Hostname: page-component-745bb68f8f-d8cs5 Total loading time: 0 Render date: 2025-02-09T08:01:35.723Z Has data issue: false hasContentIssue false

Molecular mechanisms of the GABA type A receptor function

Published online by Cambridge University Press:  14 January 2025

Michał A. Michałowski*
Affiliation:
Faculty of Medicine, Department of Biophysics and Neuroscience, Wroclaw Medical University, Wrocław, Poland
Karol Kłopotowski
Affiliation:
Faculty of Medicine, Department of Biophysics and Neuroscience, Wroclaw Medical University, Wrocław, Poland
Grzegorz Wiera
Affiliation:
Faculty of Medicine, Department of Biophysics and Neuroscience, Wroclaw Medical University, Wrocław, Poland
Marta M. Czyżewska
Affiliation:
Faculty of Medicine, Department of Biophysics and Neuroscience, Wroclaw Medical University, Wrocław, Poland
Jerzy W. Mozrzymas
Affiliation:
Faculty of Medicine, Department of Biophysics and Neuroscience, Wroclaw Medical University, Wrocław, Poland
*
Corresponding author: Michał A. Michałowski; Email: michal.michalowski@umw.edu.pl
Rights & Permissions [Opens in a new window]

Abstract

The GABA type A receptor (GABAAR) belongs to the family of pentameric ligand-gated ion channels and plays a key role in inhibition in adult mammalian brains. Dysfunction of this macromolecule may lead to epilepsy, anxiety disorders, autism, depression, and schizophrenia. GABAAR is also a target for multiple physiologically and clinically relevant modulators, such as benzodiazepines (BDZs), general anesthetics, and neurosteroids. The first GABAAR structure appeared in 2014, but the past years have brought a particularly abundant surge in structural data for these receptors with various ligands and modulators. Although the open conformation remains elusive, this novel information has pushed the structure–function studies to an unprecedented level. Electrophysiology, mutagenesis, photolabeling, and in silico simulations, guided by novel structural information, shed new light on the molecular mechanisms of receptor functioning. The main goal of this review is to present the current knowledge of GABAAR functional and structural properties. The review begins with an outline of the functional and structural studies of GABAAR, accompanied by some methodological considerations, especially biophysical methods, enabling the reader to follow how major breakthroughs in characterizing GABAAR features have been achieved. The main section provides a comprehensive analysis of the functional significance of specific structural elements in GABAARs. We additionally summarize the current knowledge on the binding sites for major GABAAR modulators, referring to the molecular underpinnings of their action. The final chapter of the review moves beyond examining GABAAR as an isolated macromolecule and describes the interactions of the receptor with other proteins in a broader context of inhibitory plasticity. In the final section, we propose a general conclusion that agonist binding to the orthosteric binding sites appears to rely on local interactions, whereas conformational transitions of bound macromolecule (gating) and allosteric modulation seem to reflect more global phenomena involving vast portions of the macromolecule.

Type
Review
Copyright
© The Author(s), 2025. Published by Cambridge University Press

GABA type A receptor

The GABA type A receptor (GABAAR) is an ionotropic receptor that plays a key role in inhibitory synaptic transmission in adult mammalian brains. Its extracellular domain (ECD, Figure 1A) contains ligand-binding sites, whereas the ion pore, together with the so-called activation channel gate, is located in the transmembrane domain (TMD, Figure 1A). Upon binding of the neurotransmitter (gamma-aminobutyric acid, GABA, Figure 2), the process of receptor activation is initiated, leading to distinct conformational transitions, including preactivation, opening/closing, and desensitization. In general, transitions between fully bound states (when agonist binding is completed) are defined as gating. GABAAR is permeant for anions, which means that under physiological conditions, chloride influx is predominant, while bicarbonate (HCO3), for which the receptor has lower permeability, is transported outwardly (Kaila et al. Reference Kaila, Voipio, Paalasmaa, Pasternack and Deisz1993). Overall, in adulthood, ion flow through GABAARs results in a change in transmembrane potential toward negative values, which underlies the inhibitory action mediated by these receptors. However, it is worth mentioning that at the early stages of development, GABAARs may mediate a depolarizing action because of the lower expression of the KCC2 exchanger, resulting in a higher intracellular chloride concentration [Cl]in and the Nernst potential for the chloride ions (VCl) depolarizing with respect to the resting membrane voltage (Rivera et al. Reference Rivera, Voipio and Kaila2005). In addition, the function of the receptor may be modulated by multiple endogenous and exogenous compounds of high physiological and pharmacological importance, including BDZs, barbiturates, general anesthetics, and neurosteroids, via a wide range of binding sites (Figures 1B and C and 2). Many of these drugs are used to treat several diseases in which dysfunction of GABAergic inhibition has been reported, for example, epilepsy, depression, anxiety disorders, autism, and schizophrenia (Brickley and Mody Reference Brickley and Mody2012; Gafford et al. Reference Gafford, Guo, Flandreau, Hazra, Rainnie and Ressler2012; Rudolph and Möhler Reference Rudolph and Möhler2006).

Figure 1. General structure of GABAAR. (A) The α1β2γ2 receptor structure viewed from the plane of the membrane. ECD and TMD are marked: α1 subunits in blue, β2 in red, and γ2 in yellow. (B) and (C): The same structure viewed from the extracellular space. Within the ECD, GABA and BDZ binding sites are located at β+ and α+ subunit interfaces, respectively. In TMD, a wide spectrum of modulatory (MOD) binding sites is located in most subunit interfaces, and a PTX-binding site is found in the ion pore. (D) Key β-strands of the ECD (black circles) and loops (red and blue circles, for principal and complementary subunits, respectively). (E) α-helices forming bundles in the TMD. Selected residues lining the ion pore are marked using x′ notation (universal for all subunits).

Figure 2. Ligands of GABAAR. Gamma aminobutyric acid (GABA) is a natural agonist that binds to orthosteric binding sites (β/α subunit interface in the ECD). Partial agonist P4S and competitive antagonists bicuculline and gabazine bind to the same site. Diazepam and Zolpidem are positive allosteric modulators binding to the allosteric binding sites (α/γ subunit interface in the ECD). Diazepam belongs to the family of BDZs together with, for example, flurazepam, lorazepam, flunitrazepam, and alprazolam. These drugs share a common benzene and diazepine ring system and differ by substitutions (at R1, R2, R3, R7, and R2’). Zolpidem belongs to the nonbenzodiazepine family, which shares a similar mode of action to BDZs, but significantly differ in structure (being an imidazopyridine). DMCM which belongs to the family of β-carbolines binds to the same site, but acts as negative modulator. General anesthetics are another large group of positive modulators, including propofol, etomidate, and barbiturates. Phenobarbital is an example of the barbiturate; other members of the family differ by R1 and R2 substitutions. Those modulators’ sites are located at subunit interfaces in the upper TMD. Neurosteroids may act as positive or negative allosteric modulators. Pregnanolone (3α5β) and its isomer allopregnanolone (3α5α) are positive modulators, whereas another isomer epipregnanolone (3β5β) is a negative modulator. The addition of sulfate ester (at C3β) to pregnanolone turns it into a negative allosteric modulator. Neurosteroids binds to multiple sites within the TMD.

GABAAR is a member of the pentameric ligand-gated ion channels (pLGICs, Figure 3A) family together with glycine receptor (GlyR), nicotinic acetylcholine receptor (nAChR), serotonin type 3 receptor (5-HT3), Erwinia ligand-gated ion channel (ELIC), and Gloeobacter ligand-gated ion channel (GLIC). There are 19 types of GABAAR subunits (Ernst and Sieghart Reference Ernst and Sieghart2015; Figure 3B): six α subunits (α1, α2, α3, α4, α5, and α6), three β subunits (β1, β2, and β3), three γ (γ1, γ2, and γ3), three ρ (ρ1, ρ2, and ρ3), and δ, ε, π, θ, subunits. Each subunit is composed of three domains: extracellular, transmembrane, and intracellular (ICD, structure not known). Functional GABAAR is an assembly of five subunits, with the most common stoichiometry being α12122 (schematic view in Figure 1B and C). The orthosteric binding sites (GABA, Figures 1B and 2) are located at the β+ interfaces (+/− notation corresponding to the principal/complementary subunit roles; the principal role is assigned to the subunit located at the left side if the receptor is viewed from the top), roughly in the middle of the ECD. A structurally similar binding site for BDZs is present at the α+ interface (Figures 1B and 2) and the role of BDZs is not to activate but rather to allosterically modulate the GABAAR. Within the subunits in the ECD, several specific structures can be distinguished; multiple β-strands (arranged predominantly in an antiparallel fashion, favoring the formation of β-sheets), which are numbered from 1 (near the extracellular N-terminal) to 10 (near the linker with TMD; Figures 1D and 3C). The fragments of these β-strands or other structures that form the orthosteric binding sites are called loops A-G (loop naming for historical reasons; Figures 1D and 3C). Loops A, B, and C are located on the principal side. Loops A and B line the deep part of the binding cavity, whereas loop C (comprising fragments of β-strands 9 and 10 and the connecting loop) closes the outer part of the cavity. At the complementary side (starting from the deep part of the binding site), loops E, D, G, and F are located, which are fragments of β-strands 6, 2, 1, and 8, respectively (Figures 1D and 3C). Another set of important ECD structures is located at the ECD/TMD interface. The β10-M1 linker is a covalent connection between the last β-strand of the ECD and the first α-helix of the TMD (Figures 1D and E and 3C). In addition, the Cys-loop (between strands 6 and 7; Figure 1D and 3C) and loop 2 (between strands 1 and 2; Figures 1D and 3C) points toward the TMD, forming a set of non-bonded interactions with the TMD (Figures 1D and 3C). In the TMD of each subunit, four membrane-spanning α-helices named M1, M2, M3, and M4 are present (Figures 1E and 3C). M1, as already mentioned, is covalently bound with the ECD via β10-M1 linker, M2 forms the ion pore and is connected with M3 via M2-M3 loop being an important part of the ECD/TMD interface, and the last helix M4 ends with the receptor’s C-terminal. M1, M2, M3, and M4 helices of each subunit form a bundle. In the pore-lining M2 helix, which is crucial for receptor function with ion permeation in particular, the residues are often numbered by x′ notation starting from the bottom (intracellular side), for example, β2A248 is -2′ residue and β2E270 (M2 top, extracellular side) is 20′ (Figures 1E and 3C). The channel gate is made of highly conserved 9′ leucine residues. At the bottom of the ion pore, another constriction is present at the -2′ residue level, and its closure is associated with receptor desensitization. As in ECD, several binding sites for various ligands are present in TMD (Figure 1C). The most important are located at the top and bottom parts of the domain at subunit interfaces. The top binding sites are located between the M3 helix (principal side) and M1 helix (complementary side) of the neighboring subunits below the M2-M3 loop, and the bottom ones are similarly located (between M3 and M1 helices) but closer to the intracellular compartment. At the TMD domain, the binding sites for various modulators are located: BDZs (in addition to the ECD-binding site), general anesthetics, and neurosteroids (Figure 2). The last domain, ICD, spans from the helices facing the inner part of the cell. The sequence of this domain is highly diversified among receptor subtypes and its three-dimensional structure is not known.

Figure 3. pLGIC and GABAAR isoforms. (A) pLGICs are categorized according to their ion selectivity. Animal pLGICs are also called Cys-loop receptors because of the characteristic disulfide bond between β-strands 6 and 7. The respective receptor isoforms are listed in parentheses. (B): There are 19 subunit types of the GABAA receptor: three β subunits that act as principal (red) subunits and six α subunits that play the role of the complementary (blue) one and three γ subunits together with ε, θ, δ, and π subunits that play a modulatory (yellow) role (when co-expressed with two β and two α subunits). The ρ subunits do not assemble with other subunit types; they form only homomers formerly categorized as GABA-type C receptors. (C) Aligned sequences of the α1, β2, and γ2 GABAAR subunits with marked β-strands (green), α-helices (blues), and key loops. Sequence of the ICD is omitted.

The exact molecular mechanism of receptor function, that is, its structure–function relationship, has been the subject of many investigations in recent years and decades. A wide spectrum of approaches, including electrophysiological, structural, and computational methods, have provided substantial information on receptor functioning; however, the molecular scenarios of specific conformational transitions have not been fully elucidated as yet. The scope of the present review is to present current knowledge on the structure–function relationship for GABAARs, together with an overview of some inhibitory plasticity mechanisms associated with distinct contributions of specific subtypes of GABAARs and alterations in their subcellular localizations. We begin by describing the basic methods that kinetically describe the activity of GABAARs at both macroscopic and single-channel levels. Next, we focus on an overview of the available structures of GABAAR, and the subsequent chapters are dedicated to a more detailed description of various regions of the receptor in relation to their functional role. The review concludes with general remarks on the GABAAR structure–function relationship.

In recent years, several reviews addressing different aspects of GABAARs and other pLGICs have been published. For the interested reader, the following papers are suggested: a variety of reviews on the structural research (Amundarain et al. Reference Amundarain, Ribeiro, Costabel and Giorgetti2019; Howard Reference Howard2021; Kim and Hibbs Reference Kim and Hibbs2021; Puthenkalam et al. Reference Puthenkalam, Hieckel, Simeone, Suwattanasophon, Feldbauer, Ecker and Ernst2016; Scott and Aricescu Reference Scott and Aricescu2019) and numerous reviews focused on more specific topics: modulation by ethanol (Förstera et al. Reference Förstera, Castro, Moraga-Cid and Aguayo2016); BDZs’ binding sites (Sigel and Ernst Reference Sigel and Ernst2018); allosteric modulators (Olsen Reference Olsen2018); ligands’ receptor subtype specificity (Sieghart and Savić Reference Sieghart and Savić2018); comparison of αβγ and αβδ receptors (Feng and Forman Reference Feng and Forman2018); receptor modeling (Germann et al. Reference Germann, Steinbach and Akk2018; Steinbach and Akk Reference Steinbach and Akk2019); the α5 subunit containing receptors (Mohamad and Has Reference Mohamad and Has2019); subtype selectivity of modulators binding at BDZ-binding site (Maramai et al. Reference Maramai, Benchekroun, Ward and Atack2020); electrophysiology of the receptor (Sallard et al. Reference Sallard, Letourneur and Legendre2021) and receptor interaction with BDZs, neurosteroids, and transmembrane proteins (Castellano et al. Reference Castellano, Shepard and Lu2021); and BDZs action (Goldschen-Ohm Reference Goldschen-Ohm2022).

Functional studies of the receptor

The time course of synaptic currents is primarily shaped by the transient of neurotransmitter concentration and postsynaptic receptor gating properties. The neurotransmitter spatial and temporal profile within the synaptic cleft is determined by the amount of neurotransmitter molecules released upon vesicle exocytosis, the mechanism of its release (need not be an instantaneous vesicle collapse but, e.g., a gradual pore formation with eventual fusion), the geometry of the synaptic cleft, and diffusion limitations around the synapse (e.g., presence of astrocyte processes or extracellular matrix structures; Barbour and Häusser Reference Barbour and Häusser1997; Bouron Reference Bouron2001; Bruns and Jahn Reference Bruns and Jahn1995; Clements Reference Clements1996; Harris and Stevens Reference Harris and Stevens1989). Key determinants of the synaptic current time course are also the binding and gating properties of the postsynaptic receptors, as well as the efficiency of the agonist uptake mechanisms. The latter factor typically has a larger impact in the case of spontaneous or evoked synaptic currents being less critical for miniature synaptic events. The agonist transient in a typical GABAergic synapse consists of a predominant rapid phase in the range of hundreds of microseconds and a slower one (tail) decaying within a millisecond or so (Mozrzymas Reference Mozrzymas2004; Mozrzymas et al. Reference Mozrzymas, Barberis, Michalak and Cherubini1999; Mozrzymas et al. Reference Mozrzymas, Wójtowicz, Piast, Lebida, Wyrembek and Mercik2007a; Pytel et al. Reference Pytel, Mercik and Mozrzymas2003; Scimemi and Beato Reference Scimemi and Beato2009). The peak concentration of synaptic agonists has been estimated to lie within the low millimolar range. However, this relatively high peak of synaptic agonist concentration is not saturating because of the brevity of the synaptic agonist transient (Mozrzymas et al. Reference Mozrzymas, Barberis, Michalak and Cherubini1999; Pytel et al. Reference Pytel, Mercik and Mozrzymas2003). Thus, synaptic conditions of postsynaptic receptor activation are characterized by a high degree of non-equilibrium, primarily due to very brief sub-millisecond exposure to the neurotransmitter (Mozrzymas Reference Mozrzymas2004; Pytel et al. Reference Pytel, Mercik and Mozrzymas2003; Scimemi and Beato Reference Scimemi and Beato2009). A similar estimation of agonist transient has also been made for the glycinergic synapse (Beato Reference Beato2008). Importantly, these specific non-equilibrium conditions of synaptic receptors activation need to be considered in pharmacological investigations. For instance, chlorpromazine, which was shown to slow down the agonist binding to the receptor, exerts a particularly strong inhibitory effect on miniature inhibitory postsynaptic currents (mIPSCs) as the time duration of synaptic GABA transient becomes insufficient to activate GABAARs to the same extent as in control conditions (Mozrzymas et al. Reference Mozrzymas, Barberis, Michalak and Cherubini1999).

Analysis of the mIPSC time course provides a clear indication that GABAA receptors are characterized by rapid gating properties. Indeed, the mIPSC 10–90% rise time is typically under 1 ms (Hájos et al. Reference Hájos, Nusser, Rancz, Freund and Mody2000) and the decay time ranges from tens to up to 100 ms for different cell types, often showing more than one kinetic component (Hájos et al. Reference Hájos, Nusser, Rancz, Freund and Mody2000; Jones and Westbrook Reference Jones and Westbrook1995; Mozrzymas et al. Reference Mozrzymas, Wójtowicz, Piast, Lebida, Wyrembek and Mercik2007a). All these observations reveal thus a dynamic nature of GABAARs activation in a typical GABAergic synapse.

The rapid binding and gating properties of GABAARs impose certain requirements on experimental methods for studying GABAAR kinetics with precision adequate to the time scale of synaptic transmission. In particular, to properly estimate receptor kinetics, we need to be able to apply the agonist with sufficient accuracy to mimic the synaptic agonist transient. This condition is reasonably fulfilled by a so-called ultrafast perfusion system based on theta (two channels) glass tubing driven by a piezoelectric translator, as described by Jonas (Reference Jonas, Sakmann and Neher1995), and then used by other investigators, including our group. The effective exchange time achieved with this system ranges between 50 and a few 100 ms for outside-out patches, which is sufficient to reasonably mimic the IPSC kinetics by current responses to rapid agonist applications (Figure 4). Analysis of current responses to rapid GABA applications has been applied for neurons, both cultured and in slices (e.g., Barberis et al. Reference Barberis, Cherubini and Mozrzymas2000; Jones and Westbrook Reference Jones and Westbrook1995; Mozrzymas et al. Reference Mozrzymas, Barberis and Vicini2007b; Mozrzymas et al. Reference Mozrzymas, Barberis, Michalak and Cherubini1999, Reference Mozrzymas, Barberis, Mercik and Zarnowska2003). However, most of these studies have been performed on cell lines (typically human embryonic kidney 293 cells, HEK293) in which recombinant GABAARs are expressed, thus enabling to attribute observed kinetic features to a specific GABAAR subtype (Boileau et al. Reference Boileau, Li, Benkwitz, Czajkowski and Pearce2003; Brodzki et al. Reference Brodzki, Rutkowski, Jatczak, Kisiel, Czyzewska and Mozrzymas2016; Czyzewska and Mozrzymas Reference Czyzewska and Mozrzymas2013; Dixon et al., Reference Dixon, Sah, Lynch and Keramidas2014; Haas and Macdonald Reference Haas and Macdonald1999). Oocytes are also used to express recombinant GABAARs with the advantage of high expression efficiency, but due to the large size of these cells, a two-electrode voltage-clamp is typically used together with perfusion systems offering at best a resolution of hundreds of milliseconds to seconds (see Figure 5 with a drawing illustrating kinetic features of a typical GABAergic current measured from oocytes). In studies on recombinant GABAARs, plasmids encoding human or rat sequences are typically used. Importantly, the human sequence contains an additional leucine at residue four (N-terminal); thus, the numbering of the sequence is shifted by one relative to the rat sequence. For clarity, the results obtained using rat sequences are marked with (r) next to the residue number; for example, α1F65 and α1F64(r) refer to the same residue.

Figure 4. IPCs. (A) Exemplary trace of IPSC between parvalbumine-positive interneuron and hippocampal CA1 pyramidal cell, measured in the whole-cell configuration. Synaptic transmission was evoked using optogenetic stimulation of presynaptic interneuron (unpublished, recorded by G. Wiera). (B) Exemplary current response to an ultrafast short application (3 ms) of saturating [GABA] (10 mM) recorded from an outside-out patch excised from a HEK293 cell expressing GABAARs. Note similar kinetics to the neuronal IPSC (unpublished, recorded by M.M. Czyżewska).

Figure 5. Macroscopic recordings. (A) Exemplary current response to an ultra-fast long application (500 ms) of saturating [GABA] (10 mM) recorded from an outside-out patch excised from a HEK293 cell expressing GABAARs (unpublished, recorded by K. Kłopotowski). (B) Exemplary drawing of a current response (saturating [GABA], long application) with kinetic features resembling typical recordings from oocytes, characterized by different kinetics (and time scale) when compared to the current recorded from a HEK293 cell (A).

The basic analysis of the time course of current responses elicited by rapid applications and measured from outside-out patches (in whole-cell or nucleated patch resolution drops several fold) is shown schematically in Figure 6 (for the most prevalent type of GABAARs, α1β2γ2; McKernan and Whiting Reference McKernan and Whiting1996). Rise time is typically measured as RT 10–90%, and for responses to saturating [GABA] it is in the range of hundreds of microseconds (Figure 6A). Prolonged pulses of saturating [GABA] reveal the so-called macroscopic desensitization (Figure 6B), which can be typically described as a sum of two (or more) exponentials and a constant value (f(t) = Aslow*e−t/τslow + Afast*e−t/τfast + C, where Aslow and Afast are the amplitudes of the desensitizing current components, τslow and τfast are the time constants, and C describes the non-desensitizing current; for normalized currents Aslow + Afast + C = 1). After agonist removal, a deactivation current is observed (Figure 6B). The protocol that most closely mimics synaptic currents is a brief (approximately 1 ms) agonist application (Figure 6C), which typically reveals fast and slow deactivation components. The deactivation time course is markedly different in protocols with short and long GABA applications, as the occupancies of receptors in different states (fully bound open, closed, and desensitized) substantially differ when GABA is washed out in these two protocols (Jones and Westbrook Reference Jones and Westbrook1995; Mozrzymas et al. Reference Mozrzymas, Wójtowicz, Piast, Lebida, Wyrembek and Mercik2007a).

Figure 6. Macroscopic recording analysis. (A) Long agonist pulse (onset phase). RT 10-90 describes the time needed for amplitude increase from 10 to 90% of its peak amplitude. Fast component of macroscopic desensitization can be fitted with a single exponential (in a limited time window), giving the amplitude of rapidly desensitizing current (Afast) and time constant (τfast). (B) Long agonist pulse. Macroscopic desensitization fitted for the entire time window of agonist application yields amplitudes and time constants of the fast and slow components. Deactivation is fitted with a single or double exponential function, depending on the characteristics of the recorded trace. (C) Short agonist pulse. Because of the short duration of the pulse, macroscopic desentization is negligible, but deactivation shows clear two-phase behavior and is fitted with a sum of two exponential functions.

While an ultrafast perfusion system is required to closely mimic synaptic conditions and carry out rigorous kinetic analysis with a resolution relevant to synaptic transmission, other approaches can also be used to study the kinetic and pharmacological features of GABAA receptors. Especially dose–response relationships proved to be particularly useful when studying different agonists or modulators (e.g., Goldschen-Ohm et al. Reference Goldschen-Ohm, Haroldson, Jones and Pearce2014; Kash et al. Reference Kash, Kim, Trudell and Harrison2004a; Pierce et al. Reference Pierce, Germann, Evers, Steinbach and Akk2020). In the simplest case, dose-responses can be fitted with the Hill equation: EC = 1/(1+(EC50/[Agonist])nh), where EC50 is the agonist concentration giving half-maximal amplitude and nh is the Hill coefficient. As already mentioned, the use of oocytes as an expression system offers relatively low temporal resolution for macroscopic responses but still allows the extraction of useful comparative information when studying the kinetic impact of specific mutations or when investigating the pharmacological effects of various compounds, taking advantage of superior expression compared to HEK293 (Akk et al. Reference Akk, Germann, Sugasawa, Pierce, Evers and Steinbach2020).

To ensure the best accuracy of kinetic macroscopic analysis, current responses must be collected for protocols revealing major kinetic features of the receptor, for example, short/long pulses of agonist, double pulses (to assess the recovery from macroscopic desensitization), applications of non-saturating GABA pulses with different durations. The larger the body of experimental evidence, the less ambiguous is the estimation of kinetic parameters. Finally, a kinetic model needs to be chosen, and the values of the respective rate constants are optimized by model fitting to experimental data. Jones and Westbrook’s model proved particularly insightful and revealed that sojourns of the receptors in the desensitized conformation contribute to relatively slow deactivation kinetics (Figure 7A; Jones and Westbrook Reference Jones and Westbrook1995). This model has been widely used by other investigators (Korol et al. Reference Korol, Jin, Jin, Bhandage, Tengholm, Gandasi, Barg, Espes, Carlsson, Laver and Birnir2018; Toyoda et al. Reference Toyoda, Saito, Sato, Tanaka, Ogawa, Yatani, Kawano, Kanematsu, Hirata and Kang2015) including our group (e.g., Brodzki and Mozrzymas Reference Brodzki and Mozrzymas2022; Szczot et al. Reference Szczot, Kisiel, Czyzewska and Mozrzymas2014; Figure 7B), but other models have also been used (Bianchi et al. Reference Bianchi, Botzolakis, Lagrange and Macdonald2009; Feng and Forman Reference Feng and Forman2018; Gielen and Corringer Reference Gielen and Corringer2018; Goldschen-Ohm et al. Reference Goldschen-Ohm, Haroldson, Jones and Pearce2014). In our hands, ChannelLab software (Synaptosoft, USA) proved sufficiently useful and versatile in optimizing even complex schemes (Brodzki and Mozrzymas Reference Brodzki and Mozrzymas2022; Kaczor et al. Reference Kaczor, Michałowski and Mozrzymas2022; Kłopotowski et al. Reference Kłopotowski, Czyżewska and Mozrzymas2021, Reference Kłopotowski, Michałowski, Gos, Mosiądz, Czyżewska and Mozrzymas2023; Szczot et al. Reference Szczot, Kisiel, Czyzewska and Mozrzymas2014; Terejko et al. Reference Terejko, Michałowski, Dominik, Andrzejczak and Mozrzymas2021a; Terejko et al. Reference Terejko, Michałowski, Iżykowska, Dominik, Brzóstowicz and Mozrzymas2021b). Using this above-described methodology we have provided extensive evidence for the presence of the preactivation (primed or flipped) state (Szczot et al. Reference Szczot, Kisiel, Czyzewska and Mozrzymas2014) connecting the fully bound closed state with (bound) open and desensitized conformations of GABAAR (Figure 7B). When focusing on gating features of the receptor under saturating conditions, the transitions originating from a single bound step (A1R) were omitted, as their occupancy under these conditions is expected to be negligible (Brodzki and Mozrzymas Reference Brodzki and Mozrzymas2022; Kłopotowski et al. Reference Kłopotowski, Czyżewska and Mozrzymas2021; Szczot et al. Reference Szczot, Kisiel, Czyzewska and Mozrzymas2014). Successful model fitting should enable close reproduction of the time course of macroscopic currents recorded using all available protocols (see Figure 7D for a long GABA application). Macroscopic data obtained from oocytes can also be modeled using cyclic models (Figure 7C); Pierce et al. Reference Pierce, Germann, Evers, Steinbach and Akk2020), but clearly, as already explained, the reproduction of receptor kinetic features underlying the rapid time course of IPSCs cannot be achieved.

Figure 7. Macroscopic modeling. (A) Kinetic model of the receptor with possible transitions (from the resting state R) to the open (A1O and A2O) and desensitized (A1D and A2D) states from the singly (A1R) and doubly (A2R) agonist-bound states. (B) Kinetic model expanded with an additional flipped/preactive state (A2F). Transitions from the singly bound state are omitted. (C) Cyclic kinetic model allowing for transition pathways between the respective states (symbols same as in A, added O and D states are spontaneously open and desensitized state). (D) Exemplary experimentally recorded trace and receptor response simulated using the kinetic model (unpublished, recorded by M.M. Czyżewska).

Although macroscopic recordings and modeling have been widely used and have proven insightful, it has been recognized that they may be vulnerable to overparameterization, meaning that optimization procedures might converge to different sets of rate constants depending on, for example, initial guesses. As correctly pointed out by Colquhoun and Lape (Reference Colquhoun and Lape2012) modeling based on the single-channel analysis is more ‘resistant’ to artifacts resulting from overparameterization, and in our practice, the optimal strategy was to use both approaches (macroscopic and single-channel (Brodzki et al. Reference Brodzki, Michałowski, Gos and Mozrzymas2020; Brodzki and Mozrzymas Reference Brodzki and Mozrzymas2022; Kaczor et al. Reference Kaczor, Wolska and Mozrzymas2021, Reference Kaczor, Michałowski and Mozrzymas2022; Kłopotowski et al. Reference Kłopotowski, Czyżewska and Mozrzymas2021; Terejko et al. Reference Terejko, Michałowski, Dominik, Andrzejczak and Mozrzymas2021a, 2021b). However, single-channel recordings are typically performed in stationary conditions, in which the vast majority of receptors may be desensitized and the process of receptor accumulation in this conformation cannot be observed, contrary to macroscopic recordings. Indeed, as we pointed out, estimations of the rate constants describing desensitization in macroscopic nonstationary and single-channel steady-state recordings may differ substantially, and for these specific rates, the estimates based on macroscopic recordings are more precise (Brodzki et al. Reference Brodzki, Michałowski, Gos and Mozrzymas2020; Kaczor et al. Reference Kaczor, Wolska and Mozrzymas2021; Kłopotowski et al. Reference Kłopotowski, Czyżewska and Mozrzymas2021). In Figure 8A, a typical single-channel activity recorded from α1β2γ2 GABAA receptors in the presence of stationary and saturating [GABA] is shown. At saturating [GABA], binding reactions are assumed to be very fast, and conformational transitions between bound conformations are rate-limiting. Limitation to transitions between fully bound states substantially simplifies the model scheme (compare Figures 7B and 9B) by removing the free binding parameters in the receptor kinetic model. In particular, at saturation, the occurrence of singly bound states is unlikely, and these transitions and states (A1R, A2F, A1D, A1O) may be omitted. Although most of the single-channel recordings are performed in steady-state conditions in the cell-attached configuration, it is also possible to do that in the dynamic conditions (Mozrzymas et al. Reference Mozrzymas, Barberis and Vicini2007b) for outside-out patches, but typically in excised patches, the noise is considerably larger, and the stability is weaker than in the cell-attached mode. As shown in Figure 8A and B, single-channel activity can be subdivided into clusters and bursts within clusters (Lema and Auerbach Reference Lema and Auerbach2006). It has been observed that GABAAR may show different modes of activity characterized by distinct Popen calculated as a percentage of dwell time in the open state (Brodzki et al. Reference Brodzki, Michałowski, Gos and Mozrzymas2020; Lema and Auerbach Reference Lema and Auerbach2006; Terejko et al. Reference Terejko, Michałowski, Dominik, Andrzejczak and Mozrzymas2021a, 2021b). The single-channel trace is then inspected to distinguish opening/closing transitions either with a threshold (typically 50%) algorithm or by trace curve fitting, which is more laborious but improves the resolution by approximately a few tens of microseconds (e.g., SCAN in DCProgs by David Colquhoun). The detected open and shut single-channel events are then presented as distributions of the open and closed dwell times (Figure 8C) that reveal the number of open and shut states of the analyzed channels, which is indicated by the number of exponential components in the probability density function fits to these distributions. The major advantage of single-channel analysis is that we can obtain insights into the briefest events (both open and closed) with durations in the tens of microseconds, far surpassing the temporal resolution of macroscopic recordings, even when using ultrafast perfusion systems. Our group has used the DCProgs (http://www.onemol.org.uk) software by David Colquhoun, which enabled us to carry out the single channel analysis of relatively low conductance (ca. 27 pS) GABAARs with fairly high resolution (50–100 μs), due to the aforementioned trace curve fitting algorithm (SCAN). This package also contains a program to construct and analyze single-channel event duration distributions (EKDIST) and optimize the rate constants of selected models (Figure 6) by fitting them to the experimental data (HJCFIT) using the maximum likelihood method (Colquhoun et al. Reference Colquhoun and Hawkes1996, Reference Colquhoun, Hatton and Hawkes2003; Colquhoun and Sigworth Reference Colquhoun, Sigworth, Sakmann and Neher1995) and by applying corrections for missed events (Hawkes et al. Reference Hawkes, Jalali and Colquhoun1992). Model fitting with HJCFIT allows the prediction of probability density functions for open and shut events and confronts them with experimental shut and open times distributions (compare Figures 8C and A). In our studies, we preferentially used Jones and Westbrook (Reference Jones and Westbrook1995), which was then upgraded with preactivation (flipping) transitions, as well as with additional open and desensitized transitions (Brodzki et al. Reference Brodzki, Rutkowski, Jatczak, Kisiel, Czyzewska and Mozrzymas2016; Szczot et al. Reference Szczot, Kisiel, Czyzewska and Mozrzymas2014; see Figure 9B). Single-channel recordings are also useful for analyzing the spontaneous activity of GABAARs, which is typically weak for wild type (WT) receptors (except for ε subunit-containing receptors, Wagner et al. Reference Wagner, Goldschen-Ohm, Hales and Jones2005), but specific mutations may strongly increase this form of activity (Jatczak-Śliwa et al. Reference Jatczak-Śliwa, Terejko, Brodzki, Michałowski, Czyzewska, Nowicka, Andrzejczak, Srinivasan and Mozrzymas2018; Kisiel et al. Reference Kisiel, Jatczak, Brodzki and Mozrzymas2018; Sexton et al. Reference Sexton, Penzinger, Mortensen, Bright and Smart2021). Using multiple single-channel recording protocols for different concentrations of the agonist allows the investigation of a complex kinetic model covering binding steps, preactivation, receptor activity in singly and doubly bound states, as well as spontaneous activity (Kisiel et al. Reference Kisiel, Jatczak, Brodzki and Mozrzymas2018, see Figure 9C).

Figure 8. Single-channel recordings. (A) and (B) Exemplary traces recorded from HEK293 cells expressing α1β2γ2 GABAA receptors, measured in the cell-attached configuration. Repetitive receptor openings form bursts that may be grouped into clusters. (C) Exemplary distributions of times in the open and closed states fitted with the sum of exponentials (two for open times and four for shut events, unpublished, recorded by K. Kłopotowski). Solid lines represent approximations of the probability density functions determined as the sums of the respective exponentials (dashed lines). The number of exponentials in a distribution indicates the number of respective states.

Figure 9. Single-channel modeling. (A) Distributions of open and shut times (solid lines) calculated using a kinetic model with optimized rate constants. Dashed lines present distributons determined with correction for missed events. Distributions are superimposed on histograms for the experimental data (unpublished, recorded by K. Kłopotowski). (B) Kinetic scheme often used to model single-channel data obtained for stationary saturating [GABA]. Tranisitions related to agonist binding reactions are omitted, but extra desensitized (A2D′) and open state (A1O′) are added (compared to the model in Figure 7B). These additional states are required to reproduce multiple open and shut states, as indicated by the number of components in the open and shut states distributions. (C) Complete kinetic model (based on (Kisiel et al. Reference Kisiel, Jatczak, Brodzki and Mozrzymas2018) of GABAAR including spontaneous and both singly/doubly agonist bound activity with and without transition through the flipped/preactivated state.

The major advantage of the above-described approach, based on macroscopic and single-channel recordings and their respective modeling, is that it offers insight into receptor gating properties with a resolution relevant to the time scale of synaptic transitions. A thorough identification and characterization of respective conformational transitions allows to ascribe the roles of distinct structural elements (addressed e.g., in mutagenesis studies) to molecular mechanisms underlying these transitions as well as the role of respective structures in pharmacological modulation of the receptor.

Structural studies of the receptor

Structural research is essential in the field of protein structure–function studies, as it provides critical insights into how a protein’s shape and architecture dictate its biological functions and interactions. For GABAA receptors, these efforts began long before any experimental structures of the receptor became available, relying primarily on electrophysiological recordings combined with mutagenesis studies. However, it has only been in recent years that a variety of GABAAR structures have been published – much later compared to other members of the pLGIC family, such as AChRs. In this chapter, we provide a brief overview of GABAAR structures that have been determined so far. A more detailed description will follow in subsequent chapters, where these structures will be analyzed in the context of functional studies, including electrophysiology and other techniques.

The nAChR was the first member of the pLGIC family with an experimentally determined structure in 2005 (Unwin Reference Unwin2005) whereas the Caenorhabditis elegans glutamate-gated chloride channel α (GluCl) was the first anionic channel from this family with structure solved in 2011 (Hibbs and Gouaux Reference Hibbs and Gouaux2011). After the nAChR structure was determined, structural studies expanded to other pLGIC members, resulting in extensivdata, particularly on bacterial channels such as GLIC and ELIC (Bocquet et al. Reference Bocquet, Prado De Carvalho, Cartaud, Neyton, Le Poupon, Taly, Grutter, Changeux and Corringer2007, Reference Bocquet, Nury, Baaden, Le Poupon, Changeux, Delarue and Corringer2009; Hilf and Dutzler Reference Hilf and Dutzler2008, Reference Hilf and Dutzler2009). These advances in related channels facilitated investigations into the GABAAR structure–function relationship through structural homology models. In 2014, the first GABAAR structure was published, depicting a β3 subunit homomeric assembly in a presumed desensitized state with bound benzamidine, determined using X-ray crystallography (Miller and Aricescu Reference Miller and Aricescu2014). Three years later, additional structures were obtained: a chimera of GLIC (ECD) and α1 GABAAR (TMD) (Laverty et al. Reference Laverty, Thomas, Field, Andersen, Gold, Biggin, Gielen and Smart2017) and a construct made of β3 (ECD) and α5 (TMD) GABAAR subunits with bound nerosteroids (Miller et al. Reference Miller, Scott, Masiulis, De Colibus, Pardon, Steyaert and Aricescu2017). However, none of these structures represented GABAAR in its typical subunit stoichiometry. In 2018, two research groups solved the first structures of this receptor in the synaptic assembly type: α1β2γ2 (Zhu et al. Reference Zhu, Noviello, Teng, Walsh, Kim and Hibbs2018) and α1β1γ2 (Phulera et al. Reference Phulera, Zhu, Yu, Claxton, Yoder, Yoshioka and Gouaux2018). Both structures were determined with bound GABA molecules using cryo-electron microscopy. However, the resolution and overall quality of these structures, particularly in the TMD region, limit their applicability. In addition, in 2018, the structure of the α5β3 receptor was published, being thus far the only structure depicting the open state of GABAAR in unusual stoichiometry with a single α5 subunit (Liu et al. Reference Liu, Xu, Guan, Liu, Cui, Zhang, Zheng, Bi, Zhou, Zhang and Ye2018).

One of the most important structures of the receptor was published in 2019 by the Aricescu group: they depicted the α1β3γ2 receptor assembly with a broad spectrum of bound ligands, including GABA, bicuculline (antagonist), picrotoxin (PTX, open channel blocker), diazepam, and alprazolam (modulators) in desensitized and closed states (Laverty et al. Reference Laverty, Desai, Uchański, Masiulis, Stec, Malinauskas, Zivanov, Pardon, Steyaert, Miller and Aricescu2019; Masiulis et al. Reference Masiulis, Desai, Uchański, Serna Martin, Laverty, Karia, Malinauskas, Zivanov, Pardon, Kotecha, Steyaert, Miller and Aricescu2019). These structures, contrary to those previously published, depicted the human receptor in physiological conditions without any substantial alterations (aside from the deletion of the ICD). Similarly, a broad set of structures (α1β2γ2) was presented by the Hibbs group 1 year later: structures with bound GABA, bicuculline, PTX, diazepam, and additionally flumazenil, etomidate, propofol, and phenobarbital (modulators) were published (Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020). These two sets of structures are currently the most used in studies aimed at describing the features of the synaptic types of GABAARs. Another ‘wave’ of receptor structures was published in 2022, including a broad set of extrasynaptic structures depicting diverse subunit assemblies: α1β3ɣ2, α4β3δ, α4β3ɣ2, β3δ, β3ɣ2 (Sente et al. Reference Sente, Desai, Naydenova, Malinauskas, Jounaidi, Miehling, Zhou, Masiulis, Hardwick, Chirgadze, Miller and Aricescu2022), structures of α1β3 assemblies with bound Zn2+ and α-cobratoxin (Kasaragod et al. Reference Kasaragod, Mortensen, Hardwick, Wahid, Dorovykh, Chirgadze, Smart and Miller2022) and additional structures of typical α1β2γ2 with bound antibodies (Noviello et al. Reference Noviello, Kreye, Teng, Prüss and Hibbs2022), and modulators: zolpidem and methyl-6,7-dimethoxyl-4-ethyl-β-carboline-3-carboxylate (DMCM, Figure 2; Zhu et al. Reference Zhu, Sridhar, Teng, Howard, Lindahl and Hibbs2022). All mentioned structures were determined using cryo-electron microscopy in the presumed closed or desensitized state. Interestingly, Sun et al. (Reference Sun, Zhu, Clark and Gouaux2023) presented structures of the receptor with a mixture of various α and β subunit types (e.g., α1 and α2 subunits present in single receptor instead of two α1’s or two α2’s) indicating that subunit assembly rules may be complex. In addition, the authors proposed a mechanism for neurosteroid potentiation. The structures of the engineered α5 homomeric receptors were obtained by Kasaragod et al. (Reference Kasaragod, Malinauskas, Wahid, Lengyel, Knoflach, Hardwick, Jones, Chen, Lucas, El Omari, Chirgadze, Aricescu, Cecere, Hernandez and Miller2023). Additionally, in 2023, the first structures of the ρ-type receptor were presented by Cowgill et al. (Reference Cowgill, Fan, Haloi, Tobiasson, Zhuang, Howard and Lindahl2023) in distinct states: resting, desensitized, and intermediate (with GABA and PTX bound). Later that year, Legesse et al. (Reference Legesse, Fan, Teng, Zhuang, Howard, Noviello, Lindahl and Hibbs2023) determined a set of α1β2γ2 receptors with different neurosteroids (both positive and negative modulators) bound, shedding light on the mechanism of action of these modulators.

Although the progress made in structural studies in the past years has paved the way for more precise functional investigations, there is still a major obstacle in these studies, as we still lack the receptor structure in a typical subunit assembly in the open state. Thus, most analyses aimed at elucidating the receptor activation mechanism are based on comparisons of the receptor structure in the desensitized versus closed state. Whereas the number of determined structures in the desensitized state seems sufficient, unfortunately, no structures have been resolved for the synaptic receptor in the resting, closed apo state, which presents significant limitations for detailed analyses. A brief summary of GABAAR structures published to date is presented in Table 1.

Table 1. Summary of published GABAAR structures

The states of the receptor are assigned according to the classification proposed by authors. Most of the structures depict the receptor in desensitized or closed states categorized according to the pore profile – with constriction on the level of 9′ residue for closed and/or -2′ for desensitized state. Some of the structures were described by the authors as intermediate/preactive/flipped/partially open; for clarity, they are specified in the table as preactive. These structures represent some features characteristic for various conformations, making unequivocal categorization impossible. For clarity, only the most important ligands are specified.

Structural determinants of the receptor function

Extracellular domain

N-terminal region

The apical part of the receptor is the least examined region of ECD, and its functional role remains unclear. Because this top region is not a separate domain, its precise boundaries have not been definitively established. However, it can be intuitively described as the region above the orthosteric ligand-binding site (Figures 1A,D and 10). Interestingly, in contrast to other parts of the receptor, this region is relatively rich in unstructured loops, indicating significant mobility. Starting from the N-terminus, the region consists of the α1 helix (Figure 10), a ~15 residue loop leading into the β1 strand, followed by another ~20 residue loop between the β2 and β3 strands (partially parallel to the α1 helix). This is followed by the short β3 strand, a second helical region (α2 helix below α1 helix) and a short loop before the β4 strand. Additionally, a short loop between strands β5′ and β6 can also be assigned to this area.

Figure 10. The N-terminal region of the receptor. α1 helices are located in the apical part of the ECD. The unstructured loop connecting the α1 helix and β1 strand of the principal subunit is located below the α1 helix of the complementary subunit, allowing for non-covalent interactions at the subunit interface (via, e.g., β2F31 and α1F15). The α1 helix is also parallel to the loop of the same subunit that connects the β2 and β3 strands located above the β strands, forming the orthosteric binding site.

Unfortunately, there has been limited research specifically focused on this region, and some of the conclusions discussed stem from studies on other pLGIC family members. Interestingly, combined voltage-clamp and fluorometry tagging at GABAAR sites α1E122C(r) and β2P120C (both located in the respective subunits’ β5′-β6 loops, Figure 10) showed changes in fluorescence during GABA stimulation (Muroi et al. Reference Muroi, Czajkowski and Jackson2006), but not with pentobarbital (Muroi et al. Reference Muroi, Theusch, Czajkowski and Jackson2009). This suggests that GABA and pentobarbital induce distinct structural rearrangements and activate GABAARs through different molecular mechanisms. Investigations of chimeric receptors (ECD of α7 nAChR and TMD of GluCl) have further highlighted the importance of this region. Deletion of the α1 helix abolished neurotransmitter binding, without affecting receptor expression or trafficking to the cell surface (Bar-Lev et al. Reference Bar-Lev, Degani-Katzav, Perelman and Paas2011). Structural studies on GlyR have also emphasized the importance of this area, identifying a novel modulator binding site between the principal subunit’s α1 helix–β1 loop and the α1 helix and β3 strand of the complementary subunit (Huang et al. Reference Huang, Shaffer, Ayube, Bregman, Chen, Lehto, Luther, Matson, McDonough, Michelsen, Plant, Schneider, Simard, Teffera, Yi, Zhang, DiMauro and Gingras2017). Most notably, structural data on α1β3γ2 GABAAR revealed significant differences in the conformation of this region when bound with GABA alone versus GABA and PTX, an open channel blocker (Masiulis et al. Reference Masiulis, Desai, Uchański, Serna Martin, Laverty, Karia, Malinauskas, Zivanov, Pardon, Kotecha, Steyaert, Miller and Aricescu2019). The authors proposed that the GABA/PTX-bound receptor is trapped in a preactive state, with the ECD adopting an active conformation, while the TMD cannot undergo activation due to PTX. This state is characterized by a rotation of the ECD, involving movement of the principal subunit’s α1-helix–β1 loop toward the complementary subunit’s α1 helix and β3 strand. Similar observations were made when comparing structures of GABA-bound versus bicuculline-bound α1β2γ2 receptors (Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020). In the GABA-bound state, the N-terminal regions of the subunit interfaces are closer together than when an antagonist is present. Altogether, the data on GABAAR and other pLGIC family members indicate the importance of the intersubunit cavity lined by the α1 helix–β1 strand loop (principal side) and α1-helix and β3 strand (complementary side).

Further insights were gained from single-channel and macroscopic recordings of α1β2γ2 GABAARs with point mutations at β2F31 (α1 helix–β1 strand loop, Figure 10) and α1F14(r) (α1 helix, Figure 10) positions (Terejko et al. Reference Terejko, Michałowski, Dominik, Andrzejczak and Mozrzymas2021a). Mutations of each of these residues to cysteine clearly affected the receptor gating as they slowed down the rise time of the current responses to saturating [GABA], decreased the macroscopic desensitization, and accelerated the deactivation kinetics. Open probability and distributions of open and shut dwell times observed in single-channel recordings were also affected. However, only minor effects on the dose–responses were observed, indicating a limited impact of these mutations on the binding process. Interestingly, double cysteine mutations (giving the possibility of disulfide bond formation) resulted in a reversal of these changes, yielding a phenotype similar to that of the WT. These functional data, together with molecular modeling, clearly indicated that these two peripheral phenylalanine residues (β2F31 and α1F14(r)) are involved in regulating receptor gating, despite the particularly large distance between these residues and the channel gate located in the TMD. Moreover, the reversal of modifications of gating in the double mutant underscores intersubunit interaction, indicating that compaction of this region, facilitated by π-stacking interactions of these phenylalanine residues, may be of key importance for activation mechanisms in this region. Surprisingly, this mechanism is possibly conserved in anionic GlyR (Huang et al. Reference Huang, Shaffer, Ayube, Bregman, Chen, Lehto, Luther, Matson, McDonough, Michelsen, Plant, Schneider, Simard, Teffera, Yi, Zhang, DiMauro and Gingras2017) but not in other members of the pLGICs family.

Orthosteric and allosteric binding sites at the ECD

In contrast to the N-terminal region, the subunit areas at the intersubunit interfaces within the central ECD have been thoroughly studied because of their early recognized role in receptor function. ECD is particularly rich in β-strands, which form a relatively rigid structure stabilized by hydrogen bonds (Figure 1D). For a given subunit, the ECD is formed by two β sheets (inner at the complementary subunit and outer at the principal subunit; Figure 1D), both of which are built from β strands organized in an antiparallel fashion (outer: β4, β7, β10, and β9; inner: β8, β1 (G), β2 (D), β6, β5, and β3). In addition, structures within the binding site were named loops A–G. These structural elements are defined loops for historical reasons, although some are at least partially β-strands. At the principal side of the binding site, the following loops are present: loop A, made of β4 strand and loop connecting it with β5 strand; loop B, consisting of β7 strand fragment and loop connecting it with β8 strand; and loop C, connecting strands β9 and β10. Loop between β6 and β7 strands is a so-called Cys-loop, a characteristic structure for the ‘Cys-loop family’ of the receptors. In the complementary side, another set of loops can be distinguished: loops D, E, and G are in the middle parts of the β2, β6, and β1 strands, respectively, and loop F is located between the two parts of the β8 strand (Figure 1D). The orthosteric neurotransmitter binding sites are located at the β(+)/α(−) subunit interfaces. Besides the neurotransmitter GABA, this area is a binding site for other agonists (Figure 2), such as muscimol (Beaumont et al. Reference Beaumont, Chilton, Yamamura and Enna1978) and partial agonist piperidine-4-sulfonic acid (P4S, Krogsgaard-Larsen et al. Reference Krogsgaard-Larsen, Falch, Schousboe, Curtist and Lodget1980) or competitive antagonists like bicuculine (Curtis et al., Reference Curtis, Duggan, Felix and Johnston1970) and gabazine (SR-95531, Wermuth and Bizière Reference Wermuth and Bizière1986). In addition to the orthosteric binding site, a secondary allosteric site – responsible for the binding of BDZs – is located at the α(+)/γ(−) interface. This allosteric site shares a similar architecture with the orthosteric site. Given the numerous comprehensive reviews on this topic, it is not covered in detail here. For further reading, we recommend the recent work by Goldschen-Ohm (Reference Goldschen-Ohm2022) and an insightful review by Sigel and Ernst (Reference Sigel and Ernst2018).

Loop A

Loop A (Figures 1A and D and 11) was investigated in α1β2 receptors expressed in oocytes by Boileau et al. (Reference Boileau, Newell and Czajkowski2002), who performed a substituted-cysteine accessibility method (SCAM) analysis for residues β2W92–β2D101. The β2W92C (forming a hydrophobic core) mutant was not functional, while the greatest effect on EC50 was observed for β2Y97C, which along with β2L99C (Figure 11A and B) was protected by GABA and SR-95531 from the methane thiosulfonate (MTS) reagent. Interestingly, mutation β2L99C resulted in spontaneous activity of β2 homomers, indicating the role of this residue in receptor gating. The general pattern of modification suggested that β2W92–β2D101 formed a β-strand conformation. Further work by Padgett et al. (Reference Padgett, Hanek, Lester, Dougherty and Lummis2007) examined the interactions between the ‘aromatic box’ and ligands, providing evidence of cation–π interactions with β2Y97 (Figure 11A and B) in α1β2 receptors. However, docking of the GABA molecule in models of the α1β2 and ρ type GABAAR indicated that in the latter receptor type, cation–π interactions were present not in loop A but in loop B (Lummis et al. Reference Lummis, Beene, Harrison, Lester and Dougherty2005). Tran et al. (Reference Tran, Laha and Wagner2011) examined a ‘tight coupling’ between aromatic β2Y97 (and also β2F200, loop C, Figure 11A and B) and several arginine residues (α1R67 loop D, α1R120 β5 strand, α1R132 loop E, and β2R207 β10 strand below loop C, Figure 11) which were previously implicated in ligand binding. By using mutant cycle analysis (MCA) on α1β2γ2 GABAARs expressed in HEK293 cells, Tran et al. (Reference Tran, Laha and Wagner2011) found that a significant functional coupling took place between β2Y97 at loop A and the following residues: β2F200, β2R207, and α1R132. They proposed that the tight coupling between β2Y97 and β2F200 plays a critical role in mediating GABA binding. In another study, the β4-β5 strand linker (below loop A) was investigated, which is thought to connect the neurotransmitter binding site to the inner β-sheet (Venkatachalan and Czajkowski Reference Venkatachalan and Czajkowski2012). They introduced multiple glycine residues into this region of the β subunit (α1β2γ2 receptor expressed in oocytes) to alter its length and flexibility. Their findings demonstrated that the more glycine residues were incorporated, the less GABA-induced activation was observed, and the competitive antagonist SR-95531 began to act as a partial agonist. This effect was not observed in the α or γ subunits, but mutations in these subunits altered the actions of pentobarbital and flurazepam, suggesting a role in receptor gating. Laha and Tran (Reference Laha and Tran2013) further examined the impact of the tyrosine β2Y97 mutation to alanine (together with β2Y157 and β2Y205, Figure 11A and B) on the expression and kinetics of α1β2γ2 and α1β2 receptors in HEK293 cells. Interestingly, the assembly or trafficking of GABAARs was impaired when tyrosine mutants were expressed as αβ receptors but not in αβγ assembly. Mutation of tyrosine β2Y97 resulted in accelerated deactivation and slower binding, providing evidence for the weakening of GABA-binding reaction. However, mutations at β2Y157 and β2Y205 were considerably more detrimental than that at β2Y97, suggesting that interactions involving multiple tyrosine residues are likely to occur during the binding process. Structural studies have confirmed the role of loop A in forming of the GABA-binding site. For instance, Masiulis et al. (Reference Masiulis, Desai, Uchański, Serna Martin, Laverty, Karia, Malinauskas, Zivanov, Pardon, Kotecha, Steyaert, Miller and Aricescu2019) demonstrated that β3Y97 (in the α1β3γ2 receptor) forms a hydrogen bond, rather than a cation–π interaction, with the GABA amino group. Collectively, these studies highlight the importance of loop A and its key residue β2Y97 in agonist binding, although the specific contributions of this residue vary between studies (compare Padgett et al. (Reference Padgett, Hanek, Lester, Dougherty and Lummis2007) and Laha and Tran (Reference Laha and Tran2013)). The involvement of β2Y97 in receptor gating remains to be fully elucidated.

Figure 11. ECD-binding site of GABAAR. (A) General view of the orthosteric binding site. The binding cavity is lined with antiparallel β-strands of the principal (4, 7, 10, and 9, loops A and B) and complementary (5, 6, 2, 1, and 8, loops E, D, G, and F) subunits and capped by loop C (between β-strands 9 and 10). (B) In the principal subunit side loops A, B, and C are present. Loops A and B are located deep in the binding cavity, whereas loop C between strand β9 and β10 is exposed to the solvent. Residues β2Y97, β2Y157, and β2F200 are forming an ‘aromatic box’ around GABA molecule, whereas β2E155 form electrostatic interaction with ligand molecule. (C) All loops (E, D, G, and F) of the complementary side of the cavity are fragments of the respective β-strands. Strand β5 and loops E and D are located similarly to loops A and B in the deeper part of the binding site, whereas loops G and F are partially exposed to the extracellular space. The ‘aromatic box’ on the complementary side is formed with α1F65 and α1F45 whereas electrostatic interaction is present between agonist and α1R67.

Loop B

Loop B is located in the outer β-sheet and parallel to loop A (Figures 1A and D and Figure 11). Amin and Weiss (Reference Amin and Weiss1993) examined mutations at key positions within loop B, including β2Y157F/N and β2T160S/A, in α1β2γ2 GABAARs expressed in oocytes (Figure 11A and B). Their results indicated that these mutations either abolished receptor responses or significantly reduced activation by GABA and muscimol, while activation by pentobarbital remained unaffected. This is consistent with the fact that the binding site for pentobarbital is located closer to the channel gate than the GABA-binding site. Further research by Newell et al. (Reference Newell, McDevitt and Czajkowski2004) used the SCAM method (α1β2 assembly expressed in oocytes) to study a large part of loop B (β2I154–β2D163). Most mutations within this region affected GABA EC50, but only β2E155C, β2S156C, and β2D163C (Figure 11A and B) also affected SR-95531 EC50, supporting their role in the binding process. Additionally, β2T160C and β2D163 mutations slowed the modification rate by 2-aminoethyl MTS hydrobromide–biotin (MTSEA-biotin) upon GABA, SR-95531, and also pentobarbital-induced activation, indicating their role in gating and long-range allosteric interactions. β2E155 was also postulated to play a role in gating, as its mutation caused spontaneous activity of the receptor.

Padgett et al. (Reference Padgett, Hanek, Lester, Dougherty and Lummis2007) further supported the role of loop B in agonist binding, demonstrating that removal of the hydroxyl group from β2Y157 significantly increased the GABA EC50. In a study of the loop C region, Venkatachalan and Czajkowski (Reference Venkatachalan and Czajkowski2008) proposed that β2E153 forms a salt bridge with β2K196 (loop C) and is thereby involved in shaping the dynamics of this loop. Laha and Tran (Reference Laha and Tran2013) examined α1β2γ2 and α1β2 receptors with the β2Y157A mutation, finding significant effects on GABA EC50, deactivation and binding rate, but only a minor impact on desensitization. This led them to propose that β2Y157 primarily contributes to agonist binding, with the hydroxyl group essential for receptor functionality. Structural research by Masiulis et al. (Reference Masiulis, Desai, Uchański, Serna Martin, Laverty, Karia, Malinauskas, Zivanov, Pardon, Kotecha, Steyaert, Miller and Aricescu2019) on α1β3γ2 receptors further highlighted the role of loop B in agonist binding. They showed that GABA forms hydrogen bonds with β3E155 carboxyl and β3S156 and β3Y157 (a residue also attributed to binding site ‘aromatic box’) main chain carbonyls. In addition, upon agonist binding, loop B moves toward the complementary subunit, and β3D163 forms a salt bridge with α1R85 located in the β3 strand. Considering a particularly large impact of the β2E155 residue mutation on binding (the largest increase in EC50) and its effect on spontaneous activity, we addressed the impact of the β2E155C mutation on receptor gating by recording macroscopic current responses to rapid GABA (or muscimol) application, as well as single-channel recordings (Jatczak-Śliwa et al. Reference Jatczak-Śliwa, Kisiel, Czyzewska, Brodzki and Mozrzymas2020). Our data confirmed the involvement of this residue in receptor gating, specifically in the preactivation transition and microscopic desensitization. Additionally, the β2E155 residue has also been proposed to act as a proton sensor at the GABA-binding site (Michałowski et al. Reference Michałowski, Czyżewska, Iżykowska and Mozrzymas2021). In conclusion, loop B is crucial for ligand binding (with β2E155 being particularly critical) and also exerts influence on receptor gating.

Loop C

The loop C (Figures 1A and D and 11) of the principal subunit (β) undergoes significant movement upon ligand binding to the orthosteric binding site, a phenomenon termed ‘capping’. This loop has drawn considerable attention due to its high mobility, which appears compatible with the structural rearrangements associated with receptor conformational transitions. Early studies on loop C were based on mutagenesis. Mutations of the β2Y205 residue (Figure 11A and B) to F, S,N in α1β2γ2 GABAARs expressed in oocytes by Amin and Weiss (Reference Amin and Weiss1993) resulted in receptors that were either non-responding or exhibited highly reduced responsiveness to the agonist, thus confirming its important functional role. Czajkowski and Wagner (Reference Czajkowski and Wagner2001) recorded currents mediated by α1β2 receptors expressed in oocytes with cysteine mutations at positions β2V199–β2S209. GABA or SR-95531 (competitive antagonist) application slowed the MTSEA-biotin modification of β2S204, β2Y205, β2R207 (β10 strand, below loop C, Figure 11A and B), and β2S209 (close to β2R207) mutants, indicating a role in forming the binding site. In addition, these studies suggested that the C loop structure was inconsistent with α- or β-helices, suggesting a form of water accessible coil. The general pattern of modifications was interpreted by these authors as β2G203–β2S209 being water accessible, whereas β2V199–β2T202 were not, although this interpretation was questioned in more recent studies (β9 strand and loop C are solvent-exposed according to structural data, for example, Miller and Aricescu (Reference Miller and Aricescu2014)). Since orthosteric binding sites in α1β2γ2 GABAARs are surrounded by different quaternary neighborhood (one is flanked by γ2 and β2, while the second is flanked by α1 and γ2 subunits), Baumann et al. (Reference Baumann, Baur and Sigel2003) used concatemers expressed in oocytes with β2Y205S point mutations at specific binding sites and found that GABA molecules bind with a higher affinity to the binding site β(+)/α(−) flanked by α1 and γ2 subunits. This non-equivalence of orthosteric binding sites in GABAARs is not surprising, as previous kinetic analyses of GABAergic currents have clearly indicated this difference (Maric et al. Reference Maric, Maric, Wen, Fritschy, Sieghart, Barker and Serafini1999; Mozrzymas et al. Reference Mozrzymas, Barberis, Mercik and Zarnowska2003). Residue β2R207 was thoroughly investigated by Wagner et al. (Reference Wagner, Czajkowski and Jones2004) by analyzing macroscopic responses to rapid agonist applications and single-channel currents of α1β2 receptors expressed in HEK293 cells. They found that macroscopic desensitization was unaffected, while deactivation was strongly accelerated. These findings, together with the lack of effect on the maximum open probability, led to the conclusion that this residue is primarily involved in binding but not in gating. Padgett et al. (Reference Padgett, Hanek, Lester, Dougherty and Lummis2007) showed that removal of the hydroxyl group from β2Y205 significantly increased GABA EC50. Loop C was further examined by Venkatachalan and Czajkowski (Reference Venkatachalan and Czajkowski2008), who performed patch-clamp recordings of α1β2γ2 GABAARs expressed in oocytes with various mutations. Specifically, charge reversal or neutralization of residues β2E153 (β7 strand below loop B, Figure 11A and B) or β2K196 (β9 strand below loop C, Figure 11A and B) increased GABA EC50, whereas the double charge switch mutation rescued it to the WT level, indicating the presence of a salt bridge between these residues. These observations were confirmed by MCA and modification with the MTS reagent. In addition, these authors proposed that β2E153 may also interact with β2E155 and β2R207. Tran et al. (Reference Tran, Laha and Wagner2011) used MCA on α1β2γ2 GABAARs expressed in HEK293 cells and showed coupling between β2F200 and each of the following residues: β2Y97 (loop A), α1R132 (loop E), and β2R207 (β10 strand, below loop C). The interactions between GABA molecule and positively charged binding site residues (arginine) were investigated by Goldschen-Ohm et al. (Reference Goldschen-ohm, Wagner and Jones2011) using macroscopic and single-channel recordings of α1β2 receptors expressed in HEK293 cells. One of them, β2R207, when mutated to alanine affected the binding and unbinding rates, indicating a major effect on agonist binding but not on gating. The β2Y205A mutation in both α1β2γ2 and α1β2 type receptors expressed in HEK293 cells was investigated by Laha and Tran (Reference Laha and Tran2013), who reported altered GABA EC50, decreased binding rate, accelerated deactivation and only minimal alterations in desensitization, indicating a major effect on agonist binding but not on gating. In addition, these authors have also shown that besides the phenyl ring, as in the case of β2Y97 (from loop A), the hydroxyl group in this residue is also essential. In our computational study (Michałowski et al. Reference Michałowski, Kraszewski and Mozrzymas2017), we showed that loop C movement is considerably larger in the principal subunit (β2) than in the complementary (α1). We may speculate that capping of loop C at the principal subunit might be related to conformational changes upon receptor gating. In another simulation study, it was shown that for successful binding, the ligand must follow the path from the solution to localization ‘underneath’ the loop C (Carpenter and Lightstone Reference Carpenter and Lightstone2016). The loop C closure movement upon GABA binding was shown by Masiulis et al. (Reference Masiulis, Desai, Uchański, Serna Martin, Laverty, Karia, Malinauskas, Zivanov, Pardon, Kotecha, Steyaert, Miller and Aricescu2019) in α1β3γ2 receptors. The authors also attributed loop C residues β3F200 and β3Y205 to the binding site ‘aromatic box’, showing that β3Y205 forms a cation–π interaction with GABA amino group and β3T202 – a hydrogen bond with ligand’s carboxylate. Taken together, the available evidence clearly demonstrates that C loop is involved in agonist binding, but its role in receptor gating remains unclear. In our recent study (Terejko et al. Reference Terejko, Kaczor, Michałowski, Dąbrowska and Mozrzymas2020), we investigated the role of loop C in gating mechanisms by analyzing the impact of β2F200 residue mutations in α1β2γ2 GABAARs. Extensive analyses and modeling of current responses to saturating agonist applications demonstrated that this mutation strongly affected preactivation, opening, closing, and desensitization, which are all considered gating steps. We also found that this mutation altered receptor sensitivity to the BDZ flurazepam, which we attributed to a change in preactivation kinetics. Thus, like loops A and B, loop C contributes to the formation of the hydrophobic binding cassette on the principal subunits, playing a crucial role not only in agonist binding but also in transmitting the activation signal to the channel gate, thereby influencing receptor gating.

Loop D

It is interesting to note that whereas loops A–C at the principal subunit are predominantly unstructured coils, the loops on the complementary subunit (α) largely form a β-sheet structure with antiparallel β-strands. This arrangement promotes the formation of numerous hydrogen bonds, which enhance the stability and rigidity of the complementary subunit structure (Figures 1A and D and 11). Early mutagenesis studies focused on identifying residues within the complementary subunit that interact with orthosteric ligands and determine the binding process. Sigel et al. (Reference Sigel, Baur, Kellenberger, Malherbel, Malherbe and Malherbel1992) examined α1β2γ2 GABAARs expressed in oocytes with point mutations at homologous positions of each subunit building up the receptor, targeting: α1F64L(r), β2Y62L, and γ2F77L (Figure 11A and C). Among these substitutions, only the mutation in the α1 subunit significantly altered the EC50 for GABA, highlighting its unique role in binding. Further work (Boileau et al. Reference Boileau, Evers, Davis and Czajkowski1999) employed recordings from HEK293 cells in addition to oocytes and α1β2 receptors with cysteine substitutions at the α1Y59(r)–α1K70(r) positions (Figure 11A and C). Mutations α1F64C(r) and α1R66C(r) (Figure 11A and C) increased GABA EC50, and in the case of α1R66C(r) and α1S68C, GABA presence protected from MTSEA-biotin modification. The general modification pattern indicated that these residues form a β-strand. Hartvig et al. (Reference Hartvig, Lükensmejer, Liljefors and Dekermendjian2002) examined the role of positively charged arginine residues in α5β2γ2 GABAARs expressed in CHO cells: residues α5R34, α5R70, α5R77, α5R79, α5R123, α5R135, α5R190, and α5R224 were mutated to lysine. Mutants of α5R70 (homolog of α1R67, Figure 11A and C) and α5R123 (homolog of α1R120, Figure 11A and C) increased GABA EC50, in agreement with earlier results on α5β2γ2 GABAARs. Previous experiments on loop D (Boileau et al. Reference Boileau, Evers, Davis and Czajkowski1999) were extended by the use of various MTS reagents (Holden and Czajkowski Reference Holden and Czajkowski2002), which confirmed earlier results and showed that α1F64(r) and α1R66(r) are located in the aqueous core of the binding site. Loop D was also investigated by Baur and Sigel (Reference Baur and Sigel2003), who showed that mutation β2Y62L (homolog of α1F65, Figure 11A and C) had a minor effect on receptor activation supporting the role of the β/α and not α/β interface in agonist binding. Using concatemers expressed in oocytes with a point mutation (α1F65L), Baumann et al. (Reference Baumann, Baur and Sigel2003) have shown that the effects of α1 subunit substitution were the same in both binding sites. Padgett et al. (Reference Padgett, Hanek, Lester, Dougherty and Lummis2007) demonstrated that no cation–π interaction exists between GABA and α1F65 (Figure 11A and C). Additionally, this residue exhibits tolerance to small chemical modifications, such as fluorination. Using docking and homology modeling, Lummis (Reference Lummis2009) postulated an electrostatic interaction between the agonist’s carboxyl group and α1R67 residue. Experiments done by Goldschen-Ohm et al. (Reference Goldschen-ohm, Wagner and Jones2011) showed that mutation α1R67A changed only binding and unbinding kinetics, indicating involvement in binding rather than in gating. Thus, the available data provided strong evidence that loop D and α1F64(r), in particular, play a key role in ligand binding, but its role in gating remained elusive. To address the issue of the involvement of this loop in receptor gating, our group examined the role of the α1F64(r) residue in gating properties of the α1β2γ2 GABAAR by analyzing current responses to rapid agonist applications. Data analysis and modeling have provided evidence that the α1F64(r) residue is involved not only in binding but also in gating properties (Szczot et al. Reference Szczot, Kisiel, Czyzewska and Mozrzymas2014). In particular, evidence was provided that mutations in this residue strongly affect the so-called flipping transition (preactivation), a state of the receptor previously postulated for other pLGICS (Jadey and Auerbach Reference Jadey and Auerbach2012; Lape et al. Reference Lape, Colquhoun and Sivilotti2008; Mukhtasimova et al. Reference Mukhtasimova, Lee, Wang and Sine2009). Mutations of this residue to L, A and C also increased the spontaneous activity of the receptor (Jatczak-Śliwa et al. Reference Jatczak-Śliwa, Terejko, Brodzki, Michałowski, Czyzewska, Nowicka, Andrzejczak, Srinivasan and Mozrzymas2018), and mutation to C affected the modulatory effect of alkaline pH (Kisiel et al. Reference Kisiel, Jatczak-Śliwa and Mozrzymas2019). This residue was also confirmed to be a part of the binding site ‘aromatic box’, whereas neighboring α1R67 was shown to form a salt bridge with GABA carboxylate (Masiulis et al. Reference Masiulis, Desai, Uchański, Serna Martin, Laverty, Karia, Malinauskas, Zivanov, Pardon, Kotecha, Steyaert, Miller and Aricescu2019). Studies on the role of residue α1F64(r) on GABAAR gating were extended to investigations of the role of preactivation in spontaneous and singly-bound states (observed at low [GABA]), and it was proposed that liganded receptors (singly and doubly bound) undergo flipping transition prior to opening, whereas spontaneous activity does not (Kisiel et al. Reference Kisiel, Jatczak, Brodzki and Mozrzymas2018. Considering that loop D is part of a rigid β-sheet, we hypothesized that this structure is particularly predisposed to convey an activatory signal from the binding site toward the channel gate. To test this possibility, we considered a glycine mutant of the α1F64(r) residue (an amino acid known to reduce the rigidity and flexibility of polypeptides). The glycine mutation of the α1F64(r) residue resulted in the strongest impact among all α1F64(r) mutations, affecting both binding and all gating steps considered in our model – preactivation, opening/closing, and desensitization (Kłopotowski et al. Reference Kłopotowski, Michałowski, Gos, Mosiądz, Czyżewska and Mozrzymas2023. This confirms the significant role of the α1F64(r) residue in receptor gating and supports the notion that the rigid β-sheet structure, which hosts the binding site loops at the complementary subunit, may facilitate efficient signal transduction from the binding site to the channel gate.

Loop E and β5 strand

Continuing the description of the orthosteric binding site at the inner sheet of the complementary subunit, adjacent to loop D, there is the loop E, which is a part of β-strand 6 (Figures 1A and D and 11). Studies on α1β2γ2 receptors expressed in Sf-9 insect cells showed that a relatively small change in the primary structure, via the α1I121V mutation (β5 strand; Westh-Hansen et al. Reference Westh-Hansen, Rasmussen, Hastrup, Nabekura, Noguchi, Akaike, Witt and Nielsen1997), markedly reduced GABAAR ligand sensitivity. In a subsequent study, the same authors reported that the α1R120K substitution (Figure 11A and C, β5 strand; Westh-Hansen et al. Reference Westh-Hansen, Witt, Dekermendjian, Liljefors, Rasmussen and Nielsen1999) resulted in a 180-fold increase in GABA EC50. Similarly, Hartvig et al. (Reference Hartvig, Lükensmejer, Liljefors and Dekermendjian2002) demonstrated that the α5R123K mutation (equivalent to α1R120) in α5β2γ2 GABAARs significantly increased GABA EC50, further confirming the key role of α5R123K (α1R120) in agonist binding.

To investigate conformational changes in the ECD upon GABA binding, Muroi et al. (Reference Muroi, Czajkowski and Jackson2006) labeled α1β2γ2 and α1β2 receptors with point cysteine mutations at positions α1E122C(r), β2P120, and γ2N135 (β5 and loop E linker; Figure 11A and C), α1L127C(r), β2L125, and γ2 L140 (loop E; Figure 11A and C) using environment-sensitive fluorophores. GABA induced fluorescence changes in both the α1 and β2 subunits (α1β2 receptor), while the antagonist SR-95531 affected only the α1. In the α1β2γ2 receptor, reduced fluorescence was observed, likely indicating that less movement was required for receptor activation, and no fluorescence was detected in the γ2 subunit. In addition, the fluorescence remained unaltered upon macroscopic desensitization onset. In a subsequent study, the same group (Muroi et al. Reference Muroi, Theusch, Czajkowski and Jackson2009) used fluorophore labeling in α1β2 GABAARs expressed in oocytes to elucidate activation-induced transitions in the ECD and TMD. Labels were placed at residues (mutations to cysteine) α1L127(r) (loop E top), β2L125 (same location), and β2L274 (M2-M3 loop). Together with previous results (Muroi et al. Reference Muroi, Czajkowski and Jackson2006), the authors found that GABA binding increased fluorescence in loop E, while pentobarbital (at high inhibitory concentrations) decreased it through a long-range allosteric mechanism. In addition, the effects on the M2-M3 loop upon GABA or pentobarbital (at lower activating/positively modulating concentrations) were different, indicating distinct activation mechanisms for these ligands.

The region encompassing β4, β5 strands and loop E was further explored by Kloda and Czajkowski (Reference Kloda and Czajkowski2007) using SCAM on α1M113(r)–α1L132(r) residues in α1β2γ2 receptors expressed in oocytes. Residues located in the β5 strand: α1N115(r), α1L117(r), and in loop E: α1L127(r), α1T129(r), and α1R131(r) (Figure 11A and C) were involved in the binding of both GABA and SR-95531. Moreover, the chemical reactivity of α1E122(r) (β5 and loop E linker (Figure 11A and C) with MTSEA-biotin was modulated by GABA and pentobarbital, but not by SR-95531, whereas α1L127C was affected by GABA and SR-95531, but not by pentobarbital. Additionally, α1E122(r), α1L127(r), and α1R131(r) chemical reactivities were all affected by flurazepam. These findings indicate which specific residues play a role only in ligand binding and which are also important for effective gating and modulation by BDZ.

Voltage-clamp fluorometry was used by Akk et al. (Reference Akk, Li, Bracamontes, Wang and Steinbach2011) for the mutant α1L127C (loop E, Figure 11A and C). Changes in fluorescence were observed upon GABA activation, but remained unaltered during desensitization onset, indicating a clear conformation-sensitivity of this method. Another interaction between ECD residues was found by Laha and Wagner (Reference Laha and Wagner2011): mutations α1R120A (β5 strand, Figure 11A and C) and β2D163A (β7–β8 loop; Figure 11A and B) of α1β2γ2 receptors (expressed in HEK293 cells) increased GABA EC50 and accelerated deactivation and unbinding, indicating the existence of a state-dependent salt bridge between these residues at the binding site. The interactions between GABA and binding site residues were investigated using macroscopic and single-channel recordings of α1β2 receptors expressed in HEK293 cells by Goldschen-Ohm et al. (Reference Goldschen-ohm, Wagner and Jones2011), who showed that the mutation α1R132A (bottom of loop E, Figure 11A and C) resulted in a new long-lived open state of the receptor, indicating the role of this residue in the gating process. In addition, the role of the α1T130 residue in GABA binding (by forming a hydrogen bond with the GABA carboxyl group) was confirmed by structural studies (Masiulis et al. Reference Masiulis, Desai, Uchański, Serna Martin, Laverty, Karia, Malinauskas, Zivanov, Pardon, Kotecha, Steyaert, Miller and Aricescu2019).

Loop F

Loop F may be defined as a structure made of strand β8 or more accurately (according to structural data, e.g., Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020; Masiulis et al. Reference Masiulis, Desai, Uchański, Serna Martin, Laverty, Karia, Malinauskas, Zivanov, Pardon, Kotecha, Steyaert, Miller and Aricescu2019) of the β8 strand (starting after the loop following the β7 strand), a coil breaking the strand, and the remaining part of the β8 strand (that can be called β8′; Figures 1A and D and 11). In addition, a fragment of loop following the β8′ leading to the β9 strand also can be attributed to loop C. Newell and Czajkowski (Reference Newell and Czajkowski2003) examined loop F (α1P174(r)–α1D191(r)) using the SCAM method on α1β2 receptors expressed in oocytes. Most of the mutations did not alter GABA EC50; however, among the studied substitutions, α1D183C had the largest effect on apparent GABA affinity. MTSEA-biotin modification indicated that the loop is unstructured and solvent exposed and in the case of α1V178C(r), α1V180C(r), and α1D183C(r) (Figure 11A and C), both GABA and SR-95531 slowed down the modification rate. Moreover, activation with pentobarbital accelerated the modification of α1V180C(r) and α1A181C(r) and slowed that of α1R186C(r), indicating that this region of the α1 subunit is involved in channel-gating transitions. Transitions inside the ECD upon ligand binding have also been examined in α1β2γ2 GABAARs expressed in oocytes using voltage-clamp fluorometry (Wang et al. Reference Wang, Pless and Lynch2010). Mutations α1R186C(r), β2I180C and γ2S159C (each residue at loop F bottom part) were investigated. In the case of the α1 subunit, agonist, partial agonist, or antagonist caused a similar increase in fluorescence, indicating binding, but not gating related effect. However, changes in fluorescence in the β2 and γ2 subunit mutants were observed only upon agonist binding, indicating an allosteric mechanism or interaction with loop C in the case of the β2 subunit. In another study, Eaton et al. (Reference Eaton, Lim, Bracamontes, Steinbach and Akk2012) used the SCAM method and found that GABA presence reduced accessibility to γ2S195 (loop F), once again indicating allosteric effects within the ECD. In a computational study by Carpenter et al. (Reference Carpenter, Lau and Lightstone2012), the authors addressed the role of loop F in two types of GABAARs: α1β2γ2 and α6β3δ, the latter being characterized by a much higher GABA sensitivity. They found that the major difference between the binding sites in the two considered receptors was the extent of loop F involvement, with a larger contribution in the case of the α6β3δ receptor. In addition, free energy calculations confirmed that the α6β3δ binding pocket is characterized by an increased affinity for GABA and indicated that the possible role of loop F in the GABAAR family is to modulate GABA affinity. Mihalik et al. (Reference Mihalik, Pálvölgyi, Bogár, Megyeri, Ling, Barkóczy, Bartha, Martinek, Gacsályi and Antoni2017) studied the action of novel GABAAR orthosteric antagonists (for αxβ2γ2 receptors) based on a tricyclic oxazolo-2,3-BDZ scaffold. In silico modeling predicted that the test compounds docked in the GABA-binding pocket would interact with various residues at the orthosteric binding site at the α- and β-subunit interface, including an interaction with loop F of the α subunit. This prediction was further investigated by replacing the amino-terminal variable segment of loop F of the α5 subunit with the corresponding residues of the α1 and α2 subunits. When tested with aforementioned inhibitors, the receptors formed by the modified α5 subunits displayed the pharmacological phenotype of the source of loop-F, thus confirming the key role of this loop in determining the pharmacological profile of studied compounds. A similar approach was employed by Pálvölgyi et al. (Reference Pálvölgyi, Móricz, Pataki, Mihalik, Gigler, Megyeri, Udvari, Gacsályi and Antoni2018), who examined the effect of replacing a short, variable segment of loop F of the GABAAR (αxβ2γ2 receptors) α5 subunit with the corresponding segment of the α2 subunit (GABAA5_LF2) and vice versa (GABAA2-LF5). When compared with their respective wild-type counterparts, GABAA5_LF2 receptors displayed enhanced sensitivity toward GABA (as for α2 subunit containing receptors), while in GABAA2-LF5 sensitivity was diminished. In summary, loop F was found to play an important role in agonist binding, and the SCAM method provided indirect evidence also for its involvement in gating transitions.

Loop G

Loop G is a neighboring β strand (β1) to loop D (β2) while both strands are oriented in the antiparallel fashion to each other forming a β sheet (Figures 1A and D and 11). Loop G was examined using electrophysiological recordings of currents mediated by α1β2γ2 GABAARs (expressed in HEK293 cells) with cysteine substitutions at residues α1D43(r)–α1T47(r) (Baptista-Hon et al. Reference Baptista-Hon, Krah, Zachariae and Hales2016). Mutations α1D43C(r) (Figure 11A and C) and α1T47C(r) reduced the apparent potency of the agonist and were accessible to MTSEA modification, whereas the α1F45C(r) mutation (Figure 11A and C) led to biphasic dose–response characteristics and spontaneous activity, but was not accessible for MTSEA modification. However, the presence of GABA reduced MTSEA modification in the α1D43C(r) and α1T47C(r) mutants. Thus, these results indicate the movement of loop G during activation of the receptor. In a continuation of their work, the authors showed that α1D43C(r) and α1T47R(r) mutations reduced the maximal open probability of the receptor, slowed down desensitization, and accelerated deactivation. In addition, the α1T47R(r) mutation affected the spontaneous gating of another mutant (β2L285R), highlighting the role of loop G not only in binding but also in gating (Baptista-Hon et al. Reference Baptista-Hon, Gulbinaite and Hales2017). In a study aimed at turning GABA β3 homomers into functional receptors responsive to GABA Gottschald Chiodi et al. (Reference Gottschald Chiodi, Baptista-Hon, Hunter and Hales2018), introduced the following mutations in the complementary side of the orthosteric binding site in loop D (Y87F and Q89R), loop E (G152T), and loop G (N66D and A70T). However, electrophysiological investigations indicated that the two substitutions from loops D and E, Q89R and G152T, in β3 GABAAR were sufficient to reconstitute GABA-mediated activation, indicating that, at least in this model, the G loop was not essential for converting β3 GABAAR into a GABA-activable receptor. To further explore the role of loop G in binding and gating, our group performed a detailed analysis of α1β2γ2 GABAARs activity with point mutations of the α1F45(r) residue (Figure 11A and C) expressed in HEK293 cells (Brodzki et al. Reference Brodzki, Michałowski, Gos and Mozrzymas2020). Macroscopic measurements showed a significant increase of GABA EC50 in the case of mutations to cysteine, glycine, and lysine and a minor increase for leucine substitution, thus confirming the involvement of this residue in the binding process of GABA. However, substitutions with C, G, and K also affected macroscopic desensitization (slower and weaker), thus providing evidence that the gating process was also affected. In single-channel recordings at saturating GABA concentrations, changes were observed in both open- and shut-time distributions. Kinetic modeling clearly indicated that, beyond its role in agonist binding, loop G is important in channel closing/opening and desensitization transitions. Molecular modeling revealed that although α1F45(r) and loop G do not directly contact with the GABA molecule at the binding site, they play an important role in shaping the interaction between the agonist and loops other than G (Brodzki et al. Reference Brodzki, Michałowski, Gos and Mozrzymas2020). Structural studies, such as those conducted by Kim et al. (Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020), further confirmed previous conclusions on the role of the loop G in the ECD-binding site. Specifically, residues α1D44(r) and α1F45(r) are located parallel to α1R66(r) and α1F64(r) (at loop D), respectively, forming the outer layer of the complementary subunit’s β-sheet. It is noteworthy that considering the substantial impact of α1F64(r) on receptor gating, it was anticipated that the neighboring residue α1F45(r), located on the same β-sheet, would also participate in GABAA receptor conformational transitions. This is because the β-sheet likely acts as a structurally cohesive element involved in transducing the activation signal to the channel gate.

Investigations on ρ-type and resistance to dieldrin (RDL) receptor-binding site

Several studies addressed the structure–function relationship of ρ receptors, formerly classified as ‘GABAC receptors’ but now they are included in the GABAAR type family. One of the reasons for such interest is that ρ receptors may form agonist-activable functional homomers (in contrast to, e.g., β3 homomers), considerably facilitating data interpretation. It has been demonstrated that in ρ-type receptors expressed in oocytes, the ρ1Y198F residue (homolog of β2Y157 on loop B, Figure 11A and B) forms a cation–π interaction with the GABA molecule (Harrison and Lummis Reference Harrison and Lummis2006a, Reference Harrison and Lummis2006b; Lummis et al. Reference Lummis, Beene, Harrison, Lester and Dougherty2005). In addition, the authors showed that ECD’s arginines are important for receptor function: ρ1R104 (homolog of α1R67 on loop D, Figure 11A and C) mutation caused a large increase in GABA EC50 as it forms electrostatic interaction with agonist carboxyl group, ρ1R158 (homolog of α1R120 on β5 strand, Figure 11A and C) and ρ1R170 (homolog of α1R132 on loop E, Figure 11A and C) mutations resulted in nonfunctional receptors or increased GABA EC50, whereas ρ1R249 (homolog of β2R207 on β10 strand, below loop C, Figure 11A and B) had small effects on GABA EC50 and was proposed to be involved in shaping the binding site via hydrogen bonds and/or salt bridges. The prominent role of ρ1R104 (homolog of α1R67, loop D, Figure 11A and C) in agonist binding was confirmed by a molecular dynamics study (Melis et al. Reference Melis, Lummis and Molteni2008). Lummis (Reference Lummis2009) using docking and homology modeling confirmed the possibility of a cation–π interaction between GABA and ρ1Y198 (homolog of β2Y157 on loop B, Figure 11A and B). The role of loop F in receptor activation was examined by Khatri et al. (Reference Khatri, Sedelnikova and Weiss2009) in ρ type receptors expressed in oocytes using fluorophore labeling (residues ρ1L216–ρ1I229, mutations to cysteine), and the loop was proposed to be involved only in binding, but not gating transitions. In addition to ρ-type receptors, RDL receptors, found in insects, was a subject of many investigations. They share similarities with GABAARs in that they are activated by GABA and permeable to anions. The cation–π interactions (shown previously to differ between α1β2γ2 and ρ type receptors) were examined in these receptors (Lummis et al. Reference Lummis, McGonigle, Ashby and Dougherty2011) using various unnatural fluorinated phenylalanine derivative mutants. The authors showed that both F206 (homolog of β2Y157 and ρ1Y198 on loop B, Figure 11A and B) and Y254 (homolog of β2Y205 and ρ1Y247 on loop C, Figure 11A and B) residues form cation–π interaction with the GABA molecule. A molecular dynamics study of the RDL-type receptor confirmed cation–π interactions between GABA and F206 (homologous to β2Y157 in loop B, Figure 11A and B) and Y254 (homologous to β2Y205 in loop C, Figure 11A and B). In addition, the authors demonstrated that the agonist forms hydrogen bonds with E204, S205, R111, S176, T251, and exhibits ionic interactions with R111 (homolog of α1R67, loop D, Figure 11A and C) and E204 (homolog of β2E155 on loop B, Figure 11A and B) (Ashby et al. Reference Ashby, McGonigle, Price, Cohen, Comitani, Dougherty, Molteni and Lummis2012). Lummis et al. (Reference Lummis, Harrison, Wang, Ashby, Millen, Beene and Dougherty2012) showed that ρ1Y198 is the only residue that forms a cation–π interaction with GABA. They further examined the roles of other tyrosine residues, showing that ρ1Y102 (loop D, homolog of α1F65, Figure 11A and C) and ρ1Y138 (homolog of β2Y97 on loop A, Figure 11A and B) stabilize the amino group of the GABA molecule. Additionally, ρ1Y241 (homologous to β2F200 on loop C, Figure 11A and B) and ρ1Y247 (homolog of β2Y205 on loop C, Figure 11A and B) form a π–π interaction, while ρ1Y241 and ρ1R104 (homolog of α1R67 on loop D, Figure 11A and B) interact via a hydrogen bond. Moreover, ρ1Y102 was not only identified as crucial for binding but also found to play an integral role in the receptor’s gating mechanism, suggesting a dual function in both ligand recognition and channel activation.

Molecular dynamics simulations of the RDL receptor model, based on the structure of the GluCl receptor, identified key residues involved in interactions with the GABA molecule: R111 (homologous to α1R67 on loop D, Figure 11A and C) and E204 (homologous to β2E155 on loop B, Figure 11A and B) form salt bridges, while Y109 (homologous to α1F65 on loop D, Figure 11A and B), F206 (homologous to β2Y157 on loop B, Figure 11A and B), and Y254 (homologous to β2Y205 on loop C, Figure 11A and B) engage in cation–π interactions (Comitani et al. Reference Comitani, Cohen, Ashby, Botten, Lummis and Molteni2014). Naffaa et al. (Reference Naffaa, Absalom, Raja Solomon, Chebib, Hibbs and Hanrahan2016) explored the role of the T244 residue in the ρ-type receptor (homologous to β2T202 on loop C, Figure 11A and B) by introducing mutations. Substitution with alanine or cysteine resulted in minimal GABA sensitivity, while substitution with serine significantly reduced this sensitivity. Using molecular modeling, the authors proposed that a hydrogen bond between GABA and the hydroxyl group of ρ1T244 is critical for ligand binding and contributes to the movement of loop C, which closes the binding site. The binding process was further investigated using funnel-metadynamics simulations and the RDL receptor model of the WT and two mutants: R111A (homolog of α1R67 on loop D, Figure 11A and C) and E204A (homolog of β2E155 on loop B, Figure 11A and B) were considered (Comitani et al. Reference Comitani, Limongelli and Molteni2016). The authors revealed that in both mutants, the free energy well present in the WT receptor-binding side, was absent, and loop C exhibited significant movement upon ligand binding. The roles of residues ρ1S168 (homolog of α1T130 on loop E, Figure 11A and C) and ρ1S243 (homolog of β2S201 on loop C, Figure 11A and B) were examined by Naffaa et al. (Reference Naffaa, Hibbs, Chebib and Hanrahan2022) using electrophysiological and computational methods. Mutations in these residues altered the dose–response relationships of GABA and other ligands, leading the authors to propose that these residues engage in a set of state-dependent interactions with other ECD residues.

Domain interfaces

The region between the ECD and TMD is known as the domain interface. It is a relatively small structure, but its impact on receptor function is critical, as it forms a link between the ligand-binding sites in the ECD and the channel gate in the TMD (Figures 1A and D and 12A). There are four main contact areas between the ECD and TMD. These main interactions can be categorized into intrasubunit and intersubunit sites. Among the intrasubunit interactions, there is a covalent linker that connects the β10 strand of the ECD to the M1 helix in the TMD (Figure 12A and B). Additionally, two non-covalent interaction sites are present: the first between the β6–β7 loop (Cys-loop, Figure 12A and C) and the β10-M1 linker, which also involves the M2-M3 loop (Figure 12A and D); and the second between the β1-β2 loop (loop 2, Figure 12A and B) and the M2-M3 loop. For the intersubunit interactions between the TMD and ECD, these occur between the M2-M3 loop of the principal subunit and the β10-M1 linker region of the adjacent complementary subunit (Figure 12A). The domain interface may also include additional structures besides the mentioned ones. Namely, one from ECD – the β8-β9 loop (F loop, Figure 1D) and one from the TMD side – C-terminal from M4 (see Bartos et al. Reference Bartos, Corradi and Bouzat2009) as a part of the general domain interface for all Cys-loop receptor superfamily members. All these structures, except for the C-terminal M4, have been studied in the context of GABAA receptors. Loop F described in the section dedicated to the ECD is primarily considered part of the ligand-binding structure rather than the domain interface. Thus, in this section, we will focus on the Cys-loop, loop 2, β10-M1 linker, and M2-M3 loop, along with their interactions, to cover the topic of the domain interface specific to GABAA receptors.

Figure 12. Domain interface of GABAAR. (A) α1β2γ2receptor structure viewed from the plane of the membrane. There are multiple areas of possible interdomain interactions: covalent connections via the β10-M1 linker and non-covalent ones via: β10-M1 linker and loop 2 (between β-strands 1 and 2), Cys-loop (between β-strands 6 and 7) and M2-M3 loop (between α-helices M2 and M3), as well as between Cys-loop and β10-M1 linker. (B) Although the β10-M1 linker and loop 2 are relatively distant from each other, some charged residue functional groups point toward each other, enabling electrostatic interactions (β2R216 and β2E52). Residues located at the bottom of the loop (β2V53/α1H56) are oriented toward the M2-M3 loop. (C) The Cys-loop is located above the M2-M3 loop and is almost parallel to the β10-M1 linker, enabling interactions with both structures. (D) The M2-M3 loop is placed below both Cys-loop and loop 2, making its apical residues (e.g., β2K279 and β2P273) crucial for interdomain interactions.

Loop 2

Loop 2 (also referred to as the β1-β2 loop, Figures 1D and 12A and B) has been extensively studied in α1β2γ2 receptors, with mutagenesis experiments targeting both the α1 and β2 subunits. Kash et al. (Reference Kash, Jenkins, Kelley, Trudell and Harrison2003) introduced mutations at three different positions in loop 2: α1D57K, α1D55K, and α1E59K, all located in the complementary α1 subunit (Figure 12B) expressed in HEK293 cells. They observed that the α1D55K mutation, which reversed the electrical charge at this position, completely disrupted receptor function, leaving no significant current response to the agonist (Kash et al. Reference Kash, Jenkins, Kelley, Trudell and Harrison2003). Mutations α1D57K and α1D59K significantly increased the EC50 value (indicating reduced GABA sensitivity), with α1D57K additionally reducing the relative efficiency of P4S-evoked currents compared to GABA, indicating alterations in gating properties (Kash et al. Reference Kash, Jenkins, Kelley, Trudell and Harrison2003).

Interestingly, in the β2 subunit, the mutation β2D56K also caused a significant increase in EC50 and reduced the relative efficiency of P4S compared to GABA, while β2E52K had no significant effect (Kash et al. Reference Kash, Kim, Trudell and Harrison2004a, Figure 12B). Kash et al. (Reference Kash, Jenkins, Kelley, Trudell and Harrison2003) also examined two other residues, β2V53 and its homolog in the α1 subunit, α1H56. Using different substitutions for each residue, they observed that only β2V53E and α1H56E significantly increased the EC50 for GABA (Kash et al. Reference Kash, Kim, Trudell and Harrison2004a). Notably, changes in P4S relative efficiency were observed only for mutations in the principal subunit (β2V53H, β2V53L, and β2V53E), but not for α1H56 mutants (Kash et al. Reference Kash, Kim, Trudell and Harrison2004a). Additionally, the relative efficiency of P4S/GABA was reduced in a negatively charged substitution β2V53E, and it increased in the case of positively charged β2V53H and non-polar β2V53L substitutions (Kash et al. Reference Kash, Kim, Trudell and Harrison2004a). Kinetic modeling of macroscopic (current responses to rapid GABA applications) and single–channel data (receptors expressed in HEK293 cells) by Kaczor et al. (Reference Kaczor, Wolska and Mozrzymas2021) revealed that α1H55(r) (Figure 12B) contributes mainly to late gating transitions and preactivation, with only minor effects on receptor binding. A similar experimental approach for mutations in the β2V53 residue (homologous to α1H55(r)) revealed varying degrees of impairment in gating transitions. While β2V53K and β2V53H resulted in mild disruptions of gating, β2V53A caused more pronounced alterations, with β2V53E showing the most severe gating defects (Kłopotowski et al. Reference Kłopotowski, Michałowski, Gos, Mosiądz, Czyżewska and Mozrzymas2023). Overall, these findings indicate the importance of electrostatic interactions mediated by loop 2 residues, particularly β2V53, with distinct gating alterations depending on the charge of the side chain (Kash et al. Reference Kash, Kim, Trudell and Harrison2004a; Kłopotowski et al. Reference Kłopotowski, Michałowski, Gos, Mosiądz, Czyżewska and Mozrzymas2023. Whereas both α1 and β2 subunits contain residues that are crucial for GABAAR function, all presented results (Kaczor et al. Reference Kaczor, Wolska and Mozrzymas2021; Kash et al. Reference Kash, Jenkins, Kelley, Trudell and Harrison2003, Reference Kash, Kim, Trudell and Harrison2004a, Kłopotowski et al. Reference Kłopotowski, Michałowski, Gos, Mosiądz, Czyżewska and Mozrzymas2023) indicate a greater involvement of the principal subunit (β2) loop 2 compared to complementary one’s (α1) in gating transitions during receptor activation.

β10-M1 linker

Within the interface structure, the β10-M1 linker (Figures 1D and 12A and B) plays a key role as it physically (via covalent bonds) connects the ECD to TMD. Mutational studies highlight its essential role in GABAA receptor function. An increased EC50 for GABA and a decreased relative efficiency of P4S and taurine were reported by Kash et al. (Reference Kash, Dizon, Trudell and Harrison2004b) for the β2K215D mutant (Figure 12B). In contrast, the β2N217D mutation (Figure 12B) had no significant effect on either EC50 or relative efficacy, while β2R216D substitution selectively decreased the relative efficiency of P4S without affecting other agonists (Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b). Further investigations by Mercado (Reference Mercado2006) on α1β2 GABAAR expressed in oocytes have explored the impact of β10-M1 linker mutations in the principal and complementary subunits. Mutations β2R216C and α1R220C(r) (Figure 12B) resulted in nonfunctional receptors (no current response to GABA). Interestingly, while the β2R216C mutant could still bind radioactive muscimol, the corresponding mutant (α1R220C(r)) in the complementary subunit could not, suggesting differences in the binding mechanisms between the principal and complementary subunits (Mercado Reference Mercado2006). Additionally, the effect of 2-(trimethylammonium)ethyl methanethiosulfonate bromide (MTSET+) application to cysteine mutants (namely β2K213C, β2K215C, and α1K219C, Figure 12B) was dependent on GABA presence, showing that both subunits undergo structural rearrangement during receptor activation (Mercado Reference Mercado2006). Moreover, the kinetic modeling of macroscopic currents for mutants ranging from 211st to 222nd residues of the α1 subunit (all in the β10-M1 linker, Figure 12B) was carried out by Keramidas et al. (Reference Keramidas, Kash and Harrison2006) for α1β2γ2 receptors expressed in HEK293 cells. Severe phenotype alteration was observed in the case of mutations that replaced polar side chain residues with aliphatic ones, particularly α1K220A (Keramidas et al. Reference Keramidas, Kash and Harrison2006). Additionally, the substitution α1V212A caused a similar effect to another mutation (α1K220D), with all three mutants (α1V212A, α1K220A, and α1K220D) showing increased EC50 and decreased relative efficiency of partial agonist (Keramidas et al. Reference Keramidas, Kash and Harrison2006). Alterations in macroscopic currents, such as slowed current onset (α1V212A, α1K220A, and α1K220D) and increased rate of deactivation in α1K220A, suggested that mutations within the β10-M1 linker can affect GABAAR gating (Keramidas et al. Reference Keramidas, Kash and Harrison2006). The lack of responsiveness to agonist applications observed in the β2R216C and α1R220C(r) mutants showed that, similar to loop 2, there are residues in the β10-M1 linker that are critically involved in GABAAR function and electrostatic interactions are crucial in the proper function of this linker.

Cys-loop

Studies on the Cys-loop (also referred to as the β6-β7 loop or loop 7, Figures 1D and 12A and C), similar to loop 2, were performed on the α1β2γ2 GABAAR, differentiating between α1 and β2 subunits. Mutations in the complementary subunit, such as α1D149K, resulted in decreased GABA sensitivity and reduced relative efficiency for partial agonists, while α1D145K showed no significant changes (Kash et al. Reference Kash, Jenkins, Kelley, Trudell and Harrison2003; Figure 12C). In the principal subunit, mutations were categorized into two groups: those that increase GABA EC50 and decrease the relative efficiency of P4S/GABA (β2D139K and β2D146K, Figure 12C), and those that did not affect these parameters (β2R141E, β2R142E and β2E147K, Figure 12C; Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b). Although the phenotypes observed with reversed-charge substitutions in the Cys-loop mirrored those observed in loop 2 (affected EC50 and relative efficiency of partial agonist), in the case of the Cys-loop, all mutants remained functional; that is, these substitutions were not critical for receptor function (like α1D55K from the β1-β2 loop, Figure 12B). Furthermore, not all residues were prone to alter receptor function after the charge-reversal substitutions (β2R141E, β2R142E, and β2E147K), suggesting that only a select few side chains in loop 7 participate in electrostatic interactions.

M2-M3 loop

The M2-M3 loop, located at the TMD side of the domain interface (Figures 1D and 12A and D), contrary to loop 2 and loop 7 placed on the ECD side, has been extensively studied not only in the α1 and β2 but also in the γ2 subunit. Bera et al. (Reference Bera, Chatav and Akabas2002) used an oocyte model and reported an increase in GABA EC50 for α1L276C(r), α1P277C(r), α1K278C(r), α1F288C(r), and α1D286C(r) by screening mutants from α1R273(r) to α1I289(r), all of which are from the M2-M3 linker, Figure 12D). An additional application of MTSET+ to cysteine mutants revealed that α1L276C(r) was exposed to the reagent only in the presence of GABA, whereas α1K278C(r) reacted with MTSET+ only in the absence of the agonist. This indicates structural changes that occur upon conformational transitions in the M2-M3 loop related to gating (Bera et al. Reference Bera, Chatav and Akabas2002). In a related study, a significant rise in EC50 (10-fold) and a decrease in the relative efficiency of the partial agonist (5-fold) were reported for α1K279D (Kash et al. Reference Kash, Jenkins, Kelley, Trudell and Harrison2003, Figure 12D). The same authors discussed the involvement of the principal subunit’s M2-M3 loop, showing that the alanine substitution (aliphatic side chain for a cationic) for β2R269 (Figure 12D) had a larger effect on receptor function than charge reversal (β2R269D) (Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b). β2K274 (Figure 12D) mutants exhibited increased EC50 and decreased relative efficiency for P4S (as β2R269A), but in this case, an altered phenotype appeared upon the charge reversal mutation β2K274D (Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b). A similar alteration, akin to that caused by α1K279D, was observed by Topf et al. (Reference Topf, Jenkins, Baron and Harrison2003) (recordings of currents from HEK293 cells evoked by a multichannel perfusion system) in the case of α1L277A (Figure 12D), where the substitution of a non-polar residue (leucine) with alanine revealed that electrostatic interactions within the domain interface are not the sole requirement for proper GABAAR function (Topf et al. Reference Topf, Jenkins, Baron and Harrison2003).

Lysine 279 (Figure 12D) in the β2 subunit was found to be critical for modulation by alkaline pH, with its mutation to β2K279A (in HEK273 cells) markedly increasing the level of spontaneous activity (Wilkins et al. Reference Wilkins, Hosie and Smart2005). Hales et al. (Reference Hales, Deeb, Tang, Bollan, King, Johnson and Connolly2006) mutated a single lysine (K289M) in the γ2 subunit (known to be responsible for familial epilepsy) and homologous residues from the α11K278M(r)) and β22K274M, Figure 12D) subunits. The mutation of the primary subunit, namely β2K274M (Figure 12D), affected receptor expression and GABA EC50 value, while α1K278M(r) (Figure 12D) and γ2K289M decreased the mean open time determined from single-channel recordings in HEK293 cells (Hales et al. Reference Hales, Deeb, Tang, Bollan, King, Johnson and Connolly2006). The M2-M3 loop was also studied for its involvement in the spontaneous activity of the α1β2γ2 receptor. For this purpose, a gain-of-function mutation α1L9’T was used, which significantly increases the unliganded open probability. This allowed Nors et al. (Reference Nors, Gupta and Goldschen-Ohm2021) to examine the impact of some M2-M3 mutants (from α1L276(r) to α1T283(r), Figure 12D) expressed in oocytes. The impaired spontaneous activity of α1V279D(r) and α1V279A(r) mutants, which additionally displayed enhanced BDZ effects, highlighted the contribution of the M2-M3 loop in gating and drug modulation in GABAAR (Nors et al. Reference Nors, Gupta and Goldschen-Ohm2021).

Extensive kinetic analysis of macroscopic and single-channel currents by Brodzki and Mozrzymas (Reference Brodzki and Mozrzymas2022) demonstrated that β2P273 (Figure 12D) substitutions (in α1β2γ2 receptors expressed in HEK293 cells) primarily affect late gating transitions, particularly desensitization. Interestingly, the functioning of the β2P273E mutant was severely disrupted, showing no clear cluster activity, which revealed a significant impact of the M2-M3 loop on the overall gating properties of GABAA receptors (Brodzki and Mozrzymas Reference Brodzki and Mozrzymas2022). The P277 residue in the α1 subunit (Figure 12D, homologous to β2P273, α1β2γ2 receptors expressed in HEK293 cells) studied by Kaczor et al. (Reference Kaczor, Michałowski and Mozrzymas2022) revealed that the effect on receptor gating was much more pronounced than alterations in binding properties, similar to observations from Brodzki and Mozrzymas (Reference Brodzki and Mozrzymas2022) on β2P273 residue role. Interestingly, the α1P277(r) mutations (α1P277A(r), α1P277E(r), α1P277K(r), and α1P277H(r)) affected GABAAR gating to a similar extent (Kaczor et al. Reference Kaczor, Michałowski and Mozrzymas2022), whereas the β2P273 substitutions (β2P273E and β2P273K) showed clear differences related to the electrostatic properties of respective substitution (Brodzki and Mozrzymas Reference Brodzki and Mozrzymas2022). In the case of α1P277(r), we proposed that for the complementary subunit, the steric interactions inside the M2-M3 loop are more important than electrostatic ones (Kaczor et al. Reference Kaczor, Michałowski and Mozrzymas2022). The impact of the M2-M3 loop function appears analogous to that observed for loop 2 and loop 7, with evidence of not only specific cationic/anionic couplings but also for general electrostatic interactions taking place inside the domain interface. Importantly, the overall impact of the β2P273E mutation was markedly greater than that of homologous mutations at the α1P277(r) residue (Brodzki and Mozrzymas Reference Brodzki and Mozrzymas2022; Kaczor et al. Reference Kaczor, Michałowski and Mozrzymas2022), indicating asymmetry in the functioning of the M2-M3 loops located in the primary and complementary subunits.

Interactions between interface structures

The substitution of a single residue offers direct insight into the role of specific amino acids in protein functioning. However, when asking about interactions between two residues, a strategy based on double mutations is often more insightful. In particular, substitutions with cysteine made simultaneously for two residues and subsequent treatment with oxidizing (Cu2+(1,10-phenanthroline)3, Cu:phen) or reducing dithiothreitol (DTT) agent may reveal disulfate bond formation, thereby allowing the assessment of the distance, mobility, and interactions between residues of interest. Double mutants with charge reversal can also be applied to explore the role of specific electrostatic interactions between specific residues. Below, we briefly summarize the results of application of these approaches to explore interactions within major structures of the receptor interface: loop 2, loop 7, M2-M3 loop, and β10-M1 linker (Figure 12A) in the α1β2γ2 receptor dynamics.

Kash et al. (Reference Kash, Jenkins, Kelley, Trudell and Harrison2003) used cysteine mutations in both loop 2 (α1D57C, Figure 12B) and the M2-M3 loop (α1K279C, Figure 12D) to determine whether these two residues are located close enough to effectively interact. Results obtained with the use of Cu:Phen showed that loop 2 and the M2-M3 loop are in close proximity, and the distance between them does not clearly change upon receptor activation (Kash et al. Reference Kash, Jenkins, Kelley, Trudell and Harrison2003). However, this was not confirmed by structural studies – in structures obtained by Kim et al. (Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020) these residues are ~10 Å apart. The charge reversal in the case of α1D57K and α1K279D (double mutant) decreased the value of EC50 for GABA to the WT level, while the EC50 of P4S further increased (compared to the single mutation of α1K279D) (Kash et al. Reference Kash, Jenkins, Kelley, Trudell and Harrison2003). Analogous studies were also performed for the principal β2 subunit, in which double cysteine mutations were introduced: β2D56C and β2K274C (Figure 12B and D), and, as in the case of the aforementioned α1D57C + α1K279C mutations, the proximity of these residues was revealed in both non-activated and activated receptors (Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b). Similar to the α1 subunit, according to structural studies, these residues are relatively distant from each other (~15 Å, according to Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020). As for the charge reversal, none of the examined double mutants (β2E52K + β2K274D, β2D56K + β2K274D) was able to rescue the GABA Ec50 value to the WT level (Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b). Thus, the above-mentioned results for both principal and complementary subunits show that the distance between loop 2 and the M2-M3 loop does not change during receptor activation and that both structures remain close to each other (Kash et al. Reference Kash, Jenkins, Kelley, Trudell and Harrison2003). However, recent structural data (Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020) indicated that the investigated residues of loop 2 (α1D57, β2D56C) and the M2-M3 loop (α1K279, β2K274) are more distant from each other than in the model proposed by Kash et al. (Reference Kash, Jenkins, Kelley, Trudell and Harrison2003), but in agreement with data based on electrophysiological recordings, the distance between them is not dependent on the receptor state (comparison with bicuculline and GABA + Diazepam-bound structures). Recently, Zarkadas et al. (Reference Zarkadas, Pebay-Peyroula, Thompson, Schoehn, Uchański, Steyaert, Chipot, Dehez, Baenziger and Nury2022) used a combination of molecular dynamics, structural data, and functional approaches to demonstrate in AChR that loop 2 and the M2-M3 loop show a ‘scissor-like’ movement in which the distance between these structures does not change, reinforcing the experimental observations and conclusions presented in (Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b). This scenario was helpful in our recent study to explain the extreme functional alteration observed in β2V53E compared to other substitutions (β2V53H, β2V53A, and β2V53K, Figure 12B, Kłopotowski et al. Reference Kłopotowski, Michałowski, Gos, Mosiądz, Czyżewska and Mozrzymas2023). Specifically, the electrostatic interaction between β2V53E (loop 2, Figure 12B) and β2K274 (M2-M3 loop, Figure 12D) was suggested as a shut state stabilizer responsible for the prevalent long-lasting shut conformation observed in the glutamate mutant (Kłopotowski et al. Reference Kłopotowski, Michałowski, Gos, Mosiądz, Czyżewska and Mozrzymas2023). This interaction was further supported by structural data (Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020; Kłopotowski et al. Reference Kłopotowski, Michałowski, Gos, Mosiądz, Czyżewska and Mozrzymas2023), showing the movement of β2V53 over β2P273 toward β2K274 during the transition from the desensitized to the shut state. Interestingly, this interaction was significantly less prominent in the complementary subunit, between α1H55(r) (loop 2) and α1P277(r) and α1K278(r) (M2-M3 loop) (Kaczor et al. Reference Kaczor, Wolska and Mozrzymas2021, Reference Kaczor, Michałowski and Mozrzymas2022).

Studies of double cysteine mutants in loop 2 and the β10-M1 linker in the primary subunit (β2D56C + β2K215C, Figure 12B) revealed that these residues are in close proximity only in the activated receptor (i.e., when Cu:Phen was administered together with GABA) (Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b). This finding is challenging to validate using structural data, as no receptor structure in the active state has been published. In the desensitized or inactive states, the functional groups in these residues are oriented in opposite directions, with their Cα atoms approximately 15 Å apart, precluding any direct interaction. Additionally, none of the double mutants with reversed charges (β2E52K + β2K215D, β2D56K + β2K215D, β2D56K + β2R216D, Figure 12B) was able to restore the GABA EC50 value or relative efficiency for P4S to levels comparable to those in the WT (Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b). These results indicate that during receptor activation, loop 2 and the β10-M1 linker interact. However, similar to the interaction between loop 2 and the M2-M3 loop, there appear to be no electrostatic interactions that would be critical for restoring GABA EC50 and P4S efficiency to WT levels. According to structural studies, the strong electrostatic coupling between loop 2 and the β10-M1 linker would involve β2E52 and β2R216 or α1R221 and α1D55 residue pairs, as their functional groups are oriented toward each other, enabling electrostatic attraction (Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020). Our unpublished results showed that mutations in these residues result in nonfunctional receptors.

The interactions between the Cys-loop (Figure 12C) and the M2-M3 loop (Figure 12D) were studied in both the α1 and β2 subunits. Kash et al. (Reference Kash, Jenkins, Kelley, Trudell and Harrison2003) demonstrated that loop 7 (α1D149) and the M2-M3 loop (α1K279) are in proximity only after treatment with an agonist, implicating the coupling of these residues (and loops) during GABAAR activation. Similar to the case of β2D56C + β2K215C, verification of this result using structural data was not feasible, as in the desensitized or inactive state, the residues are approximately 15 Å apart (Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020). An electrostatic interaction was proven to be important, as the double mutant α1D149K + α1K279D (Figure 12C and D) restored the GABA Ec50 value to levels close to those of the WT (Kash et al. Reference Kash, Jenkins, Kelley, Trudell and Harrison2003). Interestingly, the results obtained for the β2 subunit indicated that the distance between β2D139C and β2K274C, as well as between β2D146C and β2K274C (distance between residues similar to the homologous ones according to Kim et al. (Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020)), did not change in the presence of GABA, indicating that these residues remain close to each other regardless of agonist binding (Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b). In contrast to the α1 subunit, none of the reversely charged mutants between the Cys-loop and the M2-M3 loop in the principal subunit (β2D139K + β2K274D, β2D146K + β2K274D, or β2E147K + β2K274D, Figure 12C and D) were able to restore the WT GABA EC50 value (Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b). Thus, according to electrophysiology data, there is an asymmetry in the functioning of the principal and complementary subunits, as only in the latter (α1) did double mutants for loop 7 and the M2-M3 loop reveal the importance of electrostatic interactions during receptor activation. In general, the discrepancies between structural (e.g., Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020) and electrophysiological data (Kash et al. Reference Kash, Jenkins, Kelley, Trudell and Harrison2003, Reference Kash, Kim, Trudell and Harrison2004ab) regarding interactions involving the M2-M3 loop arise from the fact that in various non-open state structures, the M2-M3 loop is located below loop 2 and the Cys-loop (Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020), whereas in the model proposed by Kash et al. (Reference Kash, Dizon, Trudell and Harrison2004b), the M2-M3 loop is positioned between these two structures.

The interaction between the Cys-loop and the β10-M1 linker was studied using a double mutant strategy exclusively in the β2 subunit (Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b). Double cysteine mutants, β2D139C + β2K215C and β2D146C + β2K215C (Figure 12C and D), were modulated only upon simultaneous application of an oxidizing reagent and GABA, revealing the proximity of these residues only during receptor activation (Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b). However, structural data (Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020) suggest that a direct interaction between β2K215 and β2D139C is rather unlikely, as these residues are separated by the β7 strand and are more than 10 Å apart, whereas the interaction between β2K215 and β2D146C seems plausible. Charge-reversal mutagenesis further supports this, as only the β2D146K + β2K215D mutants restored the wild-type GABA EC50 value (Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b), providing evidence for a strong electrostatic coupling. Additionally, the β2D146K + β2K215D double mutant exhibited a relative efficiency for the partial agonist closest to WT values among other mutants (e.g., β2D139K + β2K215D and β2E147K + β2K215D), though still significantly lower (Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b). It is worth emphasizing that electrostatic coupling between loop 7 and the β10-M1 linker was observed only in the principal subunit upon receptor activation (Kash et al. Reference Kash, Dizon, Trudell and Harrison2004b).

The data presented above regarding the interface domain emphasize the role of electrostatic interactions and the proximity between specific residues in unliganded or GABA-bound GABAARs. Interestingly, while alterations in receptor function in the case of single substitutions mostly depend on the side chain charge, the results from double mutants show that only specific interactions of the aforementioned structures can rescue the double substitution phenotype (to the WT values). Bartos et al. (Reference Bartos, Corradi and Bouzat2009) proposed a diffused mechanism where electrostatic interactions take place across the whole domain interface, not only between two particular structures (as it was examined). The results obtained for different subunits also show differences between α1 and β2 interactions inside the domain interface. Interestingly, the asymmetry between the principal and complementary subunits can be appreciated by the observation that the interactions of loop 7 and the M2-M3 loop in the α1 and β2 subunits have different impacts on receptor functioning. A similar idea, that the contribution of different subunits to signal transduction across the domain interface is not equal, has been suggested for loop 2 and the M2-M3 loop (Brodzki and Mozrzymas Reference Brodzki and Mozrzymas2022; Kaczor et al. Reference Kaczor, Wolska and Mozrzymas2021, Reference Kaczor, Michałowski and Mozrzymas2022; Kash et al. Reference Kash, Kim, Trudell and Harrison2004a; Kłopotowski et al. Reference Kłopotowski, Michałowski, Gos, Mosiądz, Czyżewska and Mozrzymas2023.

Transmembrane domain

Ion pore

The TMD of each GABAA receptor subunit consists of four α-helical segments designated M1-M4, with the M2 segment lining the ion pore. As GABAAR is composed of five subunits, the channel pore comprises five vertically aligned M2 segments, which are highly homologous across receptor subunits. As already mentioned, to facilitate comparison between residues of Cys-loop family channels, a prime number notation is used: it starts from the residue at the cytoplasmic end of the M2 segment, denoted as -2′, and ends at another ring of residues at the extracellular end, labeled as 20’ (Figures 1E and 13, Laverty et al. Reference Laverty, Desai, Uchański, Masiulis, Stec, Malinauskas, Zivanov, Pardon, Steyaert, Miller and Aricescu2019; Miller Reference Miller1989). This notation is particularly useful for describing channel pore structures.

Figure 13. Ion pore of GABAAR. Residues are typically described using x′ notation, starting from the bottom of the pore. The most crucial for receptor function are 9′ residues (forming channel gate, β2L259 and α1L264) and -2′ (making additional constriction site attributed to the desensitized state, β2A248 and α1P253).

Top part

Within the top part of the ion pore, 17′ histidine of the β2 subunit (β2H267, Figure 13) has attracted considerable interest because of its role in the modulation of receptor function by pH level and Zn2+ ions (Krishek et al. Reference Krishek, Moss and Smart1998; Krishek et al. Reference Krishek, Amato, Connolly, Moss and Smart1996). H292A (17′) mutation in murine β3 subunit in homomeric (β3) and heteromeric (α1β3) GABAA receptors caused reduction of Zn2+ inhibitory effect, indicating thus that H292 in TM2 of murine β3 subunit is an important component of a Zn2-binding site on the GABAAR (Wooltorton et al. Reference Wooltorton, Mcdonald, Moss and Smart1997). The 17′ residue of the β subunit has also been implicated in receptor modulation by protons (H+). Wilkins et al. (Reference Wilkins, Hosie and Smart2002) showed that H267(A, E) mutations in the β2 subunit of α1β2 receptors abolished the potentiating effect of H+ (at GABA concentrations between 1 and 30 μM), whereas the β2H267K mutant did not affect proton modulation (compared to WT) but increased the rate of macroscopic desensitization of GABA-induced currents. Further investigation revealed that the β2H267A mutation increased the inhibition of GABA-evoked currents at pH 5.4 for the α1β2H267Aγ2 receptor to a level similar to that observed for α1β2H267A (Wilkins et al. Reference Wilkins, Hosie and Smart2005). Horenstein et al. (Reference Horenstein, Wagner, Czajkowski and Akabas2001), using two-electrode voltage-clamp on oocytes and disulfide bond trapping, examined the proximity and mobility of cysteine substitutions at the 20′, 17′, and 6′ positions of the α1 and β1 subunits (Figure 13) in resting and activated receptors. The authors found that, with or without an agonist, disulfide bonds were formed between α1N275C and β1E270C (20′), as well as between α1S272C and β1H267C (17′), indicating significant mobility of this part of the ion pore. The possible role of residue β2E270 (20′), located above β2H267, in the interaction with chloride ions was proposed in a molecular dynamics study of α1β2γ2 receptors (Michałowski et al. Reference Michałowski, Kraszewski and Mozrzymas2017). Structural studies (Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020; Masiulis et al. Reference Masiulis, Desai, Uchański, Serna Martin, Laverty, Karia, Malinauskas, Zivanov, Pardon, Kotecha, Steyaert, Miller and Aricescu2019) showed that the top part of the ion pore is wide even in the non-permeable states of the receptor (desensitized or antagonist-bound), indicating that, in contrast to the other regions of the pore (middle and bottom), no gate of any kind is present there. A recent study reported the crucial role of the β2H267 residue and its interaction with β2E270 in the activation of proton-gated chloride channels (Garifulina et al. Reference Garifulina, Friesacher, Stadler, Zangerl-Plessl, Ernst, Stary-Weinzinger, Willam and Hering2022). The authors discovered that homomers of β2 and β3 subunits formed chloride channels that were directly activated by protons. However, due to the constitutive activity of β1 homomers, the direct activation by H+ could not be reliably observed. To address this, the β1S265N (15′) mutation was examined, as it has been previously shown to reduce the spontaneous activity of homomeric β1 channels (Miko et al. Reference Miko, Werby, Sun, Healey and Zhang2004). Using a two-electrode voltage-clamp technique on oocytes, Garifulina et al. (Reference Garifulina, Friesacher, Stadler, Zangerl-Plessl, Ernst, Stary-Weinzinger, Willam and Hering2022) found that proton-activated chloride currents mediated by homomeric channels (β1S265N, β2, β3) exhibit slower activation and desensitization kinetics in comparison to αβ receptors. The β2H267A mutation completely abolished the proton sensitivity of these homomeric channels. Molecular dynamics simulations further revealed that protonation of the 17′ histidine induced hydrogen bonding with the 20′ glutamate of the adjacent subunit. Further evidence of β2H267–β2E270 interaction has been provided by homologous substitution (ρ1G331H–ρ1A334E) in homomeric ρ1 receptors. While ρ1 homomers are typically insensitive to protons, the incorporation of histidine and glutamate at positions 331 and 334 rendered the channel proton-sensitive, whereas a single mutation, ρ1G331H, was insufficient to induce this sensitivity (Garifulina et al. Reference Garifulina, Friesacher, Stadler, Zangerl-Plessl, Ernst, Stary-Weinzinger, Willam and Hering2022).

Middle part

The residues located at the 9′ and 6′ positions (Figure 13) in the middle part of the ion pore play important roles in ion permeation and GABAA receptor gating. Gurley et al. (Reference Gurley, Amin, Ross, Weiss and White1995) found that mutations at the 6′ residues in the α1β2γ2 receptor cause insensitivity to the channel blocker PTX. In a subsequent study by this group, it was reported that a single mutation of T6’ to phenylalanine in any subunit of the α1β2γ2 receptor was sufficient to confer this insensitivity (Sedelnikova et al. Reference Sedelnikova, Erkkila, Harris, Zakharkin and Weiss2006). These results underscored thus the key role of the 6′ residue in PTX binding. Studies by Tierney et al. (Reference Tierney, Birnir, Pillai, Clements, Howitt, Cox and Gage1996) demonstrated that mutations at the L9′ position, such as α1L264T or β2L259T, in α1β2 receptors caused spontaneous activity, indicating an impact on receptor gating. In the α1β1 receptor with the α1L264T mutation (9′), GABA-evoked currents were characterized by slower rise times and decay kinetics compared to native receptors. In contrast, receptors with mutations at the 9′ position of the β2 subunit or in both α1 and β1 subunits were permanently open, suggesting that this residue is key in controlling the channel activation gate. Importantly, other mutations of this conserved M2 leucine in different GABAAR subunits also led to spontaneously active receptors. Specifically, mutations such as α1L263S, β2L259S, or γ2L274S (all at 9′) shifted the GABA dose-response relationship to the left (Chang and Weiss Reference Chang and Weiss1999). The α1β2γ2L receptors containing one mutated subunit at 9′ (of any type) exhibited spontaneous activity, which was blocked by PTX (an open-channel blocker) and the competitive antagonist bicuculline (Chang and Weiss Reference Chang and Weiss1999).

High-resolution experiments using ultrafast agonist delivery provided detailed insights into the kinetic properties of evoked currents mediated by α1L264Tβ2γ2 receptors (Scheller and Forman Reference Scheller and Forman2002). Receptors with the α1L9′T substitution exhibited higher GABA sensitivity, spontaneous activity, and PTX sensitivity. Furthermore, currents mediated by these mutated receptors were characterized by extremely slow desensitization. Deactivation of currents mediated by wild-type receptors was described with two time constants, and this process slowed down after the induction of partial macroscopic desensitization (by prolonged agonist exposure), thus showing clear desensitization-deactivation coupling. In contrast, α1L264Tβ2γ2 receptors showed deactivation characterized by a single time constant, which was uncoupled from desensitization. Based on the modeling of macroscopic currents, the α1L264T mutation was found to increase the gating efficacy of GABAARs by slowing the channel closing rate (Scheller and Forman Reference Scheller and Forman2002).

Horenstein et al. (Reference Horenstein, Wagner, Czajkowski and Akabas2001) analyzed the impact of mutating various residues to cysteines of the TM2 segment of the α1 and β2 subunits. They found that, with or without GABA, disulfide bonds formed near the extracellular end of TM2 at the 20′ and 17′positions, indicating that this part of TM2 is particularly mobile and/or flexible. In contrast, the disulfide bond between α1T261C/β1T256C residues (6′) is formed only in the presence of GABA and locks the channel in the open state (Horenstein et al. Reference Horenstein, Wagner, Czajkowski and Akabas2001). Their results suggest that receptor activation must involve an asymmetric rotation of adjacent subunits toward each other.

The spontaneous activity of α1β2259Sγ2 receptors, with 9′ residue mutated, was confirmed by Mortensen et al. (Reference Mortensen, Wafford, Wingrove and Ebert2003), who demonstrated a correlation between partial agonist potency and the level of receptor spontaneous activity. Downing et al. (Reference Downing, Lee, Farb and Gibbs2005) further showed that the spontaneous activity of α1L263Sβ3γ2(r) receptors could be potentiated by various allosteric modulators. In addition to 9′ residues, also mutations at the 6′ position (Figure 13) led to spontaneous activity; for instance, the β1T256C mutation caused spontaneous activity following the formation of an intersubunit disulfide bond between adjacent β1 subunits in α1β1 receptors (Rosen et al. Reference Rosen, Bali, Horenstein and Akabas2007). Importantly, molecular dynamics simulations by Xie et al. (Reference Xie, Wang, Sha and Cheng2013) confirmed the presence of an energy barrier for chloride ion permeation at the 9′ and 6′ residue levels

Further studies by Patel et al. (Reference Patel, Mortensen and Smart2014) focused on α4β3δ receptors containing δ subunit and showed that L9′ mutations in the α4 (L297S), β3 (L284S), or δ (L288S) subunits increased GABA sensitivity. Current responses to GABA mediated by receptors with any mutated L9′ exhibited prolonged deactivation kinetics compared with the wild-type receptor (α4β3δ). Moreover, all receptors with a serine mutation at 9′ in any subunit showed spontaneous activity, which was blocked by PTX (Patel et al. Reference Patel, Mortensen and Smart2014). In contrast to the effect observed in α, β, γ, and δ subunits, Germann et al. (Reference Germann, Burbridge, Pierce and Akk2022) found that a mutation at the 9′ position of the ε subunit (L299S) in α1β2ε receptors significantly reduced constitutive activity and slightly shifted the dose–response curve to the right. Moreover, mutation of the 6′ residue to phenylalanine in the ε subunit (S299F), which is known to mediate receptor insensitivity to PTX, increased receptor sensitivity to this channel blocker (Germann et al. Reference Germann, Burbridge, Pierce and Akk2022). Finally, structural studies (Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020; Masiulis et al. Reference Masiulis, Desai, Uchański, Serna Martin, Laverty, Karia, Malinauskas, Zivanov, Pardon, Kotecha, Steyaert, Miller and Aricescu2019) confirmed that the channel gate forms at the 9′ residue level (α1L264, β2L259, and γ2L274) with pore constriction made of a conserved leucine ring in receptors bound to bicuculline or PTX.

Bottom part

Similar to some residues located in the middle segment of the ion pore, mutations in the bottom regions of the pore have also been found to cause spontaneous activity. Ueno et al. (Reference Ueno, Lin, Nikolaeva, Trudell, John Mihic, Adron Harris and Harrison2000) showed that the α2V257W mutation at the 2′ position (Figure 13) induces channel opening, enhances sensitivity to GABA, and renders the receptor insensitive to PTX. Furthermore, this mutation was shown to abolish the sensitivity to pregnanolone sulfate (PS) (Akk et al. Reference Akk, Bracamontes and Steinbach2001; Sachidanandan and Bera Reference Sachidanandan and Bera2015). Interestingly, receptors with mutations at the 2′ residues in the β or γ subunits remained susceptible to modulation by this neurosteroid.

A large number of mutations in the M2 segment of GABAAR are associated with epilepsy (for review see Hernandez and Macdonald Reference Hernandez and Macdonald2019). One severe form of epilepsy is Dravet syndrome, which is predominantly associated with a mutation in the SCN1A gene encoding the α1 subunit of the sodium channel. Recently, a pore-lining mutation at the -2′ position in the γ2 subunit (γ2P302L) has also been associated with Dravet syndrome. The γ2P302L substitution resulted in a 90% reduction in current amplitude and a rightward shift in the dose-response curve. Currents mediated by α1β3γ2P302L receptors were characterized by slower rise times, increased desensitization and faster deactivation. Moreover, single-channel recordings further revealed that the γ2P302L mutation increased the frequency of low-conductance openings, reduced the occurrence of the main-conductance openings and decreased the open probability. Measurements of mIPSCs in neurons overexpressing mutated γ2 subunit revealed reduced current amplitude, slower rise times, and prolonged decay kinetics. Collectively, the γ2P302L mutation alters gating transitions, leading to increased neuronal excitability, a key factor in the pathology of Dravet syndrome (Hernandez et al. Reference Hernandez, Kong, Hu, Zhang, Shen, Jackson, Liu, Jiang and Macdonald2017).

Role of the TMD in desensitization

In addition to the ion pore constriction located in the middle of the TMD at the 9′ residue level (Figures 1E and 13), which serves as the channel gate, another constriction has been found at the bottom of the pore (-2′, Figures 1E and 13) and has been proposed to function as the desensitization gate. In order to elucidate the mechanism underlying this process, Gielen et al. (Reference Gielen, Thomas and Smart2015) compared the desensitization profiles of α1β2 (which exhibits high desensitization) and ρ1 (which shows weak desensitization) GABAARs and their chimeras. Their findings revealed that incorporation of the bottom part of the α1 M3 helix into ρ1 homomers resulted in receptors showing profound desensitization. Moreover, additional introduction of the β2 M1-M2 intracellular loop into these chimeras reversed the response kinetics toward that of WT ρ1 homomers. This indicates the importance of interactions between the α and β subunits in shaping desensitization kinetics. In addition, in a reverse experiment, incorporating the ρ1 M1-M2 intracellular loop into the β2 subunit of α1β2 receptors resulted in highly enhanced desensitization, even stronger that in case of the WT. Additionally, inserting the bottom part of the ρ1 M3 helix into all subunits of α1β2 receptors increased desensitization with respect to α1β2 WT. However, the simultaneous addition of both the ρ1 M1-M2 loop (to the β2 subunit) and the bottom M3 region (to both β2 and α1 subunits) had a minimal effect on desensitization. These results indicate the role of the bottom TMD in desensitization and its complex mechanisms, as the observed effects were not additive. To further locate the molecular mechanism of this process, the authors examined receptors with double mutations instead of larger TMD fragment insertions. Mutants of the conserved M3 residues α1L300V(r)β2L296V and α1V296L(r)β2V292L showed decreased desensitization, while mutations α1N307(S,V)(r)β2N303(S,V) enhanced desensitization. Additionally, mutations in the M2 helix, α1G258(V, A)(r)β2G254(A,V) (located at the level of M3’s α1L300(r) and β2L296) and α1V251(I,F)(r)β2S247I (located at the level of M3’s α1N307(r) and β2N303), also increased desensitization (Figure 14). These data, together with results for other pLGICs, enabled Gielen and Corringer (Reference Gielen and Corringer2018) to propose a complex mechanism of receptor transitions into the desensitized state, indicating a key role of the -2′ residue as the second channel gate responsible for desensitization. In another study, Gielen et al. (Reference Gielen, Barilone and Corringer2020) used two-electrode voltage-clamp recordings on oocytes expressing concatemeric GABAAR construct to precisely assess the roles of each subunit in desensitization. They used mutations at the 5′ residue of the M3 helix (α1N307V(r), β2N303V, and γ2H318V, Figure 14) to enhance desensitization and observed that desensitization time constants generally decreased with an increased number of introduced mutations. However, the effects of the mutations were not additive, and some mutations had stronger effects on receptor kinetics than others. Interestingly, combining mutations produced synergistic and asymmetric effects (with respect to mutated subunits) pointing to a non-concerted mechanism, with the fast component of desensitization mediated by neighboring α1 and γ2 subunits, and the slow component by β2 subunits adjacent to γ. Further evidence supporting the role of the -2′ residue as the desensitization gate came from structural studies that depicted differences in the ion pore profile between the shut state (caused by antagonist or pore blocker binding) and the desensitized state (induced by agonist and/or positive modulators). In the desensitized state, the pore radius at the 9′ level was wide enough to allow chloride ion conduction, while a constriction (~3 Å) was observed at the -2′ residue level (α1P253, β2A248, and γ2P263, Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020; Masiulis et al. Reference Masiulis, Desai, Uchański, Serna Martin, Laverty, Karia, Malinauskas, Zivanov, Pardon, Kotecha, Steyaert, Miller and Aricescu2019). These findings indicate that in addition to the channel gate at 9′ residue, yet another one is located at the bottom of the TMD that governs receptor closure during desensitization. Our group conducted a detailed analysis of currents in HEK293 cells expressing α1β2γ2 receptors with point mutations in the β2G254V, α1G258V(r), α1L300V(r), and β2L296V residues (Terejko et al. Reference Terejko, Michałowski, Iżykowska, Dominik, Brzóstowicz and Mozrzymas2021b). Notably, our kinetic modeling based on macroscopic and single-channel recordings demonstrated that these mutations not only influenced desensitization but also affected channel opening/closing transitions. This highlights the dual role of these residues in modulating both desensitization and the broader gating behavior of the GABAA receptor

Figure 14. Residues at the bottom of the TMD involved in the desensitization transition. The residues important for the receptor’s desensitization transition are located within helical bundles at the bottom of the TMD. In addition, the -2′ residue (β2A248 and α1P253, pore lining residues, see Figure 13) forms a desensitization-dependent constriction.

Modulatory binding sites in the upper region of the TMD

The subunit interface of the upper part of the TMD is the receptor region, which is particularly interesting due to its involvement in mediating a variety of modulatory actions. The binding site located in this area is formed by the upper portions of the M2 and M3 helices of the principal subunit, which are covered by the M2-M3 loop, along with the M1 helix from the complementary subunit. Within this region of the macromolecule, cavities housing various binding sites are present (Figures 1C and E and 15A). Interestingly, in contrast to the orthosteric binding sites located at the β+ interfaces (of the ECD), almost all interfaces in the aforementioned region at the top of the TMD in the αβγ receptor assembly form functional binding sites for various ligands, including general anesthetics, BDZs (Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020; Zhu et al. Reference Zhu, Sridhar, Teng, Howard, Lindahl and Hibbs2022), and neurosteroids (NSs, no direct structural evidence, but postulated according to various electrophysiological experiments, e.g., Sugasawa et al. Reference Sugasawa, Cheng, Bracamontes, Chen, Wang, Germann, Pierce, Senneff, Krishnan, Reichert, Covey, Akk and Evers2020). NSs have a canonical binding site located at the bottom of the TMD, with additional binding sites in the upper TMD; for clarity, they will be described together in the subsequent chapter.

Binding site at the β/α interface

The analogy to the orthosteric binding site at the β/α interface in the ECD has inspired several studies aimed at characterizing the β/α interface in the upper part of the TMD. Key residues located at the helices forming the binding cavity for general anesthetics have been determined (Bali and Akabas Reference Bali and Akabas2004; Chiara et al. Reference Chiara, Dostalova, Jayakar, Zhou, Miller and Cohen2012; Jayakar et al. Reference Jayakar, Zhou, Chiara, Dostalova, Savechenkov, Bruzik, Dailey, Miller, Eckenhoff and Cohen2014; Li et al. Reference Li, Chiara, Sawyer, Husain, Olsen and Cohen2006, Reference Li, Chiara, Cohen and Olsen2010; Stewart, Hotta, Desai, et al., Reference Stewart, Hotta, Desai and Forman2013a; Stewart, Hotta, Li, et al., Reference Stewart, Hotta, Li, Desai, Chiara, Olsen and Forman2013b): β2/3M286 and β3V290 on the M3 helix (principal side) and α1L232, α1M236, α1T237, and α1I239 on the M1 helix (complementary side). Some of these residues are illustrated in Figure 15A. To further examine the role of specific residues at the M3 and M1 helices as an entrance to the binding cavity, several cysteine mutants at the respective M3 and M1 residues were analyzed, clearly demonstrating the potential for disulfide bond formation between them (Bali et al. Reference Bali, Jansen and Akabas2009; Borghese et al. Reference Borghese, Hicks, Lapid, Trudell and Harris2014), indicating thus their close interaction. In addition, mutagenesis studies revealed that mutations of β2M286 and α1M236 (Figure 15A) to tryptophan produced effects on receptor function similar to those of modulation by general anesthetics (Krasowski et al. Reference Krasowski, Nishikawa, Nikolaeva, Lin and Harrison2001; Stewart et al. Reference Stewart, Desai, Cheng, Liu and Forman2008).

Figure 15. TMD-binding sites of the GABAAR: α1β2γ2 receptor structure viewed from the plane of the membrane. (A): Upper TMD-binding site. The cavity between α-helices M2 and M3 (principal subunit side) and M1 (complementary subunit side) is marked in orange. Crucial for modulatory action, β2N265 is located in the M2 helix behind β2M286 (M3). At the same level of the complementary subunit, α1L232 is present. (B) PAM neurosteroid-binding sites: intersubunit at the bottom, and intrasubunit at the top of the TMD. Important for neurosteroid function, α1Q242 is located roughly in the middle of the M1 helix (top of the intersubunit side), and on the opposite side of the cavity, β2Y304 and α1W246 are located (bottom of the site). (C) NAM neurosteroid-binding site at the bottom of the α subunit.

The M2 helix of the β subunit lines the inner side of the cavity (Figure 15A) and its N265 residue (present in β2 and β3, but not in the β1 subunit, Figure 15A) have been identified as key players in the action of etomidate. Substituting this residue with serine (present in the β1 subunit) reduced the modulatory effect of etomidate (Belelli et al. Reference Belelli, Lambert, Peters, Wafford and Whiting1997), whereas substitution with methionine eliminated the modulation by etomidate and reduced the effect of propofol (Siegwart et al. Reference Siegwart, Jurd and Rudolph2002, Reference Siegwart, Krähenbühl, Lambert and Rudolph2003). Later research showed that this M2 residue plays an important role in both modulator binding (affinity) and its modulatory action (efficacy). The double mutant β2N265M + α1M236C showed weaker protection against thiol modification upon etomidate application than the single α1M236C mutant (Stewart et al. Reference Stewart, Pierce, Hotta, Stern and Forman2014). In addition, another double mutant β2N265M + α1M236W exhibited a reduced level of mimicking the etomidate effect compared to the single mutant (α1M236W), indicating the role of β2N265 in modulation (Stewart et al. Reference Stewart, Pierce, Hotta, Stern and Forman2014).

Binding sites at other subunit interfaces

Extensive research has targeted other upper TMD subunit interfaces. Photolabeling studies (Yip et al. Reference Yip, Chen, Edge, Smith, Dickinson, Hohenester, Townsend, Fuchs, Sieghart, Evers and Franks2013) and molecular modeling analysis (Franks Reference Franks2015) indicated that propofol could also bind to the intrasubunit cavity of the β3 subunit. This binding site is located slightly higher (near the ECD/TMD interface) compared to the cavities at the β/α interfaces, with key roles for the β3H267 residue (photolabeled site) and β3F221 (mutation to tryptophan, similar to some residues at the β/α binding sites, resulted in spontaneous activity and eliminated modulation by propofol). This intrasubunit binding site was further investigated using patch-clamp electrophysiology on β3 homomeric receptors (Eaton et al. Reference Eaton, Cao, Chen, Franks, Evers and Akk2015). The authors identified the following residues as important for propofol action at this site in β3 homomeric assembly: β3Y143, β3F221, β3Q224, and β3T266. In a subsequent study on α1β3 heteromeric GABAARs, Eaton et al. (Reference Eaton, Germann, Arora, Cao, Gao, Shin, Wu, Chiara, Cohen, Steinbach, Evers and Akk2016) summarized possible binding sites for propofol and key residues for modulation: β/α interface with β3M286, α/β interface with β3Y143, β3F221, and β3Q224, and β/β interface with the same residues as in the α/β interface. Another photolabeling study indicated that photoreactive analogs of general anesthetics can bind to intersubunit binding sites, where β subunit plays complementary role, specifically at the α/β and γ/β interfaces in the αβγ assembly type (Jayakar et al. Reference Jayakar, Zhou, Chiara, Dostalova, Savechenkov, Bruzik, Dailey, Miller, Eckenhoff and Cohen2014), with varying affinities. The photolabeled residue in this case was β3M227, a homolog of α1M236 located at the M1 helix.

In a study by the Sigel group, a set of receptor mutants was examined (in α1β2γ2 assembly): β2N265I (M2 helix, inner binding cavity, Figure 15A) and homologous point mutants: α2S269I and γ2S280I (Maldifassi et al. Reference Maldifassi, Baur and Sigel2016). They used two-electrode voltage-clamp recordings on Xenopus oocytes expressing these receptor types to assess the potentiation extent of GABA-evoked current by propofol, etomidate, pentobarbital, and low concentrations of diazepam or tetrahydrodeoxycorticosterone (THDOC). Triple mutants abolished potentiation by anesthetics and pentobarbital, but not by diazepam or THDOC (which do not bind to these sites at low concentrations). In the case of single point mutants, mutations in the β2 subunits abolished potentiation by propofol and etomidate, confirming previous data. Conversely, mutations in the γ2 subunit either lowered (for propofol) or increased (for etomidate) the potentiation, indicating the involvement of this subunit in modulation by these drugs. Interestingly, all single mutations decreased current potentiation by pentobarbital, suggesting that pentobarbital may bind to multiple interfaces and/or exert allosteric effects. In addition, the authors showed (using concatenated receptors) that etomidate (but not propofol) binding sites have different properties, and only the first β subunit (neighboring γ subunit) is important for its action. Altogether, these results indicate that the modulation of GABAAR by these compounds relies, to some extent, on allosteric coupling between the interface cavities. Similar work was performed by the Forman group; the set of mutants (to cysteine and tryptophan) at positions α1L232 and α1M236 (M1 helix, binding cavity entrance complementary side, Figure 15A) and their homologs (β3M227, β3L231, γ2L242, and γ2L246) in the α1β3γ2 receptor were examined using voltage-clamp electrophysiology in Xenopus oocytes (Nourmahnad et al. Reference Nourmahnad, Stern, Hotta, Stewart, Ziemba, Szabo and Forman2016). Tryptophan substitutions at position α1M236 and its homologs caused spontaneous channel activity and reduced GABA EC50. Cysteine substitutions demonstrated that etomidate protected both α subunit mutations, while mtFd-mPaB (R-5-allyl-1-methyl-5-(m-trifluoromethyl-diazirinylphenyl) barbituric acid) provided protection for both β subunit mutations against thiol modification. Additionally, propofol protected α1M236C and β3M227C mutants, whereas alphaxalone (a neurosteroid) did not protect against same modification. Interestingly, none of the tested modulators protected the cysteines substituted in the γ subunit. These data clearly indicated an important phenomenon – each subunit interface cavity plays a role in receptor gating (spontaneous activity induced by tryptophan substitution and effect on GABA dose-response relationship), but the α/γ interface seems to be an orphan interface without any compatible ligands. Electrophysiological recordings of currents in oocytes expressing α1β2γ2 receptors with point mutations β2M286W (at the β/α binding site, Figure 15A) and β2γY143W (at the β intrasubunit site) together with Monod–Wyman–Changeux (MWC) kinetic modeling, revealed that propofol binds to at least four distinct sites (four investigated and potentially other different sites) (Shin et al. Reference Shin, Germann, Johnson, Forman, Steinbach and Akk2018). The intersubunit sites (β/α, γ/β, and α/β) were further investigated using competitive photolabeling studies (Jayakar et al. Reference Jayakar, Zhou, Chiara, Jarava-Barrera, Savechenkov, Bruzik, Tortosa, Miller and Cohen2019). The authors found that different photoreactive analogs of general anesthetics bind to all these interfaces, but with different affinities. Most importantly, the etomidate analog demonstrated high-affinity binding to the β/α site and low-affinity binding to the γ/β and α/β sites. In contrast, the propofol analog exhibited high affinity for the β/α site and medium affinity for the γ/β and α/β sites. Conversely, mephobarbital displayed high affinity for the γ/β and α/β interfaces but low affinity for the β/α site. The intrasubunit β site was not examined. In another study, the interplay between different modulators binding in the upper TMD region was examined in concatemeric α1β2γ2 receptors expressed in oocytes (Shin et al. Reference Shin, Germann, Covey, Steinbach and Akk2019). The authors discovered that different neurosteroids (both 5α- and 5β-reduced forms, as well as natural and enantiomeric variants) share a common interaction site, whereas propofol and pentobarbital exhibit partially overlapping binding sites. This topic was further explored in another study by Szabo et al. (Reference Szabo, Nourmahnad, Halpin and Forman2019), who performed electrophysiological recordings of α1β3γ2 receptors expressed in oocytes with point mutations of residues β3N265 (and homologs in other subunits, Figure 15A) and β3M286 (and homologs, Figure 15A) and performed MWC modeling analysis. The obtained results confirmed the binding pattern of etomidate (both β/α), barbital (γ/β and α/β) and propofol (all four sites), but contrary to the proposal of Shin et al. (Reference Shin, Germann, Johnson, Forman, Steinbach and Akk2018), some allosteric coupling between the sites was suspected.

Recent cryo-EM structural imaging and molecular dynamics studies (Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020) have provided important insights into the roles of the respective binding sites in the upper TMD. Several structures of the α1β2γ2 receptor with different bound modulators have been obtained. These structures confirmed the binding of etomidate to both β/α interfaces and the binding of pentobarbital (a barbiturate) to γ/β and α/β interfaces, in agreement with most previous studies. However, propofol was found only in β/α interfaces, whereas most studies by other authors indicated binding of this compound also to other sites (see Table 2). It is important to emphasize that in the case of propofol (but not etomidate), the γ/β and α/β cavities remained expanded and showed the presence of lipids. In addition, no modulators were found in the postulated intrasubunit binding sites in the β2 subunit or in a previously determined as ‘orphan’ site at the α/γ interface. In agreement with previous studies, the residues lining the modulatory binding site were determined as follows for the β/α interface (homologs of key residues for other subunits in parenthesis, Figure 15A): β2D282, β2M286 (α1S310, γ2A291), β2F289 (α1F304, γ2Y294) at the M3 helix; β2N265 (α1S280, γ2S270), β2T262 at the M2 helix; and α1I228 (β2L223), α1L232 (β2M227), α1P233 (β2P228), and α1M236 (β2L231) at the M1 helix. A summary of the determined binding sites is presented in Table 2. Another important structural insight was provided by the work of Zhu et al. (Reference Zhu, Sridhar, Teng, Howard, Lindahl and Hibbs2022), who determined the receptor structures with zolpidem (a positive allosteric modulator) and DMCM (a negative allosteric modulator) bound at the α/γ ECD and β/α TMD interface binding sites. The key residues for these ligands binding within TMD are β2D282, β2M286, β2F289, α1I228, α1L232, and α1M236.

Table 2. Localization of the upper TMD modulatory binding sites

Presence of upper TMD modulators in respective intersubunit cavities according to studies listed in the first column.

Modulatory binding sites at the bottom part of the TMD

Neurosteroid binding in the TMD

TMD has also been the subject of numerous investigations because of its role in neurosteroid modulation. Early studies on chimeric receptors (ECD of RDL receptor and TMD of α1, β2, and γ2 GABA receptor subunits, respectively) indicated the key role of the M1 and M2 helices in modulation by neurosteroids. More specifically, the α1T236I(r) mutation reduced the direct activation by allopregnanolone and THDOC, whereas α1Q241W(r) (Figure 15B) mutant diminished both the direct activation and modulatory effects of these compounds (Hosie et al. Reference Hosie, Wilkins, Da Silva and Smart2006). Because the α1Q241(r) residue is expected to form a hydrogen bond with the hydroxyl group of neurosteroids, additional mutants were examined (Hosie et al. Reference Hosie, Wilkins, Da Silva and Smart2006). The α1Q241(H,N)(r) mutants were similar to the WT, while α1Q241(S,T,M) exhibited slightly reduced potencies of neurosteroid modulation. In contrast, α1Q241(I,L)(r) showed no susceptibility to neurosteroids whatsoever, confirming that the ability of α1Q241(r) to form a hydrogen bond with neurosteroids is crucial. On the basis of homology modeling, studies by Hosie et al. (Reference Hosie, Wilkins, Da Silva and Smart2006) indicated residues that may form contact with the opposite side of the neurosteroid molecule: α1N407(r) (Figure 15B) and α1Y410(r) and showed that mutants α1N407(A,V,D)(r) and α1Y410F(r) have reduced neurosteroids potencies. Similar analysis was performed for the α1T236(r) residue and its mutations (to I, V, and S) affected modulation by neurosteroids. In addition, β2Y284 (its mutation to F affects direct activation by neurosteroids) was found to form a neurosteroid-binding site together with α1T236(r). Altogether, Hosie et al. (Reference Hosie, Wilkins, Da Silva and Smart2006) proposed two binding sites for neurosteroids: the first made of α1Q241(r), α1N407(r), and α1Y410F(r) and the second composed of α1T236(r) and β2Y284, both spanning from the middle of the TMD up to its top part. These studies were continued for different subunit assemblies of the GABAA receptor (Hosie et al. Reference Hosie, Clarke, da Silva and Smart2009). Residues located at M4: α1N407(r) and α1Y410F(r) are highly conserved among receptor subtypes, but α1Q241(r) is specific for α subunits, indicating the key role of this subunit. All receptors in the α xβ3γ2 assembly (where x represents any of the α subunits) with a point mutation at a residue homologous to α1Q241(r) had similar phenotypes. Substitution with histidine resulted in a phenotype comparable to WT, while substitutions with threonine, serine, or methionine reduced the potentiation of GABA-evoked currents at EC10. In contrast, substitutions with leucine or tryptophan abolished potentiation (Hosie et al. Reference Hosie, Clarke, da Silva and Smart2009). These results further underlined the key role of α1Q241(r) and its ability to form hydrogen bonds with neurosteroids. Interesting results were obtained from photolabeling studies, which indicated that the neurosteroid analogs labeled the β3F301 residue (Figure 15B) located at the bottom of the M3 helix, indicating thus a different binding site for these modulators (Chen et al. Reference Chen, Manion, Townsend, Reichert, Covey, Steinbach, Sieghart, Fuchs and Evers2012)

PAM intersubunit binding site at the β/α interface and NAM intrasubunit sites in α and β subunits

An important insight into neurosteroid binding was provided by structural research on β35 chimeric GABAARs (Miller et al. Reference Miller, Scott, Masiulis, De Colibus, Pardon, Steyaert and Aricescu2017). The structure of these receptors (homomeric, β3 ECD and α5 TMD, α5T287K background mutation to reduce spontaneous activity) was resolved with bound pregnanolone and in the apo state. A neurosteroid was found between the two subunits (M3 helix of the principal and M1 of the complementary subunit) in the bottom part of the TMD. The α5Q245 residue (Figure 15B) formed a hydrogen bond with the hydroxyl group of the neurosteroid. The A ring of the modulator was located between α5I242 and α5W249 of the M1 helix, while the B and D rings were positioned between α5I305 of the M3 helix and α5W249 of the M1 helix. Additionally, the α5V246 residue in the M1 helix lined the deep part of the binding site, and the ketone of the neurosteroid formed a hydrogen bond with α5T309 in the M3 helix (Figure 15B). Mutations of these residues (α5Q245S, α5V246A, α5W249L, α5I305A, and α5T309A) reduced receptor sensitivity to pregnanolone, further indicating that the binding site containing α5Q245 (α1Q242) does not span upwardly, but rather toward the bottom of the TMD. To examine the role of the binding site residues located at the principal subunit in αβγ assembly, the modulation of α1β3L297Aγ2 (corresponding to α5I305) and α1β3LF301Aγ25T309, Figure 15B) receptors by neurosteroids was assessed. Only the α1β3L297Aγ2 mutation showed a smaller effect on modulation by pregnanolone and allopregnanolone compared to the WT, indicating a more important role of the α subunit. Docking showed that pregnanolone, allopregnanolone, and epipregnanolone bind to the same site at the β/α interface, but the epipregnanolone hydroxyl group is oriented upward and cannot coordinate with α5Q245 (Figure 15B), as in the case of other compounds. Miller et al. (Reference Miller, Scott, Masiulis, De Colibus, Pardon, Steyaert and Aricescu2017) compared also the apo (without ligand) and pregnanalone bound structures and observed that in the neurosteroid bound state the ion pore widens at -2′ residue level and loop M2-M3 moves inward. In addition, other chimeric structures (GLIC ECD and GABAAR α1 TMD, α1G258V(r) to enhance purification and increase desensitization) with THDOC, PS and without any modulators bound were obtained by Laverty et al. (Reference Laverty, Thomas, Field, Andersen, Gold, Biggin, Gielen and Smart2017). Similar to the allopregnanolone in β35 chimeric GABAAR, the THDOC hydroxyl group formed a hydrogen bond with α1Q241(r) (Figure 15B), with rings C and D aligned parallel to α1W245 (Figure 15B), and the neurosteroid ketone group formed a hydrogen bond with α1T305(r) and α1Y308(r) at the bottom of the binding site. Mutations in these residues (α1Q241L(r), α1W245L(r), and α1T305W(r)) abolished the modulatory effects of THDOC. Interestingly, Laverty et al. (Reference Laverty, Thomas, Field, Andersen, Gold, Biggin, Gielen and Smart2017) determined also the binding site of PS to be placed in the bottom TMD within single subunits M3 and M4 helices. The sulfate group of PS was oriented toward α1K390(r), and this molecule probably interacts with α1I391(r), α1A398(r), and α1F399(r) (Figure 15C). Mutations of these residues (α1K390A(r), α1I391C(r), α1A398C(r), and α1F399C(r)) reduced the inhibitory effect of PS. The binding site at the subunit interface was also confirmed by Chen et al. (Reference Chen, Wells, Arjunan, Tillman, Cohen, Xu and Tang2018), who obtained the structure of a chimeric receptor (ELIC’s ECD and α1’s subunit TMD, homomeric) with bound alphaxalone. The neurosteroid molecule was located between the principal side’s α1T306 and complementary side’s α1Q242 (Figure 15B). To further verify the location of the neurosteroid-binding sites, Ziemba et al. (Reference Ziemba, Szabo, Pierce, Haburcak, Stern, Nourmahnad, Halpin and Forman2018) used the SCAM in oocytes expressing α1β3γ2 receptors, with single point mutations aimed at residues located at the M2 and M3 helices (Figure 15A and B: β3T262, β3T266, β3M283, β3Y284, β3M286, β3G287, β3F289, β3V290, β3F293, β3L297, β3E298, and β3F301). Among these, the β3Y284C mutation caused receptors to be nonfunctional, whereas all other mutations, except β3T262C, β3M283C, β3V290C, and F β3301C, affected GABA EC50 and/or efficacy, and β3F289C additionally increased spontaneous activity. Most importantly, the β3F284C, β3F2893C, and β3L297C mutations reduced alphaxalone-induced potentiation, and in the case of β3V290C, β3F293C, β3L297C, and β3F301C, alphaxalone affected protein modification by 4-Chloromercuribenzenesulfonate (pCMBS). These data showed that the location of the neurosteroid-binding site at the bottom TMD determined by structural studies was correct. Another study confirmed that the PS-binding site consists of M3 and M4 helices of a single subunit, and this intrasubunit-binding site is functional not only within α subunits but also in β subunits (Seljeset et al. Reference Seljeset, Bright, Thomas and Smart2018).

In 2023, a number of GABAAR structures with bound neurosteroids were published, shedding light on the details of their binding sites and possible mechanisms of modulation. Sun et al. (Reference Sun, Zhu, Clark and Gouaux2023) obtained the structures of the receptor in α1β3γ2 and novel α2/3β1/2α1β1/2γ2 and α1β1/2α2/3β1/2α1γ2 (mixture of different types of α and β subunit in single receptor, isolated from mouse brain) assemblies. In addition to the unexpected subunit composition, these receptors contained bound endogenous allopregnanolone molecules, indicating their abundance in the mouse brain. Neurosteroids were present in the intersubunit-binding site at the β/α subunit interface between M3 and M1 α-helices, in agreement with previous studies. Other receptor structures (α1β3γ2 stoichiometry) have been presented by Legesse et al. (Reference Legesse, Fan, Teng, Zhuang, Howard, Noviello, Lindahl and Hibbs2023). The authors confirmed localization of the PAM neurosteroid-binding site at the subunit interface (lined by, e.g., α1Q242, α1W246, β2L301, and β2Y304, Figure 15B). Surprisingly, NAM neurosteroids (PS and DHEAS) were found at the ion pore (between the 9′ and -2′ residues, Figure 13), rather than at the previously proposed intrasubunit sites between the M3 and M4 helices (e.g., Laverty et al. Reference Laverty, Thomas, Field, Andersen, Gold, Biggin, Gielen and Smart2017; Mortensen et al. Reference Mortensen, Xu, Shehata, Krall, Ernst, Frølund and Smart2023; Seljeset et al. Reference Seljeset, Bright, Thomas and Smart2018). A summary of NS-binding sites can be found in Table 3.

Table 3. Localization of the TMD neurosteroid-binding site

Epipregnanolone and pregnanolone sulfate are NAMs, allopregnanolone is PAM, but its action via β intrasubunit site is NAM-like.

Additional PAM intrasubunit-binding sites in the top part of the TMD

Photolabeling and patch-clamp studies were performed to verify the location and number of neurosteroid-binding sites (Chen et al. Reference Chen, Bracamontes, Budelier, Germann, Shin, Kathiresan, Qian, Manion, Cheng, Reichert, Akk, Covey and Evers2019). They focused on α1β3 receptors and investigated the action of allopregnanolone and its photolabeling derivatives. Their findings provided evidence that within the receptor, there are three binding sites for non-sulfated steroids. The intersubunit-binding site is located at the bottom of the TMD and is formed by the M3 helix of the principal subunit and the M1 helix of the complementary M1 subunits (photolabeled residues: β3L294 and β3G308, binding site with previously examined α1Q242). In addition, two intrasubunit sites were located – one within the α subunit (photolabeled residues: α1N408A and α1Y415F, with mutants affecting potentiation and direct activation: α1V227W, α1N408A, and α1Y411F – see Figure 15B) and one within the β subunit (photolabeled residues: β3Y442, β3V278-A280, with no mutants affecting the neurosteroid action, Figure 15B). Since there were no mutants of the β3 intrasubunit site that reduced the allopregnanolone effect, this site is probably not functional in the α1β3 receptor assembly. As the location of neurosteroid-binding sites has been confirmed using multiple methods, the next step was to determine the roles of the residues of these sites. Photolabeling of the ELIC-α1 chimeric receptor by allopregnanolone-based reagents confirmed neurosteroid binding to both intra- and intersubunit-binding sites (photolabeled residues: α1N408, α1Y415, and α1Y309, Figure 15B), but surprisingly, mutations α1Q242L, α1Q242W, and α1W246L did not affect the photolabeling efficiency and addition of the allopregnanolone decreased photolabeling, indicating that neurosteroid binding was still possible (Sugasawa et al. Reference Sugasawa, Bracamontes, Krishnan, Covey, Reichert, Akk, Chen, Tang, Evers and Cheng2019). In addition, mutations decreased the thermal stabilization of the protein induced by neurosteroid (measured by UV absorbance intensity in size exclusion chromatography), which is proposed to be correlated with the modulatory effect of the neurosteroid. The authors then checked which residues were photolabeled by the neurosteroid analog (K220) in mutated receptors; in each case (WT and each mutant) α1N408 (Figure 15B) was labeled, but α1Y309 only in the WT receptor. Interestingly, for α1Q242L (Fig. 15B) and α1Q242W mutations, residue α1F298 was labeled and in α1W246L mutant-α1F298 or α1S299 were labeled instead of α1Y309 (Figure 15B). Residue α1F298 is located significantly above α1Y309 what together with previous data (mutations not affecting the photolabeling total efficiency) indicates, that mutations of α1Q242 and α1W246 are not affecting the neurosteroid affinity, but rather the madulator’s orientation within in the binding site. A summary of the proposed binding site locations is presented in Table 3.

Mechanism of neurosteroid action

So far, structural and electrophysiological studies have identified the locations of neurosteroid-binding sites, but the exact mechanisms of action of these compounds remain elusive. Sugasawa et al. (Reference Sugasawa, Cheng, Bracamontes, Chen, Wang, Germann, Pierce, Senneff, Krishnan, Reichert, Covey, Akk and Evers2020) examined different neurosteroids, including allopregnanolone (3α5α), epipregnanolone (3β5β), and photolabeling analogs (KK148 and KK150; for a detailed description, see Jiang et al. Reference Jiang, Shu, Krishnan, Qian, Taylor, Covey, Zorumski and Mennerick2016). Interestingly, only allopregnanolone potentiated GABA-induced responses, and all of them enhanced (KK148 and KK150 at low level) muscimol binding at the orthosteric binding site, indicating at least two independent mechanisms of modulation. In addition, the application of epipregnanolone during the steady-state phase of the current response induced by high GABA concentrations enhanced the extent of macroscopic desensitization. The latter effect was abolished by the α1V256S(r) (Figure 13) mutation, which also prevented the inhibitory action of sulfated neurosteroids (PS). Photolabeling indicated that all tested compounds were binding at intrasubunit sites, and all except epipregnanolone were found in the intersubunit site. Altogether, it has been shown that different neurosteroids bind to different sites and exert a mixture of effects on the receptor, including potentiation of the GABA-induced response, upregulation of macroscopic desensitization, and/or muscimol binding enhancement. To attribute these effects to the respective binding sites, the following mutants were examined: α1Q242L (intersubunit site, Figure 15B), α1N408A, α1Y411F, αV227W (α intrasubunit site, Figure 15B), and β3Y284F (β intrasubunit site, Figure 15B). Allopregnanolone enhanced muscimol binding by interacting with all three sites (mostly the intersubunit), but epipregnanolone and KK148 acted only via both intrasubunit sites. In addition, the 3β5β effect on desensitization was decreased by mutations in both intrasubunit sites. To check whether 3α5α also can enhance desensitization, the background mutation α1Q242L (to abolish the function of intersubunit site) was introduced. Interestingly, enhancement of desensitization by allopregnanolone was observed and was abolished by mutation in the β intrasubunit site. In summary, the authors proposed that allopregnanolone stabilizes the open state of the receptor by binding to the β/α and α intersubunit sites, and the desensitized state by binding to the β intrasubunit site. In contrast, epipregnanolone stabilizes only the desensitized state by binding to both intrasubunit sites. KK148 stabilized the desensitized state via intrasubunit sites and all states equally via the intersubunit site, and K150 stabilized all states equally via all binding sites. Owing to the novel photolabeling agent ([3H]21-pTFDBzoxy-AP), Jayakar et al. (Reference Jayakar, Chiara, Zhou, Wu, Bruzik, Miller and Cohen2020) were able to detect other residues forming the intersubunit-binding site: β3P415, β3L417, β3T418 (at M4 helix bottom), and β3R309 (at M3 helix bottom). Using multiple types of neurosteroids, they confirmed that the presence of a free 3α-OH group is crucial for binding at this interface, and that the affinity is affected by the substitution at C17. In addition, they described various effects of neurosteroid NAMs on the modulation of muscimol binding: epipregnanolone and betaxalone (epimer of alphaxalone) upregulated the binding, epiallopregnanolone and PS had no effect, and dehydroepiandrosterone sulfate (DHEAS) showed a downregulating action. Electrophysiological recordings (on α1β2γ2 receptors expressed in oocytes) allowed for a more accurate description of neurosteroids effects on gating (Pierce et al. Reference Pierce, Germann, Steinbach and Akk2022). Both PS and DHEAS reduced the current amplitude, caused faster and deeper macroscopic desensitization when coapplied with GABA, and enhanced the desensitization extent when applied at the steady-state phase of the response to GABA at saturating concentrations. Interestingly, the application of the NAM neurosteroids during the steady state (deepening the desensitization) showed also its own ‘desensitization’ (shift toward higher current values) and a rebound current once application was finished. In addition, the recovery of the current amplitude after subsequent application of GABA was faster if the first one was a co-application of GABA and NAM neurosteroid. They proposed that the closed state to which receptor transitions occur due to PS or DHEAS modulation is not a canonical desensitized state but a novel short-lived shut state. This was further supported by MWC type of kinetic modeling, in which this additional non-conducting state was incorporated. Modeling also indicated that both PS and DHEAS bind to the same site. Tateiwa et al. (Reference Tateiwa, Chintala, Chen, Wang, Amtashar, Bracamontes, Germann, Pierce, Covey, Akk and Evers2023) showed that the enantiomer of allopregnanolone, but not of pregnanolone, has lower potency due to different binding pose (180° rotation) at the intersubunit site.

Based on multiple GABAAR structures with bound neurosteroids, Sun et al. (Reference Sun, Zhu, Clark and Gouaux2023) proposed that the modulatory mechanism of neurosteroids relies on enforcing the anticlockwise rotation of the TMD, thereby stabilizing the open conformation of the ion pore. Using structural methods, Legesse et al. (Reference Legesse, Fan, Teng, Zhuang, Howard, Noviello, Lindahl and Hibbs2023) noticed that allopregnanolone affected ion pore residues; in the GABA-bound structure (without neurosteroid), β2E270 points toward α1N275, whereas in the neurosteroid-bound structure, it moves toward β2H267 (Figure 13), allowing for increased pore dilation in the top region. Based on molecular dynamics simulations, the authors also proposed that PAM neurosteroids influenced the TMD by lowering the energy barrier of ion movement and ECD via an allosteric mechanism by reducing its spread and making GABA molecules less likely to unbind.

Negative modulation by various PS analogs was examined by Mortensen et al. (Reference Mortensen, Xu, Shehata, Krall, Ernst, Frølund and Smart2023) in HEK293 cells expressing α1β3γ2 receptor. Whereas all investigated compounds retained the general effect of lowering the response amplitude, their effects on macroscopic desensitization differed. Molecular modeling allowed the authors to assess the protonation states of the ligands and to conclude that uncharged analogs had a higher potency to reduce both the amplitude (at steady state) and the macroscopic desensitization time constant. Molecular dynamics simulations indicated that charged analogs interacted with α1F295(r) residue via cation–π interaction and moved closer toward α1F399(r) (Figure 15C). α1F295A and α1F399A(r) mutations reduced the inhibitory potency of the selected charged and uncharged analogs, but increased only the charged compound effects on the desensitization time constant. To further examine the effects of neurosteroids on desensitization, the authors utilized receptor mutants with decreased (α1V296L(r)) and increased (γ2V262F) desensitization, and selected neurosteroids showed varied potencies depending on the type of incorporated mutation, indicating that the uncharged analog is more likely to bind to the receptor in the desensitized state. This finding was further explored using kinetic modeling, which showed that the uncharged analog binds to both GABA-bound closed and desensitized states, whereas the charged analog binds only to the former. In contrast to the results of the study by Mortensen et al. (Reference Mortensen, Xu, Shehata, Krall, Ernst, Frølund and Smart2023), Legesse et al. (Reference Legesse, Fan, Teng, Zhuang, Howard, Noviello, Lindahl and Hibbs2023) located the binding site of PS in the ion pore, indicating, that the mechanism of action of NAM neurosteroids involves channel blocking, rather than negative modulation.

Protein–protein interactions involving synaptic and extrasynaptic GABAARs

Extrasynaptic GABAA receptors

The description of the structural and electrophysiological properties of different GABAA receptors presented in previous chapters provides a solid foundation for the following chapter, which will explore how the molecular architecture of GABAARs and their engagement in protein–protein interactions underpin their functional diversity and adaptability. This diversity is central to their roles in distinct inhibitory signaling modes, both synaptic and extrasynaptic. As already emphasized, GABAARs are greatly diversified owing to the ample choice of subunits assembled to form functional pentamers. This diversity underpins their ability to mediate distinct forms of inhibition. Several GABAA receptor subtypes, such as those containing γ subunits, have been identified as mediators of synaptic (phasic) inhibition characterized by rapid synaptic currents’ time course. Other GABAA receptor subtypes (e.g., those containing δ subunits) are characterized by high agonist affinity and relatively slow kinetics, making them particularly suitable for mediating tonic inhibition by regulating the cell’s input resistance (Walker and Semyanov Reference Walker and Semyanov2008). Interestingly, some GABAARs, such as those containing the α5 subunit, may be involved in both tonic and phasic inhibition, depending on their interactions with various protein partners. The inhibitory regulation of neuronal networks relies thus on two major degrees of freedom: diversity due to distinct receptor stoichiometry and interactions with protein partners, which may determine their cellular localization and/or functional properties. Moreover, extrasynaptic GABAA receptors demonstrate high sensitivity to alcohol and, most importantly, to neurosteroids, such as allopregnanolone (Seljeset et al. Reference Seljeset, Bright, Thomas and Smart2018), which we extensively discussed in the context of receptor modulation. This sensitivity positions these receptors as critical mediators of the effects of endogenous modulators and various therapeutic agents that target GABAergic transmission.

Tonic GABAergic currents are generated by ambient (submicromolar) GABA concentrations that act on high-affinity GABAA receptors containing δ or α5 subunits. Interestingly, the ICD of the α5 subunit of GABAAR interacts directly with the intracellular radixin protein, solely in the extrasynaptic membranes (Loebrich et al. Reference Loebrich, Bähring, Katsuno, Tsukita and Kneussel2006). Radixin consists of a C-terminal actin-binding domain and an N-terminal FERM domain that can bind to the GABAA receptor. These domains are separated by a central α-helical region that, when not phosphorylated, adopts a closed conformation, thereby preventing interactions with partners (Senju and Tsai Reference Senju and Tsai2022). In hippocampal neurons, the phosphorylation of radixin at T564 causes a transition to the open conformation, facilitating its binding to α5GABAA receptors. Consequently, phosphorylated radixin anchors α5GABAA receptors to the actin cytoskeleton, sequestering them in the extrasynaptic region (Hausrat et al. Reference Hausrat, Muhia, Gerrow, Thomas, Hirdes, Tsukita, Heisler, Herich, Dubroqua, Breiden, Feldon, Schwarz, Yee, Smar, Triller and Kneussel2015). Upon neuronal stimulation and depolarization, radixin undergoes dephosphorylation, leading to the dissociation of α5GABAA receptors and their migration to gephyrin-containing inhibitory synapses (Davenport et al. Reference Davenport, Rajappa, Katchan, Taylor, Tsai, Smith, de Jong, Arnold, Lammel and Kramer2021). Consequently, the slower decay kinetics exhibited by inhibitory synaptic currents mediated by α5GABAA receptors retrogradely restrict long-term plasticity at nearby excitatory synapses. This observation underscores the significance of protein–protein interaction, specifically involving a distinct GABAA receptor subunit, in shaping the plasticity dynamics of non-inhibitory synapses.

Protein–protein interactions of synaptic GABAA receptors

Neurotransmitter receptors typically show lateral diffusion, a process that may be temporally interrupted or slowed down by their binding to the structures of the postsynaptic density. Within this region, they are anchored through a variety of interactions with intracellular, membrane, and extracellular proteins. At a significant number of inhibitory synapses, GABAA receptors are positioned within the synaptic membrane through direct interaction with cytoplasmic gephyrin clusters. Although structural investigations of GABAA receptors have not fully elucidated their unstructured intracellular regions, studies on interaction with gephyrin have revealed that this region can adopt a stable conformation upon interaction (Khayenko and Maric Reference Khayenko and Maric2019). Interestingly, as already mentioned in the above sections, the unstructured regions within GABAARs play a critical role in regulating kinetics of distinct conformational transitions. Additionally, these structures may play a role in various functional processes, such as biosynthesis, recycling, diffusion, and synaptic localization. It is worth noting that these regions exhibit a high degree of variability across different receptor subunits, enabling the subunit-specific regulation of GABAergic transmission. The organization of the intracellular region of the GABAA receptor is also modulated by posttranslational modifications, such as phosphorylation (Nakamura et al. Reference Nakamura, Darnieder, Deeb and Moss2015; Zacchi et al. Reference Zacchi, Antonelli and Cherubini2014), SUMOylation (Small Ubiquitin-like Modifier protein attachment), and acetylation (Ghosh et al. Reference Ghosh, Auguadri, Battaglia, Simone Thirouin, Zemoura, Messner, Acuna, Wildner, Yevenes, Dieter, Kawasaki, Hottiger, Zeilhofer, Fritschy and Tyagarajan2016), which occur in gephyrin or GABAAR subunits. In particular, phosphorylation, regulated by various kinases and phosphatases, plays a crucial role in controlling the trafficking, expression, and allosteric modulation of GABAAR (Kittler and Moss Reference Kittler and Moss2003). The addition of negative charges upon phosphorylation can significantly alter the conformation of the intracellular receptor portion and consequently affect its interactions with proteins in the postsynaptic density of inhibitory synapses.

Inhibitory synapses, similar to their excitatory counterparts, demonstrate impressive potential for plastic changes across a wide range of timescales. This plasticity can manifest through both pre- and postsynaptic expression mechanisms, including transient synaptic modifications that last for seconds and more enduring effects, relying on a larger or smaller extent of translation and transcription (Chiu et al. Reference Chiu, Barberis and Higley2019). Among various forms of long-term plasticity phenomena observed in inhibitory synapses, particular attention has been drawn to heterosynaptic postsynaptic GABAergic plasticity, which arises following the activation of NMDA receptors. This inhibitory plasticity is referred to as heterosynaptic because its induction begins with the influx of calcium ions through NMDA receptors in excitatory synapses (Chiu et al. Reference Chiu, Martenson, Yamazak, Natsume, Sakimura, Tomita, Tavalin and Higley2018). Notably, the expression of this plasticity substantially hinges on protein–protein interactions, specifically involving the GABAA receptor (Marsden et al. Reference Marsden, Beattie, Friedenthal and Carroll2007). For example, NMDA-dependent inhibitory long-term potentiation (iLTP) is induced by the aforementioned calcium influx, subsequent CaMKII activation, and its translocation to inhibitory synapses (Marsden et al. Reference Marsden, Shemesh, Bayer and Carroll2010), where diverse phosphorylation events involving gephyrin and GABAA receptors occur (Petrini et al. Reference Petrini, Ravasenga, Hausrat, Iurilli, Olcese, Racine, Sibarita, Jacob, Moss, Benfenati, Medini, Kneussel and Barberis2014). In particular, CaMKII-dependent phosphorylation of β3S383 facilitates iLTP by promoting the interaction of the receptor with gephyrin. This leads to the accumulation and immobilization of these receptors at the strengthened inhibitory synapses (Petrini et al. Reference Petrini, Ravasenga, Hausrat, Iurilli, Olcese, Racine, Sibarita, Jacob, Moss, Benfenati, Medini, Kneussel and Barberis2014). In particular, the induction of NMDA-iLTP is frequently characterized by incorporation of GABAA receptors containing specific subunits, namely β2 in the cortex (Chiu et al. Reference Chiu, Martenson, Yamazak, Natsume, Sakimura, Tomita, Tavalin and Higley2018) and α5 in the hippocampus (Brzdąk et al. Reference Brzdąk, Lebida, Wyroślak and Mozrzymas2023; Wu et al., Reference Wu, Han and Lu2022a). Hence, iLTP induction results in modification of the kinetic properties of synaptic inhibitory currents, which is attributed to the altered proportion of GABAA receptors containing distinct subunits.

Recent studies have shed light on the influence of extracellular factors on GABAergic long-term potentiation. In particular, the inhibition of matrix metalloproteinase 3 (MMP3), which is responsible for cleaving extracellular matrix constituents, has been shown to hinder the induction of iLTP (Wiera et al. Reference Wiera, Lebida, Lech, Brzdąk, Van Hove, De Groef, Moons, Petrini, Barberis and Mozrzymas2021). This finding is further supported by evidence demonstrating that the administration of exogenous active MMP3 or interference with the activity of integrin adhesion receptors leads to the potentiation of GABAergic synaptic transmission (Wiera et al. Reference Wiera, Brzdąk, Lech, Lebida, Jabłońska, Gmerek and Mozrzymas2022). Interestingly, despite the significance of extracellular protein–protein interactions in synaptic function, our understanding of the interactions involving the ECD of GABAA receptors in the central nervous system synapses remains limited. This is particularly intriguing considering the diverse array of such interactions observed with other neurotransmitter receptors (Riva et al. Reference Riva, Eibl, Volkmer, Carbone and Plested2017). However, recent investigations have expanded our knowledge in this area by exploring the interactions between GABAA receptors and various auxiliary transmembrane proteins, which play crucial roles in modulating receptor pharmacology, trafficking, and plasticity (reviewed in Han et al. Reference Han, Shepard and Lu2021). Among these auxiliary proteins, Shisa7, a single-pass transmembrane protein, has emerged as a key regulator of GABAA receptors. Shisa7 binds to the ECD of GABAARs, exerting control over their synaptic abundance (Han et al. Reference Han, Li, Pelkey, Pandey, Chen, Wang, Wu, Ge, Li, Castellano, Liu, Wu, Petralia, Lynch, McBain and Lu2019), gating kinetics (Castellano et al. Reference Castellano, Wu, Keramidas and Lu2022), sensitivity to BDZs (Han et al. Reference Han, Li, Pelkey, Pandey, Chen, Wang, Wu, Ge, Li, Castellano, Liu, Wu, Petralia, Lynch, McBain and Lu2019) and inhibitory GABAergic plasticity (Wu et al., Reference Wu, Han, Tian, Li and Lu2021b; Wu et al., Reference Wu, Shepard, Castellano, Han, Tian, Dong and Lu2022b). In the future, the integration of new proteomic and structural biology approaches is expected to be particularly valuable for advancing our understanding of GABAA receptor auxiliary proteins and extracellular interactions in the physiology of inhibitory synapses.

State-dependent phosphorylation and modulation of GABAA receptors

GABAA receptors can migrate between adjacent dendritic GABAergic synapses, particularly in prolonged desensitized states (de Luca et al. Reference de Luca, Ravasenga, Petrini, Polenghi, Nieus, Guazzi and Barberis2017). This diffusion facilitates the transfer of desensitized receptors between adjacent inhibitory synapses, thereby decreasing inhibitory synaptic currents and altering short-term synaptic plasticity (Maynard and Triller Reference Maynard and Triller2019). However, the movement of GABAARs between inhibitory synapses is controlled by excitatory activity. In response to glutamatergic activity, GABAA receptors become transiently trapped at excitatory synapses, which constrains receptor diffusion between inhibitory synapses. Consequently, excitatory activity limits the intersynaptic diffusion of desensitized GABAA receptors (de Luca et al. Reference de Luca, Ravasenga, Petrini, Polenghi, Nieus, Guazzi and Barberis2017). Future studies should determine whether transient trapping of the GABAA receptor at the excitatory synapse occurs through molecular crowding in the synapse or as a result of direct interactions with specific postsynaptic proteins. Another aspect of inhibitory plasticity involves the state-dependent phosphorylation of GABAARs, which influences their synaptic clustering. Field et al. (Reference Field, Dorovykh, Thomas and Smart2021) reported that desensitization of GABAA receptors facilitates their phosphorylation by protein kinase C (PKC), increasing their presence at inhibitory synapses. During this process, the extended occupancy of GABAA receptors in the desensitized states generates an unidentified signal that drives PKC-dependent phosphorylation, thereby potentiating inhibitory synapses. Structural investigations are necessary to precisely elucidate how the entry of receptors into desensitized states enhances GABAA receptor phosphorylation, modifies their membrane mobility, and ultimately affects synaptic efficacy (Merlaud et al. Reference Merlaud, Marques, Russeau, Saade, Tostain, Moutkine, Gielen, Corringer and Lévi2022).

Pharmacological modulators of the GABAA receptor can also impact its lateral diffusion within the membrane. For instance, the application of the inverse BDZ agonist DMCM has been observed to increase the diffusion and decrease the clustering of γ2GABAA receptors at synapses, resulting in reduced postsynaptic currents (Lévi et al. Reference Lévi, Le Roux, Eugène and Poncer2015). This effect occurred without any noticeable changes in the size of gephyrin clusters, indicating that the changes in receptor diffusion were not due to rapid disassembly of the synaptic scaffold. Furthermore, another study demonstrated that prolonged application of a high concentration of muscimol, which traps GABAA receptors in a desensitized state, alters the mobility of only the extrasynaptic receptor pool, reducing the diffusion of α2GABAARs and increasing it for α4 receptors (Hannan et al. Reference Hannan, Minere, Harris, Izquierdo, Thomas, Tench and Smart2019). As tonic inhibitory transmission exhibits plasticity in response to neuronal activity (Saliba et al. Reference Saliba, Kretschmannova and Moss2012; Wu et al. Reference Wu, Castellano, Tian and Lu2021a; Wyroślak et al. Reference Wyroślak, Lebida and Mozrzymas2021, Reference Wyroślak, Dobrzański and Mozrzymas2023), future studies should investigate how posttranslational modifications and the interaction of GABAA receptors with diverse protein partners during transport to, within, and from the membrane influence the extent of tonic inhibition. Finally, the described results clearly illustrate that structural changes occurring in the receptor after its activation, desensitization, and modulation affect the protein–protein interactions between the receptor and its intracellular partners, leading to alterations in its membrane diffusion properties. This interplay also operates in the reverse direction, in which the binding of proteins to the GABAA receptor modulates its kinetic properties, subcellular localization, and functions (Chen et al. Reference Chen, Wang, Vicini and Olsen2000; Han et al. Reference Han, Li, Pelkey, Pandey, Chen, Wang, Wu, Ge, Li, Castellano, Liu, Wu, Petralia, Lynch, McBain and Lu2019, Reference Han, Shepard and Lu2021; Pizzarelli et al. Reference Pizzarelli, Griguoli, Zacchi, Petrini, Barberis, Cattaneo and Cherubini2020). Future studies should shed light on the structural and molecular mechanisms of this process.

Concluding remarks

Recent progress and further challenges

In the past few years, a very substantial step forward was achieved in structural studies, especially of heteromeric pentamers representative for synaptic receptors, with a variety of agonists and antagonists bound to them (Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020; Masiulis et al. Reference Masiulis, Desai, Uchański, Serna Martin, Laverty, Karia, Malinauskas, Zivanov, Pardon, Kotecha, Steyaert, Miller and Aricescu2019; Miller and Aricescu Reference Miller and Aricescu2014). Particularly abundant is the body of evidence based on functional studies of WT and mutated receptors that recently was substantially enriched by particularly insightful single-channel analysis. The availability of structural and functional information, together with increasing computational power, has boosted the application of bioinformatic approaches, in particular molecular dynamics, structural studies based on homology modeling, docking, functional kinetic modeling, and so on. However, the lack of high-quality structures of GABAAR heteromers in open conformation(s) poses serious limitations in deciphering the molecular mechanisms of receptor activation. Indeed, despite the wealth of structure–function information, precise molecular scenarios of GABAAR conformational transitions remain elusive. In this final chapter, we would like to draw the reader’s attention to some novel aspects that became evident in the light of recent progress in this field.

Binding to GABAAR orthosteric binding site is local and gating is global

Considering that the orthosteric binding site is very distant (approximately 50 Å) from the channel gates made of 9′ residues forming the activation gate and -2′ residues implicated in the so-called desensitization gate, it was expected that areas close to the binding site would be involved primarily in agonist binding, whereas those near the gates would shape the receptor gating. However, surprisingly, mutations located at the ‘top’ of the ECD (β2F31 and α1F14(r)), which is even more distant from the gate than the orthosteric binding sites, clearly affected receptor gating, particularly opening/closing and microscopic desensitization (Terejko et al. Reference Terejko, Michałowski, Dominik, Andrzejczak and Mozrzymas2021a). In addition, the observations that several mutations at the hydrophobic binding cassette affect receptor gating (e.g., β2E155 at loop B, β2F200 at loop C, α1D43(r) at loop G, α1F64(r) at loop D, or α1R132 at loop E (Baptista-Hon et al. Reference Baptista-Hon, Krah, Zachariae and Hales2016; Goldschen-ohm et al. Reference Goldschen-ohm, Wagner and Jones2011; Jatczak-Śliwa et al. Reference Jatczak-Śliwa, Kisiel, Czyzewska, Brodzki and Mozrzymas2020; Szczot et al. Reference Szczot, Kisiel, Czyzewska and Mozrzymas2014; Venkatachalan and Czajkowski Reference Venkatachalan and Czajkowski2012) and induce spontaneous activity (e.g., β2L99 at loop B, β2E155 at loop B and α1F65 at loop D (Boileau et al. Reference Boileau, Newell and Czajkowski2002; Kisiel et al. Reference Kisiel, Jatczak, Brodzki and Mozrzymas2018; Newell et al. Reference Newell, McDevitt and Czajkowski2004) show that the mechanisms shaping gating encompass large portions of the macromolecule, including its orthosteric binding site. This feature of the receptor was particularly clear in our recent studies based on high-resolutiuon macroscopic and single-channel analysis. Indeed, to our surprise, mutations of α1F64(r) at loop D of the complementary subunit strongly affected not only binding but all gating transitions (preactivation, opening/closing, and desensitization, Kisiel et al. Reference Kisiel, Jatczak, Brodzki and Mozrzymas2018; Kłopotowski et al. Reference Kłopotowski, Czyżewska and Mozrzymas2021; Szczot et al. Reference Szczot, Kisiel, Czyzewska and Mozrzymas2014). A similar observation was made in the principal subunit side of the binding site, where loop C residue β2F200 was identified as important not only for agonist binding but also for gating (Terejko et al. Reference Terejko, Kaczor, Michałowski, Dąbrowska and Mozrzymas2020). Similar to the top part and the binding site in the middle of the ECD, numerous residues involved in gating were identified in the interface domain (e.g., β2V53 on loop 2 and β2P273 on M2-M3 loop, Brodzki and Mozrzymas Reference Brodzki and Mozrzymas2022; Kłopotowski et al. Reference Kłopotowski, Michałowski, Gos, Mosiądz, Czyżewska and Mozrzymas2023. These findings clearly indicate the role of the interface in transmitting the activation signal from the binding site to the TMD and the pore gate. In general, in our hands, the majority of mutations located all over the GABAAR revealed clear effects on receptor gating. This observation raises the question regarding the dependence of channel gates on other parts of the macromolecule. Since structural and mutagenesis studies clearly indicate that activation and desensitization of the receptor are particularly sensitive to mutations localized at the above-mentioned sites in the TMD (Gielen et al. Reference Gielen, Thomas and Smart2015; Gielen et al. Reference Gielen, Barilone and Corringer2020; Gielen and Corringer Reference Gielen and Corringer2018; Terejko et al. Reference Terejko, Michałowski, Iżykowska, Dominik, Brzóstowicz and Mozrzymas2021b), these two gates appear to be only a part of the molecular machinery underlying these conformational transitions. We would propose that 9′ and -2′ sites at the TMD are ‘hot spots’ for activation and desensitization mechanisms, but to a large extent these transitions are also determined by vast parts of the macromolecule, including ‘top’ of the ECD, vicinity of the orthosteric binding sites, interface and of course the TMD. These findings suggest that GABAARs gating is particularly functionally and structurally compact, meaning that mutation at any location is likely to affect gating transitions. On the other hand, studies have confirmed that the binding process to the orthosteric binding site is much more ‘local’ than receptor gating. Indeed, when mutating the residues increasingly distant from the binding site and close to the interface or within the TMD, the impact on the binding step decreased, becoming negligible for TMD close to the cytoplasmic part (Terejko et al. Reference Terejko, Michałowski, Iżykowska, Dominik, Brzóstowicz and Mozrzymas2021b).

Pharmacological modulation and global gating of GABAAR

It is well established that GABAAR is exceptionally susceptible to modulation by a variety of endogenous and exogenous modulators, many of which are of clinical importance (e.g., BDZs or general anesthetics). Notably, the binding site for BDZs bears some similarity to the orthosteric binding site, with the difference that it is localized at the interface of the α and γ subunits, but its distance from the channel gate is similar. Interestingly, while BDZs were long thought to mainly upregulate agonist binding, recently a consensus was achieved that their positive modulation of GABAARs occurs primarily as a consequence of gating modulation, particularly of the preactivation (flipping) transitions (Dixon et al. Reference Dixon, Sah, Lynch and Keramidas2014; Goldschen-Ohm et al. Reference Goldschen-Ohm, Haroldson, Jones and Pearce2014; Jatczak-Śliwa et al. Reference Jatczak-Śliwa, Terejko, Brodzki, Michałowski, Czyzewska, Nowicka, Andrzejczak, Srinivasan and Mozrzymas2018; Szczot et al. Reference Szczot, Kisiel, Czyzewska and Mozrzymas2014). Again, this finding indicates that modulation of gating is structurally highly non-local, relying on long-range mechanisms. Although other modulatory binding sites, such as extensively studied neurosteroids or barbiturates, are considerably closer to the postulated receptor’s gates (9′ and -2′), their molecular modulatory mechanisms must still comprise long-range interactions. This observation appears consistent with the above-mentioned concept of diffuse and non-local gating of GABAAR. Indeed, if substitutions of residues located distantly from the gates strongly affect receptor gating, it is not surprising that local interactions with modulators produce analogous effects, even for binding sites located distantly from the channel’s gates at the TMD. These features of GABAAR lead to the conclusion that this receptor is particularly ‘highly allosteric’, especially when considering the impact of modulating agents or mutations on receptor gating. Thus, we believe that the key to understanding the exceptional susceptibility of GABAAR to pharmacological modulation lies in its structural (and functional) compactness.

The general consensus (based on comparative studies of receptor structures in open and closed or desensitized states) is that upon ligand binding, the receptor undergoes two major structural rearrangements: its ECD structure becomes more compact (a motion often called unblooming/blooming), and the respective subunits twist relative to each other (with an anticlockwise rotation during activation that starts from the ECD and propagates to the TMD). The concept of coupled motions distributed across the entire receptor structure aligns with structural and modeling data from other pLGICs. This model was proposed for GLIC (Sauguet et al. Reference Sauguet, Shahsavar, Poitevin, Huon, Menny, Nemecz, Haouz, Changeux, Corringer and Delarue2014) and was later also confirmed for other pLGICs (Noviello et al. Reference Noviello, Gharpure, Mukhtasimova, Cabuco, Baxter, Borek, Sine and Hibbs2021; Yu et al. Reference Yu, Zhu, Lape, Greiner, Du, Lü, Sivilotti and Gouaux2021). Due to the limited data on the open structure of GABAAR, such analyses have been conducted only between closed/resting and desensitized states (Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020; Masiulis et al. Reference Masiulis, Desai, Uchański, Serna Martin, Laverty, Karia, Malinauskas, Zivanov, Pardon, Kotecha, Steyaert, Miller and Aricescu2019). Nevertheless, the scheme of ECD compaction and anticlockwise rotation of both domains was confirmed. These findings suggest that almost every residue in the receptor may significantly impact and contribute to the gating process, reflecting the global nature of these motions rather than specific pathways driven by residue-to-residue interactions. However, the roles of receptor subunits may be highly diversified. Within the orthosteric binding site in the ECD, both the principal (β) and complementary (α) subunits play crucial roles in receptor function. For example, residues from both subunits – β2E155 and α1R66 – contribute to the electrostatic environment required for agonist binding (Goldschen-ohm et al. Reference Goldschen-ohm, Wagner and Jones2011; Michałowski et al. Reference Michałowski, Czyżewska, Iżykowska and Mozrzymas2021; Newell et al. Reference Newell, McDevitt and Czajkowski2004), while other residues, such as β2F200 or α1F65, significantly impact both binding and gating (Szczot et al. Reference Szczot, Kisiel, Czyzewska and Mozrzymas2014; Terejko et al. Reference Terejko, Kaczor, Michałowski, Dąbrowska and Mozrzymas2020). On the other hand, in the allosteric binding site in the ECD formed by α (serving a principal role) and γ (complementary) subunits, while γ subunit is necessary to bind BDZ, the modulatory effect of BDZs is mediated mostly by α subunit (Tan et al. Reference Tan, Rudolph and Lüscher2011) with a prominent role of α1H101(r) (Sharkey and Czajkowski Reference Sharkey and Czajkowski2008). On the contrary, in the domain interface, it seems that the impact of the principal subunit is much more pronounced – mainly in the region of the M2-M3 loop and loop2 – where mutations of β2V53 and β2P273 had a bigger impact on gating than their homologs in the complementary subunit (α1H55(r) and α1P277C(r)) (Brodzki and Mozrzymas Reference Brodzki and Mozrzymas2022; Kaczor et al. Reference Kaczor, Wolska and Mozrzymas2021, Reference Kaczor, Michałowski and Mozrzymas2022; Kłopotowski et al. Reference Kłopotowski, Michałowski, Gos, Mosiądz, Czyżewska and Mozrzymas2023. However, within the TMD, nearly every subunit interface forms a modulatory binding site, except for the α/γ interface (Kim et al. Reference Kim, Gharpure, Teng, Zhuang, Howard, Zhu, Noviello, Walsh, Lindahl and Hibbs2020; Maldifassi et al. Reference Maldifassi, Baur and Sigel2016; Nourmahnad et al. Reference Nourmahnad, Stern, Hotta, Stewart, Ziemba, Szabo and Forman2016). Mutations in these regions affect not only modulator binding but also receptor gating, as exemplified by α1M236 and its homologs in other subunits (Nourmahnad et al. Reference Nourmahnad, Stern, Hotta, Stewart, Ziemba, Szabo and Forman2016). Similarly, the channel pore is symmetrically formed by each subunit, including the principal for gate formation residues β2L259 and α1L264.

Recent advancements in functional and structural studies of pLGICs have significantly enhanced our understanding of receptor mechanisms. Despite this progress, the lack of high-quality structures of GABAAR heteromers in open conformations limits our understanding of receptor activation. Unexpectedly, mutations distant from the channel gates affect receptor gating, indicating a widespread influence on gating mechanisms across the receptor structure. These findings suggest that GABAARs exhibit a highly allosteric nature, where mutations or modulators at various sites can impact gating transitions. This structural and functional compactness likely underlies GABAAR’s sensitivity to pharmacological modulation, emphasizing the importance of coupled motions across its entire structure.

Acknowledgments

This work was supported by funding from National Science Centre (Poland) grant MAESTRO 2015/18/A/NZ1/00395 and OPUS 2021/43/B/NZ4/0167. M.A.M. was partially supported by Wroclaw Medical University grant SUBK.A400.22.012.

Competing interest

Authors declare no conflicts of interest.

References

Akk, G, Bracamontes, J, and Steinbach, JH (2001) Pregnenolone sulfate block of GABAA receptors: mechanism and involvement of a residue in the M2 region of the α subunit. Journal of Physiology 532(3), 673684. https://doi.org/10.1111/J.1469-7793.2001.0673E.X.CrossRefGoogle ScholarPubMed
Akk, G, Germann, AL, Sugasawa, Y, Pierce, SR, Evers, AS, and Steinbach, JH (2020) Enhancement of muscimol binding and gating by allosteric modulators of the GABA A receptor: relating occupancy to state functions. Molecular Pharmacology 98(4). MOLPHARM-AR-2020-000066. https://doi.org/10.1124/molpharm.120.000066.CrossRefGoogle ScholarPubMed
Akk, G, Li, P, Bracamontes, J, Wang, M, and Steinbach, JH (2011) Pharmacology of structural changes at the GABAA receptor transmitter binding site. British Journal of Pharmacology 162(4) 840. https://doi.org/10.1111/J.1476-5381.2010.01083.X.CrossRefGoogle ScholarPubMed
Amin, J, and Weiss, DS (1993) GABAA receptor needs two homologous domains of the & beta;-subunit for activation by GABA but not by pentobarbital. Nature 366(6455), 565569. https://doi.org/10.1038/366565a0.CrossRefGoogle Scholar
Amundarain, MJ, Ribeiro, RP, Costabel, MD, and Giorgetti, A (2019) GABA A receptor family: overview on structural characterization. Future Medicinal Chemistry 11(3), 229246. https://doi.org/10.4155/fmc-2018-0336.CrossRefGoogle ScholarPubMed
Ashby, JA, McGonigle, IV, Price, KL, Cohen, N, Comitani, F, Dougherty, DA, Molteni, C, and Lummis, SCR (2012) GABA binding to an insect GABA receptor: a molecular dynamics and mutagenesis study. Biophysical Journal 103(10), 20712081. https://doi.org/10.1016/j.bpj.2012.10.016.CrossRefGoogle Scholar
Bali, M, and Akabas, MH (2004) Defining the propofol binding site location on the GABAA receptor. Molecular Pharmacology 65(1), 6876. https://doi.org/10.1124/mol.65.1.68.CrossRefGoogle ScholarPubMed
Bali, M, Jansen, M, and Akabas, MH (2009) Gaba-induced intersubunit conformational movement in the gabaa receptor α1m1-β2m3 transmembrane subunit interface: experimental basis for homology modeling of an intravenous anesthetic binding site. Journal of Neuroscience 29(10), 30833092. https://doi.org/10.1523/JNEUROSCI.6090-08.2009.CrossRefGoogle ScholarPubMed
Baptista-Hon, DT, Gulbinaite, S, and Hales, TG (2017) Loop G in the GABAA receptor α1 subunit influences gating efficacy. Journal of Physiology 595(5), 17251741. https://doi.org/10.1113/JP273752.CrossRefGoogle Scholar
Baptista-Hon, DT, Krah, A, Zachariae, U and Hales, TG (2016) A role for loop G in the α1 strand in GABAA receptor activation. The Journal of Physiology 00(19), 135. https://doi.org/10.1113/JP272463.Google Scholar
Barberis, A, Cherubini, E and Mozrzymas, JW (2000) Zinc inhibits miniature GABAergic currents by allosteric modulation of GABAA receptor gating. The Journal of Neuroscience: The Official Journal of the Society for Neuroscience 20(23), 86188627. https://doi.org/10.1523/JNEUROSCI.20-23-08618.2000.CrossRefGoogle ScholarPubMed
Barbour, B, and Häusser, M (1997). Intersynaptic diffusion of neurotransmitter. Trends in Neurosciences 20(9), 377384. https://doi.org/10.1016/S0166-2236(96)20050-5.Google ScholarPubMed
Bar-Lev, DD, Degani-Katzav, N, Perelman, A, and Paas, Y (2011) Molecular dissection of Cl --selective cys-loop receptor points to components that are dispensable or essential for channel activity. Journal of Biological Chemistry 286(51), 4383043841. https://doi.org/10.1074/jbc.M111.282715.CrossRefGoogle ScholarPubMed
Bartos, M, Corradi, J and Bouzat, C (2009) Structural basis of activation of cys-loop receptors: the extracellular–transmembrane interface as a coupling region. Molecular Neurobiology 40(3), 236252. https://doi.org/10.1007/S12035-009-8084-X.CrossRefGoogle ScholarPubMed
Baumann, SW, Baur, R, and Sigel, E (2003) Individual properties of the two functional agonist sites in GABA(A) receptors. The Journal of Neuroscience 23(35), 1115811166. https://doi.org/10.1523/JNEUROSCI.23-35-11158.2003.CrossRefGoogle ScholarPubMed
Baur, Rand Sigel, E (2003) On high- and low-affinity agonist sites in GABAA receptors. Journal of Neurochemistry 87(2), 325332. https://doi.org/10.1046/j.1471-4159.2003.01982.x.CrossRefGoogle ScholarPubMed
Beato, M (2008) The time course of transmitter at glycinergic synapses onto motoneurons. The Journal of Neuroscience : The Official Journal of the Society for Neuroscience 28(29), 74127425. https://doi.org/10.1523/JNEUROSCI.0581-08.2008.CrossRefGoogle ScholarPubMed
Beaumont, K, Chilton, WS, Yamamura, HI, and Enna, SJ (1978). Muscimol binding in rat brain: Association with synaptic GABA receptors. Brain Research 148(1), 153162. https://doi.org/10.1016/0006-8993(78)90385-2.CrossRefGoogle ScholarPubMed
Belelli, D, Lambert, JJ, Peters, JA, Wafford, K and Whiting, PJ (1997). The interaction of the general anesthetic etomidate with the γ-aminobutyric acid type A receptor is influenced by a single amino acid. Proceedings of the National Academy of Sciences of the United States of America, 94(20), 1103111036. https://doi.org/10.1073/pnas.94.20.11031.CrossRefGoogle ScholarPubMed
Bera, AK, Chatav, M and Akabas, MH (2002) GABAA receptor M2-M3 loop secondary structure and changes in accessibility during channel gating. Journal of Biological Chemistry 277(45), 4300243010. https://doi.org/10.1074/jbc.M206321200.CrossRefGoogle ScholarPubMed
Bianchi, MT, Botzolakis, EJ, Lagrange, AHand Macdonald, RL (2009) Benzodiazepine modulation of GABA(A) receptor opening frequency depends on activation context: a patch clamp and simulation study. Epilepsy Research 85(2–3), 212220. https://doi.org/10.1016/J.EPLEPSYRES.2009.03.007.CrossRefGoogle ScholarPubMed
Bocquet, N, Nury, H, Baaden, M, Le Poupon, C, Changeux, J-P, Delarue, M, and Corringer, P-J (2009). X-ray structure of a pentameric ligand-gated ion channel in an apparently open conformation. Nature 457(7225), 111114. https://doi.org/10.1038/nature07462.CrossRefGoogle Scholar
Bocquet, N, Prado De Carvalho, L, Cartaud, J, Neyton, J, Le Poupon, C, Taly, A, Grutter, T, Changeux, JP and Corringer, PJ (2007) A prokaryotic proton-gated ion channel from the nicotinic acetylcholine receptor family. Nature 445(7123), 116119. https://doi.org/10.1038/nature05371.CrossRefGoogle ScholarPubMed
Boileau, AJ, Evers, AR, Davis, AF and Czajkowski, C (1999) Mapping the agonist binding site of the GABAA receptor: evidence for a beta-strand. The Journal of Neuroscience : The Official Journal of the Society for Neuroscience 19(12), 48474854. https://doi.org/10.1523/JNEUROSCI.19-12-04847.1999.CrossRefGoogle ScholarPubMed
Boileau, AJ, Li, T, Benkwitz, C, Czajkowski, C and Pearce, RA (2003) Effects of γ2S subunit incorporation on GABAA receptor macroscopic kinetics. Neuropharmacology 44(8), 10031012. https://doi.org/10.1016/S0028-3908(03)00114-X.CrossRefGoogle ScholarPubMed
Boileau, AJ, Newell, JG, and Czajkowski, C (2002) GABAA receptor β2Tyr97 and leu99 line the GABA-binding site. Journal of Biological Chemistry 277(4), 29312937. https://doi.org/10.1074/jbc.M109334200.CrossRefGoogle Scholar
Borghese, CM, Hicks, JA, Lapid, DJ, Trudell, JR and Harris, RA (2014)GABAA receptor transmembrane amino acids are critical for alcohol action: disulfide cross-linking and alkyl methanethiosulfonate labeling reveal relative location of binding sites. Journal of Neurochemistry 128(3), 363375. https://doi.org/10.1111/jnc.12476.CrossRefGoogle ScholarPubMed
Bouron, A (2001) Modulation of spontaneous quantal release of neurotransmitters in the hippocampus. Progress in Neurobiology 63(6), 613635. https://doi.org/10.1016/S0301-0082(00)00053-8.CrossRefGoogle ScholarPubMed
Brickley, SG and Mody, I (2012) Extrasynaptic GABAA receptors: their function in the cns and implications for disease. Neuron 73(1), 2334. https://doi.org/10.1016/J.NEURON.2011.12.012.CrossRefGoogle ScholarPubMed
Brodzki, M, Michałowski, MA, Gos, M and Mozrzymas, JW (2020) Mutations of α1F45 residue of GABAA receptor loop G reveal its involvement in agonist binding and channel opening/closing transitions. Biochemical Pharmacology 177, 113917. https://doi.org/10.1016/j.bcp.2020.113917.CrossRefGoogle Scholar
Brodzki, M and Mozrzymas, JW (2022)GABAA receptor proline 273 at the interdomain interface of the β2 subunit regulates entry into desensitization and opening/closing transitions. Life Sciences 308, 120943. https://doi.org/10.1016/J.LFS.2022.120943.CrossRefGoogle ScholarPubMed
Brodzki, M, Rutkowski, R, Jatczak, M, Kisiel, M, Czyzewska, MM and Mozrzymas, JW (2016) Comparison of kinetic and pharmacological profiles of recombinant α1γ2L and α1β2γ2L GABAA receptors – a clue to the role of intersubunit interactions. European Journal of Pharmacology 784, 8189. https://doi.org/10.1016/j.ejphar.2016.05.015.CrossRefGoogle Scholar
Bruns, D and Jahn, R (1995) Real-time measurement of transmitter release from single synaptic vesicles. Nature 377(6544), 6265. https://doi.org/10.1038/377062A0.CrossRefGoogle ScholarPubMed
Brzdąk, P, Lebida, K, Wyroślak, M and Mozrzymas, JW (2023). GABAergic synapses onto SST and PV interneurons in the CA1 hippocampal region show cell-specific and integrin-dependent plasticity. Scientific Reports 13(1). https://doi.org/10.1038/S41598-023-31882-4.CrossRefGoogle ScholarPubMed
Carpenter, TS, Lau, EY and Lightstone, FC (2012) A role for loop F in modulating GABA binding affinity in the GABA A receptor. Journal of Molecular Biology 422(2), 310323. https://doi.org/10.1016/j.jmb.2012.05.025.CrossRefGoogle Scholar
Carpenter, TS and Lightstone, FC (2016)An Electrostatic Funnel in the GABA-Binding Pathway. PLOS Computational Biology 12(4), e1004831. https://doi.org/10.1371/journal.pcbi.1004831.CrossRefGoogle ScholarPubMed
Castellano, D, Shepard, RD and Lu, W (2021) Looking for novelty in an “old” receptor: recent advances toward our understanding of gabaars and their implications in receptor pharmacology. Frontiers in Neuroscience 14, 115. https://doi.org/10.3389/fnins.2020.616298.CrossRefGoogle Scholar
Castellano, D, Wu, K, Keramidas, A, and Lu, W (2022) Shisa7-dependent regulation of gaba(a) receptor single-channel gating kinetics. J Neurosci 42(47), 87588766. https://doi.org/10.1523/jneurosci.0510-22.2022.CrossRefGoogle ScholarPubMed
Chang, Y and Weiss, DS (1999) Allosteric Activation Mechanism of the α1β2γ2 γ-Aminobutyric Acid Type A Receptor Revealed by Mutation of the Conserved M2 Leucine. Biophysical Journal, 77(5), 25422551.CrossRefGoogle Scholar
Chen, L, Wang, H, Vicini, S and Olsen, RW (2000). The gamma-aminobutyric acid type A (GABAA) receptor-associated protein (GABARAP) promotes GABAA receptor clustering and modulates the channel kinetics. Proc Natl Acad Sci U S A 97(21), 1155711562. https://doi.org/10.1073/pnas.190133497.CrossRefGoogle ScholarPubMed
Chen, Q, Wells, MM, Arjunan, P, Tillman, TS, Cohen, AE, Xu, Y and Tang, P (2018) Structural basis of neurosteroid anesthetic action on GABAAreceptors. Nature Communications 9(1), 110. https://doi.org/10.1038/s41467-018-06361-4.Google Scholar
Chen, ZW, Bracamontes, JR, Budelier, MM, Germann, AL, Shin, DJ, Kathiresan, K, Qian, MX, Manion, B, Cheng, WWL, Reichert, DE, Akk, G, Covey, DF and Evers, AS (2019) Multiple functional neurosteroid binding sites on gabaa receptors. PLoS Biology 17(3), 127. https://doi.org/10.1371/journal.pbio.3000157.CrossRefGoogle ScholarPubMed
Chen, Z-W, Manion, B, Townsend, R, Reichert, DE, Covey, DF, Steinbach, JH, Sieghart, W, Fuchs, K and Evers, AS (2012) Neurosteroid analog photolabeling of a site in the third transmembrane domain of the β3 subunit of the gabaa receptor. Molecular Pharmacology 82(3), 408419. https://doi.org/10.1124/mol.112.078410.CrossRefGoogle Scholar
Chiara, DC, Dostalova, Z, Jayakar, SS, Zhou, X, Miller, KW and Cohen, JB (2012) Mapping general anesthetic binding site(s) in human α1β3γ- aminobutyric acid type a receptors with [ 3H]TDBzl-etomidate, a photoreactive etomidate analogue. Biochemistry 51(4), 836847. https://doi.org/10.1021/BI201772M/SUPPL_FILE/BI201772M_SI_001.PDF.CrossRefGoogle Scholar
Chiu, CQ, Barberis, A and Higley, MJ (2019) Preserving the balance: diverse forms of long-term GABAergic synaptic plasticity. Nature Reviews Neuroscience 20(5), 272281. https://doi.org/10.1038/s41583-019-0141-5.CrossRefGoogle ScholarPubMed
Chiu, CQ, Martenson, JS, Yamazak, M, Natsume, R, Sakimura, K, Tomita, S, Tavalin, SJ, and Higley, MJ (2018) input-specific nmdar-dependent potentiation of dendritic gabaergic inhibition. Neuron 97(2), 368https://doi.org/10.1016/j.neuron.2017.12.032.CrossRefGoogle ScholarPubMed
Clements, JD (1996) Transmitter timecourse in the synaptic cleft: its role in central synaptic function. Trends in Neurosciences 19(5), 163171. https://doi.org/10.1016/S0166-2236(96)10024-2.CrossRefGoogle ScholarPubMed
Colquhoun, D, Hawkes, AG and K, S. (1996). Joint distributions of apparent open and shut times of single-ion channels and maximum likelihood fitting of mechanisms. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 354, 25552590.Google Scholar
Colquhoun, D, Hatton, CJ and Hawkes, AG (2003) The quality of maximum likelihood estimates of ion channel rate constants. The Journal of Physiology 547( 3), 699728. https://doi.org/10.1113/jphysiol.2002.034165.CrossRefGoogle ScholarPubMed
Colquhoun, D and Lape, R (2012)Allosteric coupling in ligand-gated ion channels. The Journal of General Physiology 140(6), 599612. https://doi.org/10.1085/jgp.201210844.CrossRefGoogle ScholarPubMed
Colquhoun, D and Sigworth, FJ (1995) Fitting and statistical analysis of single channel records. In Single Channel Recording (ed. Sakmann, B. & Neher, E.), (pp. 483587), New York and London: Plenum Press. https://doi.org/10.1007/978-1-4419-1229-9_19.Google Scholar
Comitani, F, Cohen, N, Ashby, J, Botten, D, Lummis, SCR and Molteni, C (2014) Insights into the binding of GABA to the insect RDL receptor from atomistic simulations: a comparison of models. Journal of Computer-Aided Molecular Design 28(1), 3548. https://doi.org/10.1007/s10822-013-9704-0.CrossRefGoogle Scholar
Comitani, F, Limongelli, V and Molteni, C (2016) The free energy landscape of GABA binding to a pentameric ligand-gated ion channel and its disruption by mutations. Journal of Chemical Theory and Computation 12(7), 33983406. https://doi.org/10.1021/acs.jctc.6b00303.CrossRefGoogle ScholarPubMed
Cowgill, J, Fan, C, Haloi, N, Tobiasson, V, Zhuang, Y, Howard, RJ and Lindahl, E (2023)Structure and dynamics of differential ligand binding in the human ρ-type GABAA receptor. Neuron 111(21), 34503464.e5. https://doi.org/10.1016/J.NEURON.2023.08.006/ATTACHMENT/52392114-2D12-40AB-AECF-A1908D6FA0F1/MMC3.MP4.CrossRefGoogle Scholar
Curtis, DR, Duggan, AW, Felix, D and Johnston, GAR (1970) GABA, bicuculline and central inhibition. Nature. 226(5252), 12221224. https://doi.org/10.1038/2261222a0.CrossRefGoogle ScholarPubMed
Czajkowski, C and Wagner, DA (2001)Structure and dynamics of the GABA binding pocket: a narrowing cleft that constricts during activation. The Journal of Neuroscience: The Official Journal of the Society for Neuroscience 21(1), 6774. https://doi.org/10.1523/JNEUROSCI.21-01-00067.2001.Google Scholar
Czyzewska, MM and Mozrzymas, JW (2013) Monoterpene ??-thujone exerts a differential inhibitory action on GABAA receptors implicated in phasic and tonic GABAergic inhibition. European Journal of Pharmacology 702(1–3), 3843. https://doi.org/10.1016/j.ejphar.2013.01.032.CrossRefGoogle ScholarPubMed
Davenport, CM, Rajappa, R, Katchan, L, Taylor, CR, Tsai, MC, Smith, CM, de Jong, JW, Arnold, DB, Lammel, S and Kramer, RH (2021) Relocation of an extrasynaptic GABA(A) receptor to inhibitory synapses freezes excitatory synaptic strength and preserves memory. Neuron 109(1), 123134.e4. https://doi.org/10.1016/j.neuron.2020.09.037.CrossRefGoogle ScholarPubMed
de Luca, E, Ravasenga, T, Petrini, EM, Polenghi, A, Nieus, T, Guazzi, S and Barberis, A (2017) Inter-synaptic lateral diffusion of gabaa receptors shapes inhibitory synaptic currents. Neuron 95(1), 6369.e5. https://doi.org/10.1016/j.neuron.2017.06.022.CrossRefGoogle ScholarPubMed
Dixon, CL, Harrison, NL, Lynch, JW, and Keramidas, A (2015) Zolpidem and eszopiclone prime a1b2y2 GABAA receptors for longer duration of activity. British Journal of Pharmacology 172(14), 35223536. https://doi.org/10.1111/bph.13142.CrossRefGoogle ScholarPubMed
Dixon, C, Sah, P, Lynch, JW, and Keramidas, A (2014) GABAA receptor α and γ subunits shape synaptic currents via different mechanisms. The Journal of Biological Chemistry 289(9), 53995411. https://doi.org/10.1074/jbc.M113.514695CrossRefGoogle ScholarPubMed
Downing, SS, Lee, YT, Farb, DH, and Gibbs, TT (2005) Benzodiazepine modulation of partial agonist efficacy and spontaneously active GABA(A) receptors supports an allosteric model of modulation. British Journal of Pharmacology 145(7), 894906. https://doi.org/10.1038/sj.bjp.0706251CrossRefGoogle ScholarPubMed
Eaton, MM, Cao, LQ, Chen, Z, Franks, NP, Evers, AS, and Akk, G. (2015) Mutational analysis of the putative high-affinity propofol binding site in human β3 homomeric GABAA receptors. Molecular Pharmacology 88(4), 736745. https://doi.org/10.1124/MOL.115.100347CrossRefGoogle Scholar
Eaton, MM, Germann, LA, Arora, R, Cao, LQ, Gao, X, Shin, JD, Wu, A, Chiara, CD, Cohen, JB, Steinbach, JH, Evers, As, and Akk, G. (2016) Multiple Non-equivalent interfaces mediate direct activation of GABAA receptors by propofol. Current Neuropharmacology 14(7), 772780. https://doi.org/10.2174/1570159X14666160202121319CrossRefGoogle ScholarPubMed
Eaton, MM, Lim, YB, Bracamontes, J, Steinbach, JH, and Akk, G (2012) Agonist-specific Conformational Changes in the Structure of the alpha1-gamma2 Subunit Interface of the GABAA Receptor. Molecular Pharmacology 82(2), 255263. https://doi.org/10.1124/mol.112.077875CrossRefGoogle Scholar
Ernst, M, and Sieghart, W (2015) GABAA receptor subtypes: structural variety raises hope for new therapy concepts. E-Neuroforum 6(4), 97103. https://doi.org/10.1007/s13295-015-0016-9CrossRefGoogle Scholar
Feng, HJ, and Forman, SA (2018) Comparison of αβδ and αβγ GABAAreceptors: allosteric modulation and identification of subunit arrangement by site-selective general anesthetics. Pharmacological Research 133, 289300. https://doi.org/10.1016/j.phrs.2017.12.031CrossRefGoogle Scholar
Field, M, Dorovykh, V, Thomas, P, and Smart, TG (2021) Physiological role for GABAA receptor desensitization in the induction of long-term potentiation at inhibitory synapses. Nature Communications 12(1), 116. https://doi.org/10.1038/s41467-021-22420-9CrossRefGoogle ScholarPubMed
Förstera, B, Castro, PA, Moraga-Cid, G, and Aguayo, LG (2016) Potentiation of Gamma aminobutyric acid receptors (GABAAR) by ethanol: how are inhibitory receptors affected? Frontiers in Cellular Neuroscience 10 5), 117. https://doi.org/10.3389/fncel.2016.00114CrossRefGoogle ScholarPubMed
Franks, NP (2015) Structural comparisons of ligand-gated ion channels in open, closed, and desensitized states identify a novel propofol-binding site on mammalian γ-aminobutyric acid type A receptors. Anesthesiology 122(4), 787794. https://doi.org/10.1097/ALN.0000000000000588CrossRefGoogle ScholarPubMed
Gafford, GM, Guo, JD, Flandreau, EI, Hazra, R., Rainnie, DG, and Ressler, KJ (2012) Cell-type specific deletion of GABA(A) α1 in corticotropin-releasing factor-containing neurons enhances anxiety and disrupts fear extinction. Proceedings of the National Academy of Sciences of the United States of America 109(40), 1633016335. https://doi.org/10.1073/PNAS.1119261109/SUPPL_FILE/PNAS.201119261SI.PDFCrossRefGoogle ScholarPubMed
Garifulina, A, Friesacher, T, Stadler, M, Zangerl-Plessl, EM, Ernst, M, Stary-Weinzinger, A, Willam, A, and Hering, S (2022) β subunits of GABAA receptors form proton-gated chloride channels: Insights into the molecular basis. Communications Biology 5(1) https://doi.org/10.1038/s42003-022-03720-2CrossRefGoogle ScholarPubMed
Germann, AL, Burbridge, AB, Pierce, SR, and Akk, G (2022) Activation of the Rat α1β2ε GABAA receptor by orthosteric and allosteric agonists. Biomolecules 12(7) https://doi.org/10.3390/biom12070868CrossRefGoogle ScholarPubMed
Germann, AL, Steinbach, JH, and Akk, G. (2018) Application of the Co-agonist concerted transition model to analysis of GABAA receptor properties. Current Neuropharmacology 17(9), 843851. https://doi.org/10.2174/1570159x17666181206092418CrossRefGoogle Scholar
Ghosh, H, Auguadri, L, Battaglia, S, Simone Thirouin, Z, Zemoura, K, Messner, S, Acuna, MA, Wildner, H, Yevenes, GE, Dieter, A, Kawasaki, H, Hottiger, OM, Zeilhofer, HU, Fritschy, JM, and Tyagarajan, SK (2016) Several posttranslational modifications act in concert to regulate gephyrin scaffolding and GABAergic transmission. Nat Commun 7, 13365. https://doi.org/10.1038/ncomms13365CrossRefGoogle ScholarPubMed
Gielen, M, Barilone, N, and Corringer, PJ (2020) The desensitization pathway of GABAA receptors, one subunit at a time. Nature Communications 11(1) https://doi.org/10.1038/S41467-020-19218-6CrossRefGoogle ScholarPubMed
Gielen, M, and Corringer, PJ (2018) The dual-gate model for pentameric ligand-gated ion channels activation and desensitization. Journal of Physiology 596(10), 18731902. https://doi.org/10.1113/JP275100CrossRefGoogle ScholarPubMed
Gielen, M., Thomas, P., and Smart, TG (2015) The desensitization gate of inhibitory Cys-loop receptors. Nature Communications 6, 6829. https://doi.org/10.1038/ncomms7829CrossRefGoogle ScholarPubMed
Goldschen-Ohm, MP (2022) Benzodiazepine modulation of GABAA receptors: a mechanistic perspective. Biomolecules 12(12), 1784. https://doi.org/10.3390/BIOM12121784CrossRefGoogle ScholarPubMed
Goldschen-Ohm, MP, Haroldson, A., Jones, MV, and Pearce, RA (2014) A nonequilibrium binary elements-based kinetic model for benzodiazepine regulation of GABA A receptors. The Journal of General Physiology 144(1), 2739. https://doi.org/10.1085/jgp.201411183CrossRefGoogle Scholar
Goldschen-ohm, MP, Wagner, DA, and Jones, MV (2011) Three arginines in the GABAA receptor binding pocket have distinct roles in the formation and stability of agonist- versus antagonist-bound complexes. Molecular Pharmacology 80(4), 647656. https://doi.org/10.1124/mol.111.072033CrossRefGoogle ScholarPubMed
Gottschald Chiodi, C, Baptista-Hon, DT, Hunter, WN, and Hales, TG (2018) Amino acid substitutions in the human homomeric β3 GABAA receptor that enable activation by GABA. Journal of Biological Chemistry 294, jbc.RA118.006229. https://doi.org/10.1074/jbc.ra118.006229Google ScholarPubMed
Gurley, D, Amin, J, Ross, PC, Weiss, DS, and White, G (1995) Point mutations in the M2 region of the alpha, beta, or gamma subunit of the GABAA channel that abolish block by picrotoxin. Receptors & Channels 3(1), 1320. https://europepmc.org/article/med/8589989Google ScholarPubMed
Haas, KF and Macdonald, RL (1999) GABA(A) receptor subunit γ2 and δ subtypes confer unique kinetic properties on recombinant GABA(A) receptor currents in mouse fibroblasts. Journal of Physiology 514(1), 2745. https://doi.org/10.1111/j.1469-7793.1999.027af.x.CrossRefGoogle Scholar
Hájos, N., Nusser, Z., Rancz, EA, Freund, TF, and Mody, I. (2000) Cell type- and synapse-specific variability in synaptic GABAA receptor occupancy. The European Journal of Neuroscience 12(3), 810818. https://doi.org/10.1046/J.1460-9568.2000.00964.XCrossRefGoogle ScholarPubMed
Hales, TG, Deeb, TZ, Tang, H, Bollan, KA, King, DP, Johnson, SJ, and Connolly, CN (2006) An asymmetric contribution to gamma-aminobutyric type A receptor function of a conserved lysine within TM2-3 of alpha1, beta2, and gamma2 subunits. The Journal of Biological Chemistry 281(25), 1703417043. https://doi.org/10.1074/jbc.M603599200CrossRefGoogle ScholarPubMed
Han, W, Li, J, Pelkey, KA, Pandey, S, Chen, X, Wang, YX, Wu, K, Ge, L, Li, T, Castellano, D, Liu, C, Wu, LG, Petralia, RS, Lynch, JW, McBain, CJ, and Lu, W (2019) Shisa7 is a GABA(A) receptor auxiliary subunit controlling benzodiazepine actions. Science 366(6462), 246250. https://doi.org/10.1126/science.aax5719CrossRefGoogle ScholarPubMed
Han, W, Shepard, RD, and Lu, W (2021) Regulation of GABAARs by transmembrane accessory proteins. Trends in Neurosciences 44(2), 152165. https://doi.org/10.1016/j.tins.2020.10.011CrossRefGoogle ScholarPubMed
Hannan, S, Minere, M, Harris, J, Izquierdo, P, Thomas, P, Tench, B, and Smart, TG (2019) GABAAR isoform and subunit structural motifs determine synaptic and extrasynaptic receptor localisation. Neuropharmacology 107540. https://doi.org/10.1016/j.neuropharm.2019.02.022Google ScholarPubMed
Harris, KM, and Stevens, JK (1989) Dendritic spines of CA 1 pyramidal cells in the rat hippocampus: serial electron microscopy with reference to their biophysical characteristics. The Journal of Neuroscience : The Official Journal of the Society for Neuroscience 9(8), 29822997. https://doi.org/10.1523/JNEUROSCI.09-08-02982.1989CrossRefGoogle ScholarPubMed
Harrison, NJ, and Lummis, SCR (2006a) Locating the carboxylate group of GABA in the homomeric rho GABA(A) receptor ligand-binding pocket. The Journal of Biological Chemistry 281(34), 2445524461. https://doi.org/10.1074/jbc.M601775200CrossRefGoogle ScholarPubMed
Harrison, NJ, and Lummis, SCR (2006b) Molecular modeling of the GABA(C) receptor ligand-binding domain. Journal of Molecular Modeling 12(3), 317324. https://doi.org/10.1007/S00894-005-0034-6CrossRefGoogle ScholarPubMed
Hartvig, L, Lükensmejer, B, Liljefors, T, and Dekermendjian, K (2002) Two conserved arginines in the extracellular N-terminal domain of the GABAA receptor α5 subunit are crucial for receptor function. Journal of Neurochemistry 75(4), 17461753. https://doi.org/10.1046/j.1471-4159.2000.0751746.xCrossRefGoogle Scholar
Hausrat, TJ, Muhia, M, Gerrow, K, Thomas, P, Hirdes, W, Tsukita, S, Heisler, FF, Herich, L, Dubroqua, S, Breiden, P, Feldon, J, Schwarz, JR, Yee, BK, Smar, TG, Triller, A, and Kneussel, M (2015) Radixin regulates synaptic GABAA receptor density and is essential for reversal learning and short-term memory. Nat Commun 6, 6872. https://doi.org/10.1038/ncomms7872CrossRefGoogle ScholarPubMed
Hawkes, AG, Jalali, A, and Colquhoun, D (1992) Asymptotic distributions of apparent open times and shut times in a single channel record allowing for the omission of brief events. Philosophical Transactions of the Royal Society of London. Series B Biological Sciences 337(1282), 383404. https://doi.org/10.1098/rstb.1992.0116Google Scholar
Hernandez, CC, Kong, W, Hu, N, Zhang, Y, Shen, W, Jackson, L, Liu, X, Jiang, Y, and Macdonald, RL (2017) Altered channel conductance states and gating of GABAA receptors by a pore mutation linked to Dravet syndrome. ENeuro 4(1) https://doi.org/10.1523/ENEURO.0251-16.2017CrossRefGoogle ScholarPubMed
Hernandez, CC, and Macdonald, RL (2019) A Structural look at GABA A receptor mutations linked to epilepsy syndromes. In Brain Research (Vol. 1714, pp. 234247) Elsevier B.V. https://doi.org/10.1016/j.brainres.2019.03.004Google Scholar
Hibbs, RE, and Gouaux, E (2011) Principles of activation and permeation in an anion-selective Cys-loop receptor. Nature 474(7349), 5460. https://doi.org/10.1038/nature10139CrossRefGoogle Scholar
Hilf, RJC, and Dutzler, R (2008) X-ray structure of a prokaryotic pentameric ligand-gated ion channel. Nature 452(7185), 375379. https://doi.org/10.1038/nature06717CrossRefGoogle ScholarPubMed
Hilf, RJC, and Dutzler, R (2009) Structure of a potentially open state of a proton-activated pentameric ligand-gated ion channel. Nature 457(7225), 115118. https://doi.org/10.1038/nature07461CrossRefGoogle ScholarPubMed
Holden, JH, and Czajkowski, C (2002) Different residues in the GABAA receptor α1T60-α1K70 region mediate GABA and SR-95531 actions. Journal of Biological Chemistry 277(21), 1878518792. https://doi.org/10.1074/jbc.M111778200CrossRefGoogle ScholarPubMed
Horenstein, J, Wagner, DA, Czajkowski, C, and Akabas, MH (2001) Protein mobility and GABA-induced conformational changes in GABA(A) receptor pore-lining M2 segment. Nature Neuroscience 4(5), 477485. https://doi.org/10.1038/87425CrossRefGoogle ScholarPubMed
Hosie, AM, Clarke, L, da Silva, H, and Smart, TG (2009) Conserved site for neurosteroid modulation of GABAA receptors. Neuropharmacology 56(1), 149154. https://doi.org/10.1016/J.NEUROPHARM.2008.07.050CrossRefGoogle Scholar
Hosie, AM, Wilkins, ME, Da Silva, HMA, and Smart, TG (2006) Endogenous neurosteroids regulate GABAA receptors through two discrete transmembrane sites. Nature 444(7118), 486489. https://doi.org/10.1038/nature05324CrossRefGoogle ScholarPubMed
Howard, RJ (2021) Elephants in the dark: insights and incongruities in pentameric ligand-gated ion channel models. Journal of Molecular Biology 433(17), 167128. https://doi.org/10.1016/j.jmb.2021.167128CrossRefGoogle ScholarPubMed
Huang, X, Shaffer, PL, Ayube, S, Bregman, H, Chen, H, Lehto, SG, Luther, JA, Matson, DJ, McDonough, SI, Michelsen, K, Plant, MH, Schneider, S, Simard, JR, Teffera, Y, Yi, S, Zhang, M, DiMauro, EF, and Gingras, J (2017) Crystal structures of human glycine receptor α3 bound to a novel class of analgesic potentiators. Nature Structural & Molecular Biology 24(2), 108113. https://doi.org/10.1038/nsmb.3329CrossRefGoogle ScholarPubMed
Jadey, S, and Auerbach, A (2012) An integrated catch-and-hold mechanism activates nicotinic acetylcholine receptors. The Journal of General Physiology 140(1), 1728. https://doi.org/10.1085/jgp.201210801CrossRefGoogle ScholarPubMed
Jatczak-Śliwa, M, Kisiel, M, Czyzewska, MM, Brodzki, M, and Mozrzymas, JW (2020) GABAA receptor β2E155 residue located at the agonist-binding site is involved in the receptor gating. Frontiers in Cellular Neuroscience 14 2), 118. https://doi.org/10.3389/fncel.2020.00002CrossRefGoogle ScholarPubMed
Jatczak-Śliwa, M, Terejko, K, Brodzki, M, Michałowski, MAMA, Czyzewska, MMM, Nowicka, JMJM, Andrzejczak, A, Srinivasan, R, and Mozrzymas, JWJW (2018) Distinct Modulation of Spontaneous and GABA-Evoked Gating by Flurazepam Shapes Cross-Talk Between Agonist-Free and Liganded GABAA Receptor Activity. Frontiers in Cellular Neuroscience 12(8), 118. https://doi.org/10.3389/fncel.2018.00237CrossRefGoogle Scholar
Jayakar, SS, Chiara, DC, Zhou, X, Wu, B, Bruzik, KS, Miller, KW, and Cohen, JB (2020) Photoaffinity labeling identifies an intersubunit steroid-binding site in heteromeric GABA type A (GABAA) receptors. The Journal of Biological Chemistry 295(33), 1149511512. https://doi.org/10.1074/jbc.RA120.013452CrossRefGoogle Scholar
Jayakar, SS, Zhou, X, Chiara, DC, Dostalova, Z, Savechenkov, PY, Bruzik, KS, Dailey, WP, Miller, KW, Eckenhoff, RG, and Cohen, JB (2014) Multiple propofol-binding sites in a γ-aminobutyric acid type a receptor (GABAAR) identified using a photoreactive propofol analog. Journal of Biological Chemistry 289(40), 2745627468. https://doi.org/10.1074/jbc.M114.581728CrossRefGoogle Scholar
Jayakar, SS, Zhou, X, Chiara, DC, Jarava-Barrera, C, Savechenkov, PY, Bruzik, KS, Tortosa, M, Miller, KW, and Cohen, JB (2019) Identifying drugs that bind selectively to intersubunit general anesthetic sites in the α1β3γ2 GABA A R transmembrane domain. Molecular Pharmacology 02115(6), https://doi.org/10.1124/mol.118.114975Google Scholar
Jiang, X, Shu, H-J, Krishnan, K, Qian, M, Taylor, AA, Covey, DF, Zorumski, CF, and Mennerick, S (2016) A clickable neurosteroid photolabel reveals selective Golgi compartmentalization with preferential impact on proximal inhibition. Neuropharmacology 108, 193206. https://doi.org/10.1016/j.neuropharm.2016.04.031CrossRefGoogle ScholarPubMed
Jonas, P (1995) Fast application of agonists to isolated membrane patches. In Sakmann, B. & Neher, E. (Eds.), Single-Channel Recording (pp. 231243) Springer US.CrossRefGoogle Scholar
Jones, MV, and Westbrook, GL (1995) Desensitized states prolong GABAA channel responses to brief agonist pulses. Neuron 15(1), 181191. https://doi.org/10.1016/0896-6273(95)90075-6CrossRefGoogle ScholarPubMed
Kaczor, PT, Michałowski, MA, and Mozrzymas, JW (2022) α 1 Proline 277 Residues Regulate GABAAR Gating through M2-M3 Loop Interaction in the Interface Region. ACS Chemical Neuroscience 13(21), 30443056. https://doi.org/10.1021/acschemneuro.2c00401CrossRefGoogle ScholarPubMed
Kaczor, PT, Wolska, AD, and Mozrzymas, JW (2021) α1 subunit histidine 55 at the interface between extracellular and transmembrane domains affects preactivation and desensitization of the GABAA receptor. ACS Chemical Neuroscience. https://doi.org/10.1021/acschemneuro.0c00781CrossRefGoogle Scholar
Kaila, K, Voipio, J, Paalasmaa, P, Pasternack, M, and Deisz, RA (1993) The role of bicarbonate in GABAA receptor-mediated IPSPs of rat neocortical neurones. The Journal of Physiology 464(1), 273289. https://doi.org/10.1113/JPHYSIOL.1993.SP019634CrossRefGoogle ScholarPubMed
Kasaragod, VB, Malinauskas, T, Wahid, AA, Lengyel, J, Knoflach, F, Hardwick, SW, Jones, CF, Chen, WN, Lucas, X, El Omari, K, Chirgadze, DY, Aricescu, AR, Cecere, G, Hernandez, MC, and Miller, PS (2023) The molecular basis of drug selectivity for α5 subunit-containing GABAA receptors. Nature Structural & Molecular Biology 2023, 111. https://doi.org/10.1038/s41594-023-01133-1Google Scholar
Kasaragod, VB, Mortensen, M, Hardwick, SW, Wahid, AA, Dorovykh, V, Chirgadze, DY, Smart, TG, and Miller, PS (2022) Mechanisms of inhibition and activation of extrasynaptic αβ GABAA receptors. Nature 123. https://doi.org/10.1038/s41586-022-04402-zGoogle Scholar
Kash, TL, Dizon, M-JF, Trudell, JR, and Harrison, NL (2004b) Charged Residues in the β2 Subunit Involved in GABAA Receptor Activation. Journal of Biological Chemistry 279(6), 48874893. https://doi.org/10.1074/jbc.M311441200CrossRefGoogle ScholarPubMed
Kash, TL, Jenkins, A, Kelley, JC, Trudell, JR, and Harrison, NL (2003) Coupling of agonist binding to channel gating in the GABA(A) receptor. Nature 421(6920), 272275. https://doi.org/10.1038/nature01280CrossRefGoogle ScholarPubMed
Kash, TL, Kim, T, Trudell, JR, and Harrison, NL (2004a) Evaluation of a proposed mechanism of ligand-gated ion channel activation in the GABA. Neuroscience Letters 371(2–3), 230234. https://doi.org/10.1016/j.neulet.2004.09.002CrossRefGoogle Scholar
Keramidas, A, Kash, TL, and Harrison, NL (2006) The Pre-M1 Segment of the α1 Subunit is a Transduction Element in the Activation of the GABA-A Receptor. The Journal of Physiology 99(Pt 1), 133. https://doi.org/10.1113/jphysiol.2005.102756Google Scholar
Khatri, A, Sedelnikova, A, and Weiss, DS (2009) Structural rearrangements in loop F of the GABA receptor signal ligand binding, not channel activation. Biophysical Journal 96(1), 4555. https://doi.org/10.1016/j.bpj.2008.09.011CrossRefGoogle Scholar
Khayenko, V, and Maric, HM (2019) Targeting GABAAR-Associated Proteins: New Modulators, Labels and Concepts. Frontiers in Cellular Neuroscience 12, 162. https://doi.org/10.3389/fnmol.2019.00162Google Scholar
Kim, JJ, Gharpure, A, Teng, J, Zhuang, Y, Howard, RJ, Zhu, S, Noviello, CM, Walsh, RM, Lindahl, E, and Hibbs, RE (2020) Shared structural mechanisms of general anaesthetics and benzodiazepines. Nature 585(7824), 303308. https://doi.org/10.1038/s41586-020-2654-5CrossRefGoogle ScholarPubMed
Kim, JJ, and Hibbs, RE (2021) Direct Structural Insights into GABAA Receptor Pharmacology. Trends in Biochemical Sciences 46(6), 502517. https://doi.org/10.1016/j.tibs.2021.01.011CrossRefGoogle Scholar
Kisiel, M, Jatczak, M, Brodzki, M, and Mozrzymas, JW (2018) Spontaneous activity, singly bound states and the impact of alpha 1 Phe64 mutation on GABA A R gating in the novel kinetic model based on the single-channel recordings. Neuropharmacology 131, 453474. https://doi.org/10.1016/j.neuropharm.2017.11.030CrossRefGoogle Scholar
Kisiel, M, Jatczak-Śliwa, M, and Mozrzymas, JW (2019) Protons modulate gating of recombinant α1β2γ2 GABAA receptor by affecting desensitization and opening transitions. Neuropharmacology 146(10), 300315. https://doi.org/10.1016/j.neuropharm.2018.10.016CrossRefGoogle ScholarPubMed
Kittler, JT, and Moss, SJ (2003) Modulation of GABAA receptor activity by phosphorylation and receptor trafficking: implications for the efficacy of synaptic inhibition. Current Opinion in Neurobiology 13(3), 341347. https://doi.org/10.1016/s0959-4388(03)00064-3CrossRefGoogle ScholarPubMed
Kloda, JH, and Czajkowski, C. (2007) Agonist-, antagonist-, and benzodiazepine-induced structural changes in the alpha1 Met113-Leu132 region of the GABAA receptor. Molecular Pharmacology 71(2), 483493. https://doi.org/10.1124/MOL.106.028662CrossRefGoogle ScholarPubMed
Kłopotowski, K., Czyżewska, MM, and Mozrzymas, JW (2021) Glycine substitution of α1F64 residue at the loop D of GABAA receptor impairs gating - Implications for importance of binding site-channel gate linker rigidity. Biochemical Pharmacology 192. https://doi.org/10.1016/J.BCP.2021.114668CrossRefGoogle ScholarPubMed
Kłopotowski, K., Michałowski, MA, Gos, M, Mosiądz, D, Czyżewska, MM, and Mozrzymas, JW (2023) Mutation of valine 53 at the interface between extracellular and transmembrane domains of the β2 principal subunit affects the GABAA receptor gating. European Journal of Pharmacology 947, 175664. https://doi.org/10.1016/j.ejphar.2023.175664CrossRefGoogle Scholar
Korol, SV, Jin, Z, Jin, Y, Bhandage, AK, Tengholm, A, Gandasi, NR, Barg, S, Espes, D, Carlsson, PO, Laver, D, and Birnir, B (2018) Functional Characterization of Native, High-Affinity GABAA Receptors in Human Pancreatic β Cells. EBioMedicine 30, 273282. https://doi.org/10.1016/J.EBIOM.2018.03.014CrossRefGoogle ScholarPubMed
Krasowski, MD, Nishikawa, K, Nikolaeva, N, Lin, A, and Harrison, NL (2001) Methionine 286 in transmembrane domain 3 of the GABAA receptor B subunit controls a binding cavity for propofol and other alkylphenol general anesthetics. Neuropharmacology 41(8), 952964. https://doi.org/10.1016/S0028-3908(01)00141-1CrossRefGoogle Scholar
Krishek, B, Moss, S, and Smart, T (1998) Interaction of H+ and Zn2+ on recombinant and native rat neuronal GABAA receptors. Journal of Physiology 507(3), 639652. http://discovery.ucl.ac.uk/22040/CrossRefGoogle Scholar
Krishek, BJ, Amato, A, Connolly, CN, Moss, SJ, T, T. G. S, and Smart, TG (1996) Proton sensitivity of the GABAA receptor is associated with the receptor subunit composition. Journal of Physiology 431443.CrossRefGoogle ScholarPubMed
Krogsgaard-Larsen, P, Falch, E, Schousboe, A, Curtist, DR, and Lodget, D (1980) Piperidine-4-sulphonic Acid, a New Specific GAB A Agonist. Journal of Neurochemistry 34(3), 756759. https://doi.org/10.1111/J.1471-4159.1980.TB11211.XCrossRefGoogle Scholar
Laha, KT, and Tran, PN (2013) Multiple tyrosine residues at the GABA binding pocket influence surface expression and mediate kinetics of the GABAA receptor. Journal of Neurochemistry 124(2), 200209. https://doi.org/10.1111/jnc.12083CrossRefGoogle ScholarPubMed
Laha, KT, and Wagner, DA (2011) A state-dependent salt-bridge interaction exists across the β/α intersubunit interface of the GABAA receptor. Molecular Pharmacology 79(4), 662671. https://doi.org/10.1124/mol.110.068619CrossRefGoogle ScholarPubMed
Lape, R, Colquhoun, D, and Sivilotti, LG (2008) On the nature of partial agonism in the nicotinic receptor superfamily. Nature 454(7205), 722727. https://doi.org/10.1038/nature07139CrossRefGoogle ScholarPubMed
Laverty, D, Desai, R, Uchański, T, Masiulis, S, Stec, WJ, Malinauskas, T, Zivanov, J, Pardon, E, Steyaert, J, Miller, KW, and Aricescu, AR (2019) Cryo-EM structure of the human α1β3γ2 GABAA receptor in a lipid bilayer. Nature 565(7740), 516520. https://doi.org/10.1038/s41586-018-0833-4CrossRefGoogle Scholar
Laverty, D, Thomas, P, Field, M, Andersen, OJ, Gold, MG, Biggin, PC, Gielen, M, and Smart, TG (2017) Crystal structures of a GABA A -receptor chimera reveal new endogenous neurosteroid-binding sites. Nature Structural and Molecular Biology 24(11), 977985. https://doi.org/10.1038/nsmb.3477CrossRefGoogle ScholarPubMed
Legesse, DH, Fan, C, Teng, J, Zhuang, Y, Howard, RJ, Noviello, CM, Lindahl, E, and Hibbs, RE (2023) Structural insights into opposing actions of neurosteroids on GABAA receptors. Nature Communications 14(1), 113. https://doi.org/10.1038/s41467-023-40800-1CrossRefGoogle Scholar
Lema, GMCC, and Auerbach, A (2006) Modes and models of GABAA receptor gating. Journal of Physiology 572(1), 183200. https://doi.org/10.1016/j.yhbeh.2006.02.010CrossRefGoogle Scholar
Lévi, S, Le Roux, N, Eugène, E, and Poncer, JC (2015) Benzodiazepine ligands rapidly influence GABAA receptor diffusion and clustering at hippocampal inhibitory synapses. Neuropharmacology 88, 199208. https://doi.org/10.1016/j.neuropharm.2014.06.002CrossRefGoogle ScholarPubMed
Li, GD, Chiara, DC, Cohen, JB, and Olsen, RW (2010) Numerous classes of general anesthetics inhibit etomidate binding to γ-aminobutyric acid type A (GABAA) receptors. Journal of Biological Chemistry 285(12), 86158620. https://doi.org/10.1074/jbc.M109.074708CrossRefGoogle ScholarPubMed
Li, GD, Chiara, DC, Sawyer, GW, Husain, SS, Olsen, RW, and Cohen, JB (2006) Identification of a GABAA receptor anesthetic binding site at subunit interfaces by photolabeling with an etomidate analog. The Journal of Neuroscience : The Official Journal of the Society for Neuroscience 26(45), 1159911605. https://doi.org/10.1523/JNEUROSCI.3467-06.2006CrossRefGoogle ScholarPubMed
Liu, S, Xu, L, Guan, F, Liu, Y-TT, Cui, Y, Zhang, Q, Zheng, X, Bi, G-QQ, Zhou, ZH, Zhang, X, and Ye, S (2018) Cryo-EM structure of the human α5β3 GABAA receptor. Cell Research 28(9), 958961. https://doi.org/10.1038/s41422-018-0077-8CrossRefGoogle Scholar
Loebrich, S, Bähring, R, Katsuno, T, Tsukita, S, and Kneussel, M (2006) Activated radixin is essential for GABAA receptor alpha5 subunit anchoring at the actin cytoskeleton. EMBO Journal 25(5), 987999. https://doi.org/10.1038/sj.emboj.7600995CrossRefGoogle ScholarPubMed
Lummis, SCR (2009) Locating GABA in GABA receptor binding sites. Biochemical Society Transactions 37(6), 13431346. https://doi.org/10.1042/BST0371343CrossRefGoogle ScholarPubMed
Lummis, SCR, Beene, DL, Harrison, NJ, Lester, HA, and Dougherty, DA (2005) A cation-pi binding interaction with a tyrosine in the binding site of the GABAC receptor. Chemistry & Biology 12(9), 993997. https://doi.org/10.1016/J.CHEMBIOL.2005.06.012CrossRefGoogle ScholarPubMed
Lummis, SCR, Harrison, NJ, Wang, J, Ashby, JA, Millen, KS, Beene, DL, and Dougherty, DA (2012) Multiple Tyrosine Residues Contribute to GABA Binding in the GABA(C) Receptor Binding Pocket. ACS Chemical Neuroscience 3(3), 186192. https://doi.org/10.1021/CN200103NCrossRefGoogle ScholarPubMed
Lummis, SCR, McGonigle, I, Ashby, JA, and Dougherty, DA. (2011) Two amino acid residues contribute to a cation-π binding interaction in the binding site of an insect GABA receptor. The Journal of Neuroscience: The Official Journal of the Society for Neuroscience 31(34), 1237112376. https://doi.org/10.1523/JNEUROSCI.1610-11.2011CrossRefGoogle ScholarPubMed
Maldifassi, MC, Baur, R, and Sigel, E (2016) Functional sites involved in modulation of the GABAA receptor channel by the intravenous anesthetics propofol, etomidate and pentobarbital. Neuropharmacology 105, 207214. https://doi.org/10.1016/j.neuropharm.2016.01.003CrossRefGoogle ScholarPubMed
Maramai, S, Benchekroun, M, Ward, SE, and Atack, JR (2020) Subtype Selective γ-Aminobutyric Acid Type A Receptor (GABAAR) Modulators Acting at the Benzodiazepine Binding Site: An Update. Journal of Medicinal Chemistry 63(7), 34253446. https://doi.org/10.1021/acs.jmedchem.9b01312CrossRefGoogle ScholarPubMed
Maric, D, Maric, I, Wen, X, Fritschy, JM, Sieghart, W, Barker, JL, and Serafini, R (1999) GABAA receptor subunit composition and functional properties of Cl- channels with differential sensitivity to zolpidem in embryonic rat hippocampal cells. Journal of Neuroscience 19(12), 49214937. papers://9648de94-53eb-41f8-8282-2a8285ba1a70/Paper/p1508CrossRefGoogle ScholarPubMed
Marsden, KC, Beattie, JB, Friedenthal, J, and Carroll, RC (2007) NMDA receptor activation potentiates inhibitory transmission through GABA receptor-associated protein-dependent exocytosis of GABAAReceptors. The Journal of Neuroscience 27(52), 1432614337. https://doi.org/10.1523/jneurosci.4433-07.2007CrossRefGoogle Scholar
Marsden, KC, Shemesh, A, Bayer, KU, and Carroll, RC (2010) Selective translocation of Ca2+/calmodulin protein kinase IIalpha (CaMKIIalpha) to inhibitory synapses. Proceedings of the National Academy of Sciences of the United States of America 107(47), 2055920564. https://doi.org/10.1073/pnas.1010346107CrossRefGoogle ScholarPubMed
Masiulis, S, Desai, R, Uchański, T, Serna Martin, I, Laverty, D, Karia, D, Malinauskas, T, Zivanov, J, Pardon, E, Kotecha, A, Steyaert, J, Miller, KW, and Aricescu, AR (2019) GABAA receptor signalling mechanisms revealed by structural pharmacology. Nature 565(7740), 454459. https://doi.org/10.1038/s41586-018-0832-5CrossRefGoogle ScholarPubMed
Maynard, SA, and Triller, A (2019) Inhibitory Receptor Diffusion Dynamics. Frontiers in Molecular Neuroscience 12, 313. https://doi.org/10.3389/fnmol.2019.00313CrossRefGoogle ScholarPubMed
McKernan, RM, and Whiting, PJ (1996) Which GABAA-receptor subtypes really occur in the brain? Trends in Neurosciences 19(4), 139143. https://doi.org/10.1016/S0166-2236(96)80023-3CrossRefGoogle ScholarPubMed
Melis, C, Lummis, SCR, and Molteni, C (2008) Molecular dynamics simulations of GABA binding to the GABAC receptor: the role of Arg104. Biophysical Journal 95(9), 41154123. https://doi.org/10.1529/BIOPHYSJ.107.127589CrossRefGoogle Scholar
Mercado, J (2006) Charged Residues in the 1 and beta2 Pre-M1 Regions Involved in GABAA Receptor Activation. Journal of Neuroscience 26(7), 20312040. https://doi.org/10.1523/JNEUROSCI.4555-05.2006CrossRefGoogle Scholar
Merlaud, Z, Marques, X, Russeau, M, Saade, U, Tostain, M, Moutkine, I, Gielen, M, Corringer, PJ, and Lévi, S (2022) Conformational state-dependent regulation of GABAA receptor diffusion and subsynaptic domains. IScience 25(11) https://doi.org/10.1016/j.isci.2022.105467CrossRefGoogle ScholarPubMed
Michałowski, MA, Czyżewska, MM, Iżykowska, I, and Mozrzymas, JW (2021) The β2 subunit E155 residue as a proton sensor at the binding site on GABA type A receptors. European Journal of Pharmacology 906(12), 174293. https://doi.org/10.1016/j.ejphar.2021.174293CrossRefGoogle ScholarPubMed
Michałowski, MA, Kraszewski, S, and Mozrzymas, JW (2017) Binding site opening by loop C shift and chloride ion-pore interaction in the GABA A receptor model. Physical Chemistry Chemical Physics 19(21), 1366413678. https://doi.org/10.1039/C7CP00582BCrossRefGoogle Scholar
Mihalik, B, Pálvölgyi, A, Bogár, F, Megyeri, K, Ling, I, Barkóczy, J, Bartha, F, Martinek, TA, Gacsályi, I, and Antoni, FA (2017) Loop-F of the α-subunit determines the pharmacologic profile of novel competitive inhibitors of GABAA receptors. European Journal of Pharmacology 798(11), 129136. https://doi.org/10.1016/j.ejphar.2017.01.033CrossRefGoogle Scholar
Miko, A, Werby, E, Sun, H, Healey, J, and Zhang, L (2004) A TM2 residue in the β1 subunit determines spontaneous opening of homomeric and heteromeric γ-aminobutyric acid-gated ion channels. Journal of Biological Chemistry 279(22), 2283322840. https://doi.org/10.1074/jbc.M402577200CrossRefGoogle ScholarPubMed
Miller, C (1989) Genetic manipulation of ion channels: A new approach to structure and mechanism. Neuron 2(3), 11951205. https://doi.org/10.1016/0896-6273(89)90304-8CrossRefGoogle ScholarPubMed
Miller, PS, and Aricescu, AR (2014) Crystal structure of a human GABAA receptor. Nature 512(7514), 270275. https://doi.org/10.1038/nature13293CrossRefGoogle ScholarPubMed
Miller, PS, Scott, S, Masiulis, S, De Colibus, L, Pardon, E, Steyaert, J, and Aricescu, AR (2017) Structural basis for GABA A receptor potentiation by neurosteroids. Nature Structural and Molecular Biology 24(11), 986992. https://doi.org/10.1038/nsmb.3484CrossRefGoogle ScholarPubMed
Mohamad, FH, and Has, ATC (2019) The α5-Containing GABAA Receptors—a Brief Summary. Journal of Molecular Neuroscience 67(2), 343351. https://doi.org/10.1007/s12031-018-1246-4CrossRefGoogle ScholarPubMed
Mortensen, M, Wafford, KA, Wingrove, P, and Ebert, B (2003) Pharmacology of GABAA receptors exhibiting different levels of spontaneous activity. European Journal of Pharmacology 476(1–2), 1724. https://doi.org/10.1016/S0014-2999(03)02125-3CrossRefGoogle ScholarPubMed
Mortensen, M, Xu, Y, Shehata, MA, Krall, J, Ernst, M, Frølund, B, and Smart, TG (2023) Pregnenolone sulfate analogues differentially modulate GABAA receptor closed/desensitised states. British Journal of Pharmacology https://doi.org/10.1111/bph.16143CrossRefGoogle ScholarPubMed
Mozrzymas, JW (2004) Dynamism of GABAA receptor activation shapes the “personality” of inhibitory synapses. Neuropharmacology 47(7), 945960. https://doi.org/10.1016/j.neuropharm.2004.07.003CrossRefGoogle ScholarPubMed
Mozrzymas, JW, Barberis, A, Mercik, K, and Zarnowska, ED (2003) Binding Sites, Singly Bound States, and Conformation Coupling Shape GABA-Evoked Currents. Journal of Neurophysiology 89(2), 871883. https://doi.org/10.1152/jn.00951.2002CrossRefGoogle ScholarPubMed
Mozrzymas, JW, Barberis, A, Michalak, K, and Cherubini, E (1999) Chlorpromazine inhibits miniature GABAergic currents by reducing the binding and by increasing the unbinding rate of GABA(A) receptors. Journal of Neuroscience 19(7), 24742488. https://doi.org/10.1523/jneurosci.19-07-02474.1999CrossRefGoogle Scholar
Mozrzymas, JW, Barberis, A, and Vicini, S (2007b) GABAergic currents in RT and VB thalamic nuclei follow kinetic pattern of alpha3- and alpha1-subunit-containing GABAA receptors. The European Journal of Neuroscience 26(3), 657665. https://doi.org/10.1111/J.1460-9568.2007.05693.XCrossRefGoogle ScholarPubMed
Mozrzymas, JW, Wójtowicz, T, Piast, M, Lebida, K, Wyrembek, P, and Mercik, K (2007a) GABA transient sets the susceptibility of mIPSCs to modulation by benzodiazepine receptor agonists in rat hippocampal neurons. The Journal of Physiology 585( 1), 2946. https://doi.org/10.1113/jphysiol.2007.143602CrossRefGoogle ScholarPubMed
Mukhtasimova, N, Lee, WY, Wang, H-L, and Sine, SM (2009) Detection and trapping of intermediate states priming nicotinic receptor channel opening. Nature 459(7245), 451454. https://doi.org/10.1038/nature07923CrossRefGoogle ScholarPubMed
Muroi, Y, Czajkowski, C, and Jackson, MB (2006) Local and global ligand-induced changes in the structure of the GABA A receptor. Biochemistry 45(23), 70137022. https://doi.org/10.1021/bi060222vCrossRefGoogle ScholarPubMed
Muroi, Y, Theusch, CM, Czajkowski, C, and Jackson, MB (2009) Distinct structural changes in the GABAA receptor elicited by pentobarbital and GABA. Biophysical Journal 96(2), 499509. https://doi.org/10.1016/j.bpj.2008.09.037CrossRefGoogle ScholarPubMed
Naffaa, MM, Absalom, N, Raja Solomon, V, Chebib, M, Hibbs, DE, and Hanrahan, JR (2016) Investigating the Role of Loop C Hydrophilic Residue ‘T244’ in the Binding Site of ρ1 GABAC Receptors via Site Mutation and Partial Agonism. PLoS ONE 11(5), e0156618. https://doi.org/10.1371/JOURNAL.PONE.0156618CrossRefGoogle Scholar
Naffaa, MM, Hibbs, DE, Chebib, M, and Hanrahan, JR (2022) Roles of hydrophilic residues in GABA binding site of GABA-ρ1 receptor explain the addition/inhibition effects of competitive ligands. Neurochemistry International 153, 105258. https://doi.org/10.1016/J.NEUINT.2021.105258CrossRefGoogle ScholarPubMed
Nakamura, Y, Darnieder, LM, Deeb, TZ, and Moss, SJ (2015) Regulation of GABA A Rs by phosphorylation. Advances in Pharmacology 72, 97-146. https://doi.org/10.1016/bs.apha.2014.11.008CrossRefGoogle Scholar
Nakane, T, Kotecha, A, Sente, A, Mcmullan, G, Masiulis, S, Brown, PMGE, Grigoras, IT, Malinauskaite, L, Malinauskas, T, Miehling, J, Uchański, T, Yu, L, Karia, D, Pechnikova, EV, De Jong, E, Keizer, J, Bischoff, M, Mccormack, J, Tiemeijer, P, … Scheres, HW (2020) Single-particle cryo-EM at atomic resolution. Nature 587. https://doi.org/10.1038/s41586-020-2829-0CrossRefGoogle ScholarPubMed
Newell, JG, and Czajkowski, C (2003) The GABAA receptor α1 subunit Pro174–Asp191 segment is involved in GABA binding and channel gating. Journal of Biological Chemistry 278(15), 1316613172. https://doi.org/10.1074/jbc.M211905200CrossRefGoogle ScholarPubMed
Newell, JG, McDevitt, RA, and Czajkowski, C (2004) Mutation of glutamate 155 of the GABAA receptor 2 subunit produces a spontaneously open channel: a trigger for channel activation. Journal of Neuroscience 24(50), 1122611235. https://doi.org/10.1523/JNEUROSCI.3746-04.2004CrossRefGoogle Scholar
Nors, JW, Gupta, S, and Goldschen-Ohm, MP (2021) A critical residue in the α1M2–M3 linker regulating mammalian GABAA receptor pore gating by diazepam. ELife 10, 121. https://doi.org/10.7554/eLife.64400CrossRefGoogle ScholarPubMed
Nourmahnad, A, Stern, AT, Hotta, M, Stewart, DS, Ziemba, AM, Szabo, A, and Forman, SA (2016) Tryptophan and cysteine mutations in M1 helices of α1β3γ2L γ-aminobutyric acid type A receptors indicate distinct intersubunit sites for four intravenous anesthetics and one orphan site. Anesthesiology 125(6), 11441158. https://doi.org/10.1097/ALN.0000000000001390CrossRefGoogle ScholarPubMed
Noviello, CM, Gharpure, A, Mukhtasimova, N, Cabuco, R, Baxter, L, Borek, D, Sine, SM, and Hibbs, RE (2021) Structure and gating mechanism of the α7 nicotinic acetylcholine receptor. Cell 184(8), 21212134.e13. https://doi.org/10.1016/j.cell.2021.02.049CrossRefGoogle ScholarPubMed
Noviello, CM, Kreye, J, Teng, J, Prüss, H, and Hibbs, RE (2022) Structural mechanisms of GABAA receptor autoimmune encephalitis. Cell 185(14), 24692477.e13. https://doi.org/10.1016/J.CELL.2022.06.025CrossRefGoogle ScholarPubMed
Olsen, RW (2018) GABAAreceptor: positive and negative allosteric modulators. Neuropharmacology 136, 1022. https://doi.org/10.1016/j.neuropharm.2018.01.036CrossRefGoogle Scholar
Padgett, CL, Hanek, AP, Lester, HA, Dougherty, DA, and Lummis, SCRR (2007) Unnatural amino acid mutagenesis of the GABAA receptor binding site residues reveals a novel cation- interaction between GABA and beta 2Tyr97. The Journal of Neuroscience 27(4), 886892. https://doi.org/10.1523/JNEUROSCI.4791-06.2007CrossRefGoogle ScholarPubMed
Pálvölgyi, A, Móricz, K, Pataki, Á, Mihalik, B, Gigler, G, Megyeri, K, Udvari, S, Gacsályi, I, and Antoni, FA (2018) Loop F of the GABAA receptor alpha subunit governs GABA potency. Neuropharmacology 128, 408415. https://doi.org/10.1016/j.neuropharm.2017.10.042CrossRefGoogle Scholar
Patel, B, Mortensen, M, and Smart, TG (2014) Stoichiometry of δ subunit containing GABAA receptors. British Journal of Pharmacology 171(4), 985994. https://doi.org/10.1111/bph.12514CrossRefGoogle Scholar
Petrini, EM, Ravasenga, T, Hausrat, TJ, Iurilli, G, Olcese, U, Racine, V, Sibarita, JB, Jacob, TC, Moss, SJ, Benfenati, F, Medini, P, Kneussel, M, and Barberis, A (2014) Synaptic recruitment of gephyrin regulates surface GABAA receptor dynamics for the expression of inhibitory LTP. Nat Commun 5, 3921. https://doi.org/10.1038/ncomms4921CrossRefGoogle ScholarPubMed
Phulera, S, Zhu, H, Yu, J, Claxton, DP, Yoder, N, Yoshioka, C, and Gouaux, E (2018) Cryo-EM structure of the benzodiazepine-sensitive α1β1γ2S tri-heteromeric GABAA receptor in complex with GABA. ELife 7, 121. https://doi.org/10.7554/eLife.39383CrossRefGoogle ScholarPubMed
Pierce, SR, Germann, AL, Evers, AS, Steinbach, JH, and Akk, G (2020) Reduced activation of the synaptic-type GABAA receptor following prolonged exposure to low concentrations of agonists: relationship between tonic activity and desensitization. Molecular Pharmacology 98(6), 762769. https://doi.org/10.1124/molpharm.120.000088CrossRefGoogle ScholarPubMed
Pierce, SR, Germann, AL, Steinbach, JH, and Akk, G (2022) The Sulfated steroids pregnenolone sulfate and dehydroepiandrosterone sulfate inhibit the α1β3γ2L GABAA receptor by stabilizing a novel nonconducting state. Molecular Pharmacology 101(2), 6877. https://doi.org/10.1124/MOLPHARM.121.000385CrossRefGoogle ScholarPubMed
Pizzarelli, R, Griguoli, M, Zacchi, P, Petrini, EM, Barberis, A, Cattaneo, A, and Cherubini, E (2020) Tuning GABAergic inhibition: gephyrin molecular organization and functions. Neuroscience 439, 125136. https://doi.org/10.1016/j.neuroscience.2019.07.036CrossRefGoogle Scholar
Puthenkalam, R, Hieckel, M, Simeone, X, Suwattanasophon, C, Feldbauer, Rv, Ecker, GF, and Ernst, M (2016) Structural studies of GABAA receptor binding sites: which experimental structure tells us what? Frontiers in Molecular Neuroscience 9 6, 120. https://doi.org/10.3389/fnmol.2016.00044CrossRefGoogle Scholar
Pytel, M, Mercik, K, and Mozrzymas, JW (2003) Resolving the ionotropic receptor kinetics and modulation in the time scale of synaptic transmission. Cellular and Molecular Biology Letters 8, 231241.Google ScholarPubMed
Riva, I, Eibl, C, Volkmer, R, Carbone, AL, and Plested, AJ (2017) Control of AMPA receptor activity by the extracellular loops of auxiliary proteins. Elife 6. https://doi.org/10.7554/eLife.28680CrossRefGoogle ScholarPubMed
Rivera, C, Voipio, J, and Kaila, K (2005) Two developmental switches in GABAergic signalling: the K+-Cl- cotransporter KCC2 and carbonic anhydrase CAVII. The Journal of Physiology 562(Pt 1), 2736. https://doi.org/10.1113/JPHYSIOL.2004.077495CrossRefGoogle ScholarPubMed
Rosen, A, Bali, M, Horenstein, J, and Akabas, MH (2007) Channel opening by anesthetics and GABA induces similar changes in the GABAA receptor M2 segment. Biophysical Journal 92(9), 31303139. https://doi.org/10.1529/biophysj.106.094490CrossRefGoogle ScholarPubMed
Rudolph, U, and Möhler, H (2006) GABA-based therapeutic approaches: GABAA receptor subtype functions. Current Opinion in Pharmacology 6(1), 1823. https://doi.org/10.1016/j.coph.2005.10.003CrossRefGoogle ScholarPubMed
Sachidanandan, D, and Bera, AK (2015) Inhibition of the GABAA receptor by sulfated neurosteroids: a mechanistic comparison study between pregnenolone sulfate and dehydroepiandrosterone sulfate. Journal of Molecular Neuroscience 56(4), 868877. https://doi.org/10.1007/s12031-015-0527-4CrossRefGoogle ScholarPubMed
Saliba, RS, Kretschmannova, K, and Moss, SJ (2012) Activity-dependent phosphorylation of GABAA receptors regulates receptor insertion and tonic current. EMBO Journal 31(13), 29372951. https://doi.org/10.1038/emboj.2012.109CrossRefGoogle ScholarPubMed
Sallard, E, Letourneur, D, and Legendre, P (2021) Electrophysiology of ionotropic GABA receptors. Cellular and Molecular Life Sciences 78(13), 53415370. https://doi.org/10.1007/s00018-021-03846-2CrossRefGoogle ScholarPubMed
Sauguet, L, Shahsavar, A, Poitevin, F, Huon, C, Menny, A, Nemecz, A, Haouz, A, Changeux, J-P, Corringer, P-J, and Delarue, M (2014) Crystal structures of a pentameric ligand-gated ion channel provide a mechanism for activation. Proceedings of the National Academy of Sciences 111(3), 966971. https://doi.org/10.1073/pnas.1314997111CrossRefGoogle ScholarPubMed
Scheller, M, and Forman, SA (2002) Coupled and Uncoupled Gating and Desensitization Effects by Pore Domain Mutations in GABA A Receptors.CrossRefGoogle Scholar
Scimemi, A, and Beato, M (2009) Determining the neurotransmitter concentration profile at active synapses. Molecular Neurobiology 40(3), 289306. https://doi.org/10.1007/S12035-009-8087-7CrossRefGoogle ScholarPubMed
Scott, S, and Aricescu, AR (2019) A structural perspective on GABA A receptor pharmacology. Current Opinion in Structural Biology 54, 189197. https://doi.org/10.1016/j.sbi.2019.03.023CrossRefGoogle Scholar
Sedelnikova, A, Erkkila, BE, Harris, H, Zakharkin, SO, and Weiss, DS (2006) Stoichiometry of a pore mutation that abolishes picrotoxin-mediated antagonism of the GABAA receptor. The Journal of Physiology 577(Pt 2), 569577. https://doi.org/10.1113/JPHYSIOL.2006.120287CrossRefGoogle ScholarPubMed
Seljeset, S, Bright, DP, Thomas, P, and Smart, TG (2018) Probing GABAAreceptors with inhibitory neurosteroids. Neuropharmacology 136, 2336. https://doi.org/10.1016/j.neuropharm.2018.02.008CrossRefGoogle Scholar
Senju, Y, and Tsai, FC (2022) A biophysical perspective of the regulatory mechanisms of ezrin/radixin/moesin proteins. Biophysical Reviews 14(1), 199208. https://doi.org/10.1007/s12551-021-00928-0CrossRefGoogle ScholarPubMed
Sente, A, Desai, R, Naydenova, K, Malinauskas, T, Jounaidi, Y, Miehling, J, Zhou, X, Masiulis, S, Hardwick, SW, Chirgadze, DY, Miller, KW, and Aricescu, AR (2022) Differential assembly diversifies GABAA receptor structures and signalling. Nature 604(7904), 190194. https://doi.org/10.1038/s41586-022-04517-3CrossRefGoogle ScholarPubMed
Sexton, CA, Penzinger, R, Mortensen, M, Bright, DP, and Smart, TG (2021) Structural determinants and regulation of spontaneous activity in GABAA receptors. Nature Communications 12(1), 5457. https://doi.org/10.1038/s41467-021-25633-0CrossRefGoogle Scholar
Sharkey, LM, and Czajkowski, C (2008) Individually monitoring ligand-induced changes in the structure of the GABA A receptor at benzodiazepine binding site and non – binding-site interfaces. Drugs 74(1), 203212. https://doi.org/10.1124/mol.108.044891.bindingGoogle ScholarPubMed
Shin, DJ, Germann, AL, Covey, DF, Steinbach, JH, and Akk, G (2019) Analysis of GABAA receptor activation by combinations of agonists acting at the same or distinct binding sites. Molecular Pharmacology 95(1), 7081. https://doi.org/10.1124/MOL.118.113464CrossRefGoogle ScholarPubMed
Shin, DJ, Germann, AL, Johnson, AD, Forman, SA, Steinbach, JH, and Akk, G (2018) Propofol is an allosteric agonist with multiple binding sites on concatemeric ternary GABA A receptors. Molecular Pharmacology 93(2), 178189. https://doi.org/10.1124/mol.117.110403CrossRefGoogle Scholar
Sieghart, W, and Savić, MM (2018) International union of basic and clinical pharmacology. CVI: GABAA receptor subtype- and function-selective ligands: key issues in translation to humans. Pharmacological Reviews 70(4), 836878. https://doi.org/10.1124/pr.117.014449CrossRefGoogle ScholarPubMed
Siegwart, R, Jurd, R, and Rudolph, U (2002) Molecular determinants for the action of general anesthetics at recombinant alpha(2)beta(3)gamma(2)gamma-aminobutyric acid(A) receptors. Journal of Neurochemistry 80(1), 140148. https://doi.org/10.1046/J.0022-3042.2001.00682.XCrossRefGoogle Scholar
Siegwart, R, Krähenbühl, K, Lambert, S, and Rudolph, U (2003) Mutational analysis of molecular requirements for the actions of general anaesthetics at the gamma-aminobutyric acidA receptor subtype alpha1beta2gamma2. BMC Pharmacology 3. https://doi.org/10.1186/1471-2210-3-13CrossRefGoogle ScholarPubMed
Sigel, E, Baur, R, Kellenberger, S, Malherbel, P, Malherbe, P, and Malherbel, P (1992) Point mutations affecting antagonist affinity and agonist dependent gating of GABAA receptor channels. The EMBO Journal 11(6), 20172023. http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=556666&tool=pmcentrez&rendertype=abstractCrossRefGoogle ScholarPubMed
Sigel, E, and Ernst, M (2018) The benzodiazepine binding sites of GABAA receptors. Trends in Pharmacological Sciences 39(7), 659671. https://doi.org/10.1016/j.tips.2018.03.006CrossRefGoogle Scholar
Steinbach, JH, and Akk, G (2019) Applying the Monod-Wyman-Changeux allosteric activation model to pseudo–steady-state responses from GABAA receptors. Molecular Pharmacology 95(1), 106119. https://doi.org/10.1124/mol.118.113787CrossRefGoogle Scholar
Stewart, D, Desai, R, Cheng, Q, Liu, A, and Forman, SA (2008) Tryptophan mutations at Azi-etomidate photo-incorporation sites on α1 or β2 subunits enhance gabaa receptor gating and reduce etomidate modulation. Molecular Pharmacology 74(6), 16871695. https://doi.org/10.1124/MOL.108.050500CrossRefGoogle ScholarPubMed
Stewart, DS, Hotta, M, Desai, R, and Forman, SA (2013a) State-dependent etomidate occupancy of its allosteric agonist sites measured in a cysteine-substituted GABAA receptor. Molecular Pharmacology 83(6), 12001208. https://doi.org/10.1124/mol.112.084558CrossRefGoogle Scholar
Stewart, DS, Hotta, M, Li, GD, Desai, R, Chiara, DC, Olsen, RW, and Forman, SA (2013b) Cysteine substitutions define etomidate binding and gating linkages in the α-M1 domain of γ-aminobutyric acid type A (GABAA) receptors. The Journal of Biological Chemistry 288(42), 3037330386. https://doi.org/10.1074/JBC.M113.494583CrossRefGoogle ScholarPubMed
Stewart, DS, Pierce, DW, Hotta, M, Stern, AT, and Forman, SA (2014) Mutations at beta N265 in c-aminobutyric acid type a receptors alter both binding affinity and efficacy of potent anesthetics. PLoS ONE 9(10), 112. https://doi.org/10.1371/journal.pone.0111470CrossRefGoogle Scholar
Sugasawa, Y, Bracamontes, JR, Krishnan, K, Covey, DF, Reichert, DE, Akk, G, Chen, Q, Tang, P, Evers, AS, and Cheng, WWL (2019) The molecular determinants of neurosteroid binding in the GABA(A) receptor. The Journal of Steroid Biochemistry and Molecular Biology 192, 105383. https://doi.org/10.1016/J.JSBMB.2019.105383CrossRefGoogle ScholarPubMed
Sugasawa, Y, Cheng, WW, Bracamontes, JR, Chen, Z-W, Wang, L, Germann, AL, Pierce, SR, Senneff, TC, Krishnan, K, Reichert, DE, Covey, DF, Akk, G, and Evers, AS (2020) Site-specific effects of neurosteroids on GABAA receptor activation and desensitization. ELife 9, 132. https://doi.org/10.7554/eLife.55331CrossRefGoogle ScholarPubMed
Sun, C, Zhu, H, Clark, S, and Gouaux, E (2023) Cryo-EM structures reveal native GABAA receptor assemblies and pharmacology. Nature 622(7981), 195201. https://doi.org/10.1038/s41586-023-06556-wCrossRefGoogle ScholarPubMed
Szabo, A, Nourmahnad, A, Halpin, E, and Forman, SA (2019) Monod-Wyman-changeux allosteric shift analysis in mutant α1β3γ2L GABAA receptors indicates selectivity and cross-talk among intersubunit transmembrane anesthetic sites. Molecular Pharmacology 95(4), mol.118.115048. https://doi.org/10.1124/mol.118.115048CrossRefGoogle Scholar
Szczot, M, Kisiel, M, Czyzewska, MM, and Mozrzymas, JW (2014) α1F64 Residue at GABA(A) receptor binding site is involved in gating by influencing the receptor flipping transitions. The Journal of Neuroscience: The Official Journal of the Society for Neuroscience 34(9), 31933209. https://doi.org/10.1523/JNEUROSCI.2533-13.2014CrossRefGoogle ScholarPubMed
Tan, KR, Rudolph, U, and Lüscher, C (2011) Hooked on benzodiazepines: GABAAreceptor subtypes and addiction. Trends in Neurosciences 34(4), 188197. https://doi.org/10.1016/j.tins.2011.01.004CrossRefGoogle Scholar
Tateiwa, H, Chintala, SM, Chen, Z, Wang, L, Amtashar, F, Bracamontes, J, Germann, AL, Pierce, SR, Covey, DF, Akk, G, and Evers, AS (2023) The mechanism of enantioselective neurosteroid actions on GABAA receptors. Biomolecules 13(2) https://doi.org/10.3390/BIOM13020341CrossRefGoogle Scholar
Terejko, K, Kaczor, PT, Michałowski, MA, Dąbrowska, A, and Mozrzymas, JW (2020) The C loop at the orthosteric binding site is critically involved in GABAA receptor gating. Neuropharmacology 166(11), 107903. https://doi.org/10.1016/j.neuropharm.2019.107903CrossRefGoogle Scholar
Terejko, K, Michałowski, MA, Dominik, A, Andrzejczak, A, and Mozrzymas, JW (2021a) Interaction between GABAA receptor α1 and β2 subunits at the N-terminal peripheral regions is crucial for receptor binding and gating. Biochemical Pharmacology 183(9), 114338. https://doi.org/10.1016/j.bcp.2020.114338CrossRefGoogle ScholarPubMed
Terejko, K, Michałowski, MA, Iżykowska, I, Dominik, A, Brzóstowicz, A, and Mozrzymas, JW (2021b) Mutations at the M2 and M3 transmembrane helices of the GABA A Rs α 1 and β 2 subunits affect primarily late gating transitions including opening/closing and desensitization. ACS Chemical Neuroscience 12(13), 24212436. https://doi.org/10.1021/acschemneuro.1c00151CrossRefGoogle Scholar
Tierney, ML, Birnir, B, Pillai, NP, Clements, JD, Howitt, SM, Cox, GB, and Gage, PW (1996) Effects of mutating leucine to threonine in the M2 segment of α1 and β1 subunits of GABAA α1β1 receptors. Journal Membrane Biology 154 (1), 11–21.CrossRefGoogle Scholar
Topf, N, Jenkins, A, Baron, N, and Harrison, NL (2003) Effects of Isoflurane on-aminobutyric acid type A receptors activated by full and partial agonists. Anesthesiology 98, 306311.CrossRefGoogle ScholarPubMed
Toyoda, H, Saito, M, Sato, H, Tanaka, T, Ogawa, T, Yatani, H, Kawano, T, Kanematsu, T, Hirata, M, and Kang, Y (2015) Enhanced desensitization followed by unusual resensitization in GABAA receptors in phospholipase C-related catalytically inactive protein-1/2 double-knockout mice. Pflugers Archiv: European Journal of Physiology 467(2), 267284. https://doi.org/10.1007/S00424-014-1511-5CrossRefGoogle ScholarPubMed
Tran, PN, Laha, KT, and Wagner, DA (2011) A tight coupling between β 2Y97 and β 2F200 of the GABA A receptor mediates GABA binding. Journal of Neurochemistry 119(2), 283293. https://doi.org/10.1111/j.1471-4159.2011.07409.xCrossRefGoogle ScholarPubMed
Uchański, T, Masiulis, S, Fischer, B, Kalichuk, V, López-Sánchez, U, Zarkadas, E, Weckener, M, Sente, A, Ward, P, Wohlkönig, A, Zögg, T, Remaut, H, Naismith, JH, Nury, H, Vranken, W, Aricescu, AR, Pardon, E, and Steyaert, J (2021) Megabodies expand the nanobody toolkit for protein structure determination by single-particle cryo-EM. Nature Methods 18(1), 6068. https://doi.org/10.1038/s41592-020-01001-6CrossRefGoogle ScholarPubMed
Ueno, S, Lin, A, Nikolaeva, N, Trudell, JR, John Mihic, S, Adron Harris, R, and Harrison, NL (2000) Tryptophan scanning mutagenesis in TM2 of the GABA A receptor a subunit: e€ects on channel gating and regulation by ethanol 1,6. British Journal of Pharmacology 131, 296302.CrossRefGoogle Scholar
Unwin, N (2005) Refined structure of the nicotinic acetylcholine receptor at 4 A resolution. Journal of Molecular Biology 346(4), 967989. https://doi.org/10.1016/j.jmb.2004.12.031CrossRefGoogle Scholar
Venkatachalan, SP, and Czajkowski, C (2008) A conserved salt bridge critical for GABA(A) receptor function and loop C dynamics. Proceedings of the National Academy of Sciences of the United States of America 105(36), 1360413609. https://doi.org/10.1073/pnas.0801854105CrossRefGoogle ScholarPubMed
Venkatachalan, SP, and Czajkowski, C (2012) Structural link between gamma-aminobutyric acid type A (GABAA) receptor agonist binding site and inner beta-sheet governs channel activation and allosteric drug modulation. Journal of Biological Chemistry 287(9), 67146724. https://doi.org/10.1074/jbc.M111.316836CrossRefGoogle ScholarPubMed
Wagner, DA, Czajkowski, C, and Jones, MV (2004) An arginine involved in GABA binding and unbinding but not gating of the GABA(A) receptor. The Journal of Neuroscience: The Official Journal of the Society for Neuroscience 24(11), 27332741. https://doi.org/10.1523/JNEUROSCI.4316-03.2004CrossRefGoogle Scholar
Wagner, DA, Goldschen-Ohm, MP, Hales, TG, and Jones, MV (2005) Kinetics and spontaneous open probability conferred by the epsilon subunit of the GABAA receptor. The Journal of Neuroscience : The Official Journal of the Society for Neuroscience 25(45), 1046210468. https://doi.org/10.1523/JNEUROSCI.1658-05.2005CrossRefGoogle ScholarPubMed
Walker, MC, and Semyanov, A (2008) Regulation of excitability by extrasynaptic GABA(A) receptors. Results and Problems in Cell Differentiation 44, 2948. https://doi.org/10.1007/400_2007_030CrossRefGoogle ScholarPubMed
Wang, Q, Pless, SA, and Lynch, JW (2010) Ligand- and subunit-specific conformational changes in the ligand-binding domain and the TM2-TM3 linker of {alpha}1 {beta}2 {gamma}2 GABAA receptors. The Journal of Biological Chemistry 285(51), 4037340386. https://doi.org/10.1074/JBC.M110.161513CrossRefGoogle ScholarPubMed
Wermuth, CG, and Bizière, K (1986) Pyridazinyl-GABA derivatives: a new class of synthetic GABAA antagonists. Trends in Pharmacological Sciences 7(C), 421424. https://doi.org/10.1016/0165-6147(86)90408-6CrossRefGoogle Scholar
Westh-Hansen, SE, Rasmussen, PB, Hastrup, S, Nabekura, J, Noguchi, K, Akaike, N, Witt, MR and Nielsen, M (1997) Decreased agonist sensitivity of human GABAA receptors by an amino acid variant, isoleucine to valine, in the α1 subunit. European Journal of Pharmacology 329(2–3), 253257. https://doi.org/10.1016/S0014-2999(97)89186-8CrossRefGoogle Scholar
Westh-Hansen, SE, Witt, MR, Dekermendjian, K, Liljefors, T, Rasmussen, PB, and Nielsen, M (1999) Arginine residue 120 of the human GABAA receptor alpha 1, subunit is essential for GABA binding and chloride ion current gating. Neuroreport 10(11), 24172421. https://doi.org/10.1097/00001756-199908020-00036CrossRefGoogle ScholarPubMed
Wiera, G, Brzdąk, P, Lech, AM, Lebida, K, Jabłońska, J, Gmerek, P, and Mozrzymas, JW (2022) Integrins bidirectionally regulate the efficacy of inhibitory synaptic transmission and control GABAergic plasticity. Journal of Neuroscience 42(30), 58305842. https://doi.org/10.1523/jneurosci.1458-21.2022CrossRefGoogle ScholarPubMed
Wiera, G, Lebida, K, Lech, AM, Brzdąk, P, Van Hove, I, De Groef, L, Moons, L, Petrini, EM, Barberis, A, and Mozrzymas, JW (2021) Long-term plasticity of inhibitory synapses in the hippocampus and spatial learning depends on matrix metalloproteinase 3. Cellular and Molecular Life Sciences 78(5), 22792298. https://doi.org/10.1007/s00018-020-03640-6CrossRefGoogle ScholarPubMed
Wilkins, ME, Hosie, AM, and Smart, TG (2002) Identification of a beta subunit TM2 residue mediating proton modulation of GABA type A receptors. The Journal of Neuroscience: The Official Journal of the Society for Neuroscience 22(13), 53285333. https://doi.org/10.1523/JNEUROSCI.22-13-05328.2002CrossRefGoogle ScholarPubMed
Wilkins, ME, Hosie, AM, and Smart, TG (2005) Proton modulation of recombinant GABAA receptors: influence of GABA concentration and the β subunit TM2-TM3 domain. Journal of Physiology 567(2), 365377. https://doi.org/10.1113/jphysiol.2005.088823CrossRefGoogle ScholarPubMed
Wooltorton, JRA, Mcdonald, BJ, Moss, SJ, and Smart, TG (1997) Identification of a Zn2+ binding site on the murine GABAA receptor complex: dependence on the second transmembrane domain of/subunits. Journal of Physiology 505, 633640.CrossRefGoogle ScholarPubMed
Wu, K, Castellano, D, Tian, Q, and Lu, W (2021a) Distinct regulation of tonic GABAergic inhibition by NMDA receptor subtypes. Cell Reports 37(6), 109960. https://doi.org/10.1016/j.celrep.2021.109960CrossRefGoogle ScholarPubMed
Wu, K, Han, W, and Lu, W (2022a) Sleep and wake cycles dynamically modulate hippocampal inhibitory synaptic plasticity. PLoS Biology 20(11) https://doi.org/10.1371/JOURNAL.PBIO.3001812CrossRefGoogle ScholarPubMed
Wu, K, Han, W, Tian, Q, Li, Y, and Lu, W (2021b) Activity- and sleep-dependent regulation of tonic inhibition by Shisa7. Cell Reports 34(12), 108899. https://doi.org/10.1016/j.celrep.2021.108899CrossRefGoogle ScholarPubMed
Wu, K, Shepard, RD, Castellano, D, Han, W, Tian, Q, Dong, L, and Lu, W (2022b) Shisa7 phosphorylation regulates GABAergic transmission and neurodevelopmental behaviors. Neuropsychopharmacology 2022, 111. https://doi.org/10.1038/s41386-022-01334-0Google Scholar
Wyroślak, M, Dobrzański, G, and Mozrzymas, JW (2023) Bidirectional plasticity of GABAergic tonic inhibition in hippocampal somatostatin- and parvalbumin-containing interneurons. Frontiers in Cellular Neuroscience 17, 1193383. https://doi.org/10.3389/fncel.2023.1193383CrossRefGoogle ScholarPubMed
Wyroślak, M, Lebida, K, and Mozrzymas, JW (2021) Induction of inhibitory synaptic plasticity enhances tonic current by increasing the content of α5-subunit containing GABA(A) receptors in hippocampal pyramidal neurons. Neuroscience 467, 3946. https://doi.org/10.1016/j.neuroscience.2021.05.020CrossRefGoogle ScholarPubMed
Xie, HB, Wang, J, Sha, Y, and Cheng, MS (2013) Molecular dynamics investigation of Cl− transport through the closed and open states of the 2α12β2γ2 GABAA receptor. Biophysical Chemistry 180 –181, 19. https://doi.org/10.1016/J.BPC.2013.05.004CrossRefGoogle Scholar
Yip, GMS, Chen, Z-W, Edge, CJ, Smith, EH, Dickinson, R, Hohenester, E, Townsend, RR, Fuchs, K, Sieghart, W, Evers, AS, and Franks, NP (2013) A propofol binding site on mammalian GABAA receptors identified by photolabeling. Nature Chemical Biology 9(11), 715720. https://doi.org/10.1038/nchembio.1340CrossRefGoogle ScholarPubMed
Yu, J, Zhu, H, Lape, R, Greiner, T, Du, J, , W, Sivilotti, L, and Gouaux, E (2021) Mechanism of gating and partial agonist action in the glycine receptor. Cell 184(4), 957968.e21. https://doi.org/10.1016/j.cell.2021.01.026CrossRefGoogle ScholarPubMed
Zacchi, P, Antonelli, R, and Cherubini, E (2014) Gephyrin phosphorylation in the functional organization and plasticity of GABAergic synapses. Frontiers in Cellular Neuroscience 8, 103. https://doi.org/10.3389/fncel.2014.00103CrossRefGoogle ScholarPubMed
Zarkadas, E, Pebay-Peyroula, E, Thompson, MJ, Schoehn, G, Uchański, T, Steyaert, J, Chipot, C, Dehez, F, Baenziger, JE, and Nury, H (2022) Conformational transitions and ligand-binding to a muscle-type nicotinic acetylcholine receptor. Neuron 110(8), 113. https://doi.org/10.1016/j.neuron.2022.01.013CrossRefGoogle ScholarPubMed
Zhu, S, Noviello, CM, Teng, J, Walsh, RM, Kim, JJ, and Hibbs, RE (2018) Structure of a human synaptic GABAA receptor. Nature 1, https://doi.org/10.1038/s41586-018-0255-3Google Scholar
Zhu, S, Sridhar, A, Teng, J, Howard, RJ, Lindahl, E, and Hibbs, RE (2022) Structural and dynamic mechanisms of GABAA receptor modulators with opposing activities. Nature Communications 2022 13:1 13(1), 113. https://doi.org/10.1038/s41467-022-32212-4Google ScholarPubMed
Ziemba, AM, Szabo, A, Pierce, DW, Haburcak, M, Stern, AT, Nourmahnad, A, Halpin, ES, and Forman, SA (2018) Alphaxalone binds in inner transmembrane β+-α- interfaces of α1β3γ2 γ-aminobutyric acid type A receptors. Anesthesiology 128(2), 338351. https://doi.org/10.1097/ALN.0000000000001978CrossRefGoogle ScholarPubMed
Figure 0

Figure 1. General structure of GABAAR. (A) The α1β2γ2 receptor structure viewed from the plane of the membrane. ECD and TMD are marked: α1 subunits in blue, β2 in red, and γ2 in yellow. (B) and (C): The same structure viewed from the extracellular space. Within the ECD, GABA and BDZ binding sites are located at β+ and α+ subunit interfaces, respectively. In TMD, a wide spectrum of modulatory (MOD) binding sites is located in most subunit interfaces, and a PTX-binding site is found in the ion pore. (D) Key β-strands of the ECD (black circles) and loops (red and blue circles, for principal and complementary subunits, respectively). (E) α-helices forming bundles in the TMD. Selected residues lining the ion pore are marked using x′ notation (universal for all subunits).

Figure 1

Figure 2. Ligands of GABAAR. Gamma aminobutyric acid (GABA) is a natural agonist that binds to orthosteric binding sites (β/α subunit interface in the ECD). Partial agonist P4S and competitive antagonists bicuculline and gabazine bind to the same site. Diazepam and Zolpidem are positive allosteric modulators binding to the allosteric binding sites (α/γ subunit interface in the ECD). Diazepam belongs to the family of BDZs together with, for example, flurazepam, lorazepam, flunitrazepam, and alprazolam. These drugs share a common benzene and diazepine ring system and differ by substitutions (at R1, R2, R3, R7, and R2’). Zolpidem belongs to the nonbenzodiazepine family, which shares a similar mode of action to BDZs, but significantly differ in structure (being an imidazopyridine). DMCM which belongs to the family of β-carbolines binds to the same site, but acts as negative modulator. General anesthetics are another large group of positive modulators, including propofol, etomidate, and barbiturates. Phenobarbital is an example of the barbiturate; other members of the family differ by R1 and R2 substitutions. Those modulators’ sites are located at subunit interfaces in the upper TMD. Neurosteroids may act as positive or negative allosteric modulators. Pregnanolone (3α5β) and its isomer allopregnanolone (3α5α) are positive modulators, whereas another isomer epipregnanolone (3β5β) is a negative modulator. The addition of sulfate ester (at C3β) to pregnanolone turns it into a negative allosteric modulator. Neurosteroids binds to multiple sites within the TMD.

Figure 2

Figure 3. pLGIC and GABAAR isoforms. (A) pLGICs are categorized according to their ion selectivity. Animal pLGICs are also called Cys-loop receptors because of the characteristic disulfide bond between β-strands 6 and 7. The respective receptor isoforms are listed in parentheses. (B): There are 19 subunit types of the GABAA receptor: three β subunits that act as principal (red) subunits and six α subunits that play the role of the complementary (blue) one and three γ subunits together with ε, θ, δ, and π subunits that play a modulatory (yellow) role (when co-expressed with two β and two α subunits). The ρ subunits do not assemble with other subunit types; they form only homomers formerly categorized as GABA-type C receptors. (C) Aligned sequences of the α1, β2, and γ2 GABAAR subunits with marked β-strands (green), α-helices (blues), and key loops. Sequence of the ICD is omitted.

Figure 3

Figure 4. IPCs. (A) Exemplary trace of IPSC between parvalbumine-positive interneuron and hippocampal CA1 pyramidal cell, measured in the whole-cell configuration. Synaptic transmission was evoked using optogenetic stimulation of presynaptic interneuron (unpublished, recorded by G. Wiera). (B) Exemplary current response to an ultrafast short application (3 ms) of saturating [GABA] (10 mM) recorded from an outside-out patch excised from a HEK293 cell expressing GABAARs. Note similar kinetics to the neuronal IPSC (unpublished, recorded by M.M. Czyżewska).

Figure 4

Figure 5. Macroscopic recordings. (A) Exemplary current response to an ultra-fast long application (500 ms) of saturating [GABA] (10 mM) recorded from an outside-out patch excised from a HEK293 cell expressing GABAARs (unpublished, recorded by K. Kłopotowski). (B) Exemplary drawing of a current response (saturating [GABA], long application) with kinetic features resembling typical recordings from oocytes, characterized by different kinetics (and time scale) when compared to the current recorded from a HEK293 cell (A).

Figure 5

Figure 6. Macroscopic recording analysis. (A) Long agonist pulse (onset phase). RT 10-90 describes the time needed for amplitude increase from 10 to 90% of its peak amplitude. Fast component of macroscopic desensitization can be fitted with a single exponential (in a limited time window), giving the amplitude of rapidly desensitizing current (Afast) and time constant (τfast). (B) Long agonist pulse. Macroscopic desensitization fitted for the entire time window of agonist application yields amplitudes and time constants of the fast and slow components. Deactivation is fitted with a single or double exponential function, depending on the characteristics of the recorded trace. (C) Short agonist pulse. Because of the short duration of the pulse, macroscopic desentization is negligible, but deactivation shows clear two-phase behavior and is fitted with a sum of two exponential functions.

Figure 6

Figure 7. Macroscopic modeling. (A) Kinetic model of the receptor with possible transitions (from the resting state R) to the open (A1O and A2O) and desensitized (A1D and A2D) states from the singly (A1R) and doubly (A2R) agonist-bound states. (B) Kinetic model expanded with an additional flipped/preactive state (A2F). Transitions from the singly bound state are omitted. (C) Cyclic kinetic model allowing for transition pathways between the respective states (symbols same as in A, added O and D states are spontaneously open and desensitized state). (D) Exemplary experimentally recorded trace and receptor response simulated using the kinetic model (unpublished, recorded by M.M. Czyżewska).

Figure 7

Figure 8. Single-channel recordings. (A) and (B) Exemplary traces recorded from HEK293 cells expressing α1β2γ2 GABAA receptors, measured in the cell-attached configuration. Repetitive receptor openings form bursts that may be grouped into clusters. (C) Exemplary distributions of times in the open and closed states fitted with the sum of exponentials (two for open times and four for shut events, unpublished, recorded by K. Kłopotowski). Solid lines represent approximations of the probability density functions determined as the sums of the respective exponentials (dashed lines). The number of exponentials in a distribution indicates the number of respective states.

Figure 8

Figure 9. Single-channel modeling. (A) Distributions of open and shut times (solid lines) calculated using a kinetic model with optimized rate constants. Dashed lines present distributons determined with correction for missed events. Distributions are superimposed on histograms for the experimental data (unpublished, recorded by K. Kłopotowski). (B) Kinetic scheme often used to model single-channel data obtained for stationary saturating [GABA]. Tranisitions related to agonist binding reactions are omitted, but extra desensitized (A2D′) and open state (A1O′) are added (compared to the model in Figure 7B). These additional states are required to reproduce multiple open and shut states, as indicated by the number of components in the open and shut states distributions. (C) Complete kinetic model (based on (Kisiel et al. 2018) of GABAAR including spontaneous and both singly/doubly agonist bound activity with and without transition through the flipped/preactivated state.

Figure 9

Table 1. Summary of published GABAAR structures

Figure 10

Figure 10. The N-terminal region of the receptor. α1 helices are located in the apical part of the ECD. The unstructured loop connecting the α1 helix and β1 strand of the principal subunit is located below the α1 helix of the complementary subunit, allowing for non-covalent interactions at the subunit interface (via, e.g., β2F31 and α1F15). The α1 helix is also parallel to the loop of the same subunit that connects the β2 and β3 strands located above the β strands, forming the orthosteric binding site.

Figure 11

Figure 11. ECD-binding site of GABAAR. (A) General view of the orthosteric binding site. The binding cavity is lined with antiparallel β-strands of the principal (4, 7, 10, and 9, loops A and B) and complementary (5, 6, 2, 1, and 8, loops E, D, G, and F) subunits and capped by loop C (between β-strands 9 and 10). (B) In the principal subunit side loops A, B, and C are present. Loops A and B are located deep in the binding cavity, whereas loop C between strand β9 and β10 is exposed to the solvent. Residues β2Y97, β2Y157, and β2F200 are forming an ‘aromatic box’ around GABA molecule, whereas β2E155 form electrostatic interaction with ligand molecule. (C) All loops (E, D, G, and F) of the complementary side of the cavity are fragments of the respective β-strands. Strand β5 and loops E and D are located similarly to loops A and B in the deeper part of the binding site, whereas loops G and F are partially exposed to the extracellular space. The ‘aromatic box’ on the complementary side is formed with α1F65 and α1F45 whereas electrostatic interaction is present between agonist and α1R67.

Figure 12

Figure 12. Domain interface of GABAAR. (A) α1β2γ2receptor structure viewed from the plane of the membrane. There are multiple areas of possible interdomain interactions: covalent connections via the β10-M1 linker and non-covalent ones via: β10-M1 linker and loop 2 (between β-strands 1 and 2), Cys-loop (between β-strands 6 and 7) and M2-M3 loop (between α-helices M2 and M3), as well as between Cys-loop and β10-M1 linker. (B) Although the β10-M1 linker and loop 2 are relatively distant from each other, some charged residue functional groups point toward each other, enabling electrostatic interactions (β2R216 and β2E52). Residues located at the bottom of the loop (β2V53/α1H56) are oriented toward the M2-M3 loop. (C) The Cys-loop is located above the M2-M3 loop and is almost parallel to the β10-M1 linker, enabling interactions with both structures. (D) The M2-M3 loop is placed below both Cys-loop and loop 2, making its apical residues (e.g., β2K279 and β2P273) crucial for interdomain interactions.

Figure 13

Figure 13. Ion pore of GABAAR. Residues are typically described using x′ notation, starting from the bottom of the pore. The most crucial for receptor function are 9′ residues (forming channel gate, β2L259 and α1L264) and -2′ (making additional constriction site attributed to the desensitized state, β2A248 and α1P253).

Figure 14

Figure 14. Residues at the bottom of the TMD involved in the desensitization transition. The residues important for the receptor’s desensitization transition are located within helical bundles at the bottom of the TMD. In addition, the -2′ residue (β2A248 and α1P253, pore lining residues, see Figure 13) forms a desensitization-dependent constriction.

Figure 15

Figure 15. TMD-binding sites of the GABAAR: α1β2γ2 receptor structure viewed from the plane of the membrane. (A): Upper TMD-binding site. The cavity between α-helices M2 and M3 (principal subunit side) and M1 (complementary subunit side) is marked in orange. Crucial for modulatory action, β2N265 is located in the M2 helix behind β2M286 (M3). At the same level of the complementary subunit, α1L232 is present. (B) PAM neurosteroid-binding sites: intersubunit at the bottom, and intrasubunit at the top of the TMD. Important for neurosteroid function, α1Q242 is located roughly in the middle of the M1 helix (top of the intersubunit side), and on the opposite side of the cavity, β2Y304 and α1W246 are located (bottom of the site). (C) NAM neurosteroid-binding site at the bottom of the α subunit.

Figure 16

Table 2. Localization of the upper TMD modulatory binding sites

Figure 17

Table 3. Localization of the TMD neurosteroid-binding site