Hostname: page-component-76fb5796d-skm99 Total loading time: 0 Render date: 2024-04-25T10:36:19.565Z Has data issue: false hasContentIssue false

A model of high-frequency oscillatory potentials in retinal ganglion cells

Published online by Cambridge University Press:  22 January 2004

GARRETT T. KENYON
Affiliation:
P-21, Biophysics, Los Alamos National Laboratory, Los Alamos
BARTLETT MOORE
Affiliation:
P-21, Biophysics, Los Alamos National Laboratory, Los Alamos Department of Neurobiology and Anatomy, University of Texas Medical School, Houston
JANELLE JEFFS
Affiliation:
P-21, Biophysics, Los Alamos National Laboratory, Los Alamos Department of Bioengineering, University of Utah, Salt Lake City
KATE S. DENNING
Affiliation:
P-21, Biophysics, Los Alamos National Laboratory, Los Alamos
GREG J. STEPHENS
Affiliation:
P-21, Biophysics, Los Alamos National Laboratory, Los Alamos
BRYAN J. TRAVIS
Affiliation:
EES-6, Hydrology, Geochemistry, and Geology, Los Alamos National Laboratory, Los Alamos
JOHN S. GEORGE
Affiliation:
P-21, Biophysics, Los Alamos National Laboratory, Los Alamos
JAMES THEILER
Affiliation:
NIS-2, Space and Remote Sensing Sciences, Los Alamos National Laboratory, Los Alamos
DAVID W. MARSHAK
Affiliation:
Department of Neurobiology and Anatomy, University of Texas Medical School, Houston

Abstract

High-frequency oscillatory potentials (HFOPs) have been recorded from ganglion cells in cat, rabbit, frog, and mudpuppy retina and in electroretinograms (ERGs) from humans and other primates. However, the origin of HFOPs is unknown. Based on patterns of tracer coupling, we hypothesized that HFOPs could be generated, in part, by negative feedback from axon-bearing amacrine cells excited via electrical synapses with neighboring ganglion cells. Computer simulations were used to determine whether such axon-mediated feedback was consistent with the experimentally observed properties of HFOPs. (1) Periodic signals are typically absent from ganglion cell PSTHs, in part because the phases of retinal HFOPs vary randomly over time and are only weakly stimulus locked. In the retinal model, this phase variability resulted from the nonlinear properties of axon-mediated feedback in combination with synaptic noise. (2) HFOPs increase as a function of stimulus size up to several times the receptive-field center diameter. In the model, axon-mediated feedback pooled signals over a large retinal area, producing HFOPs that were similarly size dependent. (3) HFOPs are stimulus specific. In the model, gap junctions between neighboring neurons caused contiguous regions to become phase locked, but did not synchronize separate regions. Model-generated HFOPs were consistent with the receptive-field center dynamics and spatial organization of cat alpha cells. HFOPs did not depend qualitatively on the exact value of any model parameter or on the numerical precision of the integration method. We conclude that HFOPs could be mediated, in part, by circuitry consistent with known retinal anatomy.

Type
Research Article
Copyright
2003 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

Alonso, J.M., Usrey, W.M., & Reid, R.C. (1996). Precisely correlated firing in cells of the lateral geniculate nucleus. Nature 383(6603), 815819.Google Scholar
Amthor, F.R. & Grzywacz, N.M. (2001). Synchronous firing is stimulus dependent in directionally selective (DS) and non-DS rabbit ganglion cells. Investigative Ophthalmology & Visual Science 42(4), S677.Google Scholar
Ariel, M., Daw, N.W., & Rader, R.K. (1983). Rhythmicity in rabbit retinal ganglion cell responses. Vision Research 23(12), 14851493.Google Scholar
Boycott, B.B. & Wässle, H. (1974). The morphological types of ganglion cells of the domestic cat's retina. Journal of Physiology 240(2), 397419.Google Scholar
Brivanlou, I.H., Warland, D.K., & Meister, M. (1998). Mechanisms of concerted firing among retinal ganglion cells. Neuron 20(3), 527539.Google Scholar
Castelo-Branco, M., Neuenschwander, S., & Singer, W. (1998). Synchronization of visual responses between the cortex, lateral geniculate nucleus, and retina in the anesthetized cat. Journal of Neuroscience 18(16), 63956410.Google Scholar
Cohen, E.D. (1998). Interactions of inhibition and excitation in the light-evoked currents of X type retinal ganglion cells. Journal of Neurophysiology 80(6), 29752990.Google Scholar
Creutzfeldt, O.D., Sakmann, B., Scheich, H., & Korn, A. (1970). Sensitivity distribution and spatial summation within receptive-field center of retinal on-center ganglion cells and transfer function of the retina. Journal of Neurophysiology 33(5), 654671.Google Scholar
Dacey, D.M. & Brace, S. (1992). A coupled network for parasol but not midget ganglion cells in the primate retina. Visual Neuroscience 9, 279290.Google Scholar
De Carli, F., Narici, L., Canovaro, P., Carozzo, S., Agazzi, E., & Sannita, W.G. (2001). Stimulus- and frequency-specific oscillatory mass responses to visual stimulation in man. Clinical Electroencephalography 32(3), 145151.Google Scholar
Dubin, M.W. (1970). The inner plexiform layer of the vertebrate retina: A quantitative and comparative electron microscopic analysis. Journal of Comparative Neurology 140(4), 479505.Google Scholar
Euler, T. & Wassle, H. (1998). Different contributions of GABAA and GABAC receptors to rod and cone bipolar cells in a rat retinal slice preparation. Journal of Neurophysiology 79(3), 13841395.Google Scholar
Euler, T. & Masland, R.H. (2000). Light-evoked responses of bipolar cells in a mammalian retina. Journal of Neurophysiology 83(4), 18171829.Google Scholar
Foerster, M.H., van de Grind, W.A., & Grusser, O.J. (1977). The response of cat horizontal cells to flicker stimuli of different area, intensity and frequency. Experimental Brain Research 29(3–4), 367385.Google Scholar
Freed, M.A. (2000). Rate of quantal excitation to a retinal ganglion cell evoked by sensory input. Journal of Neurophysiology 83(5), 29562966.Google Scholar
Freed, M.A. & Sterling, P. (1988). The ON-alpha ganglion cell of the cat retina and its presynaptic cell types. Journal of Neuroscience 8(7), 23032320.Google Scholar
Freed, M.A., Pflug, R., Kolb, H., & Nelson, R. (1996). ON–OFF amacrine cells in cat retina. Journal of Comparative Neurology 364(3), 556566.Google Scholar
Frishman, L.J., Freeman, A.W., Troy, J.B., Schweitzer-Tong, D.E., & Enroth-Cugell, C. (1987). Spatiotemporal frequency responses of cat retinal ganglion cells. Journal of General Physiology 89(4), 599628.Google Scholar
Frishman, L.J., Saszik, S., Harwerth, R.S., Viswanathan, S., Li, Y., Smith, E.L., III, Robson, J.G., & Barnes, G. (2000). Effects of experimental glaucoma in macaques on the multifocal ERG. Multifocal ERG in laser-induced glaucoma. Documenta Ophthalmologica 100(2–3), 231251.Google Scholar
Gerstein, G.L. & Perkel, D.H. (1972). Mutual temporal relationships among neuronal spike trains. Statistical techniques for display and analysis. Biophysical Journal 12(5), 453473.Google Scholar
Guldenagel, M., Ammermuller, J., Feigenspan, A., Teubner, B., Degen, J., Sohl, G., Willecke, K., & Weiler, R. (2001). Visual transmission deficits in mice with targeted disruption of the gap junction gene connexin36. Journal of Neuroscience 21(16), 60366044.Google Scholar
Hansel, D., Mato, G., Meunier, C., & Neltner, L. (1998). On numerical simulations of integrate-and-fire neural networks. Neural Computations 10(2), 467483.Google Scholar
Hormuzdi, S.G., Pais, I., LeBeau, F.E., Towers, S.K., Rozov, A., Buhl, E.H., Whittington, M.A., & Monyer, H. (2001). Impaired electrical signaling disrupts gamma frequency oscillations in connexin 36-deficient mice. Neuron 31(3), 487495.Google Scholar
Ishikane, H., Kawana, A., & Tachibana, M. (1999). Short- and long-range synchronous activities in dimming detectors of the frog retina. Visual Neuroscience 16(6), 10011014.Google Scholar
Jacoby, R., Stafford, D., Kouyama, N., & Marshak, D. (1996). Synaptic inputs to ON parasol ganglion cells in the primate retina. Journal of Neuroscience 16(24), 80418056.Google Scholar
Kenyon, G.T. & Marshak, D.W. (1998). Gap junctions with amacrine cells provide a feedback pathway for ganglion cells within the retina. Proceedings of the Royal Society B (London) 265(1399), 919925.Google Scholar
Kenyon, G.T. & Marshak, D.W. (2000). Synchrony of ganglion cells encodes stimulus intensity in a retinal model. Society for Neuroscience 26, 1328.Google Scholar
Kenyon, G.T. & Marshak, D.W. (2001). Amacrine cells synchronize the firing of alpha ganglion cells over a wide range of stimulus intensities. Investigative Ophthalmology & Visual Science 42(4), S674.Google Scholar
Kenyon, G.T., Fetz, E.E., & Puff, R.D. (1990). Effects of firing synchrony on signal propagation in layered networks. In Advances in Neural Information Processing Systems, Vol. 2, ed. Touretzky, D.S., pp. 141148. San Mateo, CA: Morgan Kaufmann.
Kenyon, G.T., Moore, K.R., & Marshak, D.W. (1999). Stimulus-specific synchrony between alpha ganglion cells in a computer model of the mammalian retina. Society for Neuroscience 25, 1042.Google Scholar
Kolb, H. & Nelson, R. (1985). Functional neurocircuitry of amacrine cells in the cat retina. In Neurocircuitry of the Retina: A Cajal Memorial, ed. Gallego, A. & Gouras, P., pp. 215232. New York: Elsevier.
Kolb, H. & Nelson, R. (1993). OFF-alpha and OFF-beta ganglion cells in cat retina: II. Neural circuitry as revealed by electron microscopy of HRP stains. Journal of Comparative Neurology 329(1), 85110.Google Scholar
Laufer, M. & Verzeano, M. (1967). Periodic activity in the visual system of the cat. Vision Research 7(3), 215229.Google Scholar
Lee, B.B., Pokorny, J., Smith, V.C., & Kremers, J. (1994). Responses to pulses and sinusoids in macaque ganglion cells. Vision Research 34(23), 30813096.Google Scholar
Lipton, S.A. & Tauck, D.L. (1987). Voltage-dependent conductances of solitary ganglion cells dissociated from the rat retina. Journal of Physiology 385, 361391.Google Scholar
Marc, R.E. & Liu, W. (2000). Fundamental GABAergic amacrine cell circuitries in the retina: Nested feedback, concatenated inhibition, and axosomatic synapses. Journal of Comparative Neurology 425(4), 560582.Google Scholar
Mastronarde, D.N. (1983). Correlated firing of cat retinal ganglion cells. I. Spontaneously active inputs to X- and Y-cells. Journal of Neurophysiology 49(2), 303324.Google Scholar
Mastronarde, D.N. (1989). Correlated firing of retinal ganglion cells. Trends in Neurosciences 12(2), 7580.Google Scholar
Matsui, K., Hasegawa, J., & Tachibana, M. (2001). Modulation of excitatory synaptic transmission by GABA(C) receptor-mediated feedback in the mouse inner retina. Journal of Neurophysiology 86(5), 22852298.Google Scholar
Meister, M., Pine, J., & Baylor, D.A. (1994). Multi-neuronal signals from the retina: Acquisition and analysis. Journal of Neuroscience Methods 51(1), 95106.Google Scholar
Miller, R.F., Stenback, K., Henderson, D., & Sikora, M. (2002). How voltage-gated ion channels alter the functional properties of ganglion and amacrine cell dendrites. Arch Ital Biol 140(4), 347359.Google Scholar
Neuenschwander, S. & Singer, W. (1996). Long-range synchronization of oscillatory light responses in the cat retina and lateral geniculate nucleus. Nature 379(6567), 728732.Google Scholar
Neuenschwander, S., Engel, A.K., Konig, P., Singer, W., & Varela, F.J. (1996). Synchronization of neuronal responses in the optic tectum of awake pigeons. Visual Neuroscience 13(3), 575584.Google Scholar
Neuenschwander, S., Castelo-Branco, M., & Singer, W. (1999). Synchronous oscillations in the cat retina. Vision Research 39(15), 24852497.Google Scholar
Nirenberg, S. & Meister, M. (1997). The light response of retinal ganglion cells is truncated by a displaced amacrine circuit. Neuron 18(4), 637650.Google Scholar
O'Brien, B.J., Isayama, T., Richardson, R., & Berson, D.M. (2002). Intrinsic physiological properties of cat retinal ganglion cells. Journal of Physiology 538(Pt. 3), 787802.Google Scholar
Owczarzak, M.T. & Pourcho, R.G. (1999). Transmitter-specific input to OFF-alpha ganglion cells in the cat retina. Anatomical Record 255(4), 363373.3.0.CO;2-9>CrossRefGoogle Scholar
Peichl, L. (1991). Alpha ganglion cells in mammalian retinae: Common properties, species differences, and some comments on other ganglion cells. Visual Neuroscience 7, 155169.CrossRefGoogle Scholar
Peichl, L., Ott, H., & Boycott, B.B. (1987). Alpha ganglion cells in mammalian retinae. Proceedings of the Royal Society B (London) 231(1263), 169197.CrossRefGoogle Scholar
Rangaswamy, N.V., Hood, D.C., & Frishman, L.J. (2003). Regional variations in local contributions to the primate photopic flash ERG: Revealed using the slow-sequence mfERG. Investigative Ophthalmology & Visual Science 44(7), 32333247.CrossRefGoogle Scholar
Reich, D.S., Victor, J.D., Knight, B.W., Ozaki, T., & Kaplan, E. (1997). Response variability and timing precision of neuronal spike trains in vivo. Journal of Neurophysiology 77(5), 28362841.Google Scholar
Roska, B., Nemeth, E., Orzo, L., & Werblin, F.S. (2000). Three levels of lateral inhibition: A space–time study of the retina of the tiger salamander. Journal of Neuroscience 20(5), 19411951.Google Scholar
Shapley, R.M. & Victor, J.D. (1978). The effect of contrast on the transfer properties of cat retinal ganglion cells. Journal of Physiology 285, 275298.CrossRefGoogle Scholar
Shields, C.R. & Lukasiewicz, P.D. (2003). Spike-dependent GABA inputs to bipolar cell axon terminals contribute to lateral inhibition of retinal ganglion cells. Journal of Neurophysiology 89(5), 24492458.CrossRefGoogle Scholar
Smith, V.C., Pokorny, J., Lee, B.B., & Dacey, D.M. (2001). Primate horizontal cell dynamics: an analysis of sensitivity regulation in the outer retina. Journal of Neurophysiology 85(2), 545558.Google Scholar
Solessio, E., Vigh, J., Cuenca, N., Rapp, K., & Lasater, E.M. (2002). Membrane properties of an unusual intrinsically oscillating, wide-field teleost retinal amacrine cell. Journal of Physiology 544(Pt. 3), 831847.CrossRefGoogle Scholar
Solomon, S.G., Martin, P.R., White, A.J., Ruttiger, L., & Lee, B.B. (2002). Modulation sensitivity of ganglion cells in peripheral retina of macaque. Vision Research 42(27), 28932898.CrossRefGoogle Scholar
Steinberg, R.H. (1966). Oscillatory activity in the optic tract of cat and light adaptation. Journal of Neurophysiology 29(2), 139156.Google Scholar
Teeters, J., Jacobs, A., & Werblin, F. (1997). How neural interactions form neural responses in the salamander retina. Journal of Computational Neuroscience 4(1), 527.CrossRefGoogle Scholar
Troy, J.B. & Robson, J.G. (1992). Steady discharges of X and Y retinal ganglion cells of cat under photopic illuminance. Visual Neuroscience 9(6), 535553.CrossRefGoogle Scholar
Troy, J.B. & Enroth-Cugell, C. (1993). X and Y ganglion cells inform the cat's brain about contrast in the retinal image. Experimental Brain Research 93(3), 383390.CrossRefGoogle Scholar
Troy, J.B., Oh, J.K., & Enroth-Cugell, C. (1993). Effect of ambient illumination on the spatial properties of the center and surround of Y-cell receptive fields. Visual Neuroscience 10(4), 753764.CrossRefGoogle Scholar
Usrey, W.M., Alonso, J.M., & Reid, R.C. (2000). Synaptic interactions between thalamic inputs to simple cells in cat visual cortex. Journal of Neuroscience 20(14), 54615467.Google Scholar
Vaney, D.I. (1990). The mosaic of amacrine cells in the mammalian retina. Progress in Retinal Research, Vol. 9, ed. Osborne, N.N. & Chader, G.J., pp. 49100. Oxford: Pergamon Press.CrossRef
Vaney, D.I. (1994). Patterns of neuronal coupling in the retina. Progress in Retinal and Eye Research 13, 301355.CrossRefGoogle Scholar
Vardi, N., Masarachia, P.J., & Sterling, P. (1989). Structure of the starburst amacrine network in the cat retina and its association with alpha ganglion cells. Journal of Comparative Neurology 288(4), 601611.CrossRefGoogle Scholar
Vigh, J., Solessio, E., Morgans, C.W., & Lasater, E.M. (2003). Ionic mechanisms mediating oscillatory membrane potentials in wide-field retinal amacrine cells. Journal of Neurophysiology 20, 20.CrossRefGoogle Scholar
Wachtmeister, L. (1998). Oscillatory potentials in the retina: What do they reveal. Progress in Retinal and Eye Research 17(4), 485521.CrossRefGoogle Scholar
Wachtmeister, L. & Dowling, J.E. (1978). The oscillatory potentials of the mudpuppy retina. Investigative Ophthalmology and Visual Science 17(12), 11761188.Google Scholar
Waxman, S.G. & Bennett, M.V. (1972). Relative conduction velocities of small myelinated and non-myelinated fibres in the central nervous system. Nature: New Biology 238(85), 217219.Google Scholar
Xin, D. & Bloomfield, S.A. (1997). Tracer coupling pattern of amacrine and ganglion cells in the rabbit retina. Journal of Comparative Neurology 383(4), 512528.3.0.CO;2-5>CrossRefGoogle Scholar
Yoshida, K., Watanabe, D., Ishikane, H., Tachibana, M., Pastan, I., & Nakanishi, S. (2001). A key role of starburst amacrine cells in originating retinal directional selectivity and optokinetic eye movement. Neuron 30(3), 771780.CrossRefGoogle Scholar