Hostname: page-component-848d4c4894-hfldf Total loading time: 0 Render date: 2024-05-11T20:49:22.669Z Has data issue: false hasContentIssue false

Interrelations of bioenergetic and sensory functions of the retinal proteins

Published online by Cambridge University Press:  17 March 2009

Vladimir P. Skulachev
Affiliation:
Department of Bioenergetics, A. N. Belozersky Institute of Physico-Chemical Biology, Moscow State University, Moscow 119899, Russia

Extract

Rhodopsins are intrinsic membrane retinal-containing proteins composed of 7 hydrophobic a-helical transmembrane columns and hydrophilic sequences of various length connecting the helices and localized at N- and C-ends of the polypeptide. The chromophore (retinal) forms a Schiff base with a lysine residue in the middle part of the last a-helix. Absorption of a photon results in isomerization of retinal which gives rise to a conformational change in the protein moiety.

Type
Research Article
Copyright
Copyright © Cambridge University Press 1993

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

Bamberg, E., Tittor, J. & Oesterhelt, D. (1993). Light-driven proton or chloride pumping by halorhodospin. Proc. Natn. Acad. Sci. USA 90, 639643.Google Scholar
Barak, R. & Eisenbach, M. (1992). Fumarate or a fumarate metabolite restores switching ability to rotating flagella of bacterial envelopes. J. Bacteriol. 174, 643645.Google Scholar
Baryshev, V. A., Glagolev, A. N. & Skulachev, V. P. (1981). Sensing of in phototaxis of Halobacterium halobium. Nature 292, 338340.Google Scholar
Baryshev, V. A., Glagolev, A. N. & Skulachev, V. P. (1983). Interrelation of phototaxis, membrane potential and K+/Na+ gradient in Halobacterium halobium. J. Gen. Microbiol. 129, 367373.Google Scholar
Bibikov, S. I., Grishanin, R. N., Kaulen, A. D., Marwan, W., Oesterhelt, D. & Skulachev, V. P. (1992 a). Direct evidence for the involvement of membrane potential changes in the photosensory transduction of halobacteria. In Structures and Functions of Retinal Proteins (ed. Rigaud, J. L.), pp. 333336. Dourdan: Colloque INSERM.Google Scholar
Bibikov, S. I., Grishanin, R. N., Kaulen, A. D., Marwan, W., Oesterhelt, D. & Skulachev, V. P. (1992 b). Role of bacteriorhodopsin in photoreception of Halobacterium halobium. Bioorg. Khimiya 18, 14031423 [in Russian\.Google Scholar
Bibikov, S. I., Grishanin, R. N., Kaulen, A. D., Marwan, W., Oesterhelt, D. & Skulachev, V. P. (1993). Bacteriorhodopsin is involved in the system of halobacterial photoreceptor. Proc. Natn. Acad. Sci. USA (accepted).CrossRefGoogle Scholar
Bibikov, S. I., Grishanin, R. N., Marwan, W., Oesterhelt, D. & Skulachev, V. P. (1991). The proton pump bacteriorhodopsin is a photoreceptor for signal transduction in Halobacterium halobium. FEBS Lett. 295, 223226.Google Scholar
Bibikov, S. I. & Skulachev, V. P. (1989). Mechanism of phototaxis and aerotaxis in Halobacterium halobium. FEBS Lett. 243, 303306.CrossRefGoogle Scholar
Bogomolni, R. A. & Spudich, J. L. (1982). Identification of a third rhodopsin-like pigment in phototactic Halobacterium halobium. Proc. Natn. Acad. Sci. USA 79, 62506254.Google Scholar
Brown, I. I., Galperin, M. Yu., Glagolev, A. N. & Skulachev, V. P. (1983). Utilization of energy stored in the form of Na+ and K+ ion gradient by bacterial cells. Eur. J. Biochem. 134, 345349.Google Scholar
Brown, K. T. & Murakami, M. (1964). A new receptor potential of the monkey retina with no detectable latency. Nature 201, 626628.Google Scholar
Butt, H. J., Fendler, K., Bamberg, E., Tittor, J. & Oesterhelt, D. (1989). Aspartic acids 96 and 85 plays a central role in the function of bacteriorhodopsin as a proton pump. EMBO J. 4, 16571663.CrossRefGoogle Scholar
Cervetto, L., Pasino, E. & Torre, V. (1977). Electrical responses of rods in the retina of Bufo marinus. J. Physiol. 267, 1751.Google Scholar
Cone, R. A. (1967). Early receptor potential: photoreversible charge displacement in rhodopsin. Science 155, 11281131.Google Scholar
Dancshazy, Z., Drachev, L. A., Osmos, P., Nagy, K. & Skulachev, V. P. (1978). Kinetics of the blue-light inhibition of photoelectric activity of bacteriorhodopsin. FEBS Lett. 96, 5963.Google Scholar
Danshina, S. V., Drachev, L. A., Kaulen, A. D. & Skulachev, V. P. (1992 a). The inward H+ pathway in bacteriorhodopsin: the role of M412 intermediates. Photochem. Photobiol. 55, 735740.CrossRefGoogle Scholar
Danshina, S. V., Drachev, L. A., Kaulen, A. D., Khorana, H. G., Marti, T., Mogi, T. & Skulachev, V. P. (1992 b). The mechanism of H+ transport by bacteriorhodopsin: a study on Asp-96 mutants. Biokhimiya 57, 15741585 [in Russian].Google Scholar
Der, A., Szaraz, S., Toth-Boconadi, R., Tokaji, Zs., Keszthelyi, L. & Stoeckenius, W. (1991). Alternative translocation of protons and halide ions by bacteriorhodopsin. Proc. Natn. Acad. Set. USA 88, 47514755.CrossRefGoogle ScholarPubMed
Der, A., Toth-Boconadi, R. & Keszthelyi, L. (1989). Bacteriorhodopsin as a possible chloride pump. FEBS Lett. 259, 2426.Google Scholar
Drachev, A. L., Drachev, L. A., Kaulen, A. D., Khitrina, L. V., Skulachev, V. P., Lepnev, G. P. & Chekulaeva, L. N. (1989). Anion-dependent transition of two acidic forms of bacteriorhodopsin. Biochim. Biophys. Ada 976, 190195.CrossRefGoogle Scholar
Drachev, L. A., Kalamkarov, G. R., Kaulen, A. D., Ostrovsky, M. A. & Skulachev, V. P. (1981). Fast stages of photoelectric processes in biological membranes. II. Visual rhodopsin. Eur. J. Biochem. 117, 471481.CrossRefGoogle ScholarPubMed
Drachev, L. A., Kaulen, A. D., Khorana, H. G., Mogi, T., Otto, H., Skulachev, V. P., Heyn, M. P. & Holz, M. (1989 a). Participation of the Asp-96 carboxylic group in H+ transfer along the inner proton-conducting pathway. Biokhimiya 54, 14671477 [in Russian].Google Scholar
Drachev, L. A., Kaulen, A. D. & Komrakov, A. Yu. (1992). Interrelations of Mintermediates in bacteriorhodopsin photocycle. FEBS Lett. 313, 248250.CrossRefGoogle ScholarPubMed
Drachev, L. A., Kaulen, A. D., Ostroumov, S. A. & Skulachev, V. P. (1974). Electrogenesis by bacteriorhodopsin incorporated in a planar phospholipid membrane. FEBS Lett. 39, 3335.Google Scholar
Drachev, L. A., Kaulen, A. D. & Skulachev, V. P. (1978). Time-resolution of the intermediate steps in the bacteriorhodopsin-linked electrogenesis. FEBS Lett. 87, 161167.Google Scholar
Drachev, L. A., Kaulen, A. D. & Skulachev, V. P. (1984). Correlation of photochemical cycle, H+-release and uptake, and electric events in bacteriorhodopsin. FEBS Lett. 178, 331335.Google Scholar
Drachev, L. A., Kaulen, A. D. & Zorina, V. V. (1989 b). Light-scattering changes in the bacteriorhodopsin photocycle. FEBS Lett. 243, 57.Google Scholar
Dracheva, S. M., Drachev, L. A., Konstantinov, A. A., Semenov, A. Yu., Skulachev, V. P., Arutjunjan, A. M., Shuvalov, V. A. & Zaberezhnaya, S. (1988). Electrogenic steps in the redox reactions catalyzed by photosynthetic reaction center complexes from Rhodopseudomonas viridis. Eur. J. Biochem. 171, 253264.Google Scholar
Draheim, J. E. & Cassim, J. Y. (1985). Large scale global structural changes of the purple membrane during the photocycle. Biophys. J. 47, 497507.Google Scholar
Fischer, U. & Oesterhelt, D. (1979). Chromophore equilibria in bacteriorhodopsin. Biophys. J. 28, 211230.Google Scholar
Gagne, S., Roebroek, J. G. H. & Stavenga, D. G. (1989). Enigma of early receptor potential in fly eyes. Vision Res. 29, 16631670.Google Scholar
Gerwert, K., Hess, B., Soppa, J. & Oesterhelt, D. (1989). Role of aspartate-96 in proton translocation by bacteriorhodopsin. Proc. Natn. Acad. Sci. USA 86, 49434947.CrossRefGoogle ScholarPubMed
Glaeser, R. M., Baldwin, J. M., Ceska, T. A. & Henderson, R. (1986). Electron diffraction analysis of the M412 intermediate of bacteriorhodopsin. Biophys. J. 50, 913920.Google Scholar
Glagolev, A. N. (1980). Reception of the energy in bacterial taxis. J. Theor. Biol. 82, 171185.Google Scholar
Glagolev, A. N. (1984). Bacterial -sensing. Trends Biochem. Sci. 9, 397–40.Google Scholar
Hegemann, P., Oesterhelt, D. & Steiner, M. (1985). The photocycle of the chloride pump bacteriorhodopsin. I. Azide-catalyzed deprotonation of the chromophore is a side reaction of photocycle intermediate inactivating the pump. EMBO J. 4, 23472350.CrossRefGoogle Scholar
Henderson, R., Baldwin, J. M., Ceska, T. A., Zemlin, F., Beckmann, F. & Downing, K. H. (1990). Model for the structure of bacteriorhodopsin based on high-resolution electron cryo-microscopy. J. Mol. Biol. 213, 899929.Google Scholar
Holz, N., Drachev, L. A., Mogi, T., Otto, H., Kaulen, A. D., Heyn, M. P., Skulachev, V. P. & Khorana, H. G. (1989). Replacement of aspartic acid-96 by asparagine in bacteriorhodopsin slows both the decay of the M intermediate and the associated proton movement. Proc. Natn. Acad. Sci. USA 86, 21672171.Google Scholar
Hurley, J. B. (1992). Signal transduction enzymes of vertebrate photoreceptors. J. Bioenerg. Biomembr. 24, 219226.Google Scholar
Karvaly, B. & Dancshazy, Z. (1977). Bacteriorhodopsin: a molecular photoelectric regulator. Quenching of photovoltaic effect of bimolecular lipid membranes containing bacteriorhodopsin by blue light. FEBS Lett. 76, 3640.Google Scholar
Kayushin, L. P. & Skulachev, V. P. (1974). Bacteriorhodopsin as an electrogenic proton pump: reconstitution of bacteriorhodopsin proteoliposomes generating ΔΫ and ΔpH. FEBS Lett. 39, 3942.Google Scholar
Kihara, M. & Macnab, R. M. (1981). Cytoplasmic pH mediates pH taxis and weakacid repellent taxis of bacteria. J. Bacteriol. 145, 12091221.CrossRefGoogle ScholarPubMed
Lamb, T. D. & Simon, E. J. (1977). Analysis of electrical noise in turtle cones. J. Physiol. 972, 435468.Google Scholar
Lanyi, J. K. (1992). Proton transfer and energy coupling in the bacteriorhodopsin photocycle. J. Bioenerg. Biomembr. 24, 169179.Google Scholar
Lanyi, J. K., Duschl, A., Varo, G. & Zimanyi, L. (1990). Anion binding to the chloride pump bacteriorhodopsin, and its implications for the transport mechanism. FEBS Lett. 265, 16.Google Scholar
Le Courte, J., Oesterhelt, D. & Gerwert, K. (1992). The azide-effect in D96 → N/G mutated bacteriorhodopsins monitored by timeresolved FTIR-difference-spectroscopy. In Structures and Functions of Retinal Proteins (ed. Rigaud, J. L.), pp. 127130. Dourdan: Colloque INSERM.Google Scholar
Manor, D., Hasselbacher, C. A. & Spudich, J. L. (1988). Membrane potential modulates photocycling rates of bacterial rhodopsins. Biochemistry 27, 58435848.Google Scholar
Marwan, W. & Oesterhelt, D. (1987). Signal formation in the halobacterial photophobic response mediated by a fourth retinal protein (P480). J. Mol. Biol. 195, 333342.CrossRefGoogle ScholarPubMed
Marwan, W. & Oesterhelt, D. (1991). Light-induced release of the switch factor during photophobic responses of Halobacterium halobium. Naturzvissenschaften 78, 127129.Google Scholar
Marwan, W., Schäfer, W. & Oesterhelt, D. (1990). Signal transduction in Halobacterium depends on fumarate. EMBO J. 9, 355362.Google Scholar
Matsuno-Yagi, A. & Mukohata, Y. (1977). Two possible roles of bacteriorhodopsin: a comparative study of strains of Halobacterium halobium differing in pigmentation. Biochim. Biophys. Res. Commun. 78, 237243.Google Scholar
Minke, B., Hochstein, S. & Hillman, P. (1973). Early receptor potential: evidence for the existence of two thermally stable states in the barnacle visual pigment. J. Gen. Physiol. 62, 87104.Google Scholar
Mitchell, P. (1957). A general theory of membrane transport from studies of bacteria. Nature 180, 134136.Google Scholar
Mitchell, P. (1959). Structure and function in microorganisms. Biochem. Soc. Symposia 16, 7393.Google Scholar
Mitchell, P. (1991). Foundations of vectorial metabolism and osmochemistry. Biosci. Rep. 11, 297344.Google Scholar
Oesterhelt, D. & Marwan, W. (1987). Change of membrane potential is not a component of the photophobic transduction chain in Halobacterium halobium. J. Bacteriol. 169, 35153520.CrossRefGoogle Scholar
Oesterhelt, D. & Schuhmann, L. (1974). Reconstruction of bacteriorhodopsin. FEBS Lett. 44, 262265.Google Scholar
Oesterhelt, D. & Stoeckenius, W. (1971). Rhodopsin-like protein from the purplemembrane of Halobacterium halobium. Nature 233, 149152.Google Scholar
Oesterhelt, D. & Stoeckenius, W. (1973). Function of a new photoreceptor membrane. Proc. Natn. Acad. Sci. USA 70, 28532857.Google Scholar
Oesterhelt, D. & Tittor, J. (1989). Two pumps, one principle: light-driven ion transport in Halobacteria. Trends Biochem. Sci. 14, 5761.Google Scholar
Oesterhelt, D., Tittor, J. & Bamberg, E. (1992). A unifying concept for ion translocation by retinal proteins. J. Bioenerg. Biomembr. 24, 181191.CrossRefGoogle ScholarPubMed
Onsager, L. (1973). In Physics and Chemistry of Ice (ed. Whalley, E.), pp. 712. Ottawa: Royal Society of Canada.Google Scholar
Otomo, J., Tomioka, H. & Sasabe, H. (1992). Properties and the primary structure of a new halorhodopsin from halobacterial strain mex. Biochim. Biophys. Acta 1112, 713.Google Scholar
Otto, H. & Heyn, M. P. (1991). Between the ground- and M-state of bacteriorhodopsin the retinal transition dipole moment tilts out of the plane of the membrane by only 3°. FEBS Lett. 293, 111114.Google Scholar
Otto, H., Marti, T., Holz, M., Mogi, T., Khorana, H. G. & Heyn, M. P. (1989). Aspartic acid-96 is the internal proton donor in the reprotonation of the Schiff base of bacteriorhodopsin. Proc. Natn. Acad. Sci. USA 86, 92289232.Google Scholar
Ottolenghi, M. (1980). The photochemistry of rhodopsins. Adv. Photochem. 12, 97–20.Google Scholar
Pack, W. L. & Lidington, K. J. (1974). Fast electrical potential from a long-lived, long-wavelength photoproduct of fly visual pigment. J. Gen. Physiol. 63, 740756.Google Scholar
Papadopoulos, G., Dencher, N. A., Zaccai, G. & Büldt, G. (1990). Water molecules and exchangeable hydrogen ions at the active centre of bacteriorhodopsin localized by neutron diffraction. J. Mol. Biol. 214, 1519.Google Scholar
Repaske, D. R. & Adler, J. (1981). Change in the intracellular pH of Escherichia coli mediates the chemotactic response to certain attractants and repellents. J. Bacteriol. 145, 11961208.Google Scholar
Rothschild, K. J., He, Yi-WuSonar, S., Marti, T. & Khorana, H. G. (1992). Vibrational spectroscopy of bacteriorhodopsin mutants. Evidence that Thr-46 and Thr-89 form part of a transient network of hydrogen bonds. J. Biol. Chem. 267, 16151622.Google Scholar
Scharf, B., Engelhard, M. & Siebert, F. (1992). A carboxyl group is protonated during the photocycle of the photophobic receptor psR-II from Natronobacterium pharaonis. In Structures and Functions of Retinal Proteins (ed. Rigaud, J. L.), pp. 317320. Dourdan: Colloque INSERM.Google Scholar
Schertler, G. F. X., Lozier, R., Michel, H. & Oesterhelt, D. (1991). Chromophore motion during the bacteriorhodopsin photocycle. Polarized absorption spectroscopy of bacteriorhodopsin and its M-state in bacteriorhodopsin crystals. EMBO J. 10, 23532361.Google Scholar
Schöbert, B. & Lanyi, J. K. (1982). Halorhodopsin is a light-driven chloride pump. J. Biol. Chem. 257, 1030610313.Google Scholar
Skulachev, V. P. (1978). Membrane-linked energy buffering as the biological function of Na+/K+ gradient. FEBS Lett. 87, 171179.Google Scholar
Skulachev, V. P. (1982). The dual role of rhodopsin in vision: light-driven charge translocation and formation of long-lived photoproducts. FEBS Lett. 146, 244254.Google Scholar
Skulachev, V. P. (1988). Membrane Bioenergetics. Berlin: Springer-Verlag.Google Scholar
Skulachev, V. P. (1992). Rhodopsins: from ion pumps to specialized photoreceptors. In Structures and Functions of Retinal Proteins (ed. Rigaud, J. L.), pp. 229232. Dourdan: Colloque INSERM.Google Scholar
Skulachev, V. P. (1993). Bioenergetics of extreme halophiles. In: Archaea Book (ed. Kates, M.). Elsevier, Amsterdam (accepted).Google Scholar
Slonczewski, J. L., Macnab, R. M., Alger, J. R. & Castle, A. M. (1982). Effects of pH and repellent tactic stimuli of protein methylation levels in Escherichia coli. J. Bacteriol. 152, 384399.Google Scholar
Spudich, J. L. (1992). Phototransduction by sensory rhodopsin I. In Structures and Functions of Retinal Proteins (ed. Rigaud, J. L.), pp. 309312. Dourdan: Colloque INSERM.Google Scholar
Spudich, J. L. & Bogomolni, R. A. (1984). Mechanism of colour discrimination by a bacterial sensory rhodopsin. Nature 312, 509513.CrossRefGoogle ScholarPubMed
Spudich, J. L. & Bogomolni, R. A. (1988). Sensory rhodopsins of halobacteria. Ann. Rev. Biophys. Biophys. Chem. 17, 193215.Google Scholar
Spudich, J. L. & Bogomolni, R. A. (1992). Sensory rhodopsin I: Receptor activation and signal relay. J. Bioenerg. Biomembr. 24, 193199.Google Scholar
Stephenson, R. S. & Pak, W. L. (1980). Heterogenic components of a fast potential in Drosophila compound eye and their relation to visual pigment conversion. J. Gen. Physiol. 75, 353379.Google Scholar
Stoeckenius, W., Lozier, R. H. & Bogomolni, R. A. (1979). Bacteriorhodopsin and the purple membrane of Halobacteria. Biochim. Biophys. Acta 505, 215278.Google Scholar
Stoeckenius, W., Lozier, R. H. & Bogomolni, R. A. (1981). Bacteriorhodopsin photocycle and stoichiometry. In Chemiosmotic Circuits in Biological Membranes (ed. Skulachev, V. P. and Hinkle, P. C.), pp. 283309. Reading, MA: Addison-Wesley.Google Scholar
Stryer, L. (1991). Visual excitation and recovery. J. Biol. Chem. 266, 1071110714.Google Scholar
Subramaniam, S., Gerstein, M., Oesterhelt, D. & Henderson, R. (1993). Electron diffraction analysis of structural changes in the photocycle of bacteriorhodopsin. EMBO J. 12, 18.Google Scholar
Subramaniam, S., Marti, T., Rösselet, S. J., Rothschild, K. J. & Khorana, H. G. (1991). The reaction of hydroxylamine with bacteriorhodopsin studied using mutants that have altered photocycles: selective reactivity of different photointermediates. Proc. Natn. Acad. Sci. USA 88, 25832586.Google Scholar
Sucheta, A., Ackrell, B. A. C., Cochran, B. & Armstrong, F. A. (1992). Diode-like behaviour of a mitochondrial electron-transport enzyme. Nature 356, 361362.Google Scholar
Sundberg, S. A., Bogomolni, R. A. & Spudich, J. L. (1985). Selection and properties of phototaxis-deficient mutants of Halobacterium halobium. J. Bacteriol. 164, 282287.Google Scholar
Takahashi, T., Tomioka, H., Kamo, N. & Kobatake, Y. (1985 a). A photosystem other than PS370 also mediates the negative phototaxis of Halobacterium halobium. FEMS Microbiol. Lett. 28, 161164.Google Scholar
Takahashi, T., Watanabe, M., Kamo, N. & Kobatake, Y. (19856). Negative phototaxis from blue light and the role of third rhodopsin-like pigment in Halobacterium cutirubrum. Biophys. J. 48, 235240.Google Scholar
Tittor, J., Söll, C., Oesterhelt, D., Butt, H.-J. & Bamberg, E. (1989). A defective proton pump: point-mutated bacteriorhodopsin Asp-96 → Asn is fully reactivated by azide. EMBO J. 8, 34773482.Google Scholar
Tsuda, M., Nelson, B., Chang, C.-H., Govindjee, R. & Ebrey, T. G. (1985). Characterization of the chromophore of the third rhodopsin-like pigment of Halobacterium halobium and its photoproduct. Biophys. J. 47, 721724.Google Scholar
Varo, G. & Lanyi, J. K. (1989). Photoreactions of bacteriorhodopsin at acid pH. Biophys. J. 56, 11431151.Google Scholar
Wolff, E. H., Bogomolni, R. A., Scherrer, P., Hess, B. & Stoeckenius, W. (1986). Color discrimination in halobacteria: spectroscopic characterization of a second sensory receptor covering the blue-green region of the spectrum. Proc. Natn. Acad. Sci. USA 83, 72727276.Google Scholar