Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-x24gv Total loading time: 0 Render date: 2024-05-07T03:13:29.636Z Has data issue: false hasContentIssue false

1 - Embodied Brain Model for Understanding Functional Neural Development of Fetuses and Infants

from Part I - Foundations

Published online by Cambridge University Press:  26 September 2020

Jeffrey J. Lockman
Affiliation:
Tulane University, Louisiana
Catherine S. Tamis-LeMonda
Affiliation:
New York University
Get access

Summary

Early functional neural development is increasingly recognized as important for revealing the developmental origins of human cognitive-motor function and related disorders. Previous studies focusing on fetuses and neonates have revealed sophisticated behaviors and cognitive repertoires, indicating that fetuses begin learning through sensorimotor experience even inside the uterus. Despite accumulating evidence supporting the importance of sensorimotor experience in neural development as early as the fetal period, the developmental mechanisms by which intrauterine sensorimotor experience guides cortical learning, including factors in prenatal experience that are needed for normal development, remain unclear. However, investigating causal links between sensorimotor experience and cortical learning is particularly challenging in human fetuses owing to technical and ethical difficulties. Therefore, computational approaches based on comprehensive biological data about nervous system, body, and environment have been developed to probe mechanisms underlying early functional brain development. In this chapter, we show how an embodied approach focusing on interactions among brain, body, and environment offers opportunities to explore relations between functional neural development and sensorimotor experience.

Type
Chapter
Information
The Cambridge Handbook of Infant Development
Brain, Behavior, and Cultural Context
, pp. 3 - 39
Publisher: Cambridge University Press
Print publication year: 2020

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

Adolph, K. E., & Berger, S. E. (2006). Motor development. Handbook of Child Psychology, 2, 161213.Google Scholar
Akerman, C. J., Smyth, D., & Thompson, I. D. (2002). Visual experience before eye-opening and the development of the retinogeniculate pathway. Neuron, 36(5), 869879.Google Scholar
Allendoerfer, K. L., & Shatz, C. J. (1994). The subplate, a transient neocortical structure: Its role in the development of connections between thalamus and cortex. Annual Review of Neuroscience, 17(1), 185218.Google Scholar
Als, H., Lawhon, G., Brown, E., Gibes, R., Duffy, F. H., McAnulty, G., & Blickman, J. G. (1986). Individualized behavioral and environmental care for the very low birth weight preterm infant at high risk for bronchopulmonary dysplasia: Neonatal intensive care unit and developmental outcome. Pediatrics, 78(6), 11231132.CrossRefGoogle ScholarPubMed
Asada, M., MacDorman, K. F., Ishiguro, H., & Kuniyoshi, Y. (2001). Cognitive developmental robotics as a new paradigm for the design of humanoid robots. Robotics and Autonomous Systems, 37(2–3), 185193.Google Scholar
Ball, G., Srinivasan, L., Aljabar, P., Counsell, S. J., Durighel, G., Hajnal, J. V., … Edwards, A. D. (2013). Development of cortical microstructure in the preterm human brain. Proceedings of the National Academy of Sciences, 110(23), 95419546.Google Scholar
Banerjee, A., Meredith, R. M., Rodríguez-Moreno, A., Mierau, S. B., Auberson, Y. P., & Paulsen, O. (2009). Double dissociation of spike timing-dependent potentiation and depression by subunit-preferring NMDA receptor antagonists in mouse barrel cortex. Cerebral Cortex, 19(12), 29592969.Google Scholar
Barbu-Roth, M., Anderson, D. I., Desprès, A., Streeter, R. J., Cabrol, D., Trujillo, M., … Provasi, J. (2014). Air stepping in response to optic flows that move toward and away from the neonate. Developmental Psychobiology, 56(5), 11421149.Google Scholar
Bayatti, N., Moss, J. A., Sun, L., Ambrose, P., Ward, J. F., Lindsay, S., & Clowry, G. J. (2007). A molecular neuroanatomical study of the developing human neocortex from 8 to 17 postconceptional weeks revealing the early differentiation of the subplate and subventricular zone. Cerebral Cortex, 18(7), 15361548.Google Scholar
Beccaria, E., Martino, M., Briatore, E., Podestà, B., Pomero, G., Micciolo, R., … Calzolari, S. (2012). Poor repertoire general movements predict some aspects of development outcome at 2 years in very preterm infants. Early Human Development, 88(6), 393396.Google Scholar
Ben-Ari, Y., Gaiarsa, J. -L., Tyzio, R., & Khazipov, R. (2007). GABA: A pioneer transmitter that excites immature neurons and generates primitive oscillations. Physiological Reviews, 87(4), 12151284.Google Scholar
Blaesse, P., Airaksinen, M. S., Rivera, C., & Kaila, K. (2009). Cation-chloride cotransporters and neuronal function. Neuron, 61(6), 820838.CrossRefGoogle ScholarPubMed
Blankenship, A. G., & Feller, M. B. (2010). Mechanisms underlying spontaneous patterned activity in developing neural circuits. Nature Reviews Neuroscience, 11(1), 18.CrossRefGoogle ScholarPubMed
Bobet, J., & Stein, R. B. (1998). A simple model of force generation by skeletal muscle during dynamic isometric contractions. IEEE Transactions on Biomedical Engineering, 45(8), 10101016.Google Scholar
Boivin, M. J., Kakooza, A. M., Warf, B. C., Davidson, L. L., & Grigorenko, E. L. (2015). Reducing neurodevelopmental disorders and disability through research and interventions. Nature, 527(7578), S155.Google Scholar
Bosco, G., & Poppele, R. (2001). Proprioception from a spinocerebellar perspective. Physiological Reviews, 81(2), 539568.CrossRefGoogle ScholarPubMed
Bradley, R. M., & Mistretta, C. M. (1975). Fetal sensory receptors. Physiological Reviews, 55(3), 352382.CrossRefGoogle ScholarPubMed
Bremner, A. J., Lewkowicz, D. J., & Spence, C. (2012). Multisensory development. Oxford: Oxford University Press.Google Scholar
Brooks, R. A. (1991). Intelligence without representation. Artificial intelligence, 47(1–3), 139159.Google Scholar
Brumley, M. R., & Robinson, S. R. (2013). Sensory feedback alters spontaneous limb movements in newborn rats: Effects of unilateral forelimb weighting. Developmental Psychobiology, 55(4), 323333.Google Scholar
Butterworth, G., & Hopkins, B. (1988). Hand–mouth coordination in the new-born baby. British Journal of Developmental Psychology, 6(4), 303314.Google Scholar
Byrge, L., Sporns, O., & Smith, L. B. (2014). Developmental process emerges from extended brain–body–behavior networks. Trends in Cognitive Sciences, 18(8), 395403.CrossRefGoogle ScholarPubMed
Caligiore, D., Parisi, D., & Baldassarre, G. (2014). Integrating reinforcement learning, equilibrium points, and minimum variance to understand the development of reaching: A computational model. Psychological Review, 121(3), 389.Google Scholar
Cascio, C. J. (2010). Somatosensory processing in neurodevelopmental disorders. Journal of Neurodevelopmental Disorders, 2(2), 62.Google Scholar
Cheng-Yu, T. L., Poo, M. -M., & Dan, Y. (2009). Burst spiking of a single cortical neuron modifies global brain state. Science, 324(5927), 643646.Google Scholar
Clancy, B., Darlington, R., & Finlay, B. (2001). Translating developmental time across mammalian species. Neuroscience, 105(1), 717.Google Scholar
Clifton, R. K., Morrongiello, B. A., Kulig, J. W., & Dowd, J. M. (1981). Newborns’ orientation toward sound: Possible implications for cortical development. Child Development, 52(3), 833838.Google Scholar
Crisp, S. J., Evers, J. F., & Bate, M. (2011). Endogenous patterns of activity are required for the maturation of a motor network. Journal of Neuroscience, 31(29), 1044510450.Google Scholar
Cuajunco, F. (1940). Development of the neuromuscular spindle in human fetuses. Contributions to Embryology, 28, 97128.Google Scholar
de Vries, J. I., Visser, G. H., & Prechtl, H. F. (1982). The emergence of fetal behaviour. I: Qualitative aspects. Early Human Development, 7(4), 301322.Google Scholar
DeCasper, A. J., Lecanuet, J. -P., Busnel, M. -C., Granier-Deferre, C., & Maugeais, R. (1994). Fetal reactions to recurrent maternal speech. Infant Behavior and Development, 17(2), 159164.Google Scholar
Demiris, J., Rougeaux, S., Hayes, G., Berthouze, L., & Kuniyoshi, Y. (1997). Deferred imitation of human head movements by an active stereo vision head. Paper presented at 6th IEEE International Workshop on Robot and Human Communication, RO-MAN’97 SENDAI, Sendai, Japan.Google Scholar
The Developing Human Connectome Project. Retrieved from www.developingconnectome.org.Google Scholar
Dubois, J., Hertz-Pannier, L., Dehaene-Lambertz, G., Cointepas, Y., & Le Bihan, D. (2006). Assessment of the early organization and maturation of infants’ cerebral white matter fiber bundles: A feasibility study using quantitative diffusion tensor imaging and tractography. Neuroimage, 30(4), 11211132.Google Scholar
Eyre, J., Miller, S., Clowry, G., Conway, E., & Watts, C. (2000). Functional corticospinal projections are established prenatally in the human foetus permitting involvement in the development of spinal motor centres. Brain, 123(1), 5164.Google Scholar
Fombonne, E. (2009). Epidemiology of pervasive developmental disorders. Pediatric Research, 65(6), 591.Google Scholar
Fuchino, Y., Naoi, N., Shibata, M., Niwa, F., Kawai, M., Konishi, Y., … Myowa-Yamakoshi, M. (2013). Effects of preterm birth on intrinsic fluctuations in neonatal cerebral activity examined using optical imaging. PloS one, 8(6), e67432.Google Scholar
Gallagher, S., & Zahavi, D. (2007). The phenomenological mind: An introduction to philosophy of mind and cognitive science. London: Routledge.Google Scholar
Gaugler, T., Klei, L., Sanders, S. J., Bodea, C. A., Goldberg, A. P., Lee, A. B., … Reichert, J. (2014). Most genetic risk for autism resides with common variation. Nature Genetics, 46(8), 881.Google Scholar
Gerhard, D. (2013). Neuroscience. 5th edition. Yale Journal of Biology and Medicine, 86(1), 113114.Google Scholar
Ghosh, A., Antonini, A., McConnell, S. K., & Shatz, C. J. (1990). Requirement for subplate neurons in the formation of thalamocortical connections. Nature, 347(6289), 179.Google Scholar
Gibson, J. (1979). The theory of affordances. In Gibson, J. (Ed.), The ecological approach to visual perception (pp. 127143). Boston, MA: Houghton Mifflin.Google Scholar
Gima, H., Ohgi, S., Morita, S., Karasuno, H., Fujiwara, T., & Abe, K. (2011). A dynamical system analysis of the development of spontaneous lower extremity movements in newborn and young infants. Journal of Physiological Anthropology, 30(5), 179186.Google Scholar
Girvan, M., & Newman, M. E. (2002). Community structure in social and biological networks. Proceedings of the National Academy of Sciences, 99(12), 78217826.Google Scholar
Gogtay, N., Giedd, J. N., Lusk, L., Hayashi, K. M., Greenstein, D., Vaituzis, A. C., … Toga, A. W. (2004). Dynamic mapping of human cortical development during childhood through early adulthood. Proceedings of the National Academy of Sciences, 101(21), 81748179.Google Scholar
Goldman, A., & de Vignemont, F. (2009). Is social cognition embodied? Trends in Cognitive Sciences, 13(4), 154159.Google Scholar
Gonzalez-Islas, C., & Wenner, P. (2006). Spontaneous network activity in the embryonic spinal cord regulates AMPAergic and GABAergic synaptic strength. Neuron, 49(4), 563575.Google Scholar
Granmo, M., Petersson, P., & Schouenborg, J. (2008). Action-based body maps in the spinal cord emerge from a transitory floating organization. Journal of Neuroscience, 28(21), 54945503.Google Scholar
Grillner, S. (2006). Biological pattern generation: The cellular and computational logic of networks in motion. Neuron, 52(5), 751766.CrossRefGoogle ScholarPubMed
Grillner, S., & Jessell, T. M. (2009). Measured motion: Searching for simplicity in spinal locomotor networks. Current Opinion in Neurobiology, 19(6), 572586.CrossRefGoogle ScholarPubMed
Groen, S. E., de Blécourt, A. C., Postema, K., & Hadders-Algra, M. (2005). General movements in early infancy predict neuromotor development at 9 to 12 years of age. Developmental Medicine and Child Neurology, 47(11), 731738.Google Scholar
Hadders-Algra, M. (2004). General movements: A window for early identification of children at high risk for developmental disorders. Journal of Pediatrics, 145(2), S12S18.Google Scholar
Hadders-Algra, M. (2007). Putative neural substrate of normal and abnormal general movements. Neuroscience & Biobehavioral Reviews, 31(8), 11811190.CrossRefGoogle ScholarPubMed
Hadders-Algra, M., Mavinkurve-Groothuis, A. M., Groen, S. E., Stremmelaar, E. F., Martijn, A., & Butcher, P. R. (2004). Quality of general movements and the development of minor neurological dysfunction at toddler and school age. Clinical Rehabilitation, 18(3), 287299.Google Scholar
Haider, B., Duque, A., Hasenstaub, A. R., & McCormick, D. A. (2006). Neocortical network activity in vivo is generated through a dynamic balance of excitation and inhibition. Journal of Neuroscience, 26(17), 45354545.Google Scholar
Hakamada, S., Hayakawa, F., Kuno, K., & Tanaka, R. (1988). Development of the monosynaptic reflex pathway in the human spinal cord. Developmental Brain Research, 42(2), 239246.Google Scholar
Hamburger, V., Wenger, E., & Oppenheim, R. (1966). Motility in the chick embryo in the absence of sensory input. Journal of Experimental Zoology, 162(2), 133159.Google Scholar
Hanson, M. G., Milner, L. D., & Landmesser, L. T. (2008). Spontaneous rhythmic activity in early chick spinal cord influences distinct motor axon pathfinding decisions. Brain Research Reviews, 57(1), 7785.CrossRefGoogle ScholarPubMed
Hashimoto, T., Bazmi, H. H., Mirnics, K., Wu, Q., Sampson, A. R., & Lewis, D. A. (2008). Conserved regional patterns of GABA-related transcript expression in the neocortex of subjects with schizophrenia. American Journal of Psychiatry, 165(4), 479489.Google Scholar
He, J., Maltenfort, M. G., Wang, Q., & Hamm, T. M. (2001). Learning from biological systems: Modeling neural control. IEEE Control Systems, 21(4), 5569.Google Scholar
Hepper, P. G. (1996). Fetal memory: Does it exist? What does it do? Acta Paediatrica, 85, 1620.Google Scholar
Hoffmann, M., Marques, H., Arieta, A., Sumioka, H., Lungarella, M., & Pfeifer, R. (2010). Body schema in robotics: A review. IEEE Transactions on Autonomous Mental Development, 2(4), 304324.CrossRefGoogle Scholar
Honey, C. J., Kötter, R., Breakspear, M., & Sporns, O. (2007). Network structure of cerebral cortex shapes functional connectivity on multiple time scales. Proceedings of the National Academy of Sciences, 104(24), 1024010245.Google Scholar
Hooker, D. (1952). The prenatal origin of behavior. Lawrence: University of Kansas Press.Google Scholar
Hromádka, T., DeWeese, M. R., & Zador, A. M. (2008). Sparse representation of sounds in the unanesthetized auditory cortex. PLoS biology, 6(1), e16.Google Scholar
Huffman, K. J., & Krubitzer, L. (2001). Area 3a: Topographic organization and cortical connections in marmoset monkeys. Cerebral Cortex, 11(9), 849867.Google Scholar
James, D. K. (2010). Fetal learning: A critical review. Infant and Child Development: An International Journal of Research and Practice, 19(1), 4554.Google Scholar
James, D. K., Spencer, C., & Stepsis, B. (2002). Fetal learning: A prospective randomized controlled study. Ultrasound in Obstetrics & Gynecology, 20(5), 431438.Google Scholar
Kaas, J. H. (1983). What, if anything, is SI? Organization of first somatosensory area of cortex. Physiological Reviews, 63(1), 206231.Google Scholar
Kalaska, J., Cohen, D., Prud’Homme, M., & Hyde, M. (1990). Parietal area 5 neuronal activity encodes movement kinematics, not movement dynamics. Experimental Brain Research, 80(2), 351364.CrossRefGoogle Scholar
Kanazawa, H., Kawai, M., Kinai, T., Iwanaga, K., Mima, T., & Heike, T. (2014). Cortical muscle control of spontaneous movements in human neonates. European Journal of Neuroscience, 40(3), 25482553.Google Scholar
Kanold, P. O., & Luhmann, H. J. (2010). The subplate and early cortical circuits. Annual Review of Neuroscience, 33, 2348.Google Scholar
Kanold, P. O., & Shatz, C. J. (2006). Subplate neurons regulate maturation of cortical inhibition and outcome of ocular dominance plasticity. Neuron, 51(5), 627638.Google Scholar
Khazipov, R., Sirota, A., Leinekugel, X., Holmes, G. L., Ben-Ari, Y., & Buzsáki, G. (2004). Early motor activity drives spindle bursts in the developing somatosensory cortex. Nature, 432(7018), 758.Google Scholar
Kinney, H. C., Brody, B. A., Kloman, A. S., & Gilles, F. H. (1988). Sequence of central nervous system myelination in human infancy: II. Patterns of myelination in autopsied infants. Journal of Neuropathology & Experimental Neurology, 47(3), 217234.Google Scholar
Kisilevsky, B. S., Hains, S. M., Lee, K., Xie, X., Huang, H., Ye, H. H., … Wang, Z. (2003). Effects of experience on fetal voice recognition. Psychological Science, 14(3), 220224.Google Scholar
Kostović, I., & Judaš, M. (2010). The development of the subplate and thalamocortical connections in the human foetal brain. Acta Paediatrica, 99(8), 11191127.Google Scholar
Koyanagi, T., Horimoto, N., Maeda, H., Kukita, J., Minami, T., Ueda, K., & Nakano, H. (1993). Abnormal behavioral patterns in the human fetus at term: Correlation with lesion sites in the central nervous system after birth. Journal of child neurology, 8(1), 1926.Google Scholar
Kuniyoshi, Y. (1994). The science of imitation-towards physically and socially grounded intelligence. Paper presented at the Special Issue TR-94001, Real World Computing Project Joint Symposium, Tsukuba-shi, Ibaraki-ken.Google Scholar
Kuniyoshi, Y., & Berthouze, L. (1998). Neural learning of embodied interaction dynamics. Neural Networks, 11(7–8), 12591276.Google Scholar
Kuniyoshi, Y., Cheng, G., & Nagakubo, A. (2003). Etl-humanoid: A research vehicle for open-ended action imitation. Robotics Research, 6, 6782.Google Scholar
Kuniyoshi, Y., Yorozu, Y., Inaba, M., & Inoue, H. (2003). From visuo-motor self-learning to early imitation: A neural architecture for humanoid learning. Paper presented at the 2003 IEEE International Conference on Robotics and Automation (Cat. No. 03CH37422), Taipei, Japan.CrossRefGoogle Scholar
Kurjak, A., Azumendi, G., Veček, N., Kupešic, S., Solak, M., Varga, D., & Chervenak, F. (2003). Fetal hand movements and facial expression in normal pregnancy studied by four-dimensional sonography. Journal of Perinatal Medicine, 31(6), 496508.Google Scholar
Larroque, B., Ancel, P. -Y., Marret, S., Marchand, L., André, M., Arnaud, C., … Thiriez, G. (2008). Neurodevelopmental disabilities and special care of 5-year-old children born before 33 weeks of gestation (the EPIPAGE study): A longitudinal cohort study. Lancet, 371(9615), 813820.Google Scholar
Larsen, R. S., Rao, D., Manis, P. B., & Philpot, B. D. (2010). STDP in the developing sensory neocortex. Frontiers in Synaptic Neuroscience, 2, 9.Google ScholarPubMed
Lawn, J. E., Mwansa-Kambafwile, J., Horta, B. L., Barros, F. C., & Cousens, S. (2010). “Kangaroo mother care” to prevent neonatal deaths due to preterm birth complications. International Journal of Epidemiology, 39(suppl. 1), i144i154.Google Scholar
Lüchinger, A. B., Hadders-Algra, M., van Kan, C. M., & de Vries, J. I. (2008). Fetal onset of general movements. Pediatric Research, 63(2), 191.Google Scholar
Ludington-Hoe, S. M. (2013). Kangaroo care as a neonatal therapy. Newborn and Infant Nursing Reviews, 13(2), 7375.Google Scholar
Lungarella, M., Metta, G., Pfeifer, R., & Sandini, G. (2003). Developmental robotics: A survey. Connection Science, 15(4), 151190.Google Scholar
McConnell, S. K., Ghosh, A., & Shatz, C. J. (1989). Subplate neurons pioneer the first axon pathway from the cerebral cortex. Science, 245(4921), 978982.Google Scholar
McQuillen, P. S., Sheldon, R. A., Shatz, C. J., & Ferriero, D. M. (2003). Selective vulnerability of subplate neurons after early neonatal hypoxia-ischemia. Journal of Neuroscience, 23(8), 33083315.Google Scholar
Meliza, C. D., & Dan, Y. (2006). Receptive-field modification in rat visual cortex induced by paired visual stimulation and single-cell spiking. Neuron, 49(2), 183189.Google Scholar
Meltzoff, A. N., & Moore, M. K. (1977). Imitation of facial and manual gestures by human neonates. Science, 198(4312), 7578.Google Scholar
Merleau-Ponty, M., & Smith, C. (1996). Phenomenology of perception. Delhi: Motilal Banarsidass Publishers.Google Scholar
Milh, M., Kaminska, A., Huon, C., Lapillonne, A., Ben-Ari, Y., & Khazipov, R. (2006). Rapid cortical oscillations and early motor activity in premature human neonate. Cerebral Cortex, 17(7), 15821594.Google Scholar
Miller, J. A., Ding, S. -L., Sunkin, S. M., Smith, K. A., Ng, L., Szafer, A., … Aiona, K. (2014). Transcriptional landscape of the prenatal human brain. Nature, 508(7495), 199.CrossRefGoogle ScholarPubMed
Mullen, K. M., Vohr, B. R., Katz, K. H., Schneider, K. C., Lacadie, C., Hampson, M., … Ment, L. R. (2011). Preterm birth results in alterations in neural connectivity at age 16 years. Neuroimage, 54(4), 25632570.Google Scholar
Myowa-Yamakoshi, M., & Takeshita, H. (2006). Do human fetuses anticipate self-oriented actions? A study by four-dimensional (4D) ultrasonography. Infancy, 10(3), 289301.Google Scholar
Nagai, Y., Hosoda, K., Morita, A., & Asada, M. (2003). A constructive model for the development of joint attention. Connection Science, 15(4), 211229.Google Scholar
Narayanan, D. Z., & Ghazanfar, A. A. (2014). Developmental neuroscience: How twitches make sense. Current Biology, 24(19), R971R972.Google Scholar
Nebel, M. B., Joel, S. E., Muschelli, J., Barber, A. D., Caffo, B. S., Pekar, J. J., & Mostofsky, S. H. (2014). Disruption of functional organization within the primary motor cortex in children with autism. Human Brain Mapping, 35(2), 567580.Google Scholar
Newman, M. E. (2004). Fast algorithm for detecting community structure in networks. Physical review E, 69(6), 066133.Google Scholar
Ohgi, S., Morita, S., Loo, K. K., & Mizuike, C. (2007). A dynamical systems analysis of spontaneous movements in newborn infants. Journal of Motor Behavior, 39(3), 203214.Google Scholar
Ohlsson, A., & Jacobs, S. E. (2013). NIDCAP: A systematic review and meta-analyses of randomized controlled trials. Pediatrics, 131(3), e881e893.Google Scholar
Okado, N. (1984). Ontogeny of the central nervous system: Neurogenesis, fibre connection, synaptogenesis and myelination in the spinal cord. In Prechtl, H. F. R. (Ed.), Continuity of neural functions from prenatal to postnatal life (pp. 3145). London: Spastics International Medical Publications.Google Scholar
Partridge, E. A., Davey, M. G., Hornick, M. A., McGovern, P. E., Mejaddam, A. Y., Vrecenak, J. D., … Weiland, T. R. (2017). An extra-uterine system to physiologically support the extreme premature lamb. Nature Communications, 8, 15112.Google Scholar
Penfield, W., & Boldrey, E. (1937). Somatic motor and sensory representation in the cerebral cortex of man as studied by electrical stimulation. Brain, 60(4), 389443.Google Scholar
Petersson, P., Granmo, M., & Schouenborg, J. (2004). Properties of an adult spinal sensorimotor circuit shaped through early postnatal experience. Journal of Neurophysiology, 92, 280288.Google Scholar
Petersson, P., Waldenström, A., Fåhraeus, C., & Schouenborg, J. (2003). Spontaneous muscle twitches during sleep guide spinal self-organization. Nature, 424(6944), 72.Google Scholar
Pfeifer, R., & Scheier, C. (2001). Understanding intelligence. Cambridge, MA: MIT Press.Google Scholar
Piaget, J. (1952). The origins of intelligence in children. Madison, CT: International Universities Press.Google Scholar
Pitcher, J. B., Schneider, L. A., Burns, N. R., Drysdale, J. L., Higgins, R. D., Ridding, M. C., … Robinson, J. S. (2012). Reduced corticomotor excitability and motor skills development in children born preterm. Journal of Physiology, 590(22), 58275844.Google Scholar
Pitti, A., Mori, H., Yamada, Y., & Kuniyoshi, Y. (2010). A model of spatial development from parieto-hippocampal learning of body-place associations. Paper presented at the 10th International Conference on Epigenetic Robotics, Sweden.Google Scholar
Prechtl, H. F. R. (1984). Continuity and change in early neural development. In Prechtl, H. F. R. (Ed.), Continuity of neural functions from prenatal to postnatal life (pp. 115). London: Spastics International Medical Publications.Google Scholar
Prechtl, H. F. R. (1990). Qualitative changes of spontaneous movements in fetus and preterm infant are a marker of neurological dysfunction. Early Human Development, 23(3), 151158.Google Scholar
Prechtl, H. F. R. (2001). General movement assessment as a method of developmental neurology: New paradigms and their consequences. Developmental Medicine & Child Neurology, 43(12), 836842.Google Scholar
Purves, D. (2012). Neuroscience: Oxford: Oxford University Press.Google Scholar
Rausell, E., Bickford, L., Manger, P. R., Woods, T. M., & Jones, E. G. (1998). Extensive divergence and convergence in the thalamocortical projection to monkey somatosensory cortex. Journal of Neuroscience, 18(11), 42164232.Google Scholar
Reid, V. M., Dunn, K., Young, R. J., Amu, J., Donovan, T., & Reissland, N. (2017). The human fetus preferentially engages with face-like visual stimuli. Current Biology, 27(12), 18251828. e1823.Google Scholar
Reissland, N., Francis, B., Aydin, E., Mason, J., & Schaal, B. (2014). The development of anticipation in the fetus: A longitudinal account of human fetal mouth movements in reaction to and anticipation of touch. Developmental Psychobiology, 56(5), 955963.Google Scholar
Robinson, S. R., & Kleven, G. A. (2005). Learning to move before birth. In Hopkins, B. & Johnson, S. (Eds.), Prenatal development of postnatal functions (Advances in Infancy Research series) (Vol. 2, pp. 131175). Westport, CT: Praeger.Google Scholar
Robinson, S. R., Kleven, G. A., & Brumley, M. R. (2008). Prenatal development of interlimb motor learning in the rat fetus. Infancy, 13(3), 204228.Google Scholar
Rochat, P. (2009). The infant’s world. Cambridge, MA: Harvard University Press.Google Scholar
Rochat, P. (2011). The self as phenotype. Consciousness and Cognition, 20(1), 109119.Google Scholar
Rochat, P., & Hespos, S. J. (1997). Differential rooting response by neonates: Evidence for an early sense of self. Infant and Child Development, 6(3–4), 105112.Google Scholar
Rubenstein, J. L. (2010). Three hypotheses for developmental defects that may underlie some forms of autism spectrum disorder. Current Opinion in Neurology, 23(2), 118123.Google Scholar
Rubenstein, J. L. & Merzenich, M. M. (2003). Model of autism: Increased ratio of excitation/inhibition in key neural systems. Genes, Brain and Behavior, 2(5), 255267.Google Scholar
Sarnat, H. B. (2003). Functions of the corticospinal and corticobulbar tracts in the human newborn. Journal of Pediatric Neurology, 1(1), 38.Google Scholar
Sasaki, R., Yamada, Y., Tsukahara, Y., & Kuniyoshi, Y. (2013). Tactile stimuli from amniotic fluid guides the development of somatosensory cortex with hierarchical structure using human fetus simulation. Paper presented at the 2013 IEEE Third Joint International Conference on Development and Learning and Epigenetic Robotics (ICDL), Osaka, Japan.Google Scholar
Schaal, B., Marlier, L., & Soussignan, R. (1998). Olfactory function in the human fetus: Evidence from selective neonatal responsiveness to the odor of amniotic fluid. Behavioral Neuroscience, 112(6), 1438.Google Scholar
Shirado, H., Konyo, M., & Maeno, T. (2007). Modeling of tactile texture recognition mechanism. Nihon Kikai Gakkai Ronbunshu, C Hen/Transactions of the Japan Society of Mechanical Engineers, Part C, 73(9), 25142522.Google Scholar
Sizun, J., & Westrup, B. (2004). Early developmental care for preterm neonates: A call for more research. Archives of Disease in Childhood: Fetal and Neonatal Edition, 89(5), F384F388.Google Scholar
Smyser, C. D., Inder, T. E., Shimony, J. S., Hill, J. E., Degnan, A. J., Snyder, A. Z., & Neil, J. J. (2010). Longitudinal analysis of neural network development in preterm infants. Cerebral Cortex, 20(12), 28522862.Google Scholar
Softky, W. R., & Koch, C. (1993). The highly irregular firing of cortical cells is inconsistent with temporal integration of random EPSPs. Journal of Neuroscience, 13(1), 334350.Google Scholar
Spittle, A., Orton, J., Doyle, L. W., & Boyd, R. (2007). Early developmental intervention programs post hospital discharge to prevent motor and cognitive impairments in preterm infants. Cochrane Database of Systematic Reviews, 2, CD005495. doi: 005491-CD005495.005471.Google Scholar
Spitzer, N. C. (2006). Electrical activity in early neuronal development. Nature, 444(7120), 707.Google Scholar
Sporns, O. (2010). Networks of the brain. Cambridge, MA: MIT Press.Google Scholar
Sretavan, D. W., Shatz, C. J., & Stryker, M. P. (1988). Modification of retinal ganglion cell axon morphology by prenatal infusion of tetrodotoxin. Nature, 336(6198), 468.Google Scholar
Stephen, D. G., Hsu, W. -H., Young, D., Saltzman, E. L., Holt, K. G., Newman, D. J., … Goldfield, E. C. (2012). Multifractal fluctuations in joint angles during infant spontaneous kicking reveal multiplicativity-driven coordination. Chaos, Solitons & Fractals, 45(9–10), 12011219.Google Scholar
Suster, M. L., & Bate, M. (2002). Embryonic assembly of a central pattern generator without sensory input. Nature, 416(6877), 174.Google Scholar
Symington, A. J., & Pinelli, J. (2006). Developmental care for promoting development and preventing morbidity in preterm infants. Cochrane Database Systematic Review, 4, CD001814.Google Scholar
Takahashi, E., Folkerth, R. D., Galaburda, A. M., & Grant, P. E. (2011). Emerging cerebral connectivity in the human fetal brain: An MR tractography study. Cerebral Cortex, 22(2), 455464.Google Scholar
Takahashi, E., Hayashi, E., Schmahmann, J. D., & Grant, P. E. (2014). Development of cerebellar connectivity in human fetal brains revealed by high angular resolution diffusion tractography. Neuroimage, 96, 326333.Google Scholar
Tau, G. Z., & Peterson, B. S. (2010). Normal development of brain circuits. Neuropsychopharmacology, 35(1), 147.Google Scholar
Teramae, J. -N., Tsubo, Y., & Fukai, T. (2012). Optimal spike-based communication in excitable networks with strong-sparse and weak-dense links. Scientific Reports, 2, 485.Google Scholar
Thelen, E., & Smith, L. B. (1994). A dynamic systems approach to the development of cognition and action. Cambridge, MA: MIT Press.Google Scholar
Toga, A. W., Thompson, P. M., & Sowell, E. R. (2006). Mapping brain maturation. TRENDS in Neurosciences, 29(3), 148159.Google Scholar
Tomasello, M. (2009). The cultural origins of human cognition: Cambridge, MA: Harvard University Press.Google Scholar
Tripodi, M., Stepien, A. E., & Arber, S. (2011). Motor antagonism exposed by spatial segregation and timing of neurogenesis. Nature, 479(7371), 61.Google Scholar
Vaal, J., van Soest, A., Hopkins, B., Sie, L., & van der Knaap, M. (2000). Development of spontaneous leg movements in infants with and without periventricular leukomalacia. Experimental Brain Research, 135(1), 94105.Google Scholar
van den Heuvel, M. P., Kersbergen, K. J., de Reus, M. A., Keunen, K., Kahn, R. S., Groenendaal, F., … Benders, M. J. (2014). The neonatal connectome during preterm brain development. Cerebral Cortex, 25(9), 30003013.Google Scholar
van der Meer, A. L. (1997). Keeping the arm in the limelight: Advanced visual control of arm movements in neonates. European Journal of Paediatric Neurology, 1(4), 103108.Google Scholar
Varela, F. J., Rosch, E., & Thompson, E. (1992). The embodied mind: Cognitive science and human experience. Cambridge, MA: MIT Press.Google Scholar
Vauclair, J. (2012). Developpment du jeune enfant, Motricite, Perception, Cognition. Paris: Belin.Google Scholar
von Hofsten, C. (1982). Eye–hand coordination in the newborn. Developmental Psychology, 18(3), 450.Google Scholar
von Hofsten, C. (2007). Action in development. Developmental Science, 10(1), 5460.Google Scholar
Waldmeier, S., Grunt, S., Delgado-Eckert, E., Latzin, P., Steinlin, M., Fuhrer, K., & Frey, U. (2013). Correlation properties of spontaneous motor activity in healthy infants: A new computer-assisted method to evaluate neurological maturation. Experimental Brain Research, 227(4), 433446.CrossRefGoogle ScholarPubMed
Wallin, L., & Eriksson, M. (2009). Newborn individual development care and assessment program (NIDCAP): A systematic review of the literature. Worldviews on Evidence-Based Nursing, 6(2), 5469.Google Scholar
Warp, E., Agarwal, G., Wyart, C., Friedmann, D., Oldfield, C. S., Conner, A., … Isacoff, E. Y. (2012). Emergence of patterned activity in the developing zebrafish spinal cord. Current Biology, 22(2), 93102.Google Scholar
Watts, D. J., & Strogatz, S. H. (1998). Collective dynamics of “small-world” networks. Nature, 393(6684), 440.Google Scholar
Weng, J., McClelland, J., Pentland, A., Sporns, O., Stockman, I., Sur, M., & Thelen, E. (2001). Autonomous mental development by robots and animals. Science, 291(5504), 599600.Google Scholar
White, L. E., Coppola, D. M., & Fitzpatrick, D. (2001). The contribution of sensory experience to the maturation of orientation selectivity in ferret visual cortex. Nature, 411(6841), 1049.Google Scholar
Yamada, Y., Fujii, K., & Kuniyoshi, Y. (2013). Impacts of environment, nervous system and movements of preterms on body map development: Fetus simulation with spiking neural network. Paper presented at the Development and Learning and Epigenetic Robotics (ICDL), 2013 IEEE Third Joint International Conference, Osaka, Japan.Google Scholar
Yamada, Y., Kanazawa, H., Iwasaki, S., Tsukahara, Y., Iwata, O., Yamada, S., & Kuniyoshi, Y. (2016). An embodied brain model of the human foetus. Scientific Reports, 6, 27893.Google Scholar
Yamada, Y., & Kuniyoshi, Y. (2012a). Embodiment guides motor and spinal circuit development in vertebrate embryo and fetus. Paper presented at the Development and Learning and Epigenetic Robotics (ICDL), 2012 IEEE International Conference, San Diego, California.Google Scholar
Yamada, Y., & Kuniyoshi, Y. (2012b). Emergent spontaneous movements based on embodiment: Toward a general principle for early development. Paper presented at the Post-Graduate Conference on Robotics and Development of Cognition, Lausanne, Switzerland.Google Scholar
Yvert, B., Branchereau, P., & Meyrand, P. (2004). Multiple spontaneous rhythmic activity patterns generated by the embryonic mouse spinal cord occur within a specific developmental time window. Journal of Neurophysiology, 91(5), 21012109.Google Scholar
Zoia, S., Blason, L., D’Ottavio, G., Bulgheroni, M., Pezzetta, E., Scabar, A., & Castiello, U. (2007). Evidence of early development of action planning in the human foetus: a kinematic study. Experimental Brain Research, 176(2), 217226.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
×