The structural patterns of the potentiation and the blockade of inhibitory cys-loop receptors through the transmembrane domain

Cover Page

Cite item

Full Text

Abstract

Anion-conducting cys-loop receptors activated by γ-aminobutyric acid (GABAАRs) and glycine (GlyRs) have inhibitory activity in the brain and spinal cord. GABAАRs and GlyRs are targets for various substances that potentiate or inhibit the receptor functions. Many of these substances are clinically significant agents to treat neurological and psychiatric conditions.

The review covers both our results and literature data on electrophysiology, mutations, and biochemistry of non-competitive antagonists, general anesthetics, barbiturates, and fenamates modulating GABAАRs and GlyRs. We focused on our own molecular modeling to determine the sites and the characteristics of binding of these substances to the GABAАR and GlyR transmembrane domain. With the structural patterns of the binding, we have identified possible molecular mechanisms of action for these substances.

About the authors

Alexey V. Rossokhin

Research Center of Neurolgy

Author for correspondence.
Email: alrossokhin@yandex.ru
ORCID iD: 0000-0001-7024-7461

Cand. Sci. (Phys.-Math.), leading researcher, Laboratory of functional synaptology, Brain Institute

Russian Federation, 105064, Moscow, Obukha per., 5

References

  1. Webb T.I., Lynch J.W. Molecular pharmacology of the glycine receptor chloride channel. Curr. Pharm. Des. 2007; 13(23): 2350–2367. doi: 10.2174/138161207781368693
  2. Braat S., Kooy R.F. The GABAA receptor as a therapeutic target for neurodevelopmental disorders. Neuron. 2015; 86(5): 1119–1130. doi: 10.1016/j.neuron.2015.03.042
  3. Sieghart W. Allosteric modulation of GABAA receptors via multiple drug–binding sites. Adv. Pharmacol. 2015; 72: 53–96. doi: 10.1016/bs.apha.2014.10.002
  4. Olsen R.W. GABAA receptor: positive and negative allosteric modulators. Neuropharmacology. 2018; 136(Pt A): 10–22. doi: 10.1016/j.neuropharm.2018.01.036
  5. Hille B. Ionic channels of excitable membrane. 3 ed. Sunderland; 2001.
  6. Jones A.K., Sattelle D.B. The cys-loop ligand-gated ion channel gene superfamily of the red flour beetle, Tribolium castaneum. BMC Genomics. 2007; 8: 327. doi: 10.1186/1471-2164-8-327
  7. Corringer P.J., Baaden M., Bocquet N. et al. Atomic structure and dynamics of pentameric ligand–gated ion channels: new insight from bacterial homologues. J. Physiol. 2010; 588(Pt 4): 565–572. doi: 10.1113/jphysiol.2009.183160
  8. Jaiteh M., Taly A., Henin J. Evolution of pentameric ligand-gated ion channels: pro-loop receptors. PLoS One. 2016; 11(3): e0151934. doi: 10.1371/journal.pone.0151934
  9. Hevers W., Luddens H. The diversity of GABAA receptors. Pharmacological and electrophysiological properties of GABAA channel subtypes. Mol. Neurobiol. 1998; 18(1): 35–86. doi: 10.1007/BF02741459
  10. Sieghart W. Structure, pharmacology, and function of GABAA receptor subtypes. Adv. Pharmacol. 2006; 54: 231–263. doi: 10.1016/s1054–3589(06)54010-4
  11. Aroeira R.I., Ribeiro J.A., Sebastiao A.M. et al. Age-related changes of glycine receptor at the rat hippocampus: from the embryo to the adult. J. Neurochem. 2011; 118(3): 339–353. doi: 10.1111/j.1471–4159.2011.07197.x
  12. Dutertre S., Becker C.M., Betz H. Inhibitory glycine receptors: an update. J. Biol. Chem. 2012; 287(48): 40216–40223. doi: 10.1074/jbc.R112.408229
  13. Jonsson S., Morud J., Pickering C. et al. Changes in glycine receptor subunit expression in forebrain regions of the Wistar rat over development. Brain Res. 2012; 1446: 12–21. doi: 10.1016/j.brainres.2012.01.050
  14. Yu H., Bai X.C., Wang W. Characterization of the subunit composition and structure of adult human glycine receptors. Neuron. 2021; 109(17): 2707–2716 e2706. doi: 10.1016/j.neuron.2021.08.019.
  15. Miller P.S., Smart T.G. Binding, activation and modulation of Cys-loop receptors. Trends Pharmacol. Sci. 2010; 31(4): 161–174. doi: 10.1016/j.tips.2009.12.005
  16. Sauguet L., Shahsavar A., Poitevin F. et al. Crystal structures of a pentameric ligand–gated ion channel provide a mechanism for activation. Proc. Natl. Acad. Sci. USA. 2014; 111(3): 966–971. doi: 10.1073/pnas.1314997111
  17. Masiulis S., Desai R., Uchanski T. et al. GABAA receptor signalling mechanisms revealed by structural pharmacology. Nature. 2019; 565(7740): 454–459. doi: 10.1038/s41586-018-0832-5
  18. Breitinger U., Breitinger H.G. Modulators of the inhibitory glycine receptor. ACS Chem. Neurosci. 2020. 11(12): 1706–1725. doi: 10.1021/acschemneuro.0c00054
  19. Kim J.J., Hibbs R.E. Direct structural insights into GABAA receptor pharmacology. Trends Biochem. Sci. 2021. 46(6): 502–517. doi: 10.1016/j.tibs.2021.01.011
  20. Miller P.S., Aricescu A.R. Crystal structure of a human GABAA receptor. Nature. 2014; 512(7514): 270–275. doi: 10.1038/nature13293
  21. Du J., Lu W., Wu S. et al. Glycine receptor mechanism elucidated by electron cryo–microscopy. Nature. 2015; 526(7572): 224–229. doi: 10.1038/nature14853
  22. Huang X., Chen H., Michelsen K. et al. Crystal structure of human glycine receptor–alpha3 bound to antagonist strychnine. Nature. 2015; 526(7572): 277–280. doi: 10.1038/nature14972
  23. Laverty D., Desai R., Uchanski T. et al. Cryo–EM structure of the human alpha1beta3gamma2 GABAA receptor in a lipid bilayer. Nature. 2019; 565(7740): 516–520. doi: 10.1038/s41586-018-0833-4
  24. Kim J. J., Gharpure A., Teng J. et al. Shared structural mechanisms of general anaesthetics and benzodiazepines. Nature. 2020; 585(7824): 303–308. doi: 10.1038/s41586-020-2654-5
  25. Rossokhin A.V., Zhorov B.S. Side chain flexibility and the pore dimensions in the GABAA receptor. J. Comput. Aided Mol. Des. 2016; 30(7): 559–567. doi: 10.1007/s10822-016-9929-9
  26. Rossokhin A.V. Homology modeling of the transmembrane domain of the GABAA receptor. Biophysics. 2017; 62(5): 708–716. doi: 10.1134/s0006350917050190
  27. Miller C. Genetic manipulation of ion channels: a new approach to structure and mechanism. Neuron. 1989; 2(3): 1195–1205. doi: 10.1016/0896-6273(89)90304–8
  28. Gielen M., Corringer P.J. The dual-gate model for pentameric ligand–gated ion channels activation and desensitization. J. Physiol. 2018; 596(10): 1873–1902. doi: 10.1113/JP275100
  29. Chovancova E., Pavelka A., Benes P. et al. CAVER 3.0: a tool for the analysis of transport pathways in dynamic protein structures. PLoS Comput. Biol. 2012; 8(10): e1002708. doi: 10.1371/journal.pcbi.1002708
  30. Hilf R. J., Dutzler R. X-ray structure of a prokaryotic pentameric ligand–gated ion channel. Nature. 2008; 452(7185): 375–379. doi: 10.1038/nature06717
  31. Bocquet N., Nury H., Baaden M. et al. X-ray structure of a pentameric ligand–gated ion channel in an apparently open conformation. Nature. 2009; 457(7225): 111–114. doi: 10.1038/nature07462
  32. Hibbs R.E., Gouaux E. Principles of activation and permeation in an anion–selective Cys–loop receptor. Nature. 2011; 474(7349): 54–60. doi: 10.1038/nature10139
  33. Shannon R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallographica Section A. 1976; 32(5): 751–767. doi: 10.1107/s0567739476001551
  34. Lang P.F., Smith B.C. Ionic radii for Group 1 and Group 2 halide, hydride, fluoride, oxide, sulfide, selenide and telluride crystals. Dalton Trans. 2010; 39(33): 7786–7791. doi: 10.1039/c0dt00401d
  35. Conway D.E. Ionic hydration in chemistry and biophysics. New York; 1981.
  36. Michalowski M.A., Kraszewski S., Mozrzymas J.W. Binding site opening by loop C shift and chloride ion–pore interaction in the GABAA receptor model. Phys. Chem. Chem. Phys. 2017; 19(21): 13664–13678. doi: 10.1039/c7cp00582b
  37. Yang Z., Cromer B.A., Harvey R.J. et al. A proposed structural basis for picrotoxinin and picrotin binding in the glycine receptor pore. J. Neurochem. 2007; 103(2): 580–589. doi: 10.1111/j.1471-4159.2007.04850.x
  38. Wang D.S., Mangin J.M., Moonen G. et al. Mechanisms for picrotoxin block of alpha2 homomeric glycine receptors. J. Biol. Chem. 2006; 281(7): 3841–3855. doi: 10.1074/jbc.M511022200
  39. Bali M., Akabas M.H. The location of a closed channel gate in the GABAA receptor channel. J. Gen. Physiol. 2007; 129(2): 145–159. doi: 10.1085/jgp.200609639
  40. Chang Y., Weiss D.S. Site-specific fluorescence reveals distinct structural changes with GABA receptor activation and antagonism. Nat. Neurosci. 2002; 5(11): 1163–1168. doi: 10.1038/nn926
  41. Goutman J.D., Calvo D.J. Studies on the mechanisms of action of picrotoxin, quercetin and pregnanolone at the GABA rho 1 receptor. Br. J. Pharmacol. 2004; 141(4): 717–727. doi: 10.1038/sj.bjp.0705657
  42. Xu M., Covey D.F., Akabas M.H. Interaction of picrotoxin with GABAA receptor channel-lining residues probed in cysteine mutants. Biophys. J. 1995; 69(5): 1858–1867. doi: 10.1016/s0006-3495(95)80056-1
  43. Chen L., Durkin K.A., Casida J.E. Structural model for gamma–aminobutyric acid receptor noncompetitive antagonist binding: widely diverse structures fit the same site. Proc. Natl. Acad. Sci. USA. 2006; 103(13): 5185–5190. doi: 10.1073/pnas.0600370103
  44. Sedelnikova A., Erkkila B.E., Harris H. et al. Stoichiometry of a pore mutation that abolishes picrotoxin–mediated antagonism of the GABAA receptor. J. Physiol. 2006; 577(2): 569–577. doi: 10.1113/jphysiol.2006.120287
  45. Zhorov B.S., Bregestovski P.D. Chloride channels of glycine and GABA receptors with blockers: Monte Carlo minimization and structure-activity relationships. Biophys. J. 2000; 78(4): 1786–1803. doi: 10.1016/S0006–3495(00)76729-4
  46. Erkkila B.E., Sedelnikova A.V., Weiss D.S. Stoichiometric pore mutations of the GABAA reveal a pattern of hydrogen bonding with picrotoxin. Biophys. J. 2008; 94(11): 4299–4306. doi: 10.1529/biophysj.107.118455
  47. Tokutomi N., Agopyan N., Akaike N. Penicillin-induced potentiation of glycine receptor–operated chloride current in rat ventro-medial hypothalamic neurones. Br. J. Pharmacol. 1992; 106(1): 73–78. doi: 10.1111/j.1476-5381.1992.tb14295.x
  48. Kumamoto E., Murata Y. Action of furosemide on GABA- and glycine currents in rat septal cholinergic neurons in culture. Brain Res. 1997; 776: 246–249. doi: 10.1016/s0006-8993(97)01083-4
  49. Kolbaev S.N., Sharonova I.N., Vorobjev V.S. et al. Mechanisms of GABAA receptor blockade by millimolar concentrations of furosemide in isolated rat Purkinje cells. Neuropharmacology. 2002; 42(7): 913–921. doi: 10.1016/s0028-3908(02)00042-4
  50. Rossokhin A.V., Sharonova I.N., Bukanova J.V. et al. Block of GABA receptor ion channel by penicillin: Electrophysiological and modeling insights toward the mechanism. Mol. Cell Neurosci. 2014; 63: 72–82. doi: 10.1016/j.mcn.2014.10.001
  51. Curtis D.R., Game C.J., Johnston G.A. et al. Convulsive action of penicillin. Brain Res. 1972; 43(1): 242–245. doi: 10.1016/0006-8993(72)90288-0
  52. Chow K.M., Hui A.C., Szeto C.C. Neurotoxicity induced by beta–lactam antibiotics: from bench to bedside. Eur. J. Clin. Microbiol. Infect. Dis. 2005; 24(10): 649–653. doi: 10.1007/s10096-005-0021-y
  53. Neher E., Steinbach J.H. Local anaesthetics transiently block currents through single acetylcholine-receptor channels. J. Physiol. 1978; 277: 153–176. doi: 10.1113/jphysiol.1978.sp012267
  54. Vorobjev V.S., Sharonova I.N. Tetrahydroaminoacridine blocks and prolongs NMDA receptor–mediated responses in a voltage-dependent manner. Eur. J. Pharmacol. 1994; 253(1–2): 1–8. doi: 10.1016/0014-2999(94)90750-1
  55. Li Z., Scheraga H.A. Monte Carlo-minimization approach to the multiple-minima problem in protein folding. Proc. Natl. Acad. Sci. USA. 1987; 84(19): 6611–6615. doi: 10.1073/pnas.84.19.6611
  56. Rossokhin A., Teodorescu G., Grissmer S. et al. Interaction of d–tubocurarine with potassium channels: molecular modeling and ligand binding. Mol. Pharmacol. 2006; 69(4): 1356–1365. doi: 10.1124/mol.105.017970
  57. Rossokhin A., Dreker T., Grissmer S. et al. Why does the inner–helix mutation A413C double the stoichiometry of Kv1.3 channel block by emopamil but not by verapamil? Mol. Pharmacol. 2011; 79(4): 681–691. doi: 10.1124/mol.110.068031
  58. Franks N.P. Molecular targets underlying general anaesthesia. Br. J. Pharmacol. 2006; 147 Suppl 1: S72–S81. doi: 10.1038/sj.bjp.0706441
  59. Loscher W., Rogawski M.A. How theories evolved concerning the mechanism of action of barbiturates. Epilepsia. 2012; 53 Suppl 8: 12–25. doi: 10.1111/epi.12025
  60. Pistis M., Belelli D., Peters J.A. et al. The interaction of general anaesthetics with recombinant GABAA and glycine receptors expressed in Xenopus laevis oocytes: a comparative study. Br. J. Pharmacol. 1997; 122(8): 1707–1719. doi: 10.1038/sj.bjp.0701563
  61. Germann A.L., Shin D.J., Manion B.D. et al. Activation and modulation of recombinant glycine and GABAA receptors by 4-halogenated analogues of propofol. Br. J. Pharmacol. 2016; 173(21): 3110–3120. doi: 10.1111/bph.13566
  62. Frolich M., Lachinsky N., Moolenaar A.J. The influence of combined cyproterone acetate–ethinyl oestradiol therapy on serum levels of dehydroepiandrosterone, androstenedione, and testosterone in hirsute women. Acta Endocrinol. (Copenh.). 1977; 84(2): 333–342. doi: 10.1530/acta.0.0840333
  63. Woodward R.M., Polenzani L., Miledi R. Effects of fenamates and other nonsteroidal anti–inflammatory drugs on rat brain GABAA receptors expressed in Xenopus oocytes. J. Pharmacol. Exp. Ther. 1994; 268(2): 806–817.
  64. Zhang Z.X., Lü H., Dong X.P. et al. Kinetics of etomidate actions on GABAA receptors in the rat spinal dorsal horn neurons. Brain Res. 2002; 953(1–2): 93–100. doi: 10.1016/s0006–8993(02)03274–2
  65. Halliwell R.F., Thomas P., Patten D. et al. Subunit–selective modulation of GABAA receptors by the non–steroidal anti–inflammatory agent, mefenamic acid. Eur. J. Neurosci. 1999; 11(8): 2897–2905. doi: 10.1046/j.1460-9568.1999.00709.x
  66. Sharonova I.N., Dvorzhak A.Y. Blockade of GABAA receptor channels by niflumic acid prevents agonist dissociation. Biochemistry (Moscow) Suppl. Ser. A: Membrane and Cell Biol. 2013; 7(1): 37–44. doi: 10.1134/s1990747812050169.
  67. Rossokhin A.V., Sharonova I.N., Dvorzhak A. et al. The mechanisms of potentiation and inhibition of GABAA receptors by non-steroidal anti-inflammatory drugs, mefenamic and niflumic acids. Neuropharmacology. 2019; 160: 107795. doi: 10.1016/j.neuropharm.2019.107795
  68. Coyne L., Su J., Patten D. et al. Characterization of the interaction between fenamates and hippocampal neuron GABAA receptors. Neurochem. Int. 2007; 51(6–7): 440–446. doi: 10.1016/j.neuint.2007.04.017
  69. Maleeva G., Peiretti F., Zhorov B.S. et al. Voltage-dependent inhibition of glycine receptor channels by niflumic acid. Front. Mol. Neurosci. 2017; 10: 125. doi: 10.3389/fnmol.2017.00125
  70. Eaton M.M., Germann A.L., Arora R. et al. Multiple non-equivalent interfaces mediate direct activation of GABAA receptors by propofol. Curr. Neuropharmacol. 2016; 14(7): 772–780. doi: 10.2174/1570159x14666160202121319
  71. Stewart D.S., Hotta M., Li G.D. et al. Cysteine substitutions define etomidate binding and gating linkages in the alpha-M1 domain of gamma-aminobutyric acid type A (GABAA) receptors. J. Biol. Chem. 2013; 288(42): 30373–30386. doi: 10.1074/jbc.M113.494583
  72. Orser B.A., Wang L.Y., Pennefather P.S. et al. Propofol modulates activation and desensitization of GABAA receptors in cultured murine hippocampal neurons. J. Neurosci. 1994; 14(12): 7747–7760.
  73. Lu H., Xu T.L. The general anesthetic pentobarbital slows desensitization and deactivation of the glycine receptor in the rat spinal dorsal horn neurons. J. Biol. Chem. 2002; 277(44): 41369–41378. doi: 10.1074/jbc.M206768200
  74. Mathers D.A., Wan X., Puil E. Barbiturate activation and modulation of GABAA receptors in neocortex. Neuropharmacology. 2007; 52(4): 1160–1168. doi: 10.1016/j.neuropharm.2006.12.004
  75. Wakita M., Kotani N., Akaike N. Effects of propofol on glycinergic neurotransmission in a single spinal nerve synapse preparation. Brain Res. 2016; 1631: 147–156. doi: 10.1016/j.brainres.2015.11.030
  76. Parker I., Gundersen C.B., Miledi R. Actions of pentobarbital on rat brain receptors expressed in Xenopus oocytes. J. Neurosci. 1986; 6(8): 2290–2297. doi: 10.1523/JNEUROSCI.06-08-02290.1986
  77. Yang J., Uchida I. Mechanisms of etomidate potentiation of GABAA receptor-gated currents in cultured postnatal hippocampal neurons. Neuroscience. 1996; 73(1): 69–78. doi: 10.1016/0306-4522(96)00018-8
  78. Kitamura A., Sato R., Marszalec W. et al. Halothane and propofol modulation of gamma–aminobutyric acidA receptor single–channel currents. Anesth. Analg. 2004; 99(2): 409–415. doi: 10.1213/01.ANE.0000131969.46439.71
  79. Dvorzhak A.Y. Effects of fenamate on inhibitory postsynaptic currents in Purkinje’s cells. Bull. Exp. Biol. Med. 2008; 145(5): 564–568. doi: 10.1007/s10517-008-0144-0
  80. Belelli D., Lambert J.J., Peters J.A. et al. The interaction of the general anesthetic etomidate with the gamma-aminobutyric acid type A receptor is influenced by a single amino acid. Proc. Natl. Acad. Sci. USA. 1997; 94: 11031–11036. doi: 10.1073/pnas.94.20.11031
  81. Smith A.J., Oxley B., Malpas S. et al. Compounds exhibiting selective efficacy for different beta subunits of human recombinant gamma-aminobutyric acid A receptors. J. Pharmacol. Exp. Ther. 2004; 311(2): 601–609. doi: 10.1124/jpet.104.070342
  82. Li G.D., Chiara D.C., Sawyer G.W. et al. Identification of a GABAA receptor anesthetic binding site at subunit interfaces by photolabeling with an etomidate analog. J. Neurosci. 2006; 26(45): 11599–11605. doi: 10.1523/JNEUROSCI.3467-06.2006
  83. Krasowskia M.D., Nishikawac K., Nikolaevaa N. et al. Methionine 286 in transmembrane domain 3 of the GABAA receptor β subunit controls a binding cavity for propofol and other alkylphenol general anesthetics. Neuropharmacology. 2001; 41(8): 952–964. doi: 10.1016/s0028-3908(01)00141-1
  84. Bali M., Akabas M.H. Defining the propofol binding site location on the GABAA receptor. Mol. Pharmacol. 2004; 65(1): 68–76. doi: 10.1124/mol.65.1.68
  85. Laurie D.J., Seeburg P.H., Wisden W. The distribution of 13 GABAA receptor subunit mRNAs in the rat brain. II. Olfactory bulb and cerebellum. J. Neurosci. 1992; 12(3): 1063–1076. doi: 10.1523/JNEUROSCI.12-03-01063.1992
  86. Rossokhin A. The general anesthetic etomidate and fenamate mefenamic acid oppositely affect GABAA and GlyR: a structural explanation. Eur. Biophys. J. 2020; 49(7): 591–607. doi: 10.1007/s00249-020-01464-7
  87. Rossokhin A.V., Sharonova I.N. Structural pharmacology of GABAA receptors. Ann. Clin. Exp. Neurol. 2021; 15(4): 44–53. doi: 10.54101/acen.2021.4.5
  88. Bode A., Lynch J.W. Analysis of hyperekplexia mutations identifies transmembrane domain rearrangements that mediate glycine receptor activation. J. Biol. Chem. 2013; 288(47): 33760–33771. doi: 10.1074/jbc.M113.513804
  89. Scott S., Lynch J.W., Keramidas A. Correlating structural and energetic changes in glycine receptor activation. J. Biol. Chem. 2015; 290(9): 5621–5634. doi: 10.1074/jbc.M114.616573
  90. Bode A., Lynch J.W. The impact of human hyperekplexia mutations on glycine receptor structure and function. Mol. Brain. 2014; 7: 2. doi: 10.1186/1756-6606-7-2
  91. Durisic N., Godin A.G., Wever C.M. et al. Stoichiometry of the human glycine receptor revealed by direct subunit counting. J. Neurosci. 2012; 32(37): 12915–12920. doi: 10.1523/JNEUROSCI.2050-12.2012
  92. Low S.E., Ito D., Hirata H. Characterization of the zebrafish glycine receptor family reveals insights into glycine receptor structure function and stoichiometry. Front. Mol. Neurosci. 2018; 11: 286. doi: 10.3389/fnmol.2018.00286

Supplementary files

Supplementary Files
Action
1. JATS XML
2. Fig. 1. Ligand-activated pentamer receptor architecture. А — the view of the receptor in the membrane plane. The frontal subunit is omitted to demonstrate a pore. The arrow indicates the receptor central axis. The transmembrane domain (TMD) and the extracellular domain (ECD) are pointed out; B — the extracellular view of TMD. Transmembrane chains М1–М4 are indicated on the colored subunit. Intersubunit interfaces are pointed with + and –; C — the alignment of GABAАR and GlyR segment M1–M3 aminoacid sequences. The GABAАR_α1 P14867, GABAАR_β1 P18505, GABAАR_β2 P47870, GABAАR_γ2 P18507, GlyR_α1 P23415, GlyR_α2 P23416, and GlyR_β P48167 sequences were sourced from the UniProt Database and aligned relative to the highly conserved Arg 0' residues. Intersubunit potentiating site residues are marked with numbers and strokes above the sentences.

Download (408KB)
3. Fig. 2. GABAАR pore. А — visualized space inside the pore of the α1β2γ2 GABAАR models that are homologous to the closed α3 GlyR (5CFB), open α1 GlyR (3JAE), and desensitized β3 GABAАR (4COF) structures. Only М2 segments are depicted. The frontal subunit was omitted for clarity. Horizontal lines indicate residue levels -2′, 9′, and 20′ for main pore constrictions. Some subunits are highlighted in cyan (α1), orange (β2), and purple (γ2); В — dependence of the pore diameter on the pore depth in the closed, open, and desensitized receptor states. Vertical lines indicate pore levels -2′, 9′, and 20′.

Download (992KB)
4. Fig. 3. GABAАR non-competitive antagonists. А — cryo-EM structure of GABAАR (6X40) complexed with picrotoxin; B, C — binding of penicillin and furosemide in the pore of the GABAАR model that is homologous to the bacterial receptor GLIC (3HEZ) structure. A–C present views from the extracellular space (top) and in the plane of membrane (bottom). The frontal subunit is omitted for clarity on the side-view images. The images show side chains of residues 2′, 6′, 9′, and 13′ that most significantly contribute to the blocker-receptor interaction. Hydrogen bonds are depicted as red dashed lines. Color subunit presentation complies with Fig. 2.

Download (274KB)
5. Fig. 4. Fenamate GABAАR potentiation and inhibition. The potentiating sites of binding meclofenamic acid (MFA; А) and niflumic acid (NFA; В) in β(+)/α(−) transmembrane interface of the α1β2γ2 GABAАR model that is homologous to the 3JAE α1 GlyR structure. C — superposition of the α1β1γ2 и α1β2γ2 GABAАR models in the β(+)/α(−) interface region. R19′ side chain in β2 subunit is green; D — NFA energy profile in the pore of the α1β2γ2 GABAАR model; E — structural models of NFA binding in the pore at the upper and lower sites. А, B, and E present residue side chains that significantly contribute (>1 kcal/mol) to the energy of ligand-receptor interaction. Hydrogen bonds are depicted as red dashed lines. Color presentation of subunits complies with Fig. 2.

Download (306KB)
6. Fig. 5. Structural determinants of GlyR potentiation. А, B — TMD of α1 GlyR in open (3JAE) and closed (3JAD) states. The images present residue side chains R19′ and Q(–26′). 3JAE and 3JAD structures are shown as they are laid out in Protein Data Bank without preliminary energy minimization. С–Е — view of the transmembrane β(+)/α(−) (С), α(+)/β(−) (D), and β(+)/β(−) (E) interfaces of the α1β GlyR model in the membrane plane that are homologous to the 3JAE structure. The figure presents the parts of the M1–M3 helices and the residue side chains that are homologous to those involved in MFA binding on the β(+)/α(−) interface of α1β2γ2 GABAАR. Hydrogen bonds are depicted as red dashed lines. Some subunits are highlighted in cyan (α1) and orange (β).

Download (321KB)

Copyright (c) 2022 Rossokhin A.V.

Creative Commons License
This work is licensed under a Creative Commons Attribution 4.0 International License.

This website uses cookies

You consent to our cookies if you continue to use our website.

About Cookies