Hostname: page-component-848d4c4894-jbqgn Total loading time: 0 Render date: 2024-06-25T06:11:02.786Z Has data issue: false hasContentIssue false

A retinal dark-light switch: A review of the evidence

Published online by Cambridge University Press:  02 June 2009

Ian G. Morgan
Affiliation:
Visual Sciences Group, Centre for Visual Science and Research School of Biological Sciences, Australian National University, Canberra, ACT 2601, Australia
Meeuwis K. Boelen
Affiliation:
Centre for Research on Ageing and Health, Latrobe University, Bendigo, Vic 3550, Australia

Abstract

We propose that there exists within the avian, and perhaps more generally in the vertebrate retina, a two-state nonadapting flip-flop circuit, based on reciprocal inhibitory interactions between the photoreceptors, releasing melatonin, the dopaminergic amacrine cells, and amacrine cells which contain enkephalin-, neurotensin-, and somatostatin-like immunoreactivity (the ENSL1 amacrine cells). This circuit consists of two loops, one based on the photoreceptors and dopaminergic amacrine cells, and the other on the dopaminergic and ENSLI amacrine cells. In the dark, the photoreceptors and ENSL1 amacrine cells are active, with the dopaminergic amacrine cells inactive. In the light, the dopaminergic amacrine cells are active, with the photoreceptors and ENSLI amacrine cells inactive. The transition from dark to light state occurs over a narrow (<1 log unit) range of low light intensities, and we postulate that this transition is driven by a graded, adapting pathway from photoreceptors, releasing glutamate, to ON-bipolar cells to dopaminergic amacrine cells. The properties of this pathway suggest that, once released from the reciprocal inhibitory controls of the dark state, dopamine release will show graded, adapting characteristics. Thus, we postulate that retinal function will be divided into two phases: a dopamine-independent phase at low light intensities, and a dopamine-dependent phase at higher light intensities. Dopamine-dependent functions may show two-state properties, or two-state properties on which are superimposed graded, adapting characteristics. Functions dependent upon melatonin, the enkephalins, neurotensin, and somatostatin may tend to show simpler two-state properties. We propose that the dark-light switch may have a role in a range of light-adaptive phenomena, in signalling night-day transitions to the suprachiasmatic nucleus and the pineal, and in the control of eye growth during development.

Type
Review 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

Altschuler, R.A., Mosinger, J.L., Hoffman, D.W. & Parakkal, M.H. (1982). Immunocytochemical localization of enkephalin-like immunoreactivity in lhe retina of the guinea pig. Proceedings of the National Academy of Sciences of the U.S.A. 79, 23982400.CrossRefGoogle Scholar
Ariano, M.A., Kang, H.C., Haugland, R.P. & Sibley, D.R. (1991). Multiple fluorescent ligands for dopamine receptors. II. Visualization in neural tissues. Brain Research 547, 208222.Google Scholar
Bartmann, M., Schaeffel, F., Hagel, G. & Zrenner, E. (1994). Constant light affects retinal dopamine levels and blocks deprivation myopia but not lens-induced refractive errors in chickens. Visual Neuroscience 11, 199208.Google Scholar
Besharse, J.C. (1992). The “ON”-bipolar agonist, L-2-amino4-phosphonobutyrate, blocks light-evoked cone contraction in Xenopus eye cups. Neurochemical Research 17, 7580.Google Scholar
Besharse, J.C. & Iuvone, P.M. (1992). Is dopamine a light-adaptive or a dark-adaptive modulator in retina? Neurochemistry International 20, 193199.CrossRefGoogle ScholarPubMed
Besharse, J.C., Iuvone, P.M. & Pierce, M.E. (1988). Regulation of rhythmic photoreceptor metabolism: A role for post-receptoral neurons. Progress in Retinal Research 7, 2161.CrossRefGoogle Scholar
Binkley, S., Hryshchyshyn, M. & Reilly, K. (1979). N-acetyltransferase activity responds to environmental lighting in the eye as well as in the pineal gland. Nature 281, 479481.Google Scholar
Boatright, J.H., Gordon, J.R. & Iuvone, P.M. (1994 a). Inhibition of endogenous dopamine release in amphibian retina by L-2-amino-4-phosphonobutyric acid (L-AP4) and trans-2-aminocyclopentane-1,3-dicarboxylate (ACPD). Brain Research 649, 339342.Google Scholar
Boatrioht, J.H., Rubim, N.M. & Iuvone, P.M. (1994 b). Regulation of endogenous dopamine release in amphibian retina by melatonin: The role of GABA. Visual Neuroscience 11, 10131018.CrossRefGoogle Scholar
Bock, G.R. & Widdows, K. (1990). Myopia and the control of eye growth. CIBA Foundation Symposium 155. Chichester: John Wiley.Google Scholar
Boelen, M.K., Dowton, M. & Chubb, I.W. (1989). The release of [Leu5]-enkephalin-like immunoreactivity from chicken retina is reduced by light in vitro. Brain Research 29, 4348.Google Scholar
Boelen, M.K., Wellard, J., Dowton, M., Chubb, I.W. & Morgan, I.G. (1991). Glycinergic control of [Leu5]-enkephalin levels in chicken retina. Brain Research 23, 221226.CrossRefGoogle Scholar
Boelen, M.K., Dowton, M. & Morgan, I.G. (1993). [Leu5]-enkephalin-like immunoreactive amacrine cells are under nicotinic excitatory control during darkness in chicken retina. Brain Research 624, 137142.Google Scholar
Boelen, M.K., Wellard, J., Dowton, M. & Morgan, I.G. (1994). Endogenous dopamine inhibits the release of enkephalin-like immunoreactivity from amacrine cells of the chicken retina in the light. Brain Research 645, 240246.CrossRefGoogle ScholarPubMed
Brainard, G.C. & Morgan, W.W. (1987). Light-induced stimulation of retinal dopamine: A dose-response relationship. Brain Research 424, 199203.CrossRefGoogle ScholarPubMed
Brann, M.R. & Young, W.S. (1986). Dopamine receptors are located on rods in bovine retina. Neuroscience Letters 69, 221226.Google Scholar
Brecha, N., Karten, H.J. & Laverack, C. (1979). Enkephalin-containing amacrine cells in the avian retina: Immunohistochemical localization. Proceedings of the National Academy of Sciences of the U.S.A. 76, 30103014.CrossRefGoogle ScholarPubMed
Brecha, N., Karten, H.J. & Schenker, C. (1981). Neurotensin-like and somatostatin-like immunoreactivity within amacrine cells of the retina. Neuroscience 6, 13291340.CrossRefGoogle ScholarPubMed
Brunken, W.J., Witkovsky, P. & Karten, H.J. (1986). Retinal neurochemistry of three elasmobranch species: An immunohistochemical approach. Journal of Comparative Neurology 243, 112.Google Scholar
Bubenik, G.A., Brown, G.M., Uhlir, I. & Grota, L.J. (1974). Immunohistological localization of N-acetylindolealkylamines in pineal gland, retina and cerebellum. Brain Research 81, 233242.Google Scholar
Bubenik, G.A., Purtill, R.A., Brown, G.M. & Grota, L.J. (1978). Melatonin in the retina and the Harderian gland. Ontogeny, diurnal variations and melatonin treatment. Experimental Eye Research 27, 323333.Google Scholar
Buckerfield, M., Oliver, J., Chubb, I.W. & Morgan, I.G. (1981). Somatostatin-like immunoreactivity in amacrine cells of the chicken retina. Neuroscience 6, 689695.Google Scholar
Cahill, G.M. & Besharse, J.C. (1991). Resetting the circadian clock in cultured Xenopus eyecups: Regulation of retinal melatonin rhythms by light and D2 dopamine receptors. Journal of Neuroscience 11, 29592971.Google Scholar
Cahill, G.M. & Besharse, J.C. (1992). Light-sensitive melatonin synthesis by Xenopus photoreceptors after destruction of the inner retina. Visual Neuroscience 8, 487490.Google Scholar
Cahill, G.M. & Besharse, J.C. (1993). Circadian clock functions localized in xenopus retinal photoreceptors. Neuron 10, 573577.Google Scholar
Cahill, G.M. & Besharse, J.C. (1995). Circadian rhythmicity in vertebrate retinas: Regulation by a photoreceptor oscillator. Progress in Retinal Research 14, 267291.Google Scholar
Cahill, G.M., Grace, M.S. & Besharse, J.C. (1991). Rhythmic regulation of retinal melatonin: Metabolic pathways, neurochemical mechanisms, and the ocular circadian clock. Cellular and Molecular Neurobiology 11, 529560.CrossRefGoogle ScholarPubMed
Chiba, C. & Saito, T. (1994). APB (2-amino-4-phosphonobutyric acid) activates a chloride conductance in ganglion cells isolated from newt retina. Neuro Report 5, 489492.Google Scholar
Cohen, J., Iuvone, P.M. & Neff, N.H. (1981). Neuroleptic drugs activate tyrosine hydroxylase of retinal amacrine cells. Journal of Pharmacology and Experimental Therapeutics 218, 390394.Google Scholar
Cohen, A.I., Todd, R.D., Harmon, S. & O'Malley, K.L. (1994). Photoreceptors of mouse retinas possess D4 receptors coupled to adenylate cyclase. Proceedings of the National Academy of Sciences of the U.S.A. 89, 1209312097.Google Scholar
Critz, S.D. & Marc, R.E. (1992). Glutamate antagonists that block hyperpolarizing bipolar cells increase the release of dopamine from turtle retina. Visual Neuroscience 9, 271278.Google Scholar
Dearry, A., Falardeau, P., Shores, C. & Caron, M.G. (1991). D2 dopamine receptors in the human retina: Cloning of cDNA and localization of m RNA. Cellular and Molecular Neurobiology 11, 437453.Google Scholar
Denis, P., Elena, P.P., Nordmann, J.P., Saraux, H. & Lapalus, P. (1990). Autoradiographic localization of Dl and D2 dopamine binding sites in the human retina. Neuroscience Letters 14, 8186.Google Scholar
Djamgoz, M.B. & Wagner, H.J. (1992). Localization and function of dopamine in the adult vertebrate retina. Neurochemistry International 20, 139191.Google Scholar
Djamgoz, M.B., Stell, W.K., Chin, C.A. & Lam, D.M. (1981). An opiate system in the goldfish retina. Nature 292, 620623.Google Scholar
Dong, C.J. & McReynolds, J.S. (1991). The relationship between light, dopamine release and horizontal cell coupling in the mudpuppy retina. Journal of Physiology 440, 291309.Google Scholar
Dong, C.J. & McReynolds, J.S. (1992). Comparison of the effects of flickering and steady light on dopamine release and horizontal cell coupling in the mudpuppy retina. Journal of Neurophysiology 67, 364372.Google Scholar
Dowton, M., Boelen, M.K., Morgan, I.G. & Chubb, I.W. (1990). Light inhibits the release of both [Met5]enkephalin and [Met5]enkephalin-containing peptides in chicken retina, but not their syntheses. Neuroscience 38, 187193.CrossRefGoogle Scholar
Dubocovich, M.L. (1983). Melatonin is a potent modulator of dopamine release in the retina. Nature 306, 782784.CrossRefGoogle ScholarPubMed
Dubocovich, M.L. (1988). Role of melatonin in retina. Progress in Retinal Research 8, 129151.Google Scholar
Dubocovich, M.L. & Weiner, N. (1981). Modulation of the stimulation-evoked release of [3H]dopamine in the rabbit retina. Journal of Pharmacology and Experimental Therapeutics 219, 701707.Google Scholar
Dubocovich, M.L. & Weiner, N. (1983). Enkephalins modulate [3H]dopamine release from rabbit retina in vitro. Journal of Pharmacology and Experimental Therapeutics 224, 634639.Google Scholar
Eldred, W.D. & Karten, H.J. (1983). Characterization and quantification of peptidergic amacrine cells in the turtle retina: Enkephalin, neurotensin, and glucagon. Journal of Comparative Neurology 221, 371381.Google Scholar
Glasener, G., Himstedt, W., Weiler, R. & Matute, C. (1988). Putative neurotransmitters in the retinae of three urodele species (Triturus alpestris, Salamandra salamandra, Pleurodeles waltli). Cell and Tissue Research 252, 317328.Google Scholar
Goridis, C. & Virmaux, N. (1974). Light-regulated guanosine 3′,5′-mono-phosphate phosphodiesterase of bovine retina. Nature 248, 5758.Google Scholar
Hadjiconstantinou, M., Cohen, j., Rubenstein, J.S. & Neff, N.H. (1984). An endogenous ligand modulates dopamine-containing neurons of retina via alpha-2 adrenoceptors. Journal of Pharmacology and Experimental Therapeutics 229, 381385.Google ScholarPubMed
Hadjiconstantinou, M., Rossetti, Z., Silvia, C., Krajnc, D. & Neff, N.H. (1988). Aromatic L-amino acid decarboxylase activity of the rat retina is modulated in vivo by environmental light. Journal of Neurochemistry 51, 15601564.Google Scholar
Hamano, K., Katayama-Kumoi, Y., Kiyama, H., Ishimoto, I., Manabe, R. & Tohyama, M. (1989). Coexistence of enkephalin and somatostatin in the chicken retina. Brain Research 489, 254260.Google Scholar
Hamm, H.E. & Menaker, M. (1980). Retinal rhythms in chicks: Circadian variation in melatonin and serotonin N-acetyltransferase activity. Proceedings of the National Academy of Sciences of the U.S.A. 77, 49985002.Google Scholar
Hampson, B.C., Vaney, D.I. & Weiler, R. (1992). Dopaminergic modulation of gap junction permeability between amacrine cells in mammalian retina. Journal of Neuroscience 12, 49114922.Google Scholar
Howells, R.D., Groth, J., Hiller, J.M. & Simon, E.J. (1980). Opiate binding sites in the retina: Properties and distribution. Journal of Pharmacology and Experimental Therapeutics 215, 6064.Google Scholar
Ishimoto, I., Fukuda, M., Kuwayama, Y., Shimizu, Y., Shiosaka, S., Taragi, H., Senba, E., Sakanaka, M., Inagaki, S., Takatsuki, K., Minagawa, H. & Tohyama, M. (1982). Phylogenetical development of somatostatin-containing cells in the retina from teleosts to mammals: Immunohistochemical analysis. Journal fur Hirnforschung 23, 127132.Google Scholar
Ishimoto, I., Millar, T., Chubb, I.W. & Morgan, I.G. (1986). Somatostatin-immunoreactive amacrine cells of chicken retina: Retinal mosaic, ultrastructural features, and light-driven variations in peptide metabolism. Neuroscience 17, 12171233.Google Scholar
Iuvone, P.M. (1986). Evidence for a D2 dopamine receptor in frog retina that decreases cyclic AMP accumulation and serotonin N-acetyltransferase activity. Life Sciences 38, 331342.Google Scholar
Iuvone, P.M. & Besharse, J.C. (1983). Regulation of indoleamine N-acetyltransferase activity in the retina: Effects of light and dark, protein synthesis inhibitors and cyclic nucleotide analogs. Brain Research 273, 111119.Google Scholar
Iuvone, P.M. & Rauch, A.L. (1983). Alpha 2-adrenergic receptors influence tyrosine hydroxylase activity in retinal dopamine neurons. Life Sciences 33, 24552463.Google Scholar
Iuvone, P.M., Galli, C.L., Garrison-Gund, C.K. & Neff, N.H. (1978 a). Light stimulates tyrosine hydroxylase activity and dopamine synthesis in retinal amacrine neurons. Science 202, 901902.Google Scholar
Iuvone, P.M., Galli, C.L. & Neff, N.H. (1978 b). Retinal tyrosine hydroxylase: Comparison of short-term and long-term stimulation by light. Molecular Pharmacology 14, 12121219.Google Scholar
Iuvone, P.M., Joh, T.H. & Neff, N.H. (1979). Regulation of retinal tyrosine hydroxylase: Long-term exposure to light increased the apparent V max without a concomitant increase of immunotitratable enzyme molecules. Brain Research 178, 191195.Google Scholar
Iuvone, P.M., Rauch, A.L., Marshburn, P.B., Glass, D.B. & Neff, N.H. (1982). Activation of retinal tyrosine hydroxylase in vitro by cyclic AMP-dependent protein kinase: Characterization and comparison to activation in vivo by photic stimulation. Journal of Neurochemistry 39, 16321640.Google Scholar
Iuvone, P.M., Boatright, J.H. & Bloom, M.M. (1987). Dopamine mediates the light-evoked suppression of serotonin N-acetyltransferase activity in retina. Brain Research 418, 314324.CrossRefGoogle ScholarPubMed
Iuvone, P.M., Tigges, M., Fernandes, A. & Tigges, J. (1989). Dopamine synthesis and metabolism in rhesus monkey retina: Development, aging, and the effects of monocular visual deprivation. Visual Neuroscience 2, 465471.CrossRefGoogle ScholarPubMed
Iuvone, P.M., Avendano, G., Butler, B.J. & Adler, R. (1990). Cyclic AMP-dependent induction of serotonin N-acetyltransferase activity in photoreceptor-enriched chick retinal cell cultures: Characterization and inhibition by dopamine. Journal of Neurochemistry 55, 673682.Google Scholar
Kamp, C.W. & Morgan, W.W. (1983). Effects of excitatory amino acids on dopamine synthesis in the rat retina. European Journal of Pharmacology 92, 139142.Google Scholar
Kato, S., Teranishi, T., Kuo, C.H. & Negishi, K. (1982). 5-hydroxy-tryptamine stimulates [3H] dopamine release from the fish retina. Journal of Neurochemistry 39, 493498.Google Scholar
Kirsch, M. & Wagner, H.J. (1989). Release pattern of endogenous dopamine in teleost retinae during light adaptation and pharmacological stimulation. Vision Research 29, 147154.Google Scholar
Kolbinger, W. & Weiler, R. (1993). Modulation of endogenous dopamine release in the turtle retina: Effects of light, calcium, and neuro-transmitters. Visual Neuroscience 10, 10351041.Google Scholar
Kolbinger, W., Kohler, K., Oetting, H. & Weiler, R. (1990). Endogenous dopamine and cyclic events in the fish retina, I: HPLC assay of total content, release, and metabolic turnover during different light/dark cycles. Visual Neuroscience 5, 143149.Google Scholar
Kramer, S.G. (1971). Dopamine: A retinal neurotransmitter. I. Retinal uptake, storage, and light-stimulated release of HS-dopamine in vivo. Investigative Ophthalmology and Visual Science 10, 438452.Google Scholar
Krizaj, D., Akopian, A., & Witkovsky, P. (1994). The effects of L-glutamate, AMPA, quisqualate, and kainate on retinal horizontal cells depend on adaptational state: Implications for rod-cone interactions. Journal of Neuroscience 14, 56615671.Google Scholar
Laufer, M., Negishi, K. & Drujan, B.D. (1981). Pharmacological manipulation of spatial properties of S-potentials. Vision Research 21, 16571660.Google Scholar
Li, H.B., Watt, C.B. & Lam, D.M. (1990). Double-label analyses of somatostatin's coexistence with enkephalin and gamma-aminobutyric acid in amacrine cells of the chicken retina. Brain Research 525, 304309.Google Scholar
Li, X.X., Schaeffel, F., Kohler, K. & Zrenner, E. (1992). Dose-dependent effects of 6-hydroxy dopamine on deprivation myopia, electroretinograms, and dopaminergic amacrine cells in chickens. Visual Neuroscience 9, 483492.Google Scholar
Lin, Z.S. & Yazulla, S. (1994 a). Depletion of retinal dopamine increases brightness perception in goldfish. Visual Neuroscience 11, 683693.Google Scholar
Lin, Z.S. & Yazulla, S. (1994 b). Depletion of retinal dopamine does not affect the ERG b–wave increment threshold function in goldfish in vivo. Visual Neuroscience 11, 695702.Google Scholar
Maguire, G. & Hamasaki, D.I. (1994). The retinal dopamine network alters the adaptational properties of retinal ganglion cells in the cat. Journal of Neurophysiology 72, 730741.Google Scholar
Mangel, S.C. & Dowling, J.E. (1985). Responsiveness and receptive field size of carp horizontal cells are reduced by prolonged darkness and dopamine. Science 229, 11071109.Google Scholar
McCormack, C.A. & Burnside, B. (1993). Light and circadian modulation of teleost retinal tyrosine hydroxylase activity. Investigative Ophthalmology and Visual Science 34, 18531860.Google Scholar
Megaw, P., Cluff, M.P., Morgan, I.G. & Boelen, M.K. (1994). Deprivation of form vision reduces, and temporal contrast restores the levels of [leu]enkephalin in chicken retina. Proceedings of the Australian Neuroscience Society 5, 200.Google Scholar
Millar, T.J., Salipan, N., Oliver, J.O., Morgan, I.G. & Chubb, I.W. (1984). The concentration of enkephalin-like material in the chick retina is light dependent. Neuroscience 13, 221226.Google Scholar
Morgan, W.W. & Kamp, C.W. (1980). A GABAergic influence on the light-induced increase in dopamine turnover in the dark-adapted rat retina in vivo. Journal of Neurochemistry 34, 10821086.Google Scholar
Morgan, W.W. & Kamp, C.W. (1983). Effect of strychnine and of bicuculline on dopamine synthesis in retinas of dark-maintained rats. Brain Research 278, 362365.Google Scholar
Morgan, I.G., Wellard, J.W. & Boelen, M.K. (1994). A role for the enkephalin-immunoreactive amacrine cells of the chicken retina in adaptation to light and dark. Neuroscience Letters 174, 6466.Google Scholar
Morgan, I.G., Boelen, M.K. & Miethke, P. (1995 a). Pineal activity is under the control of retinal D1-dopaminergic pathways. Neuro-Report 6, 446448.Google Scholar
Morgan, I.G., Boelen, M.K. & Miethke, P. (1995 b). Parallel suppression of retinal and pineal melatonin synthesis by retinally mediated light. NeuroReport 6, 15301532.Google Scholar
Muresan, Z. & Besharse, J.C. (1993). D2-like dopamine receptors in amphibian retina: Localization with fluorescent ligands. Journal of Comparative Neurology 331, 149160.Google Scholar
Myhr, K.L., Dong, C.J. & McReynolds, J.S. (1994). Cones contribute to light-evoked, dopamine-mediated uncoupling of horizontal cells in the mudpuppy retina. Journal of Neurophysiology 72, 5662.Google Scholar
Naarendorp, F., Hitchcock, P.F. & Sieving, P.A. (1993). Dopaminergic modulation of rod pathway signals does not affect the scotopic ERG of cat at dark-adapted threshold. Journal of Neurophysiology 10, 16811691.CrossRefGoogle Scholar
Osborne, H.H. & Herz, A. (1983). Opiate binding sites in bovine retina: Evidence for benzomorphan selective binding sites. European Journal of Pharmacology 86, 373378.Google Scholar
Pierce, M.E. & Besharse, J.C. (1987). Melatonin and rhythmic photo-receptor metabolism: Melatonin-induced cone elongation is blocked at high light intensity. Brain Research 405, 400404.Google Scholar
Pugh, E.N. & Lamb, T.D. (1990). Cyclic GMP and calcium: The internal messengers of excitation and adaptation in vertebrate photoreceptors. Vision Research 30, 19231948.Google Scholar
Rohrer, B., Spira, A.W. & Stell, W.K. (1993). Apomorphine blocks form-deprivation myopia in chickens by a dopamine D2-receptor mechanism acting in retina or pigmented epithelium. Visual Neuroscience 10, 447453.Google Scholar
Rohrer, B., Iuvone, P.M. & Stell, W.K. (1995). Stimulation of dopa-minergic amacrine cells by stroboscopie illumination or fibroblast growth factor (bFGF, FGF-2) injections: Possible roles in prevention of form-deprivation myopia in the chick. Brain Research 686, 169181.Google Scholar
Sait, M.L., Shirley, G.M., Boelen, M.G., Morgan, I.G. & Boelen, M.K. (1994). Excitatory amino acid-control of the ENSLl amacrine cells in chicken retina. Proceedings of the Australian Neuroscience Society 5, 201.Google Scholar
Schorderet, M. & Nowak, J.Z. (1990). Retinal dopamine D1 and D2 receptors: Characterization by binding or pharmacological studies and physiological functions. Cellular and Molecular Neurobiology 10, 303325.Google Scholar
Seltner, R. P.L., Grant, V. & Stell, W.K. (1994). [Met5]enkephalin and form deprivation myopia. Investigative Ophthalmology and Visual Science 35, 2069.Google Scholar
Slaughter, M.M. & Maler, R.F. (1981). 2-amino-4-phosphonobutyric acid: A new pharmacological tool for retina research. Science 211, 182185.Google Scholar
Slaughter, M.M., Mattler, J.A. & Gottlieb, D.I. (1985). Opiate binding sites in the chick, rabbit and goldfish retina. Brain Research 339, 3947.Google Scholar
Stone, R.A., Lin, T., Laties, A.M. & Iuvone, P.M. (1989). Retinal dopamine and form-deprivation myopia. Proceedings of the National Academy of Sciences of the U.S.A. 86, 704706.Google Scholar
Su, Y.Y. & Watt, C.B. (1987). Interaction between enkephalin and dopamine in the avian retina. Brain Research 423, 6370.Google Scholar
Takahashi, K. & Copenhagen, D.R. (1992). APB suppresses synaptic input to retinal horizontal cells modulated by intracellular pH. Journal of Neurophysiology 67, 16331642.CrossRefGoogle ScholarPubMed
Teranishi, T., Negishi, K. & Kato, S. (1983). Dopamine modulates S-potential amplitude and dye-coupling between external horizontal cells in carp retina. Nature 301, 243246.Google Scholar
Thomas, K.B., Tigges, M. & Iuvone, P.M. (1993). Melatonin synthesis and circadian tryptophan hydroxylase activity in chicken retina following destruction of serotonin immunoreactive amacrine and bipolar cells by kainic acid. Brain Research 601, 303307.Google Scholar
Tornqvist, K., Uddman, R., Sundler, F. & Ehinger, B. (1982). Soma-tostatin and VIP neurons in the retina of different species. Histochemistry 76, 137152.Google Scholar
Tran, V.T. & Dickman, M. (1992). Differential localization of dopamine D1 and D2 receptors in rat retina. Investigative Ophthalmology and Visual Science 33, 16201626.Google ScholarPubMed
Umino, O. & Dowling, J.E. (1991). Dopamine release from interplexi-form cells in the retina: Effects of GnRH, FMRFamide, bicuculline, and enkephalin on horizontal cell activity. Journal of Neuroscience 11, 30343046.Google Scholar
Umino, O., Lee, Y. & Dowling, J.E. (1991). Effects of light stimuli on the release of dopamine from interplexiform cells in the white perch retina. Visual Neuroscience 7, 451458.Google Scholar
Vuvan, T., Geffard, M., Denis, P., Simon, A. & Nguyen-Legros, J. (1993). Radioimmunoligand characterization and immunohistochemical localization of dopamine D2 receptors on rods in the rat retina. Brain Research 614, 5764.Google Scholar
Wagner, H.J. & Behrens, U.D. (1993). Microanatomy of the dopa-minergic system in the rainbow trout retina. Vision Research 33, 13451358.Google Scholar
Wagner, H.J. & Zeutzius, I. (1987). Amacrine cells with neurotensin- and somatostatin-like immunoreactivities in three species of teleosts with different color vision. Cell and Tissue Research 248, 663673.Google Scholar
Wagner, H.J., Luo, B.C., Ariano, M.A., Sibley, D.R. & Stell, W.K. (1993). Localization of D2 dopamine receptors in vertebrate retinae with anti-peptide antibodies. Journal of Comparative Neurology 331, 469481.Google Scholar
Wallman, J. (1990). Retinal control of eye growth and refraction. Progress in Retinal Research 12, 133153.Google Scholar
Watt, C.B. & Florack, V.J. (1994). A triple-label analysis demonstrating that enkephalin-, somatostatin- and neurotensin-like immunoreactivities are expressed by a single population of amacrine cells in the chicken retina. Brain Research 634, 310316.Google Scholar
Watt, C.B., Glazebrook, P.A. & Li, H.B. (1991 a). Coexistence of somatostatin and neurotensin in amacrine cells of the chicken retina. Brain Research 546, 166170.Google Scholar
Watt, C.B., Florack, V.J. & Lam, D.M. (1991 b). A double-label analysis demonstrating that all enkephalin-immunoreactive amacrine cells in the chicken retina express neurotensin immunoreactivity. Brain Research 566, 337341.Google Scholar
Weiler, R. & Ball, A.K. (1989). Enkephalinergic modulation of the dopamine system in the turtle retina. Visual Neuroscience 3, 455461.Google Scholar
Weiler, R., Kohler, K., Kirsch, M. & Wagner, H.J. (1988). Glutamate and dopamine modulate synaptic plasticity in horizontal cell dendrites of fish retina. Neuroscience Letters 87, 205209.Google Scholar
Weiss, S. & Schaeffel, F. (1993). Diurnal growth rhythms in the chicken eye: Relation to myopia development and retinal dopamine levels. Journal of Comparative Physiology A 172, 263270.Google Scholar
Wiechmann, A.F. & Craft, C.M. (1993). Localization of mRNA encoding the indolamine synthesizing enzyme, hydroxyindole-O-methyl-transferase, in chicken pineal gland and retina by in situ hybridization. Neuroscience Letters 150, 207211.Google Scholar
Wirz-Justice, A., Daprada, M. & Reme, C. (1984). Circadian rhythm in rat retinal dopamine. Neuroscience Letters 45, 2125.Google Scholar
Witkovsky, P. & Dearry, A. (1992). Functional roles of dopamine in the vertebrate retina. Progress in Retinal Research 11, 247292.Google Scholar
Witkovsky, P., Stone, S. & Besharse, J.C. (1988). Dopamine modifies the balance of rod and cone inputs to horizontal cells of the Xenopus retina. Brain Research 449, 332336.Google Scholar
Witkovsky, P., Nicholson, C., Rice, M.E., Bohmaker, K. & Meller, E. (1993). Extracellular dopamine concentration in the retina of the clawed frog, Xenopus laevis. Proceedings of the National Academy of Sciences of the U.S.A. 90, 56675671.Google Scholar
Wulle, I., Kirsch, M. & Wagner, H.J. (1990). Cyclic changes in dopamine and DOPAC content, and tyrosine hydroxylase activity in the retina of a cichlid fish. Brain Research 515, 163167.Google Scholar
Yang, X.L., Tornqvist, K. & Dowling, J.E. (1988). Modulation of cone horizontal cell activity in the teleost fish retina. I. Effects of prolonged darkness and background illumination on light responsiveness. Journal of Neuroscience 8, 22592268.Google Scholar
Zawilska, J.B. & Iuvone, P.M. (1992). Melatonin synthesis in chicken retina: Effect of kainic acid-induced lesions on the diurnal rhythm and D2-dopamine receptor-mediated regulation of serotonin N-acetyl-transferase activity. Neuroscience Letters 135, 7174.Google Scholar
Zawilska, J. & Nowak, J.Z. (1991). Regulation of melatonin biosynthesis in vertebrate retina: Involvement of dopamine in the suppressive effects of light. Folio Histochemica et Cytobiologica 29, 313.Google Scholar
Zawilska, J.B. & Nowak, J.Z. (1994). Does D4 dopamine receptor mediate the inhibitory effect of light on melatonin biosynthesis in chick retina? Neuroscience Letters 166, 203206.Google Scholar
Zawilska, J.B., Derbiszewska, T. & Nowak, J.Z. (1994). Clozapine and other neuroleptic drugs antagonize the light-evoked suppression of melatonin biosynthesis in chick retina: Involvement of the D4-like dopamine receptor. Journal of Neural Transmission 97, 107117.Google Scholar