Structural foundations of optogenetics: Determinants of channelrhodopsin ion selectivity

Significance Channelrhodopsins are membrane proteins that enable cellular regulation of transmembrane ion conductance through light-gated pores; these proteins have found application in optogenetics. This paper tests the hypothesis that selectivity of channelrhodopsins is determined by surface potential of the pore region: Cations are conducted by a negatively charged pore, and chloride ions are conducted by a pore that has neutral and positively charged residues. In confirming this hypothesis and applying the resulting principles, we engineer improved chloride-conducting channels with higher chloride selectivity and conductivity. We also provide insights into the distinct mechanisms underlying inhibition mediated by higher-efficiency chloride channels compared with ion pumps. Finally, we demonstrate initial utility of light-gated microbial opsin-based chloride channels in controlling behavior of freely moving animals. The structure-guided design of chloride-conducting channelrhodopsins has illuminated mechanisms underlying ion selectivity of this remarkable family of light-activated ion channels. The first generation of chloride-conducting channelrhodopsins, guided in part by development of a structure-informed electrostatic model for pore selectivity, included both the introduction of amino acids with positively charged side chains into the ion conduction pathway and the removal of residues hypothesized to support negatively charged binding sites for cations. Engineered channels indeed became chloride selective, reversing near −65 mV and enabling a new kind of optogenetic inhibition; however, these first-generation chloride-conducting channels displayed small photocurrents and were not tested for optogenetic inhibition of behavior. Here we report the validation and further development of the channelrhodopsin pore model via crystal structure-guided engineering of next-generation light-activated chloride channels (iC++) and a bistable variant (SwiChR++) with net photocurrents increased more than 15-fold under physiological conditions, reversal potential further decreased by another ∼15 mV, inhibition of spiking faithfully tracking chloride gradients and intrinsic cell properties, strong expression in vivo, and the initial microbial opsin channel-inhibitor–based control of freely moving behavior. We further show that inhibition by light-gated chloride channels is mediated mainly by shunting effects, which exert optogenetic control much more efficiently than the hyperpolarization induced by light-activated chloride pumps. The design and functional features of these next-generation chloride-conducting channelrhodopsins provide both chronic and acute timescale tools for reversible optogenetic inhibition, confirm fundamental predictions of the ion selectivity model, and further elucidate electrostatic and steric structure–function relationships of the light-gated pore.

[1]  Satoshi P. Tsunoda,et al.  Conversion of Channelrhodopsin into a Light-Gated Chloride Channel , 2014, Science.

[2]  Lief E. Fenno,et al.  Principles for applying optogenetic tools derived from direct comparative analysis of microbial opsins , 2011, Nature Methods.

[3]  S. Lummis,et al.  Conversion of the Ion Selectivity of the 5-HT3AReceptor from Cationic to Anionic Reveals a Conserved Feature of the Ligand-gated Ion Channel Superfamily* , 2001, The Journal of Biological Chemistry.

[4]  Alcino J. Silva,et al.  CREB regulates excitability and the allocation of memory to subsets of neurons in the amygdala , 2009, Nature Neuroscience.

[5]  K. Deisseroth Circuit dynamics of adaptive and maladaptive behaviour , 2014, Nature.

[6]  Hideaki E. Kato,et al.  Crystal structure of the channelrhodopsin light-gated cation channel , 2012, Nature.

[7]  V. Stein,et al.  Molecular structure and physiological function of chloride channels. , 2002, Physiological reviews.

[8]  S. Ikemoto,et al.  Similar Roles of Substantia Nigra and Ventral Tegmental Dopamine Neurons in Reward and Aversion , 2014, The Journal of Neuroscience.

[9]  Bruno Bontempi,et al.  Selective Erasure of a Fear Memory , 2009, Science.

[10]  Lief E. Fenno,et al.  Neocortical excitation/inhibition balance in information processing and social dysfunction , 2011, Nature.

[11]  P. Hegemann,et al.  The nature of rhodopsin-triggered photocurrents in Chlamydomonas. I. Kinetics and influence of divalent ions. , 1996, Biophysical journal.

[12]  René Hen,et al.  Targeted gene expression in dopamine and serotonin neurons of the mouse brain , 2005, Journal of Neuroscience Methods.

[13]  Jadin C. Jackson,et al.  Quantitative measures of cluster quality for use in extracellular recordings , 2005, Neuroscience.

[14]  J. Spudich,et al.  Natural light-gated anion channels: A family of microbial rhodopsins for advanced optogenetics , 2015, Science.

[15]  Y. Jan,et al.  TMEM16F Forms a Ca2+-Activated Cation Channel Required for Lipid Scrambling in Platelets during Blood Coagulation , 2012, Cell.

[16]  Lief E. Fenno,et al.  The Microbial Opsin Family of Optogenetic Tools , 2011, Cell.

[17]  Lloyd H. Michael,et al.  The Guide for the Care and Use of Laboratory Animals. , 2016, ILAR journal.

[18]  Joseph E LeDoux Emotion circuits in the brain. , 2009, Annual review of neuroscience.

[19]  Nathan A. Baker,et al.  Electrostatics of nanosystems: Application to microtubules and the ribosome , 2001, Proceedings of the National Academy of Sciences of the United States of America.

[20]  Stephen Maren The Amygdala, Synaptic Plasticity, and Fear Memory , 2003, Annals of the New York Academy of Sciences.

[21]  K. Deisseroth,et al.  Manipulating a “Cocaine Engram” in Mice , 2014, The Journal of Neuroscience.

[22]  Kwang-Hwi Cho,et al.  BIOPHYSICS AND COMPUTATIONAL BIOLOGY , 2009 .

[23]  K. L. Montgomery,et al.  Virally mediated optogenetic excitation and inhibition of pain in freely moving non-transgenic mice , 2014, Nature Biotechnology.

[24]  Karl Deisseroth,et al.  Structure-Guided Transformation of Channelrhodopsin into a Light-Activated Chloride Channel , 2014, Science.

[25]  George Paxinos,et al.  The Mouse Brain in Stereotaxic Coordinates , 2001 .

[26]  E. Bamberg,et al.  Channelrhodopsin-1: A Light-Gated Proton Channel in Green Algae , 2002, Science.

[27]  Jean-Luc Galzi,et al.  Mutations in the channel domain of a neuronal nicotinic receptor convert ion selectivity from cationic to anionic , 1992, Nature.

[28]  K. Deisseroth,et al.  Bi-stable neural state switches , 2009, Nature Neuroscience.

[29]  S. Arber,et al.  A Developmental Switch in the Response of DRG Neurons to ETS Transcription Factor Signaling , 2005, PLoS biology.

[30]  Kelly R. Tan,et al.  GABA Neurons of the VTA Drive Conditioned Place Aversion , 2012, Neuron.

[31]  Paul W. Frankland,et al.  Neurons Are Recruited to a Memory Trace Based on Relative Neuronal Excitability Immediately before Training , 2014, Neuron.

[32]  Amene Shahrokhi,et al.  Oxytocin-Mediated GABA Inhibition During Delivery Attenuates Autism Pathogenesis in Rodent Offspring , 2014, Science.

[33]  L. Jan,et al.  International Union of Basic and Clinical Pharmacology. LXXXV: Calcium-Activated Chloride Channels , 2012, Pharmacological Reviews.

[34]  John F. Guzowski,et al.  Neuronal Competition and Selection During Memory Formation , 2006, Science.

[35]  Zengcai V. Guo,et al.  A motor cortex circuit for motor planning and movement , 2015, Nature.

[36]  Massimo Scanziani,et al.  An improved chloride-conducting channelrhodopsin for light-induced inhibition of neuronal activity in vivo , 2015, Scientific Reports.

[37]  A. Rowan Guide for the Care and Use of Laboratory Animals , 1979 .

[38]  Gisella Vetere,et al.  Spine growth in the anterior cingulate cortex is necessary for the consolidation of contextual fear memory , 2011, Proceedings of the National Academy of Sciences.

[39]  Michael Davis,et al.  The role of the amygdala in fear and anxiety. , 1992, Annual review of neuroscience.

[40]  Y. Ben-Ari,et al.  Multiple facets of GABAergic neurons and synapses: multiple fates of GABA signalling in epilepsies , 2005, Trends in Neurosciences.

[41]  Karl Deisseroth,et al.  Expanding the optogenetics toolkit , 2018, Nature Methods.

[42]  K. Deisseroth,et al.  Molecular and Cellular Approaches for Diversifying and Extending Optogenetics , 2010, Cell.

[43]  P. Jonas,et al.  Shunting Inhibition Improves Robustness of Gamma Oscillations in Hippocampal Interneuron Networks by Homogenizing Firing Rates , 2006, Neuron.

[44]  B. Hoffer,et al.  Characterization of a mouse strain expressing Cre recombinase from the 3′ untranslated region of the dopamine transporter locus , 2006, Genesis.

[45]  Gisella Vetere,et al.  MEF2 negatively regulates learning-induced structural plasticity and memory formation , 2012, Nature Neuroscience.

[46]  Kevin J. Staley,et al.  Differences in Cortical versus Subcortical GABAergic Signaling: A Candidate Mechanism of Electroclinical Uncoupling of Neonatal Seizures , 2009, Neuron.

[47]  K. Deisseroth,et al.  When the electricity (and the lights) go out: transient changes in excitability , 2012, Nature Neuroscience.

[48]  Anatol C. Kreitzer,et al.  Regulation of parkinsonian motor behaviours by optogenetic control of basal ganglia circuitry , 2010, Nature.