• Nem Talált Eredményt

Due to the wealth and the variety of the observed gap junc-tions,Vaney (1994)suggested that the vertebrate retina should be viewed as a stack of planar neuronal arrays with vertical links. In that context, the present review detailed how each retinal cell type forms electrically coupled planar mosaics, how coupled ar-rays are interlaced with each other and how the electrically cou-pled neuronal networks act upon vertical signal processing throughout the retina. This article has also summarized results from various vertebrates, and has aimed to reveal evolutionary trends. However, a comprehensive phylogenetic comparison was not possible, since the available dataset was rather incomplete:(i) the early electrophysiological studies were primarily carried out in lower vertebrates (due to technical difficulties); (ii) recent ad-vances in molecular biology provided the opportunity to identify connexins but mainly popular mammalian models were studied;

(iii) in this dataset, lower vertebrates are represented by only a few model species. Nevertheless, we can still point out a number of general features and trends followed by most examined verte-brate species.

First, the cell types that contact each other via electrical syn-apses are conserved throughout vertebrate evolution. Thus, rod-to-rod, rod-to-cone, cone-to-cone, horizontal-to-horizontal, bipolar-to-bipolar, amacrine-to-amacrine, amacrine-to-ganglion and ganglion-to-ganglion cell contacts are ubiquitous across all verte-brate retinas. The only (known) exception is the mammalian pri-mary rod pathway that incorporates two gap junctions (AIIeAII and AIIeON cone bipolar cell), which are consequently uniquely mammalian features. Apart from the latter AII-bipolar cell contact, no other amacrine-to-bipolar cell gap junction has been discovered in vertebrates, not even in mammals.

Second, gap junctions located in particular retinal circuits play largely the same roles in retinal information processing in all examined vertebrates, again, with the sole exception of the mam-malian specific rod pathway. Gap junctions in this latter pathway process feed forward signaling, which is a unique feature in the retina and very rare within the entire central nervous system. Be-sides mammalian rod pathways, however, the functions of the rest of the retinal gap junctions are peculiarly similar in all vertebrates. It is clear, for example that all vertebrate horizontal cells utilize con-nexin subunits (Cx50, Cx57 or their orthologs) that possess high unitary conductance and/or are capable of forming extensive pla-ques to provide extremely high transmembrane conductance (in the nSemS range) for intercellular signaling. In contrast, other retinal cell types form gap junctions with moderate (1000e2000 pS) or low conductances (<800 pS), including most neurons in the inner retina and photoreceptors in the outer retina, respectively. These latter contacts are composed of Cx36, Cx45 subunits, some of which show about the lowest unitary conductances. This obvious difference in gap junctional conductance between horizontal cells and other retinal cell types corresponds nicely with the observed difference in the diameters of tracer-coupled arrays of these reti-nal neurons. Whereas horizontal cells display extended syncytia in the outer retina, most inner retinal neurons and photoreceptors form rather constrained tracer-coupled arrays restricted to direct neighbors. One exception from this tendency though is the exten-ded coupling of so-called wide-field amacrine cells in the inner retina of all examined species. These cells often form extended lattices involving up to a few hundred amacrine cells in one tracer coupled array. However, connexin makeup and conductance levels of the underlying gap junctions are yet to be determined. Generally speaking, apart from hard wiring provided by the fast synaptic transmitters and soft wiring contributed primarily by aminergic, NOergic and peptidergic systems, electrical coupling may subserve

as the skeleton of lateral processing, determining important func-tions such as signal averaging and synchronization.

Third, the modulation of gap junctions formed by retinal neurons follows the same patterns, including the dopamine/D1 mediated reduction or D2mediated increase in gap junction conductance. In addition, some contacts, including those between horizontal cells, are modulated by changes in extracellular NO concentrations as well. Since the production and release of both dopamine and NO are under the control of light adaptation levels and/or circadian rhyth-micity, the conductance of various gap junctions are dynamically modulated. Apart from the effects of the ambient background light level, conductance of most gap junctions is sensitive to changes in pH and/or the intracellular concentration of Caþþ. Therefore, the functioning of gap junctions is controlled by the changes in the environment as well as the internal condition of the living system.

Fourth, the observed species’ differences in the connexin makeup of gap junctions are also minor. The highest variety is found among horizontal-to-horizontal cell contacts of vertebrates but this diversity is largely due to the numerous ortholog connexins used in horizontal cell lattices of various species. It is important to point out that one canfind many different connexins in the inner retina as well (Cx36, Cx45, Cx30.2 etc.), although this variety is likely due to the diversity of inner retinal cell types rather than divergent evolution. In fact, we found that the genetical distances among connexin paralogs are larger than those of the orthologs expressed in the same circuit of various vertebrates (Fig. 1). This is in line with thefinding ofEastman et al. (2006)who detected high sequence similarity of even the most variable C-terminal tail of the

>400 myr separated zebrafish and mammalian ortholog Cx se-quences, even though many paralogs were not close relatives.

Based on theirfindings, they suggested that a stronger selection occurred for C-terminus function in orthologs, as opposed to a relaxed selection on one of the paralogs following gene duplica-tions. Such relaxed selection likely allowed divergence of connexins and facilitated novel gene evolution. The close homology of orthologs but not paralogs supports the view that the function determines the utilized connexin (Cruciani and Mikalsen, 2006) and not the phylogenetic group (DeBoer and Van der Heyden, 2005). In fact, connexin orthologs display very similar biophysical and physiological properties; in consequence, the fact that they are utilized in similar contacts further reinforces the equivalent roles they play in the vision of various vertebrates. In summary, all reviewed aspects of vertebrate retinal gap junctions appear remarkably conservative. This implies that the architecture of the electrically coupled neuronal networks in the retina is as old as the retinal circuitry itself. It seems fair to conclude, therefore, that gap junctions have been an integral component of retinal circuitry from the inception of vertebrate eye formation.

Acknowledgments

This study was supported by SROP-4.2.2/B-10/1-2010-0029 and SROP-4.2.1.B-10/2/KONV-2010-0002, and the Hungarian Academy by awarding OTKA K100144 to R.G and OTKA K105247 and the János Bolyai Fellowship to B.V. Authors are indebted to Dr. Paul Witkovsky with his style and grammar corrections and for pro-fessional advices.

References

Aboelela, S.W., Robinson, D.W., 2004. Physiological response properties of displaced amacrine cells of the adult ferret retina. Vis. Neurosci. 21, 135e144.

Ackert, J.M., Wu, S.H., Lee, J.C., Abrams, J., Hu, E.H., Perlman, I., Bloomfield, S.A., 2006. Light-induced changes in spike synchronization between coupled ON direction selective ganglion cells in the mammalian retina. J. Neurosci. 26, 4206e4215.

Al-Ubaidi, M.R., White, T.W., Ripps, H., Poras, I., Avner, P., Gome`s, D., Bruzzone, R., 2000. Functional properties, developmental regulation, and chromosomal localization of murine connexin36, a gap-junctional protein expressed prefer-entially in retina and brain. J. Neurosci. Res. 59, 813e826.

Arai, I., Tanaka, M., Tachibana, M., 2010. Active roles of electrically coupled bipolar cell network in the adult retina. J. Neurosci. 30, 9260e9270.

Baldridge, W.H., Ball, A.K., 1991. Background illumination reduces horizontal cell receptive-field size in both normal and 6-hydroxydopamine-lesioned goldfish retinas. Vis. Neurosci. 7, 441e450.

Baylor, D.A., Fuortes, M.G.F., O’Bryan, P.M., 1971. Receptivefields of single cones in the retina of the turtle. J. Physiol. (Lond.) 24, 265e294.

Becker, D.L., Bonness, V., Catsicas, M., Mobbs, P., 2002. Changing patterns of gan-glion cell coupling and connexin expression during chick retinal development.

J. Neurobiol. 52, 280e293.

Bennett, M.V.L., Verselis, V.K., 1992. Biophysics of gap junctions. Semin. Cell. Biol. 3, 29e47.

Bennett, M.V.L., Barrio, L.C., Bargiello, T.A., Spray, D.C., Hertzberg, E., Saez, J.C., 1991.

Gap junctions: new tools, new answers, new questions. Neuron 6, 305e320.

Beyer, E.C., Lipkind, G.M., Kyle, J.W., Berthoud, V.M., 2012. Structural organization of intercellular channels II. Amino terminal domain of the connexins: sequence, functional roles, and structure. Biocim Biophys. Acta 1818, 1823e1830.

Bloomfield, S.A., 1992. Relationship between receptive and dendriticfield size of amacrine cells in the rabbit retina. J. Neurophysiol. 68, 711e725.

Bloomfield, S.A., Völgyi, B., 2004. Function and plasticity of homologous coupling between AII amacrine cells. Vision Res. 44, 3297e3306.

Bloomfield, S.A., Völgyi, B., 2007. Response properties of a unique subtype of wide-field amacrine cell in the rabbit retina. Vis. Neurosci. 24, 459e469.

Bloomfield, S.A., Völgyi, B., 2009. The diverse functional roles and regulation of neuronal gap junctions in the retina. Nat. Rev. Neurosci. 10, 495e506.

Bloomfield, S.A., Xin, D., Persky, S.E., 1995. A comparison of receptivefield and tracer coupling size of horizontal cells in the rabbit retina. Vis. Neurosci. 12, 985e999.

Brightman, M.W., Reese, T.S., 1969. Junctions between intimately apposed cell membranes in the vertebrate brain. J. Cell. Biol. 40, 648e677.

Brivanlou, I.H., Warland, D.K., Meister, M., 1998. Mechanisms of concertedfiring among retinal ganglion cells. Neuron 20, 527e539.

Bugnon, O., Schaad, N.C., Schorderet, M., 1994. Nitric oxide modulates endogenous dopamine release in bovine retina. Neuroreport 5, 401e404.

Cha, J., Kim, H.-L., Pan, F., Chun, M.-H., Massey, S.C., Kim, I.-B., 2012. Variety of horizontal cell gap junctions in the rabbit retina. Neurosci. Lett. 510, 99e103.

Chesler, M., 2003. Regulation and modulation of pH in the brain. Physiol. Rev. 83, 1183e1221.

Chun, M.H., Han, S.H., Chung, J.W., Wässle, H., 1993. Electron microscopic analysis of the rod pathway of the rat retina. J. Comp. Neurol. 332, 421e432.

Church, J., Baimbridge, K.G., 1991. Exposure to high-pH medium increases the incidence and extent of dye coupling between rat hippocampal CA1 pyramidal neurons in vitro. J. Neurosci. 11, 3289e3295.

Ciolofan, C., Lynn, B.D., Wellershaus, K., Willecke, K., Nagy, J.I., 2007. Spatial re-lationships of connexin36, connexin57 and zonula occludens-1 in the outer plexiform layer of mouse retina. Neuroscience 148, 473e488.

Cobbett, P., Hatton, G.I., 1984. Dye coupling in hypothalamic slices: dependence on in vivo hydration state and osmolality of incubation medium. J. Neurosci. 4, 3034e3038.

Cohen, E., Sterling, P., 1990. Convergence and divergence of cones onto bipolar cells in the central area of cat retina. Philos. Trans. R. Soc. Lond B Biol. Sci. 330, 323e 328.

Connaughton, V.P., Nelson, R., 2010. Spectral responses in zebrafish horizontal cells include a tetraphasic response and a novel UV-dominated triphasic response.

J. Neurophysiol. 104, 2407e2422.

Cook, J.E., Becker, D.L., 1995. Gap junctions in the vertebrate retina. Microsc. Res.

Tech. 31, 408e419.

Cook, J.E., Becker, D.L., 2009. Gap-junction proteins in retinal development: new roles for the“nexus”. Physiology (Bethesda) 24, 219e230.

Cooper, N.G., McLaughlin, B.J., 1981. Gap junctions in the outer plexiform layer of the chick retina: thin section and freeze-fracture studies. J. Neurocytol. 10, 515e 529.

Copenhagen, D.R., Owen, W.G., 1976. Coupling between rod photoreceptors in a vertebrate retina. Nature 260, 57e59.

Cruciani, V., Mikalsen, S.-O., 2006. The vertebrate connexin family. Cell. Mol. Life Sci.

63, 1125e1140.

Cruciani, V., Mikalsen, S.O., 2007. Evolutionary selection pressure and family re-lationships among connexin genes. Biol. Chem. 388, 253e264.

Dacheux, R.F., Raviola, E., 1982. Horizontal cells in the retina of the rabbit.

J. Neurosci. 2, 1486e1493.

Dacheux, R.F., Raviola, E., 1986. The rod pathway in the rabbit retina: a depolarizing bipolar and amacrine cell. J. Neurosci. 6, 331e345.

Deans, M.R., Völgyi, B., Goodenough, D.A., Bloomfield, S.A., Paul, D.L., 2002. Con-nexin36 is essential for transmission of rod-mediated visual signals in the mammalian retina. Neuron 36, 703e712.

DeBoer, T.P., Van der Heyden, M.A.G., 2005. Xenopus connexins: how frogs bridge the gap. Differentiation 73, 330e340.

Dedek, K., Schultz, K., Pieper, M., Dirks, P., Maxeiner, S., Willecke, K., Weiler, R., Janssen-Bienhold, U., 2006. Localization of heterotypic gap junctions composed of connexin45 and connexin36 in the rod pathway of the mouse retina. Eur. J.

Neurosci. 24, 1675e1686.

Dedek, K., Breuninger, T., de Sevilla Müller, L.P., Maxeiner, S., Schultz, K., Janssen-Bienhold, U., Willecke, K., Euler, T., Weiler, R., 2009. A novel type of inter-plexiform amacrine cell in the mouse retina. Eur. J. Neurosci. 30, 217e228.

Dehal, P., Boore, J.L., 2005. Two rounds of whole genome duplication in the ancestral vertebrate. Plos Biol. 3, e314. 1700e1708.

Demb, J.B., Pugh, E.N., 2002. Connexin36 forms synapses essential for night vision.

Neuron 36, 551e553.

Dermietzel, R., Kremer, M., Paputsoglu, G., Stang, A., Skerrett, I.M., Gomes, D., Srinivas, M., Janssen-Bienhold, U., Weiler, R., Nicholson, B.J., Bruzzone, R., Spray, D.C., 2000. Molecular and functional diversity of neural connexins in the retina. J. Neurosci. 20, 8331e8343.

Detwiler, P.B., Hodgkin, A.L., 1979. Electrical coupling between cones in turtle retina.

J. Physiol. 291, 75e100.

DeVries, S.H., 1999. Correlatedfiring in rabbit retinal ganglion cells. J. Neurophysiol.

81, 908e920.

DeVries, S.H., Schwartz, E.A., 1989. Modulation of an electrical synapse between solitary pairs of catfish horizontal cells by dopamine and second messengers.

J. Physiol. 414, 351e375.

DeVries, S.H., Qi, X., Smith, R., Makous, W., Sterling, P., 2002. Electrical coupling between mammalian cones. Curr. Biol. 12, 1900e1907.

Doolittle, R.F., Anderson, K.L., Feng, D.F., 1989. Estimating the prokaryoticeeukary-otic divergence time from protein sequences. In: The Hierarchy of Life. Excerpta Medica, New York, pp. 73e85.

Eastman, S.D., Chen, T.H.-P., Falk, M.M., Mendelson, T.C., Iovine, M.K., 2006. Phylo-genetic analysis of three complete gap junction gene families reveals lineage-specific duplications and highly supported gene classes. Genomics 87, 265e274.

Fain, G.L., Gold, G.H., Dowling, J.E., 1976. Receptor coupling in the toad retina. Cold Spring Harbor Symp. Quant. Biol. 60, 547e561.

Farquhar, M.G., Palade, G.E., 1965. Cell junctions in amphibian skin. J. Cell. Biol. 26, 263e291.

Feigenspan, A., Teubner, B., Willecke, K., Weiler, R., 2001. Expression of neuronal connexin36 in AII amacrine cells of the mammalian retina. J. Neurosci. 21, 230e 239.

Feigenspan, A., Janssen-Bienhold, U., Hormuzdi, S., Monyer, H., Degen, J., Söhl, G., Willecke, K., Ammermüller, J., Weiler, R., 2004. Expression of connexin36 in cone pedicles and OFF-cone bipolar cells of the mouse retina. J. Neurosci. 24, 3325e3334.

Firsov, M.L., Green, D.G., 1998. Photoreceptor coupling in turtle retina. Vis. Neurosci.

15, 755e764.

Furshpan, E.J., Potter, D.D., 1957. Mechanism of nerve-impulse transmission at a crayfish synapse. Nature 180, 342e343.

Fyk-Kolodziej, B., Qin, P., Pourcho, R.G., 2003. Identification of a cone bipolar cell in cat retina which has input from both rod and cone photoreceptors. J. Comp.

Neurol. 464, 104e113.

Galarreta, M., Hestrin, S., 2002. Electrical and chemical synapses among parvalbu-min fast-spiking GABAergic interneurons in adult mouse neocortex. Proc. Natl.

Acad. Sci. U. S. A. 99, 12438e12443.

Ghai, K., Zelinka, C., Fischer, A.J., 2009. Serotonin released from amacrine neurons is scavenged and degraded in bipolar neurons in the retina. J. Neurochem. 111, 1e 14.

Gibson, J.R., Beierlein, M., Connors, B.W., 1999. Two networks of electrically coupled inhibitory neurons in neocortex. Nature 402, 75e79.

Gold, G.H., Dowling, J.E., 1979. Photoreceptor coupling in retina of the toad, Bufo marinus. I. Anatomy. J. Neurophysiol. 42, 292e310.

Goodenough, D.A., Revel, J.P., 1970. Afine structural analysis of intercellular junc-tions in the mouse liver. J. Cell. Biol. 45, 272e290.

Greschner, M., Shlens, J., Bakolitsa, C., Field, G.D., Gauthier, J.L., Jepson, L.H., Sher, A., Litke, A.M., Chichilnisky, E.J., 2011. Correlatedfiring among major ganglion cell types in primate retina. J. Physiol. 589, 75e86.

Güldenagel, M., Ammermüller, J., Feigenspan, A., Teubner, B., Degen, J., Söhl, G., Willecke, K., Weiler, R., 2001. Visual transmission deficits in mice with targeted disruption of the gap junction gene connexin36. J. Neurosci. 21, 6036e6044.

Hack, I., Peichl, L., Brandstätter, J.H., 1999. An alternative pathway for rod signals in the rodent retina: rod photoreceptors, cone bipolar cells, and the localization of glutamate receptors. Proc. Natl. Acad. Sci. U. S. A. 96, 14130e14135.

Hack, I., Frech, M., Dick, O., Peichl, L., Brandstätter, J.H., 2001. Heterogeneous dis-tribution of AMPA glutamate receptor subunits at the photoreceptor synapses of rodent retina. Eur. J. Neurosci. 13, 15e24.

Hampson, E.C., Vaney, D.I., Weiler, R., 1992. Dopaminergic modulation of gap junction permeability between amacrine cells in mammalian retina. J. Neurosci.

12, 4911e4922.

Han, Y., Massey, S.C., 2005. Electrical synapses in retinal ON cone bipolar cells:

subtype-specific expression of connexins. Proc. Natl. Acad. Sci. U. S. A. 102, 13313e13318.

Hansen, K.A., Torborg, C.L., Elstrott, J., Feller, M.B., 2005. Expression and function of the neuronal gap junction protein connexin 36 in developing mammalian ret-ina. J. Comp. Neurol. 493, 309e320.

Harris, A.L., 2001. Emerging issues of connexin channels: biophysicsfills the gap.

Q. Rev. Biophys. 34, 325e472.

Hidaka, S., Maehara, M., Umino, O., Lu, Y., Hashimoto, Y., 1993. Lateral gap junction connections between retinal amacrine cells summating sustained responses.

Neuroreport 5, 29e32.

Hidaka, S., Kato, T., Miyachi, E., 2002. Expression of gap junction connexin36 in adult rat retinal ganglion cells. J. Integr. Neurosci. 1, 3e22.

Hidaka, S., Akahori, Y., Kurosawa, Y., 2004. Dendrodendritic electrical synapses between mammalian retinal ganglion cells. J. Neurosci. 24, 10553e10567.

Hidaka, S., Kato, T., Hashimoto, Y., 2005. Structural and functional properties of homologous electrical synapses between retinal amacrine cells. J. Integr. Neu-rosci. 4, 313e340.

Hidaka, S., 2008. Intracellular cyclic-AMP suppresses the permeability of gap junctions between retinal amacrine cells. J. Integr. Neurosci. 7, 29e48.

Hidaka, S., 2012. Suppression of electrical synapses between retinal amacrine cells of goldfish by intracellular cyclic-AMP. Brain Res. 1449, 1e14.

Hilgen, G., von Maltzahn, J., Willecke, K., Weiler, R., Dedek, K., 2011. Subcellular distribution of connexin45 in OFF bipolar cells of the mouse retina. J. Comp.

Neurol. 519, 433e450.

Hitchcock, P.F., 1993. Mature, growing ganglion cells acquire new synapses in the retina of the goldfish. Vis. Neurosci. 10, 219e224.

Hornstein, E.P., Verweij, J., Schnapf, J.L., 2004. Electrical coupling between red and green cones in primate retina. Nat. Neurosci. 7, 745e750.

Hornstein, E.P., Verweij, J., Li, P.H., Schnapf, J.L., 2005. Gap-junctional coupling and absolute sensitivity of photoreceptors in macaque retina. J. Neurosci. 25, 11201e 11209.

Hoshi, H., O’Brien, J., Mills, S.L., 2006. A novelfluorescent tracer for visualizing coupled cells in neural circuits of living tissue. J. Histochem. Cytochem. 54, 1169e1176.

Hu, E.H., Bloomfield, S.A., 2003. Gap junctional coupling underlies the short-latency spike synchrony of retinal alpha ganglion cells. J. Neurosci. 23, 6768e6777.

Hu, E.H., Pan, F., Völgyi, B., Bloomfield, S.A., 2010. Light increases the gap junctional coupling of retinal ganglion cells. J. Physiol. 588, 4145e4163.

Huang, H., Li, H., He, S.G., 2005. Identification of connexin 50 and 57 mRNA in A-type horizontal cells of the rabbit retina. Cell. Res. 15, 207e211.

Ishikane, H., Gangi, M., Honda, S., Tachibana, M., 2005. Synchronized retinal oscil-lations encode essential information for escape behavior in frogs. Nat. Neurosci.

8, 1087e1095.

Jacobs, A., Werblin, F., 1994. Modelling emergent image processing in the retina with cellular neural networks. Invest. Ophthalmol. Vis. Sci. 35, 21e27.

Jaillon, O., Aury, J.-M., Brunet, F., Petit, J.-L., Stange-Thomann, N., Mauceli, E., Bouneau, L., Fischer, C., Ozouf-Costaz, C., Bernot, A., Nicaud, S., Jaffe, D., Fisher, S., Lutfalla, G., Dossat, C., Segurens, B., Dasilva, C., Salanoubat, M., Levy, M., Boudet, N., Castellano, S., Anthouard, V., Jubin, C., Castelli, V., Katinka, M., Vacherie, B., Biémont, C., Skalli, Z., Cattolico, L., Poulain, J., de Berardinis, V., Cruaud, C., Duprat, S., Brottier, P., Coutanceau, J.-P., Gouzy, J., Parra, G., Lardier, G., Chapple, C., McKernan, K.J., McEwan, P., Bosak, S., Kellis, M., Volff, J.-N., Guigo, R., Zody, M.C., Mesirov, J., Lindblad-Toh, K., Birren, B., Nusbaum, C., Kahn, D., Robinson-Rechavi, M., Laudet, V., Schachter, V., Quetier, F., Saurin, W., Scarpelli, C., Wincker, P., Lander, E.S., Weissenbach, J., Crollius, H.R., 2004.

Genome duplication in the teleostfishTetraodon nigroviridisreveals the early vertebrate proto-karyotype. Nature 431, 946e957.

Janssen-Bienhold, U., Dermietzel, R., Weiler, R., 1998. Distribution of connexin43 immunoreactivity in the retinas of different vertebrates. J. Comp. Neurol. 396, 310e321.

Janssen-Bienhold, U., Trümpler, J., Hilgen, G., Schultz, K., Müller, L.P., Sonntag, S., Dedek, K., Dirks, P., Willecke, K., Weiler, R., 2009. Connexin57 is expressed in dendro-dendritic and axo-axonal gap junctions of mouse horizontal cells and its distribution is modulated by light. J. Comp. Neurol. 513, 363e374.

Johansson, K., Bruun, A., Ehinger, B., 1999. Gap junction protein connexin43 is heterogeneously expressed among glial cells in the adult rabbit retina. J. Comp.

Neurol. 407, 395e403.

Jouhou, H., Yamamoto, K., Iwasaki, M., Yamada, M., 2007. Acidification decouples gap junctions but enlarges the receptivefield size of horizontal cells in carp retina. Neurosci. Res. 57, 203e209.

Kaneko, A., 1971. Electrical connexions between horizontal cells in the dogfish retina. J. Physiol. 213, 95e105.

Kenyon, G.T., Marshak, D.W., 1998. Gap junctions with amacrine cells provide a feedback pathway for ganglion cells within the retina. Proc. Biol. Sci. 265, 919e925.

Kenyon, G.T., Theiler, J., George, J.S., Travis, B.J., Marshak, D.W., 2004. Correlated firing improves stimulus discrimination in a retinal model. Neural Comput. 16, 2261e2291.

Kerr, N.M., Johnson, C.S., de Souza, C.F., Chee, K.S., Good, W.R., Green, C.R., Danesh-Meyer, H.V., 2010. Immunolocalization of gap junction protein connexin43 (GJA1) in the human retina and optic nerve. Invest. Ophthalmol. Vis. Sci. 51, 4028e4034.

Kihara, A.H., de Castro, L.M., Moriscot, A.S., Hamassaki, D.E., 2006a. Prolonged dark adaptation changes connexin expression in the mouse retina. J. Neurosci. Res.

83, 1331e1341.

Kihara, A.H., Mantovani de Castro, L., Belmonte, M.A., Yan, C.Y., Moriscot, A.S., Hamassaki, D.E., 2006b. Expression of connexins 36, 43, and 45 during post-natal development of the mouse retina. J. Neurobiol. 66, 1397e1410.

Kihara, A.H., Paschon, V., Akamine, P.S., Saito, K.C., Leonelli, M., Jiang, J.X., Hamassaki, D.E., Britto, L.R., 2008. Differential expression of connexins during histogenesis of the chick retina. Dev. Neurobiol. 68, 1287e1302.

Kihara, A.H., Paschon, V., Cardoso, C.M., Higa, G.S., Castro, L.M., Hamassaki, D.E., Britto, L.R., 2009. Connexin36, an essential element in the rod pathway, is highly expressed in the essentially rodless retina ofGallus gallus. J. Comp.

Neurol. 512, 651e663.

Kohler, K., Kolbinger, W., Kurz-Isler, G., Weiler, R., 1990. Endogenous dopamine and cyclic events in thefish retina, II: correlation of retinomotor movement, spinule

formation, and connexon density of gap junctions with dopamine activity during light/dark cycles. Vis. Neurosci. 5, 417e428.

Kolb, H., 1979. The inner plexiform layer in the retina of the cat: electron micro-scopic observations. J. Neurocytol. 8, 295e329.

Kolb, H., Wang, H.H., 1985. The distribution of photoreceptors, dopaminergic amacrine cells and ganglion cells in the retina of the North American opossum (Didelphis virginiana). Vision Res. 25, 1207e1221.

Kothmann, W.W., Massey, S.C., O’Brien, J., 2009. Dopamine-stimulated dephos-phorylation of connexin 36 mediates AII amacrine cell uncoupling. J. Neurosci.

29, 14903e14911.

Kothmann, W.W., Trexler, B.E., Christopher, M., Whitaker, C.M., Li, W., Massey, S.C., O’Brien, J., 2012. Nonsynaptic NMDA receptors mediate activity-dependent plasticity of gap junctional coupling in the AII amacrine cell network.

J. Neurosci. 32, 6747e6759.

Krizaj, D., Gábriel, R., Owen, W.G., Witkovsky, P., 1998. Dopamine D2 receptor-mediated modulation of rodecone coupling in theXenopus retina. J. Comp.

Neurol. 398, 529e538.

Kurz-Isler, G., Voigt, T., Wolburg, H., 1992. Modulation of connexon densities in gap

Kurz-Isler, G., Voigt, T., Wolburg, H., 1992. Modulation of connexon densities in gap