Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-sjtt6 Total loading time: 0 Render date: 2024-06-28T19:51:51.140Z Has data issue: false hasContentIssue false

7 - Neurotransmitters and neuromodulators

from Section 1 - Making of the brain

Published online by Cambridge University Press:  01 March 2011

Hugo Lagercrantz
Affiliation:
Karolinska Institutet, Stockholm
M. A. Hanson
Affiliation:
Southampton General Hospital
Laura R. Ment
Affiliation:
Yale University, Connecticut
Donald M. Peebles
Affiliation:
University College London
Get access

Summary

Introduction

Neuronal communication is mainly mediated by myriads of synapses via neurotransmitters, although there are also electrical synapses. Neurotransmitters can be defined as chemicals released from neurons that act on specific receptors. However, neuromodulators released from other cells, such as adenosine triphosphate (ATP), adenosine, and prostaglandin, can also affect neuronal signaling. Neurotransmitters and their receptor subtypes are expressed in high amounts during certain stages of development, but then persist in only a few synapses (Parnavelas & Cavanagh, 1988).

The transient increase in expression of a transmitter and receptor subtype in the central nervous system (CNS) may occur during a susceptible developmental time window. (A transient increase occurs, but it may not always occur during a critical window.) Development is characterized by timing and spacing or critical periods when some kind of stimulation is essential for correct development. The transmitters and modulators affect formation of synaptic contacts, maturation of synapses, and structural refinement of connectivity by regulating electrical activity, excitability, and release of neurotrophins (Zhang & Poo, 2001). In particular, at birth a cascade of neurotransmitters and transcriptional factors is activated. For example, the norepinephrine surge at birth may be important for initiating the bonding of the infant to the mother by increasing the ability to sense odors (Sullivan et al., 1994). Imprinting at birth and visual input to form the ocular dominance columns also occur during critical periods, and are probably dependent on the switching on and off of neurotransmitters.

Type
Chapter
Information
The Newborn Brain
Neuroscience and Clinical Applications
, pp. 99 - 120
Publisher: Cambridge University Press
Print publication year: 2010

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

Almqvist, P. M., Akesson, E., Wahlberg, L. U., et al.(1996). First trimester development of the human nigrostriatal dopamine system. Experimental Neurology, 139, 227–37.CrossRefGoogle ScholarPubMed
Alvaro, R. E., Hasan, S. U., Chemtob, S., et al. (2004). Prostaglandins are responsible for the inhibition of breathing observed with a placental extract in fetal sheep. Respiratory Physiology and Neurobiology, 144, 35–44.CrossRefGoogle ScholarPubMed
Andang, M. & Lendahl, U. (2008). Ion fluxes and neurotransmitters signaling in neural development. Current Opinion in Neurobiology, 18, 1–5.CrossRefGoogle ScholarPubMed
Andang, M., Hjerling-Leffler, J., Moliner, A., et al. (2008). Histone H2AX-dependent GABA(A) receptor regulation of stem cell proliferation. Nature, 451, 460–4.CrossRefGoogle ScholarPubMed
Arslan, G., Kontny, E., & Fredholm, B. B. (1997). Down-regulation of adenosine A2A receptors upon NGF-induced differentiation of PC12 cells. Neuropharmacology, 36, 1319–26.CrossRefGoogle ScholarPubMed
Audero, E., Coppi, E., Mlinar, B., et al. (2008). Sporadic autonomic dysregulation and death associated with excessive serotonin autoinhibition. Science, 321, 130–3.CrossRefGoogle ScholarPubMed
Aurich, J. E., Dobrinski, I., Hoppen, H. O., et al. (1990). Beta-endorphin and met-enkephalin in plasma of cattle during pregnancy, parturition and the neonatal period. Journal of Reproduction and Fertility, 89, 605–12.CrossRefGoogle ScholarPubMed
Baraban, S. C., Hollopeter, G., Erickson, J. C., et al. (1997). Knock-out mice reveal a critical antiepileptic role for neuropeptide Y. Journal of Neuroscience, 17, 8927–36.CrossRefGoogle ScholarPubMed
Bar-Peled, O., Gross-Isseroff, R., Ben-Hur, H, et al. (1991). Fetal human brain exhibits a prenatal peak in the density of serotonin 5-HT1A receptors. Neuroscience Letters, 127, 173–6.CrossRefGoogle ScholarPubMed
Bedecs, K., Berthold, M., & Bartfai, T. (1995). Galanin – 10 years with a neuroendocrine peptide. International Journal of Biochemistry and Cell Biology, 27, 337–49.CrossRefGoogle ScholarPubMed
Belhage, B., Hansen, G. H., Elster, L., et al. (1998). Effects of gamma-aminobutyric acid (GABA) on synaptogenesis and synaptic function. Perspectives on Developmental Neurobiology, 5, 235–46.Google ScholarPubMed
Ben-Ari, Y., Khazipov, R., Leinekugel, X., et al. (1997). GABAA, NMDA and AMPA receptors: a developmentally regulated “menage a trois.”Trends in Neurosciences, 20, 523–9.CrossRefGoogle Scholar
Benitez-Diaz, P., Miranda-Contreras, L., Mendoza-Briceno, R. V., et al. (2003). Prenatal and postnatal contents of amino acid neurotransmitters in mouse parietal cortex. Developmental Neuroscience, 25, 366–74.CrossRefGoogle ScholarPubMed
Benke, D., Michel, C., & Mohler, H. (2002). Structure of GABAB receptors in rat retina. Journal of Receptor and Signal Transduction Research, 22, 253–66.CrossRefGoogle ScholarPubMed
Berger-Sweeney, J. & Hohmann, C. F. (1997). Behavioral consequences of abnormal cortical development: insights into developmental disabilities. Behavioural Brain Research, 86, 121–42.CrossRefGoogle ScholarPubMed
Bergstrom, L., Lagercrantz, H., & Terenius, L. (1984). Post-mortem analyses of neuropeptides in brains from sudden infant death victims. Brain Research, 323, 279–85.CrossRefGoogle ScholarPubMed
Berne, R. M. (1986). Adenosine: an important physiological regulator. Trends in Neuroscience, 1, 163–7.Google Scholar
Blaesse, P., Airaksinen, M. S., Rivera, C., et al. (2009). Cation-chloride cotransporters and neuronal function. Neuron, 61, 820–38.CrossRefGoogle ScholarPubMed
Boehning, D. & Snyder, S. H. (2003). Novel neural modulators. Annual Review of Neuroscience, 26, 105–31.CrossRefGoogle ScholarPubMed
Bourgeois, F. (2002). Synaptogenesis. In The Newborn Brain: Neuroscience and Clinical Applications, 1st edn., eds. Lagercrantz, H., Evrard, P., Hanson, M.et al. Cambridge: Cambridge University Press.Google Scholar
Boutrel, B., Franc, B., Hen, R., et al. (1999). Key role of 5-HT1B receptors in the regulation of paradoxical sleep as evidenced in 5-HT1B knock-out mice. Journal of Neuroscience, 19, 3204–12.CrossRefGoogle ScholarPubMed
Bowers, C. W. (1994). Superfluous neurotransmitters?Trends in Neurosciences, 17, 315–20.CrossRefGoogle ScholarPubMed
Boyson, S. J. & Adams, C. E. (1997). D1 and D2 dopamine receptors in perinatal and adult basal ganglia. Pediatric Research, 41, 822–31.CrossRefGoogle ScholarPubMed
Brana, C., Charron, G., Aubert, I., et al. (1995). Ontogeny of the striatal neurons expressing neuropeptide genes in the human fetus and neonate. Journal of Comparative Neurology, 360, 488–505.CrossRefGoogle ScholarPubMed
Burnstock, G. (2008). Purinergic signalling and disorders of the central nervous system. Nature Reviews Drug Discovery, 7, 575–90.CrossRefGoogle ScholarPubMed
Caceres, M., Lachuer, J., Zapala, M. A., et al. (2003). Elevated gene expression levels distinguish human from non-human primate brains. Proceedings of the National Academy of Sciences of the U S A, 100, 13030–5.CrossRefGoogle ScholarPubMed
Cases, O., Vitalis, T., Seif, I., et al. (1996). Lack of barrels in the somatosensory cortex of monoamine oxidase A-deficient mice: role of a serotonin excess during the critical period. Neuron, 16, 297–307.CrossRefGoogle ScholarPubMed
Cavanagh, M. E. & Parnavelas, J. G. (1988). Neurotransmitter differentiation in cortical neurons. In The Making of the Nervous System, eds. Parnavelas, J. G., Stern, C. D. & Stirling, R. V.. London: Oxford University Press, pp. 435–53.Google Scholar
Changeux, J.-P. & Edelstein, S. (2005). Nicotinic Acetylcholine Receptors. New York: Odile Jacob.Google Scholar
Cherubini, E., Gaiarsa, J. L., & Ben-Ari, Y. (1991). GABA: an excitatory transmitter in early postnatal life. Trends in Neurosciences, 14, 515–19.CrossRefGoogle ScholarPubMed
Chugani, D. C. (2002). Role of altered brain serotonin mechanisms in autism. Molecular Psychiatry, 7 (Suppl. 2), S16–17.CrossRefGoogle ScholarPubMed
Cohen, G., Han, Z. Y., Grailhe, R., et al. (2002). Beta 2 nicotinic acetylcholine receptor subunit modulates protective responses to stress: a receptor basis for sleep-disordered breathing after nicotine exposure. Proceedings of the National Academy of Sciences of the U S A, 99, 13272–7.CrossRefGoogle ScholarPubMed
Cohen, G., Roux, J. C., Grailhe, R., et al. (2005). Perinatal exposure to nicotine causes deficits associated with a loss of nicotinic receptor function. Proceedings of the National Academy of Sciences of the U S A, 102, 3817–21.CrossRefGoogle ScholarPubMed
Condie, B. G., Bain, G., Gottlieb, D. I., et al. (1997). Cleft palate in mice with a targeted mutation in the gamma-aminobutyric acid-producing enzyme glutamic acid decarboxylase 67. Proceedings of the National Academy of Sciences of the U S A, 94, 11451–5.CrossRefGoogle ScholarPubMed
Cooper, J. R., Bloom, F. E., & Roth, R. H. (2003). The Biochemical Basis of Neuropharmacology. New York: Oxford University Press.Google Scholar
Corsi, M., Fina, P., & Trist, D. G. (1996). Co-agonism in drug-receptor interaction: illustrated by the NMDA receptors. Trends in Pharmacological Sciences, 17, 220–2.CrossRefGoogle ScholarPubMed
Cote, F., Fligny, C., Bayard, E., et al. (2007). Maternal serotonin is crucial for murine embryonic development. Proceedings of the National Academy of Sciences of the U S A, 104, 329–34.CrossRefGoogle ScholarPubMed
Dare, E., Fetissov, S., Hokfelt, T., et al. (2003). Effects of prenatal exposure to methylmercury on dopamine-mediated locomotor activity and dopamine D2 receptor binding. Naunyn Schmiedebergs Archives of Pharmacology, 367, 500–8.CrossRefGoogle ScholarPubMed
Demarque, M., Represa, A., Becq, H., et al. (2002). Paracrine intercellular communication by a Ca2+- and SNARE-independent release of GABA and glutamate prior to synapse formation. Neuron, 36, 1051–61.CrossRefGoogle ScholarPubMed
Diamond, A. (1996). Evidence for the importance of dopamine for prefrontal cortex functions early in life. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 351, 1483–93; discussion 1494.CrossRefGoogle ScholarPubMed
Diamond, A. (2007). Consequences of variations in genes that affect dopamine in prefrontal cortex. Cerebral Cortex, 17 (Suppl. 1), i161–70.CrossRefGoogle ScholarPubMed
Diamond, A., Briand, L., Fossella, J., et al. (2004). Genetic and neurochemical modulation of prefrontal cognitive functions in children. American Journal of Psychiatry, 161, 125–32.CrossRefGoogle ScholarPubMed
Diaz, R., Fuxe, K., & Ogren, S. O. (1997). Prenatal corticosterone treatment induces long-term changes in spontaneous and apomorphine-mediated motor activity in male and female rats. Neuroscience, 81, 129–40.CrossRefGoogle ScholarPubMed
Dwyer, J. B., Broide, R. S., & Leslie, F. M. (2008). Nicotine and brain development. Birth Defects Research Part C Embryo Today, 84, 30–44.CrossRefGoogle ScholarPubMed
Edwards, H. E. & Burnham, W. M. (2001). The impact of corticosteroids on the developing animal. Pediatric Research, 50, 433–40.CrossRefGoogle ScholarPubMed
Elias, L. A. & Kriegstein, A. R. (2008). Gap junctions: multifaceted regulators of embryonic cortical development. Trends in Neurosciences, 31, 243–50.CrossRefGoogle ScholarPubMed
El-Khodor, B. F. & Boksa, P. (2002). Birth insult and stress interact to alter dopamine transporter binding in rat brain. Neuroreport, 13, 201–6.CrossRefGoogle ScholarPubMed
Fitzgerald, M. (1991). The development of descending brainstem control of spinal cord sensory processing. In The Fetal and Neonatal Brain Stem; Developmental and Clinical Issues, ed. Hanson, M. A.. Cambridge: Cambridge University Press, pp. 127–36.Google Scholar
Foster, G. A. (1992). Ontogeny of transmitters and peptides in the CNS. In Ontogeny of Transmitters and Peptides in the CNS, eds. Björklund, A., Hökfelt, T., & Tohyama, M.. Amsterdam: Elsevier.Google Scholar
Fox, K., Henley, J., & Isaac, J. (1999). Experience-dependent development of NMDA receptor transmission [news; comment]. Nature Neuroscience, 2, 297–9.CrossRefGoogle Scholar
Fredholm, B. B. (1997). Adenosine and neuroprotection. International Review of Neurobiology, 40, 259–80.CrossRefGoogle ScholarPubMed
Fredholm, B. B. & Svenningsson, P. (2003). Adenosine-dopamine interactions: development of a concept and some comments on therapeutic possibilities. Neurology, 61, S5–9.CrossRefGoogle ScholarPubMed
Fredholm, B. B., IJzerman, A. P., Jacobson, K. A., et al. (2001). International Union of Pharmacology. XXV. Nomenclature and classification of adenosine receptors. Pharmacological Reviews, 53, 527–52.Google ScholarPubMed
Frimpter, G. W., Andelman, R. J., & George, W. F. (1969). Vitamin B6-dependency syndromes. New horizons in nutrition. American Journal of Clinical Nutrition, 22, 794–805.CrossRefGoogle ScholarPubMed
Fritschy, J. M., Paysan, J., Enna, A., et al. (1994). Switch in the expression of rat GABAA-receptor subtypes during postnatal development: an immunohistochemical study. Journal of Neuroscience, 14, 5302–24.CrossRefGoogle Scholar
Fujii, M., Umezawa, K., & Arata, A. (2006). Adrenaline contributes to prenatal respiratory maturation in rat medulla-spinal cord preparation. Brain Research, 1090, 45–50.CrossRefGoogle ScholarPubMed
Gallo, V. & Haydar, T. (2003). GABA: exciting again in its own right. Journal of Physiology, 550, 665.CrossRefGoogle ScholarPubMed
Gaspar, P., Cases, O., & Maroteaux, L. (2003). The developmental role of serotonin: news from mouse molecular genetics. Nature Reviews Neuroscience, 4, 1002–12.CrossRefGoogle ScholarPubMed
Gaveriaux-Ruff, C. & Kieffer, B. L. (2002). Opioid receptor genes inactivated in mice: the highlights. Neuropeptides, 36, 62–71.CrossRefGoogle ScholarPubMed
Gilmore, J. H. & Jarskog, L. F. (1997). Exposure to infection and brain development: cytokines in the pathogenesis of schizophrenia. Schizophrenia Research, 24, 365–7.CrossRefGoogle ScholarPubMed
Goridis, C. & Rohrer, H. (2002). Specification of catecholaminergic and serotonergic neurons. Nature Reviews, 3, 531–41.CrossRefGoogle ScholarPubMed
Hagberg, H., Bona, E., Gilland, E., et al. (1997). Hypoxia-ischaemia model in the 7-day-old rat: possibilities and shortcomings. Acta Paediatrica Supplement, 422, 85–8.CrossRefGoogle ScholarPubMed
Hansel, D. E., Eipper, B. A., & Ronnett, G. V. (2001). Neuropeptide Y functions as a neuroproliferative factor. Nature, 410, 940–4.CrossRefGoogle ScholarPubMed
Happe, H. K., Coulter, C. L., Gerety, M. E., et al. (2004). Alpha-2 adrenergic receptor development in rat CNS: an autoradiographic study. Neuroscience, 123, 167–78.CrossRefGoogle Scholar
Hellstrom-Lindahl, E., Gorbounova, O., Seiger, A., et al. (1998). Regional distribution of nicotinic receptors during prenatal development of human brain and spinal cord. Brain Research Developmental Brain Research, 108, 147–60.CrossRefGoogle ScholarPubMed
Henschel, O., Gipson, K. E., & Bordey, A. (2008). GABAA receptors, anesthetics and anticonvulsants in brain development. CNS Neurological Disorders Drug Targets, 7, 211–24.CrossRefGoogle ScholarPubMed
Herlenius, E. & Lagercrantz, H. (1999). Adenosinergic modulation of respiratory neurones in the neonatal rat brainstem in vitro. Journal of Physiology, 518 (Pt 1), 159–72.CrossRefGoogle ScholarPubMed
Herlenius, E. & Lagercrantz, H. (2004). Development of neurotransmitter systems during critical periods. Experimental Neurology, 190 (Suppl. 1), S8–21.CrossRefGoogle ScholarPubMed
Herlenius, E., Aden, U., Tang, L. Q., et al. (2002). Perinatal respiratory control and its modulation by adenosine and caffeine in the rat. Pediatric Research, 51, 4–12.CrossRefGoogle ScholarPubMed
Herlitze, S., Villaroel, A., Witzemann, V., et al. (1996). Structural determinants of channel conductance in fetal and adult rat muscle acetylcholine receptors. Journal of Physiology, 492, 775–87.CrossRefGoogle ScholarPubMed
Herregodts, P., Velkeniers, B., Ebinger, G., et al. (1990). Development of monoaminergic neurotransmitters in fetal and postnatal rat brain: analysis by HPLC with electrochemical detection. Journal of Neurochemistry, 55, 774–9.CrossRefGoogle ScholarPubMed
Herschkowitz, N., Kagan, J., & Zilles, K. (1997). Neurobiological bases of behavioral development in the first year. Neuropediatrics, 28, 296–306.CrossRefGoogle ScholarPubMed
Hofstetter, A. O., Saha, S., Siljehav, V., et al. (2007). The induced prostaglandin E2 pathway is a key regulator of the respiratory response to infection and hypoxia in neonates. Proceedings of the National Academy of Sciences of the U S A, 104, 9894–9.CrossRefGoogle ScholarPubMed
Hokfelt, T., Bartfai, T., & Bloom, F. (2003). Neuropeptides: opportunities for drug discovery. Lancet Neurology, 2, 463–72.CrossRefGoogle ScholarPubMed
Hoyer, D., Hannon, J. P., & Martin, G. R. (2002). Molecular, pharmacological and functional diversity of 5-HT receptors. Pharmacology Biochemistry and Behavior, 71, 533–54.CrossRefGoogle ScholarPubMed
Iglesias, R., Dahl, G., Qiu, F., et al. (2009). Pannexin 1: the molecular substrate of astrocyte “hemichannels”. Journal of Neuroscience, 29, 7092–7.CrossRefGoogle ScholarPubMed
Insel, T. R. & Young, L. J. (2001). The neurobiology of attachment. Nature Reviews Neuroscience, 2, 129–36.CrossRefGoogle ScholarPubMed
Jarskog, L. F., Xiao, H., Wilkie, M. B., et al. (1997). Cytokine regulation of embryonic rat dopamine and serotonin neuronal survival in vitro. International Journal of Developmental Neuroscience, 15, 711–16.CrossRefGoogle ScholarPubMed
Johnston, M. V. & Silverstein, F. S. (1998). Development of neurotransmitters. In Fetal and Neonatal Physiology, eds. Polin, R. A. & Fox, W. W.. Philadelphia: Saunders, W. B., pp. 2116–17.
Koch, C. (2004). The Quest for Consciousness. Englewood:Roberts and Company.Google Scholar
Kopecky, E. A. & Koren, G. (1998). Maternal drug abuse: effects on the fetus and neonate. In Fetal and Neonatal Physiology, eds. Polin, R. A. & Fox, W. W.. Philadelphia: W. B. Saunders, pp. 203–20.Google Scholar
Kriegstein, A. R. & Owens, D. F. (2001). GABA may act as a self-limiting trophic factor at developing synapses. Science's STKE: Signal Transduction Knowledge Environment, 95, pe1.Google Scholar
Kurosawa, A., Kageyama, H., John, T. M., et al. (1980). Effect of neonatal hydrocortisone treatment on brain monoamines in developing rats. Japanese Journal of Pharmacology, 30, 213–20.CrossRefGoogle ScholarPubMed
Lagercrantz, H. (1996). Stress, arousal and gene activation at birth. News in Physiological Sciences, 11, 214–18.Google Scholar
Lattimore, K. A., Donn, S. M., Kaciroti, N., et al. (2005). Selective serotonin reuptake inhibitor (SSRI) use during pregnancy and effects on the fetus and newborn: a meta-analysis. Journal of Perinatology, 25, 595–604.CrossRefGoogle ScholarPubMed
Douarin, N. M. (1981). Plasticity in the development of the peripheral nervous system. Ciba Foundation Symposium, 83, 19–50.Google ScholarPubMed
Lebrand, C., Cases, O., Adelbrecht, C., et al. (1996). Transient uptake and storage of serotonin in developing thalamic neurons. Neuron, 17, 823–35.CrossRefGoogle ScholarPubMed
Letinic, K., Zoncu, R., & Rakic, P. (2002). Origin of GABAergic neurons in the human neocortex. Nature, 417, 645–9.CrossRefGoogle ScholarPubMed
Li, L. & Moore, P. K. (2007). An overview of the biological significance of endogenous gases: new roles for old molecules. Biochemical Society Transactions, 35, 1138–41.CrossRefGoogle ScholarPubMed
Lichtensteiger, W. (1998). Developmental neuropharmacology. In Fetal and Neonatal Physiology, eds. Polin, R. A. & Fox, W. W.. Philadelphia: W. B. Saunders, pp. 226–39.Google Scholar
Lipton, S. A. & Nakanishi, N. (1999). Shakespeare in love – with NMDA receptors?Nature Medicine, 5, 270–1.CrossRefGoogle ScholarPubMed
Liu, X., Hashimoto-Torii, K., Torii, M., et al. (2008). The role of ATP signaling in the migration of intermediate neuronal progenitors to the neocortical subventricular zone. Proceedings of the National Academy of Sciences of the U S A, 105, 11802–7.CrossRefGoogle ScholarPubMed
Maggi, L., Magueresse, C., Changeux, J. P., et al. (2003). Nicotine activates immature “silent” connections in the developing hippocampus. Proceedings of the National Academy of Sciences of the U S A, 100, 2059–64.CrossRefGoogle ScholarPubMed
Mangelsdorf, D. J., Thummel, C., Beato, M., et al. (1995). The nuclear receptor superfamily: the second decade. Cell, 83, 835–9.CrossRefGoogle ScholarPubMed
Marin, O. & Rubenstein, J. L. (2001). A long, remarkable journey: tangential migration in the telencephalon. Nature Reviews Neuroscience, 2, 780–90.CrossRefGoogle ScholarPubMed
Matsuishi, T., Nagamitsu, S., Yamashita, Y., et al. (1997). Decreased cerebrospinal fluid levels of substance P in patients with Rett syndrome. Annals of Neurology, 42, 978–81.CrossRefGoogle ScholarPubMed
Matthews, S. G. (2001). Antenatal glucocorticoids and the developing brain: mechanisms of action. Seminars in Neonatology, 6, 309–17.CrossRefGoogle ScholarPubMed
McDonald, J. W. & Johnston, M. V. (1990). Physiological and pathophysiological roles of excitatory amino acids during central nervous system development. Brain Research Brain Research Reviews, 15, 41–70.CrossRefGoogle ScholarPubMed
Meyer, U. & Feldon, J. (2009). Neural basis of psychosis-related behaviour in the infection model of schizophrenia. Behavioural Brain Research, 204, 322–34.CrossRefGoogle ScholarPubMed
Miles, R. (1999). Neurobiology. A homeostatic switch. Nature, 397, 215–16.CrossRefGoogle ScholarPubMed
Morita, Y. (1992). Ontogenic and differential expression of the preproenkephalin and predynorphin genes in the rat brain. In Ontogeny of Transmitters and Peptides in the CNS, eds. Björklund, A, Hökfelt, T, & Tohyama, M. Amsterdam: Elsevier, pp. 257–95.Google Scholar
Naeff, B., Schlumpf, M., & Lichtensteiger, W. (1992). Pre- and postnatal development of high-affinity [3H]nicotine binding sites in rat brain regions: an autoradiographic study. Brain Research Developmental Brain Research, 68, 163–74.CrossRefGoogle Scholar
Naqui, S. Z. H., Harris, B. S., Thomaidou, D., et al. (1999). The noradrenergic system influences in fate of Cajal-Retzius cells in the developing cerebral cortex. Developmental Brain Research, 113, 75–82.CrossRefGoogle ScholarPubMed
Noctor, S. C., Flint, A. C., Weissman, T. A., et al. (2001). Neurons derived from radial glial cells establish radial units in neocortex. Nature, 409, 714–20.CrossRefGoogle ScholarPubMed
Noctor, S. C., Martinez-Cerdeno, V., Ivic, L., et al. (2004). Cortical neurons arise in symmetric and asymmetric division zones and migrate through specific phases. Nature Neuroscience, 7, 136–44.CrossRefGoogle ScholarPubMed
O'Brien, R. A., Da Prada, M., & Pletscher, A. (1972). The ontogenesis of catecholamines and adenosine-5′-triphosphate in the adrenal medulla. Life Sciences, 11, 749–59.CrossRefGoogle ScholarPubMed
Olney, J. W., Wozniak, D. F., Farber, N. B., et al. (2002). The enigma of fetal alcohol neurotoxicity. Annals of Medicine, 34, 109–19.CrossRefGoogle ScholarPubMed
Olson, L. & Seiger, A. (1972). Early prenatal ontogeny of central monoamine neurons in the rat: fluorescence histochemical observations. Zeitschrift für Anatomie und Entwicklungsgeschichte, 137, 301–16.CrossRefGoogle ScholarPubMed
Onimaru, H., Herlenius, E., & Homma, I. (1999). GABA-dependent responses of respiratory neurons in the fetal rat medulla. Neuroscience Research, 23 (Suppl.), 77.Google Scholar
Owens, D. F. & Kriegstein, A. R. (2002a). Developmental neurotransmitters?Neuron, 36, 989–91.CrossRefGoogle ScholarPubMed
Owens, D. F. & Kriegstein, A. R. (2002b). Is there more to GABA than synaptic inhibition?Nature Reviews Neuroscience, 3, 715–27.CrossRefGoogle ScholarPubMed
Parnavelas, J. G. & Cavanagh, M. E. (1988). Transient expression of neurotransmitters in the developing neocortex. Trends in Neurosciences, 11, 92–3.CrossRefGoogle ScholarPubMed
Paterson, D. S., Trachtenberg, F. L., Thompson, E. G., et al. (2006). Multiple serotonergic brainstem abnormalities in sudden infant death syndrome. JAMA Journal of the American Medical Association, 296, 2124–32.CrossRefGoogle ScholarPubMed
Patterson, P. H. & Chun, L. L. (1977). The induction of acetylcholine synthesis in primary cultures of dissociated rat sympathetic neurons. II. Developmental aspects. Developmental Biology, 60, 473–81.CrossRefGoogle ScholarPubMed
Pauly, J. R. & Slotkin, T. A. (2008). Maternal tobacco smoking, nicotine replacement and neurobehavioural development. Acta Paediatrica, 97, 1331–7.CrossRefGoogle ScholarPubMed
Pendleton, R. G., Rasheed, A., Roychowdhury, R., et al. (1998). A new role for catecholamines: ontogenesis. Trends in Pharmacological Sciences, 19, 248–51.Google ScholarPubMed
Peyronnet, J., Dalmaz, Y., Ehrstrom, M., et al. (2002). Long-lasting adverse effects of prenatal hypoxia on developing autonomic nervous system and cardiovascular parameters in rats. Pflügers Archiv, 443, 858–65.CrossRefGoogle ScholarPubMed
Qu, Y., Vadivelu, S., Choi, L., et al. (2003). Neurons derived from embryonic stem (ES) cells resemble normal neurons in their vulnerability to excitotoxic death. Experimental Neurology, 184, 326–36.CrossRefGoogle ScholarPubMed
Ramanathan, S., Puig, M. M., & Turndorf, H. (1989). Plasma beta-endorphin levels in the umbilical cord blood of preterm human neonates. Biology of the Neonate, 56, 117–20.CrossRefGoogle ScholarPubMed
Rasmussen, S. G., Choi, H. J., Rosenbaum, D. M., et al. (2007). Crystal structure of the human beta2 adrenergic G-protein-coupled receptor. Nature, 450, 383–7.CrossRefGoogle ScholarPubMed
Ringstedt, T., Tang, L. Q., Persson, H., et al. (1995). Expression of c-fos, tyrosine hydroxylase, and neuropeptide mRNA in the rat brain around birth: effects of hypoxia and hypothermia. Pediatric Research, 37, 15–20.CrossRefGoogle ScholarPubMed
Rivera, C., Voipio, J., Payne, J. A., et al. (1999). The K+/Cl– co-transporter KCC2 renders GABA hyperpolarizing during neuronal maturation. Nature, 397, 251–5.CrossRefGoogle Scholar
Rivera, C., Li, H., Thomas-Crusells, J., et al. (2002). BDNF-induced TrkB activation down-regulates the K+/Cl– cotransporter KCC2 and impairs neuronal Cl− extrusion. Journal of Cell Biology, 159, 747–52.CrossRefGoogle Scholar
Sakanaka, M. (1992). Development of neuronal elements with substance P-like immunoreactivity in the central nervous system. In Ontogeny of Transmitters and Peptides in the CNS, eds. Björklund, A, Hökfelt, T., & Tohyama, M.. Amsterdam: Elsevier, pp. 197–250.Google Scholar
Sapolsky, R. M. (1997). The importance of a well-groomed child [comment]. Science, 277, 1620–1.CrossRefGoogle Scholar
Sayer, A. A., Cooper, C., & Barker, D. J. (1997). Is lifespan determined in utero?Archives of Disease in Childhood Fetal and Neonatal Edition, 77, F162–4.CrossRefGoogle ScholarPubMed
Schmidt, B. (1999). Methylxanthine therapy in premature infants: sound practice, disaster, or fruitless byway?Journal of Pediatrics, 135, 526–8.CrossRefGoogle ScholarPubMed
Schmidt, B., Roberts, R. S., Davis, P., et al. (2006). Caffeine therapy for apnea of prematurity. New England Journal of Medicine, 354, 2112–21.CrossRefGoogle ScholarPubMed
Schmidt, B., Roberts, R. S., Davis, P., et al. (2007). Long-term effects of caffeine therapy for apnea of prematurity. New England Journal of Medicine, 357, 1893–902.CrossRefGoogle ScholarPubMed
Semba, K. (1992). Development of central cholinergic neurons. In Ontogeny of Transmitters and Peptides in the CNS, eds. Björklund, A., Hökfelt, T, & Tohyama, M.. Amsterdam: Elsevier, pp. 33–62.Google Scholar
Semba, K. (2004). Phylogenetic and ontogenetic aspects of the basal forebrain cholinergic neurons and their innervation of the cerebral cortex. Progress in Brain Research, 145, 3–43.Google ScholarPubMed
Sernagor, E., Young, C., & Eglen, S. J. (2003). Developmental modulation of retinal wave dynamics: shedding light on the GABA saga. Journal of Neuroscience, 23, 7621–9.CrossRefGoogle ScholarPubMed
Sieghart, W. & Sperk, G. (2002). Subunit composition, distribution and function of GABA(A) receptor subtypes. Current Topics in Medicinal Chemistry, 2, 795–816.CrossRefGoogle ScholarPubMed
Simons, C. T., Gogineni, A. G., Iodi Carstens, M., et al. (2002). Reduced aversion to oral capsaicin following neurotoxic destruction of superficial medullary neurons expressing NK-1 receptors. Brain Research, 945, 139–43.CrossRefGoogle ScholarPubMed
Stefan, M. D. & Murray, R. M. (1997). Schizophrenia: developmental disturbance of brain and mind?Acta Paediatrica Supplement, 422, 112–16.CrossRefGoogle ScholarPubMed
Stevens, B., Porta, S., Haak, L. L., et al. (2002). Adenosine: a neuron-glial transmitter promoting myelination in the CNS in response to action potentials. Neuron, 36, 855–68.CrossRefGoogle ScholarPubMed
Sullivan, R. M., Wilson, D. A., Lemon, C., et al. (1994). Bilateral 6-OHDA lesions of the locus coeruleus impair associative olfactory learning in newborn rats. Brain Research, 643, 306–9.CrossRefGoogle ScholarPubMed
Sundstrom, E., Kolare, S., Souverbie, F., et al. (1993). Neurochemical differentiation of human bulbospinal monoaminergic neurons during the first trimester. Brain Research Developmental Brain Research, 75, 1–12.CrossRefGoogle ScholarPubMed
Tang, Y. P., Shimizu, E., Dube, G. R., et al. (1999). Genetic enhancement of learning and memory in mice. Nature, 401, 63–9.CrossRefGoogle ScholarPubMed
Thomas, S. A. & Palmiter, R. D. (1997). Impaired maternal behavior in mice lacking norepinephrine and epinephrine. Cell, 91, 583–92.CrossRefGoogle ScholarPubMed
Thomas, S. A., Matsumoto, A. M., & Palmiter, R. D. (1995). Noradrenaline is essential for mouse fetal development. Nature, 374, 643–6.CrossRefGoogle ScholarPubMed
Trosko, J. E. (2007). Gap junctional intercellular communication as a biological “rosetta stone” in understanding, in a systems biological manner, stem cell behavior, mechanisms of epigenetic toxicology, chemoprevention and chemotherapy. Journal of Membrane Biology, 218, 93–100.CrossRefGoogle Scholar
Tyzio, R., Cossart, R., Khalilov, I., et al. (2006). Maternal oxytocin triggers a transient inhibitory switch in GABA signaling in the fetal brain during delivery. Science, 314, 1788–92.CrossRefGoogle ScholarPubMed
Ubink, R., Calza, L., & Hokfelt, T. (2003). “Neuro”-peptides in glia: focus on NPY and galanin. Trends in Neurosciences, 26, 604–9.CrossRefGoogle ScholarPubMed
Verhage, M., Maia, A. S., Plomp, J. J., et al. (2000). Synaptic assembly of the brain in the absence of neurotransmitter secretion. Science, 287, 864–9.CrossRefGoogle ScholarPubMed
Verney, C., Lebrand, C., & Gaspar, P. (2002). Changing distribution of monoaminergic markers in the developing human cerebral cortex with special emphasis on the serotonin transporter. Anatomical Record, 267, 87–93.CrossRefGoogle ScholarPubMed
Wang, D. D. & Kriegstein, A. R. (2008). GABA regulates excitatory synapse formation in the neocortex via NMDA receptor activation. Journal of Neuroscience, 28, 5547–58.CrossRefGoogle ScholarPubMed
Wang, F. & Lidow, M. S. (1997). Alpha 2A-adrenergic receptors are expressed by diverse cell types in the fetal primate cerebral wall. Journal of Comparative Neurology, 378, 493–507.3.0.CO;2-Y>CrossRefGoogle ScholarPubMed
Wang, Y. Q., Li, J. S., Li, H. M., et al. (1992). Ontogeny of pro-opiomelanocortin (POMC)-derived peptides in the brain and pituitary. In Ontogeny of Transmitters and Peptides in the CNS, eds. Björklund, A., Hökfelt, T., & Tohyama, M.. Amsterdam: Elsevier, pp. 297–323.Google Scholar
Weinstock, M. (1997). Does prenatal stress impair coping and regulation of hypothalamic-pituitary-adrenal axis?Neuroscience and Biobehavioral Reviews, 21, 1–10.CrossRefGoogle ScholarPubMed
Weitzman, M., Gortmaker, S., & Sobol, A. (1992). Maternal smoking and behavior problems of children. Pediatrics, 90, 342–9.Google ScholarPubMed
Wickström, H. R., Holgert, H., Hokfelt, T., et al. (1999). Birth-related expression of c-fos, c-jun and substance P mRNAs in the rat brainstem and pia mater: possible relationship to changes in central chemosensitivity. Brain Research Developmental Brain Research, 112, 255–66.CrossRefGoogle ScholarPubMed
Williams, G. V. & Goldman-Rakic, P. S. (1995). Modulation of memory fields by dopamine D1 receptors in prefrontal cortex. Nature, 376, 572–5.CrossRefGoogle ScholarPubMed
Winn, H. R., Rubio, G. R., & Berne, R. M. (1981a). The role of adenosine in the regulation of cerebral blood flow. Journal of Cerebral Blood Flow and Metabolism, 1, 239–44.CrossRefGoogle ScholarPubMed
Winn, H. R., Rubio, R., & Berne, R. M. (1981b). Brain adenosine concentration during hypoxia in rats. American Journal of Physiology, 241, H235–42.Google ScholarPubMed
Young, C. & Olney, J. W. (2006). Neuroapoptosis in the infant mouse brain triggered by a transient small increase in blood alcohol concentration. Neurobiology of Disease, 22, 548–54.CrossRefGoogle ScholarPubMed
Zhang, L. I. & Poo, M. M. (2001). Electrical activity and development of neural circuits. Nature Neuroscience, 4 (Suppl.), 1207–14.CrossRefGoogle ScholarPubMed
Zheng, T. M., Zhu, W. J., Puia, G., et al. (1994). Changes in gamma-aminobutyrate type A receptor subunit mRNAs, translation product expression, and receptor function during neuronal maturation in vitro. Proceedings of the National Academy of Sciences of the U S A, 91, 10952–6.CrossRefGoogle ScholarPubMed
Zhou, Q. Y. & Palmiter, R. D. (1995). Dopamine-deficient mice are severely hypoactive, adipsic, and aphagic. Cell, 83, 1197–209.CrossRefGoogle ScholarPubMed
Zhou, Q. Y., Quaife, C. J., & Palmiter, R. D. (1995). Targeted disruption of the tyrosine hydroxylase gene reveals that catecholamines are required for mouse fetal development. Nature, 374, 640–3.CrossRefGoogle ScholarPubMed

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
×