, pp. 7797.CrossRefGoogle Scholar
Rose, J.E. (1942) The ontogenetic development of the rabbit’s diencephalon. J Comp Neurol 77:61129.Google Scholar
Roth, G., Grunwald, W., & Dicke, U. (2003) Morphology, axonal projection pattern, and responses to optic nerve stimulation of thalamic neurons in the fire-bellied toad Bombina orientalis. J Comp Neurol 461:91110.Google Scholar
Roth, G., Laberge, F., Mühlenbrock-Lenter, S., & Grunwald, W. (2007) Organization of the pallium in the fire-bellied toad Bombina orientalis. I: morphology and axonal projection pattern of neurons revealed by intracellular biocytin labeling. J Comp Neurol 501:443464.Google Scholar
Roth, T.C. II, Krochmal, A.R., & LaDage, L.D. (2019) Reptilian cognition: a more complex picture via integration of neurological mechanisms, behavioral constraints, and evolutionary context. BioEssays 41. doi:10.1002/bies.201900033.Google Scholar
Rowe, T.B. (2017) The emergence of mammals. In: Kaas, J.H., & Herculano-Houzel, S. (Eds.), Evolution of Nervous Systems. Vol. 2: The Nervous Systems of Early Mammals and Their Evolution, 2nd ed.Elsevier, Amsterdam, pp. 152.Google Scholar
Sherman, S.M., & Guillery, R.W. (2013) Functional Connections of Cortical Areas: A New View from the Thalamus. MIT Press, Cambridge, MA.Google Scholar
Shi, W., Xianyu, A., Han, Z., Tang, X., Li, Z., Zhong, H., Mao, T., Huang, K., & Shi, S.-H. (2017) Ontogenetic establishment of order-specific nuclear organization in the mammalian thalamus. Nat Neurosci 20:516528.Google Scholar
Simpson, G.G. (1961) Principles of Animal Taxonomy. Columbia University Press, New York.Google Scholar
Smith, H. (1967) Biological similarities and homologies. Syst Biol 16:101102.Google Scholar
Smith-Fernandez, A., Pieau, C., Repérant, J., Boncinelli, E., & Wassef, M. (1998) Expression of the Emx-1 and Dlx-1 homeobox genes define three molecularly distinct domains in the telencephalon of mouse, chick, turtle and frog embryos: implications for the evolution of telencephalic subdivisions in amniotes. Devel 2111:20992111.Google Scholar
Sovrano, V.A., & Bisazza, A. (2008) Recognition of partly occluded objects by fish. Anim Cogn 11:161166.Google Scholar
Sovrano, V.A., & Bisazza, A. (2009) Perception of subjective contours in fish. Percep 38:579590.Google Scholar
Stacho, M., Herold, C., Rook, N., Wagner, H., Axer, M., Amunts, K., & Gunturkun, O. (2020) A cortex-like canonical circuit in the avian forebrain. Sci 369:eabc5534.Google Scholar
Stancher, G., Rugani, R., Regolin, L., & Vallortigara, G. (2015) Numerical discrimination by frogs (Bombina orientalis). Anim Cogn 18:219229.Google Scholar
Stein, B.E., & Meredith, M.A. (1993) The Merging of the Senses. MIT Press, Cambridge, MA.Google Scholar
Steriade, M., Jones, E.G., & McCormick, D.A. (1997) Thalamus. Vol. 1: Organization and Function. Elsevier, Amsterdam.Google Scholar
Steriade, M., & Paré, D. (2007) Gating in Cerebral Networks. Cambridge University Press, Cambridge, UK.CrossRefGoogle Scholar
Striedter, G. (1991) Auditory, electrosensory, and mechanosensory lateral line pathways through the forebrain in channel catfishes. J Comp Neurol 312:311331.Google Scholar
Striedter, G.F. (1999) Homology in the nervous system: of characters, embryology and levels of analysis. In: Bock, G.R., & Cardew, G. (Eds.), Homology (Novartis Foundation Symposium 222). John Wiley & Sons, Chichester, UK, pp. 158172.Google Scholar
Striedter, G.F., & Northcutt, R.G. (2020) Brains through Time: A Natural History of Vertebrates. Oxford University Press, Oxford, UK.Google Scholar
Ströhmann, B., Schwarz, D.W.F., & Puil, E. (1994) Mode of firing and rectifying properties of nucleus ovoidalis neurons in the avian auditory thalamus. J Neurophysiol 71:13511360.Google Scholar
Suárez, J., Andreu, M.J., Heredia, R., Dávila, J.C., & Guirado, S. (2002) A putative striato-dorsal thalamic pathway in lizards. Brain Res Bull 57:533535.Google Scholar
Suárez, J., Dávila, J.C., Real, M.Á., & Guirado, S. (2005) Distribution of GABA, calbindin and nitric oxide synthase in the developing chick entopallium. Brain Res Bull 66:441444.Google Scholar
Suárez, J., Dávila, J.C., Real, M.Á., Guirado, S., & Medina, L. (2006) Calcium-binding proteins, neuronal nitric oxide synthase, and GABA help to distinguish different pallial areas in the developing and adult chicken. I. Hippocampal formation and hyperpallium. J Comp Neurol 497:751771.CrossRefGoogle ScholarPubMed
Sugahara, F., Murakami, Y., Adachi, N., & Kuratani, S. (2013) Evolution of the regionalization and patterning of the vertebrate telencephalon: what can we learn from cyclostomes? Curr Opin Genet Devel 23:475483.Google Scholar
Sugahara, F., Murakami, Y., Pascual-Anaya, J., & Kuratani, S. (2017) Reconstructing the ancestral vertebrate brain. Develop Growth Differ 59:163174.Google Scholar
Sugahara, F., Pascual-Anaya, J., Oisi, Y., Kuraku, S., Aota, S., Adachi, N., Takagi, W., Hirai, T., Sato, H., Murakami, Y., & Kuratani, S. (2016) Evidence from cyclostomes for complex regionalization of the ancestral vertebrate brain. Nature 531: 97100.Google Scholar
Sur, M., & Rubenstein, J.L.R. (2005) Patterning and plasticity of the cerebral cortex. Sci 310:805810.Google Scholar
Suryanarayana, S.M., Pérez-Fernández, J., Robertson, B., & Grillner, S. (2020) The evolutionary origin of visual and somatosensory representation in the vertebrate pallium. Nat Ecol Evol 4:639651.Google Scholar
Suryanarayana, S.M., Robertson, B., Wallén, P., & Grillner, S. (2017) The lamprey pallium provides a blueprint of the mammalian layered cortex. Curr Biol 27:32643277.Google Scholar
Suzuki, I.K., & Hirata, T. (2012) Evolutionary conservation of neocortical neurogenetic program in the mammals and birds. Bioarchitec 2: 124129.CrossRefGoogle ScholarPubMed
Suzuki, I.K., & Hirata, T. (2013) Neocortical neurogenesis is not really “neo”: a new evolutionary model derived from a comparative study of chick pallial development. Dev Growth Differ 55:173187.Google Scholar
Suzuki, I.K., Kawasaki, T., Gojobori, T., & Hirata, T. (2012) The temporal sequence of the mammalian neocortical neurogenic program drives mediolateral pattern in the chick pallium. Dev Cell 22:863870.Google Scholar
Swanson, L.W., & Petrovich, G.D. (1998) What is the amygdala? Trends Neurosci 21:323331.Google Scholar
Szele, F.G., Chin, H.K., Rowlson, M.A., & Cepko, C.L. (2002) Sox-9 and cDachsund-2 expression in the developing chick telencephalon. Mech Dev 112:179182.Google Scholar
Tosa, Y., Hirao, A., Matsubara, I., Kawaguchi, M., Fukui, M., Kuratani, S., & Murakami, Y. (2015) Development of the thalamo-dorsal ventricular ridge tract in the Chinese soft-shelled turtle, Pelodiscus sinensis. Develop Growth Differ 57:4057.Google Scholar
Tosches, M.A., Yamawaki, T.M., Naumann, R.K., Jacobi, A.A., Tushev, G., & Laurent, G. (2018) Evolution of pallium, hippocampus, and cortical cell types revealed by single-cell transcriptomes in reptiles. Sci 360:881888.Google Scholar
Ulinski, P.S. (1983) Dorsal Ventricular Ridge: A Treatise on Forebrain Organization in Reptiles and Birds. John Wiley, New York.Google Scholar
van der Meij, J., Martinez-Gonzalez, D., Beckers, G.J.L., & Rattenborg, N.C. (2019) Intra-“cortical” activity during avian non-REM and REM sleep: variant and invariant traits between birds and mammals. Sleep J 42:113.Google ScholarPubMed
Veenman, C.L., Medina, L., & Reiner, A. (1997) Avian homologues of mammalian intralaminar, mediodorsal and midline thalamic nuclei: immunohistochemical and hodological evidence. Brain Behav Evol 49:7898.Google Scholar
Vernier, P. (2017) The brains of teleost fishes. In: Kaas, J.H., & Striedter, G. (Eds.), Evolution of Nervous Systems. Vol. 1: The Evolution of the Nervous Systems in Nonmammalian Vertebrates, 2nd ed.Elsevier, Amsterdam, pp. 5975.Google Scholar
Vila Pouca, C., Gervais, C., Reed, J., Michard, J., & Brown, C. (2019) Quantity discrimination in Port Jackson sharks incubated under elevated temperatures. Behav Ecol Sociobiol 73:93. doi:10.1007/s00265-019-2706-8.Google Scholar
von Bayern, A.M.P., Danel, S., Auersperg, A.M.I., Mioduszewska, B., & Kacelnik, A. (2018) Compound tool construction by New Caledonian crows. Sci Reports 8:15676. doi:10.1038/s41598-018-33458-z.Google Scholar
von Fersen, L., Wynne, C.D.L., & Delius, J.D. (1990) Deductive reasoning in pigeons. Naturwissench 77:548549.Google Scholar
Wagner, G.P. (1999) A research programme for testing the biological homology concept. In Bock, G.R., & Cardew, G. (Eds.), Homology (Novartis Foundation Symposium 222). John Wiley & Sons, Chichester, pp. 125140.Google Scholar
Wagner, G.P. (2014) Homology, Genes, and Evolutionary Innovation. Princeton University Press, Princeton, NJ.Google Scholar
Wagner, G.P., & Misof, B.Y. (1992) Evolutionary modification of regenerative capability in vertebrates: a comparative study on teleost pectoral fin regeneration. J Exp Zool 261:6278.Google Scholar
Watanabe, M. (1987) Synaptic organization of the nucleus dorsolateralis anterior thalami in the Japanese quail (Coturnix coturnis japonica). Brain Res 401:92112.Google Scholar
Watanabe, S., Sakamoto, J., & Wakita, M. (1995) Pigeons’ discrimination of paintings by Monet and Picasso. J Exp Anal Behav 63:165174.Google Scholar
Westhoff, G., Roth, G., & Straka, H. (2004) Topographic representation of vestibular and somatosensory signals in the anuran thalamus. Neurosci 124:669683.Google Scholar
Wild, J.M. (1987) The avian somatosensory system: connections of regions of body representation in the forebrain of the pigeon. Brain Res 412:205223.Google Scholar
Wild, J.M. (1997) The avian somatosensory system: the pathway from wing to Wulst in a passerine (Chloris chloris). Brain Res 759:122134.Google Scholar
Wild, J.M., Arends, J.J., & Zeigler, H.P. (1985) Telencephalic connections of the trigeminal system in the pigeon (Columba livia): a trigeminal sensorimotor circuit. J Comp Neurol 234:441464.Google Scholar
Wild, J.M., Karten, H.J., & Frost, B.J. (1993) Connections of the auditory forebrain in the pigeon (Columba livia). J Comp Neurol 337:3262.Google Scholar
Wild, J.M., & Williams, M.N. (2000) Rostral Wulst in passerine birds. I. Origin, course, and termination of an avian pyramidal tract. J Comp Neurol 416:429450.Google Scholar
Wyzisk, K., & Neumeyer, C. (2007) Perception of illusory surfaces and contours in goldfish. Vis Neurosci 24:291298.Google Scholar
Yamamoto, K., & Bloch, S. (2017) Overview of brain evolution: lobe-finned fish vs. ray-finned fish. In: Watanabe, S., Hofman, M.A., & Shimizu, T. (Eds.), Evolution of the Brain, Cognition, and Emotion in Vertebrates. Springer Japan KK, Tokyo, pp. 333.Google Scholar
Yamamoto, K., & Reiner, A. (2007) Is the avian dorsal ventricular ridge (DVR) homologous to the mammalian cerebral cortex or to the amygdala? Evaluating hypotheses by assessing homologous projection pathways to the telencephalon. In Watanabe, S., & Hofman, M.A. (Eds.), Integration of Comparative Neuroanatomy and Cognition. Keio University Press, Tokyo, pp. 7596.Google Scholar
Zeier, H., & Karten, H.J. (1971) The archistriatum of the pigeon: organization of afferent and efferent connections. Brain Res 31:313326.Google Scholar
Zhu, D., Lustig, K.H., Bifulco, K., & Keifer, J. (2005) Thalamocortical connections in the pond turtle Pseudemys scripta elegans. Brain Behav Evol 65:278292.Google Scholar

References

Benito-Gutiérrez, È., Stemmer, M., Rohr, S.D., Schuhmacher, L.N., Tang, J., Marconi, A., Jékely, G., and Arendt, D. (2018). Patterning of a telencephalon-like region in the adult brain of amphioxus. bioRxiv, 307629.Google Scholar
Bloch, S., Hagio, H., Thomas, M., Heuze, A., Hermel, J.M., Lasserre, E., Colin, I., Saka, K., Affaticati, P., Jenett, A., et al. (2020). Non-thalamic origin of zebrafish sensory nuclei implies convergent evolution of visual pathways in amniotes and teleosts. eLife 9, e54945.Google Scholar
Bourassa, J., and Deschênes, M. (1995). Corticothalamic projections from the primary visual cortex in rats: a single fiber study using biocytin as an anterograde tracer. Neuroscience 66, 253263.Google Scholar
Braasch, I., Gehrke, A.R., Smith, J.J., Kawasaki, K., Manousaki, T., Pasquier, J., Amores, A., Desvignes, T., Batzel, P., Catchen, J., et al. (2016). The spotted gar genome illuminates vertebrate evolution and facilitates human-teleost comparisons. Nat Genet 48, 427437.Google Scholar
Braford, M.R.J., and Northcutt, R.G. (1983). Organization of the diencephalon and pretectum of the ray-finned fishes. In Neurobiology, Vol 2: Higher Brain Areas and Functions, Davis, R.E., and Northcutt, R.G., eds. (Ann Arbor: University of Michigan Press), pp. 117164.Google Scholar
Briscoe, S.D., and Ragsdale, C.W. (2019). Evolution of the chordate telencephalon. Curr Biol 29, R647R662.Google Scholar
Butler, A.B. (1994). The evolution of the dorsal thalamus of jawed vertebrates, including mammals: cladistic analysis and a new hypothesis. Brain Res Brain Res Rev 19, 2965.Google Scholar
Butler, A.B. (1995). The dorsal thalamus of jawed vertebrates: a comparative viewpoint. Brain Behav Evol 46, 209223.Google Scholar
Butler, A.B. (2008). Evolution of brains, cognition, and consciousness. Brain Res Bull 75, 442449.Google Scholar
Butler, A.B., and Hodos, W. (2005). Comparative Vertebrate Neuroanatomy: Evolution and Adaptation, 2nd edn (New Jersey: Wiley).Google Scholar
Butler, A.B., and Saidel, W.M. (1993). Retinal projections in teleost fishes: Patterns, variations, and questions. Comp Biochem Physiol Part A 104, 431442.Google Scholar
Capantini, L., von Twickel, A., Robertson, B., and Grillner, S. (2017). The pretectal connectome in lamprey. J Comp Neurol 525, 753772.Google Scholar
Cleland, B.G., Dubin, M.W., and Levick, W.R. (1971). Simultaneous recording of input and output of lateral geniculate neurones. Nat New Biol 231, 191192.CrossRefGoogle ScholarPubMed
Cohen, D.H., Duff, T.A., and Ebbesson, S.O. (1973). Electrophysiological identification of a visual area in shark telencephalon. Science 182, 492494.Google Scholar
Comoli, E., Das Neves Favaro, P., Vautrelle, N., Leriche, M., Overton, P.G., and Redgrave, P. (2012). Segregated anatomical input to sub-regions of the rodent superior colliculus associated with approach and defense. Front Neuroanat 6, 9.Google Scholar
Crabtree, J.W. (2018). Functional diversity of thalamic reticular subnetworks. Front Syst Neurosci 12, 41.Google Scholar
Deschênes, M., Bourassa, J., and Pinault, D. (1994). Corticothalamic projections from layer V cells in rat are collaterals of long-range corticofugal axons. Brain Res 664, 215219.Google Scholar
Drager, U.C. (1975). Receptive fields of single cells and topography in mouse visual cortex. J Comp Neurol 160, 269290.Google Scholar
Dubuc, R., Bongianni, F., Ohta, Y., and Grillner, S. (1993a). Anatomical and physiological study of brainstem nuclei relaying dorsal column inputs in lampreys. J Comp Neurol 327, 260270.Google Scholar
Dubuc, R., Bongianni, F., Ohta, Y., and Grillner, S. (1993b). Dorsal root and dorsal column mediated synaptic inputs to reticulospinal neurons in lampreys: involvement of glutamatergic, glycinergic, and GABAergic transmission. J Comp Neurol 327, 251259.Google Scholar
Dugas-Ford, J., and Ragsdale, C.W. (2015). Levels of homology and the problem of neocortex. Annu Rev Neurosci 38, 351368.Google Scholar
Ebbesson, S.O., and Hodde, K.C. (1981). Ascending spinal systems in the nurse shark, Ginglymostoma cirratum. Cell Tissue Res 216, 313331.Google Scholar
Ebbesson, S.O., and Schroeder, D.M. (1971). Connections of the nurse shark’s telencephalon. Science 173, 254256.Google Scholar
Echteler, S.M., and Saidel, W.M. (1981). Forebrain connections in the goldfish support telencephalic homologies with land vertebrates. Science 212, 683685.Google Scholar
Erisir, A., Van Horn, S.C., and Sherman, S.M. (1997). Relative numbers of cortical and brainstem inputs to the lateral geniculate nucleus. Proc Natl Acad Sci USA 94, 15171520.Google Scholar
Feinberg, T.E., and Mallatt, J. (2017). Corrigendum to “The Nature of Primary Consciousness. A New Synthesis” [Conscious Cogn. 43 (2016) 113–127]. Conscious Cogn 48, 293.Google Scholar
Fournier, J., Muller, C.M., Schneider, I., and Laurent, G. (2018). Spatial information in a non-retinotopic visual cortex. Neuron 97, 164–180 e167.Google Scholar
González, A., Lopez, J.M., Morona, R., and Moreno, N. (2020). The organization of the central nervous system of amphibians. In Evolutionary Neuroscience, Kaas, J.H., ed. (Oxford: Academic Press), pp. 125157.Google Scholar
Grillner, S., and Robertson, B. (2016). The basal ganglia over 500 million years. Curr Biol 26, R1088R1100.Google Scholar
Grillner, S., Robertson, B., and Kotaleski Hellgren, J. (2020). Basal ganglia—a motion perspective. Compr Physiol 10, 12411275.Google Scholar
Guillery, R. (2017). The Brain as a Tool. A Neuroscientist’s Account (Oxford: Oxford University Press).Google Scholar
Hall, J.A., Foster, R.E., Ebner, F.F., and Hall, W.C. (1977). Visual cortex in a reptile, the turtle (Pseudemys scripta and Chrysemys picta). Brain Res 130, 197216.Google Scholar
Heap, L.A.L., Vanwalleghem, G., Thompson, A.W., Favre-Bulle, I.A., and Scott, E.K. (2018). Luminance changes drive directional startle through a thalamic pathway. Neuron 99, 293–301 e294.Google Scholar
Heier, P. (1948). Fundamental properties in the structure of the brain. A study of the brain of Petromyzon marinus. Acta Anat 8, 3213.Google Scholar
Heimberg, A.M., Cowper-Sal-lari, R., Semon, M., Donoghue, P.C., and Peterson, K.J. (2010). microRNAs reveal the interrelationships of hagfish, lampreys, and gnathostomes and the nature of the ancestral vertebrate. Proc Natl Acad Sci USA 107, 1937919383.Google Scholar
Helmbrecht, T.O., Dal Maschio, M., Donovan, J.C., Koutsouli, S., and Baier, H. (2018). Topography of a visuomotor transformation. Neuron 100, 1429–1445 e1424.Google Scholar
Herrick, C.J. (1948). The Brain of the Tiger Salamander, Ambystoma tigrinum (Chicago: University of Chicago Press).Google Scholar
Ishikawa, Y., Yamamoto, N., Yoshimoto, M., Yasuda, T., Maruyama, K., Kage, T., Takeda, H., and Ito, H. (2007). Developmental origin of diencephalic sensory relay nuclei in teleosts. Brain Behav Evol 69, 8795.Google Scholar
Ito, T., and Atoji, Y. (2016). Tectothalamic inhibitory projection neurons in the avian torus semicircularis. J Comp Neurol 524, 26042622.Google Scholar
Jones, M.R., Grillner, S., and Robertson, B. (2009). Selective projection patterns from subtypes of retinal ganglion cells to tectum and pretectum: distribution and relation to behavior. J Comp Neurol 517, 257275.Google Scholar
Kaas, J.H. (2004). Somatosensory system. In The Human Nervous System, Paxinos, G., and Mai, J.K., eds. (New York: Elsevier), pp. 10591092.Google Scholar
Kaas, J.H., Nelson, R.J., Sur, M., Lin, C.S., and Merzenich, M.M. (1979). Multiple representations of the body within the primary somatosensory cortex of primates. Science 204, 521523.Google Scholar
Karten, H.J. (2015). Vertebrate brains and evolutionary connectomics: on the origins of the mammalian “neocortex.” Philos Trans R Soc Lond B Biol Sci 370.Google Scholar
Kenigfest, N.B., and Belekhova, M.G. (2009). [Evolutionary significance of reciprocal connections in the turtle tectofugal visual system]. Zh Evol Biokhim Fiziol 45, 334342.Google Scholar
Krubitzer, L.A., and Kaas, J.H. (1992). The somatosensory thalamus of monkeys: cortical connections and a redefinition of nuclei in marmosets. J Comp Neurol 319, 123140.Google Scholar
Kumar, S., and Hedges, S.B. (1998). A molecular timescale for vertebrate evolution. Nature 392, 917920.Google Scholar
Kunst, M., Laurell, E., Mokayes, N., Kramer, A., Kubo, F., Fernandes, A.M., Forster, D., Dal Maschio, M., and Baier, H. (2019). A Cellular-resolution atlas of the larval zebrafish brain. Neuron 103, 21–38 e25.Google Scholar
Lamanna, F, Hervas-Sotomayor F, A.P. O, Jandzik D, Sobrido-Cameán D, Martik ML, Green SA, Brüning T, Mößinger K, Schmidt J, et al.: Reconstructing the ancestral vertebrate brain using a lamprey neural cell type atlas. BioRxiv 2022.Google Scholar
Luiten, P.G. (1981a). Two visual pathways to the telencephalon in the nurse shark (Ginglymostoma cirratum). I. Retinal projections. J Comp Neurol 196, 531538.Google Scholar
Luiten, P.G. (1981b). Two visual pathways to the telencephalon in the nurse shark (Ginglymostoma cirratum). II. Ascending thalamo-telencephalic connections. J Comp Neurol 196, 539548.Google Scholar
Ma, M., Kler, S., and Pan, Y.A. (2019). Structural neural connectivity analysis in zebrafish with restricted anterograde transneuronal viral labeling and quantitative brain mapping. Front Neural Circuits 13, 85.Google Scholar
Marin, O., Gonzalez, A., and Smeets, W.J. (1997a). Basal ganglia organization in amphibians: afferent connections to the striatum and the nucleus accumbens. J Comp Neurol 378, 1649.Google Scholar
Marin, O., Gonzalez, A., and Smeets, W.J. (1997b). Basal ganglia organization in amphibians: efferent connections of the striatum and the nucleus accumbens. J Comp Neurol 380, 2350.Google Scholar
Matesz, C., and Szekely, G. (1978). The motor column and sensory projections of the branchial cranial nerves in the frog. J Comp Neurol 178, 157176.Google Scholar
Medina, L., and Reiner, A. (2000). Do birds possess homologues of mammalian primary visual, somatosensory and motor cortices? Trends Neurosci 23, 112.Google Scholar
Medina, M., Reperant, J., Ward, R., Rio, J.P., and Lemire, M. (1993). The primary visual system of flatfish: an evolutionary perspective. Anat Embryol (Berl) 187, 167191.Google Scholar
Miyashita, T., Coates, M.I., Farrar, R., Larson, P., Manning, P.L., Wogelius, R.A., Edwards, N.P., Anne, J., Bergmann, U., Palmer, A.R., et al. (2019). Hagfish from the Cretaceous Tethys Sea and a reconciliation of the morphological-molecular conflict in early vertebrate phylogeny. Proc Natl Acad Sci USA 116, 21462151.Google Scholar
Morona, R., and Gonzalez, A. (2008). Calbindin-D28k and calretinin expression in the forebrain of anuran and urodele amphibians: further support for newly identified subdivisions. J Comp Neurol 511, 187220.Google Scholar
Morona, R., Lopez, J.M., and Gonzalez, A. (2011). Localization of calbindin-d28k and calretinin in the brain of Dermophis mexicanus (amphibia: gymnophiona) and its bearing on the interpretation of newly recognized neuroanatomical regions. Brain Behav Evol 77, 231269.Google Scholar
Mueller, T. (2012). What is the thalamus in zebrafish? Front Neurosci 6, 64.Google Scholar
Mundell, N.A., Beier, K.T., Pan, Y.A., Lapan, S.W., Goz Ayturk, D., Berezovskii, V.K., Wark, A.R., Drokhlyansky, E., Bielecki, J., Born, R.T., et al. (2015). Vesicular stomatitis virus enables gene transfer and transsynaptic tracing in a wide range of organisms. J Comp Neurol 523, 16391663.Google Scholar
Munoz, A., Munoz, M., Gonzalez, A., and Ten Donkelaar, H.J. (1995). Anuran dorsal column nucleus: organization, immunohistochemical characterization, and fiber connections in Rana perezi and Xenopus laevis. J Comp Neurol 363, 197220.Google Scholar
Nieuwenhuys, R., and Nicholson, C. (1998). Lampreys (Petromyzontoidea). In The Central Nervous System of Vertebrates, Nieuwenhuys, R., Donkelaar, H.J.T., and Nicholson, C., eds. (Berlin, Heidelberg: Springer), pp. 397495.Google Scholar
Nieuwenhuys, R., ten Donkelaar, H.J., and Nicholson, C. (1998). The Central Nervous System of Vertebrates (Berlin, Heidelberg: Springer).Google Scholar
Northcutt, R.G., and Butler, A.B. (1976). Retinofugal pathways in the lingnose gar Lepisosteus osseus (linnaeus). J Comp Neurol 166, 115.Google Scholar
Northcutt, R.G., and Butler, A.B. (1980). Projections of the optic tectum in the longnose gar, Lepisosteus osseus. Brain Res 190, 333346.Google Scholar
Northcutt, R.G., and Kicliter, E. (1980). Organization of the amphibian telencephalon. In Comparative Neurology of the Telencephalon, Ebbesson, S.O.E., ed. (Boston, MA: Springer).Google Scholar
Northcutt, R.G., and Wicht, H. (1997). Afferent and efferent connections of the lateral and medial pallia of the silver lamprey. Brain Behav Evol 49, 119.Google Scholar
Ocana, F.M., Suryanarayana, S.M., Saitoh, K., Kardamakis, A.A., Capantini, L., Robertson, B., and Grillner, S. (2015). The lamprey pallium provides a blueprint of the mammalian motor projections from cortex. Curr Biol 25, 413423.Google Scholar
Parichy, D.M. (2016). The gar is a fish… is a bird… is a mammal? Nat Genet 48, 344345.Google Scholar
Patel, M.B., Sons, S., Yudintsev, G., Lesicko, A.M., Yang, L., Taha, G.A., Pierce, S.M., and Llano, D.A. (2017). Anatomical characterization of subcortical descending projections to the inferior colliculus in mouse. J Comp Neurol 525, 885900.Google Scholar
Perez-Fernandez, J., Kardamakis, A.A., Suzuki, D.G., Robertson, B., and Grillner, S. (2017). Direct dopaminergic projections from the SNc modulate visuomotor transformation in the lamprey tectum. Neuron 96, 910–924 e915.Google Scholar
Perez-Fernandez, J., Stephenson-Jones, M., Suryanarayana, S.M., Robertson, B., and Grillner, S. (2014). Evolutionarily conserved organization of the dopaminergic system in lamprey: SNc/VTA afferent and efferent connectivity and D2 receptor expression. J Comp Neurol 522, 37753794.Google Scholar
Pombal, M.A., and Puelles, L. (1999). Prosomeric map of the lamprey forebrain based on calretinin immunocytochemistry, Nissl stain, and ancillary markers. J Comp Neurol 414, 391422.Google Scholar
Puelles, L. (2017). Comments on the updated tetrapartite pallium model in the mouse and chick, featuring a homologous claustro-insular complex. Brain Behav Evol 90, 171189.Google Scholar
Ravi, V., and Venkatesh, B. (2018). The divergent genomes of teleosts. Annu Rev Anim Biosci 6, 4768.Google Scholar
Reiner, A. (1993). Neurotransmitter organization and connections of turtle cortex: implications for the evolution of mammalian isocortex. Comp Biochem Physiol Comp Physiol 104, 735748.Google Scholar
Reiner, A., Yamamoto, K., and Karten, H.J. (2005). Organization and evolution of the avian forebrain. Anat Rec A Discov Mol Cell Evol Biol 287, 10801102.Google Scholar
Roth, G., Blanke, J., and Ohle, M. (1995). Brain size and morphology in miniaturized plethodontid salamanders. Brain Behav Evol 45, 8495.Google Scholar
Roth, G., Nishikawa, K.C., Naujoks-Manteuffel, C., Schmidt, A., and Wake, D.B. (1993). Paedomorphosis and simplification in the nervous system of salamanders. Brain Behav Evol 42, 137170.Google Scholar
Schneider, G.E. (1969). Two visual systems. Science 163, 895902.Google Scholar
Sincich, L.C., Zhang, Y., Tiruveedhula, P., Horton, J.C., and Roorda, A. (2009). Resolving single cone inputs to visual receptive fields. Nat Neurosci 12, 967969.Google Scholar
Smeets, W.J. (1982). The afferent connections of the tectum mesencephali in two chondrichthyans, the shark Scyliorhinus canicula and the ray Raja clavata. J Comp Neurol 205, 139152.Google Scholar
Smeets, W.J. (1992). Comparative aspects of basal forebrain organization in vertebrates. Eur J Morphol 30, 2336.Google Scholar
Stephenson-Jones, M., Ericsson, J., Robertson, B., and Grillner, S. (2012). Evolution of the basal ganglia: dual-output pathways conserved throughout vertebrate phylogeny. J Comp Neurol 520, 29572973.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
Striedter, G.F. (1990a). The diencephalon of the channel catfish, Ictalurus punctatus. I. Nuclear organization. Brain Behav Evol 36, 329354.Google Scholar
Striedter, G.F. (1990b). The diencephalon of the channel catfish, Ictalurus punctatus. II. Retinal, tectal, cerebellar and telencephalic connections. Brain Behav Evol 36, 355377.Google Scholar
Sugahara, F., Murakami, Y., Pascual-Anaya, J., and Kuratani, S. (2017). Reconstructing the ancestral vertebrate brain. Dev Growth Differ 59, 163174.Google Scholar
Sugahara, F., Pascual-Anaya, J., Oisi, Y., Kuraku, S., Aota, S., Adachi, N., Takagi, W., Hirai, T., Sato, N., Murakami, Y., et al. (2016). Evidence from cyclostomes for complex regionalization of the ancestral vertebrate brain. Nature 531, 97100.Google Scholar
Suryanarayana, S.M., Perez-Fernandez, J., Robertson, B., and Grillner, S. (2020). The evolutionary origin of visual and somatosensory representation in the vertebrate pallium. Nat Ecol Evol 4, 639651.Google Scholar
Suryanarayana, S.M., Pérez-Fernández, J., Robertson, B., and Grillner, S. (2021). Olfaction in lamprey pallium revisited—dual projections of mitral and tufted cells. Cell Reports 34.Google Scholar
Suryanarayana, S.M., Robertson, B., Wallen, P., and Grillner, S. (2017). The lamprey pallium provides a blueprint of the mammalian layered cortex. Curr Biol 27, 3264–3277 e3265.Google Scholar
ten Donkelaar, H.J. (1998). Urodeles. In The Central Nervous System of Vertebrates, Nieuwenhuys, R., ten Donkelaar, H.J., and Nicolson, C., eds. (Berlin, Heidelberg: Springer), pp. 10451150.Google Scholar
Ulinski, P.S. (1986). Organization of corticogeniculate projections in the turtle, Pseudemys scripta. J Comp Neurol 254, 529542.Google Scholar
Villar-Cervino, V., Barreiro-Iglesias, A., Mazan, S., Rodicio, M.C., and Anadon, R. (2011). Glutamatergic neuronal populations in the forebrain of the sea lamprey, Petromyzon marinus: an in situ hybridization and immunocytochemical study. J Comp Neurol 519, 17121735.Google Scholar
Westhoff, G., and Roth, G. (2002). Morphology and projection pattern of medial and dorsal pallial neurons in the frog Discoglossus pictus and the salamander Plethodon jordani. J Comp Neurol 445, 97121.Google Scholar
Westhoff, G., Roth, G., and Straka, H. (2004). Topographic representation of vestibular and somatosensory signals in the anuran thalamus. Neuroscience 124, 669683.Google Scholar
Wicht, H., and Himstedt, W. (1988). Topologic and connectional analysis of the dorsal thalamus of Triturus alpestris (amphibia, urodela, salamandridae). J Comp Neurol 267, 545561.Google Scholar
Wild, J.M. (1997). The avian somatosensory system: the pathway from wing to Wulst in a passerine (Chloris chloris). Brain Res 759, 122134.Google Scholar
Wullimann, M.F., and Mueller, T. (2004). Teleostean and mammalian forebrains contrasted: Evidence from genes to behavior. J Comp Neurol 475, 143162.Google Scholar
Wullimann, M.F., and Northcutt, R.G. (1990). Visual and electrosensory circuits of the diencephalon in mormyrids: an evolutionary perspective. J Comp Neurol 297, 537552.Google Scholar
Yamamoto, N., and Ito, H. (2008). Visual, lateral line, and auditory ascending pathways to the dorsal telencephalic area through the rostrolateral region of the lateral preglomerular nucleus in cyprinids. J Comp Neurol 508, 615647.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.

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.

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.

Available formats
×