Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-xtgtn Total loading time: 0 Render date: 2024-04-23T17:34:02.442Z Has data issue: false hasContentIssue false

Part III - Optogenetics in Neurobiology, Brain Circuits, and Plasticity

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. 167 - 238
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

Bergey, G. K., Morrell, M. J., et al., (2015). Long-term treatment with responsive brain stimulation in adults with refractory partial seizures. Neurology 84(8): 810817.CrossRefGoogle ScholarPubMed
Berndt, A., Lee, S. Y., et al., (2014). Structure-guided transformation of channelrhodopsin into a light-activated chloride channel. Science 344(6182): 420424.CrossRefGoogle ScholarPubMed
Boyden, E. S., Zhang, F., et al., (2005). Millisecond-timescale, genetically targeted optical control of neural activity. Nat Neurosci 8(9): 12631268.CrossRefGoogle ScholarPubMed
Carr, F. B. and Zachariou, V. (2014). Nociception and pain: lessons from optogenetics. Front Behav Neurosci 8: 69.CrossRefGoogle ScholarPubMed
Cavanaugh, J., Monosov, I. E., et al., (2012). Optogenetic inactivation modifies monkey visuomotor behavior. Neuron 76(5): 901907.CrossRefGoogle ScholarPubMed
Chiang, C. C., Ladas, T. P., et al., (2014). Seizure suppression by high frequency optogenetic stimulation using in vitro and in vivo animal models of epilepsy. Brain Stimul 7(6): 890899.CrossRefGoogle ScholarPubMed
Gerits, A., Farivar, R., et al., (2012). Optogenetically induced behavioral and functional network changes in primates. Curr Biol 22(18): 17221726.CrossRefGoogle ScholarPubMed
Gerits, A. and Vanduffel, W. (2013). Optogenetics in primates: a shining future? Trends Genet 29(7): 403411.CrossRefGoogle ScholarPubMed
Gradinaru, V., Mogri, M., et al., (2009). Optical deconstruction of Parkinsonian neural circuitry. Science 324(5925): 354359.CrossRefGoogle ScholarPubMed
Gradinaru, V., Zhang, F., et al., (2010). Molecular and cellular approaches for diversifying and extending optogenetics. Cell 141(1): 154165.CrossRefGoogle ScholarPubMed
Gunaydin, L. A., Yizhar, O., et al., (2010). Ultrafast optogenetic control. Nat Neurosci 13(3): 387392.CrossRefGoogle ScholarPubMed
Han, X. (2012). Optogenetics in the nonhuman primate. Prog Brain Res 196: 215233.CrossRefGoogle ScholarPubMed
Han, X., Chow, B. Y., et al., (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
Jazayeri, M., Lindbloom-Brown, Z., et al., (2012). Saccadic eye movements evoked by optogenetic activation of primate V1. Nat Neurosci 15(10): 13681370.CrossRefGoogle ScholarPubMed
Jin, X., Tecuapetla, F., et al., (2014). Basal ganglia subcircuits distinctively encode the parsing and concatenation of action sequences. Nat Neurosci 17(3): 423430.CrossRefGoogle ScholarPubMed
Krook-Magnuson, E., Armstrong, C., et al., (2013). On-demand optogenetic control of spontaneous seizures in temporal lobe epilepsy. Nat Commun 4: 1376.CrossRefGoogle ScholarPubMed
Krook-Magnuson, E. and Soltesz, I. (2015). Beyond the hammer and the scalpel: selective circuit control for the epilepsies. Nat Neurosci 18(3): 331338.CrossRefGoogle Scholar
Krook-Magnuson, E., Szabo, G. G., et al., (2014). Cerebellar directed optogenetic intervention inhibits spontaneous hippocampal seizures in a mouse model of temporal lobe epilepsy. eNeuro 1(1): e.2014.CrossRefGoogle Scholar
Kros, L., Rooda, O. H. Eelkman, et al., (2015). Cerebellar output controls generalized spike-and-wave discharge occurrence. Ann Neurol 77(6): 10271049.CrossRefGoogle ScholarPubMed
Lammel, S., Tye, K. M., et al., (2014). Progress in understanding mood d isorders: optogenetic dissection of neural circuits. Genes Brain Behav 13(1): 3851.CrossRefGoogle Scholar
Montgomery, K. L., Yeh, A. J., et al., (2015). Wirelessly powered, fully internal optogenetics for brain, spinal and peripheral circuits in mice. Nat Methods 12(10): 969974.CrossRefGoogle ScholarPubMed
Nature Video. (2010). Method of the Year 2010: Optogenetics. from https://www.youtube.com/watch?v=I64X7vHSHOE.Google Scholar
Ozden, I., Wang, J., et al., (2013). A coaxial optrode as multifunction write-read probe for optogenetic studies in non-human primates. J Neurosci Methods 219(1): 142154.CrossRefGoogle ScholarPubMed
Paz, J. T., Davidson, T. J., et al., (2013). Closed-loop optogenetic control of thalamus as a tool for interrupting seizures after cortical injury. Nat Neurosci 16(1): 6470.CrossRefGoogle ScholarPubMed
Raimondo, J. V., Kay, L., et al., (2012). Optogenetic silencing strategies differ in their effects on inhibitory synaptic transmission. Nat Neurosci 15(8): 11021104.CrossRefGoogle ScholarPubMed
Tonnesen, J., Sorensen, A. T., et al., (2009). Optogenetic control of epileptiform activity. Proc Natl Acad Sci U S A 106(29): 1216212167.CrossRefGoogle ScholarPubMed
Wykes, R. C., Heeroma, J. H., et al., (2012). Optogenetic and potassium channel gene therapy in a rodent model of focal neocortical epilepsy. Sci Transl Med 4(161): 161ra152.CrossRefGoogle Scholar
Zemelman, B. V., Lee, G. A., et al., (2002). Selective photostimulation of genetically chARGed neurons. Neuron 33(1): 1522.CrossRefGoogle ScholarPubMed

References

Airan, RD, Thompson, KR, Fenno, LE, Bernstein, H, Deisseroth, K (2009) Temporally precise in vivo control of intracellular signalling. Nature 458:10251029.CrossRefGoogle ScholarPubMed
Attwell, D, Buchan, AM, Charpak, S, Lauritzen, M, Macvicar, BA, Newman, EA (2010) Glial and neuronal control of brain blood flow. Nature 468:232243.CrossRefGoogle ScholarPubMed
Bailes, HJ, Zhuang, LY, Lucas, RJ (2012) Reproducible and sustained regulation of Galphas signalling using a metazoan opsin as an optogenetic tool. PLoS One 7:e30774.CrossRefGoogle ScholarPubMed
Bekar, LK, He, W, Nedergaard, M (2008) Locus coeruleus alpha-adrenergic-mediated activation of cortical astrocytes in vivo. Cereb Cortex 18:27892795.CrossRefGoogle ScholarPubMed
Beppu, K, Sasaki, T, Tanaka, KF, Yamanaka, A, Fukazawa, Y, Shigemoto, R, Matsui, K (2014) Optogenetic countering of glial acidosis suppresses glial glutamate release and ischemic brain damage. Neuron 81:314320.CrossRefGoogle ScholarPubMed
Berndt, A, Schoenenberger, P, Mattis, J, Tye, KM, Deisseroth, K, Hegemann, P, Oertner, TG (2011) High-efficiency channelrhodopsins for fast neuronal stimulation at low light levels. Proc Natl Acad Sci U S A 108:75957600.CrossRefGoogle ScholarPubMed
Berndt, A, Yizhar, O, Gunaydin, LA, Hegemann, P, Deisseroth, K (2009) Bi-stable neural state switches. Nat Neurosci 12:229234.CrossRefGoogle ScholarPubMed
Chen, J, Tan, Z, Zeng, L, Zhang, X, He, Y, Gao, W, Wu, X, Li, Y, Bu, B, Wang, W, Duan, S (2013) Heterosynaptic long-term depression mediated by ATP released from astrocytes. Glia 61:178191.CrossRefGoogle ScholarPubMed
Christie, IN, Wells, JA, Southern, P, Marina, N, Kasparov, S, Gourine, AV, Lythgoe, MF (2013) fMRI response to blue light delivery in the naive brain: implications for combined optogenetic fMRI studies. Neuroimage 66:634641.CrossRefGoogle ScholarPubMed
Chuong, AS, et al. (2014) Noninvasive optical inhibition with a red-shifted microbial rhodopsin. Nat Neurosci 17:11231129.CrossRefGoogle ScholarPubMed
Deisseroth, K, Feng, G, Majewska, AK, Miesenbock, G, Ting, A, Schnitzer, MJ (2006) Next-generation optical technologies for illuminating genetically targeted brain circuits. J Neurosci 26:1038010386.CrossRefGoogle ScholarPubMed
Duale, H, Kasparov, S, Paton, JF, Teschemacher, AG (2005) Differences in transductional tropism of adenoviral and lentiviral vectors in the rat brainstem. Exp Physiol 90:7178.CrossRefGoogle ScholarPubMed
Ferrero, JJ, Alvarez, AM, Ramirez-Franco, J, Godino, MC, Bartolome-Martin, D, Aguado, C, Torres, M, Lujan, R, Ciruela, F, Sanchez-Prieto, J (2013) Beta-adrenergic receptors activate exchange protein directly activated by cAMP (Epac), translocate Munc13-1, and enhance the Rab3A–RIM1alpha interaction to potentiate glutamate release at cerebrocortical nerve terminals. J Biol Chem 288:3137031385.CrossRefGoogle ScholarPubMed
Fields, RD, Burnstock, G (2006) Purinergic signalling in neuron-glia interactions. Nat Rev Neurosci 7:423436.CrossRefGoogle ScholarPubMed
Figueiredo, M, Lane, S, Stout, RF Jr., Liu, B, Parpura, V, Teschemacher, AG, Kasparov, S (2014) Comparative analysis of optogenetic actuators in cultured astrocytes. Cell Calcium 56:208214.CrossRefGoogle ScholarPubMed
Figueiredo, M, Lane, S, Tang, F, Liu, BH, Hewinson, J, Marina, N, Kasymov, V, Souslova, EA, Chudakov, DM, Gourine, AV, Teschemacher, AG, Kasparov, S (2011) Optogenetic experimentation on astrocytes. Exp Physiol 96:4050.CrossRefGoogle ScholarPubMed
Fleischer, W, Theiss, S, Slotta, J, Holland, C, Schnitzler, A (2015) High-frequency voltage oscillations in cultured astrocytes. Physiol Rep 3:e12400.CrossRefGoogle ScholarPubMed
Gourine, AV, Kasymov, V, Marina, N, Tang, F, Figueiredo, MF, Lane, S, Teschemacher, AG, Spyer, KM, Deisseroth, K, Kasparov, S (2010) Astrocytes control breathing through pH-dependent release of ATP. Science 329:571575.CrossRefGoogle ScholarPubMed
Gradinaru, V, Mogri, M, Thompson, KR, Henderson, JM, Deisseroth, K (2009) Optical deconstruction of parkinsonian neural circuitry. Science 324:354359.CrossRefGoogle ScholarPubMed
Han, X, Chow, BY, Zhou, H, Klapoetke, NC, Chuong, A, Rajimehr, R, Yang, A, Baratta, MV, Winkle, J, Desimone, R, Boyden, ES (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
Hertz, L, Dringen, R, Schousboe, A, Robinson, SR (1999) Astrocytes: glutamate producers for neurons. J Neurosci Res 57:417428.3.0.CO;2-N>CrossRefGoogle ScholarPubMed
Hertz, L, Lovatt, D, Goldman, SA, Nedergaard, M (2010) Adrenoceptors in brain: cellular gene expression and effects on astrocytic metabolism and [Ca2+]i. Neurochem Int 57:411420.CrossRefGoogle Scholar
Hill, RA, Tong, L, Yuan, P, Murikinati, S, Gupta, S, Grutzendler, J (2015) Regional blood flow in the normal and ischemic brain is controlled by arteriolar smooth muscle cell contractility and not by capillary pericytes. Neuron 87:95110.CrossRefGoogle Scholar
Huckstepp, RT, id Bihi, R, Eason, R, Spyer, KM, Dicke, N, Willecke, K, Marina, N, Gourine, AV, Dale, N (2010) Connexin hemichannel-mediated CO2-dependent release of ATP in the medulla oblongata contributes to central respiratory chemosensitivity. J Physiol 588:39013920.CrossRefGoogle ScholarPubMed
Ji, ZG, Wang, H (2015) Optogenetic control of astrocytes: is it possible to treat astrocyte-related epilepsy? Brain Res Bull 110:2025.CrossRefGoogle ScholarPubMed
Kleinlogel, S, Feldbauer, K, Dempski, RE, Fotis, H, Wood, PG, Bamann, C, Bamberg, E (2011) Ultra light-sensitive and fast neuronal activation with the Ca2+-permeable channelrhodopsin CatCh. Nat Neurosci 14:513518.CrossRefGoogle ScholarPubMed
Latour, I, Hamid, J, Beedle, AM, Zamponi, GW, Macvicar, BA (2003) Expression of voltage-gated Ca2+ channel subtypes in cultured astrocytes. Glia 41:347353.CrossRefGoogle ScholarPubMed
Li, D, Agulhon, C, Schmidt, E, Oheim, M, Ropert, N (2013) New tools for investigating astrocyte-to-neuron communication. Front Cell Neurosci 7:193.CrossRefGoogle ScholarPubMed
Li, D, Herault, K, Isacoff, EY, Oheim, M, Ropert, N (2012) Optogenetic activation of LiGluR-expressing astrocytes evokes anion channel-mediated glutamate release. J Physiol 590:855873.CrossRefGoogle ScholarPubMed
Lin, JY (2011) A user’s guide to channelrhodopsin variants: features, limitations and future developments. Exp Physiol 96:1925.CrossRefGoogle ScholarPubMed
Lin, JY, Knutsen, PM, Muller, A, Kleinfeld, D, Tsien, RY (2013) ReaChR: a red-shifted variant of channelrhodopsin enables deep transcranial optogenetic excitation. Nat Neurosci 16:14991508.CrossRefGoogle ScholarPubMed
Liu, B, Paton, JF, Kasparov, S (2008) Viral vectors based on bidirectional cell-specific mammalian promoters and transcriptional amplification strategy for use in vitro and in vivo. BMC Biotechnol 8:49.CrossRefGoogle ScholarPubMed
Liu, BH, Yang, Y, Paton, JF, Li, F, Boulaire, J, Kasparov, S, Wang, S (2006) GAL4-NF-kappaB fusion protein augments transgene expression from neuronal promoters in the rat brain. Mol Ther 14:872882.CrossRefGoogle ScholarPubMed
Marina, N, Ang, R, Machhada, A, Kasymov, V, Karagiannis, A, Hosford, PS, Mosienko, V, Teschemacher, AG, Vihko, P, Paton, JF, Kasparov, S, Gourine, AV (2015) Brainstem hypoxia contributes to the development of hypertension in the spontaneously hypertensive rat. Hypertension 65:775783.CrossRefGoogle Scholar
Marina, N, Tang, F, Figueiredo, M, Mastitskaya, S, Kasimov, V, Mohamed-Ali, V, Roloff, E, Teschemacher, AG, Gourine, AV, Kasparov, S (2013) Purinergic signalling in the rostral ventro-lateral medulla controls sympathetic drive and contributes to the progression of heart failure following myocardial infarction in rats. Basic Res Cardiol 108:317.CrossRefGoogle Scholar
Masamoto, K, Unekawa, M, Watanabe, T, Toriumi, H, Takuwa, H, Kawaguchi, H, Kanno, I, Matsui, K, Tanaka, KF, Tomita, Y, Suzuki, N (2015) Unveiling astrocytic control of cerebral blood flow with optogenetics. Sci Rep 5:11455.CrossRefGoogle ScholarPubMed
Mohanty, SK, Reinscheid, RK, Liu, X, Okamura, N, Krasieva, TB, Berns, MW (2008) In-depth activation of channelrhodopsin 2-sensitized excitable cells with high spatial resolution using two-photon excitation with a near-infrared laser microbeam. Biophys J 95:39163926.CrossRefGoogle ScholarPubMed
Nagel, G, Brauner, M, Liewald, JF, Adeishvili, N, Bamberg, E, Gottschalk, A (2005a) Light activation of channelrhodopsin-2 in excitable cells of Caenorhabditis elegans triggers rapid behavioral responses. Curr Biol 15:22792284.CrossRefGoogle ScholarPubMed
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 U S A 100:1394013945.CrossRefGoogle ScholarPubMed
Nagel, G, Szellas, T, Kateriya, S, Adeishvili, N, Hegemann, P, Bamberg, E (2005b) Channelrhodopsins: directly light-gated cation channels. Biochem Soc Trans 33:863866.CrossRefGoogle ScholarPubMed
Pan, ZH, Ganjawala, TH, Lu, Q, Ivanova, E, Zhang, Z (2014) ChR2 mutants at L132 and T159 with improved operational light sensitivity for vision restoration. PLoS One 9:e98924.CrossRefGoogle ScholarPubMed
Papagiakoumou, E, Anselmi, F, Begue, A, de Sars, V, Gluckstad, J, Isacoff, EY, Emiliani, V (2010) Scanless two-photon excitation of channelrhodopsin-2. Nat Methods 7:848854.CrossRefGoogle ScholarPubMed
Perea, G, Yang, A, Boyden, ES, Sur, M (2014) Optogenetic astrocyte activation modulates response selectivity of visual cortex neurons in vivo. Nat Commun 5:3262.CrossRefGoogle ScholarPubMed
Prakash, R, Yizhar, O, Grewe, B, Ramakrishnan, C, Wang, N, Goshen, I, Packer, AM, Peterka, DS, Yuste, R, Schnitzer, MJ, Deisseroth, K (2012) Two-photon optogenetic toolbox for fast inhibition, excitation and bistable modulation. Nat Methods 9:11711179.CrossRefGoogle ScholarPubMed
Sasaki, T, Beppu, K, Tanaka, KF, Fukazawa, Y, Shigemoto, R, Matsui, K (2012) Application of an optogenetic byway for perturbing neuronal activity via glial photostimulation. Proc Natl Acad Sci U S A 109:2072020725.CrossRefGoogle ScholarPubMed
Steinhäuser, C, Seifert, G, Deitmer, JW (2013) Physiology of astrocytes: Ion channels and ion transporters. In: Neuroglia (Kettenmann, H, Ransom, B, eds), pp. 185196. Oxford: Oxford University Press.Google Scholar
Stierl, M, Stumpf, P, Udwari, D, Gueta, R, Hagedorn, R, Losi, A, Gartner, W, Petereit, L, Efetova, M, Schwarzel, M, Oertner, TG, Nagel, G, Hegemann, P (2011) Light-modulation of cellular cAMP by a small bacterial photoactivated adenylyl cyclase, bPAC, of the soil bacterium Beggiatoa. J Biol Chem 286:11811188.CrossRefGoogle ScholarPubMed
Suzuki, A, Stern, SA, Bozdagi, O, Huntley, GW, Walker, RH, Magistretti, PJ, Alberini, CM (2011) Astrocyte-neuron lactate transport is required for long-term memory formation. Cell 144:810823.CrossRefGoogle ScholarPubMed
Tang, F, Lane, S, Korsak, A, Paton, JF, Gourine, AV, Kasparov, S, Teschemacher, AG (2014) Lactate-mediated glia-neuronal signaling in the mammalian brain. Nat Commun 5:3284.CrossRefGoogle ScholarPubMed
Verkhratsky, A, Nedergaard, M, Hertz, L (2015) Why are astrocytes important? Neurochem Res 40:389401.CrossRefGoogle ScholarPubMed
Volgraf, M, Gorostiza, P, Numano, R, Kramer, RH, Isacoff, EY, Trauner, D (2006) Allosteric control of an ionotropic glutamate receptor with an optical switch. Nat Chem Biol 2:4752.CrossRefGoogle ScholarPubMed
Wang, S, Benamer, N, Zanella, S, Kumar, NN, Shi, Y, Bevengut, M, Penton, D, Guyenet, PG, Lesage, F, Gestreau, C, Barhanin, J, Bayliss, DA (2013) TASK-2 channels contribute to pH sensitivity of retrotrapezoid nucleus chemoreceptor neurons. J Neurosci 33:1603316044.CrossRefGoogle ScholarPubMed
Wells, JA, Christie, IN, Hosford, PS, Huckstepp, RT, Angelova, PR, Vihko, P, Cork, SC, Abramov, AY, Teschemacher, AG, Kasparov, S, Lythgoe, MF, Gourine, AV (2015) A critical role for purinergic signalling in the mechanisms underlying generation of BOLD fMRI responses. J Neurosci 35:52845292.CrossRefGoogle ScholarPubMed
Yarkoni, O, Donlon, L, Frankel, D (2012) Creating a bio-hybrid signal transduction pathway: opening a new channel of communication between cells and machines. Bioinspir Biomim 7:046017.CrossRefGoogle ScholarPubMed
Zemelman, BV, Nesnas, N, Lee, GA, Miesenbock, G (2003) Photochemical gating of heterologous ion channels: remote control over genetically designated populations of neurons. Proc Natl Acad Sci U S A 100:13521357.CrossRefGoogle ScholarPubMed

References

Apergis-Schoute, J., Iordanidou, P., Faure, C., Jego, S., Schöne, C., Aitta-Aho, T., Adamantidis, A., Burdakov, D. (2015). Optogenetic evidence for inhibitory signaling from orexin to MCH neurons via local microcircuits. Journal of Neuroscience, 35, 54355441.CrossRefGoogle ScholarPubMed
Aponte, Y., Atasoy, D., Sternson, S.M. (2011). AGRP neurons are sufficient to orchestrate feeding behavior rapidly and without training. Nature Neuroscience, 14, 351355.CrossRefGoogle ScholarPubMed
Atasoy, D., Betley, J.N., Su, H.H., Sternson, S.M. (2012). Deconstruction of a neural circuit for hunger. Nature, 488, 172177.CrossRefGoogle ScholarPubMed
Busnelli, M., Saulière, A., Manning, M., Bouvier, M., Galés, C., Chini, B. (2012). Functional selective oxytocin-derived agonists discriminate between individual G protein family subtypes. Journal of Biological Chemistry, 287, 36173629.CrossRefGoogle ScholarPubMed
Choe, H.K., Reed, M.D., Benavidez, N., Montgomery, D., Soares, N., Yim, Y.S., Choi, G.B. (2015). Oxytocin mediates entrainment of sensory stimuli to social cues of opposing valence. Neuron, 87, 152163.CrossRefGoogle ScholarPubMed
de Kock, C.P., Wierda, K.D., Bosman, L.W., Min, R., Koksma, J.J., Mansvelder, H.D., Verhage, M., Brussaard, A.B. (2003). Somatodendritic secretion in oxytocin neurons is upregulated during the female reproductive cycle. Journal of Neuroscience, 23, 27262734.CrossRefGoogle ScholarPubMed
Dreifuss, J.J. (1975). A review on neurosecretory granules: their contents and mechanisms of release. Annals of New York Academy of Sciences, 248, 184201.CrossRefGoogle ScholarPubMed
Eliava, M., Melchior, M., Knobloch-Bollmann, H.S., Wahis, J., da Silva Gouveia, M., Tang, Y., Ciobanu, A.C., Triana del Rio, R., Roth, L.C., Althammer, F., Chavant, V., Goumon, Y., Gruber, T., Busnelli, M., Chini, B., Tan, L., Mitre, M., Froemke, R.C., Chao, M.V., Giese, G., Sprengel, R., Kuner, R., Poisbeau, P., Seeburg, P.H., Stoop, R., Charlet, A., Grinevich, V. (2016). A new population of parvocellular oxytocin neurons controlling magnocellular neuron activity and inflammatory pain processing. Neuron, 89, 12911304.CrossRefGoogle ScholarPubMed
Eriksson, M., Ceccatelli, S., Uvnäs-Moberg, K., Iadarola, M., Hökfelt, T. (1996). Expression of fos-related antigens, oxytocin, dynorphin and galanin in the paraventricular and supraoptic nuclei of lactating rats. Neuroendocrinology, 63, 356367.CrossRefGoogle ScholarPubMed
Fields, R.L, Ponzio, T.A., Kawasaki, M., Gainer, H. (2012). Cell-type specific oxytocin gene expression from AAV delivered promoter deletion constructs into the rat supraoptic nucleus in vivo. PLoS One, 7, e32085.CrossRefGoogle ScholarPubMed
Grinevich, V., Desarménien, M.G., Chini, B., Tauber, M., Muscatelli, F. (2015). Ontogenesis of oxytocin pathways in the mammalian brain: late maturation and psychosocial disorders. Frontiers in Neuroanatomy, 8, 164.CrossRefGoogle ScholarPubMed
Knobloch, H.S., Charlet, A., Hoffmann, L.C., Eliava, M., Khrulev, S., Cetin, A.H., Osten, P., Schwarz, M. K., Seeburg, P.H., Stoop, R., Grinevich, V. (2012). Evoked axonal oxytocin release in the central amygdala attenuates fear response. Neuron, 73, 553566.CrossRefGoogle ScholarPubMed
Landry, M., Vila-Porcile, E., Hökfelt, T., Calas, A. (2003). Differential routing of coexisting neuropeptides in vasopressin neurons. European Journal of Neuroscience, 17, 579589.CrossRefGoogle ScholarPubMed
Lee, H.J., Macbeth, A.H., Pagani, J.H., Young, W.S. 3rd. (2009). Oxytocin: the great facilitator of life. Progress in Neurobiology, 88, 127151.Google ScholarPubMed
Ludwig, M., Sabatier, N., Dayanithi, G., Russell, J.A., Leng, G. (2002). The active role of dendrites in the regulation of magnocellular neurosecretory cell behavior. Progress in Brain Research, 139, 247256.CrossRefGoogle ScholarPubMed
Ludwig, M., Bull, P.M., Tobin, V.A., Sabatier, N., Landgraf, R., Dayanithi, G., Leng, G. (2005). Regulation of activity-dependent dendritic vasopressin release from rat supraoptic neurons. Journal of Physiology, 564, 515522.CrossRefGoogle Scholar
Marlin, B.J., Mitre, M., D’amour, J.A., Chao, M.V., Froemke, R.C. (2015). Oxytocin enables maternal behaviour by balancing cortical inhibition. Nature, 520, 499504.CrossRefGoogle ScholarPubMed
McCall, J.G., Al-Hasani, R., Siuda, E.R., Hong, D.Y., Norris, A.J., Ford, C.P., Bruchas, M.R. (2015). CRH engagement of the locus coeruleus noradrenergic system mediates stress-Induced anxiety. Neuron, 87, 605620.CrossRefGoogle ScholarPubMed
Mendell, L.M., Wall, P.D. (1965). Responses of single dorsal cord cells to peripheral cutaneous unmyelinated fibers. Nature, 206, 9799.CrossRefGoogle Scholar
Miesenböck, G., De Angelis, D.A., Rothman, J. E. (1998). Vizualizing secretion and synaptic transmission with pH-sensitive green fluorescent protein. Nature, 394, 192195.CrossRefGoogle Scholar
Mong, J.A, Pfaff, D.W. (2004). Hormonal symphony: steroid orchestration of gene modules for sociosexual behaviors. Molecular Psychiatry, 9, 550556.CrossRefGoogle ScholarPubMed
Oliet, S.H., Piet, R., Poulain, D.A, Theodosis, D.T. (2004). Glial modulation of synaptic transmission: insights from the supraoptic nucleus of the hypothalamus. Glia, 47, 258267.CrossRefGoogle ScholarPubMed
Owen, S.F., Tuncdemir, S.N., Bader, P.L., Tirko, N.N., Fishell, G., Tsien, R.W. (2013). Oxytocin enhances hippocampal spike transmission by modulating fast-spiking interneurons. Nature, 500, 458462.CrossRefGoogle ScholarPubMed
Park, S.J., Borghuis, B.G., Rahmani, P., Zeng, Q., Kim, I.J., Demb, J.B. (2015). Function and circuitry of VIP+ interneurons in the mouse retina. Journal of Neuroscience, 35, 1068510700.CrossRefGoogle ScholarPubMed
Tovote, P., Lüthi, A. (2012). Curbing fear by axonal oxytocin release in the amygdala. Neuron, 73, 407410.CrossRefGoogle ScholarPubMed
Scott, N., Prigge, M., Yizhar, O., Kimchi, T. (2015). A sexually dimorphic hypothalamic circuit controls maternal care and oxytocin secretion. Nature, 525, 519522.CrossRefGoogle ScholarPubMed
Sears, R.M., Fink, A.E., Wigestrand, M.B., Farb, C.R., de Lecea, L., Ledoux, J.E. (2013). Orexin/hypocretin system modulates amygdala-dependent threat learning through the locus coeruleus. Proceeding of National Academy of Sciences of USA, 110, 2026020265.CrossRefGoogle ScholarPubMed
Watson, S.J., Akil, H., Fischli, W., Goldstein, A., Zimmerman, E., Nilaver, G., van wimersma Griedanus, T.B. (1982). Dynorphin and vasopressin: common localization in magnocellular neurons. Science, 216, 8587.CrossRefGoogle ScholarPubMed
Wu, Z., Autry, A.E., Bergan, J.F., Watabe-Uchida, M., Dulac, C.G. (2014). Galanin neurons in the medial preoptic area govern parental behaviour. Nature, 509, 325330.CrossRefGoogle ScholarPubMed

References

Atasoy, D., Aponte, Y., Su, H.H. et al. (2008). A FLEX switch targets channelrhodopsin-2 to multiple cell types for imaging and long-range circuit mapping. The Journal of Neuroscience 28, 70257030.CrossRefGoogle ScholarPubMed
Berndt, A., Lee, S.Y., Ramakrishnan, C. et al. (2014). Structure-guided transformation of channelrhodopsin into a light-activated chloride channel. Science 344, 420424.CrossRefGoogle ScholarPubMed
Berndt, A., Yizhar, O., Gunaydin, L.A. et al. (2009). Bi-stable neural state switches. Nature Neuroscience 12, 229234.CrossRefGoogle ScholarPubMed
Chow, B.Y., Han, X., Dobry, A.S. et al. (2010). High-performance genetically targetable optical neural silencing by light-driven proton pumps. Nature 463, 98102.CrossRefGoogle ScholarPubMed
Chuong, A.S., Miri, M.L., Busskamp, V. et al. (2014). Noninvasive optical inhibition with a red-shifted microbial rhodopsin. Nature Neuroscience 17, 11231129.CrossRefGoogle ScholarPubMed
Gradinaru, V., Thompson, K.R., and Deisseroth, K. (2008). eNpHR: a Natronomonas halorhodopsin enhanced for optogenetic applications. Brain Cell Biology 36, 129139.CrossRefGoogle ScholarPubMed
Gradinaru, V., Zhang, F., Ramakrishnan, C. et al. (2010). Molecular and cellular approaches for diversifying and extending optogenetics. Cell 141, 154165.CrossRefGoogle ScholarPubMed
Gunaydin, L.A., Yizhar, O., Berndt, A. et al. (2010). Ultrafast optogenetic control. Nature Neuroscience 13, 387392.CrossRefGoogle ScholarPubMed
Herman, A.M., Huang, L., Murphey, D.K. et al. (2014). Cell type-specific and time-dependent light exposure contribute to silencing in neurons expressing channelrhodopsin-2. eLife 3, e01481.CrossRefGoogle ScholarPubMed
Klapoetke, N.C., Murata, Y., Kim, S.S. et al. (2014). Independent optical excitation of distinct neural populations. Nature Methods 11, 338346.CrossRefGoogle ScholarPubMed
Lin, J.Y. (2011). A user’s guide to channelrhodopsin variants: features, limitations and future developments. Experimental Physiology 96, 1925.CrossRefGoogle ScholarPubMed
Lin, J.Y., Knutsen, P.M., Muller, A. et al. (2013). ReaChR: a red-shifted variant of channelrhodopsin enables deep transcranial optogenetic excitation. Nature Neuroscience 16, 14991508.CrossRefGoogle ScholarPubMed
Mastakov, M.Y., Baer, K., Symes, C.W. et al. (2002). Immunological aspects of recombinant adeno-associated virus delivery to the mammalian brain. Journal of Virology 76, 84468454.CrossRefGoogle Scholar
Mattis, J., Tye, K.M., Ferenczi, E.A. et al. (2012). Principles for applying optogenetic tools derived from direct comparative analysis of microbial opsins. Nature Methods 9, 159172.CrossRefGoogle Scholar
Miyashita, T., Shao, Y.R., Chung, J. et al. (2013). Long-term channelrhodopsin-2 (ChR2) expression can induce abnormal axonal morphology and targeting in cerebral cortex. Frontiers in Neural Circuits 7, 8.Google ScholarPubMed
Nagel, G., Szellas, T., Huhn, W. et al. (2003). Channelrhodopsin-2, a directly light-gated cation-selective membrane channel. PNAS 100, 1394013945.CrossRefGoogle ScholarPubMed
Nathanson, J.L., Jappelli, R., Scheeff, E.D. et al. (2009). Short promoters in viral vectors drive selective expression in mammalian inhibitory neurons, but do not restrict activity to specific inhibitory cell-types. Frontiers in Neural Circuits 3, 19.CrossRefGoogle Scholar
Osakada, F., Mori, T., Cetin, A.H. et al. (2011). New rabies virus variants for monitoring and manipulating activity and gene expression in defined neural circuits. Neuron 71, 617631.CrossRefGoogle ScholarPubMed
Prakash, R., Yizhar, O., Grewe, B. et al. (2012). Two-photon optogenetic toolbox for fast inhibition, excitation and bistable modulation. Nature Methods 9, 11711179.CrossRefGoogle ScholarPubMed
Rajasethupathy, P., Sankaran, S., Marshel, J.H. et al. (2015). Projections from neocortex mediate top-down control of memory retrieval. Nature 526, 653659.CrossRefGoogle ScholarPubMed
Schobert, B., and Lanyi, J.K. (1982). Halorhodopsin is a light-driven chloride pump. The Journal of Biological Chemistry 257, 1030610313.CrossRefGoogle ScholarPubMed
Schwarz, L.A., Miyamichi, K., Gao, X.J. et al. (2015). Viral-genetic tracing of the input–output organization of a central noradrenaline circuit. Nature 524, 8892.CrossRefGoogle ScholarPubMed
Shaner, N.C., Steinbach, P.A., and Tsien, R.Y. (2005). A guide to choosing fluorescent proteins. Nature Methods 2, 905909.CrossRefGoogle ScholarPubMed
Shemiakina, II, Ermakova, G.V., Cranfill, P.J. et al. (2012). A monomeric red fluorescent protein with low cytotoxicity. Nature Communications 3, 1204.CrossRefGoogle ScholarPubMed
Soudais, C., Laplace-Builhe, C., Kissa, K. et al. (2001). Preferential transduction of neurons by canine adenovirus vectors and their efficient retrograde transport in vivo. FASEB Journal 15, 22832285.CrossRefGoogle ScholarPubMed
Takatoh, J., Nelson, A., Zhou, X. et al. (2013). New modules are added to vibrissal premotor circuitry with the emergence of exploratory whisking. Neuron 77, 346360.CrossRefGoogle ScholarPubMed
Tighilet, B., Hashikawa, T., and Jones, E.G. (1998). Cell- and lamina-specific expression and activity-dependent regulation of type II calcium/calmodulin-dependent protein kinase isoforms in monkey visual cortex. The Journal of Neuroscience 18, 21292146.CrossRefGoogle ScholarPubMed
Wickersham, I.R., Lyon, D.C., Barnard, R.J. et al. (2007). Monosynaptic restriction of transsynaptic tracing from single, genetically targeted neurons. Neuron 53, 639647.CrossRefGoogle ScholarPubMed
Yizhar, O., Fenno, L.E., Prigge, M. et al. (2011). Neocortical excitation/inhibition balance in information processing and social dysfunction. Nature 477, 171178.CrossRefGoogle ScholarPubMed
Zhang, F., Wang, L.P., Brauner, M. et al. (2007). Multimodal fast optical interrogation of neural circuitry. Nature 446, 633639.CrossRefGoogle ScholarPubMed

References

Adamantidis, A.R. et al., 2007. Neural substrates of awakening probed with optogenetic control of hypocretin neurons. Nature, 450(7168), pp.420–4.CrossRefGoogle ScholarPubMed
Anastasiades, P.G. et al., 2016. GABAergic interneurons form transient, layer-specific circuits in early postnatal neocortex. Nat Commun, 4, p.7.Google Scholar
Andrasfalvy, B.K. et al., 2010. Two-photon single-cell optogenetic control of neuronal activity by sculpted light. Proc Natl Acad Sci U S A, 107(26), pp.11981–6.CrossRefGoogle ScholarPubMed
Bamann, C. et al., 2008. Spectral characteristics of the photocycle of channelrhodopsin-2 and its implication for channel function. J Mol Biol, 375(3), pp.686–94.CrossRefGoogle ScholarPubMed
Bi, A. et al., 2006. Ectopic expression of a microbial-type rhodopsin restores visual responses in mice with photoreceptor degeneration. Neuron, 50(1), pp.2333.CrossRefGoogle ScholarPubMed
Binzegger, T., Douglas, R.J. & Martin, K.A., 2004. A quantitative map of the circuit of cat primary visual cortex. J Neurosci, 24(39), pp.8441–53.CrossRefGoogle ScholarPubMed
Bliss, T.V.P. & Lømo, T., 1973. Long-lasting potentiation of synaptic transmission in the dentate area of the anaesthetized rabbit following stimulation of the perforant path. The Journal of Physiology, 232(2), pp.331–56.Google ScholarPubMed
Boyden, E.S. et al., 2005. Millisecond-timescale, genetically targeted optical control of neural activity. Nat Neurosci, 8(9), pp.1263–8.CrossRefGoogle ScholarPubMed
Brill, J. & Huguenard, J.R., 2009. Robust short-latency perisomatic inhibition onto neocortical pyramidal cells detected by laser-scanning photostimulation. J Neurosci, 29(23), pp.7413–23.CrossRefGoogle ScholarPubMed
Brown, S.P. & Hestrin, S., 2009. Intracortical circuits of pyramidal neurons reflect their long-range axonal targets. Nature, 457(7233), pp.1133–6.CrossRefGoogle ScholarPubMed
Callaway, E.M. & Katz, L.C., 1993. Photostimulation using caged glutamate reveals functional circuitry in living brain slices. Proc Natl Acad Sci U S A, 90(16), pp.7661–5.CrossRefGoogle ScholarPubMed
Clyne, J.D. & Miesenböck, G., 2008. Sex-specific control and tuning of the pattern generator for courtship song in Drosophila. Cell, 133(2), pp.354–63.CrossRefGoogle ScholarPubMed
Crick, F., 1999. The impact of molecular biology on neuroscience. Philos Trans R Soc Lond B, 354, pp.2021–5.CrossRefGoogle ScholarPubMed
Dantzker, J.L. & Callaway, E.M., 2000. Laminar sources of synaptic input to cortical inhibitory interneurons and pyramidal neurons. Nat Neurosci, 3(7), pp.701–7.CrossRefGoogle ScholarPubMed
Deisseroth, K., 2014. Circuit dynamics of adaptive and maladaptive behaviour. Nature, 505(7483), pp.309–17.CrossRefGoogle ScholarPubMed
Denk, W., 1994. Two-photon scanning photochemical microscopy: mapping ligand-gated ion channel distributions. Proc Natl Acad Sci U S A, 91(14), pp.6629–33.CrossRefGoogle ScholarPubMed
Denk, W., Strickler, J.H. & Webb, W.W., 1990. Two-photon laser scanning fluorescence microscopy. Science, 248(4951), pp.73–6.CrossRefGoogle ScholarPubMed
Fino, E. & Yuste, R., 2011. Dense inhibitory connectivity in neocortex. Neuron, 69(6), pp.1188–203.CrossRefGoogle ScholarPubMed
Gu, Z. & Yakel, J.L., 2011. Timing-dependent septal cholinergic induction of dynamic hippocampal synaptic plasticity. Neuron, 71(1), pp.155–65.CrossRefGoogle ScholarPubMed
Gupta, A., Wang, Y. & Markram, H., 2000. Organizing principles for a diversity of GABAergic interneurons and synapses in the neocortex. Science, 287(5451), pp.273–8.CrossRefGoogle ScholarPubMed
Helmstaedter, M., Sakmann, B. & Feldmeyer, D., 2009a. L2/3 interneuron groups defined by multiparameter analysis of axonal projection, dendritic geometry, and electrical excitability. Cereb Cortex, 19(4), pp.951–62.Google ScholarPubMed
Helmstaedter, M., Sakmann, B. & Feldmeyer, D., 2009b. Neuronal correlates of local, lateral, and translaminar inhibition with reference to cortical columns. Cereb Cortex, 19(4), pp.926–37.CrossRefGoogle ScholarPubMed
Helmstaedter, M., Sakmann, B. & Feldmeyer, D., 2009c. The relation between dendritic geometry, electrical excitability, and axonal projections of L2/3 interneurons in rat barrel cortex. Cereb Cortex, 19(4), pp.938–50.Google ScholarPubMed
Katz, L.C. & Dalva, M.B., 1994. Scanning laser photostimulation: a new approach for analyzing brain circuits. J Neurosci Methods, 54(2), pp.205–18.CrossRefGoogle ScholarPubMed
Kätzel, D. et al., 2011. The columnar and laminar organization of inhibitory connections to neocortical excitatory cells. Nat Neurosci, 14(1), pp.100–7.CrossRefGoogle ScholarPubMed
Kätzel, D. & Miesenböck, G., 2014. Experience-dependent rewiring of specific inhibitory connections in adult neocortex. PLoS Biol, 12(2), p. e1001798.CrossRefGoogle ScholarPubMed
Kohl, M.M. et al., 2011. Hemisphere-specific optogenetic stimulation reveals left–right asymmetry of hippocampal plasticity. Nat Neurosci, 14, pp.1413–5.Google ScholarPubMed
Kozloski, J., Hamzei-Sichani, F. & Yuste, R., 2001. Stereotyped position of local synaptic targets in neocortex. Science, 293(5531), pp.868–72.CrossRefGoogle ScholarPubMed
Kuhlman, S.J. & Huang, Z.J., 2008. High-resolution labeling and functional manipulation of specific neuron types in mouse brain by Cre-activated viral gene expression. PLoS ONE, 3(4), p. e2005.CrossRefGoogle ScholarPubMed
Lesch, K.-P. & Waider, J., 2012. Serotonin in the modulation of neural plasticity and networks: implications for neurodevelopmental disorders. Neuron, 76(1), pp.175–91.CrossRefGoogle ScholarPubMed
Li, X. et al., 2005. Fast noninvasive activation and inhibition of neural and network activity by vertebrate rhodopsin and green algae channelrhodopsin. Proc Natl Acad Sci U S A, 102(49), pp.17816–21.CrossRefGoogle ScholarPubMed
Lima, S.Q. & Miesenböck, G., 2005. Remote control of behavior through genetically targeted photostimulation of neurons. Cell, 121(1), pp.141–52.CrossRefGoogle ScholarPubMed
Madisen, L. et al., 2012. A toolbox of Cre-dependent optogenetic transgenic mice for light-induced activation and silencing. Nat Neurosci, 15(5), pp.793802.CrossRefGoogle ScholarPubMed
Martin, K.A.C., 2009. The road ahead for brain-circuit reconstruction. Nature, 462(7272), pp.411.CrossRefGoogle ScholarPubMed
Martin, S.J., Grimwood, P.D., & Morris, R.G., 2000. Synaptic plasticity and memory: an evaluation of the hypothesis. Annu Rev Neurosci, 23(1), pp.649711.CrossRefGoogle ScholarPubMed
Matsuzaki, M. et al., 2001. Dendritic spine geometry is critical for AMPA receptor expression in hippocampal CA1 pyramidal neurons. Nat Neurosci, 4(11), pp.1086–92.CrossRefGoogle ScholarPubMed
Mohanty, S.K. et al., 2008. In-depth activation of channelrhodopsin 2-sensitized excitable cells with high spatial resolution using two-photon excitation with a near-infrared laser microbeam. Biophys J, 95(8), pp.3916–26.CrossRefGoogle ScholarPubMed
Monyer, H. & Markram, H., 2004. Interneuron diversity series: molecular and genetic tools to study GABAergic interneuron diversity and function. Trends Neurosci, 27(2), pp.90–7.CrossRefGoogle ScholarPubMed
Nabavi, S. et al., 2014. Engineering a memory with LTD and LTP. Nature, 511(7509), pp.348–52.CrossRefGoogle ScholarPubMed
Nagel, G. et al., 2003. Channelrhodopsin-2, a directly light-gated cation-selective membrane channel. Proc Natl Acad Sci U S A, 100(24), pp.13940–5.CrossRefGoogle ScholarPubMed
Nagel, G. et al., 2005. Light activation of channelrhodopsin-2 in excitable cells of Caenorhabditis elegans triggers rapid behavioral responses. Curr Biol, 15(24), pp.2279–84.CrossRefGoogle ScholarPubMed
Nikolenko, V. et al., 2008. SLM microscopy: scanless two-photon imaging and photostimulation using spatial light modulators. Front Neural Circuits, 2, p. 5.CrossRefGoogle ScholarPubMed
Nikolenko, V., Poskanzer, K.E. & Yuste, R., 2007. Two-photon photostimulation and imaging of neural circuits. Nat Methods, 4(11), pp.943–50.CrossRefGoogle ScholarPubMed
Packer, A.M. et al., 2012. Two-photon optogenetics of dendritic spines and neural circuits in 3D. Nat Methods, 9(12), pp.1202–5.CrossRefGoogle Scholar
Packer, A.M. & Yuste, R., 2011. Dense, unspecific connectivity of neocortical parvalbumin-positive interneurons: a canonical microcircuit for inhibition? J Neurosci, 31(37), pp.13260–71.CrossRefGoogle ScholarPubMed
Papagiakoumou, E. et al., 2010. Scanless two-photon excitation of channelrhodopsin-2. Nat Methods, 7(10), pp.848–54.CrossRefGoogle ScholarPubMed
Pascoli, V., Turiault, M. & Luscher, C., 2012. Reversal of cocaine-evoked synaptic potentiation resets drug-induced adaptive behaviour. Nature, 481(7379), pp.71–5.CrossRefGoogle Scholar
Petreanu, L. et al., 2007. Channelrhodopsin-2-assisted circuit mapping of long-range callosal projections. Nat Neurosci, 10(5), pp.663–8.CrossRefGoogle ScholarPubMed
Petreanu, L. et al., 2009. The subcellular organization of neocortical excitatory connections. Nature, 457(7233), pp.1142–5.CrossRefGoogle ScholarPubMed
Pignatelli, M. & Bonci, A., 2015. Role of dopamine neurons in reward and aversion: a synaptic plasticity perspective. Neuron, 86(5), pp.1145–57.CrossRefGoogle ScholarPubMed
Prakash, R. et al., 2012. Two-photon optogenetic toolbox for fast inhibition, excitation and bistable modulation. Nat Methods, 9(12), pp.1171–9.CrossRefGoogle ScholarPubMed
Reijmers, L.G. et al., 2007. Localization of a stable neural correlate of associative memory. Science, 317(5842), pp.1230–3.CrossRefGoogle ScholarPubMed
Rickgauer, J.P. & Tank, D.W., 2009. Two-photon excitation of channelrhodopsin-2 at saturation. Proc Natl Acad Sci U S A, 106(35), pp.15025–30.CrossRefGoogle ScholarPubMed
Rosen, Z.B., Cheung, S. & Siegelbaum, S.A., 2015. Midbrain dopamine neurons bidirectionally regulate CA3–CA1 synaptic drive. Nat Neurosci, 18(12), pp.1763–71.CrossRefGoogle ScholarPubMed
Scanziani, M. & Häusser, M., 2009. Electrophysiology in the age of light. Nature, 461(7266), pp.930–9.CrossRefGoogle ScholarPubMed
Schubert, D. et al., 2001. Layer-specific intracolumnar and transcolumnar functional connectivity of layer V pyramidal cells in rat barrel cortex. J Neurosci, 21(10), pp.3580–92.CrossRefGoogle Scholar
Schubert, D., Kötter, R. & Staiger, J., 2007. Mapping functional connectivity in barrel-related columns reveals layer- and cell type-specific microcircuits. Brain Struct Funct, 212(2), pp.107–19.CrossRefGoogle ScholarPubMed
Seol, G.H. et al., 2007. Neuromodulators control the polarity of spike-timing-dependent synaptic plasticity. Neuron, 55(6), pp.919–29.CrossRefGoogle ScholarPubMed
Shepherd, G.M. et al., 2005. Geometric and functional organization of cortical circuits. Nat Neurosci, 8(6), pp.782–90.CrossRefGoogle ScholarPubMed
Shepherd, G.M. & Svoboda, K., 2005. Laminar and columnar organization of ascending excitatory projections to layer 2/3 pyramidal neurons in rat barrel cortex. J Neurosci, 25(24), pp.5670–9.CrossRefGoogle ScholarPubMed
Shipton, O.A. et al., 2014. Left–right dissociation of hippocampal memory processes in mice. Proc Natl Acad Sci U S A, 111(42), pp.15238–43.CrossRefGoogle ScholarPubMed
Teles-Grilo Ruivo, L. & Mellor, J., 2013. Cholinergic modulation of hippocampal network function. Front Synaptic Neurosci, 5, p. 2.CrossRefGoogle ScholarPubMed
Thomson, A.M. et al., 1996. Single axon IPSPs elicited in pyramidal cells by three classes of interneurones in slices of rat neocortex. J Physiol, 496(Pt 1), pp.81102.CrossRefGoogle ScholarPubMed
Thomson, A.M. et al., 2002. Synaptic connections and small circuits involving excitatory and inhibitory neurons in layers 2–5 of adult rat and cat neocortex: triple intracellular recordings and biocytin labelling in vitro. Cereb Cortex, 12(9), pp.936–53.CrossRefGoogle ScholarPubMed
Thomson, A.M., Deuchars, J. & West, D.C., 1996. Neocortical local synaptic circuitry revealed with dual intracellular recordings and biocytin-filling. J Physiol Paris, 90(3–4), pp.211–5.CrossRefGoogle ScholarPubMed
Walker, J.W., McCray, J.A. & Hess, G.P., 1986. Photolabile protecting groups for an acetylcholine receptor ligand. Synthesis and photochemistry of a new class of o-nitrobenzyl derivatives and their effects on receptor function. Biochemistry, 25(7), pp.1799–805.CrossRefGoogle ScholarPubMed
Wang, Y. et al., 2004. Anatomical, physiological and molecular properties of Martinotti cells in the somatosensory cortex of the juvenile rat. J Physiol, 561(1), pp.6590.CrossRefGoogle ScholarPubMed
Weiler, N. et al., 2008. Top-down laminar organization of the excitatory network in motor cortex. Nat Neurosci, 11(3), pp.360–6.CrossRefGoogle ScholarPubMed
Wieboldt, R. et al., 1994a. Photolabile precursors of glutamate: synthesis, photochemical properties, and activation of glutamate receptors on a microsecond time scale. Proc Natl Acad Sci U S A, 91(19), pp.8752–6.CrossRefGoogle ScholarPubMed
Wieboldt, R. et al., 1994b. Synthesis and photochemistry of photolabile derivatives of gamma-aminobutyric acid for chemical kinetic investigations of the gamma-aminobutyric acid receptor in the millisecond time region. Biochemistry, 33(6), pp.1526–33.CrossRefGoogle ScholarPubMed
Wilcox, M. et al., 1990. Synthesis of photolabile precursors of amino acid neurotransmitters. J Org Chem, 55(5), pp.1585–9.CrossRefGoogle Scholar
Xiong, W. & Jin, X., 2012. Optogenetic field potential recording in cortical slices. J Neurosci Methods, 210(2), pp.119–24.CrossRefGoogle ScholarPubMed
Xu, X. & Callaway, E.M., 2009. Laminar specificity of functional input to distinct types of inhibitory cortical neurons. J Neurosci, 29(1), pp.7085.CrossRefGoogle ScholarPubMed
Yoshimura, Y. & Callaway, E.M., 2005. Fine-scale specificity of cortical networks depends on inhibitory cell type and connectivity. Nat Neurosci, 8(11), pp.1552–9.CrossRefGoogle ScholarPubMed
Yoshimura, Y., Dantzker, J.L. & Callaway, E.M., 2005. Excitatory cortical neurons form fine-scale functional networks. Nature, 433(7028), pp.868–73.CrossRefGoogle ScholarPubMed
Zemelman, B.V. et al., 2003. Photochemical gating of heterologous ion channels: remote control over genetically designated populations of neurons. Proc Natl Acad Sci U S A, 100(3), pp.1352–7.CrossRefGoogle ScholarPubMed
Zemelman, B.V. et al., 2002. Selective photostimulation of genetically chARGed neurons. Neuron, 33(1), pp.1522.CrossRefGoogle ScholarPubMed
Zemelman, B.V. & Miesenböck, G., 2001. Genetic schemes and schemata in neurophysiology. Curr Opin Neurobiol, 11(4), pp.409–14.CrossRefGoogle ScholarPubMed
Zhu, P. et al., 2009. Optogenetic dissection of neuronal circuits in zebrafish using viral gene transfer and the Tet system. Front Neural Circuits, 3, p. 21.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
×