Hostname: page-component-848d4c4894-pftt2 Total loading time: 0 Render date: 2024-06-10T12:21:05.461Z Has data issue: false hasContentIssue false

Evolution and overview of classical transmitter molecules and their receptors

Published online by Cambridge University Press:  06 April 2009

R. J. Walker
Affiliation:
Department of Physiology and Pharmacology, Biomedical Sciences, Bassett Crescent East, The University of Southampton, Southampton, SO16 7PX, UK
H. L. Brooks
Affiliation:
Department of Physiology and Pharmacology, Biomedical Sciences, Bassett Crescent East, The University of Southampton, Southampton, SO16 7PX, UK
L. Holden-Dye
Affiliation:
Department of Physiology and Pharmacology, Biomedical Sciences, Bassett Crescent East, The University of Southampton, Southampton, SO16 7PX, UK

Summary

All the classical transmitter ligand molecules evolved at least 1000 million years ago. With the possible exception of the Porifera and coelenterates (Cnidaria), they occur in all the remaining phyla. All transmitters have evolved the ability to activate a range of ion channels, resulting in excitation, inhibition and biphasic or multiphasic responses. All transmitters can be synthesised in all three basic types of neurones, i.e. sensory, interneurone and motoneurone. However their relative importance as sensory, interneurone or motor transmitters varies widely between the phyla. It is likely that all neurons contain more than one type of releasable molecule, often a combination of a classical transmitter and a neuroactive peptide. Second messengers, i.e. G proteins and phospholipase C systems, appeared early in evolution and occur in all phyla that have been investigated. Although the evidence is incomplete, it is likely that all the classical transmitter receptor subtypes identified in mammals, also occur throughout the phyla. The invertebrate receptors so far cloned show some interesting homologies both between those from different invertebrate phyla and with mammalian receptors. This indicates that many of the basic receptor subtypes, including benzodiazepine subunits, evolved at an early period, probably at least 800 million years ago. Overall, the evidence stresses the similarity between the major phyla rather than their differences, supporting a common origin from primitive helminth stock.

Type
Research Article
Copyright
Copyright © Cambridge University Press 1996

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

REFERENCES

Abalis, I. M., Eldefrawi, M. E. & Eldefrawi, A. T. (1983). Biochemical identification of putative GABA/benzodiazepine receptors in housefly thorax muscles. Pesticide Biochemistry and Physiology 20, 3948.CrossRefGoogle Scholar
Ajuh, P. M. & Egwang, T. G. (1994). Cloning of a cDNA encoding a putative nicotinic acetylcholine receptor subunit of the human filarial parasite Onchocerca volvulus. Gene, 144, 127–9.Google ScholarPubMed
Arakawa, S., Gocayne, J. D., Mccrombie, W. R., Urquhart, D. A., Hall, L. M., Fraser, C. M. & Venter, J. C. (1990). Cloning, localization and permanent expression of a Drosophila octopmaine receptor. Neuron, 2, 343–54.CrossRefGoogle Scholar
Arena, J. P., Liu, K. K., Paress, P. S. & Cully, D. F. (1991). Avermectin-sensitive chloride currents by Caenorhabditis elegans RNA in Xenopus oocytes. Molecular Pharmacology 40, 368–74.Google Scholar
Arena, J. P., Liu, K. K., Paress, P. S., Schaeffer, J. M. & Cully, D. F. (1992). Expression of a glutamateactivated chloride current in xenopus oocytes injected with Caenorhabditis elegans RNA: evidence for modulation by avermectin. Molecular Brain Research 15, 339–48.CrossRefGoogle ScholarPubMed
Arshavsky, Y. I., Deliagina, T. G., Gamkrelidze, G. N., Orlovsky, G. N., Panchin, Y. V., Popova, L. B. & Shupliakov, O. V. (1993). Pharmacologically induced elements of the hunting and feeding behaviour in the pteropod mollusc, Clione limacine. I. Effects of GABA. Journal of Neurophysiology 69, 512–21.CrossRefGoogle Scholar
Bai, D. & Sattelle, D. B. (1995). A GABA-B receptor on an identified insect motor neurone. Journal of Experimental Biology 198, 889–94.CrossRefGoogle Scholar
Barkas, T., Mauron, A., Roth, B., Alliod, C., Tzartps, S. J.. & Ballivet, M. (1987). Mapping the main immunogenic region and toxin-binding site of the nicotinic acetylcholine receptor. Science, 235, 7780.Google Scholar
Barnard, E. A., Burnstock, G. & Webb, T. E. (1994). Gprotein coupled receptors for ATP and other nucleotides: a new receptor family. Trends in Pharmacological Sciences 15, 6770.CrossRefGoogle Scholar
Barnes, R. S. K., Calow, P. & Olive, P. J. W. (1988). The Invertebrates: A New Synthesis, pp. 582. Oxford, Blackwell Scientific Publications.Google Scholar
Bascal, Z. A., Holden-Dye, L., Willis, R. J., Smith, S. W. G. & Walker, R. J.. (1996 a). Novel azole derivatives are antagonists at the inhibitory GABA receptor on the somatic muscle cells of the parasitic nematode Ascaris suum. Parasitology, 112, 253–9.CrossRefGoogle ScholarPubMed
Bascal, Z. A., Montgomery, A., Holden-Dye, L., Williams, R. G., Thorndyke, M. C. & Walker, R. J. (1996 b). NADPH diaphorase activity in peptidergic neurones of the parasitic nematode, Ascaris suum. Parasitology 112, 125–34.CrossRefGoogle ScholarPubMed
Bascal, Z. A., Montgomery, A., Holden-Dye, L., Williams, R. G. & Walker, R. J. (1995). Histochemical mapping of NADPH diaphorase in the nervous system of the parasitic nematode, Ascaris suum. Parasitology 110, 625–37.Google Scholar
Battelle, B. A., Calman, B. G., Andrews, A. W., Grieco, F. D., Mieziva, M. B., Callaway, J. C. & Stuart, A. E. (1991). Histamine: a putative afferent neurotransmitter in Limulus eyes. Journal of Comparative Neurology 305, 527–42.CrossRefGoogle ScholarPubMed
Baux, G., Fossier, P., Trudeau, L. E. & Tauc, L. (1992). Presynaptic receptors for FMRFamide, histamine and buccalin regulate acetylcholine release at a neuroneuronal synapse of Aplysia by modulating N-type Ca++ channels. Journal of Physiology (Paris) 86, 3—13.Google Scholar
Benjamin, P. R. & Elliott, C. J. H. (1989). Snail feeding oscillator: the central pattern generator and its control by modulatory interneurones. In Neuronal and Cellular Oscillators (ed Jacklet, J. W.), pp. 173–4. New York, Dekker.Google Scholar
Berridge, M. (1987). Inositol triphosphate and diacylglycerol: Two interacting second messengers. Annual Review of Biochemistry 56, 159—93.CrossRefGoogle Scholar
Berry, M. S. & Cottrell, G. A. (1975). Excitatory, inhibitory and biphasic synaptic potentials mediated by an identified dopamine-containing neurone. Journal of Physiology 244, 589612.CrossRefGoogle ScholarPubMed
Bertrand, D., Ballivet, M., Gomez, M., Bertrand, S., Phannavong, B. & Gundelfinger, E. D. (1994). Physiological properties of neuronal nicotinic receptors reconstituted from the vertebrate beta-2 subunit and Drosophila alpha-subunits. European Journal of Neuroscience 6, 869–75.CrossRefGoogle ScholarPubMed
Bertrand, D., Bertrand, S. & Ballivet, M. (1995). Reconstitution of a functional homomeric neuronal nicotinic acetylcholine receptor from C. elegans. Society for Neuroscience Abstracts 21, 334—6.Google Scholar
Bhandal, N. S., Ramsey, R. L., Harvaey, R. J., Darlison, M. G. & Usherwood, P. N. R. (1995). Channel gating in the absence of agonist by a homo-oligomeric molluscan GABA receptor expressed in Xenopus oocytes from a cloned cDNA. Invertebrate Neuroscience 1, 267–72.CrossRefGoogle ScholarPubMed
Bolshakov, V. YU., Gapaon, S. & Magazanik, L. G. (1991). Different types of glutamate receptors in isolated and identified neurones of the mollusc, Planorbarius corneus. Journal of Physiology 439, 1535.Google Scholar
Bossy, B., Ballivet, M. & Spierer, P. (1988). Conservation of neural nicotinic acetylcholine receptors from Drosophila to vertebrate central nervous systems. European Molecular Biology Organization Journal 7, 611–18.CrossRefGoogle ScholarPubMed
Brake, A. J., Wagenbach, M. J. & Julius, D. (1994). New structural motif for ligand-gated ion channels defined by an ionotropic ATP receptor. Nature 371, 519–23.CrossRefGoogle ScholarPubMed
Breer, H. & Benke, D. (1985). Synthesis of acetylcholine receptors in Xenopus oocytes induced by poly AmRNA from locust nervous tissue. Naturwissenschaften 72, 213–14.Google Scholar
Breer, H. & Sattelle, D. B. (1987). Molecular properties and functions of insect acetylcholine receptors. Journal of Insect Physiology 33, 771–90.CrossRefGoogle Scholar
Brezina, V., Bank, B., Cropper, E. C., Rosen, S., Vilim, F. S., Kupfermann, I. & Weiss, K. R. (1995). Nine members of the Myomodulin family of peptide cotransmitters at the B—16—ARC neuromuscular junction of Aplysia. Journal of Neurophysiology 74, 5471.Google Scholar
Brooks, H. L., Foreman, R. C., Burke, J. F., Holden-Dye, L. & Walker, R. J.. (1995). Identification of nicotinic acetylcholine receptor (nAChR) genes from the parasitic nematode Ascaris suum. Society for Neuroscience Abstracts 21, 334–1.Google Scholar
Brown, G. D. & Willows, A. O. D. (1993). The role of glutamate in the swim neural circuit of Tritonia. American Society for Neuroscience 19, 1700.Google Scholar
Buchner, E., Buchner, S., Burg, M. G., Hofbauer, A., Pak, W. L. & Pollack, I. (1993). Histamine is a major mechanosensory neurotransmitter candidate in Drosophila melanogaster. Cell and Tissue Research 273, 119–25.CrossRefGoogle Scholar
Burg, M. G., Sarthy, P. V., Koliantz, G. & Pak, W. L. (1993). Genetic and molecular identification of a Drosophila histidine decarboxylase gene required in photoreceptor transmitter synthesis. European Molecular Biology Organization Journal 12, 911–9.CrossRefGoogle ScholarPubMed
Callaway, J. C. & Stuart, A. E. (1989). Biochemical and physiological evidence that histamine is the transmitter of barnacle photoreceptors. Vision Neuroscience 3, 311–25.CrossRefGoogle ScholarPubMed
Carr, W. E. S., Ache, B. W. & Gleeson, R. A. (1987). Chemoreceptors of crustaceans: Similarities to receptors for neuroactive substances in internal tissues. Environmental Health Perspectives 71, 31—46.CrossRefGoogle ScholarPubMed
Chen, R., Belelli, D., Lambert, J. J., Peters, J. A., Reyes, A. & Lan, N. C. (1994). Cloning and functional expression of a Drosophila gamma-aminobutyric acid receptor. Proceedings of the National Academy of Sciences, USA 91, 6069–73.CrossRefGoogle ScholarPubMed
Chichery, R. & Chichery, M. P. (1994). NADPHdiaphorase in a cephalopod brain (Sepia): presence in an analogue of the cerebellum. Neuroreport 5, 1273–6.CrossRefGoogle Scholar
Colas, J. F., Launay, J. M., Kellermann, O., Rosay, P. & Maroteaux, L. (1995). Drosophila 5-HT-2 serotonin receptor: Coexpression with fushi-tarazu during segmentation. Proceedings of the National Academy of Sciences, USA 92, 5441–5.CrossRefGoogle Scholar
Colbert, E. H. & Morales, M. M. (1991). Evolution of the Vertebrates (eds Colbert, E. H. & Morales, M. M.). vol. 4, New York, Wiley-Liss, Inc.Google Scholar
Colquhoun, L., Holden-Dye, L. & Walker, R. J. W. (1991). The pharmacology of cholinoceptors on the somatic muscle cells of the parasitic nematode Ascaris suum. Journal of Experimental Biology 158, 509–30.Google Scholar
Colquhoun, L. M., Holden-Dye, L. & Walker, R. J. (1992). The action of nicotinic receptor specific toxins on the somatic muscle cells of the parasitic nematode, Ascaris suum. Molecular Neuropharmacology 3, 1116.Google Scholar
Conte, A. & Ottaviani, E. (1995). Nitric oxide synthase activity in molluscan hemocytes. Federation of European Biochemical Societies Letters 365, 120–4.CrossRefGoogle ScholarPubMed
Cox, R. T. L. & Walker, R. J.. (1987). An analysis of the adenosine receptors for modulation of an excitatory acetylcholine response on an identified Helix neurone. Comparative Biochemistry and Physiology 88C, 121–30.Google Scholar
Cully, D. F., Vassilatis, D. K., Liu, K. K., Paress, P. S., Van Der Ploeg, L. H. T., Schaeffer, J. M. & Arena, J. P. (1994). Cloning of an avermectin-sensitive glutamategated chloride channel from Caenorhabditis elegans. Nature 371, 707–11.Google Scholar
Culotti, J. G. & Klein, w. L. (1983). Occurence of muscarinic acetylcholine receptors in wild type and cholinergic mutants of Caenorhabditis elegans. Journal of Neuroscience 3, 359—68.CrossRefGoogle Scholar
Dal, N. & Kandel, E. R. (1993). L-Glutamate may be the fast excitatory transmitter of Aplysia sensory neurons. Proceedings of the National Academy of Sciences, USA 90, 7163–67.CrossRefGoogle Scholar
Dale, H. H. (1914). The action of certain esters and ethers of choline, and their relation to muscarine. Journal of Pharmacology and Experimental Therapeutics 6, 147–90.Google Scholar
Dale, H. H. (1935). Pharmacology and nerve endings. Proceedings of the Royal Society of Medicine 28, 319–32.CrossRefGoogle ScholarPubMed
Dale, N. & Roberts, A. (1984). Excitatory amino acid receptors in Xenopus embryo spinal cord and their role in the activation of Swimming. Journal of Physiology 348, 527–43.CrossRefGoogle ScholarPubMed
Darlison, M. G. (1992). Invertebrate GABA and glutamate receptors: molecular biology reveals predictable structures but some unusual pharmacologies. Trends in Neurophysiological Sciences 15, 469–74.Google ScholarPubMed
Darlison, M. G. & Albrecht, B. E. (1995). GABA-A receptor subtypes: which, where and why? Seminars in Neuroscience 7, 115–26.CrossRefGoogle Scholar
Darlison, M. G., Hutton, M. L. & Harvey, R. J. (1993). Molluscan ligand-gated ion-channel receptors. In Comparative Molecular Neurobiology (ed Pichon, Y.), pp. 4864. Basel, Birkhauser Verlag.CrossRefGoogle Scholar
Darlison, M. G., Stuhmer, T., Harvey, R. J., Dautzenberg, F. M., Kim, H. C., Zimmerman, C., Amar, M., Bermudez, I. & Van Minnen, J. (1994). Molecular characterization and functional expression of molluscan ion-channel receptors that can be activated by either γ—aminobutyric acid or L-glutamate. Netherlands Journal of Zoology 44, 473–85.Google Scholar
David, J. A. & Pitman, R. M. (1993). The pharmacology of α—bungarotoxin-resistant acetylcholine receptors on an identified cockroach motoneurone. Journal of Comparative Physiology A 172, 359–68.CrossRefGoogle Scholar
David, J. A. & Sattelle, D. B. (1984). Actions of cholinergic pharmacological agents on the cell body membrane of the fast coxal depressor motoneurone of the cockroach (Periplaneta Americana). Journal of Experimental Biology 108, 119–36.CrossRefGoogle Scholar
Dawson, T. M. & Snyder, S. H. (1994). Gases as biological messengers: Nitric oxide and carbon monoxide. Journal of Neuroscience 14, 5147—59.CrossRefGoogle ScholarPubMed
De Camilli, p. & Navone, F. (1987). Regulated secretory pathways of neurones and their relation to the regulated secretory pathways of endocrine cells. Annals of the New York Academy of Sciences 493, 461–79.CrossRefGoogle Scholar
Djamgoz, M. B. A. (1995). Diversity of GABA receptors in the vertebrate outer retina. Trends in Neurosciences 18, 118–20.Google Scholar
Downer, R. G. H., Hiripi, L. & Juhos, S. (1993). Characterization of the tyraminergic system in the central nervous system of the locust, Locusta migratoria migratoides. Neurochemistry Research 18, 1245–8.CrossRefGoogle ScholarPubMed
Dubas, F. (1991). Actions of putative amino acid neurotransmitters on the neuropile arborizations of locust flight motoneurones. Journal of Experimental Biology 155, 337–56.CrossRefGoogle Scholar
Dudai, Y. & Zvi, S. (1982). Aminergic receptors in Drosophila melanogaster. Properties of (3H)dihydroergotamine binding sites. Journal of Neurochemistry 38, 1551–8.CrossRefGoogle ScholarPubMed
Dunbar, s. J.. & Piek, T. (1983). The action of iontophoretically applied L-glutamate on an insect visceral muscle. Archives of Insect Biochemistry and Physiology 1, 93103.CrossRefGoogle Scholar
Elofsson, R., Carlberg, M., Moroz, L. L., Nezlin, L. & Sakharov, D. (1993). Is Nitric Oxide (NO) produced by invertebrate neurones? Neuroreport 4, 279–82.CrossRefGoogle ScholarPubMed
Elphick, M. R., Green, I. C. & O’shea, M. (1993 a). Nitric oxide synthesis and action in an invertebrate brain. Brain Research 619, 344–6.CrossRefGoogle Scholar
Elphick, M. R., Green, I. C. & O’shea, M. (1994). Nitric oxide signalling in the insect nervous system. In Insect Neurochemistry and Neurophysiology 1993 (eds Borkovec, A. B. & Loeb, M. J.), pp. 129–32; Boca Raton, USA, CRC Press Inc.Google Scholar
Elphick, M. R., Kemenes, G., Staras, K. & O’shea, M. (1995 a). Behavioural role for nitric oxide in chemosensory activation of feeding in a mollusc. Journal of Neuroscience 15, 7653–64.Google Scholar
Elphick, M. R., Rayne, R. C., Riveros-Moreno, V., Moncada, S. & O’shea, M. (1995 b). Nitric oxide synthesis in locust olfactory interneurones. Journal of Experimental Biology 198, 821–9.CrossRefGoogle ScholarPubMed
Elphick, M. R., Riveros-Moreno, V., Moncada, S. & O’shea, M. (1993 b). Identification of nitrergic neurons in invertebrates. Endothelium 1 (supplement), s57.Google Scholar
Evans, P. B., Reale, V., Merzon, R. M. & Villegas, J. (1991). N-methyl-D-aspartate (NMDA) and non- NMDA type glutamate receptors are present on squid axon Schwann cells. Journal of Experimental Biology 157, 593600.CrossRefGoogle ScholarPubMed
Evans, P. D. (1981). Multiple receptor types for octopamine in the locust. Journal of Physiology 318, 99122.CrossRefGoogle ScholarPubMed
Evans, P. D. (1984). A modulatory octopaminergic neurone increases cyclic nucleotide levels in locust muscle. Journal of Physiology 348, 307–24.CrossRefGoogle Scholar
Evans, P. D. (1993). Molecular studies on insect octopamine receptors. In Invertebrate Molecular Biology (ed. Pichon, Y.), pp. 286–96, Basel, Birkhauser Verlag.Google Scholar
Evans, P. D. & O’shea, M. (1978). The identification of an octopaminergic neurone and the modulation of a myogenic rhythm in the locust. Journal of Experimental Biology 73, 235–60.Google Scholar
Evans, P. D. & Robb, S. (1993). Octopamine receptor subtypes and their modes of action. Neurochemistry Research 18, 869–74.CrossRefGoogle ScholarPubMed
Evans, P. D., Robb, S., Cheek, T. R., Reale, V., Hannan, F. L., Swales, L. S., Hall, L. M. & Midgley, J. M. (1995). Agonist-specific coupling of G-protein-coupled receptors to second messenger systems. Progress in Brain Research 106, (in press).CrossRefGoogle ScholarPubMed
Ffrench-Constant, R. H., Mortlock, D. P., Schaffer, C. D., Macintyre, R. J.. & Roush, R. T. (1991). Molecular cloning and transformation of cyclodience resistance in Drosophila: An invertebrate gamma-aminobutyric acid subtype A receptor locus. Proceedings of the National Academy of Sciences, USA 88, 7209–13.CrossRefGoogle Scholar
Ffrench-Constant, R. T. & Rochelleau, T. (1992). Drosophila cyclodiene resistance gene shows conserved genomic organization with vertebrate gamma-aminobutyric acid A receptors. Journal of Neurochemistry 59, 1562–5.CrossRefGoogle Scholar
Ffrench-Constant, R. H., Rocheleau, T. A., Steichen, J. C. & Chalmers, A. E. (1993). A point mutation in a Drosophila GABA receptor confers insecticide resistance. Nature 363, 449–51.CrossRefGoogle Scholar
Fredholm, B. B., Abbracchio, M. P.., Burnstock, G., Daly, J. W., Harden, T. K., Jacobson, K. A., Leff, P. & Williams, M. (1994). VI. Nomenclature and classification of purinoceptors. Pharmacological Reviews 46, 143–56.Google Scholar
Gelperin, A. (1994 a). Nitric oxide mediates network oscillations of olfactory interneurons in a terrestrial mollusc. Nature 369, 61–3.CrossRefGoogle Scholar
Gelperin, A. (1994 b). Nitric oxide, odour processing and plasticity. Netherlands Journal of Zoology 44, 159–69.CrossRefGoogle Scholar
Glantz, R. M. & Pfeiffer-Linn, C. (1992). NMDA Receptors in Invertebrates. Comparative Biochemistry and Physiology 103C, 243–8.Google Scholar
Gotzes, F., Balfanz, s. & Baumann, A. (1994). Primary structure and functional characterization of a Drosophila dopamine receptor with high homology to human Dl, 5 receptors. Receptors & Channels 2, 131–41.Google Scholar
Gould, S. J.. (1994). The evolution of life on the earth. Scientific American October, 63–9.Google Scholar
Grimmelikhuijzen, C. J.P. (1983). Coexistence of neuropeptides in Hydra. Neuroscience 9, 837—45.CrossRefGoogle ScholarPubMed
Grimmelikhuijzen, C. J. P., Darmer, D., Schmutzler, C., Carstensen, K., Moosler, A., Nothacker, H. P., Reinscheid, R. K., Vollert, H., Rinehart, K. L. & Mcfarlane, I. D. (1992). The peptidergic nervous system of coelenterates. In Progress in Brain Research 92, 127–48.(eds Joosse, J., Buijs, R. M. & Tilders, F. J. H.). Amsterdam, Elsevier Science Publishers B. V.Google Scholar
Gundelfinger, E. D. (1992). How complex is the nicotinic receptor system of insects? Trends in Neurosciences 15, 206–11.CrossRefGoogle ScholarPubMed
Hannan, F. & Hall, L. B. (1993). Muscarinic acetylcholine receptors in invertebrates: Comparisons with homologous receptors from vertebrates. In Comparative Molecular Neurobiology (ed Pichon, Y.). pp 98145; Basel. Birkhauser Verlag.CrossRefGoogle Scholar
Hansson, B. S., Ljungberg, H., Hallberg, E. & Loftstedt, C. (1992). Functional specialization of olfactory glomeruli in a moth. Science 256, 1313–15.CrossRefGoogle Scholar
Hardie, R. C. (1987). Is histamine a neurotransmitter in insect photoreceptors? Journal of Comparative Physiology A 161, 201–13.CrossRefGoogle ScholarPubMed
Hart, A. C., Sims, S. & Kaplan, J. M. (1995). Synaptic code for sensory modalities revealed by C. elegans GLR-1 glutamate receptor. Nature 378, 82–5.CrossRefGoogle Scholar
Harvey, R. J., Schmitt, B., Hermans-Borgmeyer, I., Gundelfinger, E. D., Betz, H. & Darlison, M. G. (1994). Sequence of a Drosophila ligand-gated ionchannel polypeptide with an unusual amino-terminal extracellular domain. Journal of Neurochemistry 62, 2480–3.CrossRefGoogle ScholarPubMed
Harvey, R. J., Vreugdenhil, E., Zaman, S. H., Bhandal, N. S., Usherwood, P. N. R., Barnard, E. A. & Darlison, M. G. (1991). Sequence of a functional invertebrate GABA—A receptor subunit which can form a chimeric receptor with a vertebrate α subunit. European Molecular Biology Organization Journal 10, 3239—45.CrossRefGoogle Scholar
Hashezadeh-Gargari, H. & Freschi, J. E. (1992). Histamine activates chloride conductance in motor neurons of the lobster cardiac ganglion. Journal of Neurophysiology 68, 915.CrossRefGoogle Scholar
Henderson, J. E., Soderlund, D. M. & Knipple, D. C. (1993). Characterization of a putative gammaaminobutyric acid (GABA) receptor subunit gene from Drosophila melanogaster. Biochemical and Biophysical Research Communications 193, 474–82.Google Scholar
Hermans-Borgmeyer, I., Zopf, D., Ryseck, R-P., Hovemann, B., Betz, H. & Gundelfinger, E. D. (1986). Primary structure of a developmentally regulated nicotinic acetylcholine receptor protein from Drosophila. European Molecular Biology Organization Journals, 5 1503–8.Google Scholar
Hetherington, M. S., Mckenzie, J. D., Dean, H. G. & Winlow, W. (1994). A quantitative analysis of the biogenic amines in the central ganglia of the pond snail, Lymnaea stagnalis (L.). Comparative Biochemistry and Physiology 107C, 8393.Google Scholar
Hildebrand, J. G., Christensen, T. A., Arbas, E. A., Hayashi, J. H., Homberg, U., Kanzaki, R. & Stengl, M. (1992). Olfaction in Manduca sexta: Cellular mechanisms of responses to sex pheromones; In Neurotox ‘91: Molecular Basis of Drug & Pesticide Action (ed. Duce, I. R.), pp. 323–38. London, Elsevier Applied Science.CrossRefGoogle Scholar
Hill, S. J. (1990). Distribution, properties and functional characteristics of three classes of histamine receptors. Pharmacogical Reviews 42, 4683.Google Scholar
Hill, R. B. & Huddart, H. (1995). Actions of GTP on molluscan buccal, cardiac and visceral smooth muscle. Comparative Biochemistry and Physiology 111C 389–96.Google Scholar
Hillman, G. R., Ewert, A., Westerfield, L. & Grim, O. (1983). Effects of selected cholinergic and anticholinergic drugs on Brugia malayi (Nematoda). Comparative Biochemistry and Physiology 74C, 299301.Google Scholar
Hirning, L. D., Fox, A. P., Mcclesky, E. W., Olivera, B. M., Thater, S. A., Miller, R. J. & Tsien, R. W. (1988). Dominant role of N-type Ca++ channels in evoked release of norepinephrine from sympathetic neurons. Science 239, 5761.CrossRefGoogle Scholar
Hollmann, M. & Heinemann, S. (1994). Cloned glutamate receptors. Annual Review of Neuroscience 17, 31108.Google Scholar
Hosie, A. M. & Sattelle, D. B. (1996). AlloSteric modulation of an expressed homo-oligomeric GABAgated chloride channel of Drosophila melanogaster. British Journal of Pharmacology 117, 1129–37.CrossRefGoogle ScholarPubMed
Hoyle, C. H. V. & Greenberg, M. J.. (1988). Actions of adenylyl compounds in invertebrates from several phyla: Evidence for internal purinoceptors. Comparative Biochemistry and Physiology 90C, 113–22.Google Scholar
Huddart, H. & Hill, R. B. (1996). Electrical and mechanical characteristics of the atrium of the whelk, Busycon canaliculatum. General Pharmacology 27, 809–18.CrossRefGoogle ScholarPubMed
Hudman, D. & Mcfarlane, I. D. (1995). The role of L-DOPA in the nervous system of sea anemones: A putative inhibitory transmitter in tentacles. Journal of Experimental Biology 198, 1045–50.Google Scholar
Hutton, M. L., Harvey, R. J., Barnard, E. A. & Darlison, M. G. (1991). Cloning of a cDNA that encodes an invertebrate glutamate receptor subunit. Federation of European Biochemical Societies Letters 92, 111–14.Google Scholar
Hutton, M. L., Harvey, R. J., Earley, F. G. P., Barnard, E. A. & Darlison, M. G. (1993). A novel invertebrate GABA-A receptor-like polypeptide: Sequence and pattern of gene expression. Federation of European Biochemical Societies Letters 326, 112—16.CrossRefGoogle ScholarPubMed
Jacklet, J. W. (1995). Nitric oxide is used as an orthograde co-transmitter at identified histaminergic synapses. Journal of Neurophysiology 74, 891–95.CrossRefGoogle Scholar
Jacklet, J. W. & Gruhn, M. (1994). Nitric oxide as a putative transmitter in Aplysia: neural circuits and membrane effects. Netherlands Journal of Zoology 44, 524–34.Google Scholar
Jonas, P. E., Phannavong, B., Schuster, R., Schroder, C. & Gundelfinger, E. D. (1994). Expression of the ligand-binding nicotinic acetylcholine receptor subunit Dα-2 in the Drosophila central nervous system. Journal of Neurobiology 25, 1494–508.Google Scholar
Kawai, N., Saito, M. & Ohsako, s. (1988). Differential expression of glutamate receptors in Xenopus oocytes injected with messenger RNA from lobster. Neuroscience Letters 95, 203—7.Google Scholar
Kehoe, J. S. (1972). The physiological role of three acetylcholine receptors in Aplysia neurones. Journal of Physiology 225, 147–72.Google Scholar
Kehoe, J. S. (1994). Glutamate activates a potassium conductance increase in Aplysia neurons that appear to be independent of G proteins. Neuron 13, 691702.Google Scholar
Knight, G. E., Hoyle, G. H. V. & Burnstock, G. (1992). Responses of the rectum and oesophagus of the snail, Helix aspersa, to purine nucleotides and nucleosides. Comparative Biochemistry and Physiology 103C, 535–40.Google Scholar
Knol, J. C., Ramnatsingh, S., Vankesteren, E. R., Vanminnen, J., Planta, R. J., Van Heerikhuizen, H. & Vreugdenhil, E. (1995). Cloning of a molluscan G protein alpha subunit of the G-q class which is expressed differentially in identified neurons. European Journal of Biochemistry 230, 193—9.CrossRefGoogle Scholar
Laughton, D. L., Amar, M., Thomas, P., Towner, P., Harris, P., Lunt, G. G. & Wolstenholme, A. J. (1994). Cloning of a putative inhibitory amino acid receptor subunit from the parasitic nematode, Haemonchus contortus. Receptors and Channels 2, 155–63.Google Scholar
Laughton, D. L., Wheeler, S. V., Lunt, G. G. & Wolstenholme, A. J. (1995). The β subunit of Caenorhabditis elegans avermectin receptor responds to glycine and is encoded by chromosome International Journal of Neurochemistry 64, 2354–7.CrossRefGoogle ScholarPubMed
Lea, T. J. & Usherwood, P. N. R. (1973). The site of action of ibotenic acid and the identification of two populations of glutamate receptors on insect muscle fibres. Comparative and General Pharmacology 4, 333–50.CrossRefGoogle ScholarPubMed
Leake, L. D. & Moroz, L. L. (1996). Putative nitric oxide synthase (NOS)-containing cells in the central nervous system of the leech, Hirudo medicinalis: NADPH-diaphorase histochemistry. Brain Research (in press).CrossRefGoogle Scholar
Leake, L. D. & Walker, R. J. (1980). Invertebrate Neuropharmacology. Glasgow, U.K. Blackie.Google Scholar
Le Corronc, H., Lapied, B. & Hue, B. (1991). M-2-like presynaptic receptors modulate acetylcholine release in the cockroach (Periplaneta Americana) central nervous system. Journal of Insect Physiology 37, 647–52.Google Scholar
Lee, Y. L., Dobbs, M. B., Verardi, M. L. & Hyde, D. R. (1990). dgq: a Drosophila gene encoding a visual system-specific G molecule. Neuron 5, 889–98.Google Scholar
Lees, G., Beadle, D. J., Neumann, R. & Benson, J. A. (1987). Responses to GABA by isolated insect neuronal somata: pharmacology and modulation by a benzodiazepine and a barbiturate. Brain Research 401, 267–78.Google Scholar
Leinders-Zufall, T., Shepherd, G. M. & Zufall, F. (1995). Regulation of cyclic nucleotide-gated channels and membrane excitability in olfactory receptor cells by carbon monoxide. Journal of Neurophysiology 74, 1498–508.Google Scholar
Li, X. C., Giot, J. F., Kuhl, D., Hen, R. & Kandel, E. R. (1995). Cloning and characterization of two related serotonergic receptors from the brain and the reproductive system of Aplysia that activate phospholipase C. Journal of Neuroscience 15, 7585–91.Google Scholar
Linden, J. (1994). Cloned adenosine A-3 receptors: pharmacological properties, species differences and receptor functions. Trends in Pharmacological Sciences 15, 298306.CrossRefGoogle Scholar
Liu, W. C., Yoon, J., Burg, M., Chen, L. & Pak, W. L. (1995). Molecular characterization of two Drosophila guanylate cyclases expressed in the nervous system. Journal of Biological Chemistry 270, 12418–27.CrossRefGoogle ScholarPubMed
Mccaman, R. E. & Weinreich, D. (1985). Histaminergic synaptic transmission in the cerebral ganglion of Aplysia. Journal of Neurophysiology 53, 1016–37.Google Scholar
Magoski, N. S., Bauce, L. G., Syed, N. I. & Bulloch, A. G. M. (1995). Dopaminergic transmission between identified neurons from the mollusc, Lymnaea stagnalis. Journal of Neurophysiology 74, 1287–300.Google Scholar
Maines, M. D. (1988). Heme oxygenase: function, multiplicity, regulatory mechanisms, and clinical applications. Federation of American Societies for Experimental Biology Journal 2, 2557–68.Google Scholar
Maines, M. D. (1993). Carbon monoxide: an emerging regulator of cGMP in the brain. Molecular Cell Neuroscience 4, 389–97.Google Scholar
Maricq, A. V., Peckol, E., Driscoll, M. & Bargmann, C. I. (1995). Mechanosensory signalling in C. elegans mediated by the GLR—1 glutamate receptor. Nature 378, 7881.Google Scholar
Marshall, J., Buckingham, S. D., Shingai, R., Lunt, G. G., Goosey, M. W., Darlinson, M. G., Sattelle, D. B. & Barnard, E. A. (1990). Sequence and functional expression of a single a-subunit of an insect nicotinic acetylcholine receptor. European Molecular Biology Organization Journal 9, 4391–8.Google Scholar
Martinez, A. (1995). Nitric oxide synthase in invertebrates. Histochemistry Journal 27, 770–6.CrossRefGoogle ScholarPubMed
Mat Jais, A. M., Sharma, R. P., Kerkut, G. A. & Walker, R. J. (1984). The homomethyl-ketone derivative of L-Gluy-DL-Ala-CH2C1 and N-methyl-D-asparatate as selective antagonists against L-glutamate and kainate excitations respectively on Retzius cells of the leech, Hirudo medicinalis. Comparative Biochemistry and Physiology. 77C, 385–98.Google Scholar
Matzel, L. D., Muzzio, I. A. & Rogers, R. F. (1995). Diverse current and voltage responses to Baclofen in an identifed molluscan photoreceptor. Journal of Neurophysiology 74, 506–18.Google Scholar
Mei, L., Roeske, W. R. & Yamamura, H. I. (1989). Molecular pharmacology of muscarinic receptor heterogeneity. Life Sciences 45, 1831–51.CrossRefGoogle ScholarPubMed
Meulemans, A., Mothet, J. P., Schirar, A., Fossier, P., Tauc, L. & Baux, G. (1995). A nitric oxide synthase activity is involved in the modulation of acetylcholine release in Aplysia ganglion neurons: a histological, voltammetric and electrophysiological study. Neuroscience 69, 985–95.CrossRefGoogle ScholarPubMed
Miwa, A., Ui, M. & Kawai, N. (1990). G-protein is coupled to presynaptic glutamate and GABA receptors in lobster neuromuscular synapse. Journal of Neurophysiology 63, 173–80.Google Scholar
Moghadam, H. F., Winlow, W. & Moroz, L. I. (1996). Effects of hydrogen peroxide and nitric oxide (NO) on neuronal discharges and intracellular calcium concentration in the molluscan CNS. Acta Biologica Hungarica 46, 145–53.Google Scholar
Moncada, S., Palmer, R. M. J. & Higgs, E. A. (1991). Nitric oxide: physiology, pathophysiology and pharmacology. Pharmacological Reviews 43, 109–42.Google ScholarPubMed
Moroz, L. L., Gyori, J. & Salanki, J. (1993). NMDA-like receptors in the central nervous system of molluscs. Neuroreport 4, 201–4.Google Scholar
Moroz, L. L., Maslova, G. T., Potapovich, A. I., Rubakhin, S. S., Romanova, E. V., Petrashevskaya, N. N., Budko, D. Y., Winlow, W. & Kostyuk, V. A. (1994a). Molluscan central neurones are highly resistant to free radical damage due to their well developed anti-oxidant defence system. Journal of Physiology 477, 92P.Google Scholar
Moroz, L. L., Park, j.-H. & Winlow, W. (1993). Nitric oxide activates buccal motor patterns in Lymnaea stagnalis. Neuroreport 4, 643–6.Google Scholar
Moroz, L. L., Park, J. H., Budko, D. Y. & Winlow, W. (1994b). Effects of hydrogen peroxide on central identified neurones of the mollusc, Lymnaea stagnalis (L.). Journal of Physiology 475, 139–40.Google Scholar
Moroz, L. L., Winlow, W., Turner, R. W., Bulloch, A. G. M., Lukowiak, K. & Syed, N. I. (1994). Nitric oxide synthase-immunoreactive cells in the CNS and periphery of Lymnaea. Neuroreport 5, 1277–80.CrossRefGoogle ScholarPubMed
Morton, D. B. & Evans, P. D. (1984). Octopamine release from an identified neurone in the locust. Journal of Experimental Biology 113, 177–83.CrossRefGoogle Scholar
Muller, U. & Hildebrandt, H. (1995). The Nitric Oxide/cGMP system in the antennal lobe of Apis mellifera is implicated in integrative processing of chemosensory stimuli. European Journal of Neuroscience 7, 2240–48.CrossRefGoogle ScholarPubMed
Murray, T. F. & Mpitsos, G. J. (1988). Evidence for heterogeneity of muscarinic receptors in the mollusc, Pleurobranchaea. Brain Research Bulletin 21, 181–90.CrossRefGoogle ScholarPubMed
Nathanson, J. A., Kantham, L. & Hunnicutt, E. J. (1989). Isolation and iV-terminal amino acid sequence of an octopamine ligand binding protein. Federation of European Biochemical Societies Letters 259, 177–90.CrossRefGoogle ScholarPubMed
Nistri, A. & Constanti, A. (1979). Pharmacological characterization of different types of GABA and glutamate receptors in vertebrates and invertebrates. Progress in Neurobiology 13, 117–35.Google Scholar
Noda, M., Takahashi, H., Tanabe, T., Toyosato, M., Furutani, Y., Hirose, T., Asai, M., Inayama, S., Miyata, T. & Numa, s. (1982). Primary structure of α-subunit precursor of Torpedo californica acetylcholine receptor deduced from cDNA sequence. Nature 299, 793–7.Google Scholar
Onai, T., Fitzgerald, M. G., Arakawa, S., Gocayne, J. D., Urquhart, D. A., Hall, L. M., Fraser, C. M., McCombie, W. R. & Venter, J. C. (1989). Cloning, sequence analysis and chromosome localization of a Drosophila muscarinic receptor. Federation of European Biochemical Societies Letters 255, 219–25.Google Scholar
Orona, E. & Ache, B. W. (1992). Physiological and pharmacological evidence for histamine as a neurotransmitter in the olfactory CNS of the spiny lobster. Brain Research 590, 136–43.CrossRefGoogle ScholarPubMed
Ortells, M. O. & Lunt, G. G. (1995). Evolutionary history of the ligand-gated ion-channel superfamily of receptors. Trends in Neurosciences 18, 121—7.Google Scholar
Parker, D. (1994). Glutamatergic transmission between antagonistic motor neurones in the locust. Journal of Comparative Physiology A 175, 737–48.Google Scholar
Peroutka, S. J. (1994). 5-Hydroxytryptamine receptors in vertebrates and invertebrates: why are there so many? Neurochemistry International 25, 533–6.Google Scholar
Peroutka, S. J. & Howell, T. A. (1994). The molecular evolution of G protein-coupled receptors: focus on 5-hydroxytryptamine receptors. Neuropharmacology 33, 319–24.Google Scholar
Pfeiffer-Linn, C. & Glantz, R. M. (1991). An arthropod NMDA receptor. Synapse 9, 3542.Google Scholar
Piomelli, D., Volterra, A., Dale, N., Siegelbaum, S. A., Kandel, E. R., Schwartz, J. H. & Belardetti, F. (1987). Lipoxygenase metabolites of arachidonic acid as second messengers for presynaptic inhibition of Aplysia sensory cells. Nature 328, 3843.Google Scholar
Pollack, I. & Hofbauer, A. (1991). Histamine-like immunoreactivity in the visual system and brain of Drosophila melanogaster. Cell and Tissue Research 266 391–8.Google Scholar
Quinlan, E. M. & Murphy, A. D. (1991). Glutamate as a putative neurotransmitter in the buccal central pattern generator of Helisoma trivolvis. Journal of Neurophysiology 66, 1264–71.CrossRefGoogle ScholarPubMed
Ramirez, J. M. & Orchard, I. (1990). Octopaminergic modulation of the forewing stretch receptor in the locust, Locusta migratoria. Journal of Experimental Biology 149, 255–79.Google Scholar
Regulski, M. & Tully, T. (1995). Molecular and biochemical characterization of dNOS: A Drosophila Ca++/calmodulin-dependent nitric oxide synthase. Proceedings of the National Academy of Sciences, USA 92, 9072–6.Google Scholar
Richmond, J. E., Bulloch, A. G. M., Bauce, L. & Lukowiak, K. (1991). Evidence for the presence, synthesis immunoreactivity, and uptake of GABA in the nervous system of the snail, Helisoma trivolvis. Journal of Comparative Neurology 307, 131–43.CrossRefGoogle ScholarPubMed
Richmond, J. E., Murphy, A. D., Lukowiak, K. & Bulloch, A. G. M. (1994). GABA regulates the buccal motor output of Helisoma by two pharmacologically distinct actions. Journal of Comparative Physiology A 174, 593600.Google Scholar
Robb, S., Cheek, T. R., Hannan, F. L., Hall, L. M., Midgley, J. M. & Evans, P. D. (1994). Agonist-specific coupling of a cloned Drosophila octopamine/tyramine receptor to multiple second messenger systems. European Molecular Biology Organization Journal 13, 1325–30.Google Scholar
Roeder, T. (1992). A new octopamine receptor class in locust nervous tissue, the octopamine-3 (OA—3) receptor. Life Sciences 50, 21–8.CrossRefGoogle Scholar
Ryba, N. J. P., Findlay, J. B. C. & Reid, J. D. (1993). The molecular cloning of the squid (Loligo forbesi) visual Gqa subunit and its expression in Saccharomyces cerevisiae. Biochemical Journal 292, 333–41.Google Scholar
Sattelle, D. B., Pinnock, R. D., Wafford, K. A. & David, J. A. (1988). GABA receptors on the cell-body membrane of an identified insect motor neuron. Proceedings of the Royal Society of London, Series BBiological Sciences 232, 443–56.Google Scholar
Saudou, F., Amlaiky, N., Plassat, J. L., Borrelli, E. & Hen, R. (1990). Cloning and characterization of Drosophila tyramine receptor. European Molecular Biology Organization Journal 9, 3611–17.CrossRefGoogle ScholarPubMed
Saudou, F., Boschert, U., Amlaiky, N., Plassat, J. L. & Hen, R. (1992). A family of Drosophila serotonin receptors with distinct intracellular signalling properties and expression patterns. European Molecular Biology Organization Journal 11, 717.Google Scholar
Saudou, F. & Hen, R. (1994). 5-Hydroxytryptamine receptor subtypes in vertebrates and invertebrates. Neurochemistry International 25, 503–32.Google Scholar
Sawada, M., McADOO, D. J., Blankenship, J. E. & Price, C. H. (1981). Modulation of contraction of arterial muscle of Aplysia by glycine and neurone R—4. Brain Research 207, 486–90.CrossRefGoogle Scholar
Sawada, M., Mcadoo, D. J., Ichinose, M. & Priee, C. H. (1984). Influence of glycine and neuron R-14 on contraction of the anterior aorta of Aplysia. Japanese Journal of Physiology 34, 747–67.Google Scholar
Sawada, M., Ichinose, M. & Hara, N. (1995). Nitric oxide induces an increased Na+ conductance in identified neurons of Aplysia. Brain Research 670, 248–56.Google Scholar
Sawruk, E., Hermans-Borgmeyer, I., Betz, H. & Gundelfinger, E. D. (1988). Characterization of an invertebrate nicotinic acetylcholine receptor gene: the ard gene of Drosophila melanogaster. Federation of European Biochemical Societies Letters 235, 40–6.CrossRefGoogle ScholarPubMed
Schloss, P., Betz, H., Schroder, C. & Gundelfinger, E. D. (1991). Neuronal nicotinic acetylcholine receptors in Drosophila: antibodies against an a-like and a non α-subunit recognize the same high-affinity α-bungarotoxin binding complex. Journal of Neurochemistry 75, 1556–62.Google Scholar
Schloss, P., Mayser, W., Gundelfinger, E. D. & Betz, H. (1992). Cross-linking of 125I-α-bungarotoxin to Drosophila head membranes identifies a 42 kDa toxinbinding polypeptide. Neuroscience Letters 145, 63–6.CrossRefGoogle Scholar
Schmid, A. & Duncker, M. (1993). Histamine immunoreactivity in the central nervous system of the spider, Cupiennius salei. Cell and Tissue Research 273, 533–45.CrossRefGoogle Scholar
Schuman, E. M. & Madison, D. V. (1991). The intercellular messenger nitric oxide is required for long-term potentiation. Science 254, 1503–6.Google Scholar
Schuster, C. M., Ultsch, A., Scholoss, P., Cox, J. A., Schmitt, B. & Betz, H. (1991). Molecular cloning of an invertebrate glutamate receptor subunit expressed in Drosophila muscle. Science 254, 112–14.CrossRefGoogle ScholarPubMed
Schuster, R., Phannavong, B., Schroder, C. & Gundelfinger, E. D. (1993 a). Immunohistochemical localization of a ligand-binding and a structural subunit of nicotinic acetylcholine receptors in the central nervous system of Drosophila melanogaster. Journal of Comparative Neurology 335, 149–62.Google Scholar
Schuster, C. M., Ultsch, A., Schmitt, B. & Betz, H. (1993 b). Molecular analysis of Drosophila glutamate receptors. In Comparative Molecular Neurobiology (ed. Pichon, Y.), pp. 234–40. Basel, Birkhauser VerlagCrossRefGoogle Scholar
Schwartz, J. C., Arrang, J. M., Garbarg, M. & Traiffort, E. (1995). Histamine; In Psychopharmacology: The Fourth Generation of Progress (eds Bloom, F. E. & Kupfer, D. J.), pp. 397405.Google Scholar
Shapiro, R. A., Scherer, N. M., Habecker, B. A., Subers, E. B. & Nathanson, N. M. (1988). Isolation, sequence and functional expression of the mouse M1 muscarinic acetylcholine receptor gene. Journal of Biological Chemistry 263, 397403.Google Scholar
Shapiro, R. A., Wakimoto, B. T., Subers, E. B. & Nathanson, N. M. (1989). Characterization and functional expression in mammalian cells of genomic and cDNA clones encoding a Drosophila muscarinic acetylcholine receptor. Proceedings of the National Academy of Sciences, USA 80, 9039–43.Google Scholar
Shimizu, N., Akaike, N., Oomura, Y., Maruhashi, J. & Klee, R. M. (1983). GABA and lioresal actions on the identified Onchidium neuron. Japanese Journal of Physiology 33, 459–67.Google Scholar
Shinozaki, H. & Ishida, M. (1992). A metabotropic Lglutamate receptor agonist: Pharmacological difference between rat central neurons and crayfish neuromuscular junction. Comparative Biochemistry and Physiology 103C, 13–7.Google Scholar
Shortridge, R. D. & McKay, R. A. (1995). Invertebrate phosphatidylinositol-specific phospholipases C and their role in cell signalling. Invertebrate Neuroscience 1, 199–6.Google Scholar
Shotkoski, F., Lee, H. J., Zhang, H. G., Jackson, M. B. & Ffrench-Constant, R. H. (1994). Functional expression of insecticide-resistant GABA receptors from the mosquito, Aedes aegypti. Insect Molecular Biology 3, 283–7.Google Scholar
Skingsley, D. R., Laughlin, S. B. & Hardie, R. C. (1995). Properties of histamine-activated chloride channels in the large monopolar cells of the dipteran compound eye: a comparative study. Journal of Comparative Physiology A 176, 611–23.Google Scholar
Squire, M. D., Tornoe, C., Baylis, H. A., Fleming, J. T., Barnard, E. A. & Sattelle, D. B. (1995). Molecular cloning and functional co-expression of a Caenorhabditis elegans nicotinic acetylcholine receptor subunit (acr—2). Receptors and Channels 3, 107–15.Google Scholar
Stettmeier, H. & Finger, W. (1983). Excitatory postsynaptic channels operated by quisqualate in crayfish muscle. Pflugers Archives 397, 237–42.Google Scholar
Stevenson, P. A. & Pfluger, H. J. (1994). Co-localization of octopamine and FMRFamide-related peptides in identified heart projecting (DUM) neurones in the locust revealed by immunocytochemistry. Brain Research 638, 117–25.CrossRefGoogle Scholar
Sugamori, K. S., Demchyshyn, L. L., McConkey, F., Forte, M. A. & Niznik, H. B. (1995). A primordial dopamine D-l-like adenylyl cyclase-linked receptor from Drosophila melanogaster displaying poor affinity for benzazepines. Federation of European Biochemical Societies Letters 362, 131–8.Google Scholar
Sugamori, K. S., Sunahara, R. K., Guan, H. C., Bulloch, A. G. M., Tensen, C. P., Seeman, P., Niznik, H. B. & Vantol, H. H. M. (1993). Serotonin receptor cDNA cloned from Lymnaea stagnalis. Proceedings of the National Academy of Sciences, USA 90, 11–5.CrossRefGoogle ScholarPubMed
Sun, Y., Rotenberg, M. O. & Maines, M. D. (1990). Development expression of heme oxygenase isozymes in rat brain: two HO—2mRNAs are detected. Journal of Biological Chemistry 265, 8212–7.CrossRefGoogle ScholarPubMed
Takeuchi, H. (1992). Sensitivities of Achatina giant neurones to putative amino acid transmitters. Comparative Biochemistry and Physiology 103C, 112.Google Scholar
Takeuchi, A. & Takeuchi, N. (1964). The effect on crayfish muscle of iontophoretically-applied glutamate. Journal of Physiology 170, 296–17.Google Scholar
Thompson, M., Shotkoski, F. & Ffrench-Constant, R. (1993). Cloning and sequencing of the cyclodience insecticide resistance gene from the yellow fever mosquito, Aedes aegypti: Conservation of the gene and resistance associated mutation with Drosophila. Federation of European Biochemical Societies Letters 325, 187–90.Google Scholar
Thompson, K. J.. & Siegler, M. V. S. (1991). Anatomy and physiology of spiking local and intersegmental interneurones in the median neuroblast lineage of the grasshopper. Journal of Comparative Neurology 305, 659–75.CrossRefGoogle ScholarPubMed
Thorogood, M. S. E. & Brodfuehrer, P. D. (1995). The role of glutamate in swim initiation in the medicinal leech. Invertebrate Neuroscience 1, 223–33.Google Scholar
Tu, Y. & Budelmann, B. u. (1994). The effect of Lglutamate on the afferent resting activity in the cephalopod statocyst. Brain Research 642, 4758.CrossRefGoogle ScholarPubMed
Ultsch, A., Schuster, C. M., Laube, B., Betz, H. & Schmitt, B. (1993). Glutamate receptor of Drosophila melanogaster: primary structure of a putative NMDA receptor protein expressed in the head of the adult fly. Federation of European Biochemical Societies 324, 171–7.Google Scholar
Ultsch, A., Schuster, C. M., Laube, B., Schloss, P., Schmitt, B. & Betz, H. (1992). Glutamate receptors of Drosophila melanogaster: cloning of a kainate-selective subunit expressed in the central nervous system. Proceedings of the National Academy of Sciences, USA 89, 10484–8.Google Scholar
Usherwood, P. N. R. (1994). Insect glutamate receptors. Advances in Insect Physiology 24, 309–41.Google Scholar
Valera, S., Hussy, N., Evans, R. J., Adami, N., North, R. A., Surprenant, A. & Buell, G. (1994). A new class of ligand-gated ion channel defined by P-2X receptor for extracellular ATP. Nature 371, 516–9.Google Scholar
Vanden Broeck, J., Vulsteke, V., Huybrechts, R. & De Loof, A. (1995). Characterization of a cloned locust tyramine receptor cDNA by functional expression in permanently transformed Drosophila S2 cells. Journal of Neurochemistry 64, 2387–95.Google Scholar
Van Golen, F. A., Li, K. W., De Lange, R. P. J., Van Kesteren, R. E., Van Der Schors, R. C. & Geraerts, W. P. M. (1995). Co-localized neuropeptides conopressin and Ala-Pro-Gly-Trp-NH2 have antagonistic effects on the vas deferens of Lymnaea. Neuroscience 69, 1275–87.Google Scholar
Wafford, K. A., Sattelle, D. B., Abalis, I., Eldefrawi, A. T. & Eldefrawi, M. E. (1987). Gamma-aminobutyric acid-activated CL-36-influx- a functional in vitro assay for CNS gamma-aminobutyric acid receptors of insects. Journal of Neurochemistry 48, 177–80.Google Scholar
Wafford, K. A. & Sattelle, D. B. (1989). L-Glutamate receptors on the cell body membrane of an identified insect motor neurone. Journal of Experimental Biology 144; 449–62.Google Scholar
Wafford, K. A., Bai, D. & Sattelle, D. B. (1992). A novel kainate receptor in the insect nervous system. Neuroscience Letters 141, 273–6.CrossRefGoogle ScholarPubMed
Walker, R. J. (1976). The actions of kainic acid and quisqualic acid on the glutamate receptors of three identifiable neurones from the brain of the snail, Helix aspersa. Comparative Biochemistry and Physiology 55C, 61–7.Google Scholar
Walker, R. J. & Holden-Dye, L. (1991). Evolutionary aspects of transmitter molecules, their receptors and channels. Parasitology 102, S7–S29.Google Scholar
Watson, A. D. H. & Seymour-Laurent, K. J. (1993). The distribution of glutamate-like immunoreactivity in the thoracic ganglia of the locust (Schistocerca gregaria). Cell and Tissue Research 273, 557–70.Google Scholar
Whim, M. D. & Lloyd, p. E. (1989). Frequency-dependent release of peptide co-transmitters from identified cholinergic motor neurons in Aplysia. Proceedings of the National Academy of Sciences, USA 86, 9034–8.Google Scholar
Wisden, W. & Seeburg, P. H. (1992). GABA-A receptor channels: from subunits to functional entities. Current Opinions in Neurobiology 2, 263–9.CrossRefGoogle ScholarPubMed
Witz, p., Amlaiky, N., Plassat, J. L., Maroteaux, L., Borrelli, E. & Hen, R. (1990). Cloning and characterization of a Drosophila serotonin receptor that activates adenylate cyclase. Proceedings of the National Academy of Sciences, USA 87, 8940–4.Google Scholar
Yamamura, H. I. & Snyder, S. H. (1974). Muscarinic cholinergic binding in rat brain. Proceedings of the National Academy of Sciences, USA 71, 1725–9Google Scholar
Yeoman, M. S., Parish, D. C. & Benjamin, P. R. (1993). A cholinergic modulatory interneurone in the feeding system of the snail, Lymnaea. Journal of Neurophysiology 70, 3750.CrossRefGoogle ScholarPubMed