Hostname: page-component-848d4c4894-cjp7w Total loading time: 0 Render date: 2024-07-01T06:01:37.864Z Has data issue: false hasContentIssue false

The Role of Cyclic Nucleotides in the CNS

Published online by Cambridge University Press:  18 September 2015

John W. Phillis*
Affiliation:
Department of Physiology, College of Medicine, University of Saskatchewan, Saskatoon, Saskatchewan
*
Dept. of Physiology, College of Medicine, University of Saskatchewan, Saskatoon, Saskatchewan, Canada. S7N 0W0
Rights & Permissions [Opens in a new window]

Summary:

Core share and HTML view are not available for this content. However, as you have access to this content, a full PDF is available via the ‘Save PDF’ action button.

On the basis of the information presented in this review, it is difficult to reach any firm decision regarding the role of cyclic AMP (or cyclic GMP) in synaptic transmission in the brain. While it is clear that cyclic nucleotide levels can be altered by the exposure of neural tissues to various neurotransmitters, it would be premature to claim that these nucleotides are, or are not, essential to the transmission process in the pre- or postsynaptic components of the synapse. In future experiments with cyclic AMP it will be necessary to consider more critically whether the extracellularly applied nucleotide merely provides a source of adenosine and is thus activating an extracellularly located adenosine receptor, or whether it is actually reaching the hypothetical sites at which it might act as a second messenger. The application of cyclic AMP by intracellular injection techniques should minimize this particular problem, although possibly at the expense of new difficulties. Prior blockade of the adenosine receptor with agents such as theophylline or adenine xylofuranoside may also assist in the categorization of responses to extracellularly applied cyclic AMP as being a result either of activation of the adenosine receptor or of some other mechanism. Ultimately, the development of highly specific inhibitors for adenylate cyclase should provide a firm basis from which to draw conclusions about the role of cyclic AMP in synaptic transmission. Similar considerations apply to the actions of cyclic GMP and the role of its synthesizing enzyme, guanylale cyclase.

The use of phosphodiesterase inhibitors in studies on cyclic nucleotides must also be approached with caution. The diverse actions of many of these compounds, which include calcium mobilization and block of adenosine uptake, could account for many of the results that have been reported in the literature.

Type
Research Article
Copyright
Copyright © Canadian Neurological Sciences Federation 1977

References

REFERENCES

Ahn, H.S. Mishra, R.K. Demirjian, C. and Makman, M.H. (1976). Catecholamine-sensitive adenylate cyclase in frontal cortex of primate brain. Brain. Res. 116, 437454. Google Scholar
Amer, M.S. and Kreighbaum, E. (1975). Cyclic nucleotide phosphodiesterase: properties, activators, inhibitors, structure-activity relationships and possible role in drug development. J. Pharm. Sci. 64, 137. CrossRefGoogle Scholar
Andén, N.-E. Dahlström, A., Fuxe, K., Larsson, K., Olson, L. and Ungerstedt, U. (1966). Ascending monoamine neurons to the telencephalon and diencephalon. Acta physiol. scand. 67, 313326.Google Scholar
Amderson, E.G., Haas, H. and Hosli, L. (1973). Comparison of effects of noradrenaline and histamine with cyclic AMP on brain stem neurones. Brain Res. 49, 471475.Google Scholar
Anlezark, G.M., Crow, T.J. and Greenway, A.P. (1973). Impaired learning and decreased cortical norepinephrine after bilateral locus coeruleus lesions. Science 181, 682684.Google Scholar
Appleman, M.M. and Terasaki, W.L. (1975). Regulation of cyclic nucleotide phosphodiesterase. Advanc. Cyclic Nucleotide Res. 5, 153162.Google Scholar
Appleman, M.M., Thompson, W.J. and Russell, T.R. (1973). Cyclic nucleotide phosphodiesterases. Advanc. Cyclic Nucleotide Res. 3, 6698.Google Scholar
Akasu, T. and Koketsu, K. Effects of dibutyryl cyclic AMP and theophylline on the bullfrog sympathetic ganglion cells. Br. J. Pharmac. (in press).Google Scholar
Araki, T. and Terzuolo, C.A. (1962). Membrane currents in spinal neurones associated with the action potential and synaptic activity. J. Neurophysiol. 25. 772789.Google Scholar
Ashman, D.F., Lipton, R., Melicow, M.M. and Price, T.D. (1963). Isolation of adenosine 3’ 5’-monophosphate and guanosine 3’, 5’-monophosphate from rat urine. Biochem. Biophys. Res. Commun. 11, 330334.CrossRefGoogle Scholar
Baker, P.F. (1972). Transport and metabolism of calcium ions. Prog. Biophys. Mol. Biol. 24, 177223.CrossRefGoogle ScholarPubMed
Baudry, M. Martres, M.P. and Schwartz, J.C. (1975). H1 and H2 receptors in the histamine-induced accumulation of cyclic AMP in guinea pig brain slices. Nature 253, 362364.Google Scholar
Berger, B., Tassin, J.P., Blanc, O., Moyne, M.A. and Thierry, A.M. (1974). Histochemical confirmation for dopaminergic innervation of rat cerebral cortex after destruction of the noradrenergic ascending pathways. Brain Res. 81, 332337. CrossRefGoogle ScholarPubMed
Berkowitz, B.A., Tarver, J.H. and Spector, S. (1970). Release of norepinephrine in central nervous system by theophylline and caffeine. Eur. J. Pharmac. 10, 6471.Google Scholar
Bernardi, G., Floris, V., Marciani, M.G., Morocutti, C. and Stanzione, P. (1976). The action of acetylcholine and L-glutamic acid on rat caudate neurons. Brain Res. 114, 134138.Google Scholar
Berne, R.M., Rubio, R. and Curnish, R. (1974). Release of adenosine from is-chaemic brain. Circ. Res. 35, 262272.Google Scholar
Berridge, M.J. (1975). The interaction of cyclic nucleotides and calcium in the control of cellular activity. Advanc. Cyclic Nucleotide Res. 6, 198.Google ScholarPubMed
Berti, F., Trabucchi, M., Bernareggi, V. and Fumagalli, R. (1972). The effects of prostaglandins on cyclic-AMP formation in cerebral cortex of different mammalian species. Pharmacol. Res. Commun. 4, 253259.Google Scholar
Bevan, P., Bradshaw, C.M. and Szabadi, E. (1976). Neuronal responses to adrenoceptor agonists in the cerebral cortex: evidence for excitatory (α -adrenoceptors and inhibitory β-adrenoceptors. Br. J. Pharmac. 58, 418 P.Google Scholar
Blinks, J.R., Olsen, C.B., Jewell, B.R. and Braveny, P. (1972). Influence of caffeine and other methylxanthines on mechanical properties of isolated mammalian heart muscle. Evidence for a dual mechanism of action. Circ. Res. 30, 367391.Google Scholar
Bloom, F.E. (1974). To spritz or not to spritz: the doubtful value of aimless iontophoresis. Life Sci. 14, 18191834.Google Scholar
Bloom, F.E. (1975).The role of cyclic nucleotides in central synaptic function. Rev. Physiol. Biochem. Pharmacol. 74. 1103.Google ScholarPubMed
Bloom, F.E., Hoffer, B.J., Battenberg, E.R., Siggins, G.R., Steiner, A.L., Parker, C.W. and Wedner, H.J. (1972). Adenosine 3’, 5’-monophosphate is localized in cerebellar neurons: immunofluorescence evidence. Science 177, 436438.Google Scholar
Bloom, F.E., Hoffer, B.J. and Siggins, G.R. (1971). Studies on norepinephrine-containing afférents to cerebellar Purkinje cells of rat cerebellum. I. Localization of the fibers and their synapses. Brain Res. 25, 501521.CrossRefGoogle Scholar
Bloom, F.E., Siggins, G.R., Hoffer, B.J., Segal, M. and Oliver, A.P. (1975). Cyclic nucleotides in the central synaptic actions of catecholamines. J. Cyclic Nucleotide Res. 5, 603618.Google Scholar
Blume, A.J. and Foster, C.J. (1975). Mouse neuroblastoma adenylate cyclase. Adenosine and adenosine analogues as potent effectors of adenylate cyclase activity. J. Biol. Chem. 250, 50035008.CrossRefGoogle ScholarPubMed
Bianchi, C.P. (1968). Cell Calcium. New York: Appleton-Century-Crofts.Google Scholar
Borle, A.B. (1973). Calcium metabolism at the cellular level. Fedn. Proc. 32, 19441950.Google Scholar
Borle, A.B. (1974). Cyclic AMP stimulation of calcium efflux from kidney, liver, and heart mitochondria. J. Memb. Biol. 16, 221236.Google Scholar
Bradham, L.S., Holt, D.A. and Sims, M. (1970). The effect of Ca2+ on the adenyl cyclase of calf brain. Biochem. Biophys. Acta, 201, 250260.CrossRefGoogle ScholarPubMed
Bradshaw, C.M., Szabadi, E. and Roberts, M.H.T. (1973). The reflection of ejecting and retaining currents in the time course of neuronal responses to microelectrophoretically applied drugs. J. Pharm. Pharmacol. 25, 513520.Google Scholar
Breckenridge, B.M. (1964). The measurement of cyclic adenylate in tissues. Proc. Natl. Acad. Sci. (U.S.A.) 57, 15801586.Google Scholar
Breckenridge, B.M., Burn, J.H. and Matschinsky, F.M. (1967). Theophylline, epinephrine and neostigmine facilitation of neuromuscular transmission. Proc. Natl. Acad. Sci. (U.S.A.) 57, 18931897.Google Scholar
Breckenridge, B.M. and Johnston, R.E. (1969). Cyclic 3’, 5’-nucleotide phosphodiesterase in brain. J. Histochem. Cytochem. 17, 505511.Google Scholar
Bressler, B.H., Phillis, J.W. and Kozachuk, W. (1975). Noradrenaline stimulation of a membrane pump in frog skeletal muscle. Eur. J. Pharmac. 33, 201204.Google Scholar
Bülbring, E. and Kuriyama, H. (1963). Effects of changes in ionic environment on the action of acetylcholine and adrenaline on the smooth muscle cells of guinea-pig taenia coli. J. Physiol. (Lond.) 166, 5974.Google Scholar
Bülbring, E. and Tomita, T. (1969a). Increase of membrane conductance by adrenaline in the smooth muscle of guinea-pig taenia coli. Proc. Roy. Soc. B. 172, 89102.Google Scholar
Bülbring, E. and Tomita, T. (1969b). Suppression of spontaneous spike generation by catecholamines in the smooth muscle of the guinea-pig taenia coli. Proc. Roy. Soc. B. 172, 103119.Google Scholar
Bülbring, E. and Tomita, T. (1969c). Effect of calcium, barium and manganese on the action of adrenaline in the smooth muscle of the guinea pig taenia coli. Proc. Roy. Soc. B. 172, 121136.Google Scholar
Bunney, B.S. and Aghajanian, G.K. (1973). Electrophysiological effects of amphetamine in dopaminergic neurons. In: Frontiers in Catecholamine Research, Usdin, E. and Snyder, S.H. Eds., New York: Pergamon Press, pp. 957962.Google Scholar
Bunney, B.S. and Aghajanian, G.K. (1976). Dopamine and norepinephrine innervated cells in the rat prefrontal cortex: pharmacological differentiation using mic-roiontophoretic techniques. Life Sci. 19, 17831792.Google Scholar
Burkard, W.P. (1972). Catecholamine induced increase of cyclic adenosine 3’, 5’-monophosphate in rat brain in vivo. J. Neurochem. 19, 26152619.Google Scholar
Burkard, W.P. (1975). Adenylate cyclase in the central nervous system. Progr. Neurobiol. 4, 241267.Google Scholar
Burkard, W.P. Pieri, L. and Haefely, W. (1976). In vivo changes of guanosine 3’, 5’-cyclic phosphate in rat cerebellum by dopaminergic mechanisms. J. Neurochem. 27. 297298.Google Scholar
Burnstock, G. (1976). Do some nerve cells release more than one transmitter? Neuroscience 1, 239248.Google Scholar
Burr, I.M., Slonim, A.E., Burke, V. and Fletcher, T. (1976). Extracellular calcium and adrenergic and cholinergic effects on islet β -cell function. Amer. J. Physiol. 231, 12461249.Google Scholar
Chasin, M., Mamrak, F. and Samaiego, S.G. (1974). Preparation and properties of a cell-free, hormonally responsive adenylate cyclase from guinea pig brain. J. Neurochem. 22, 10311038.Google Scholar
Chasin, M., Rivkin, I., Mamrak, F., Samaniego, G. and Hess, S.M. (1971). α- AND β -Adrenergic receptors as mediators of accumulation of cyclic adenosine 3’, 5’-monophosphate in specific areas of guinea pig brain. J. Biol. Chem. 246, 30373041.Google Scholar
Cheng, L.C, Rogus, E.M. and Zierler, K. (1977). Catechol, a structural requirement for Na+, K+ -ATPase stimulation in rat skeletal muscle membrane. Biochim. Biophys. Acta. 464, 338346.CrossRefGoogle ScholarPubMed
Chlapowski, F.J., Kelly, L.A. and Butcher, R.W. (1975). Cyclic nucleotides in cultured cells. Advanc. Cyclic Nucleotide Res. 6, 245338.Google Scholar
Chou, W.S., Ho, A.K.S. and Loh, H.H. (1971). Neurohormones on brain adenyl cyclase activity in vivo. Nature New Biol. 233, 280281.Google Scholar
Christ, D.D. and Nishi, S. (1971). Site of adrenaline blockage in the superior cervical ganglion of the rabbit. J. Physiol. 213, 107117.Google Scholar
Ciani, S. and Edwards, C. (1963). The effect of acetylcholine on neuromuscular transmission in the frog. J. Pharmac. Exp. Ther. 142, 2123.Google ScholarPubMed
Clark, R.B. and Gross, R. (1974). Regulation of adenosine 3’, 5’-monophosphate content in human astrocytoma cells by adenosine and the adenine nucleotides. J. Biol. Chem. 249, 52965302.Google Scholar
Clark, R.B. and Perkins, J.P. (1971). Regulation of adenosine 3’, 5’-cyclic monophosphate concentration in cultured human astrocytoma cells by catecholamines and histamine. Proc. Natl. Acad. Sci. 68, 27572760.Google Scholar
Clark, R.B. and Seney, M.N. (1976). Regulation of adenylate cyclase in cultured human cell lines by adenosine. J. Biol. Chem. 251, 42394246.Google Scholar
Clement-Cormier, Y.C, Kebabian, J.W., Petzold, G.L. and Greengard, P. (1974). Dopamine-sensitive adenylate cyclase in mammalian brain: a possible site of action of antipsychotic drugs. Proc. Nat. Acad. Sci. (Wash.) 71, 11131171.Google Scholar
Clement-Cormier, Y.C, Parrish, R.G., Petzold, G.L., Kebabian, J.W. and Greengard, P. (1975). Characterization of a dopamine-sensitive adenylate cyclase in the rat caudate nucleus. J. Neurochem. 25, 143149.Google Scholar
Connor, J.D. (1970). Caudate nucleus neurones: correlation of the effects of substantia nigra stimulation with iontophoretic dopamine. J. Physiol. (Lond.) 208, 691703.Google Scholar
Connor, J.D. (1975). Electrophysiology of the nigro caudate dopamine pathway. Pharmac. Therap. B. 1, 357370.Google Scholar
Conrad, L.C.A., Leonard, C.M. and Pfaff, D.W. (1974). Connections of the median and dorsal raphé nuclei in the rat: An autoradiographic and degeneration study. J. Comp. Neurol. 156, 179206.Google Scholar
Coombs, J.S., Eccles, J.C. and Fatt, P. (1955). The specific ionic conductances and the ionic movements across the motoneuronal membrane that produce the inhibitory post-synaptic potential. J. Physiol. (Lond.) 130, 326373.Google Scholar
Coote, J.H. and Macleod, V.H. (1974). The influence of bulbospinal monoaminer-gic pathways on sympathetic nerve activity. J. Physiol. 241, 453475.CrossRefGoogle ScholarPubMed
Cramer, H., Johnson, D.G., Hanbauer, I., Silberstein, S.D. and Kopin, I.J. (1973). Accumulation of adenosine 3’, 5’-monophosphate induced by catecholamines in the rat superior cervical ganglion in vitro. Brain Res. 53, 97104.Google Scholar
Crawford, J.M., Curtis, D.R., Voorhoeve, P.E. and Wilson, V.J. (1966). Acetylcholine sensitivity of cerebellar neurones in the cat. J. Physiol. (Lond.) 186, 139165.Google Scholar
Crow, T.J. (1973). Catecholamine-containing neurones and electrical self-stimulation: 2. A theoretical interpretation and some psychiatric implications. Psychological Med. 3, 6673.Google Scholar
Cubbedu, L.X., Barnes, E. and Weiner, N. (1975). Release of norepinephrine and dopamine-β-hydroxylase by nerve stimulation. IV. An evaluation of a role for cyclic adenosine monophosphate. J. Pharmac. Exp. Ther. 193, 105127.Google Scholar
DahlstrÔM, A. and Fuxe, K. (1965). Experimentally induced changes in the in-traneuronal amine levels of bulbospinal neurone systems. Acta Physiol. Scand. 64, suppl. 247, 536.Google Scholar
Dalton, C, Crowley, H.J., Sheppard, H. and Schallek, W. (1974). Regional cyclic nucleotide phosphodiesterase activity in cat central nervous system: Effects of benzodiazepines. Proc. Soc. Exp. Biol. Med. 145, 407410.CrossRefGoogle ScholarPubMed
Daly, J. (1975). Role of cyclic nucleotides in the nervous system. In: Handbook of Psychopharmacology, Iverson, L.L. Iverson, S.D. and Snyder, S.H. Eds., New York: Plenum.Google Scholar
Daly, J.W. (1976). Minireview. The nature of receptors regulating the formation of cyclic AMP in brain tissue. Life Sci. 18. 13491358.Google Scholar
Daly, J.W. Huang, M. and Shimizu, H. (1972). Regulation of cyclic AMP levels in brain tissue. Advanc. Cyclic Nucleotide Res. 1, 375387.Google Scholar
Desaiah, D. and Ho, I.K. (1976). Effect of biogenic amines and GABA on ATPase activities in mouse tissues. Eur. J. Pharmac. 40, 255261.Google Scholar
Descarries, L. and Lapierre, Y. (1973). Noradrenergic axon terminals in the cerebral cortex of ral. 1. Radioauiographic visualization after topical application of DL-[3H]norepinephrine. Brain Res. 51. 141160.CrossRefGoogle Scholar
Dismukes, K. and Daly, J.W. (1974). Norepinephrine-sensitive systems generating adenosine 3’, 5’-monophosphate: increased responses in cerebral cortical slices from reserpine-treated rats. Molec. Pharmacol. 10, 933940.Google Scholar
Dismukes, K. and Daly, J.W. (1975). Accumulation of adenosine 3’, 5’-monophosphate in rat brain slices: Effects of prostaglandins. Life. Sci. 17, 199210.Google Scholar
Dismukes, R.K., Ghosh, P., Creveling, C.R. and Daly, J.W. (1975). Altered responsiveness of adenosine 3’, 5’-monophosphate-generating systems in rat cortical slices after lesions of the medial forebrain bundle. Exp. Neurol. 49, 725735.Google Scholar
Dismukes, R.K. and Mulder, A.H. (1976). Cyclic AMP and Ci-receptor-mediated modulation of noradrenaline release from rat brain slices. Eur. J. Pharmac. 39, 383388.Google Scholar
Dismukes, K., Rogers, M. and Daly, J.W. (1976). Cyclic adenosine 3’, 5’-monophosphate formation in guinea-pig brain slices: Effects of Hi and H2-histaminergic agonists. J. Neurochem. 26, 785790.Google Scholar
Donaldson, J., St. Pierre, T., Minnich, J. and Barbeau, A. (1971). Seizures in rats associated with divalent cation inhibition of Na+- K+-ATPase. Can. J. Biochem. 49, 12171224.Google Scholar
Dowd, F.J., Pitts, B.J.R. and Schwartz, A. (1976). Phosphorylation of a low molecular weight polypeptide in beef heart Na+, K+:a TPase preparations. Arch. Biochem. Biophys. 175, 321331.Google Scholar
Dowd, F. and Schwartz, A. (1975). The presence of cyclic AMP-stimulated protein kinase substrates and evidence for endogenous protein kinase activity in various Na+, K+-ATPase preparations from brain, heart and kidney. J. Molec. Cell. Cardiol. 7, 483497.Google Scholar
Drahota, Z., Carafoli, E., Rossi, C.S., Gamble, R.L. and Lehninger, A.L. (1965). The steady state maintenance of accumulated Ca+ + in rat liver mitochondria. J. Biol. Chem. 240, 27122720.Google Scholar
Drummond, G.I., Eng, D.Y. and Mcintosh, C.A. (1971). Ribonucleoside 2’. 3’-cyclic phosphate diesterase activity and cerebroside levels in vertebrate and invertebrate nerve. Brain. Res. 28. 153163.Google Scholar
Drummond, G.I. and May, (1971). Metabolism and functions of cyclic AMP in nerve. Progr. Neurobiol. 2, 121176.Google Scholar
Drummond, G.I. and Powell, C.A. (1970). Analogues of adenosine 3’. 5’-cyclic phosphate as activators of phosphorylase β kinase and as substrates for cyclic 3’, 5’-nucleotide phosphodiesterase. Mol. Pharmac. 6. 2430.Google Scholar
Drummond, G.I. and Yamamoto, M. (1971). Nucleoside cyclic phosphate diesterases. The Enzymes 4. 355371.Google Scholar
Duffy, M.J., Wong, J. and Powell, D. (1975). Stimulation of adenylate cyclase in ral and human brain by substance P. Biochem. Soc. Trans. 2, 12621264.CrossRefGoogle Scholar
Dun, N.J. and Karczmar, A.G. (1977). A comparison of the effect of theophylline and adenosine 3’, 5’-cyclic monophosphate on the superior cervical ganglion of the rabbit by means of the sucrose-gap method. J. Pharmac. Exp. Ther. (in press.).Google Scholar
Dun, N. and Nishi, S. (1974). Effects of dopamine on the superior cervical ganglion of the rabbit. J. Physiol. 239, 155164.Google Scholar
Ebadi, M.S., Weiss, B. and Costa, E. (1971). Distribution of cyclic adenosine monophosphate in rat brain. Arch. Neurol. 24, 353357.Google Scholar
Eccles, J.C. (1963). Postsynaptic and presynaptic inhibitory actions in the spinal cord. Progr. Brain Res. 1, 118.Google Scholar
Edstrom, J.P. and Phillis, J.W. (1976). The effects of AMP on the potential of rat cerebral cortical neurones. Can. J. Physiol. Pharmacol. 54, 787790.Google Scholar
Engberg, I., Flatman, J.A. and Kadzielawa, K. (1974). The hyperpolarization of motoneurones by electrophoretically applied amines and other agents. Acta physiol. scand. 91, 2–4A.Google Scholar
Eränko, O. (1976). Histochemical demonstration of catecholamines in sympathetic ganglia. Ann. Histochim. 21, 83100.Google Scholar
Faiers, A.A. and Mogenson, G.J. (1976). Electrophysiological identification of neurona in locus coeruleus. Exp. Neurol. 53, 254266.CrossRefGoogle Scholar
Ferrendelli, J.A., Chang, M.M. and Kinscherf, D.A. (1974). Elevation of cyclic GMP levels in central nervous system by excitatory and inhibitory amino acids. J. Neurochem. 22, 535540.Google Scholar
Ferrendelli, J.A., Kinscherf, D.A. and Chang, M.M. (1975). Comparison of the effects of biogenic amines on cyclic GMP and cyclic AMP levels in mouse cerebellum in vitro. Brain Res. 84, 6373.Google Scholar
Ferrendelli, J.A.. Kinscherf, D.A. and Kipnis, D.M. (1972). Effects of amphetamine, chlorpromazine and reser-pine on cyclic GMP and cyclic AMP levels in mouse cerebellum. Biochem. Biophys. Res. Commun. 46, 21142120.Google Scholar
Ferrendelli, J.A., Steiner, A.L.. McDougal, D.B. and Kipnis, D.M. (1970). The effect of oxotremorine and atropine on cGMP and cAMP levels in mouse cerebral cortex and cerebellum. Biochem. Biophys. Res. Commun. 41. 10611067.Google Scholar
Florendo, N.T., Barrnett, R.J. and Greengard, P. (1971). Cyclic 3’, 5’-nucleotide phosphodiesterase: cyto-chemical localization in cerebral cortex. Science 173. 745747.Google Scholar
Forn, J. and Krishna, G. (1971). Effect of norepinephrine, histamine and other drugs on cyclic 3’. 5’-AMP formation in brain slices of various animal species. Pharmacol. 5. 193204.Google Scholar
Forn, J., Krueger, B.K. and Greengard, P. (1974). Adenosine 3’. 5’-monophosphate content in ral caudate nucleus. Demonstration of deopaminergic and adrenergic receptors. Science 186. 11181119.Google Scholar
Forn, J. and Valdecasas, F.G. (1971). Effects of lithium on brain adenyl cyclase activity. Biochem. Pharmac. 20. 27732779.Google Scholar
Frederickson, R.C.A., Hewes, C.R. and Norris, F.H. (1975). Microiontophoresis of a selective neuronal uptake inhibitor [3-( ρ -trifluoromethylphenoxy)-N-methyl-3-phenylpropy lamine, Lilly 110140]: Assessing serotonin (5-HT) as a cortical neurotransmitter. Fedn. Proc. 34. 801.Google Scholar
Freedman, R. and Hoffer, B.J. (1975). Phenothiazine antagonism of the noradrenergic inhibition of cerebellar Purkinje neurons. J. Neurobiol. 6. 277288.Google Scholar
Freedman, R., Hoffer, B.J. and Woodward, D.J. (1975). A quantitative microiontophoretic analysis of the responses of central neurons to noradrenaline: interactions with cobalt, manganese, verapamil and dichloroisoprenaline. Br. J. Pharmac. 54, 529539.Google Scholar
Fuxe, K., Hokfelt, T.. Johansson, O.. Jonsson, O., Lidbrink, P. and Ljungdahl, A. (1974). The origin of dopamine nerve terminals in limbic and frontal cortex. Evidence for meso-cortical dopamine neurons. Brain Res. 82. 349355.Google Scholar
Garbarg, M., Barbin, G.. Bischoff, S., Pollard, H. and Schwartz, J.C. (1976). Dual localization of histamine in an ascending neuronal pathway and in non-neuronal cells evidenced by lesions in the lateral hypothalamic area. Brain Res. 106. 333348.Google Scholar
Garvan, F.L. (1964). Metal chelates of ethylenediaminetetracelic acid and related substances. In: Chelating Agents and Metal Chelates. Dwyer, F.P. and Mellor, D.P. Eds.. New York: Academic Press, pp. 283333.Google Scholar
George, W.J.. Polson, J.B.. O’toole, A.G. and Goldberg, N.D. (1970) . Elevation of guanosine 3’, 5’-cyclic phosphate in rat heart after perfusion with acetylcholine. Proc. Natl. Acad. Sci. U.S.A. 66. 398403 Google Scholar
Gilbert, J.C. Wyllie, M.G. and Davison, D.V. (1975). Nerve terminal ATPase as possible trigger for neurotransmitter release. Nature (Lond.) 255. 237238.CrossRefGoogle ScholarPubMed
Gilman, A.G. and Nirenberg, M. (1971) . Effect of calecholamines on the adenosine monophosphate metabolism in cultured neuroblastoma cells. Nature 234. 356358.Google Scholar
Ginsborg, B.L. (1971). On the presynaptic acetylcholine receptors in sympathetic ganglia of the frog. J. Physiol. (Lond.) 216. 237246.Google Scholar
Ginsborg, B.L. and Hirst, G.D.S. (1972). The effect of adenosine on the release of the transmitter from the phrenic nerve of the rat. J. Physiol. (Lond.) 224. 629645.Google Scholar
Godfrainq, J.M.. Kawamura, H.. Krnjevic, K. and Pumain, R. (1971). Actions of dinitrophenol and some other metabolic inhibitors on cortical neurones. J. Physiol. (Lond.) 215, 199222.Google Scholar
Godfraind, J.M. and Pumain, R. (1971). Cyclic adenosine monophosphate and norepinephrine: effect on Purkinje cells in rat cerebellar cortex. Science 174, 12571258.Google Scholar
Godfraind, T., Koch, M.-C. and Verbeke, N. (1974). The action of EGTA on the catecholamines stimulation of rat brain Na-K-ATPase. Biochem. Pharmacol. 23, 35053511.Google Scholar
Goldberg, A.L. and Singer, J.J. (1969). Evidence fora role of cyclic AMP in neuromuscular transmission. Proc. Nat. Acad. Sci. U.S.A. 64, 134141.Google Scholar
Goldberg, N.D., Lust, W.D., O’dea, R.F., Wei, S. and O’toole, A G. (1970). A role of cyclic nucleotides in brain metabolism. Advanc. Biochem. Psychopharmacol. 3, 6787.Google Scholar
Goldberg, N.D., O’dea, R.F. and Haddox, M.K. (1973). Cyclic GMP. Advanc. Cyclic Nucleotide Res. 3, 155223.Google Scholar
Goldberg, N.D. and O’toole, A.G. (1969). The properties of glycogen synthetase and regulation of glycogen biosynthesis in rat brain. J. Biol. Chem. 244, 30533061.Google Scholar
Gonzalez-Vegas, J.A. (1974). Antagonism of dopamine-mediated inhibition in the nigro-striatal pathway: a mode of action of some catatonia-inducing drugs. Brain Res. 80, 219228.Google Scholar
Green, R.D. and Stanberry, L.R. (1977). Elevation of cyclic AMP in C-1300 murine neuroblastoma by adenosine and related compounds and the antagonism of this response by methylxanthines. Biochem. Pharmac. 26, 3743.Google Scholar
Greengard, P. (1976). Possible role for cyclic nucleotides and phosphorylated membrane proteins in postsynaptic actions of neurotransmitters. Nature (Lond.) 260, 101108.Google Scholar
Greengard, P. and Kebabian, J.W. (1974). Role of cyclic AMP in synaptic transmission in the mammalian peripheral nervous system. Fedn. Proc. 33, 10591067.Google Scholar
Greengard, P. and Kuo, J.F. (1970). On the mechanism of action of cyclic AMP. Advan. Biochem. Psychopharmacol. 3, 287306.Google Scholar
Groat, W.C. De, and Theobald, R.J. (1976). Effects of ATP, cyclic AMP and related nucleotides on transmission in parasympathetic ganglia. Pharmacologist 18, 185.Google Scholar
Guidotti, A. and Costa, E. (1973). Involvement of adenosine 3’, 5’-mono-phosphate in the activation of tyrosine hydroxylase elicited by drugs. Science 179, 902904.Google Scholar
Guidotti, A., Hanbauer, I. and Costa, E. (1975). Role of cyclic nucleotides in the induction of tyrosine hydroxylase. Advanc. Cyclic Nucleotide Res. 5, 619639.Google ScholarPubMed
Gumulka, S.W., Dinnendahl, V., Schonhofer, P.S. and Stock, K. (1976). Dopaminergic stimulants and cyclic nucleotides in mouse brain. Effects of dopaminergic antagonists, cholinolytics and GABA agonists. Arch. Pharmac. 295, 2126.Google Scholar
Hadhazy, P. Vizi, E.S., Magyar, K. and Knoll, J. (1976). Inhibition of adrenergic neurotransmission by prostaglandin E1 (PGE1) in the rabbit ear artery. Neuropharmac. 15, 245250.Google Scholar
Hagiwara, S. (1973). Ca spike. Advan. Biophys. 4, 71102.Google Scholar
Hamprecht, B. and Schultz, J. (1973). Stimulation of prostaglandin El of adenosine 3’, 5’-cyclic monophosphate formation in neuroblastoma cells in the presence of phosphodiesterase inhibitors. FEBS Letters 34, 8589.Google Scholar
Hardman, J.G. and Sutherland, E.W. (1969). Guanyl cyclase, an enzyme catalyzing the formation of guanosine 3’, 5’-monophosphate from guanosine triphosphate. J. Biol. Chem. 244, 63636370.Google Scholar
Hays, E.T., Dwyer, T.M., Horowicz, P. and Swift, J.G., (1974). Epinephrine action on sodium fluxes in frog striated muscle. Amer. J. Physiol. 227, 13401347.Google Scholar
Hedqvist, P. (1976). Further evidence that prostaglandins inhibit the release of noradrenaline from adrenergic nerve terminals by restriction of availability of calcium. Br. J. Pharmac. 58, 599603.Google Scholar
Hedqvist, P. and Fredholm, B.B. (1976). Effects of adenosine on adrenergic neurotransmission; prejunctional inhibition and postjunctional enhancement. Arch. Pharmacol. 293, 217223.Google Scholar
Hegstrand, L.R., Kanof, P.D. and Greengard, P. (1976). Histamine-sensitive adenylate cyclase in mammalian brain. Nature 260, 163165.Google Scholar
Hexum, T.D. (1974). Studies on the reaction catalyzed by transport (Na, K) adenosine triphosphatase. I. Effects of divalent metals. Biochem. Pharmac. 23, 34413447.Google Scholar
Hoffer, B.J., Freedman, R., Woodward, D.J., Daly, J.W. and Skolnick, P. (1976). /3β-Adrenergic control of cyclic AMP generating systems in cerebellum: pharmacological heterogeneity confirmed by destruction of interneurons. Exp. Neurol. 51, 653667.Google Scholar
Hoffer, B.J., Siggins, G.R., Oliver, A.P. and Bloom, F.E. (1971). Cyclic AMP mediation of norepinephrine inhibition in rat cerebellar cortex: A unique class of synaptic responses. Ann. N.Y. Acad. Sci. 185, 531549.Google Scholar
Hoffer, B.J., Siggins, G.R., Oliver, A.P. and Bloom, F.E. (1973). Activation of the pathway from locus coeruleus to rat cerebellar Purkinje neurons: pharmacological evidence of noradrenergic central inhibition. J. Pharmacol, exp. Ther. 184, 553569.Google Scholar
Hokfelt, T. and Ungerstedt, U. (1973). Specificity of 6-hydroxydopamine-induced degeneration of central monoamine neurons: electron and fluorescence microscopic study with special reference to intracerebral injection of the nigrostriatal dopamine system. Brain Res. 60, 296298.Google Scholar
Horn, A.S., Cuello, A.C. and Miller, R.J. (1974). Dopamine in the mesolimbic system of rat brain: endogenous levels and the effect of drugs on the uptake mechanisms and stimulation of adenylate cyclase activity. J. Neurochem. 22, 265270.Google Scholar
Hornykiewicz, O. (1966). Dopamine (3-hydroxy-tyramine) and brain function. Pharmac. Rev. 18, 925964.Google Scholar
Hosie, R.J.A. (1965). The localization of ATPase in morphologically characterized subcellular fractions of guinea pig brain. Biochem. J. 96, 404412.Google Scholar
Hotta, Y. and Tsukiu, R. (1968). Effect on the guinea-pig taenia coli of the substitution of strontium or barium ions for calcium ions. Nature 217, 867869.Google Scholar
Huang, M. and Daly, J.W. (1974). Adenosine-elicited accumulation of cyclic AMP in brain slices: Potentiation by agents which inhibit uptake of adenosine. Life Sci. 14, 489503.Google Scholar
Huang, M. and Drummond, G.I. (1976). Effect of adenosine on cyclic AMP accumulation in ventricular myocardium. Biochem. Pharmac. 25, 27132719.Google Scholar
Huang, M., Ho, A.K.S. and Daly, J.W. (1973). Accumulation of adenosine cyclic 3’5’-monophosphate in rat cerebral cortical slices. Stimulatory effects of alpha and beta adrenergic agents after treatment with 6-hydroxydopamine, 2,3,5-trihydroxy-phenethylamine and dihydroxytryptamines. Molec. Pharmac. 9, 711717.Google Scholar
Huang, M., Shimizu, H. and Daly, J.W. (1971). Regulation of adenosine cyclic 3’,5’-phosphate formation in cerebral cortical slices: Interaction among norepinephrine, histamine, serotonin. Molec. Pharmac. 7, 155162.Google ScholarPubMed
Huang, M., Shimizu, H. and Daly, J.W. (1972). Accumulation of cyclic adenosine monophosphate in incubated slices of brain tissue. 2. Effects of depolarizing agents, membrane stabilizers, phospho-diesterase inhibitors, and adenosine analogs. J. Med. Chem. 15, 462466.Google Scholar
Hubbard, J.I., Schmidt, R.F. and Yokota, T. (1965). The effect of acetylcholine upon mammalian motor nerve terminals. J. Physiol. (Lond.) 181. 810829.Google Scholar
Ichida, S., Kuo, C.H., Matsuda, T. and Yoshida, H. (1976). Effects of La+ + +, Mn+ +and ruthenium red on Mg-Ca-ATPhase activity and ATP-dependent Ca-binding of the synaptic plasma membrane. Jap. J. Pharmac. 26. 3943.Google Scholar
Inoue, Y., Yamamura, H. and Nishizuka, Y. (1973).Protamine kinase independent of adenosine 3’,5’-monophosphate from rat brain cytosol. Biochem. Biophys. Res. Commun. 50, 228236.Google Scholar
Isaacson, A. and Sandow, A. (1976). Quinine and caffeine effects on calcium movements in frog sartorius muscle. J. Gen. Physiol. 50, 21092128.Google Scholar
Israel, Y. (1970). Cellular effects of alcohol. A review. Quart. J. Stud. Ale. 31, 293316.Google Scholar
Iverson, L.L. (1975). Dopamine receptors in the brain. Science 188, 10841089.Google Scholar
Iwangoff, P., Enz, A. and Chappuis, A. (1974). Effect of adrenergic blockers on the activation of brain ATPase by noradrenaline. Experientia 30, 688.Google Scholar
Jaanus, S.D. and Rubin, R.P. (1974). Analysis of the role of cyclic adenosine 30-monophosphate in catecholamine release. J. Physiol. (Lond.) 237, 465476.Google Scholar
Jack, J.J.B., Miller, S., Porter, R. and Redman, S.J. (1971). The time course of minimal excitatory post-synaptic potentials evoked in spinal motoneurones by group la afferent fibres. J. Physiol. 215, 353380.Google Scholar
Jacobowitz, D. (1970). Catecholamine fluorescence studies of adrenergic neurons and chromaffin cells in sympathetic ganglia. Fedn. Proc. 29, 19291944.Google Scholar
Jenkinson, D.H., Stamenovic, B.A. and Whitaker, B.D.L. (1968). The effect of noradrenaline on the end-plate potential in twitch fibres of the frog. J. Physiol. 195, 743754.Google Scholar
Johnson, E.M., Maeno, H. and Greengard, P. (1971). Phosphorylation of endogenous protein of rat brain by a cyclic adenosine 3’, 5’-monophosphate-dependent protein kinase. J. Biol. Chem. 246, 77317739.Google Scholar
Johnson, G.A., Boukma, S.J., Lahti, R.A. and Mathews, J. (1973). Cyclic AMP and phosphodiesterase in synaptic vesicles from mouse brain. J. Neurochem. 20, 13871392.Google Scholar
Johnson, P.N. and Inesi, A. (1969). The effect of methylxanthines and local anaesthetics on fragmented sarcoplasmic reticulum. J. Pharmacol. Exp. Therp. 169, 308314.Google Scholar
Kakiuchi, S. and Rall, T.W. (1968a). Studies on adenosine 3’, 5’-phosphate in rabbit cerebral cortex. Mol. Pharmacol. 4, 379388.Google Scholar
Kakiuchi, S. and Rall, T.W. (1968b). The influence of chemical agents on the accumulation of adenosine 3’, 5’-phosphate in slices of rabbit cerebellum. Mol. Pharmacol. 4, 367378.Google Scholar
Kakiuchi, S., Yamazaki, R., Teshima, Y., Uenishi, K. and Miyamoto, E. (1975). Ca2+/Mg2+ -dependent cyclic nucleotide phosphodiesterase and its activator protein. Advanc. Cyclic Nucleotide Res. 5, 163178.Google Scholar
Kalix, P., McAfee, D.A., Schorderet, M. and Greengard, P. (1974). Pharmacological analysis of synaptically mediated increase in cyclic adenosine monophosphate in rabbit superior cervical ganglion. J. Pharmac. Exp. Therp. 188, 676687.Google Scholar
Kass, R.S. and Tsien, R.W. (1975). Multiple effects of calcium antagonists on plateau currents in cardiac Purkinje fibers. J. Gen. Physiol. 66, 169192.Google Scholar
Kataoka, K. and De Robertis, E. (1967). Histamine in isolated small nerve endings and synaptic vesicles of rat brain cortex. J. Pharmac. Exp. Therp. 156, 114125.Google Scholar
Kebabian, J.W., Bloom, F.E., Steiner, A.L. and Greengard, P. (1975). Neurotransmitters increase cyclic nucleotides in postganglionic neurons: im-munocytochemical demonstration. Science 190, 157159.Google Scholar
Kebabian, J.W. and Greengard, P. (1971). Dopamine-sensitive adenyl cyclase: possible role in synaptic transmission. Science 174, 13461349.Google Scholar
Kebabian, J.W., Steiner, A.L. and Greengard, P. (1975). Muscarinic cholinergic regulation of cyclic guanosine 3’, 5’-monophosphate in autonomic ganglia: Possible role in synaptic transmission. J. Pharmac. Exp. Ther. 1973, 474488.Google Scholar
Kelly, J.S. (1975). Microiontophoretic application of drugs onto single neurons. In: Handbook of Psychopharmacology, Vol. 2, Iverson, L., Iverson, S.D. and Snyder, S.H. Eds., pp. 2967, New York: Plenum Press.Google Scholar
Kety, S. (1972). Brain catecholamines: affective states and memory. Advan. Behavioral Biol. 4, 6580.Google Scholar
Kim, T.S., Shulman, J. and Levine, R.A. (1968). Relaxant effect of cyclic adenosine 3’, 5’-monophosphate on the isolated rabbit ileum. J. Pharmac. Exp. Ther. 163, 3642.Google Scholar
Kinscherf, D.A., Chang, M.M., Rubin, E.H., Schneider, D.R. and Ferrendelli, J.A. (1976). Comparison of the effects of depolarizing agents and neurotransmitters on regional CNS cyclic GMP levels in various animals. J. Neurochem. 26, 527530.Google Scholar
Kitai, S.T., Sugimori, M. and Kocsis, J.D. (1976). Excitatory nature of dopamine in the nigro-caudate pathway. Exp. Brain Res. 24, 351363.Google Scholar
Klainer, L.M., Chi, Y.-M., Freidberg, S.L., Rall, T.W. and Sutherland, E.W. (1962). Adenyl cyclase: IV. The effects of neurohormones on the formation of adenosine 3’, 5’-phosphate by preparations from brain and other tissues. J. Biol. Chem. 237, 12391243.Google Scholar
Kobayashi, H. and Libet, B. (1968). Generation of slow post-synaptic potentials without increases in ionic conductance. Proc. Nat. Acad. Sci. U.S.A. 60, 13041311.Google Scholar
Kobayashi, H. and Libet, B. (1970). Actions of noradrenaline and acetylcholine on sympathetic ganglion cells. J. Physiol. (Lond.) 208, 353372.Google Scholar
Kobayashi, H. and Libet, B. (1974). Is inactivation of potassium conductance involved in slow postsynaptic excitation of sympathetic ganglion cells? Effects of nicotine. Life Sci. 14, 18711883.Google Scholar
Kober, T.E. and Cooper, G.P. (1976). Lead competitively inhibits calcium-dependent synaptic transmission in the bullfrog sympathetic ganglion. Nature (Lond.) 262, 704705.Google Scholar
Koketsu, K. and Nakamura, M. (1976). The electrogenesis of adrenaline-hyperpolarization of sympathetic ganglion cells in bullfrogs. Jap. J. Physiol. 26, 6377.Google Scholar
Koketsu, K. and Ohta, Y. (1976). Acceleration of the electrogenic Na+ pump by adrenaline in frog skeletal muscle fibres. Life Sci. 19, 10091014.Google Scholar
Koketsu, K. and Shiras’awa, Y. (1974). 5-HT and the electrogenic sodium pump. Experientia 30, 10341035.Google Scholar
Koketsu, K., Shoji, T. and Nishi, S. (1973). Slow inhibitory postsynaptic potentials of bullfrog sympathetic ganglia in sodium-free media. Life Sci. 13, 453458.Google Scholar
Kostopoulos, G.K., Limacher, J.J. and Phillis, J.W. (1975). Action of various adenine derivatives on cerebellar Purkinje cells. Brain Res. 88, 162165.Google Scholar
Krnjevic, K. and Lisiewicz, A. (1972). Injections of calcium ions into spinal motoneurons. J. Physiol. (Lond.) 225, 363390. yGoogle Scholar
Krnjevic, K. and Miledi, R. (1958). Some effects produced by adrenaline upon neuromuscular propagation in rats. J. Physiol. (Lond.) 141, 291304.Google Scholar
Krnjevic, K. and Phillis, J.W. (1963). Actions of certain amines on cerebral cortical neurones. Br. J. Pharmac. Chemother. 20, 471490.Google Scholar
Krnjevic, K., Puil, E. and Werman, R. (1975). Evidence for Ca2+-activated K+ conductance in cat spinal motoneurons from intracellular EGTA injections. Can. J. Physiol. Pharmacol. 53, 12141218.Google Scholar
Krnjevic, K., Puil, E. and Werman, R. (1976). Is cyclic guanosine monophosphate the internal ‘second messenger’ for cholinergic actions on central neurons? Can. J. Physiol. Pharmacol. 54, 172176.Google Scholar
Krnjevic, K., Pumain, R. and Renaud, L. (1971a). The mechanism of excitation by acetylcholine in the cerebral cortex. J. Physiol. (Lond.), 215, 247268.Google Scholar
Krnjevic, K., Pumain, R. and Renaud, L. (1971b). Effects of Ba2+ and tetraethylammonium on cortical neurones. J. Physiol. (Lond.) 215, 223245.Google Scholar
Krnjevic, K. and Van Meter, W.G. (1976). Cyclic nucleotides in spinal cells. Can. J. Physiol. Pharmacol. 54, 416421.Google Scholar
Kuba, K. and Nishi, S. (1976). Rhythmic hyperpolarizations and depolarization of sympathetic ganglion cells induced by caffeine. J. Neurophysiol. 39, 547563.Google Scholar
Kuehl, F.A., Humes, J.L., Cirillo, V.J. and Ham, E.A. (1972). Cyclic AMP and prostaglandins in hormone actions. Advanc. Cyclic Nucleotide Res. 1, 493502.Google Scholar
Kuhar, M.J., Aghajanian, G.K. and Roth, R.H. (1972). Tryptophan hydroxylase activity and synaptosomal uptake of serotonin in discrete brain regions after midbrain raphe lesions: correlations with serotonin levels and histochemical fluorescence. Brain Res. 44, 165176.Google Scholar
Kuo, J.F. (1974). Guanosine 3’, 5’-mono-phosphate-dependent protein kinases in mammalian tissues. Proc. Nat. Acad. Sci. (U.S.A.) 71, 40374041.Google Scholar
Kuo, J.F. and Greengard, P. (1969a). Cyclic nucleotide-dependent protein kinases. IV. Widespread occurrence of adenosine 3’, 5’-monophosphate dependent protein kinase in various tissues and phyla of the animal kingdom. Proc. Nat. Acad. Sci. (U.S.A.) 64, 13491355.Google Scholar
Kuo, J.F. and Greengard, P. (1969b). Adenosine 3’, 5’-monophosphate dependent protein kinase from brain. Science 165, 6365.Google Scholar
Kuo, J.F. and Greengard, P. (1970a). Stimulation of adenosine 3’, 5’-mono-phosphate-dependent protein kinases by some analogs of adenosine 3’, 5’-mono-phosphate. Biochem. Biophys. Res. Commun. 40, 10321038.Google Scholar
Kuo, J.F. and Greengard, P. (1970b). Cyclic nucleotide-dependent protein kinases. VII. Comparison of various histories as substrates for adenosine 3’, 5’-monophosphate-dependent and guanosine 3’, 5’-monophosphate-dependent protein kinases. Biochim. Biophys. Acta. 212, 434440.Google Scholar
Kuo, J.F. and Greengard, P. (1973). Stimulation of cyclic GMP dependent protein kinase by a protein fraction which inhibits cyclic-AMP-dependent protein kinases. Fedn. Proc. 30, 1089 Abs.Google Scholar
Kuo, J.F., Lee, T.-P., Reyes, P.L., Wakton, K.G., Donnelly, T.E. and Greengard, P. (1972). Cyclic nucleotide-dependent protein kinases. X. An assay method for the measurement of guanosine 3’, 5’-monophosphate in various biological materials and a study of agents regulating its levels in heart and brain. J. Biol. Chem. 247, 1622.Google Scholar
Kuroda, Y., Saito, M. and Kobayashi, K. (1976a). Concomitant changes in cyclic AMP level and postsynaptic potentials of olfactory cortex slices induced by adenosine derivatives. Brain Res. 109, 196201.Google Scholar
Kuroda, Y., Saito, M. and Kobayashi, K. (1976b). High concent-rations of calcium prevent the inhibition of postsynaptic potentials and the accumulation of cyclic AMP induced by adenosine in brain slices. Proc. Jap. Acad. 52, 8689.Google Scholar
Kurokawa, M., Sakamoto, T. and Kato, M. (1965). Distribution of Na-plus-K-stimulated ATPase activity in isolated nerve ending particles. Biochem. J. 97, 833844.Google Scholar
Kuypers, H.G.J.M. and Maisky, V.A. (1975). Retrograde axonal transport of horseradish peroxidase from spinal cord to brainstem cell groups in the cat. Neurosci. Lett. 1, 914.Google Scholar
Lake, N. and Jordan, L.M. (1974). Failure to confirm cyclic AMP as second messenger for norepinephrine in rat cerebellum. Sci. 183, 663664.Google Scholar
Lake, N., Jordan, L.M. and Phillis, J.W. (1973). Evidence against cyclic adenosine 3’, 5’-monophosphate (AMP) mediation of noradrenaline depression of cerebral cortical neurones. Brain. Res. 60, 411421.Google Scholar
Lee, T.-P., Kuo, J.F. and Greengard, P. (1972). Role of muscarinic cholinergic receptors in regulation of guanosine 3’, 5′-cyclic monophosphate content in mammalian brain, heart muscle and intestinal smooth muscle. Pros. Nat. Acad. Sci. U.S.A. 69, 32873291.Google Scholar
Libet, B. (1970). Generation of slow inhibitory and excitatory postsynaptic potentials. Fedn. Proc. 29, 19451949.Google Scholar
Libet, B. and Kobayashi, H. (1974). Adrenergic mediation of slow inhibitory postsynaptic potential in sympathetic ganglia of the frog. J. Neurophysiol. 37, 805814.Google Scholar
Libet, B., Kobayashi, H. and Tanaka, T. (1975). Synaptic coupling into the production and storage of a neuronal memory trace. Nature (Lond.) 258, 155157.Google Scholar
Libet, B. and Owman, C.H. (1974). Concomitant changes in formaldehyde-induced fluorescence of dopamine interneurones and in slow inhibitory post-synaptic potentials of the rabbit superior cervical ganglion, induced by stimulation of the preganglionic nerve or by a muscarinic agent. J. Physiol. (Lond.) 237, 635662.Google Scholar
Libet, B. and Tosaka, T. (1970). Dopamine as a synaptic transmitter and modulator in sympathetic ganglia: A different mode of synaptic action. Proc. Nat. Acad. Sci. U.S.A. 67, 667673.Google Scholar
Lindl, T. and Cramer, H. (1975). Evidence against dopamine as the mediator of the rise of cyclic AMP in the superior cervical ganglion. Biochem. Biophys. Res. Comm. 65, 731739.Google Scholar
Lindvall, O., Bjôrklund, A., Moore, R.Y. and Stenevi, U. (1974). Mesencephalic dopamine neurons projecting to neocortex. Brain Res. 81, 325331.Google Scholar
Llinas, R. and Baker, R. (1972). A chloride-dependent inhibitory postsynaptic potential in cat trochlear motoneurons. J. Neurophysiol. 35, 484492.Google Scholar
Logan, J.G. and O’donovan, D.J. (1976). The effects of ouabain and the activation of neural membrane ATPase by biogenic amines. J. Neurochem. 27, 185189.Google Scholar
Lorens, S.A. and Goldberg, H.C. (1974). Regional 5-hydroxytryptamine following midbrain raphé lesions in the rat. Brain Res. 78, 4565.Google Scholar
Lust, W.D., Passonneau, J.V. and Veech, R.L. (1973). Cyclic adenosine monophosphate metabolites and phos-phorylase in neuronal tissue. A comparison of methods of fixation. Science 181, 280282.Google Scholar
Lux, H.D. (1971). Ammonium and chloride extrusion: Hyperpolarizing synaptic inhibition in spinal motoneurons. Science 173, 555557.Google Scholar
Lux, H.D., Loracher, C. and Neher, E. (1970). The action of ammonium on postsynaptic inhibition of cat spinal motoneurons. Exp. Brain Res. 11, 431447.Google Scholar
Madiera, V.M. and Antunes-Madiera, M.C. (1973). Interaction of Ca+ + and Mg+ + with synaptic plasma membranes. Biochim. Biophys. Acta. 323, 396407.Google Scholar
Maeno, H. and Greengard, P. (1972). Phosphoprotein phosphatases from rat cerebral cortex. J. Biol. Chem. 247, 32693277.Google Scholar
Maeno, H., Johnson, E.M. and Greengard, P. (1971). Subcellular distribution of adenosine 3’, 5’-monophosphate-de pendent protein kinase in rat brain. J. Biol. Chem. 246, 134142.Google Scholar
Maeno, H., Ueda, T. and Greengard, P. (1975). Adenosine 3’, 5’-monophosphate dependent protein phosphatase activity in synaptic membrane fractions. J. Cyclic Nucleotide Res. 1, 3748.Google Scholar
Magaribuchi, T.and Kuriyama, H. (1972). Effects of noradrenaline and iso-prenaline on the electrical and mechanical activities of guinea pig depolarized taenia coli. Jap. J. Physiol. 22, 253270.Google Scholar
Mah, H.D. and Daly, J.W. (1976). Adenosine-dependent formation of cyclic AMP in brain slices. Pharmac. Res. Commun. 8, 6579.Google Scholar
Manalis, R.S. and Cooper, G.P. (1973) . Presynaptic and postsynaptic effects of lead at the frog neuromuscular junction. Nature (Lond.) 243, 354356.Google Scholar
Mao, C.C, Guidotti, A. and Costa, E. (1974a). Inhibition by diazepam of the tremor and the increase of cerebellar cGMP content elicited by harmaline. Brain. Res. 83, 526529.Google Scholar
Mao, C.C, Guidotti, A. and Costa, E. (1974b). Interactions between Y-aminobutyric acid and guanosine 3’, 5’-monophosphate in rat cerebellum. Molec. Pharmac. 10, 736745.Google Scholar
Matthews, M.R. and Raisman, G. (1969). The ultrastructure and somatic efferent synapses of small granule-containing cells in the superior cervical ganglion. J. Anat. 105, 255282.Google Scholar
McAfee, D.A. and Greengard, P. (1972). Adenosine 3’, 5’-monosphosphate: Electrophysiological evidence for a role in synaptic transmission. Science 178, 310312.Google Scholar
McAfee, D.A., Schorderet, M. and Greengard, P. (1971). Adenosine 3’, 5’-monophosphate in nervous tissue: increase associated with synaptic transmission. Science 171, 11561158.Google Scholar
McLennan, H. (1970). Synaptic transmission. Philadelphia: Saunders.Google Scholar
Melamed, E., Lahav, M. and Atlas, D. (1976). Direct localization of a β -adrenoceptor site in rat cerebellum by a new fluorescent analogue of propranolol. Nature (Lond.) 261. 420422.Google Scholar
Miledi, R. (1973). Transmitter release induced by injection of calcium ions into nerve terminals. Proc. Roy. Soc. Ser. B. 183, 421425.Google Scholar
Miller, J.P., Boswell, K.H., Muneyana, K., Simon, L.N., Robins, R.K. and Shuman, D.A. (1973). Synthesis and biochemical studies of various 8-substituted derivatives of guanosine 3’. 5’-cyclic phosphate and xanthosine 3’, 5’-cyclic phosphate. Biochemistry 12, 53105319.Google Scholar
Miller, R.J., Horn, A.S. and Iversen, L.L. (1974). The action of neuroleptic drugs on dopamine-stimulated adenosine 3’, 5’-monophosphate production in rat neostriatum and limbic forebrain. Molec. Pharmac. 10, 759766.Google Scholar
Miller, R.J. and Iversen, L.L. (1974). Effect of psychoactive drugs on dopamine (3. 4-dihydroxyphenethylamine) sensitive adenylate cyclase activity in corpus striatum of rat brain. Biochem. Soc. Transact. 2, 256259.Google Scholar
Minna, J.D. and Gilman, A.G. (1973). Expression of genes for metabolism of cyclic adenosine 3’. 5’-monophosphate in somatic cells. II. Effects of prostaglandin EI and theophylline on parental and hybrid cells. J. Biol. Chem. 248, 66186625.Google Scholar
Miyamoto, E.. Kuo, J.F. and Greengard, P. (1969a). Adenosine 3’, 5’-mono-phosphate-dependent protein kinase from brain. Science 165. 6365.Google Scholar
Miyamoto, E. Kuo, J.G and Greengard, P. (1969b). Cyclic nucleotide-dependenl protein kinases. 1. Purification and properties of adenosine 3’. 5’-mono-phosphate-dependent protein kinase from bovine brain. J. Biol. Chem. 244, 63956402.Google Scholar
Miyamoto, E., Petzold, G.L., Harris, J.S. and Greengard, P. (1971). Dissociation and concomitant activation of adenosine 3’, 5’-monophosphate-dependent protein kinase by histone. Biochem. Biophys. Res. Commun. 44, 305312.Google Scholar
Miyamoto, M.D. and Breckenridge, B. McL. (1974). A cyclic adenosine monophosphate link in the catecholamine enhancement of transmitter release at the neuromuscular junction. J. Gen. Physiol. 63. 609624.Google Scholar
Moritoki, H., Morita, M. and Kanbe, T. (1976). Effects of methylxan-thines and imidazole on the contractions of guinea-pig ileum induced by transmural stimulation. Eur. J. Pharmac. 35, 185198.Google Scholar
Nahorski, S.R. and Rogers, K.J. (1976). Inhibition of 3’, 5’-nucleotide phosphodiesterase and the stimulation of cerebral cyclic AMP formation by biogenic amines in vitro and in vivo. Neuropharmac. 15, 609612.Google Scholar
Nahorski, S.R., Rogers, K.J. and Smith, B.M. (1974). Histamine H2 receptors and cyclic AMP in brain. Life Sci. 15, 18871894.Google Scholar
Nakamura, S. and Iwama, K. (1975). Antidromic activation of the rat locus coeruleus neurons from hippocampus, cerebral and cerebellar cortices. Brain Res. 99, 372376.Google Scholar
Nakazawa, K. and Sano, M. (1974). Studies on guanylate cyclase. A new assay method for guanylate cyclase and properties of the cyclase from rat brain. J. Biol. Chem. 249, 42074211.Google Scholar
Nathanson, J.A. and Bloom, F.E. (1976). Heavy metals and adenosine cyclic 3’, 5’-monophosphate metabolism: possible relevance to heavy metal toxicity. Molec. Pharmac. 12, 390398.Google Scholar
Nathanson, J.A., Freedman, R. and Hoffer, B.J. (1976). Lanthanum inhibits brain adenylate cyclase and blocks noradrenergic depression of Purkinje cell discharge independent of calcium. Nature 261, 330332.Google Scholar
Nayler, W.G. and Harris, J.P. (1976). Inhibition by lanthanum of the Na+K+-activated ouabain-sensitive adenosine triphosphatase enzyme. J. Molec. Cell. Cardiol. 8, 811822.Google Scholar
Nicoll, R.A. (1976). Promising peptides. Society for Neuroscience Symposia 1, 99122.Google Scholar
Nikodijevic, O., Nikodijevic, B., Zinder, O., Yu, M.-Y.W., Guroff, G. and Pollard, H.B. (1976). Control of adenylate cyclase from secretory vesicle membranes by β -adrenergic agents and nerve growth factor. Proc. Nat. Acad. Sci. U.S.A. 73, 771, 774.Google Scholar
Nishi, S. (1970). Cholinergic and adrenergic receptors at sympathetic preganglionic nerve terminals. Fedn. Proc. 29, 19571965.Google Scholar
Nishi, S. and Koketsu, K. (1968). Analysis of slow inhibitory post-synaptic potential of bullfrog sympathetic ganglion. J. Neurophysiol. 31, 717728.Google Scholar
Nishi, S., Soeda, H. and Koketsu, K. (1969). Unusual nature of ganglionic slow EPSP studied by a voltage clamp method. Life Sci. 8, 3342.Google Scholar
Noon, J.P., McAfee, D.A. and Roth, R.H. (1975). Norepinephrine release from nerve terminals within the rabbit superior cervical ganglion. Arch. Pharmac. 291, 139162.Google Scholar
Obata, K. and Yoshida, M. (1973). Caudate-evoked inhibition and actions of GABA and other substances on cat pallidal neurons. Brain. Res. 64, 455459.Google Scholar
Oliver, A.P. and Segal, M. (1974). Transmembrane changes in hippocampal neurons: hyperpolarizing actions of norepinephrine, cyclic AMP and locus coeruleus. Proc. Soc. Neurosci. 361.Google Scholar
Olson, L. and Fuxe, K. (1971). On the projections from the locus coeruleus noradrenaline neurons. Brain Res. 28, 165171.Google Scholar
Opler, L.A. and Makman, M.H. (1972). Mediation by cyclic AMP of hormone-stimulated glycogenolysis in cultured rat astrocytoma cells. Biochem. Biophys. Res. Commun. 46, 11401145.Google Scholar
Opmeer, F.A., Gumulka, S.W.. Dinnendahl, V. and Schönhöfer, P.S., (1976). Effects of stimulatory and depressant drugs on cyclic guanosine 3’. 5’-monophosphate and adenosine 3'. 5'-monophosphate levels in mouse brain. Arch. Pharmacol. 292. 259265.Google Scholar
Otten, U., Mueller, R.A., Oesch, F. and Thoenen, H. (1974). Location of an isoproterenol-responsive cyclic AMP-pool in adrenergic nerve cell bodies and its relationship to tyrosine hydroxylase induction. Proc. Nat. Acad. Sci. U.S.A. 71. 22172221.Google Scholar
Palmer, G.C. (1973). Adenyl cyclase in neuronal and glial-enriched fractions from rat and rabbit brain. Res. Commun. Chem. Pathol. Pharmac. 5, 603613.Google Scholar
Palmer, G.C. and Duszynski, C.R. (1975). Regional cyclic GMP content in incubated tissue slices of rat brain. Eur. J. Pharmac. 32, 375379.Google Scholar
Palmer, G.C, Sulser, F. and Robi-SON, G.A. (1973). Effects of neurohumoral and adrenergic agents on cyclic AMP levels in various areas of the rat brain in vitro. Neuropharmacol. 12. 327337.Google Scholar
Parfitt, A., Weller, J.L.. Klein, D.C., Sakai, K.K. and Marks, B.H. (1975). Blockage by ouabain or elevated potassium ion concentration of the adrenergic and adenosine cyclic 3’. 5’-monophosphate-induced stimulation of pineal serotonin N-acetyltransferase activity. Molec. Pharmac. 11. 241255.Google Scholar
Peach, M.J. (1972). Stimulation of release of adrenal catecholamine by adenosine 3’. 5’-monophosphate and theophylline in ab-sense of extracellular Ca2+ + . Proc. Nat. Acad. Sci. U.S.A. 69, 834836.Google Scholar
Penit, J., Huot, J. and Jard, S. (1976). Neuroblastoma cell adenylate cyclase: direct activation by adenosine and prostaglandins. J. Neurochem. 26, 265273.Google Scholar
Perkins, J.P. (1973). Adenyl cyclase. Advan. Cyclic Nucleotide Res. 3. 164.Google Scholar
Perkins, J.P., Macintyre, E.H., Riley, W.D. and Clark, R.B. (1971). Adenyl cyclase, phosphodiesterase and cyclic AMP dependent protein kinase of malignant glial cells in culture. Life Sci. 10. 10691080.Google Scholar
Perkins, J.P. and Moore, M.M. (1971). Adenyl cyclase of rat cerebral cortex. J. Biol. Chem. 246, 6268.Google Scholar
Perkins, J.P. and Moore, M.M. (1973). Characterization of the adrenergic receptors mediating a rise in cyclic 3’. 5’-adenosine monophosphates in rat cerebral cortex. J. Pharmac. exp. Ther. 185. 371378.Google Scholar
Phillis, J.W. (1970). The Pharmacology of Synapses. Oxford: Pergamon Press.Google Scholar
Phillis, J.W. (1974a). The role of calcium in the central effects of biogenic amines. Life Sci. 14, 11891201.Google Scholar
Phillis, J.W. (1974b). Neomycin and ruthenium red antagonism of monoaminer-gic depression of cerebral cortical neurones. Life Sci. 15, 213222.Google Scholar
Phillis, J.W. (1974c). Evidence for cholinergic transmission in the cerebral cortex. Advanc. Behavioral Biol. 10, 5777.Google Scholar
Phillis, J.W. (1976). An involvement of calcium and Na, K-ATPase in the inhibitory actions of various compounds on central neurons. In: Taurine, R. Huxtable, and Barbeau, A. Eds., pp. 209223, New York: Raven Press.Google Scholar
Phillis, J.W. (1977). Substance p and related peptides. Soc. Neuroscience Symposia II. (in press).Google Scholar
Phillis, J.W. and Edstrom, J.P. (1976). Effects of adenosine analogs on rat cerebral cortical neurons. Life Sci. 19, 10411054.Google Scholar
Phillis, J.W. and Kostopoulos, G.K. (1975). Adenosine as a putative transmitter in the cerebral cortex. Studies with potentiators and antagonists. Life Sci. 17, 10851094.Google Scholar
Phillis, J.W. and Kostopoulos, G.K. (1977). Activation of a noradrenergic pathway from the brain stem to rat cerebral cortex. Gen. Pharmac. (in press).Google Scholar
Phillis, J.W., Kostopoulos, G.K. and Limacher, J.J. (1974). Depression of corticospinal cells by various purines and pyrimidines. Can. J. Physiol. Pharmacol. 52. 12261229.Google Scholar
Phillis, J.W., Kostopoulos, G.K. and Limacher, J.J. (1975). A potent depressant action of adenine derivatives on cerebral cortical neurones. Eur. J. Pharmac. 30, 125129.Google Scholar
Phillis, J.W., Limacher, J.J. (1974). Effects of some metallic cations on cerebral cortical neurones and their interactions with biogenic amines. Can. J. Physiol. Pharmacol. 52, 566574.Google Scholar
Phillis, J.W., Tebecis, A.K. and York, D.H. (1968a). Depression of spinal motoneurones by noradrenaline, 5-hydroxytryptamine and histamine. Eur. J. Pharmac. 7, 471475.Google Scholar
Phillis, J.W., Tebecis, A.K. and York, D.H. (1968b). Histamine and some anti-histamines: their actions on cerebral cortical neurones. Brit. J. Pharmac. 33, 426440.Google Scholar
Pickel, V.M., Segal, M. and Bloom, F.E. (1974). A radioautographic study of the efferent pathways of the nucleus locus coeruleus. J. Comp. Neurol. 155, 1542.Google Scholar
Pin, D., Jones, B. and Jouvet, M. (1968). Topographie des neurones monoaminergiques du tronc cerebral du chat: etude par histofluorescence. C. R. Soc. Biol. (Paris) 162, 21362141.Google Scholar
Prakash, N.J., Fontana, J. and Henkin, R.I. (1973). Effect of transitional metal ions on (Na+ + K+) ATPase activity and the uptake of norepinephrine and choline by brain synaptosomes. Life. Sci. 12, 249259.Google Scholar
Prasad, K.N., Gilmer, K.N. and Sahu, S.K. (1974). Demonstration of acetylcholine-sensitive adenyl cyclase in malignant neuroblastoma cells in culture. Nature (Lond.) 249, 765767.Google Scholar
Rahwan, K.G. and Borowitz, J.L. (1973). Mechanisms of stimulus-secretion coupling in adrenal medulla. J. Pharm. Sciences 62, 19111923.Google Scholar
Rall, T.W. (1972). Role of adenosine 3’, 5’-monophosphate (cyclic AMP) in actions of catecholamines. Pharmac. Rev. 24, 399409.Google Scholar
Rall, T.W. and Sattin, A. (1970). Factors influencing the accumulation of cyclic AMP in brain tissue. Advanc. Biochem. Psychopharmacol. 3, 113133.Google Scholar
Rall, T.W. and Sutherland, E.W. (1962). Adenyl cyclase: II. Enzymatically catalyzed formation of adenosine 3’, 5’-phosphate and inorganic pyrophosphate from adenosine triphosphate. J. Biol. Chem. 237, 12281232.Google Scholar
Rall, W., Burke, R.E., Smith, T.-G., Nelson, P.G. and Frank, K. (1967). Dendritic location of synapses and possible mechanisms for the monosynaptic EPSP in motoneurons. J. Neurophysiol. 30, 11691193.Google Scholar
Ramsay, A.G., Gallagher, D.L., Shoemaker, R.L. and Sachs, G. (1976). Barium inhibition of sodium ion transport in toad bladder. Biochim. Biophys. Acta. 436, 617627.Google Scholar
Rasmussen, H. (1970). Cell communication, calcium ion and cyclic adenosine monophosphate. Science 170, 404412.Google Scholar
Rasmussen, H., Goodman, D.B P. and Tenenhouse, A. (1972). The role of cyclic AMP and calcium in cell activation. Critical Rev. Biochem. 1, 95148.Google Scholar
Reader, T.A., De Champlain, J. and Jasper, H. (1976). Catecholamines released from cerebral cortex in the cat; decrease during sensory stimulation. Brain Res. 111, 95108.Google Scholar
Robertis, , Arnaiz, E. De, , G. R. D.-L., Butcher, R.W. and Sutherland, E.W. (1967). Subcellular distribution of adenyl cyclase and cyclic phosphodiesterase in rat brain cortex. J. Biol. Chem. 242, 34873493.Google Scholar
Robison, G.A., Butcher, R.W. and Sutherland, E.W. (1971). Cyclic AMP. New York: Academic Press.Google Scholar
Rodbell, M., Lin, M.C, Salomon, Y., Londos, C, Harwood, J.P., Martin, B.R., Rendell, M. and Berman, M. (1975). Role of adenine and guanine nucleotides in the activity and response of adenylate cyclase systems to hormones. Evidence for multisite transition states. Advanc. Cyclic Nucleotide Res. 5, 329.Google Scholar
Rogus, E.M., Cheng, L.C. and Zierler, K. (1977). β -Adrenergic effect on Na+ -I- K+ transport in rat skeletal muscle. Biochim. Biophys. Acta. 464, 347355.Google Scholar
Roufogalis, B.D. and Belleau, B. (1969). Inhibition of sodium-potassium activated brain adenosine triphosphatase (Na+-K+-ATPase) by adrenergic blocking alkylating agents. Life Sciences 8, 911918.Google Scholar
Sahu, S.K. and Prasad, K.N. (1975). Effect of neurotransmitters and prostaglandin El on cyclic AMP levels in various clones of neuroblastoma cells in culture. J. Neurochem. 24, 12671269.Google Scholar
Sastry, B.S.R. and Phillis, J.W. (1976a). Depression of rat cerebral cortical neurons by Hi and H2 histamine receptor agonists. Europ. J. Pharmac. 38, 269273.Google Scholar
Sastry, B.S.R. and Phillis, J.W. (1976b). Evidence for an ascending inhibitory histaminergic pathway to the cerebral cortex. Can. J. Physiol. Pharmacol. 54, 782786.Google Scholar
Sastry, B.S.R. and Phillis, J.W. (1977a). Metergoline as a selective 5-hydroxytryptamine antagonist in the cerebral cortex. Can. J. Physiol. Pharmacol. 55, 130133.Google Scholar
Sastry, B.S.R. and Phillis, J.W. (1977b). Antagonism of biogenic amine-induced depression of cerebral cortical neurones by Na+-K, K+-K -ATPase inhibitors. Can. J. Physiol. Pharmacol. 55, 170179.Google Scholar
Sattin, A. and Rall, T.W. (1970). The effect of adenosine and adenine nucleotides on the cyclic adenosine 3’, 5’-phosphate content of guinea pig cerebral cortex slices. Molec. Pharmacol. 6, 1323.Google Scholar
Sattin, A. and Rall, T.W. and Zanella, J. (1975). Regulation of cyclic adenosine 3’, 5’-monophosphate levels in guinea pig cerebral cortex by interaction of alpha adrenergic and adenosine receptor activity. J. Pharmacol. Exp. Ther. 192, 2232.Google Scholar
Schaefer, A., Unyi, G. and Pfeifer, A.K. (1972). The effects of a soluble factor and of catecholamines on the activity of adenosine triphosphatase in subcellular fractions of rat brain. Biochem. Pharmac. 21, 22892294.CrossRefGoogle ScholarPubMed
Schmidt, M.J., Palmer, E.C, Dettbarn, W.-D. and Robison, G.A. (1970). Cyclic AMP and adenyl cyclase in the developing rat brain. Develop. Psychobiol. 3, 5367.Google Scholar
Schmidt, M.J., Schmidt, D.E. and Robison, G.A. (1971). Cyclic adenosine monophosphate in brain areas: Microwave irradiation as a means of tissue fixation. Science 173, 11421143.Google Scholar
Schmidt, M.J. and Sokoloff, L. (1973). Activity of cyclic AMP-dependent microsomal protein kinase and phosphorylation of ribosomal protein in rat brain during postnatal development. J. Neurochem. 21, 11931205.Google Scholar
Schmidt, R.F. (1971). Presynaptic inhibition in the vertebrate central nervous system. Ergebn. Physiol. 63, 20107.Google Scholar
Schubert, D., Tarikas, H. and Lacorbiere, M. (1976). Neurotransmitter regulation of adenosine 3’, 5’-monophosphate in clonal nerve, glia and muscle cell lines. Science 192, 471472.Google Scholar
Schultz, J. and Daly, J.W. (1973a). Accumulation of cyclic adenosine 3’, 5’-monophosphate in cerebral cortical slices from rat and mouse: stimulatory effect of Oi - and β -adrenergic agents and adenosine. J. Neurochem. 21, 13191326.Google Scholar
Schultz, J. and Daly, J.W. (1973b). Adenosine 3’, 5’-monophosphate in guinea pig cerebral cortical slices: effects of d-and β -adrenergic agents, histamine, serotonin and adenosine. J. Neurochem. 21, 573579.Google Scholar
Schultz, J. and Hamprecht, B. (1973). Adenosine 3’, 5’-monophosphate in cultured neuroblastoma cells: Effect of adenosine phosphodiesterase inhibitors and benzapines. Arch. Pharmacol. 278, 215225.Google Scholar
Schwartz, J.C. (1975). Histamine as a transmitter in brain. Life Sci. 17, 503518.Google Scholar
Schwartz, J.P. (1976). Catecholamine-mediated elevation of cyclic GMP in the rat C-6 glioma cell line. J. Cyclic Nucleotide Res. 2, 287296.Google Scholar
Schwartz, J.P., Morris, N.R. and Breckenridge, B.M. (1973). Adenosine 3’, 5’-monophosphate in glial tumor cells. J. Biol. Chem. 248, 26992704.Google Scholar
Seeds, N.W. (1971). Biochemical differentiation in reaggregating brain cell culture. Proc. Nat. Acad. Sci. U.S.A. 68, 18581861.Google Scholar
Seeman, P. (1972). The membrane actions of anaesthetics and tranquilizers. Pharmac. Rev. 24, 583655.Google Scholar
Segal, M. (1974). Lithium and the monoamine neurotransmitters in the rat hippocampus. Nature 250, 7173.Google Scholar
Segal, M. (1975). Physiological and pharmacological evidence for a serotonergic projection to the hippocampus. Brain Res. 94, 115131.Google Scholar
Segal, M. and Bloom, F.E. (1974). The activation of norepinephrine in the rat hippocampus. II. Activation of the input pathway. Brain Res. 72, 99114.Google Scholar
Serck-Hanssen, G. (1974). Effects of theophylline and propranolol on acetylcholine-induced release of adrenal medullary catecholamines. Biochem. Pharmac. 23, 22252234.Google Scholar
Shanta, T.R., Woods, W.D., Waitz-Man, M.B. and Bourne, G.H. (1966). Histochemical method for localization of cyclic 3’, 5’-nucleotide phosphodiesterase. Histochemie 7, 177190.Google Scholar
Shimizu, H., Ichishita, H. and Odagiri, H. (1974). Stimulated formation of cyclic adenosine 3’, 5’-monophosphate by aspartate and glutamate in cerebral cortical slices of guinea pig. J. Biol. Chem. 249, 59555962.Google Scholar
Shimizu, H., Tanaka, S., Suzuki, T. and Matsukado, Y. (1971). The response of human cerebrum adenyl cyclase to biogenic amines. J. Neurochem. 18, 11571161.Google Scholar
Shinnick-Gallagher, P., Williams, B.J. and Gallagher, J.P. (1976). Biochemical and electrophysiological studies of cyclic nucleotides and their effects in the rat superior cervical ganglion. Proc. Soc. Neurosci. 800.Google Scholar
Shoemaker, W.J., Balentine, L.T., Siggins, G.R., Hoffer, B.J., Henriksen, S.J. and Bloom, F.E. (1975). Characteristics of the release of adenosine 3’, 5’-monophosphate from micropipets by microiontophoresis. J. Cyclic Nucleotide Res. 1, 97106.Google Scholar
Siggins, G.R., Battenberg, E.F., Hoffer, B.J., Bloom, F.E. and Steiner, A.L. (1973). Noradrenergic stimulation of cyclic adenosine monophosphate in rat Purkinje neurons: An im-munocytochemical study. Science 179, 585588.Google Scholar
Siggins, G.R. and Henriksen, S.J. (1975). Analogs of cyclic adenosine monophosphate: correlation of inhibition of Purkinje neurons with protein kinase activation. Science 189, 559561.Google Scholar
Siggins, G.R., Hoffer, G.J. and Bloom, F.E. (1971a). Studies on norepinephrine-containing afferents to Purkinje cells of rat cerebellum. III. Evidence for mediation of norepinephrine effects by cyclic 3’ 5’-adenosine monophosphate. Brain Res. 25, 535553.Google Scholar
Siggins, G.R., Hoffer, B. and Bloom, F. (1971b). Prostaglandin-norepinephrine interactions in brain: microelectrophoretic and histochemical correlates. Ann. N. Y. Acad. Sci. 180, 302323.Google Scholar
Siggins, G.R., Hoffer, B.J. and Ungerstedt, U. (1974). Electrophysiological evidence for involvement of cyclic adenosine monophosphate in dopamine responses of caudate neurons. Life Sci. 15, 779792.Google Scholar
Simon, L.N., Shuman, D.A. and Robins, R.K. (1973). The chemistry and biological properties of nucleotides related to nucleoside 3’, 5’-cyclic phosphates. Advanc. Cyclic Nucleotide Res. 3, 225353.Google Scholar
Skolnick, P., Daly, J.W., Freedman, R. and Hoffer, B.J. (1976). Interrelationship between catecholamine-stimulated formation of adenosine 3’, 5’-monophosphate in cerebellar slices and inhibitory effects on cerebellar Purkinje cells. Antagonism by neuroleptic compounds. J. Pharm. Exp. Ther. 197, 280292.Google Scholar
Skolnick, P., Huang, M., Daly, J. and Hoffer, B.J. (1973). Accumulation of adenosine 3’, 5’-monophosphate in incubated slices from discrete regions of squirrel monkey cortex: effect of norepinephrine, serotonin and adenosine. J. Neurochem. 21, 237240.Google Scholar
Skou, J.C. (1965). Enzymatic basis for active transport of Na+-K and K+-Kacross cell membrane. Physiol. Rev. 45, 596617.Google Scholar
Smith, T.G., Wuerker, R.B. and Frank, K. (1967). Membrane impedance changes during synaptic transmission in cat spinal motoneurons. J. Neurophysiol. 30, 10721096.Google Scholar
Snyder, F.F. and Seegmiller, J.E. (1976). The adenosine-like effect of exogenous cyclic AMP upon nucleotide and PP-ribose-P concentrations of cultured human lymphoblasts. FEBS Letters 66, 102106.Google Scholar
Somlyo, A.P. AND Somlyo, A.V. (1969). Pharmacology of excitation-contraction coupling in vascular smooth muscle and in avian slow muscle. Fedn. Proc. 28, 16341642.Google Scholar
Specht, S.C. and Robinson, J.D. (1973). Stimulation of the (Na+-K +-K K+-K+-K)-dependent adenosine triphosphatase by amino acids and phosphatidylserine: chelation of trace metal inhibitors. Arch. Biochem. Biophys. 154, 314323.Google Scholar
Stahl, W.L. and Broderson, S.H. (1976). Localization of Na+-K, K+-K-ATPase in brain. Fedn. Proc. 35. 12601265.Google Scholar
Standaert, F.G., Dretchen, K.L., Skirboll, L.R. and Mogenroth, V.H. (1976a). Effects of cyclic nucleotides on mammalian motor nerve terminals. J. Pharmac. Exp. Ther. 199. 544552.Google Scholar
Standaert, F.G., Dretchen, K.L., Skirboll, L.R. and Mogenroth, V.H. (1976b). A role of cyclic nucleotides in neuromuscular transmission. J. Pharmacol. Exp. Ther. 199, 553564.Google Scholar
Stein, L., Belluzzi, J.D. and Wise, C.D. (1975). Memory enhancement by central administration of norepinephrine. Brain Res. 84, 329335.Google Scholar
Steiner, A.L., Ferrendelli, J.A. and Kipnis, D.M. (1972). Radioimmunoassay for cyclic nucleotides. III. Effect of ischemia, changes during development and regional distribution of adenosine 3’, 5’-monophosphate and guanosine 3’, 5’-monophosphate in mouse brain. J. Biol. Chem. 247, 11211124.Google Scholar
Stitt, J.T. and Hardy, J.D. (1975). Microelectrophoresis of PGE1 onto single units in the rabbit hypothalamus. Amer. J. Physiol. 229, 240245.Google Scholar
Stjärne, L. (1976). Relative importance of calcium and cyclic AMP for noradrenaline secretion from sympathetic nerves of guinea-pig vas deferens and for prostaglandin ?-induced depression of noradrenaline secretion. Neuroscience 1, 1922.Google Scholar
St. Louis, P. and Sulakhe, P.V. (1976). Adenylate cyclase, guanylate cyclase and cyclic nucleotide phosphodiesterases of guinea-pig cardiac sarcolemma. Biochem. J. 158, 535541.Google Scholar
Stone, T.W. (1973). Pharmacology of pyramidal tract cells in the cerebral cortex. Noradrenaline and related substances. Arch. Pharmacol. 278, 333346.Google Scholar
Stone, T.W., Taylor, D.A. and Bloom, F.E. (1975). Cyclic AMP and cyclic GMP may mediate opposite neuronal responses in the rat cerebral cortex. Science 187, 845846.Google Scholar
Sturgill, T.W., Schrier, B.K. and Gilman, A.G. (1975). Stimulation of cyclic AMP accumulation by 2-chloroadenosine: lack of incorporation of nucleoside into cyclic nucleotides. J. Cyclic Nucleotide Res. 1, 2130.Google Scholar
Sulakhe, P.V., Jan, S.-H. and Sulakhe, S.J. (1977). Studies on the stimulation of (Na+-K'-K +-K) ATPase of neural tissues by catecholamines. Gen. Pharmac. 8. 3741.Google Scholar
Sulakhe, P.V. and St. LOUIS, P.J. (1976). Membrane phosphorylation and calcium transport in cardiac and skeletal muscle membranes. Gen. Pharmac. 7, 313319.Google Scholar
Sulakhe, P.V., Sulakhe, S.J., Leung, N.L.-K., St. Louis, P.J. and Hickie, R.A. (1976). Guanylate cyclase. Subcellular distribution in cardiac muscle, skeletal muscle, cerebral cortex and liver. Biochem. J. 157, 705712.Google Scholar
Sutherland, E.W. and Rall, T.W. (1958). Fractionation and characterization of a cyclic adenine ribonucleotide formed by tissue particles. J. Biol. Chem. 232. 10771091.Google Scholar
Sutherland, E.W., Rall, T.W. and Menon, T. (1962). Adenyl cyclase: I. Distribution, preparation and properties. J. Biol. Chem. 237, 12201227.Google Scholar
Sutherland, E.W., Robison, G.A. and Butcher, R.W. (1968). Some aspects of the biological role of adenosine 3’, 5’-monophosphate (cyclic AMP). Circulation 37. 279306.Google Scholar
Swartz, B. and Woody, C.D. (1976). Correlated effects of acetylcholine (ACh) and cyclic GMP(cGMP) on input resistance (Rm) of neocortical neurons in awake cats. Neuroscience Abstr. II 1156. p. 802.Google Scholar
Swislocki, N.I. and Tierney, J. (1973). Solubilization, stabilization, and partial purification of brain adenylate cyclase from rat. Biochemistry 12, 18621866.Google Scholar
Szerb, J.C. and Somogyi, G.T. (1973). Depression of acetylcholine release from cerebral cortical slices by cholinesterase inhibition and by oxotremorine. Nature New Biol. (Lond.) 241, 121122.Google Scholar
Tanaka, C, Inagaki, C. and Fujiwara, H. (1976). Labeled noradrenaline release from rat cerebral cortex following electrical stimulation of locus coeruleus. Brain Res. 106, 384389.Google Scholar
Taxi, J. and Mikulajova, M. (1976). Some cytochemical and cytological features of the so-called SIF cells of the superior cervical ganglion of the rat. J. Neurocytol. 5. 283295.Google Scholar
Taylor, D., Nathanson, J., Hoffer, B., Olson, L. and Seiger, A. (1976). Lead blockade of noradrenergic inhibition in cerebellar Purkinje neurons. Neuroscience Abstracts II, (1). 507.Google Scholar
Ting-Beall, H.P., Clark, D.A., Suelter, C.H. and Wells, W.W. (1973). Studies on the interaction of chick brain microsomal (Na+-K -1- K+-K)-ATPase with copper. Biochim. Biophys. Acta. 291. 229236.Google Scholar
Tomiyama, A., Takayanagi, I., Saeki, M. and Takagi, K. (1973). Beta-adrenergic blocking action of ruthenium red. Jap. J. Pharmacol. 23, 889891.Google Scholar
Triggle, D.J. (1971). Neurotransmitter-Receptor Interactions. New York: Academic Press.Google Scholar
Tsuzuki, J. and Newburgh, R.W. (1975). Inhibition of 5’-nucleotidase in rat brain by methylxanthines. J. Neurochem. 25, 895896.Google Scholar
Ueda, T., Maeno, H. and Greengard, P. (1973). Regulation of endogenous phosphorylation of specific proteins in synaptic membrane fractions from rat brain by adenosine 3’, 5’-monophosphate. J. Biol. Chem. 248, 82958305.Google Scholar
Ungerstedt, U. (1971). Stereotaxic mapping of the monoamine pathways in the rat brain. Acta physiol. scand. Suppl. 367, 148.Google Scholar
Uzunov, P. and Weiss, B. (1972). Psychopharmacological agents and the cyclic AMP system of rat brain. Advanc. Cyclic Nucleotide Res. 1, 435453.Google Scholar
Uzunov, P., Shein, H.M. and Weiss, B. (1973). Cyclic AMP phosphodiesterase in cloned astrocytoma cells: norepinephrine induces a specific enzyme form. Science 180, 304306.Google Scholar
Von Hungen, K. and Roberts, S. (1974). Neurotransmitter-sensitive adenylate cyclase systems in the brain. Rev. Neurosci. 1, 231281.Google Scholar
Von Hungen, K., Roberts, S. and Hill, D.F. (1974). Developmental and regional variations in neurotransmitter-sensitive adenylate cyclase systems in cell-free preparations from rat brain. J. Neurochem. 22, 811819.Google Scholar
Von Hungen, K., Roberts, S. and Hill, D.F. (1975). Serotonin-sensitive adenylate cyclase activity in immature rat brain. Brain Res. 84, 257267.Google Scholar
Walton, K.G. and Baldessarini, R.J. (1976). Effects of Mn2+ and other divalent cations on adenylate cyclase activity in rat brain. J. Neurochem. 27, 557564.Google Scholar
Watanabe, S. and Koketsu, K. (1973). 5-HT hyperpolarization of bullfrog sympathetic ganglion cell membrane. Experientia 29, 13701372.Google Scholar
Weight, F.F. and Padjen, A. (1973a). Slow synaptic inhibition in sympathetic ganglion cells: evidence for synaptic inac-tivation of sodium conductance in sympathetic ganglion cells. Brain Res. 55, 219224.Google Scholar
Weight, F.F. and Padjen, A. (1973b). Acetylcholine and slow synaptic inhibition in frog sympathetic ganglion cells. Brain Res. 55, 225228.Google Scholar
Weight, F.F., Petzold, G. and Greengard, P. (1974). Guanosine 3’. 5’-monophosphate in sympathetic ganglia: increase associated with synaptic transmission. Science. N.Y.. 186, 942944.Google Scholar
Weight, F.F. and Votava, J. (1970). Slow synaptic excitation in sympathetic ganglion cells: Evidence for synaptic inac-tivation of potassium conductance. Science. N.Y., 755758.Google Scholar
Weir, M.C.L. and McLennan, H. (1963). The action of catecholamines in sympathetic ganglia. Can. J. Biochem. Physiol. 41, 26272636.Google Scholar
Weiss, B. (1973). Selective regulation of the multiple forms of cyclic nucleotide phosphodiesterase by norepinephrine and other agents. In: Frontiers in Catecholamine Research, Edit. Usdin, E. and Snyder, S.H. New York: Pergamon Press, pp. 327333.Google Scholar
Weiss, B. (1975). Differential activation and inhibition of the multiple forms of cyclic nucleotide phosphodiesterase. Advanc. Cyclic Nucleotide Res. 5. 195211.Google Scholar
Weiss, B. and Costa, E. (1968). Regional and subcellular distribution of adenyl cyclase and 3’, 5’-cyclic nucleotide phosphodiesterase in brain and pineal gland. Biochem. Pharmacol. 17. 21072116.Google Scholar
Weiss, B. and Strada, J.J. (1972). Neuroendocrine control of the cyclic AMP system of brain and pineal gland. Advanc. Cyclic Nucleotide Res. 1. 357374.Google Scholar
Weller, M. and Rodnight, R. (1973). Protein kinase activity in membrane preparations from ox brain: stimulation of intrinsic activity by adenosine 3’. 5’-cyclic monophosphate. Biochem. J. 132. 483492.Google Scholar
Wellmann, W. and Schwabe, U. (1973). Effects of prostaglandins E1. E2 and F2Q on cyclic AMP levels in brain in vivo. Brain Res. 59. 371378.Google Scholar
Westfall, T.C. Kitay, D. and Wahl, G. (1976). The effect of cyclic nucleotides on the release of 3H-dopamine from rat striatal slices. J. Pharmac. Exp. Ther. 199, 149157.Google Scholar
Werman, R. (1966). A Review - Criteria for identification of a central nervous system transmitter. Comp. Biochem. Physiol. 18, 745766.Google Scholar
Williams, T.H., Black, A.C. Chiba, T. and Bhalla, R.C. (1975). Morphology and biochemistry of small intensely fluorescent cells of sympathetic ganglia. Nature 256. 315317.Google Scholar
Wilson, D.F. (1974). The effects of di-butyryl cyclic adenosine 3’. 5’-mono-phosphate, theophylline and aminophylline on neuromuscular transmission in the rat. J. Pharmacol. Exp. Ther. 188. 447452.Google Scholar
Woods, W.T. and Lieberman, E.M. (1976). The effects of papaverine on sodium-potassium adenosine triphosphatase and the ouabain sensitive electrical properties of crayfish nerve. Neuroscience 1. 383390.Google Scholar
Wooten, F.G., Thoa, N.B., Kopin, I.J. and Axelrod, J. (1973). Enhanced release of dopamine-/3-hydroxylase and norepinephrine from sympathetic nerves by dibutyryl cyclic adenosine 3’, 5’ monophosphate and theophylline. Molec. Pharmacol. 9, 178183.Google Scholar
Yarbrough, G.G. (1976). Ouabain antagonism of noradrenaline inhibitions of cerebellar Purkinje cells and dopamine inhibitions of caudate neurons. Neurophar-mac. 15, 335338.Google Scholar
Yarbrough, G.G., Lake, N. and Phillis, J.W. (1974). Calcium antagonism and its effect on the inhibitory actions of biogenic amines on cerebral cortical neurones. Brain Res. 67, 7788.Google Scholar
York, D.H. (1975). Amine receptors in CNS. II. Dopamine. In: Handbook of Psychopharmacology, Vol. 6. Iverson, L.L. Iverson, S.D. and Snyder, S.H. Eds., pp. 2361, New York: Plenum Press.Google Scholar
Yoshimura, K. (1973). Activation of Na-K activated ATPase in rat brain by catecholamine. Biochem, J. (Tokyo) 74, 389391.Google Scholar
Yount, R.G. (1975). ATP analogs. Advances in Enzymology. 43. 156.Google Scholar
Zanella, J. and Rall, T.W. (1973). Evaluation of electrical pulses and elevated levels of potassium ions as stimulants of adenosine 3’. 5’-monophosphate (cyclic AMP) accumulation in guinea pig brain. J. Pharmac. Exp. Ther. 186. 241252.Google Scholar
Zieglgänsberger, W. and Reiter, Ch. (1974). A cholinergic mechanism in the spinal cord of cats. Neuropharmacol. 13, 519527.Google Scholar