Hostname: page-component-8448b6f56d-m8qmq Total loading time: 0 Render date: 2024-04-24T01:04:31.524Z Has data issue: false hasContentIssue false

Screening of gap junction antagonists on dye coupling in the rabbit retina

Published online by Cambridge University Press:  22 August 2007

FENG PAN
Affiliation:
Department of Ophthalmology and Visual Science, University of Texas Medical School at Houston, Houston Texas
STEPHEN L. MILLS
Affiliation:
Department of Ophthalmology and Visual Science, University of Texas Medical School at Houston, Houston Texas
STEPHEN C. MASSEY
Affiliation:
Department of Ophthalmology and Visual Science, University of Texas Medical School at Houston, Houston Texas

Abstract

Many cell types in the retina are coupled via gap junctions and so there is a pressing need for a potent and reversible gap junction antagonist. We screened a series of potential gap junction antagonists by evaluating their effects on dye coupling in the network of A-type horizontal cells. We evaluated the following compounds: meclofenamic acid (MFA), mefloquine, 2-aminoethyldiphenyl borate (2-APB), 18-α-glycyrrhetinic acid, 18-β-glycyrrhetinic acid (18-β-GA), retinoic acid, flufenamic acid, niflumic acid, and carbenoxolone. The efficacy of each drug was determined by measuring the diffusion coefficient for Neurobiotin (Mills & Massey, 1998). MFA, 18-β-GA, 2-APB and mefloquine were the most effective antagonists, completely eliminating A-type horizontal cell coupling at a concentration of 200 μM. Niflumic acid, flufenamic acid, and carbenoxolone were less potent. Additionally, carbenoxolone was difficult to wash out and also may be harmful, as the retina became opaque and swollen. MFA, 18-β-GA, 2-APB and mefloquine also blocked coupling in B-type horizontal cells and AII amacrine cells. Because these cell types express different connexins, this suggests that the antagonists were relatively non-selective across several different types of gap junction. It should be emphasized that MFA was water-soluble and its effects on dye coupling were easily reversible. In contrast, the other gap junction antagonists, except carbenoxolone, required DMSO to make stock solutions and were difficult to wash out of the preparation at the doses required to block coupling in A-type HCs. The combination of potency, water solubility and reversibility suggest that MFA may be a useful compound to manipulate gap junction coupling.

Type
Research Article
Copyright
© 2007 Cambridge University Press

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

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. Journal of Neuroscience 26, 42064215.Google Scholar
Bai, D., del Corsso, C., Srinivas, M. & Spray, D.C. (2006). Block of specific gap junction channel subtypes by 2-aminoethoxydiphenyl borate (2-APB). Journal of Pharmacology and Experimental Therapeutics 319, 14521458.Google Scholar
Bloomfield, S.A., Xin, D. & Osborne, T. (1997). Light-induced modulation of coupling between AII amacrine cells in the rabbit retina. Visual Neuroscience 14, 565576.Google Scholar
Boycott, B.B., Peichl, L. & Wässle, H. (1978). Morphological types of horizontal cell in the retina of the domestic cat. Proceeding of the Royal Society London B: Biological Science 203, 229245.Google Scholar
Chen, Q., Olney, J.W., Lukasiewicz, P.D., Almli, T. & Romano, C. (1998). Fenamates protect neurons against ischemic and excitotoxic injury in chick embryo retina. Neuroscience Letters 242, 163166.Google Scholar
Cruikshank, S.J., Hopperstad, M., Younger, M., Connors, B.W., Spray, D.C. & Srinivas, M. (2004). Potent block of Cx36 and Cx50 gap junction channels by mefloquine. Proceedings of the National Academy of Sciences 0402044101.
Dacheux, R.F. & Raviola, E. (1982). Horizontal cells in the retina of the rabbit. Journal of Neuroscience 2, 14861493.Google Scholar
Deans, M.R., Volgyi, B., Goodenough, D.A., Bloomfield, S.A. & Paul, D.L. (2002). Connexin36 is essential for transmission of rod-mediated visual signals in the mammalian retina. Neuron 36, 703712.Google Scholar
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. European Journal of Neuroscience 24, 16751686.Google Scholar
Dobrydneva, Y. & Blackmore, P. (2001). 2-Aminoethoxydiphenyl borate directly inhibits store-operated calcium entry channels in human platelets. Molecular Pharmacology 60, 541552.Google Scholar
Eskandari, S., Zampighi, G.A., Leung, D.W., Wright, E.M. & Loo, D.D. (2002). Inhibition of gap junction hemichannels by chloride channel blockers. Journal of Membrane Biology 185, 93102.Google Scholar
Feigenspan, A., Janssen-Bienhold, U., Hormuzdi, S., Monyer, H., Degen, J., Sohl, G., Willecke, K., Ammermuller, J. & Weiler, R. (2004). Expression of connexin36 in cone pedicles and OFF-cone bipolar cells of the mouse retina. Journal of Neuroscience 24, 33253334.Google Scholar
Feigenspan, A., Teubner, B., Willecke, K. & Weiler, R. (2001). Expression of Neuronal Connexin36 in AII Amacrine Cells of the Mammalian Retina. Journal of Neuroscience 21, 230239.Google Scholar
Goodenough, D.A. & Paul, D.L. (2003). Beyond the gap: Functions of unpaired connexon channels. Nature Reviews Molecular Cell Biology 4, 285294.Google Scholar
Guldenagel, M., Sohl, G., Plum, A., Traub, O., Teubner, B., Weiler, R. & Willecke, K. (2000). Expression patterns of connexin genes in mouse retina. Journal of Comparative Neurology 425, 193201.Google Scholar
Han, Y. & Massey, S.C. (2005). Electrical synapses in retinal ON cone bipolar cells: Subtype-specific expression of connexins. Proceedings of the National Academy of Sciences 102, 1331313318.Google Scholar
Harks, E.G., de Roos, A.D., Peters, P.H., de Haan, L.H., Brouwer, A., Ypey, D.L., van Zoelen, E.J. & Theuvenet, A.P. (2001). Fenamates: A novel class of reversible gap junction blockers. Journal of Pharmacology and Experimental Therapeutics 298, 10331041.Google Scholar
Hombach, S., Janssen-Bienhold, U., Sohl, G., Schubert, T., Bussow, H., Ott, T., Weiler, R. & Willecke, K. (2004). Functional expression of connexin57 in horizontal cells of the mouse retina. European Journal of Neuroscience 19, 26332640.Google Scholar
Lin, B., Jakobs, T.C. & Masland, R.H. (2005). Different functional types of bipolar cells use different gap-junctional proteins. Journal of Neuroscience 25, 66966701.Google Scholar
Lu, C., Zhang, D.Q. & McMahon, D.G. (1999). Electrical coupling of retinal horizontal cells mediated by distinct voltage-independent junctions. Visual Neuroscience 16, 811818.Google Scholar
Massey, S.C. & Mills, S.L. (1996). A calbindin-immunoreactive cone bipolar cell type in the rabbit retina. Journal of Comparative Neurology 366, 1533.Google Scholar
Massey, S.C. & Mills, S.L. (1999). Gap junctions between AII amacrine cells and calbindin-positive bipolar cells in the rabbit retina. Visual Neuroscience 16, 11811189.Google Scholar
Massey, S.C., O'Brien, J.J., Trexler, E.B., Li, W., Keung, J.W., Mills, S.L. & O'Brien, J. (2003). Multiple neuronal connexins in the mammalian retina. Cell Communication and Adhesion 10, 425430.Google Scholar
Maxeiner, S., Dedek, K., Janssen-Bienhold, U., Ammermuller, J., Brune, H., Kirsch, T., Pieper, M., Degen, J., Kruger, O., Willecke, K. & Weiler, R. (2005). Deletion of connexin45 in mouse retinal neurons disrupts the rod/cone signaling pathway between AII amacrine and ON cone bipolar cells and leads to impaired visual transmission. Journal of Neuroscience 25, 566576.Google Scholar
Mills, S.L. & Massey, S.C. (1994). Distribution and coverage of A- and B-type horizontal cells stained with Neurobiotin in the rabbit retina. Visual Neuroscience 11, 549560.Google Scholar
Mills, S.L. & Massey, S.C. (1995). Differential properties of two gap junctional pathways made by AII amacrine cells. Nature 377, 734737.Google Scholar
Mills, S.L. & Massey, S.C. (1998). The kinetics of tracer movement through homologous gap junctions in the rabbit retina. Visual Neuroscience 15, 765777.Google Scholar
Mills, S.L. & Massey, S.C. (2000). A series of biotinylated tracers distinguishes three types of gap junction in retina. Journal of Neuroscience 20, 86298636.Google Scholar
Mills, S.L., O'Brien, J.J., Li, W., O'Brien, J. & Massey, S.C. (2001). Rod pathways in the mammalian retina use connexin36. Journal of Comparative Neurology 436, 336350.Google Scholar
O'Brien, J., Nguyen, H.B. & Mills, S.L. (2004). Cone photoreceptors in bass retina use two connexins to mediate electrical coupling. Journal of Neuroscience 24, 56325642.Google Scholar
O'Brien, J.J., Li, W., Pan, F., Keung, J., O'Brien, J. & Massey, S.C. (2006). Coupling between A-type horizontal cells is mediated by connexin 50 gap junctions in the rabbit retina. Journal of Neuroscience 26, 1162411636.Google Scholar
Packer, O.S. & Dacey, D.M. (2002). Receptive field structure of H1 horizontal cells in macaque monkey retina. Journal of Vision 2, 272292.Google Scholar
Peichl, L. & Gonzalez-Soriano, J. (1994). Morphological types of horizontal cell in rodent retinae: A comparison of rat, mouse, gerbil, and guinea pig. Visual Neuroscience 11, 501517.Google Scholar
Peretz, A., Degani, N., Nachman, R., Uziyel, Y., Gibor, G., Shabat, D. & Attali, B. (2005). Meclofenamic acid and diclofenac, novel templates of KCNQ2/Q3 potassium channel openers, depress cortical neuron activity and exhibit anticonvulsant properties. Molecular Pharmacology 67, 10531066.Google Scholar
Raviola, E. & Dacheux, R.F. (1990). Axonless horizontal cells of the rabbit retina: Synaptic connections and origin of the rod aftereffect. Journal of Neurocytology 19, 731736.Google Scholar
Reitsamer, H.A., Pflug, R., Franz, M. & Huber, S. (2006). Dopaminergic modulation of horizontal-cell-axon-terminal receptive field size in the mammalian retina. Vision Research 46, 467474.Google Scholar
Rodieck, R. (1998). The First Steps in Seeing. Sunderland, MA: Sinauer Associates, Inc.
Schubert, T., Maxeiner, S., Kruger, O., Willecke, K. & Weiler, R. (2005). Connexin45 mediates gap junctional coupling of bistratified ganglion cells in the mouse retina. Journal of Comparative Neurology 490, 2939.Google Scholar
Sohl, G., Maxeiner, S. & Willecke, K. (2005). Expression and functions of neuronal gap junctions. Nature Reviews Neuroscience 6, 191200.Google Scholar
Srinivas, M., Costa, M., Gao, Y., Fort, A., Fishman, G.I. & Spray, D.C. (1999a). Voltage dependence of macroscopic and unitary currents of gap junction channels formed by mouse connexin50 expressed in rat neuroblastoma cells. Journal of Physiology 517 (3), 673689.Google Scholar
Srinivas, M., Hopperstad, M.G. & Spray, D.C. (2001). Quinine blocks specific gap junction channel subtypes. Proceedings of the National Academy of Sciences 98, 1094210947.CrossRefGoogle Scholar
Srinivas, M., Kronengold, J., Bukauskas, F.F., Bargiello, T.A. & Verselis, V.K. (2005). Correlative studies of gating in Cx46 and Cx50 hemichannels and gap junction channels. Biophysics Journal 88, 17251739.CrossRefGoogle Scholar
Srinivas, M., Rozental, R., Kojima, T., Dermietzel, R., Mehler, M., Condorelli, D.F., Kessler, J.A. & Spray, D.C. (1999b). Functional properties of channels formed by the neuronal gap junction protein connexin36. Journal of Neuroscience 19, 98489855.Google Scholar
Srinivas, M. & Spray, D.C. (2003). Closure of gap junction channels by arylaminobenzoates. Molecular Pharmacology 63, 13891397.CrossRefGoogle Scholar
Suryanarayanan, A. & Slaughter, M.M. (2006). Synaptic transmission mediated by internal calcium stores in rod photoreceptors. Journal of Neuroscience 26, 17591766.CrossRefGoogle Scholar
Trexler, E.B., Li, W. & Massey, S.C. (2005). Simultaneous contribution of two rod pathways to AII amacrine and cone bipolar cell light responses. Journal of Neurophysiology 93, 14761485.Google Scholar
Vaney, D.I. (1994). Patterns of neuronal coupling in the retina. Progress in Retinal and Eye Research 13, 301355.CrossRefGoogle Scholar
Vaney, D.I., Nelson, J.C. & Pow, D.V. (1998). Neurotransmitter coupling through gap junctions in the retina. Journal of Neuroscience 18, 1059410602.Google Scholar
Verweij, J., Hornstein, E.P. & Schnapf, J.L. (2003). Surround antagonism in macaque cone photoreceptors. Journal of Neuroscience 23, 1024910257.Google Scholar
Vessey, J.P., Lalonde, M.R., Mizan, H.A., Welch, N.C., Kelly, M.E. & Barnes, S. (2004). Carbenoxolone inhibition of voltage-gated Ca channels and synaptic transmission in the retina. Journal of Neurophysiology 92, 12521256.CrossRefGoogle Scholar
Volgyi, B., Abrams, J., Paul, D.L. & Bloomfield, S.A. (2005). Morphology and tracer coupling pattern of alpha ganglion cells in the mouse retina. Journal of Comparative Neurology 492, 6677.CrossRefGoogle Scholar
Wässle, H., Boycott, B.B. & Peichl, L. (1978). Receptor contacts of horizontal cells in the retina of the domestic cat. Proceeding of the Royal Society London B: Biological Science 203, 247267.CrossRefGoogle Scholar
Willecke, K., Eiberger, J., Degen, J., Eckardt, D., Romualdi, A., Guldenagel, M., Deutsch, U. & Sohl, G. (2002). Structural and functional diversity of connexin genes in the mouse and human genome. Journal of Biological Chemistry 383, 725737.CrossRefGoogle Scholar
Xia, Y. & Nawy, S. (2003). The gap junction blockers carbenoxolone and 18beta-glycyrrhetinic acid antagonize cone-driven light responses in the mouse retina. Visual Neuroscience 20, 429435.CrossRefGoogle Scholar
Zhang, J. & Wu, S.M. (2004). Connexin35/36 gap junction proteins are expressed in photoreceptors of the tiger salamander retina. Journal of Compartive Neurology 470, 112.CrossRefGoogle Scholar
Zimmerman, A.L. & Rose, B. (1985). Permeability properties of cell-to-cell channels: Kinetics of fluorescent tracer diffusion through a cell junction. Journal of Membrane Biology 84, 269283.CrossRefGoogle Scholar