Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-mp689 Total loading time: 0 Render date: 2024-04-25T04:47:26.569Z Has data issue: false hasContentIssue false

Part II - Opsin Biology, Tools, and Technology Platform

Published online by Cambridge University Press:  28 April 2017

Krishnarao Appasani
Affiliation:
GeneExpression Systems, Inc., Massachusetts
Get access
Type
Chapter
Information
Optogenetics
From Neuronal Function to Mapping and Disease Biology
, pp. 77 - 166
Publisher: Cambridge University Press
Print publication year: 2017

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

Balashov, S.P., Imasheva, E.S., Boichenko, V.A., Antón, J., Wang, J.M., Lanyi, J.K., (2005). Xanthorhodopsin: a proton pump with a light-harvesting carotenoid antenna. Science 309, 20612064.CrossRefGoogle ScholarPubMed
Balashov, S.P., Imasheva, E.S., Dioumaev, A.K., Wang, J.M., Jung, K.-H., Lanyi, J.K., (2014). Light-driven Na+ pump from Gillisia limnaea: a high-affinity Na+ binding site is formed transiently in the photocycle. Biochem. (Mosc.) 53, 75497561.CrossRefGoogle ScholarPubMed
Balashov, S.P., Petrovskaya, L.E., Imasheva, E.S., Lukashev, E.P., Dioumaev, A.K., Wang, J.M., Sychev, S.V., Dolgikh, D.A., Rubin, A.B., Kirpichnikov, M.P., Lanyi, J.K., (2013). Breaking the carboxyl rule lysine 96 facilitates reprotonation of the Schiff base in the photocycle of a retinal protein from Exiguobacterium sibiricum. J. Biol. Chem. 288, 2125421265.CrossRefGoogle ScholarPubMed
Béjà, O., Aravind, L., Koonin, E.V., Suzuki, M.T., Hadd, A., Nguyen, L.P., Jovanovich, S.B., Gates, C.M., Feldman, R.A., Spudich, J.L., Spudich, E.N., DeLong, E.F., (2000). Bacterial rhodopsin: evidence for a new type of phototrophy in the sea. Science 289, 19021906.CrossRefGoogle ScholarPubMed
Bertsova, Y.V., Bogachev, A.V., Skulachev, V.P., (2015). Proteorhodopsin from Dokdonia sp. PRO95 is a light-driven Na+-pump. Biochem. Mosc. 80, 449454.CrossRefGoogle Scholar
Bogomolni, R.A., Spudich, J.L., (1982). Identification of a third rhodopsin-like pigment in phototactic Halobacterium halobium. Proc. Natl. Acad. Sci. 79, 62506254.CrossRefGoogle ScholarPubMed
de la Torre, J.R. Christianson, L.M., Béjà, O., Suzuki, M.T., Karl, D.M., Heidelberg, J., DeLong, E.F., (2003). Proteorhodopsin genes are distributed among divergent marine bacterial taxa. Proc. Natl. Acad. Sci. 100, 1283012835.CrossRefGoogle ScholarPubMed
DeLong, E.F., Karl, D.M., (2005). Genomic perspectives in microbial oceanography. Nature 437, 336342.CrossRefGoogle ScholarPubMed
Ernst, O.P., Lodowski, D.T., Elstner, M., Hegemann, P., Brown, L.S., Kandori, H., (2014). Microbial and animal rhodopsins: structures, functions, and molecular mechanisms. Chem. Rev. 114, 126163.CrossRefGoogle ScholarPubMed
Gordeliy, V.I., Labahn, J., Moukhametzianov, R., Efremov, R., Granzin, J., Schlesinger, R., Büldt, G., Savopol, T., Scheidig, A.J., Klare, J.P., Engelhard, M., (2002). Molecular basis of transmembrane signalling by sensory rhodopsin II–transducer complex. Nature 419, 484487.CrossRefGoogle ScholarPubMed
Greene, R.V., Lanyi, J.K., (1979). Proton movements in response to a light-driven electrogenic pump for sodium ions in Halobacterium halobium membranes. J. Biol. Chem. 254, 1098610994.CrossRefGoogle ScholarPubMed
Gushchin, I., Chervakov, P., Kuzmichev, P., Popov, A.N., Round, E., Borshchevskiy, V., Ishchenko, A., Petrovskaya, L., Chupin, V., Dolgikh, D.A., Arseniev, A.S., Kirpichnikov, M., Gordeliy, V., (2013). Structural insights into the proton pumping by unusual proteorhodopsin from nonmarine bacteria. Proc. Natl. Acad. Sci. 110, 1263112636.CrossRefGoogle Scholar
Gushchin, I., Shevchenko, V., Polovinkin, V., Borshchevskiy, V., Buslaev, P., Bamberg, E., Gordeliy, V., (2015a). Structure of the light-driven sodium pump KR2 and its implications for optogenetics. FEBS J. 283, 12321238.CrossRefGoogle ScholarPubMed
Gushchin, I., Shevchenko, V., Polovinkin, V., Kovalev, K., Alekseev, A., Round, E., Borshchevskiy, V., Balandin, T., Popov, A., Gensch, T., Fahlke, C., Bamann, C., Willbold, D., Büldt, G., Bamberg, E., Gordeliy, V., (2015b). Crystal structure of a light-driven sodium pump. Nat. Struct. Mol. Biol. 22, 390395.CrossRefGoogle Scholar
Han, X., Chow, B.Y., Zhou, H., Klapoetke, N.C., Chuong, A., Rajimehr, R., Yang, A., Baratta, M.V., Winkle, J., Desimone, R., Boyden, E.S., (2011). A high-light sensitivity optical neural silencer: development and application to optogenetic control of non-human primate cortex. Front. Syst. Neurosci. 5, 18.CrossRefGoogle ScholarPubMed
Henderson, R., Baldwin, J.M., Ceska, T.A., Zemlin, F., Beckmann, E., Downing, K.H., (1990). Model for the structure of bacteriorhodopsin based on high-resolution electron cryo-microscopy. J. Mol. Biol. 213, 899929.CrossRefGoogle Scholar
Inoue, K., Konno, M., Abe-Yoshizumi, R., Kandori, H., (2015). The role of the NDQ motif in sodium-pumping rhodopsins. Angew. Chem. 127, 1169811701.CrossRefGoogle Scholar
Inoue, K., Ono, H., Abe-Yoshizumi, R., Yoshizawa, S., Ito, H., Kogure, K., Kandori, H., (2013). A light-driven sodium ion pump in marine bacteria. Nat. Commun. 4, 1678.CrossRefGoogle ScholarPubMed
Ivanova, N., Rohde, C., Munk, C., Nolan, M., Lucas, S., Del Rio, T.G., Tice, H., Deshpande, S., Cheng, J.-F., Tapia, R., Han, C., Goodwin, L., Pitluck, S., Liolios, K., Mavromatis, K., Mikhailova, N., Pati, A., Chen, A., Palaniappan, K., Land, M., Hauser, L., Chang, Y.-J., Jeffries, C.D., Brambilla, E., Rohde, M., Göker, M., Tindall, B.J., Woyke, T., Bristow, J., Eisen, J.A., Markowitz, V., Hugenholtz, P., Kyrpides, N.C., Klenk, H.-P., Lapidus, A., (2011). Complete genome sequence of Truepera radiovictrix type strain (RQ-24T). Stand. Genomic Sci. 4, 9199.CrossRefGoogle Scholar
Kanada, S., Takeguchi, Y., Murakami, M., Ihara, K., Kouyama, T., (2011). Crystal structures of an O-like blue form and an anion-free yellow form of pharaonis halorhodopsin. J. Mol. Biol. 413, 162176.CrossRefGoogle Scholar
Kato, H.E., Inoue, K., Abe-Yoshizumi, R., Kato, Y., Ono, H., Konno, M., Hososhima, S., Ishizuka, T., Hoque, M.R., Kunitomo, H., Ito, J., Yoshizawa, S., Yamashita, K., Takemoto, M., Nishizawa, T., Taniguchi, R., Kogure, K., Maturana, A.D., Iino, Y., Yawo, H., Ishitani, R., Kandori, H., Nureki, O., (2015). Structural basis for Na+ transport mechanism by a light-driven Na+ pump. Nature 521, 4853.CrossRefGoogle ScholarPubMed
Kato, Y., Inoue, K., Kandori, H., (2015). Kinetic analysis of H+–Na+ selectivity in a light-driven Na+-pumping rhodopsin. J. Phys. Chem. Lett. 6, 51115115.CrossRefGoogle Scholar
Kolbe, M., Besir, H., Essen, L.-O., Oesterhelt, D., (2000). Structure of the light-driven chloride pump halorhodopsin at 1.8 Å resolution. Science 288, 13901396.CrossRefGoogle ScholarPubMed
Konno, M., Kato, Y., Kato, H.E., Inoue, K., Nureki, O., Kandori, H., (2016). Mutant of a light-driven sodium ion pump can transport cesium ions. J. Phys. Chem. Lett. 7, 5155.CrossRefGoogle ScholarPubMed
Kouyama, T., Kanada, S., Takeguchi, Y., Narusawa, A., Murakami, M., Ihara, K., (2010). Crystal structure of the light-driven chloride pump halorhodopsin from Natronomonas pharaonis. J. Mol. Biol. 396, 564579.CrossRefGoogle ScholarPubMed
Kwon, Y.M., Kim, S.-Y., Jung, K.-H., Kim, S.-J., (2016). Diversity and functional analysis of light-driven pumping rhodopsins in marine Flavobacteria. Microbiologyopen 5, 212223.CrossRefGoogle ScholarPubMed
Lindley, E.V., MacDonald, R.E., (1979). A second mechanism for sodium extrusion in Halobacterium halobium: a light-driven sodium pump. Biochem. Biophys. Res. Commun. 88, 491499.CrossRefGoogle ScholarPubMed
Luecke, H., Schobert, B., Richter, H.-T., Cartailler, J.-P., Lanyi, J.K., (1999). Structure of bacteriorhodopsin at 1.55 Å resolution 1. J. Mol. Biol. 291, 899911.CrossRefGoogle Scholar
Luecke, H., Schobert, B., Stagno, J., Imasheva, E.S., Wang, J.M., Balashov, S.P., Lanyi, J.K., (2008). Crystallographic structure of xanthorhodopsin, the light-driven proton pump with a dual chromophore. Proc. Natl. Acad. Sci. 105, 1656116565.CrossRefGoogle ScholarPubMed
MacDonald, R.E., Greene, R.V., Clark, R.D., Lindley, E.V., (1979). Characterization of the light-driven sodium pump of Halobacterium halobium. J Biol Chem 254, 1183111838.CrossRefGoogle ScholarPubMed
Mongodin, E.F., Nelson, K.E., Daugherty, S., DeBoy, R.T., Wister, J., Khouri, H., Weidman, J., Walsh, D.A., Papke, R.T., Perez, G.S., Sharma, A.K., Nesbø, C.L., MacLeod, D., Bapteste, E., Doolittle, W.F., Charlebois, R.L., Legault, B., Rodriguez-Valera, F., (2005). The genome of Salinibacter ruber: convergence and gene exchange among hyperhalophilic bacteria and archaea. Proc. Natl. Acad. Sci. 102, 1814718152.CrossRefGoogle ScholarPubMed
Moukhametzianov, R., Klare, J.P., Efremov, R., Baeken, C., Göppner, A., Labahn, J., Engelhard, M., Büldt, G., Gordeliy, V.I., (2006). Development of the signal in sensory rhodopsin and its transfer to the cognate transducer. Nature 440, 115119.CrossRefGoogle Scholar
Nagel, G., Szellas, T., Huhn, W., Kateriya, S., Adeishvili, N., Berthold, P., Ollig, D., Hegemann, P., Bamberg, E., (2003). Channelrhodopsin-2, a directly light-gated cation-selective membrane channel. Proc. Natl. Acad. Sci. 100, 1394013945.CrossRefGoogle ScholarPubMed
Nakanishi, M., Meirelles, P., Suzuki, R., Takatani, N., Mino, S., Suda, W., Oshima, K., Hattori, M., Ohkuma, M., Hosokawa, M., Miyashita, K., Thompson, F.L., Niwa, A., Sawabe, T., Sawabe, T., (2014). Draft genome sequences of marine flavobacterium nonlabens strains NR17, NR24, NR27, NR32, NR33, and Ara13. Genome Announc. 2, e01165-14.CrossRefGoogle Scholar
Oesterhelt, D., Stoeckenius, W., (1971). Rhodopsin-like protein from the purple membrane of Halobacterium halobium. Nature 233, 149152.Google ScholarPubMed
O’Malley, M.A., (2007). Exploratory experimentation and scientific practice: metagenomics and the proteorhodopsin case. Hist. Philos. Life Sci. 29, 337360.Google ScholarPubMed
Riedel, T., Held, B., Nolan, M., Lucas, S., Lapidus, A., Tice, H., Del Rio, T.G., Cheng, J.-F., Han, C., Tapia, R., Goodwin, L.A., Pitluck, S., Liolios, K., Mavromatis, K., Pagani, I., Ivanova, N., Mikhailova, N., Pati, A., Chen, A., Palaniappan, K., Land, M., Rohde, M., Tindall, B.J., Detter, J.C., Göker, M., Bristow, J., Eisen, J.A., Markowitz, V., Hugenholtz, P., Kyrpides, N.C., Klenk, H.-P., Woyke, T., (2012). Genome sequence of the Antarctic rhodopsins-containing flavobacterium Gillisia limnaea type strain (R-8282(T)). Stand. Genomic Sci. 7, 107119.CrossRefGoogle Scholar
Schobert, B., Lanyi, J.K., (1982). Halorhodopsin is a light-driven chloride pump. J. Biol. Chem. 257, 1030610313.CrossRefGoogle ScholarPubMed
Singh, A., Kumar Jangir, P., Sharma, R., Singh, A., Kumar Pinnaka, A., Shivaji, S., (2013). Draft genome sequence of Indibacter alkaliphilus strain LW1T, isolated from Lonar Lake, a haloalkaline lake in the Buldana District of Maharashtra, India. Genome Announc. 1, e00515-13.CrossRefGoogle Scholar
Stoeckenius, W., Lozier, R.H., Bogomolni, R.A., (1979). Bacteriorhodopsin and the purple membrane of halobacteria. Biochim. Biophys. Acta BBA Rev. Bioenerg. 505, 215278.CrossRefGoogle Scholar
Wickstrand, C., Dods, R., Royant, A., Neutze, R., (2015). Bacteriorhodopsin: would the real structural intermediates please stand up? Biochim. Biophys. Acta 1850, 536553.CrossRefGoogle ScholarPubMed
Yoshizawa, S., Kumagai, Y., Kim, H., Ogura, Y., Hayashi, T., Iwasaki, W., DeLong, E.F., Kogure, K., (2014). Functional characterization of Flavobacteria rhodopsins reveals a unique class of light-driven chloride pump in bacteria. Proc. Natl. Acad. Sci. 111, 67326737.CrossRefGoogle ScholarPubMed
Zhao, S., Cunha, C., Zhang, F., Liu, Q., Gloss, B., Deisseroth, K., Augustine, G.J., Feng, G., (2008). Improved expression of halorhodopsin for light-induced silencing of neuronal activity. Brain Cell Biol. 36, 141154.CrossRefGoogle ScholarPubMed

References

Ahrens, M.B., Orger, M.B., Robson, D.N., Li, J.M., Keller, P.J., 2013. Whole-brain functional imaging at cellular resolution using light-sheet microscopy. Nat. Methods 10, 413420.CrossRefGoogle ScholarPubMed
Berndt, A., Yizhar, O., Gunaydin, L.A., Hegemann, P., Deisseroth, K., 2009. Bi-stable neural state switches. Nat. Neurosci. 12, 229234.CrossRefGoogle ScholarPubMed
Boyden, E.S., Zhang, F., Bamberg, E., Nagel, G., Deisseroth, K., 2005. Millisecond-timescale, genetically targeted optical control of neural activity. Nat. Neurosci. 8, 12631268.CrossRefGoogle ScholarPubMed
Cao, H., Gu, L., Mohanty, S.K., Chiao, J.-C., 2013. An integrated μLED optrode for optogenetic stimulation and electrical recording. IEEE Trans. Biomed. Eng. 60, 225229.CrossRefGoogle ScholarPubMed
Chow, B.Y., Han, X., Dobry, A.S., Qian, X., Chuong, A.S., Li, M., Henninger, M.A., Belfort, G.M., Lin, Y., Monahan, P.E., Boyden, E.S., 2010. High-performance genetically targetable optical neural silencing by light-driven proton pumps. Nature 463, 98102.CrossRefGoogle Scholar
Clancy, K.B., Schnepel, P., Rao, A.T., Feldman, D.E., 2015. Structure of a single whisker representation in layer 2 of mouse somatosensory cortex. J. Neurosci. 35, 39463958.CrossRefGoogle ScholarPubMed
Deisseroth, K., Feng, G., Majewska, A.K., Miesenböck, G., Ting, A., Schnitzer, M.J., 2006. Next-generation optical technologies for illuminating genetically targeted brain circuits. J. Neurosci. 26, 1038010386.CrossRefGoogle Scholar
Drake, K.L., Wise, K.D., Farraye, J., Anderson, D.J., BeMent, S.L., 1988. Performance of planar multisite microprobes in recording extracellular single-unit intracortical activity. IEEE Trans. Biomed. Eng. 35, 719732.CrossRefGoogle ScholarPubMed
Eggermann, E., Kremer, Y., Crochet, S., Petersen, C.C.H., 2014. Cholinergic signals in mouse barrel cortex during active whisker sensing. Cell Rep. 9, 16541660.CrossRefGoogle ScholarPubMed
Favre-Bulle, I.A., Preece, D., Nieminen, T.A., Heap, L.A., Scott, E.K., Rubinsztein-Dunlop, H., 2015. Scattering of sculpted light in intact brain tissue, with implications for optogenetics. Sci. Rep. 5, 11501.CrossRefGoogle ScholarPubMed
Fendyur, A., Spira, M.E., 2012. Toward on-chip, in-cell recordings from cultured cardiomyocytes by arrays of gold mushroom-shaped microelectrodes. Front. Neuroengineering 5, 21.CrossRefGoogle ScholarPubMed
Galvani, L., Volta, A., Zambelli, J., 1791. Aloysii Galvani De Viribus Electricitatis in Motu Musculari Commentarius. Bononiae: Ex Typographia Instituti Scientiarium.CrossRefGoogle Scholar
Gentet, L.J., Kremer, Y., Taniguchi, H., Huang, Z.J., Staiger, J.F., Petersen, C.C.H., 2012. Unique functional properties of somatostatin-expressing GABAergic neurons in mouse barrel cortex. Nat. Neurosci. 15, 607612.CrossRefGoogle ScholarPubMed
Ghosh, K.K., Burns, L.D., Cocker, E.D., Nimmerjahn, A., Ziv, Y., Gamal, A.E., Schnitzer, M.J., 2011. Miniaturized integration of a fluorescence microscope. Nat. Methods 8, 871878.CrossRefGoogle ScholarPubMed
Gradinaru, V., Thompson, K.R., Zhang, F., Mogri, M., Kay, K., Schneider, M.B., Deisseroth, K., 2007. Targeting and readout strategies for fast optical neural control in vitro and in vivo. J. Neurosci. 27, 1423114238.CrossRefGoogle ScholarPubMed
Grinvald, A., Cohen, L.B., Lesher, S., Boyle, M.B., 1981. Simultaneous optical monitoring of activity of many neurons in invertebrate ganglia using a 124-element photodiode array. J. Neurophysiol. 45, 829840.CrossRefGoogle ScholarPubMed
Guo, Z.V., Hart, A.C., Ramanathan, S., 2009. Optical interrogation of neural circuits in Caenorhabditis elegans. Nat. Methods 6, 891896.CrossRefGoogle ScholarPubMed
Guo, Z.V., Hires, S.A., Li, N., O’Connor, D.H., Komiyama, T., Ophir, E., Huber, D., Bonardi, C., Morandell, K., Gutnisky, D., Peron, S., Xu, N., Cox, J., Svoboda, K., 2014. Procedures for behavioral experiments in head-fixed mice. PLoS One 9, e88678.CrossRefGoogle ScholarPubMed
Han, X., Chow, B.Y., Zhou, H., Klapoetke, N.C., Chuong, A., Rajimehr, R., Yang, A., Baratta, M.V., Winkle, J., Desimone, R., Boyden, E.S., 2011. A high-light sensitivity optical neural silencer: development and application to optogenetic control of non-human primate cortex. Front. Syst. Neurosci. 5, 18.CrossRefGoogle ScholarPubMed
Hegemann, P., Möglich, A., 2011. Channelrhodopsin engineering and exploration of new optogenetic tools. Nat. Methods 8, 3942.CrossRefGoogle ScholarPubMed
Henze, D.A., Borhegyi, Z., Csicsvari, J., Mamiya, A., Harris, K.D., Buzsáki, G., 2000. Intracellular features predicted by extracellular recordings in the hippocampus in vivo. J. Neurophysiol. 84, 390400.CrossRefGoogle ScholarPubMed
Hochbaum, D.R., Zhao, Y., Farhi, S.L., Klapoetke, N., Werley, C.A., Kapoor, V., Zou, P., Kralj, J.M., Maclaurin, D., Smedemark-Margulies, N., Saulnier, J.L., Boulting, G.L., Straub, C., Cho, Y.K., Melkonian, M., Wong, G.K.-S., Harrison, D.J., Murthy, V.N., Sabatini, B.L., Boyden, E.S., Campbell, R.E., Cohen, A.E., 2014. All-optical electrophysiology in mammalian neurons using engineered microbial rhodopsins. Nat. Methods 11, 825833.CrossRefGoogle ScholarPubMed
Huber, D., Petreanu, L., Ghitani, N., Ranade, S., Hromádka, T., Mainen, Z., Svoboda, K., 2008. Sparse optical microstimulation in barrel cortex drives learned behaviour in freely moving mice. Nature 451, 6164.CrossRefGoogle Scholar
Iwai, Y., Honda, S., Ozeki, H., Hashimoto, M., Hirase, H., 2011. A simple head-mountable LED device for chronic stimulation of optogenetic molecules in freely moving mice. Neurosci. Res. 70, 124127.CrossRefGoogle ScholarPubMed
Jouhanneau, J.-S., Ferrarese, L., Estebanez, L., Audette, N.J., Brecht, M., Barth, A.L., Poulet, J.F.A., 2014. Cortical fosGFP expression reveals broad receptive field excitatory neurons targeted by POm. Neuron 84, 10651078.CrossRefGoogle ScholarPubMed
Katz, Y., Yizhar, O., Staiger, J., Lampl, I., 2013. Optopatcher – an electrode holder for simultaneous intracellular patch-clamp recording and optical manipulation. J. Neurosci. Methods 214, 113117.CrossRefGoogle ScholarPubMed
Kim, T., McCall, J.G., Jung, Y.H., Huang, X., Siuda, E.R., Li, Y., Song, J., Song, Y.M., Pao, H.A., Kim, R.-H., Lu, C., Lee, S.D., Song, I.-S., Shin, G., Al-Hasani, R., Kim, S., Tan, M.P., Huang, Y., Omenetto, F.G., Rogers, J.A., Bruchas, M.R., 2013. Injectable, cellular-scale optoelectronics with applications for wireless optogenetics. Science 340, 211216.CrossRefGoogle ScholarPubMed
Kipke, D.R., Shain, W., Buzsáki, G., Fetz, E., Henderson, J.M., Hetke, J.F., Schalk, G., 2008. Advanced neurotechnologies for chronic neural interfaces: new horizons and clinical opportunities. J. Neurosci. 28, 1183011838.CrossRefGoogle ScholarPubMed
Klapoetke, N.C., Murata, Y., Kim, S.S., Pulver, S.R., Birdsey-Benson, A., Cho, Y.K., Morimoto, T.K., Chuong, A.S., Carpenter, E.J., Tian, Z., Wang, J., Xie, Y., Yan, Z., Zhang, Y., Chow, B.Y., Surek, B., Melkonian, M., Jayaraman, V., Constantine-Paton, M., Wong, G.K.-S., Boyden, E.S., 2014. Independent optical excitation of distinct neural populations. Nat. Methods 11, 338346.CrossRefGoogle ScholarPubMed
Kwon, K.Y., Lee, H.-M., Ghovanloo, M., Weber, A., Li, W., 2015. Design, fabrication, and packaging of an integrated, wirelessly-powered optrode array for optogenetics application. Front. Syst. Neurosci. 9, 69.CrossRefGoogle Scholar
Lee, D., Shtengel, G., Osborne, J.E., Lee, A.K., 2014. Anesthetized- and awake-patched whole-cell recordings in freely moving rats using UV-cured collar-based electrode stabilization. Nat. Protoc. 9, 27842795.CrossRefGoogle ScholarPubMed
Lima, S.Q., Miesenböck, G., 2005. Remote control of behavior through genetically targeted photostimulation of neurons. Cell 121, 141152.CrossRefGoogle ScholarPubMed
Lin, J.Y., Knutsen, P.M., Muller, A., Kleinfeld, D., Tsien, R.Y., 2013. ReaChR: a red-shifted variant of channelrhodopsin enables deep transcranial optogenetic excitation. Nat. Neurosci. 16, 14991508.CrossRefGoogle ScholarPubMed
Lin, J.Y., Lin, M.Z., Steinbach, P., Tsien, R.Y., 2009. Characterization of engineered channelrhodopsin variants with improved properties and kinetics. Biophys. J. 96, 18031814.CrossRefGoogle ScholarPubMed
Manita, S., Suzuki, T., Homma, C., Matsumoto, T., Odagawa, M., Yamada, K., Ota, K., Matsubara, C., Inutsuka, A., Sato, M., Ohkura, M., Yamanaka, A., Yanagawa, Y., Nakai, J., Hayashi, Y., Larkum, M.E., Murayama, M., 2015. A top-down cortical circuit for accurate sensory perception. Neuron 86, 13041316.CrossRefGoogle ScholarPubMed
McAlinden, N., Gu, E., Dawson, M.D., Sakata, S., Mathieson, K., 2015. Optogenetic activation of neocortical neurons in vivo with a sapphire-based micro-scale LED probe. Front. Neural Circuits 9, 25.CrossRefGoogle ScholarPubMed
Miesenböck, G., De Angelis, D.A., Rothman, J.E., 1998. Visualizing secretion and synaptic transmission with pH-sensitive green fluorescent proteins. Nature 394, 192195.CrossRefGoogle Scholar
Miesenböck, G., Rothman, J.E., 1997. Patterns of synaptic activity in neural networks recorded by light emission from synaptolucins. Proc. Natl. Acad. Sci. U. S. A. 94, 34023407.CrossRefGoogle ScholarPubMed
Montgomery, K.L., Yeh, A.J., Ho, J.S., Tsao, V., Mohan Iyer, S., Grosenick, L., Ferenczi, E.A., Tanabe, Y., Deisseroth, K., Delp, S.L., Poon, A.S.Y., 2015. Wirelessly powered, fully internal optogenetics for brain, spinal and peripheral circuits in mice. Nat. Methods 12, 969974.CrossRefGoogle ScholarPubMed
Muñoz, W., Tremblay, R., Rudy, B., 2014. Channelrhodopsin-assisted patching: in vivo recording of genetically and morphologically identified neurons throughout the brain. Cell Rep. 9, 23042316.CrossRefGoogle ScholarPubMed
Musall, S., von der Behrens, W., Mayrhofer, J.M., Weber, B., Helmchen, F., Haiss, F., 2014. Tactile frequency discrimination is enhanced by circumventing neocortical adaptation. Nat. Neurosci. 17, 15671573.CrossRefGoogle ScholarPubMed
Nakai, J., Ohkura, M., Imoto, K., 2001. A high signal-to-noise Ca2+ probe composed of a single green fluorescent protein. Nat. Biotechnol. 19, 137141.CrossRefGoogle Scholar
Nakamura, S., Baratta, M.V., Pomrenze, M.B., Dolzani, S.D., Cooper, D.C., 2012. High fidelity optogenetic control of individual prefrontal cortical pyramidal neurons in vivo. F1000Research 1, 7.CrossRefGoogle ScholarPubMed
Ollerenshaw, D.R., Zheng, H.J.V., Millard, D.C., Wang, Q., Stanley, G.B., 2014. The adaptive trade-off between detection and discrimination in cortical representations and behavior. Neuron 81, 11521164.CrossRefGoogle ScholarPubMed
Packer, A.M., Russell, L.E., Dalgleish, H.W.P., Häusser, M., 2015. Simultaneous all-optical manipulation and recording of neural circuit activity with cellular resolution in vivo. Nat. Methods 12, 140146.CrossRefGoogle Scholar
Pala, A., Petersen, C.C.H., 2015. In vivo measurement of cell-type-specific synaptic connectivity and synaptic transmission in layer 2/3 mouse barrel cortex. Neuron 85, 6875.CrossRefGoogle ScholarPubMed
Reutsky-Gefen, I., Golan, L., Farah, N., Schejter, A., Tsur, L., Brosh, I., Shoham, S., 2013. Holographic optogenetic stimulation of patterned neuronal activity for vision restoration. Nat. Commun. 4, 1509.CrossRefGoogle ScholarPubMed
Royer, S., Zemelman, B.V., Barbic, M., Losonczy, A., Buzsáki, G., Magee, J.C., 2010. Multi-array silicon probes with integrated optical fibers: light-assisted perturbation and recording of local neural circuits in the behaving animal. Eur. J. Neurosci. 31, 22792291.CrossRefGoogle Scholar
Rubehn, B., Wolff, S.B.E., Tovote, P., Lüthi, A., Stieglitz, T., 2013. A polymer-based neural microimplant for optogenetic applications: design and first in vivo study. Lab. Chip 13, 579588.CrossRefGoogle ScholarPubMed
Ruiz, O., Lustig, B.R., Nassi, J.J., Cetin, A., Reynolds, J.H., Albright, T.D., Callaway, E.M., Stoner, G.R., Roe, A.W., 2013. Optogenetics through windows on the brain in the nonhuman primate. J. Neurophysiol. 110, 14551467.CrossRefGoogle ScholarPubMed
Sato, M., Ito, M., Nagase, M., Sugimura, Y.K., Takahashi, Y., Watabe, A.M., Kato, F., 2015. The lateral parabrachial nucleus is actively involved in the acquisition of fear memory in mice. Mol. Brain 8, 22.CrossRefGoogle ScholarPubMed
Schiemann, J., Puggioni, P., Dacre, J., Pelko, M., Domanski, A., van Rossum, M.C.W., Duguid, I., 2015. Cellular mechanisms underlying behavioral state-dependent bidirectional modulation of motor cortex output. Cell Rep. 11, 13191330.CrossRefGoogle ScholarPubMed
Siegle, J.H., Carlen, M., Meletis, K., Tsai, L.-H., Moore, C.I., Ritt, J., 2011. Chronically implanted hyperdrive for cortical recording and optogenetic control in behaving mice. Conf. Proc. IEEE Eng. Med. Biol. Soc. 2011, 75297532.Google ScholarPubMed
Stark, E., Koos, T., Buzsáki, G., 2012. Diode probes for spatiotemporal optical control of multiple neurons in freely moving animals. J. Neurophysiol. 108, 349363.CrossRefGoogle ScholarPubMed
Sudo, Y., Okazaki, A., Ono, H., Yagasaki, J., Sugo, S., Kamiya, M., Reissig, L., Inoue, K., Ihara, K., Kandori, H., Takagi, S., Hayashi, S., 2013. A blue-shifted light-driven proton pump for neural silencing. J. Biol. Chem. 288, 2062420632.CrossRefGoogle ScholarPubMed
Tolhurst, D.J., Smyth, D., Thompson, I.D., 2009. The sparseness of neuronal responses in ferret primary visual cortex. J. Neurosci. 29, 23552370.CrossRefGoogle Scholar
Wang, J., Wagner, F., Borton, D.A., Zhang, J., Ozden, I., Burwell, R.D., Nurmikko, A.V., van Wagenen, R., Diester, I., Deisseroth, K., 2012. Integrated device for combined optical neuromodulation and electrical recording for chronic in vivo applications. J. Neural Eng. 9, 016001.CrossRefGoogle Scholar
Wang, Y., Gong, Q., Li, Y.Y., Li, A.Z., Zhang, Y.G., Cao, C.F., Xu, H.X., Cui, J., Gao, J.J., 2015. A wireless remote high-power laser device for optogenetic experiments. Laser Phys. 25, 045601.CrossRefGoogle Scholar
Wentz, C.T., Bernstein, J.G., Monahan, P., Guerra, A., Rodriguez, A., Boyden, E.S., 2011. A wirelessly powered and controlled device for optical neural control of freely-behaving animals. J. Neural Eng. 8, 046021.CrossRefGoogle ScholarPubMed
Yoon, I., Hamaguchi, K., Borzenets, I.V., Finkelstein, G., Mooney, R., Donald, B.R., 2013. Intracellular neural recording with pure carbon nanotube probes. PLoS One 8, e65715.CrossRefGoogle ScholarPubMed
Zemelman, B.V., Lee, G.A., Ng, M., Miesenböck, G., 2002. Selective photostimulation of genetically chARGed neurons. Neuron 33, 1522.CrossRefGoogle Scholar
Zhang, F., Prigge, M., Beyrière, F., Tsunoda, S.P., Mattis, J., Yizhar, O., Hegemann, P., Deisseroth, K., 2008. Red-shifted optogenetic excitation: a tool for fast neural control derived from Volvox carteri. Nat. Neurosci. 11, 631633.CrossRefGoogle Scholar
Zhang, J., Laiwalla, F., Kim, J.A., Urabe, H., Van Wagenen, R., Song, Y.-K., Connors, B.W., Zhang, F., Deisseroth, K., Nurmikko, A.V., 2009. Integrated device for optical stimulation and spatiotemporal electrical recording of neural activity in light-sensitized brain tissue. J. Neural Eng. 6, 055007.CrossRefGoogle ScholarPubMed
Zorzos, A.N., Boyden, E.S., Fonstad, C.G., 2010. Multiwaveguide implantable probe for light delivery to sets of distributed brain targets. Opt. Lett. 35, 41334135.CrossRefGoogle ScholarPubMed
Zorzos, A.N., Scholvin, J., Boyden, E.S., Fonstad, C.G., 2012. Three-dimensional multiwaveguide probe array for light delivery to distributed brain circuits. Opt. Lett. 37, 48414843.CrossRefGoogle Scholar

References

Boyden, ES, Zhang, F, Bamberg, E, Nagel, G, Deisseroth, K (2005) Millisecond-timescale, genetically targeted optical control of neural activity. Nat Neurosci 8:12631268.CrossRefGoogle ScholarPubMed
Callaway, EM, Katz, LC (1990) Emergence and refinement of clustered horizontal connections in cat striate cortex. J Neurosci 10:11341153.CrossRefGoogle ScholarPubMed
Callaway, EM, Katz, LC (1991) Effects of binocular deprivation on the development of clustered horizontal connections in cat striate cortex. Proc Natl Acad Sci U S A 88:745749.CrossRefGoogle ScholarPubMed
Cohan, CS, Kater, SB (1986) Suppression of neurite elongation and growth cone motility by electrical activity. Science 232:16381640.CrossRefGoogle Scholar
Cohan, CS, Connor, JA, Kater, SB (1987) Electrically and chemically mediated increases in intracellular calcium in neuronal growth cones. J Neurosci 7:35883599.CrossRefGoogle ScholarPubMed
Durack, JC, Katz, LC (1996) Development of horizontal projections in layer 2/3 of ferret visual cortex. Cereb Cortex 6:178183.CrossRefGoogle ScholarPubMed
Fields, RD, Neale, EA, Nelson, PG (1990) Effects of patterned electrical activity on neurite outgrowth from mouse sensory neurons. J Neurosci 10:29502964.CrossRefGoogle ScholarPubMed
Gilbert, CD, Wiesel, TN (1989) Columnar specificity of intrinsic horizontal and corticocortical connections in cat visual cortex. J Neurosci 9:24322442.CrossRefGoogle Scholar
Gomez, TM, Spitzer, NC (1999) In vivo regulation of axon extension and pathfinding by growth-cone calcium transients. Nature 397:350355.CrossRefGoogle Scholar
Huberman, AD, Feller, MB, Chapman, B (2008) Mechanisms underlying development of visual maps and receptive fields. Annu Rev Neurosci 31:479509.CrossRefGoogle ScholarPubMed
Ibarretxe, G, Perrais, D, Jaskolski, F, Vimeney, A, Mulle, C (2007) Fast regulation of axonal growth cone motility by electrical activity. J Neurosci 27:76847695.CrossRefGoogle Scholar
Itoh, K, Ozaki, M, Stevens, B, Fields, RD (1997) Activity-dependent regulation of N-cadherin in DRG neurons: differential regulation of N-cadherin, NCAM, and L1 by distinct patterns of action potentials. J Neurobiol 33:735748.3.0.CO;2-A>CrossRefGoogle Scholar
Katz, LC, Shatz, CJ (1996) Synaptic activity and the construction of cortical circuits. Science 274:11331138.CrossRefGoogle ScholarPubMed
Lowel, S, Singer, W (1992) Selection of intrinsic horizontal connections in the visual cortex by correlated neuronal activity. Science 255:209212.CrossRefGoogle Scholar
Malyshevskaya, O, Shiraishi, Y, Kimura, F, Yamamoto, N (2013) Role of electrical activity in horizontal axon growth in the developing cortex: a time-lapse study using optogenetic stimulation. PLoS One 8:e82954.CrossRefGoogle ScholarPubMed
Nagel, G, Brauner, M, Liewald, JF, Adeishvili, N, Bamberg, E, Gottschalk, A (2005) Light activation of channelrhodopsin-2 in excitable cells of Caenorhabditis elegans triggers rapid behavioral responses. Curr Biol 15:22792284.CrossRefGoogle ScholarPubMed
Nirenberg, S, Pandarinath, C (2012) Retinal prosthetic strategy with the capacity to restore normal vision. Proc Natl Acad Sci U S A 109:1501215017.CrossRefGoogle ScholarPubMed
Ruthazer, ES, Stryker, MP (1996) The role of activity in the development of long-range horizontal connections in area 17 of the ferret. J Neurosci 16:72537269.CrossRefGoogle ScholarPubMed
Trachtenberg, JT, Chen, BE, Knott, GW, Feng, G, Sanes, JR, Welker, E, Svoboda, K (2002) Long-term in vivo imaging of experience-dependent synaptic plasticity in adult cortex. Nature 420:788794.CrossRefGoogle ScholarPubMed
Uesaka, N, Hayano, Y, Yamada, A, Yamamoto, N (2007) Interplay between laminar specificity and activity-dependent mechanisms of thalamocortical axon branching. J Neurosci 27:52155223.CrossRefGoogle ScholarPubMed
Uesaka, N, Hirai, S, Maruyama, T, Ruthazer, ES, Yamamoto, N (2005) Activity dependence of cortical axon branch formation: a morphological and electrophysiological study using organotypic slice cultures. J Neurosci 25:19.CrossRefGoogle ScholarPubMed
Yamamoto, N, Kurotani, T, Toyama, K (1989) Neural connections between the lateral geniculate nucleus and visual cortex in vitro. Science 245:192194.CrossRefGoogle Scholar
Yamamoto, N, Yamada, K, Kurotani, T, Toyama, K (1992) Laminar specificity of extrinsic cortical connections studied in coculture preparations. Neuron 9:217228.CrossRefGoogle ScholarPubMed
Yuste, R, Peinado, A, Katz, LC (1992) Neuronal domains in developing neocortex. Science 257:665669.CrossRefGoogle ScholarPubMed

References

Ausubel, F.M., Kingston, R.E., Seidman, F.G., et al. (ed.) (1995). Current Protocols in Molecular Biology, New York, NY: John Wiley and Sons.Google Scholar
Baarlink, C., Wang, H. and Grosse, R. (2013). Nuclear actin network assembly by formins regulates the SRF coactivator MAL. Science, 340, 864867.CrossRefGoogle ScholarPubMed
Belle, A., Tanay, A., Bitincka, L., et al. (2006). Quantification of protein half-lives in the budding yeast proteome. Proc Natl Acad Sci U S A, 103, 1300413009.CrossRefGoogle ScholarPubMed
Bonger, K.M., Rakhit, R., Payumo, A.Y., et al. (2014). General method for regulating protein stability with light. ACS Chem Biol, 9, 111115.CrossRefGoogle ScholarPubMed
Chong, Y.T., Koh, J.L., Friesen, H., et al. (2015). Yeast proteome dynamics from single cell imaging and automated analysis. Cell, 161, 14131424.CrossRefGoogle Scholar
Chudakov, D.M., Matz, M.V., Lukyanov, S., et al. (2010). Fluorescent proteins and their applications in imaging living cells and tissues. Physiol Rev, 90, 11031163.CrossRefGoogle ScholarPubMed
Deisseroth, K., Feng, G., Majewska, A.K., et al. (2006). Next-generation optical technologies for illuminating genetically targeted brain circuits. J Neurosci, 26, 1038010386.CrossRefGoogle ScholarPubMed
Delacour, Q., Li, C., Plamont, M.A., et al. (2015). Light-activated proteolysis for the spatiotemporal control of proteins. ACS Chem Biol, 10, 16431647.CrossRefGoogle ScholarPubMed
Descenzo, R.A. and Minocha, S.C. (1993). Modulation of cellular polyamines in tobacco by transfer and expression of mouse ornithine decarboxylase cDNA. Plant Mol Biol, 22, 113127.CrossRefGoogle ScholarPubMed
Gautier, A., Gauron, C., Volovitch, M., et al. (2014). How to control proteins with light in living systems. Nat Chem Biol, 10, 533541.CrossRefGoogle ScholarPubMed
Ghaemmaghami, S., Huh, W.K., Bower, K., et al. (2003). Global analysis of protein expression in yeast. Nature, 425, 737741.CrossRefGoogle ScholarPubMed
Ghoda, L., Van Daalen Wetters, T., Macrae, M., et al. (1989). Prevention of rapid intracellular degradation of ODC by a carboxyl-terminal truncation. Science, 243, 14931495.CrossRefGoogle Scholar
Goldberg, A.L. (2007). Functions of the proteasome: from protein degradation and immune surveillance to cancer therapy. Biochem Soc Trans, 35, 1217.CrossRefGoogle ScholarPubMed
Guthrie, C., Fink, G., Abelson, J., et al. (1991). Guide to yeast genetics and molecular biology. Methods Enzymol, 194, 1863.Google Scholar
Harper, S.M., Christie, J.M. and Gardner, K.H. (2004). Disruption of the LOV-Jalpha helix interaction activates phototropin kinase activity. Biochemistry, 43, 1618416192.CrossRefGoogle ScholarPubMed
Hausser, M. (2014). Optogenetics: the age of light. Nat Methods, 11, 10121014.CrossRefGoogle Scholar
Hermann, A., Liewald, J.F. and Gottschalk, A. (2015). A photosensitive degron enables acute light-induced protein degradation in the nervous system. Curr Biol, 25, R749R750.CrossRefGoogle ScholarPubMed
Hoyt, M.A., Zhang, M. and Coffino, P. (2003). Ubiquitin-independent mechanisms of mouse ornithine decarboxylase degradation are conserved between mammalian and fungal cells. J Biol Chem, 278, 1213512143.CrossRefGoogle ScholarPubMed
Janke, C., Magiera, M.M., Rathfelder, N., et al. (2004). A versatile toolbox for PCR-based tagging of yeast genes: new fluorescent proteins, more markers and promoter substitution cassettes. Yeast, 21, 947962.CrossRefGoogle Scholar
Jungbluth, M., Renicke, C. and Taxis, C. (2010). Targeted protein depletion in Saccharomyces cerevisiae by activation of a bidirectional degron. BMC Syst Biol, 4, 176.CrossRefGoogle ScholarPubMed
Kim, B. and Lin, M.Z. (2013). Optobiology: optical control of biological processes via protein engineering. Biochem Soc Trans, 41, 11831188.CrossRefGoogle ScholarPubMed
Konermann, S., Brigham, M.D., Trevino, A.E., et al. (2013). Optical control of mammalian endogenous transcription and epigenetic states. Nature, 500, 472476.CrossRefGoogle ScholarPubMed
Kulak, N.A., Pichler, G., Paron, I., et al. (2014). Minimal, encapsulated proteomic-sample processing applied to copy-number estimation in eukaryotic cells. Nat Methods, 11, 319324.CrossRefGoogle ScholarPubMed
Laemmli, U.K. (1970). Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature, 227, 680685.CrossRefGoogle ScholarPubMed
Levskaya, A., Chevalier, A.A., Tabor, J.J., et al. (2005). Synthetic biology: engineering Escherichia coli to see light. Nature, 438, 441442.CrossRefGoogle ScholarPubMed
Lin, J.Y. (2011). A user’s guide to channelrhodopsin variants: features, limitations and future developments. Exp Physiol, 96, 1925.CrossRefGoogle ScholarPubMed
Loetscher, P., Pratt, G. and Rechsteiner, M. (1991). The C terminus of mouse ornithine decarboxylase confers rapid degradation on dihydrofolate reductase. Support for the pest hypothesis. J Biol Chem, 266, 1121311220.CrossRefGoogle Scholar
Matsuzawa, S., Cuddy, M., Fukushima, T., et al. (2005). Method for targeting protein destruction by using a ubiquitin-independent, proteasome-mediated degradation pathway. Proc Natl Acad Sci U S A, 102, 1498214987.CrossRefGoogle Scholar
Miesenbock, G. (2009). The optogenetic catechism. Science, 326, 395399.CrossRefGoogle ScholarPubMed
Neiman, A.M. (2005). Ascospore formation in the yeast Saccharomyces cerevisiae. Microbiol Mol Biol Rev, 69, 565584.CrossRefGoogle ScholarPubMed
Newman, J.R., Ghaemmaghami, S., Ihmels, J., et al. (2006). Single-cell proteomic analysis of S. cerevisiae reveals the architecture of biological noise. Nature, 441, 840846.CrossRefGoogle ScholarPubMed
Ohlendorf, R., Vidavski, R.R., Eldar, A., et al. (2012). From dusk till dawn: one-plasmid systems for light-regulated gene expression. J Mol Biol, 416, 534542.CrossRefGoogle ScholarPubMed
Pathak, G.P., Strickland, D., Vrana, J.D., et al. (2014). Benchmarking of optical dimerizer systems. ACS Synth Biol, 3, 832838.CrossRefGoogle ScholarPubMed
Paul, V.D., Muhlenhoff, U., Stumpfig, M., et al. (2015). The deca-GX3 proteins Yae1-Lto1 function as adaptors recruiting the ABC protein Rli1 for iron–sulfur cluster insertion. Elife, 4, e08231.CrossRefGoogle ScholarPubMed
Polstein, L.R. and Gersbach, C.A. (2012). Light-inducible spatiotemporal control of gene activation by customizable zinc finger transcription factors. J Am Chem Soc, 134, 1648016483.CrossRefGoogle Scholar
Pudasaini, A., El-Arab, K.K. and Zoltowski, B.D. (2015). LOV-based optogenetic devices: light-driven modules to impart photoregulated control of cellular signaling. Front Mol Biosci, 2, 18.CrossRefGoogle ScholarPubMed
Ravid, T. and Hochstrasser, M. (2008). Diversity of degradation signals in the ubiquitin-proteasome system. Nat Rev Mol Cell Biol, 9, 679690.CrossRefGoogle Scholar
Renicke, C., Schuster, D., Usherenko, S., et al. (2013). A LOV2 domain-based optogenetic tool to control protein degradation and cellular function. Chem Biol, 20, 619626.CrossRefGoogle Scholar
Selevsek, N., Chang, C.Y., Gillet, L.C., et al. (2015). Reproducible and consistent quantification of the Saccharomyces cerevisiae proteome by SWATH-mass spectrometry. Mol Cell Proteomics, 14, 739749.CrossRefGoogle Scholar
Shimizu-Sato, S., Huq, E., Tepperman, J.M., et al. (2002). A light-switchable gene promoter system. Nat Biotechnol, 20, 10411044.CrossRefGoogle Scholar
Sorokina, O., Kapus, A., Terecskei, K., et al. (2009). A switchable light-input, light-output system modelled and constructed in yeast. J Biol Eng, 3, 15.CrossRefGoogle ScholarPubMed
Takeuchi, J., Chen, H. and Coffino, P. (2007). Proteasome substrate degradation requires association plus extended peptide. EMBO J, 26, 123131.CrossRefGoogle ScholarPubMed
Takeuchi, J., Chen, H., Hoyt, M.A., et al. (2008). Structural elements of the ubiquitin-independent proteasome degron of ornithine decarboxylase. Biochem J, 410, 401407.CrossRefGoogle ScholarPubMed
Towbin, H., Staehelin, T. and Gordon, J. (1979). Electrophoretic transfer of proteins from polyacrylamide gels to nitrocellulose sheets: procedure and some applications. Proc Natl Acad Sci U S A, 76, 43504354.CrossRefGoogle ScholarPubMed
Tsien, R.Y. (1998). The green fluorescent protein. Annu Rev Biochem, 67, 509544.CrossRefGoogle ScholarPubMed
Usherenko, S., Stibbe, H., Musco, M., et al. (2014). Photo-sensitive degron variants for tuning protein stability by light. BMC Syst Biol, 8, 128.CrossRefGoogle ScholarPubMed
Wang, X., Chen, X. and Yang, Y. (2012). Spatiotemporal control of gene expression by a light-switchable transgene system. Nat Methods, 9, 266269.CrossRefGoogle ScholarPubMed
Wu, Y.I., Frey, D., Lungu, O.I., et al. (2009). A genetically encoded photoactivatable Rac controls the motility of living cells. Nature, 461, 104108.CrossRefGoogle ScholarPubMed
Yaffe, M.P. and Schatz, G. (1984). Two nuclear mutations that block mitochondrial protein import in yeast. Proc Natl Acad Sci U S A, 81, 48194823.CrossRefGoogle Scholar
Zhang, K. and Cui, B. (2015). Optogenetic control of intracellular signaling pathways. Trends Biotechnol, 33, 92100.CrossRefGoogle ScholarPubMed
Ziegler, T. and Moglich, A. (2015). Photoreceptor engineering. Front Mol Biosci, 2, 30.CrossRefGoogle Scholar

References

Abdul-Sater, A.A., Grajkowski, A., Erdjument-Bromage, H., Plumlee, C., Levi, A., Schreiber, M.T., Lee, C., Shuman, H., Beaucage, S.L., and Schindler, C. (2012). The overlapping host responses to bacterial cyclic dinucleotides. Microbes Infect, 14, 188–97.CrossRefGoogle ScholarPubMed
Airan, R.D., Thompson, K.R., Fenno, L.E., Bernstein, H., and Deisseroth, K. (2009). Temporally precise in vivo control of intracellular signalling. Nature, 458, 1025–9.CrossRefGoogle Scholar
Arrenberg, A.B., Stainier, D.Y., Baier, H., and Huisken, J. (2010). Optogenetic control of cardiac function. Science, 330, 971–4.CrossRefGoogle ScholarPubMed
Auslander, S., Wieland, M., and Fussenegger, M. (2012). Smart medication through combination of synthetic biology and cell microencapsulation. Metab Eng, 14, 252–60.CrossRefGoogle ScholarPubMed
Bagheri, A., Gabran, S.R., Salam, M.T., Perez Velazquez, J.L., Mansour, R.R., Salama, M.M., and Genov, R. (2013). Massively-parallel neuromonitoring and neurostimulation rodent headset with nanotextured flexible microelectrodes. IEEE Trans Biomed Circuits Syst, 7, 601–9.CrossRefGoogle ScholarPubMed
Bailes, H.J., Zhuang, L.Y., and Lucas, R.J. (2012). Reproducible and sustained regulation of Galphas signalling using a metazoan opsin as an optogenetic tool. PLoS One, 7, e30774.CrossRefGoogle ScholarPubMed
Basar, M.R., Ahmad, M.Y., Cho, J., and Ibrahim, F. (2014). Application of wireless power transmission systems in wireless capsule endoscopy: an overview. Sensors (Basel), 14, 10929–51.CrossRefGoogle ScholarPubMed
Baumann, P., Spulber, M., Dinu, I.A., and Palivan, C.G. (2014). Cellular Trojan horse based polymer nanoreactors with light-sensitive activity. J. Phys. Chem. B, 118, 9361–70.CrossRefGoogle ScholarPubMed
Bellmann, D., Richardt, A., Freyberger, R., Nuwal, N., Schwarzel, M., Fiala, A., and Stortkuhl, K.F. (2010). Optogenetically induced olfactory stimulation in Drosophila larvae reveals the neuronal basis of odor-aversion behavior. Front. Behav. Neurosci., 4, 27.CrossRefGoogle ScholarPubMed
Boyden, E.S., Zhang, F., Bamberg, E., Nagel, G., and Deisseroth, K. (2005). Millisecond-timescale, genetically targeted optical control of neural activity. Nat. Neurosci., 8, 1263–8.CrossRefGoogle ScholarPubMed
Bruegmann, T., Malan, D., Hesse, M., Beiert, T., Fuegemann, C.J., Fleischmann, B.K., and Sasse, P. (2010). Optogenetic control of heart muscle in vitro and in vivo. Nat Methods, 7, 897900.CrossRefGoogle ScholarPubMed
Bucher, D. and Buchner, E. (2009). Stimulating PACalpha increases miniature excitatory junction potential frequency at the Drosophila neuromuscular junction. J. Neurogenet., 23, 220–4.CrossRefGoogle ScholarPubMed
Bugaj, L.J., Choksi, A.T., Mesuda, C.K., Kane, R.S., and Schaffer, D.V. (2013). Optogenetic protein clustering and signaling activation in mammalian cells. Nat Methods, 10, 249–52.CrossRefGoogle ScholarPubMed
Cao, Z., Livoti, E., Losi, A., and Gartner, W. (2010). A blue light-inducible phosphodiesterase activity in the cyanobacterium Synechococcus elongatus. Photochem. Photobiol., 86, 606–11.CrossRefGoogle ScholarPubMed
Chen, Z.H., Raffelberg, S., Losi, A., Schaap, P., and Gartner, W. (2014). A cyanobacterial light activated adenylyl cyclase partially restores development of a Dictyostelium discoideum, adenylyl cyclase a null mutant. J. Biotechnol., 191, 246–9.CrossRefGoogle ScholarPubMed
Christen, M., Kulasekara, H.D., Christen, B., Kulasekara, B.R., Hoffman, L.R., and Miller, S.I. (2010). Asymmetrical distribution of the second messenger c-di-GMP upon bacterial cell division. Science, 328, 1295–7.CrossRefGoogle ScholarPubMed
Clement, R.G., Bugler, K.E., and Oliver, C.W. (2011). Bionic prosthetic hands: a review of present technology and future aspirations. Surgeon, 9, 336–40.CrossRefGoogle Scholar
Daly, J.J. and Wolpaw, J.R. (2008). Brain–computer interfaces in neurological rehabilitation. Lancet Neurol, 7, 1032–43.CrossRefGoogle ScholarPubMed
Diring, S., Wang, D.O., Kim, C., Kondo, M., Chen, Y., Kitagawa, S., Kamei, K., and Furukawa, S. (2013). Localized cell stimulation by nitric oxide using a photoactive porous coordination polymer platform. Nat Commun, 4, 2684.CrossRefGoogle Scholar
Efetova, M., Petereit, L., Rosiewicz, K., Overend, G., Haussig, F., Hovemann, B.T., Cabrero, P., Dow, J.A., and Schwarzel, M. (2013). Separate roles of PKA and EPAC in renal function unraveled by the optogenetic control of cAMP levels in vivo. J. Cell Sci., 126, 778–88.Google ScholarPubMed
Enomoto, G., Ni-Ni-Win, R. Narikawa, and Ikeuchi, M. (2015). Three cyanobacteriochromes work together to form a light color-sensitive input system for c-di-GMP signaling of cell aggregation. Proc. Natl. Acad. Sci. U. S. A., 112, 8082–7.CrossRefGoogle Scholar
Escalante, C.R., Nistal-Villan, E., Shen, L., Garcia-Sastre, A., and Aggarwal, A.K. (2007). Structure of IRF-3 bound to the PRDIII-I regulatory element of the human interferon-beta enhancer. Mol Cell, 26, 703–16.CrossRefGoogle Scholar
Folcher, M., Oesterle, S., Zwicky, K., Thekkottil, T., Heymoz, J., Hohmann, M., Christen, M., Daoud El-Baba, M., Buchmann, P., and Fussenegger, M. (2014). Mind-controlled transgene expression by a wireless-powered optogenetic designer cell implant. Nat Commun, 5, 5392.CrossRefGoogle Scholar
Galan, F., Nuttin, M., Lew, E., Ferrez, P.W., Vanacker, G., Philips, J., and Millan Jdel, R. (2008). A brain-actuated wheelchair: asynchronous and non-invasive brain–computer interfaces for continuous control of robots. Clin Neurophysiol, 119, 2159–69.CrossRefGoogle ScholarPubMed
Gasser, C., Taiber, S., Yeh, C.M., Wittig, C.H., Hegemann, P., Ryu, S., Wunder, F., and Moglich, A. (2014). Engineering of a red-light-activated human cAMP/cGMP-specific phosphodiesterase. Proc. Natl. Acad. Sci. U. S. A., 111, 8803–8.CrossRefGoogle Scholar
Gomelsky, M. (2011). cAMP, c-di-GMP, c-di-AMP and now cGMP: bacteria use them all! Mol. Microbiol., 79, 562–5.CrossRefGoogle Scholar
Gomelsky, M. and Klug, G. (2002). BLUF: a novel FAD-binding domain involved in sensory transduction in microorganisms. Trends Biochem. Sci., 27, 497500.CrossRefGoogle ScholarPubMed
Guenther, T., Lovell, N.H., and Suaning, G.J. (2012). Bionic vision: system architectures: a review. Expert Rev. Med. Devices, 9, 3348.CrossRefGoogle ScholarPubMed
Gutierrez-Triana, J.A., Herget, U., Castillo-Ramirez, L.A., Lutz, M., Yeh, C.M., De Marco, R.J., and Ryu, S. (2015). Manipulation of interrenal cell function in developing zebrafish using genetically targeted ablation and an optogenetic tool. Endocrinology, 156, 3394–401.CrossRefGoogle Scholar
Hartmann, A., Arroyo-Olarte, R.D., Imkeller, K., Hegemann, P., Lucius, R., and Gupta, N. (2013). Optogenetic modulation of an adenylate cyclase in Toxoplasma gondii demonstrates a requirement of the parasite cAMP for host-cell invasion and stage differentiation. J. Biol. Chem., 288, 13705–17.CrossRefGoogle ScholarPubMed
He, J., Zhang, P., Babu, T., Liu, Y., Gong, J., and Nie, Z. (2013). Near-infrared light-responsive vesicles of Au nanoflowers. Chem. Commun. (Camb.), 49, 576–8.Google Scholar
Herrou, J. and Crosson, S. (2011). Function, structure and mechanism of bacterial photosensory LOV proteins. Nat. Rev. Microbiol., 9, 713–23.CrossRefGoogle ScholarPubMed
Ieda, N., Hotta, Y., Miyata, N., Kimura, K., and Nakagawa, H. (2014). Photomanipulation of vasodilation with a blue-light-controllable nitric oxide releaser. J. Am. Chem. Soc., 136, 7085–91.CrossRefGoogle ScholarPubMed
Jansen, V., Alvarez, L., Balbach, M., Strunker, T., Hegemann, P., Kaupp, U.B., and Wachten, D. (2015). Controlling fertilization and cAMP signaling in sperm by optogenetics. Elife, 4, e05161.CrossRefGoogle ScholarPubMed
Jenal, U. and Malone, J. (2006). Mechanisms of cyclic-di-GMP signaling in bacteria. Annu. Rev. Genet., 40, 385407.CrossRefGoogle ScholarPubMed
Jordheim, L.P., Durantel, D., Zoulim, F., and Dumontet, C. (2013). Advances in the development of nucleoside and nucleotide analogues for cancer and viral diseases. Nat. Rev. Drug Discov., 12, 447–64.CrossRefGoogle ScholarPubMed
Joung, Y.H. (2013). Development of implantable medical devices: from an engineering perspective. Int. Neurourol. J., 17, 98106.CrossRefGoogle Scholar
Kale, R.P., Kouzani, A.Z., Walder, K., Berk, M., and Tye, S.J. (2015). Evolution of optogenetic microdevices. Neurophotonics, 2, 031206.CrossRefGoogle Scholar
Karami, A., Eyjolfsdottir, H., Vijayaraghavan, S., Lind, G., Almqvist, P., Kadir, A., Linderoth, B., Andreasen, N., Blennow, K., Wall, A., Westman, E., Ferreira, D., Wiberg, M. Kristoffersen, Wahlund, L.O., Seiger, A., Nordberg, A., Wahlberg, L., Darreh-Shori, T., and Eriksdotter, M. (2015). Changes in CSF cholinergic biomarkers in response to cell therapy with NGF in patients with Alzheimer’s disease. Alzheimers Dement.CrossRefGoogle Scholar
Kasahara, M., Unno, T., Yashiro, K., and Ohmori, M. (2001). CyaG, a novel cyanobacterial adenylyl cyclase and a possible ancestor of mammalian guanylyl cyclases. J. Biol. Chem., 276, 10564–9.CrossRefGoogle Scholar
Kim, J.M., Hwa, J., Garriga, P., Reeves, P.J., RajBhandary, U.L., and Khorana, H.G. (2005). Light-driven activation of beta 2-adrenergic receptor signaling by a chimeric rhodopsin containing the beta 2-adrenergic receptor cytoplasmic loops. Biochemistry, 44, 2284–92.CrossRefGoogle Scholar
Kim, T., Folcher, M., Charpin-El Hamri, G., and Fussenegger, M. (2015a). A synthetic cGMP-sensitive gene switch providing Viagra®-controlled gene expression in mammalian cells and mice. Metab. Eng, 29, 169–79.CrossRefGoogle ScholarPubMed
Kim, T., Folcher, M., Doaud-El Baba, M., and Fussenegger, M. (2015b). A synthetic erectile optogenetic stimulator enabling blue-light-inducible penile erection. Angew. Chem. Int. Ed. Engl., 54, 5933–8.CrossRefGoogle Scholar
Kim, T.I., McCall, J.G., Jung, Y.H., Huang, X., Siuda, E.R., Li, Y., Song, J., Song, Y.M., Pao, H.A., Kim, R.H., Lu, C., Lee, S.D., Song, I.S., Shin, G., Al-Hasani, R., Kim, S., Tan, M.P., Huang, Y., Omenetto, F.G., Rogers, J.A., and Bruchas, M.R. (2013). Injectable, cellular-scale optoelectronics with applications for wireless optogenetics. Science, 340, 211–6.CrossRefGoogle ScholarPubMed
Leung, D.W., Otomo, C., Chory, J., and Rosen, M.K. (2008). Genetically encoded photoswitching of actin assembly through the Cdc42–WASP–Arp2/3 complex pathway. Proc. Natl. Acad. Sci. U. S. A., 105, 12797–802.CrossRefGoogle ScholarPubMed
Looser, J., Schroder-Lang, S., Hegemann, P., and Nagel, G. (2009). Mechanistic insights in light-induced cAMP production by photoactivated adenylyl cyclase alpha (PACalpha). Biol. Chem., 390, 1105–11.CrossRefGoogle Scholar
Losi, A. and Gartner, W. (2008). Bacterial bilin- and flavin-binding photoreceptors. Photochem. Photobiol. Sci., 7, 1168–78.CrossRefGoogle ScholarPubMed
Mandalari, C., Losi, A., and Gartner, W. (2013). Distance-tree analysis, distribution and co-presence of bilin- and flavin-binding prokaryotic photoreceptors for visible light. Photochem Photobiol Sci, 12, 1144–57.CrossRefGoogle ScholarPubMed
Marden, J.N., Dong, Q., Roychowdhury, S., Berleman, J.E., and Bauer, C.E. (2011). Cyclic GMP controls Rhodospirillum centenum cyst development. Mol. Microbiol., 79, 600–15.CrossRefGoogle ScholarPubMed
Moglich, A. and Moffat, K. (2010). Engineered photoreceptors as novel optogenetic tools. Photochem. Photobiol. Sci., 9, 1286–300.CrossRefGoogle ScholarPubMed
Montgomery, K.L., Yeh, A.J., Ho, J.S., Tsao, V., Mohan Iyer, S., Grosenick, L., Ferenczi, E.A., Tanabe, Y., Deisseroth, K., Delp, S.L., and Poon, A.S. (2015). Wirelessly powered, fully internal optogenetics for brain, spinal and peripheral circuits in mice. Nat Methods, 12, 969–74.CrossRefGoogle ScholarPubMed
Motta-Mena, L.B., Reade, A., Mallory, M.J., Glantz, S., Weiner, O.D., Lynch, K.W., and Gardner, K.H. (2014). An optogenetic gene expression system with rapid activation and deactivation kinetics. Nat. Chem. Biol., 10, 196202.CrossRefGoogle ScholarPubMed
Myakishev-Rempel, M., Stadler, I., Brondon, P., Axe, D.R., Friedman, M., Nardia, F.B., and Lanzafame, R. (2012). A preliminary study of the safety of red light phototherapy of tissues harboring cancer. Photomed. Laser Surg., 30, 551–8.CrossRefGoogle ScholarPubMed
Nagahama, T., Suzuki, T., Yoshikawa, S., and Iseki, M. (2007). Functional transplant of photoactivated adenylyl cyclase (PAC) into Aplysia sensory neurons. Neurosci. Res., 59, 81–8.CrossRefGoogle ScholarPubMed
Oplander, C., Deck, A., Volkmar, C.M., Kirsch, M., Liebmann, J., Born, M., van Abeelen, F., van Faassen, E.E., Kroncke, K.D., Windolf, J., and Suschek, C.V. (2013). Mechanism and biological relevance of blue-light (420–453 nm)-induced nonenzymatic nitric oxide generation from photolabile nitric oxide derivates in human skin in vitro and in vivo. Free Radic. Biol. Med., 65, 1363–77.CrossRefGoogle ScholarPubMed
Piatkevich, K.D., Subach, F.V., and Verkhusha, V.V. (2013). Far-red light photoactivatable near-infrared fluorescent proteins engineered from a bacterial phytochrome. Nat Commun, 4, 2153.CrossRefGoogle ScholarPubMed
Roembke, B.T., Zhou, J., Zheng, Y., Sayre, D., Lizardo, A., Bernard, L., and Sintim, H.O. (2014). A cyclic dinucleotide containing 2-aminopurine is a general fluorescent sensor for c-di-GMP and 3′,3′-cGAMP. Mol. Biosyst., 10, 1568–75.CrossRefGoogle Scholar
Ryu, M.H., Kang, I.H., Nelson, M.D., Jensen, T.M., Lyuksyutova, A.I., Siltberg-Liberles, J., Raizen, D.M., and Gomelsky, M. (2014). Engineering adenylate cyclases regulated by near-infrared window light. Proc Natl Acad Sci U S A, 111, 10167–72.CrossRefGoogle ScholarPubMed
Ryu, M.H., Moskvin, O.V., Siltberg-Liberles, J., and Gomelsky, M. (2010). Natural and engineered photoactivated nucleotidyl cyclases for optogenetic applications. J. Biol. Chem., 285, 41501–8.CrossRefGoogle ScholarPubMed
Ryu, M.H., Youn, H., Kang, I.H., and Gomelsky, M. (2015). Identification of bacterial guanylate cyclases. Proteins, 83, 799804.CrossRefGoogle ScholarPubMed
Samanta, A., Thunemann, M., Feil, R., and Stafforst, T. (2014). Upon the photostability of 8-nitro-cGMP and its caging as a 7-dimethylaminocoumarinyl ester. Chem. Commun. (Camb.), 50, 7120–3.CrossRefGoogle ScholarPubMed
Scheib, U., Stehfest, K., Gee, C.E., Korschen, H.G., Fudim, R., Oertner, T.G., and Hegemann, P. (2015). The rhodopsin-guanylyl cyclase of the aquatic fungus Blastocladiella emersonii enables fast optical control of cGMP signaling. Sci Signal, 8, rs8.CrossRefGoogle ScholarPubMed
Schroder-Lang, S., Schwarzel, M., Seifert, R., Strunker, T., Kateriya, S., Looser, J., Watanabe, M., Kaupp, U.B., Hegemann, P., and Nagel, G. (2007). Fast manipulation of cellular cAMP level by light in vivo. Nat Methods, 4, 3942.CrossRefGoogle Scholar
Sikka, G., Hussmann, G.P., Pandey, D., Cao, S., Hori, D., Park, J.T., Steppan, J., Kim, J.H., Barodka, V., Myers, A.C., Santhanam, L., Nyhan, D., Halushka, M.K., Koehler, R.C., Snyder, S.H., Shimoda, L.A., and Berkowitz, D.E. (2014). Melanopsin mediates light-dependent relaxation in blood vessels. Proc. Natl. Acad. Sci. U. S. A., 111, 17977–82.CrossRefGoogle Scholar
Sinha, S.C. and Sprang, S.R. (2006). Structures, mechanism, regulation and evolution of class III nucleotidyl cyclases. Rev. Physiol. Biochem. Pharmacol., 157, 105–40.CrossRefGoogle ScholarPubMed
Stierl, M., Penzkofer, A., Kennis, J.T., Hegemann, P., and Mathes, T. (2014). Key residues for the light regulation of the blue light-activated adenylyl cyclase from Beggiatoa sp. Biochemistry, 53, 5121–30.CrossRefGoogle ScholarPubMed
Stortkuhl, K.F. and Fiala, A. (2011). The smell of blue light: a new approach toward understanding an olfactory neuronal network. Front. Neurosci., 5, 72.CrossRefGoogle Scholar
Sun, L., Wu, J., Du, F., Chen, X., and Chen, Z.J. (2013). Cyclic GMP–AMP synthase is a cytosolic DNA sensor that activates the type I interferon pathway. Science, 339, 786–91.CrossRefGoogle ScholarPubMed
Takala, H., Bjorling, A., Berntsson, O., Lehtivuori, H., Niebling, S., Hoernke, M., Kosheleva, I., Henning, R., Menzel, A., Ihalainen, J.A., and Westenhoff, S. (2014). Signal amplification and transduction in phytochrome photosensors. Nature, 509, 245–8.CrossRefGoogle ScholarPubMed
Tang, X., Zhang, J., Sun, J., Wang, Y., Wu, J., and Zhang, L. (2013). Caged nucleotides/nucleosides and their photochemical biology. Org. Biomol. Chem., 11, 7814–24.CrossRefGoogle ScholarPubMed
Weissenberger, S., Schultheis, C., Liewald, J.F., Erbguth, K., Nagel, G., and Gottschalk, A. (2011). PACalpha – an optogenetic tool for in vivo manipulation of cellular cAMP levels, neurotransmitter release, and behavior in Caenorhabditis elegans. J. Neurochem., 116, 616–25.CrossRefGoogle ScholarPubMed
Werner, G.S., Schaefer, C., Dirks, R., Figulla, H.R., and Kreuzer, H. (1993). Doppler echocardiographic assessment of left ventricular filling in idiopathic dilated cardiomyopathy during a one-year follow-up: relation to the clinical course of disease. Am. Heart J., 126, 1408–16.CrossRefGoogle Scholar
Wu, Y., Li, S.S., Jin, X., Cui, N., Zhang, S., and Jiang, C. (2015). Optogenetic approach for functional assays of the cardiovascular system by light activation of the vascular smooth muscle. Vascul. Pharmacol., 71, 192200.CrossRefGoogle ScholarPubMed
Yasukawa, H., Sato, A., Kita, A., Kodaira, K., Iseki, M., Takahashi, T., Shibusawa, M., Watanabe, M., and Yagita, K. (2013). Identification of photoactivated adenylyl cyclases in Naegleria australiensis and BLUF-containing protein in Naegleria fowleri. J. Gen. Appl. Microbiol., 59, 361–9.CrossRefGoogle ScholarPubMed
Ye, H., Baba, M. Daoud-El, Peng, R.W., and Fussenegger, M. (2011). A synthetic optogenetic transcription device enhances blood-glucose homeostasis in mice. Science, 332, 1565–8.CrossRefGoogle Scholar
Yizhar, O., Fenno, L.E., Davidson, T.J., Mogri, M., and Deisseroth, K. (2011). Optogenetics in neural systems. Neuron, 71, 934.CrossRefGoogle ScholarPubMed

References

Atasoy, D., Aponte, Y., Su, H.H. and Sternson, S.M. (2008) A FLEX switch targets channelrhodopsin-2 to multiple cell types for imaging and long-range circuit mapping. The Journal of Neuroscience, 28, 7025–30.CrossRefGoogle ScholarPubMed
Baubet, V., Le Mouellic, H., Campbell, A.K., Lucas-Meunier, E., Fossier, P. and Brúlet, P. (2000) Chimeric green fluorescent protein-aequorin as bioluminescent Ca2+ reporters at the single-cell level. Proceedings of the National Academy of Sciences of the United States of America, 97, 7260–5.Google Scholar
Berglund, K., Birkner, E., Augustine, G.J. and Hochgeschwender, U. (2013) Light-emitting channelrhodopsins for combined optogenetic and chemical-genetic control of neurons. PLoS ONE, 8, e59759.CrossRefGoogle ScholarPubMed
Berglund, K., Gutekunst, C.-A., Tung, J., Hochgeschwender, U. and Gross, R.E. (2015) Step-function luminopsin for prolonged activation of neurons by bioluminescence. Society for Neuroscience Abstracts.Google Scholar
Berglund, K., Tung, J.K., Higashikubo, B., Gross, R.E., Moore, C.I. and Hochgeschwender, U. (2016) Combined optogenetic and chemogenetic control of neurons. Methods in Molecular Biology, 1408, 207–25.CrossRefGoogle ScholarPubMed
Berndt, A., Lee, S.Y., Ramakrishnan, C. and Deisseroth, K. (2014) Structure-guided transformation of channelrhodopsin into a light-activated chloride channel. Science (New York, N.Y.), 344, 420–4.CrossRefGoogle ScholarPubMed
Berndt, A., Schoenenberger, P., Mattis, J., Tye, K.M., Deisseroth, K., Hegemann, P., et al. (2011) High-efficiency channelrhodopsins for fast neuronal stimulation at low light levels. Proceedings of the National Academy of Sciences of the United States of America, 108, 7595–600.Google ScholarPubMed
Berndt, A., Yizhar, O., Gunaydin, L.A., Hegemann, P. and Deisseroth, K. (2009) Bi-stable neural state switches. Nature Neuroscience, 12, 229–34.CrossRefGoogle Scholar
Birkner, E., Berglund, K., Klein, M.E., Augustine, G.J. and Hochgeschwender, U. (2014) Non-invasive activation of optogenetic actuators. SPIE Proceedings, 8928, 89282F.Google ScholarPubMed
Boyden, E.S., Zhang, F., Bamberg, E., Nagel, G. and Deisseroth, K. (2005) Millisecond-timescale, genetically targeted optical control of neural activity. Nature neuroscience, 8, 1263–8.CrossRefGoogle Scholar
Campbell, A.K. (1974) Extraction, partial purification and properties of obelin, the calcium-activated luminescent protein from the hydroid Obelia geniculata. The Biochemical Journal, 143, 411–8.CrossRefGoogle ScholarPubMed
Chou, W.C., Liao, K.W., Lo, Y.C., Jiang, S.Y., Yeh, M.Y. and Roffler, S.R. (1999) Expression of chimeric monomer and dimer proteins on the plasma membrane of mammalian cells. Biotechnology and Bioengineering, 65, 160–9.3.0.CO;2-U>CrossRefGoogle ScholarPubMed
Chow, B.Y., Han, X., Dobry, A.S., Qian, X., Chuong, A.S., Li, M., et al. (2010) High-performance genetically targetable optical neural silencing by light-driven proton pumps. Nature, 463, 98102.CrossRefGoogle Scholar
Chuong, A.S., Miri, M.L., Busskamp, V., Matthews, G.A.C., Acker, L.C., Sørensen, A.T., et al. (2014) Noninvasive optical inhibition with a red-shifted microbial rhodopsin. Nature Neuroscience, 17, 1123–9.CrossRefGoogle ScholarPubMed
Clissold, K.A., Berglund, K., Klein, M.E., Prevosto, V., Koval, M., Abzug, Z.M., et al. (2014) In vivo bioluminescence-driven optogenetics for neuronal activation and inhibition. Society for Neuroscience Abstracts.Google Scholar
Deluca, M. (1976) Firefly luciferase. Advances in Enzymology and Related Areas of Molecular Biology, 44, 3768.CrossRefGoogle ScholarPubMed
Ernst, O.P., Sánchez Murcia, P.A., Daldrop, P., Tsunoda, S.P., Kateriya, S. and Hegemann, P. (2008) Photoactivation of channelrhodopsin. The Journal of Biological Chemistry, 283, 1637–43.CrossRefGoogle Scholar
Fairless, R., Masius, H., Rohlmann, A., Heupel, K., Ahmad, M., Reissner, C., et al. (2008) Polarized targeting of neurexins to synapses is regulated by their C-terminal sequences. The Journal of Neuroscience, 28, 12969–81.CrossRefGoogle Scholar
Feinberg, E.H., Vanhoven, M.K., Bendesky, A., Wang, G., Fetter, R.D., Shen, K., et al. (2008) GFP reconstitution across synaptic partners (GRASP) defines cell contacts and synapses in living nervous systems. Neuron, 57, 353–63.CrossRefGoogle ScholarPubMed
Govorunova, E.G., Sineshchekov, O.A., Janz, R., Liu, X. and Spudich, J.L. (2015) Natural light-gated anion channels: a family of microbial rhodopsins for advanced optogenetics. Science, 349, 647–50.CrossRefGoogle ScholarPubMed
Haddock, S.H.D., Moline, M.A. and Case, J.F. (2010) Bioluminescence in the sea. Annual Review of Marine Science, 2, 443–93.CrossRefGoogle ScholarPubMed
Hall, M.P., Unch, J., Binkowski, B.F., Valley, M.P., Butler, B.L., Wood, M.G., et al. (2012) Engineered luciferase reporter from a deep sea shrimp utilizing a novel imidazopyrazinone substrate. ACS Chemical Biology, 7, 1848–57.CrossRefGoogle ScholarPubMed
Han, X. and Boyden, E.S. (2007) Multiple-color optical activation, silencing, and desynchronization of neural activity, with single-spike temporal resolution. PLoS One, 2, e299.CrossRefGoogle ScholarPubMed
Han, X., Chow, B.Y., Zhou, H., Klapoetke, N.C., Chuong, A., Rajimehr, R., et al. (2011) A high-light sensitivity optical neural silencer: development and application to optogenetic control of non-human primate cortex. Frontiers in Systems Neuroscience, 5, 18.CrossRefGoogle Scholar
Higashikubo, B., McDonnell, E., Hochgeschwender, U. and Moore, C.I. (2015) Multi-timescale In vivo regulation of the thalamic reticular nucleus using bioluminescent optogenetics (BL-OG). Society for Neuroscience Abstracts.Google Scholar
Hoshino, H., Nakajima, Y. and Ohmiya, Y. (2007) Luciferase-YFP fusion tag with enhanced emission for single-cell luminescence imaging. Nature Methods, 4, 637–9.CrossRefGoogle ScholarPubMed
Hu, C.-D. and Kerppola, T.K. (2003) Simultaneous visualization of multiple protein interactions in living cells using multicolor fluorescence complementation analysis. Nature Biotechnology, 21, 539–45.CrossRefGoogle ScholarPubMed
Ichtchenko, K., Hata, Y., Nguyen, T., Ullrich, B., Missler, M., Moomaw, C., et al. (1995) Neuroligin 1: a splice site-specific ligand for beta-neurexins. Cell, 81, 435–43.CrossRefGoogle ScholarPubMed
Kendall, J.M., Badminton, M.N., Dormer, R.L. and Campbell, A.K. (1994) Changes in free calcium in the endoplasmic reticulum of living cells detected using targeted aequorin. Analytical Biochemistry, 221, 173–81.CrossRefGoogle Scholar
Kerppola, T.K. (2006) Design and implementation of bimolecular fluorescence complementation (BiFC) assays for the visualization of protein interactions in living cells. Nature Protocols, 1, 12781286.CrossRefGoogle Scholar
Kim, S.B., Suzuki, H., Sato, M. and Tao, H. (2011) Superluminescent variants of marine luciferases for bioassays. Analytical Chemistry, 83, 8732–40.CrossRefGoogle Scholar
Kim, S.B., Torimura, M. and Tao, H. (2013) Creation of artificial luciferases for bioassays. Bioconjugate Chemistry, 24, 2067–75.CrossRefGoogle ScholarPubMed
Kim, J., Zhao, T., Petralia, R.S., Yu, Y., Peng, H., Myers, E., et al. (2012) mGRASP enables mapping mammalian synaptic connectivity with light microscopy. Nature Methods, 9, 96102.CrossRefGoogle Scholar
Klapoetke, N.C., Murata, Y., Kim, S.S., Pulver, S.R., Birdsey-Benson, A., Cho, Y.K., et al. (2014) Independent optical excitation of distinct neural populations. Nature Methods, 11, 338–46.Google ScholarPubMed
Kopparaju, R., Lin, S.H., Chen, Y.C., Hochgeschwender, U. and Chen, C.C. (2015) Intimate touch at a distance. Society for Neuroscience Abstracts.Google Scholar
Lanyi, J.K., Duschl, A., Hatfield, G.W., May, K. and Oesterhelt, D. (1990) The primary structure of a halorhodopsin from Natronobacterium pharaonis. Structural, functional and evolutionary implications for bacterial rhodopsins and halorhodopsins. The Journal of Biological Chemistry, 265, 1253–60.CrossRefGoogle ScholarPubMed
Li, X., Gutierrez, D. V, Hanson, M.G., Han, J., Mark, M.D., Chiel, H., et al. (2005) Fast noninvasive activation and inhibition of neural and network activity by vertebrate rhodopsin and green algae channelrhodopsin. Proceedings of the National Academy of Sciences of the United States of America, 102, 17816–21.Google ScholarPubMed
Lin, J.Y., Knutsen, P.M., Muller, A., Kleinfeld, D. and Tsien, R.Y. (2013) ReaChR: a red-shifted variant of channelrhodopsin enables deep transcranial optogenetic excitation. Nature Neuroscience, 16, 1499–508.CrossRefGoogle ScholarPubMed
Martin, J.-R., Rogers, K.L., Chagneau, C. and Brûlet, P. (2007) In vivo bioluminescence imaging of Ca signalling in the brain of Drosophila. PLoS One, 2, e275.CrossRefGoogle ScholarPubMed
Le Masson, G., Przedborski, S. and Abbott, L.F. (2014) A computational model of motor neuron degeneration. Neuron, 83, 975–88.Google ScholarPubMed
Morise, H., Shimomura, O., Johnson, F.H. and Winant, J. (1974) Intermolecular energy transfer in the bioluminescent system of aequorea. Biochemistry, 13, 2656–62.CrossRefGoogle Scholar
Nagel, G., Szellas, T., Huhn, W., Kateriya, S., Adeishvili, N., Berthold, P., et al. (2003) Channelrhodopsin-2, a directly light-gated cation-selective membrane channel. Proceedings of the National Academy of Sciences of the United States of America, 100, 13940–5.Google ScholarPubMed
Naumann, E.A., Kampff, A.R., Prober, D.A., Schier, A.F. and Engert, F. (2010) Monitoring neural activity with bioluminescence during natural behavior. Nature Neuroscience, 13, 513–20.CrossRefGoogle Scholar
Rangaraju, V., Calloway, N. and Ryan, T.A. (2014) Activity-driven local ATP synthesis is required for synaptic function. Cell, 156, 825–35.CrossRefGoogle Scholar
Rogers, K.L., Picaud, S., Roncali, E., Boisgard, R., Colasante, C., Stinnakre, J., et al. (2007) Non-invasive in vivo imaging of calcium signaling in mice. PLoS One, 2, e974.CrossRefGoogle ScholarPubMed
Rogers, K.L., Stinnakre, J., Agulhon, C., Jublot, D., Shorte, S.L., Kremer, E.J., et al. (2005) Visualization of local Ca2+ dynamics with genetically encoded bioluminescent reporters. The European Journal of Neuroscience, 21, 597610.CrossRefGoogle ScholarPubMed
Saito, K., Chang, Y.-F., Horikawa, K., Hatsugai, N., Higuchi, Y., Hashida, M., et al. (2012) Luminescent proteins for high-speed single-cell and whole-body imaging. Nature Communications, 3, 1262.CrossRefGoogle Scholar
Schnütgen, F., Doerflinger, N., Calléja, C., Wendling, O., Chambon, P. and Ghyselinck, N.B. (2003) A directional strategy for monitoring Cre-mediated recombination at the cellular level in the mouse. Nature Biotechnology, 21, 562–5.CrossRefGoogle Scholar
Shimomura, O. (1985) Bioluminescence in the sea: photoprotein systems. Symposia of the Society for Experimental Biology, 39, 351–72.Google Scholar
Shimomura, O., Johnson, F.H. and Saiga, Y. (1962) Extraction, purification and properties of aequorin, a bioluminescent protein from the luminous hydromedusan, aequorea. Journal of Cellular and Comparative Physiology, 59, 223–39.CrossRefGoogle Scholar
Shimomura, O., Masugi, T., Johnson, F.H. and Haneda, Y. (1978) Properties and reaction mechanism of the bioluminescence system of the deep-sea shrimp Oplophorus gracilorostris. Biochemistry, 17, 994–8.CrossRefGoogle ScholarPubMed
Sternson, S.M. and Roth, B.L. (2014) Chemogenetic tools to interrogate brain functions. Annual Review of Neuroscience, 37, 387407.CrossRefGoogle ScholarPubMed
Takai, A., Nakano, M., Saito, K., Haruno, R., Watanabe, T.M., Ohyanagi, T., et al. (2015) Expanded palette of nano-lanterns for real-time multicolor luminescence imaging. Proceedings of the National Academy of Sciences of the United States of America, 112, 4352–6.Google ScholarPubMed
Tannous, B.A., Kim, D.-E., Fernandez, J.L., Weissleder, R. and Breakefield, X.O. (2005) Codon-optimized Gaussia luciferase cDNA for mammalian gene expression in culture and in vivo. Molecular Therapy, 11, 435–43.CrossRefGoogle ScholarPubMed
Teranishi, K. and Shimomura, O. (1997) Solubilizing coelenterazine in water with hydroxypropyl-.BETA.-cyclodextrin. Bioscience, Biotechnology, and Biochemistry, 61, 12191220.CrossRefGoogle Scholar
Tung, J.K., Gutekunst, C.-A. and Gross, R.E. (2015) Inhibitory luminopsins: genetically-encoded bioluminescent opsins for versatile, scalable, and hardware-independent optogenetic inhibition. Scientific Reports, 5, 14366.CrossRefGoogle ScholarPubMed
Verhaegen, M. and Christopoulos, T.K. (2002) Recombinant Gaussia luciferase. Overexpression, purification, and analytical application of a bioluminescent reporter for DNA hybridization. Analytical Chemistry, 74, 4378–85.CrossRefGoogle Scholar
Wang, H., Peca, J., Matsuzaki, M., Matsuzaki, K., Noguchi, J., Qiu, L., et al. (2007) High-speed mapping of synaptic connectivity using photostimulation in channelrhodopsin-2 transgenic mice. Proceedings of the National Academy of Sciences of the United States of America, 104, 8143–8.Google Scholar
Ward, W.W. and Cormier, M.J. (1976) In vitro energy transfer in Renilla bioluminescence. The Journal of Physical Chemsitry, 80, 2289–91.Google Scholar
Welsh, J.P., Patel, K.G., Manthiram, K. and Swartz, J.R. (2009) Multiply mutated Gaussia luciferases provide prolonged and intense bioluminescence. Biochemical and Biophysical Research Communications, 389, 563–8.CrossRefGoogle ScholarPubMed
Wen, L., Park, S.Y., Clissold, K.A., Berglund, K., Yin, H.H., Augustine, G.J., et al. (2015) Luminopsins allow neuronal activation over a range of spatial and temporal scales. Society for Neuroscience Abstracts.Google Scholar
Wietek, J., Beltramo, R., Scanziani, M., Hegemann, P., Oertner, T.G. and Simon Wiegert, J. (2015) An improved chloride-conducting channelrhodopsin for light-induced inhibition of neuronal activity in vivo. Scientific Reports, 5, 14807.CrossRefGoogle ScholarPubMed
Wietek, J., Wiegert, J.S., Adeishvili, N., Schneider, F., Watanabe, H., Tsunoda, S.P., et al. (2014) Conversion of channelrhodopsin into a light-gated chloride channel. Science (New York, N.Y.), 344, 409–12.CrossRefGoogle ScholarPubMed
Yamagata, M. and Sanes, J.R. (2012) Transgenic strategy for identifying synaptic connections in mice by fluorescence complementation (GRASP). Frontiers in Molecular Neuroscience, 5, 18.CrossRefGoogle ScholarPubMed
Zhang, F., Prigge, M., Beyrière, F., Tsunoda, S.P., Mattis, J., Yizhar, O., et al. (2008) Red-shifted optogenetic excitation: a tool for fast neural control derived from Volvox carteri. Nature Neuroscience, 11, 631–3.CrossRefGoogle ScholarPubMed
Zhang, F., Wang, L.-P., Brauner, M., Liewald, J.F., Kay, K., Watzke, N., et al. (2007) Multimodal fast optical interrogation of neural circuitry. Nature, 446, 633–9.CrossRefGoogle ScholarPubMed
Zhao, H., Doyle, T.C., Wong, R.J., Cao, Y., Stevenson, D.K., Piwnica-Worms, D., et al. (2004) Characterization of coelenterazine analogs for measurements of Renilla luciferase activity in live cells and living animals. Molecular Imaging, 3, 4354.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×