Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Dec 18.
Published in final edited form as: Neuropharmacology. 2008 Aug 8;56(1):141–148. doi: 10.1016/j.neuropharm.2008.07.045

GABAA Receptors: Subtypes Provide Diversity of Function and Pharmacology

Richard W Olsen 1, Werner Sieghart 2
PMCID: PMC3525320  NIHMSID: NIHMS86837  PMID: 18760291

GABAA Receptors

Gamma-aminobutyric acid (GABA), the major inhibitory neurotransmitter in the brain, mediates rapid inhibition via GABAA receptors (GABAA-R). These are ligand-gated chloride ion channels that were first identified pharmacologically as being activated by GABA and the selective agonist muscimol, blocked by bicuculline and picrotoxin, and modulated by benzodiazepines, barbiturates, and certain other CNS depressants (Macdonald & Olsen, 1994; Sieghart, 1995). GABAA-R mediate rapid phasic inhibitory synaptic transmission, and also tonic inhibition by producing currents in extrasynaptic and perisynaptic locations (Mody & Pearce, 2004; Farrant & Nusser, 2005). Due to their widespread localization throughout the mammalian nervous system, GABAA-R play a major role in virtually all brain physiological functions and serve as targets of numerous classes of drugs, both used clinically and important as research tools (Hevers & Lüddens, 1998; Olsen & Martin, 2000).

Heterogeneity of GABAA receptors

GABAA receptors are composed of five protein subunits that can belong to different subunit classes. There are 19 genes for GABAA-R subunits (Simon et al., 2004). These include 16 subunits (α1–6, β1–3, γ1–3, δ, ε, θ, π) combined as GABAA, and 3 rho (ρ) subunits, which contribute to what have sometimes been called GABAC receptors; the latter are considered by the Nomenclature Committee of IUPHAR to be subtypes of GABAA-R containing the ρ subunits, and they recommend against using the term ‘GABAC receptors’, especially in the title, abstract, or initial mention of these receptors in publications (Barnard et al., 1998; Olsen & Sieghart, 2008).

The assembly of GABAA-R as heteropentamers produces complex heterogeneity in their structure, which is the major determinant of their pharmacological profile (Barnard et al., 1998; Olsen & Sieghart, 2008). The various receptor subtypes differ in abundance in cells throughout the nervous system and thus in functions related to the circuits involved. Available techniques do not allow us to unequivocally establish the subunit composition and arrangement of GABAA receptors in vivo, but this review summarizes our attempts to establish criteria for designating a proposed receptor heterooligomeric subtype as a native receptor, and our current opinion of which subtypes meet these criteria (Olsen & Sieghart, 2008).

Identification of native GABAA receptor subtypes by their regional and cellular distribution

In situ hybridization (Wisden et al., 1992; Persohn et al., 1992) and immunohistochemical studies (Pirker et al., 2000; Fritschy et al., 1992) have indicated that α1, β1, β2, β3, and γ2 subunits are found throughout the brain, although differences in their distribution were observed. Subunits α2, α3, α4, α5, α6, γ1, and δ are more confined to certain brain areas and in some brain regions, a complementary distribution of α2, α4, β3, and δ versus α1, β2, and γ2 subunits was detected (Sieghart and Sperk, 2002). The δ subunit is frequently co-distributed with the α4 subunit, e.g. in the thalamus, striatum, outer layers of the cortex and in the dentate molecular layer. In the cerebellum, however, it is co-distributed with the α6 subunit (Pirker et al., 2000).

An important criterion for association of subunit isoforms into oligomeric native receptors is co-localization of the subunits. Immunocytochemical studies investigating the co-localization of subunits in GABAA receptor clusters on neuronal membranes (Fritschy et al., 1992; Bohlhalter et al., 1996), as well as electron microscopic studies (Nusser et al., 1995; Somogyi et al., 1996), indicate that the majority of GABAA receptors present in the brain are composed of α, β, and γ subunits. Receptors composed of α1, β2, and γ2 subunits are extensively co-localized among subsets of GABAergic interneurons in hippocampus and other brain regions, e.g. calretinin-, neuropeptideY-, somatostatin-positive cells, but not on calbindin-D28k-, cholecystokinin- and vasoactive intestinal peptide-containing cells (Gao and Fritschy, 1994), supporting the conclusion that α1β2γ2 receptors are the most abundant GABAA receptors in the brain (Pirker et al., 2000).

In the raphe nuclei the vast majority of serotonergic neurons express strong α3 subunit immunoreactivity but are devoid of α1 subunit staining. In contrast, both the α1 and the α3 subunits are expressed in GABAergic neurons (Gao and Fritschy, 1993). In other studies it was demonstrated that 84–95% of the cholinergic neurons in the basal forebrain expressed the α3 subunit but not the α1 subunit. In contrast, parvalbumin-positive GABAergic neurons in these brain regions were frequently co-stained with the α1-subunit and to a lesser extent with the α3 subunit antibody (Gao et al., 1995). The α3 subunit, however, not only is associated with serotonergic or cholinergic neurons, but also with noradrenergic and dopaminergic neurons in the brainstem (Gao et al., 1995). In addition, data indicating an overlapping distribution of α3, θ, and ε GABAA receptor subunits in the dorsal raphe and the locus coeruleus (Moragues et al., 2000; Moragues et al., 2002), suggest that novel GABAA receptor subtypes that so far have not been studied in detail, may regulate neuroendocrine and modulatory systems in the brain.

Identification of native GABAA receptor subtypes by their synaptic and extrasynaptic localization

Other studies indicated that the individual subunits exhibit a distinct subcellular distribution. For instance, in cerebellar granule cells α1, α6, β2/3 and γ2 subunits have been found by immunogold localizations to be concentrated in GABAergic Golgi synapses and also are present in the extrasynaptic membrane at lower concentration. In constrast, δ subunits could not be detected in synaptic junctions, although they were abundantly present in the extrasynaptic dendritic and somatic membranes (Nusser et al., 1998). Receptors containing the δ subunit also contain α6 and β subunits (Jechlinger et al., 1998; Pöltl et al., 2003). Receptors containing δ subunits exhibit a smaller single channel conductance and a much longer open time, and do not desensitize on the prolonged presence of GABA (Saxena and Macdonald, 1994). Together with the exclusive extrasynaptic localization of these receptors, these properties indicate that tonic inhibition observed in these cells is mediated mainly by the persistent activation of α6βδ receptors by GABA, that is present in the extracellular space of glomeruli (Nusser et al., 1998; Brickley et al., 1999).

Extrasynaptic receptors have been identified also in other brain regions. In the forebrain it is assumed that they are predominantly composed of α4βδ. Experiments indicating that tonic conductance sometimes can also be enhanced by benzodiazepines suggest that tonic inhibition can also be produced by gamma subunit-containing receptors. In the forebrain evidence has accumulated that such receptors might be composed of α5βγ2 subunits. But it can be assumed that receptors with other subunit composition also can occur extrasynaptically (Semyanov et al., 2004). Due to the much larger cell surface area, the charge carried by the activation of tonically active GABAA-R can be more than three times larger than that produced by phasic inhibition (Nusser and Mody, 2002; Rossi et al., 2003).

Identification of native GABAA receptor subtypes by their subunit composition

A variety of GABAA receptor subunit-specific antibodies have been generated and have been used for purifying GABAA receptor subtypes from brain membrane extracts by immunoprecipitation or immunoaffinity chromatography. These studies indicated an extreme promiscuity of the various subunits. Although the antibodies used were highly specific for the respective subunits, most if not all of the other subunits investigated could be co-purified with antibodies directed against an individual α or β subunit, suggesting that α and β subunits can combine with most of the other subunits to form a variety of different receptor subtypes. These studies also indicate that two different α and two different β subunits can be present in GABAA receptors. In contrast, most studies agreed that γ subunits could not be co-precipitated with other γ subunits. Similarly, δ subunits seem not to be present together with γ subunits in the same receptors (for review see Sieghart and Sperk, 2002). From these results a subunit stoichiometry of 2α, 2β and 1γ or 1δ subunit can be deduced for native receptors. This conclusion was confirmed in several recombinant receptor studies, some of which also were able to determine the subunit arrangement of receptors composed of 2α, 2βand 1γ subunit (Tretter et al., 1997; Baumann et al., 2002).

Studies investigating the abundance of GABAA-R subtypes using subunit-specific antibodies, confirmed that receptors composed of α1β2βγ2 subunits are the most abundant GABAA-R in the brain. Similarly, α2βγ2, α3βγ2, α4βγ2, α5βγ2, α6βγ2, α4βδ and α6βδ seem to be abundant, but to a lower extent (for review see Sieghart and Sperk, 2002). Given the promiscuity of subunits discussed above, and the receptor subtypes so far identified, it was estimated that more than 800 distinct GABAA-R subtypes might exist in the brain (Barnard et al., 1998). Most of these receptors are not very abundant, but due to the widespread distribution and quantitative importance of the GABAergic system, even minor GABAA-R subtypes probably exhibit an abundance comparable with that of major norepinephrine, dopamine, serotonin or peptide receptors.

Structural basis of GABAA receptor pharmacology

The GABAA-R are members of the Cys-loop pentameric LGIC superfamily, including nicotinic acetylcholine receptors, inhibitory glycine receptors, and ionotropic 5-HT3 (serotonin) receptors. They differ in structure from two additional LGIC families: the tetrameric glutamate receptors, and the trimeric purine receptors (see mini-reviews on these families in this volume). All of the 44 subunit members of the Cys-loop pentameric LGIC superfamily (Collingridge et al., 2008) show sequence homology on the order of 30 % identity, but even greater similarity at the level of secondary and tertiary structure. All are organized as pentameric membrane-spanning proteins surrounding a central pore which forms the ion channel through the membrane. They all use similar sequences and functional domains to establish membrane topology, ion channel structure, agonist binding sites, and even binding sites for diverse allosteric ligands (Corringer et al., 2000; Sine & Engel, 2006; Sigel & Buhr, 1997; Li et al., 2006; Bali & Akabas,2004; Hosie et al., 2006). Each subunit consists of a long N-terminal extracellular hydrophilic region, followed by four trans-membrane (M) α-helices with a large intracellular loop between M3 and M4, and ends with a relatively short extracellular C-terminal domain, and M2 forms the lining of the ion channel. The structure of the Cys-loop pentameric LGIC superfamily has been resolved (currently 4 Å) for the nicotinic acetylcholine receptor of Torpedo marmorata obtained by cryo-electron microscopy and image reconstruction (Unwin, 2005). This has been combined with X-ray crystallography data on the homologous snail soluble acetylcholine binding protein (AChBP; Brejc et al., 2001), recently followed by that of the water soluble portion of the muscle nicotinic acetylcholine receptor (Dellisanti et al., 2007), and that of a related bacterial membrane LGIC receptor (Hilf & Dutzler, 2008).

The homologous sequence and structure allows homology modeling of the various functional domains of GABAA-Rs using the structures of the AChBP and the nAChR as templates (Cromer et al., 2002; Ernst et al., 2003; Ernst et al., 2005). The resulting models not only identified the absolute arrangement of the 1γ, 2α and 2β subunits within the receptor, but also the structural features of the extracellular domain and their binding pockets. In the extracellular domain, there are 5 pockets at the 5 subunit interfaces. The two pockets at the β/α interfaces form the 2 GABA binding sites, and the pocket at the α/γ interface forms the benzodiazepine binding site. The GABA pocket is formed by the so-called “loops” A, B, and C of the β+ (“principal”) side and the so-called “loops” D, E, and F of the α- (“complementary”) side (Ernst et al., 2003). Residues homologous to those in the agonist binding loops at the β/α interface, are homologous in all members of the superfamily (Smith & Olsen, 1995; Sigel & Buhr, 1997; Corringer et al., 2000) and are also homologous to those of the respective “loops” A, B, C, of the α+ and D, E, and F of the γ- side forming the benzodiazepine binding site (Ernst et al., 2003).

Homology modeling of the extracellular as well as the transmembrane domain of the GABAA receptor was based on the 4 Ǻ structure of the nAChR (Unwin, 2003), and indicated that GABAA receptors contain additional cavities in the transmembrane domain in addition to those found extracellularly (Ernst et al., 2005). One set of putative pockets is actually an extension of the extracellular pockets into the lipid bilayer of the transmembrane domain. The extracellular and transmembrane intersubunit pockets, however, are not necessarily connected with each other and could also represent completely separate entities. Another type of cavity is found contained inside each of the subunits surrounded by the four helices that make up the transmembrane domain. These multiple cavities are probably needed for conformational changes of the receptors, but could also serve as drug binding sites. Binding of drugs into any one of these sites could either stabilize or induce conformational changes of the receptor and thus, enhance or reduce GABA-induced chloride flux. These multiple binding sites at a single receptor subtype, thus, could explain the extremely complex pharmacology of GABAA receptors (Ernst et al., 2005).

Identification of native GABAA receptor subtypes by their pharmacology

Benzodiazepines

GABAA-R are the site of action of a variety of pharmacologically and clinically important drugs. Their interaction with benzodiazepines has been most thoroughly investigated. The location of the benzodiazepine binding site at the α+/γ- interface indicates that the benzodiazepine pharmacology of receptor subtypes is mainly determined by the α and γ isoform forming this site. The classical benzodiazepines, such as diazepam or flunitrazepam, predominantly interact with receptors composed of α1βγ2, α2βγ2, α3βγ2, or α5βγ2. They exhibit no activity on α4βγ2 or α6βγ2 receptors and a reduced activity on receptors containing γ1 or γ3 subunits (Sieghart, 1995; Hevers and Lüddens, 1998). However, the pharmacology of γ1 and γ3 containing receptors so far has not been extensively investigated. The classical benzodiazepines cannot distinguish between the benzodiazepine sites of different GABAA-R subtypes, but over time, some compounds, such as the benzodiazepines quazepam and cinolazepam (Sieghart, 1989) or non-benzodiazepines such as zolpidem, Cl218872, abecarnil, zaleplon and indiplon (for review see Möhler, 2006), have been developed that exhibit a preferential selectivity for α1βγ2 receptors. Most of these compounds exhibit sedative and hypnotic properties. Other compounds, such as SL651498 (Griebel et al., 2003) or TPA023 (Atack et al., 2006) seem to exhibit a certain selectivity for α2βγ2 and α3βγ2 receptors, whereas L-838,417 (McKernan et al., 2000) exhibits a selectivity for α2βγ2, α3βγ2 and α5βγ2 receptors. All these compounds exhibit anxiolytic properties. One of the few non-sedative, anxiolytic compounds that is not α2 subunit-selective is ocinaplon (Basile et al., 2004; Atack, 2005). The compound L-655 708 is a partial inverse agonist with preference for α5βγ2 receptors and exhibits memory-enhancing properties (Sternfeld et al., 2004;Chambers et al., 2004), α3IA is a weak inverse agonist at α3βγ2 receptors that has anxiogenic actions (Atack et al., 2005). The overall selectivity of most of these drugs, however, is not yet sufficient for a selective activation of a single GABAA receptor subtype, allowing the unequivocal identification of these receptor subtypes and their function in the brain.

Other benzodiazepine site ligands, such as the imidazobenzodiazepines flumazenil or Ro15-4513, are also able to interact with α4βγ2 or α6βγ2 receptors. These drugs have been used for identification of these receptor subtypes in the brain with ligand binding studies and autoradiography, using [3H]Ro15-4513 in the absence, or presence, of high concentrations of diazepam or flumazenil, to mask receptors containing the α1, α2, α3, or α5 subunits. (Korpi et al., 2002).

The function of at least some GABAA-R subtypes in the brain recently has been identified using a combined molecular genetic and pharmacological approach (Rudolph & Möhler, 2004; Whiting, 2006). Thus, based on the evidence that most of the actions of diazepam are mediated via receptors composed of α1βγ2, α2βγ2, α3βγ2 and α5βγ2 subunits, a point mutation was introduced into the genes of the individual α subunits rendering the respective receptors insensitive to allosteric modulation by diazepam. A comparison of drug-induced behavioral responses in the mutated and wild-type mice then allowed the identification of diazepam effects that were missing or reduced in the mutant mice. Using this approach, it was demonstrated that α1βγ2 receptors mediate the sedative, anterograde amnestic and in part the anticonvulsant actions of diazepam (Rudolph et al., 1999; McKernan et al., 2000). The anxiolytic activity of diazepam is mediated primarily by GABAA-Rs composed of α2 βγ2 subunits (Low et al., 2000), and, under conditions of high receptor occupancy, also by α3 GABAA-Rs (Dias et al., 2005; Yee et al., 2005). The α2βγ2 receptors are also implicated in some of the muscle relaxant activities of diazepam (Low et al., 2000). Receptors containing the α3 subunit seem to mediate the anti-absence effects of clonazepam, as indicated by the respective point mutated mouse (Sohal et al., 2003) and the α3 global knockout mice displayed a hyperdopaminergic phenotype relevant for GABAergic control of psychotic-like symptoms (Yee et al., 2005). The α5βγ2 receptors seem to influence learning and memory, shown by improved spatial memory in mice with knockout of α5 subunits (Collinson et al., 2002), and improved trace fear conditioning in the point-mutated α5 knockin mouse (Crestani et al., 2002). These and other studies for the first time indicated a possible function of specific GABAA-R subtypes in the brain and provided additional evidence for their actual existence in vivo.

Other allosteric ligands

In addition to benzodiazepines, other allosteric modulators of GABAA receptors such as general anesthetics of diverse chemical structure, including volatile agents and intravenous anesthetics, such as barbiturates, long-chain alcohols, neuroactive steroids, etomidate, propofol, as well as ethanol (Olsen & Sieghart, 2008) can be used for the identification of specific GABAA receptor subtypes.

The differential sensitivity of GABAA receptor subtypes to anesthetics led to the identification of residues in the trans-membrane region of α and β subunits that are critical for modulation of GABAA-R function by volatile anesthetics like isoflurane, and high dose (≥100 mM) ethanol (Mihic et al., 1997). These residues are located at TM2-15’ and TM3-4’ (helical numbering convention), and proposed to be part of a single intrasubunit water-filled binding pocket (Yamakura et al., 2001), the existence of which has also been predicted by homology modeling studies (Ernst et al., 2005). The TM2-15’ had been found independently to affect sensitivity to loreclezole and the structurally related intravenous anesthetic etomidate, in β subunits (Belelli et al., 1997).

The βM3-4’residue was further demonstrated to be part of the binding site for the intravenous anesthetic propofol (Bali & Akabas, 2004) and to be affinity labeled with an analogue of etomidate (Li et al., 2006), which also identified a second residue in αM1 (11’). Residues αM1 (11’) and βM3-4’were proposed to contribute to a single intersubunit binding pocket at the β/α subunit interface, 50 Ǻ below the GABA binding pocket in the extracellular domain (Li et al., 2006). The residues in αM1 and βM3 were shown to be facing each other by cross-linking of helical cysteine-substituted residues (Jansen & Akabas, 2006; Bali et al., 2007) supporting the conclusion that residues αM1-11’ and βM3-4’ are likely to be positioned appropriately to participate in an intersubunit pocket. This conclusion is also consistent with homology modeling using the Unwin (2005) structure of the nAChR as a template. Further experiments will have to clarify whether βM3-4’contributes to an intrahelical or interhelical pocket.

The sensitivity of drugs towards the respective residues βM2-15’ in β2- and β3-containing receptors led to the production of knock-in mice which exhibited a reduced sensitivity to the sedative (β2N265S: Reynolds et al., 2003) and anesthetic actions of etomidate in vivo (β3N265M: Jurd et al., 2003). These studies demonstrated not only that GABAA-Rs mediate the action of the general anesthetic etomidate, but also that the behavioral effects depend differentially on β2- or β3-containing GABAA-R, presumably due to their anatomical localization in brain structures contributing to the respective actions.

Another, related, mutagenesis study utilizing homology models of the TMD region of GABAAR identified two potential binding pockets important for neurosteroid action: one intrasubunit site shared within αM1and αM4 mediating enhancement of GABA, and a second, intersubunit site αM1/βM3, involved in direct gating of channels, respectively (Hosie et al., 2006). The steroid and etomidate site models have the intersubunit pockets near each other but steroids do not inhibit etomidate binding (Li et al.,2006). It is possible that the residues needed for steroid action identified by mutagenesis are involved in conformational coupling and not binding. More information is needed to understand how these important residues participate in structure and function. Better models of GABAAR structure will not only be useful for future rational drug design but also for aiding our understanding of their functional mechanism of action. The generation of knock-in mice carrying a mutation of the respective residues will be invaluable in a more detailed clarification of the action of these drugs and help to identify the receptor subtypes involved in their effects.

Which of the many possible GABAA receptor subtypes have been unequivocally identified?

The frequent co-localization of α, β, and γ, or α4/α6,β, and δ subunits in the brain and in synaptic or extrasynaptic sites of specific neurons, respectively, the evidence that these subunits could be co-precipitated from brain membrane extracts, that they form defined recombinant receptors with a specific pharmacology, their identification by receptor binding and electrophysiological studies and the identification of their function using molecular-genetic knock-in or knock-out and pharmacological experiments, suggest the actual existence of 8 different GABAA-R subtypes (α1-6βγ2, α4βδ / α6βδ).

The ρ subunits seem to be preferentially expressed in the retina (Enz et al., 1996;Koulen et al., 1998). mRNA encoding ρ subunits, however, is present also in the superior colliculus, dorsal lateral geniculate nucleus and cerebellar Purkinje cells (Boue-Grabot et al., 1998; Wegelius et al., 1998). In addition, pharmacological effects characteristic for ρ subunit containing receptors have been reported in various brain regions (for review see Sieghart and Sperk, 2002). The absence of antibodies or a pharmacology that could distinguish between receptors containing different ρ subunits, did not allow more thorough studies on their possible existence. Since the mere existence of a subunit and the fact that it can form receptors in recombinant systems does not necessarily imply that it actually forms receptors in vivo and since such subunits could also have other functions in the brain, at the time being it only seems safe to assume the existence of one ρ-containing native receptor in the brain.

Therefore, using a stringent classification, we only have 9 different GABAA receptor subtypes (α1-6βγ2, α4βδ, α6βδ, ρ) unequivocally identified in the brain (Olsen and Sieghart, 2008). To keep the possibility alive that there are also receptors composed of ρ1, ρ2, and ρ3 subunits, we added these three receptor types to a list of “tentative receptors”. In this list we did not include receptors containing mixed ρ subunits, because only recombinant receptor studies so far provided evidence for their possible existence, and in such studies receptors also can be formed that might not be formed under native conditions (Olsen and Sieghart, 2008). Out of the same reason, we also did not include possible receptors formed from a combination of ρ and γ2 or glycine receptor subunits (Pan et al., 2000; Milligan et al., 2004).

As mentioned above, not many studies have been performed investigating receptors containing any one of the minor GABAA receptors subunits (γ1, γ3, θ, ε, π). These subunits have been demonstrated to have a distinct regional distribution, in some cases, co-precipitation with other GABAA receptor subunits has been demonstrated, and recombinant receptors have been generated exhibiting specific properties (for review see Sieghart and Sperk, 2002). But the lack of a selective pharmacology and function in the brain indicated that receptors possibly formed from these subunits at present are quite tentative, justifying their inclusion into the list of “tentative receptors”.

Finally, the question arose, whether receptors containing 2 distinct α and/or 2 distinct β subunits actually occur in the brain. The existence of such receptors is suggested by the overlapping regional and cellular distribution of the α and β subunits, and by co-precipitation studies (Sieghart and Sperk, 2002). However, neither an overlapping regional and cellular distribution nor a co-precipitation of subunits necessarily means that the subunits are present in the same receptor subtypes. In the latter case they also could be present in different receptors associated with each other either directly or via their associated proteins. In addition, receptors containing 2 distinct α and/or 2 distinct β subunit so far have not been demonstrated to exhibit a specific pharmacology and function in the brain. But since Benke et al.(2004), using knock-in mice containing point-mutated α subunits, provided additional evidence for the existence of receptors containing different α subunits and reported upon the abundance of these receptors in an in vivo system, we decided to include at least receptors containing 2 different α subunits on the list of “tentative receptors”.

At this time, however, we decided that the evidence for the existence of receptors containing 2 different β subunits, or containing five different subunits (Sieghart and Sperk, 2002) is not strong enough yet for justifying inclusion of such receptor types in any list of possible receptors. The same holds true for receptors containing two different γ, δ, or ε per pentamer, or those containing a γ and a δ subunit. So in summary, in the list of “tentative receptors” we only included 9 different receptors (ρ1, ρ2, ρ3, αβγ1, αβγ3, αβθ, αβε, αβπ, αxαyβγ2).

However, Nusser et al., (1998) provided additional evidence for the existence of receptors composed of α1α6βγ2 or α1α6βδ subunits by demonstrating separate as well as co-localization of α1 and α6 subunits in the cerebellum granule cells. This prompted us to place at least one receptor composed of α1α6βγ / α1α6βδ into a third list of “receptors that exist with high probability” (Category B). This list, in between “receptors unequivocally identified” (Category A) and “tentative receptors” (Category C), defines receptors for which additional evidence for their existence has been accumulated, but for which evidence seemed not strong enough to include them into the list of “identified receptors”.

Additional evidence from co-localization, co-immunprecipitation and neuronal electrophysiology studies also suggests that γ2 or δ subunits partner with more than one β subunit (one at a time). In addition, point mutations in the M2 domain (M2-15´) that can produce β subunit selectivity for loreclezole or etomidate (Cestari and Yang, 1996; Belelli et al., 1997; Rudolph and Antkowiak, 2004) provided evidence for the existence of receptor subtypes with specific β subunits. So far, these results did not provide evidence on whether the γ2 or δ subunit-containing receptors, or both mediate the etomidate effects identified, although the δ-containing receptors are more sensitive to modulation. In addition, the respective α subunits present in these receptor also could not be identified.

The δ subunit-containing GABAA receptors, partnered with the α4, or α6 subunits, and especially with β3, are more sensitive than γ2-containing receptors to general anesthetics, neurosteroids, GABA analogs like THIP (Gaboxadol) (Wohlfarth et al., 2002; Brown et al., 2002; Chandra et al., 2006), as well as taurine (Jia et al., 2008) and ethanol (Wallner et al., 2003; Hanchar et al., 2005).The α4β2δ and α4β3δ receptors differ in sensitivity to modulators when recombinantly expressed in cells and both clearly occur naturally, because some brain areas express α4 and δ but not β2, or β3, subunits and display a pharmacology that distinguishes between these receptors. Thus, thalamic relay nuclei mainly express the β2 subunit and the moderately ethanol-sensitive α4β2δ receptor definitely mediates the tonic current (Chandra et al., 2006). Dentate granule cells express high levels of highly ethanol-sensitive, presumably α4β3δ isoforms (Liang et al., 2006), but additionally express the etomidate-sensitive α4β2δ isoforms (Herd et al., 2008). Partnering of either β2, or β3, with both the α4δ and α6δ subunits thus appears highly likely, even conclusive. Therefore, we added α4β2δ and α4β3δ, α6β2δ and α6β3δ subtypes to the list of “receptors identified”, replacing the generic α4βδδand α6βδ.

Recently, Glykys et al., 2007, demonstrated that the δ subunit can pair with the α1β in hippocampal interneurons, based on co-localization with immuno-staining, electropharmacological properties and changes in knockout mice. While this single report is convincing, we decided that this subtype did not meet sufficient criteria to be included in the list of “identified” subtypes. We therefore included the α1βδ receptor on the category B list, “identified with high probability”.

Another candidate qualifying for this list is the α5β3γ2 receptor subtype. The α5 and β3 subunits appear to be co-depleted in mice lacking either the β3, or α5 subunits (Olsen and Homanics, 2000), and the properties of recombinant α5β3γ2 receptors appear to reflect those of a native subtype found in CA1 pyramidal neurons (that are enriched in both α5 and β3 subunits) and are distinct from receptors in dentate gyrus granule cells (Burgard et al., 1996; Sur et al., 1998; McClellan and Twyman, 1999; Stell et al., 2003;Caraiscos et al., 2004). Such studies are not conclusive evidence for this combination, so we place this subtype in the second list. Another example in this category is the α1β3γ2 subtype. If we specify that β2 is the most common partner of α1γ2, and list it in category A, we need to ask whether α1 can partner with other β subunits. While strong evidence for β1 is lacking, the properties of α1β3γ2 receptors in recombinant studies resemble those in certain neurons that express these subunits, some of which lack the β2 subunit (Whiting et al., 2000; Sieghart and Sperk, 2002).

Although the β1 subunit is less abundant than the other β subunits, it is likely that some β1-containing GABAA receptors exist, based on regional distribution and coimmunoprecipitation data (Li and DeBlas, 1997; Jechlinger et al., 1998). A potential pharmacological fingerprint, salicylidene salicylhydrazide, has been reported to be a negative allosteric modulator selective for β1 versus β2 or β3 subunit-containing receptors (Thompson et al., 2004). There is some evidence (Whiting et al., 2000) that cultured astrocytes express GABAA-R with α2β1γ1 combination, and they are uniquely enhanced by the benzodiazepine site ligand DMCM, a β-carboline. However, more evidence is needed to establish if this subtype actually exists in the brain. We therefore placed just one subtype for αβ1γ/ αβ1δ in category B.

Finally, evidence accumulated for the existence of receptors composed of α and β subunits, only. First, Bencsits et al., (1999) demonstrated that a large part of the α4 receptors (about 50%) are not associated with γ, or δ, subunits and are possibly composed of α and β subunits only. This was in contrast to α1 receptors. Second, in several knockout mice it was demonstrated that receptors composed of only α and β subunits do exist, for example in γ2 knockout mice (Günther et al., 1995), δ knockout mice (Tretter et al., 2001), or α1 knockout mice (Ogris et al., 2006). Third, in several brain regions α and β occur in the absence of γ and δ subunits. And fourth, Mortensen and Smart (2006) demonstrated by electrophysiological studies that there are extrasynaptic αβ receptors on rat hippocampal pyramidal neurons. In the absence of any evidence for a specific αβ subunit combination, we placed one undefined αβ subtype into category B.

Conclusions

The importance of GABAA-Rs in different brain regions and behaviors has been reinforced in recent years, with recognition that GABAA-Rs come in many different flavors differing in subunit combination, resulting in subtypes that vary in physiology, pharmacology, and most importantly, in location. Both the circuitry and the subcellular location on different postsynaptic membranes and extrasynaptic membranes is important for GABAA-R control of neuronal excitability, circuit, and network activity.

As knowledge of structure and function of GABAA-Rs increases, we find that allosteric modulation by clinically important CNS drugs can be explained by micro-domains within the receptor protein structure, which differ slightly but importantly among receptor subtypes. Modulation of GABAA-R by drugs with anxiolytic, sedative/hypnotic, anesthetic, and antiepileptic profiles continues to provide a target for current and pharmaceutical pipeline candidates. These actions involve multiple categories of drug binding sites on GABAA-R, the benzodiazepine sites and probably several other distinct sites for drugs like other anxiolytics, neurosteroids, general anesthetics, including barbiturates, etomidate, propofol, and volatile agents, and ethanol. The existence of receptor subtypes is now established, and exploitation for products with specific targets is in progress. It is expected that the NC-IUPHAR database on receptor classification and nomenclature will soon be expanded to the LGIC receptors, provided at no cost to the entire community.

The recent review on GABAA-R (Olsen & Sieghart, 2008) provides a compendium of criteria for inclusion of candidate subtypes on a list of native receptors, and the resulting receptor list. Whether all these receptor subtypes as well as those still to be identified are biologically relevant currently cannot be answered. A specific subunit composition could allow, for example, specific targeting to a certain subcellular localization, specific regulation by protein kinases and phosphatases, or interaction with other associated proteins or other receptors. Thus we will need to continue to develop careful approaches for defining which receptor subtypes are native, especially the lower abundance subtypes. Three categories of subtypes are designated: Identified; Existence with High Probability; and Tentative. A ‘working list’ of receptor subtypes totaling 26 is provided in Olsen & Sieghart (2008), reproduced here (Table 1). The 19 subunit genes also are presented in Olsen & Sieghart (2008), and the subunit list for all LGIC genes is included in the introductory chapter for this volume (Collingridge et al., 2008).

Table 1.

GABAA Receptor List 2008.

A. Identified
    α1β2γ2 α4βγ2 α5βγ2 α6β3δ
    α2βγ2 α4β2δ α6βγ2 ρ
    α3βγ2 α4β3δ α6β2δ
B. Existence with High Probability
    α1β3γ2 α5β3γ2 αβ
    α1βδ αβ1γ/ αβ1δ α1α6βγ/ α1α6βδ
C. Tentative
    ρ1 αβγ1 αβθ
    ρ2 αβγ3 αβπ
    ρ3 αβε αxαyβγ2

The criteria for inclusion on the native receptors list and evidence for the subtypes were first described in Olsen & Sieghart, Pharmacol. Rev. 2008.

Current evidence suggests that only 11 subtypes can be listed as conclusively identified, and these are reasonably abundant. Also listed are several subtypes for which the evidence is strong but not conclusive (6 subtypes). Finally, subtypes are included containing one of each of the minor subunits, evidence for whose native existence is tentative (another 8, plus one subtype with two kinds of α subunit, for a total of 9, and a grand total of 26). Note that one (ρ) overlaps with list A. An additional similar number of subtypes which are relatively rare, but are likely to exist, are not listed at this time, but the list will continue to grow as more information becomes available. Each subtype could play a significant role in the cells in which they occur. It should be noted that even these minor GABAA-R subtypes are present in amounts comparable to, or greater than, receptor subtypes for other accepted brain neurotransmitters other than glutamate; that is the biogenic amines and acetylcholine. Clearly, the heteromeric ligand-gated ion channels offer much greater heterogeneity than other known receptor subtypes.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Contributor Information

Richard W. Olsen, Department of Molecular & Medical Pharmacology Geffen School of Medicine at UCLA, Los Angeles, California, USA

Werner Sieghart, Division of Biochemistry and Molecular Biology Center for Brain Research, Medical University of Vienna Vienna, Austria.

References

  1. Atack JR. The benzodiazepine binding site of GABAA receptors as a target for the development of novel anxiolytics. Expert Opin Investig Drugs. 2005;14:599–616. doi: 10.1517/13543784.14.5.601. [DOI] [PubMed] [Google Scholar]
  2. Atack JR, Hutson PH, Collinson N, Marshall G, Bentley G, Moyes C, Cook SM, Collins I, Wafford K, McKernan RM, Dawson GR. Anxiogenic properties of an inverse agonist selective for alpha3 subunit-containing GABAA receptors. Br J Pharmacol. 2005;144:357–366. doi: 10.1038/sj.bjp.0706056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Atack JR, Wafford KA, Tye SJ, Cook SM, Sohal B, Pike A, Sur C, Melillo D, Bristow L, Bromidge F, Ragan I, Kerby J, Street L, Carling R, Castro JL, Whiting P, Dawson GR, McKernan RM. TPA023 [7-(1,1-dimethylethyl)-6-(2-ethyl-2H-1,2,4-triazol-3-ylmethoxy)-3-(2-fluor ophenyl)-1,2,4-triazolo[4,3-b]pyridazine], an agonist selective for alpha2- and alpha3-containing GABAA receptors, is a nonsedating anxiolytic in rodents and primates. J Pharmacol Exp Ther. 2006;316:410–422. doi: 10.1124/jpet.105.089920. [DOI] [PubMed] [Google Scholar]
  4. Bali M, Akabas MH. Defining the propofol binding site location on the GABAA receptor. Mol Pharmacol. 2004;65:68–76. doi: 10.1124/mol.65.1.68. [DOI] [PubMed] [Google Scholar]
  5. Bali M, Jansen M, Akabas MH. Spatial relationship between the GABAA receptor β2 M3 region and the α1 subunit preM1 region. SfN abstracts. 2007 #441.8. [Google Scholar]
  6. Barnard EA, Skolnick P, Olsen RW, Mohler H, Sieghart W, Biggio G, Braestrup C, Bateson AN, Langer SZ. International Union of Pharmacology. XV. Subtypes of γ-aminobutyric acidA receptors: classification on the basis of subunit structure and receptor function. Pharmacol Rev. 1998;50:291–313. [PubMed] [Google Scholar]
  7. Basile AS, Lippa AS, Skolnick P. Anxioselective anxiolytics: can less be more? Eur J Pharmacol. 2004;500:441–451. doi: 10.1016/j.ejphar.2004.07.043. [DOI] [PubMed] [Google Scholar]
  8. Baumann SW, Baur R, Sigel E. Forced subunit assembly in alpha1beta2gamma2 GABAA receptors. Insight into the absolute arrangement. J Biol Chem. 2002;277:46020–46025. doi: 10.1074/jbc.M207663200. [DOI] [PubMed] [Google Scholar]
  9. Belelli D, Lambert JJ, Peters JA, Wafford K, Whiting PJ. The interaction of the general anesthetic etomidate with the gamma-aminobutyric acid type A receptor is influenced by a single amino acid. Proc Natl Acad Sci U S A. 1997;94:11031–11036. doi: 10.1073/pnas.94.20.11031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Bencsits E, Ebert V, Tretter V, Sieghart W. A significant part of native gamma -aminobutyric acidA receptors containing alpha 4 subunits do not contain gamma or delta subunits. J Biol Chem. 1999;274:19613–19616. doi: 10.1074/jbc.274.28.19613. [DOI] [PubMed] [Google Scholar]
  11. Benke D, Fakitsas P, Roggenmoser C, Michel C, Rudolph U, Mohler H. Analysis of the presence and abundance of GABAA receptors containing two different types of alpha subunits in murine brain using point-mutated alpha subunits. J Biol Chem. 2004;279:43654–43660. doi: 10.1074/jbc.M407154200. [DOI] [PubMed] [Google Scholar]
  12. Bohlhalter S, Weinmann O, Mohler H, Fritschy JM. Laminar compartmentalization of GABAA-receptor subtypes in the spinal cord: an immunohistochemical study. J Neurosci. 1996;16:283–297. doi: 10.1523/JNEUROSCI.16-01-00283.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Boue-Grabot E, Roudbaraki M, Bascles L, Tramu G, Bloch B, Garret M. Expression of GABA receptor rho subunits in rat brain. J Neurochem. 1998;70:899–907. doi: 10.1046/j.1471-4159.1998.70030899.x. [DOI] [PubMed] [Google Scholar]
  14. Brejc K, van Dijk WJ, Klaassen RV, Schuurmans M, van der Oost J, Smit AB, Sixma TK. Crystal structure of an ACh-binding protein reveals the ligand-binding domain of nicotinic receptors. Nature. 2001;411:269–276. doi: 10.1038/35077011. [DOI] [PubMed] [Google Scholar]
  15. Brickley SG, Cull-Candy SG, Farrant M. Single-channel properties of synaptic and extrasynaptic GABAA receptors suggest differential targeting of receptor subtypes. J Neurosci. 1999;19:2960–2973. doi: 10.1523/JNEUROSCI.19-08-02960.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Brown N, Kerby J, Bonnert TP, Whiting PJ, Wafford KA. Pharmacological characterization of a novel cell line expressing human alpha(4)beta(3)delta GABA(A) receptors. Br J Pharmacol. 2002;136:965–974. doi: 10.1038/sj.bjp.0704795. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Burgard EC, Tietz EI, Neelands TR, Macdonald RL. Properties of recombinant gamma-aminobutyric acid A receptor isoforms containing the alpha 5 subunit subtype. Mol Pharmacol. 1996;50:119–127. [PubMed] [Google Scholar]
  18. Caraiscos VB, Elliott EM, You-Ten KE, Cheng VY, Belelli D, Newell JG, Jackson MF, Lambert JJ, Rosahl TW, Wafford KA, MacDonald JF, Orser BA. Tonic inhibition in mouse hippocampal CA1 pyramidal neurons is mediated by alpha5 subunit-containing gamma-aminobutyric acid type A receptors. Proc Natl Acad Sci U S A. 2004;101:3662–3667. doi: 10.1073/pnas.0307231101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Cestari IN, Uchida I, Li L, Burt D, Yang J. The agonistic action of pentobarbital on GABAA beta-subunit homomeric receptors. Neuroreport. 1996;7:943–947. doi: 10.1097/00001756-199603220-00023. [DOI] [PubMed] [Google Scholar]
  20. Chambers MS, Atack JR, Carling RW, Collinson N, Cook SM, Dawson GR, Ferris P, Hobbs SC, O'Connor D, Marshall G, Rycroft W, Macleod AM. An orally bioavailable, functionally selective inverse agonist at the benzodiazepine site of GABAA alpha5 receptors with cognition enhancing properties. J Med Chem. 2004;47:5829–5832. doi: 10.1021/jm040863t. [DOI] [PubMed] [Google Scholar]
  21. Chandra D, Jia F, Liang J, Peng Z, Suryanarayanan A, Werner DF, Spigelman I, Houser CR, Olsen RW, Harrison NL, Homanics GE. GABAA receptor alpha 4 subunits mediate extrasynaptic inhibition in thalamus and dentate gyrus and the action of gaboxadol. Proc Natl Acad Sci U S A. 2006;103:15230–15235. doi: 10.1073/pnas.0604304103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Collingridge GL, Olsen RW, Peters J, Spedding M. A nomenclature for ligand-gated ion channels. Neuropharmacology. 2009 doi: 10.1016/j.neuropharm.2008.06.063. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Collinson N, Kuenzi FM, Jarolimek W, Maubach KA, Cothliff R, Sur C, Smith A, Otu FM, Howell O, Atack JR, McKernan RM, Seabrook GR, Dawson GR, Whiting PJ, Rosahl TW. Enhanced learning and memory and altered GABAergic synaptic transmission in mice lacking the alpha 5 subunit of the GABAA receptor. J Neurosci. 2002;22:5572–5580. doi: 10.1523/JNEUROSCI.22-13-05572.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Corringer PJ, Le Novere N, Changeux JP. Nicotinic receptors at the amino acid level. Annu Rev Pharmacol Toxicol. 2000;40:431–458. doi: 10.1146/annurev.pharmtox.40.1.431. [DOI] [PubMed] [Google Scholar]
  25. Crestani F, Keist R, Fritschy J-M, Benke D, Vogt K, Prut L, Bluthmann H, Mohler H, Rudolph U. Trace fear conditioning involves hippocampal alpha5 GABA(A) receptors. Proc Nat Acad Sci USA. 2002;99:8980–8985. doi: 10.1073/pnas.142288699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Cromer BA, Morton CJ, Parker MW. Anxiety over GABAA receptor structure relieved by AChBP. Trends Biochem Sci. 2002;27:280–287. doi: 10.1016/s0968-0004(02)02092-3. [DOI] [PubMed] [Google Scholar]
  27. Dellisanti CD, Yao Y, Stroud JC, Wang ZZ, Chen L. Crystal structure of the extracellular domain of nAChR alpha1 bound to alpha-bungarotoxin at 1.94 A resolution. Nat Neurosci. 2007;10:953–962. doi: 10.1038/nn1942. [DOI] [PubMed] [Google Scholar]
  28. Dias R, Sheppard WF, Fradley RL, Garrett EM, Stanley JL, Tye SJ, Goodacre S, Lincoln RJ, Cook SM, Conley R, Hallett D, Humphries AC, Thompson SA, Wafford KA, Street LJ, Castro JL, Whiting PJ, Rosahl TW, Atack JR, McKernan RM, Dawson GR, Reynolds DS. Evidence for a significant role of alpha 3-containing GABAA receptors in mediating the anxiolytic effects of benzodiazepines. J Neurosci. 2005;25:10682–10688. doi: 10.1523/JNEUROSCI.1166-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Enz R, Brandstatter JH, Wassle H, Bormann J. Immunocytochemical localization of the GABAC receptor rho subunits in the mammalian retina. J Neurosci. 1996;16:4479–4490. doi: 10.1523/JNEUROSCI.16-14-04479.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Ernst M, Brauchart D, Boresch S, Sieghart W. Comparative modeling of GABAA receptors: limits, insights, future developments. Neuroscience. 2003;119:933–943. doi: 10.1016/s0306-4522(03)00288-4. [DOI] [PubMed] [Google Scholar]
  31. Ernst M, Bruckner S, Boresch S, Sieghart W. Comparative models of GABAA receptor extracellular and transmembrane domains: important insights in pharmacology and function. Mol Pharmacol. 2005;68:1291–1300. doi: 10.1124/mol.105.015982. [DOI] [PubMed] [Google Scholar]
  32. Farrant M, Nusser Z. Variations on an inhibitory theme: phasic and tonic activation of GABA(A) receptors. Nat Rev Neurosci. 2005;6:215–229. doi: 10.1038/nrn1625. [DOI] [PubMed] [Google Scholar]
  33. Fritschy JM, Benke D, Mertens S, Oertel WH, Bachi T, Mohler H. Five subtypes of type A gamma-aminobutyric acid receptors identified in neurons by double and triple immunofluorescence staining with subunit-specific antibodies. Proc Natl Acad Sci U S A. 1992;89:6726–6730. doi: 10.1073/pnas.89.15.6726. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Gao B, Fritschy JM. Selective allocation of GABAA receptors containing the alpha 1 subunit to neurochemically distinct subpopulations of rat hippocampal interneurons. Eur J Neurosci. 1994;6:837–853. doi: 10.1111/j.1460-9568.1994.tb00994.x. [DOI] [PubMed] [Google Scholar]
  35. Gao B, Fritschy JM, Benke D, Mohler H. Neuron-specific expression of GABAA receptor subtypes: differential association of the alpha 1 and alpha 3 subunits with serotonergic and GABAergic neurons. Neuroscience. 1993;54:881–892. doi: 10.1016/0306-4522(93)90582-z. [DOI] [PubMed] [Google Scholar]
  36. Gao B, Hornung JP, Fritschy JM. Identification of distinct GABAA-receptor subtypes in cholinergic and parvalbumin-positive neurons of the rat and marmoset medial septum-diagonal band complex. Neuroscience. 1995;65:101–117. doi: 10.1016/0306-4522(94)00480-s. [DOI] [PubMed] [Google Scholar]
  37. Glykys J, Peng Z, Chandra D, Homanics GE, Houser CR, Mody I. A new naturally occurring GABA(A) receptor subunit partnership with high sensitivity to ethanol. Nat Neurosci. 2007;10:40–48. doi: 10.1038/nn1813. [DOI] [PubMed] [Google Scholar]
  38. Griebel G, Perrault G, Simiand J, Cohen C, Granger P, Depoortere H, Francon D, Avenet P, Schoemaker H, Evanno Y, Sevrin M, George P, Scatton B. SL651498, a GABAA receptor agonist with subtype-selective efficacy, as a potential treatment for generalized anxiety disorder and muscle spasms. CNS Drug Rev. 2003;9:3–20. doi: 10.1111/j.1527-3458.2003.tb00241.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Gunther U, Benson J, Benke D, Fritschy JM, Reyes G, Knoflach F, Crestani F, Aguzzi A, Arigoni M, Lang Y, Bluethmann H, Mohler H, Luscher B. Benzodiazepine-insensitive mice generated by targeted disruption of the gamma 2 subunit gene of gamma-aminobutyric acid type A receptors. Proc Natl Acad Sci USA. 1995;92:7749–7753. doi: 10.1073/pnas.92.17.7749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Hanchar HJ, Chutsrinopkun P, Meera P, Supavilai P, Sieghart W, Wallner M, Olsen RW. Ethanol potently and competitively inhibits binding of the alcohol antagonist Ro15-4513 to α4/6β3δ GABAA receptors. Proc Natl Acad Sci USA. 2006;103:8546–8551. doi: 10.1073/pnas.0509903103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Herd MB, Haythornthwaite AR, Rosahl TW, Wafford KA, Homanics GE, Lambert JJ, Belelli D. The expression of GABAA beta subunit isoforms in synaptic and extrasynaptic receptor populations of mouse dentate gyrus granule cells. J Physiol. 2008;586:989–1004. doi: 10.1113/jphysiol.2007.146746. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Hevers W, Luddens H. The diversity of GABAA receptors. Pharmacological and electrophysiological properties of GABAA channel subtypes. Mol Neurobiol. 1998;18:35–86. doi: 10.1007/BF02741459. [DOI] [PubMed] [Google Scholar]
  43. Hilf RJ, Dutzler R. X-ray structure of a prokaryotic pentameric ligand-gated ion channel. Nature. 2008;452:375–379. doi: 10.1038/nature06717. [DOI] [PubMed] [Google Scholar]
  44. Hosie AM, Wilkins ME, da Silva HM, Smart TG. Endogenous neurosteroids regulate GABAA receptors through two discrete transmembrane sites. Nature. 2006;444:486–489. doi: 10.1038/nature05324. [DOI] [PubMed] [Google Scholar]
  45. Jansen M, Akabas MH. State-dependent cross-linking of the M2 and M3 segments: functional basis for the alignment of GABAA and acetylcholine receptor M3 segments. J Neurosci. 2006;26:4492–4499. doi: 10.1523/JNEUROSCI.0224-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Jechlinger M, Pelz R, Tretter V, Klausberger T, Sieghart W. Subunit composition and quantitative importance of hetero-oligomeric receptors: GABAA receptors containing alpha6 subunits. J Neurosci. 1998;18:2449–2457. doi: 10.1523/JNEUROSCI.18-07-02449.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Jia F, Yue M, Chandra D, Keramidas A, Goldstein PA, Homanics GE, Harrison NL. Taurine is a potent activator of extrasynaptic GABA(A) receptors in the thalamus. J Neurosci. 2008;28:106–115. doi: 10.1523/JNEUROSCI.3996-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Jurd R, Arras M, Lambert S, Drexler B, Siegwart R, Crestani F, Zaugg M, Vogt KE, Ledermann B, Antkowiak B, Rudolph U. General anesthetic actions in vivo strongly attenuated by a point mutation in the GABA(A) receptor beta3 subunit. Faseb J. 2003;17:250–252. doi: 10.1096/fj.02-0611fje. [DOI] [PubMed] [Google Scholar]
  49. Korpi ER, Mihalek RM, Sinkkonen ST, Hauer B, Hevers W, Homanics GE, Sieghart W, Luddens H. Altered receptor subtypes in the forebrain of GABA(A) receptor delta subunit-deficient mice: recruitment of gamma 2 subunits. Neuroscience. 2002;109:733–743. doi: 10.1016/s0306-4522(01)00527-9. [DOI] [PubMed] [Google Scholar]
  50. Koulen P, Brandstatter JH, Enz R, Bormann J, Wassle H. Synaptic clustering of GABA(C) receptor rho-subunits in the rat retina. Eur J Neurosci. 1998;10:115–127. doi: 10.1046/j.1460-9568.1998.00005.x. [DOI] [PubMed] [Google Scholar]
  51. Li G-D, Chiara DC, Sawyer GW, Husain SS, Olsen RW, Cohen JB. GABAA receptor anesthetic binding sites identified at subunit interfaces by photolabeling. J. Neurosci. 2006;26:11599–11605. doi: 10.1523/JNEUROSCI.3467-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Liang J, Zhang N, Cagetti E, Houser CR, Olsen RW, Spigelman I. Chronic intermittent ethanol-induced switch of ethanol actions from extrasynaptic to synaptic hippocampal GABAA receptors. J Neurosci. 2006;26:1749–1758. doi: 10.1523/JNEUROSCI.4702-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Low K, Crestani F, Keist R, Benke D, Brunig I, Benson JA, Fritschy JM, Rulicke T, Bluethmann H, Mohler H, Rudolph U. Molecular and neuronal substrate for the selective attenuation of anxiety. Science. 2000;290:131–134. doi: 10.1126/science.290.5489.131. [DOI] [PubMed] [Google Scholar]
  54. Macdonald RL, Olsen RW. GABAA receptor channels. Annu Rev Neurosci. 1994;17:569–602. doi: 10.1146/annurev.ne.17.030194.003033. [DOI] [PubMed] [Google Scholar]
  55. McClellan AM, Twyman RE. Receptor system response kinetics reveal functional subtypes of native murine and recombinant human GABAA receptors. J Physiol. 1999;515(Pt 3):711–727. doi: 10.1111/j.1469-7793.1999.711ab.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. McKernan RM, Rosahl TW, Reynolds DS, Sur C, Wafford KA, Atack JR, Farrar S, Myers J, Cook G, Ferris P, Garrett L, Bristow L, Marshall G, Macaulay A, Brown N, Howell O, Moore KW, Carling RW, Street LJ, Castro JL, Ragan CI, Dawson GR, Whiting PJ. Sedative but not anxiolytic properties of benzodiazepines are mediated by the GABA(A) receptor alpha1 subtype. Nat Neurosci. 2000;3:587–592. doi: 10.1038/75761. [DOI] [PubMed] [Google Scholar]
  57. Mihic SJ, Ye Q, Wick MJ, Koltchine VV, Krasowski MD, Finn SE, Mascia MP, Valenzuela CF, Hanson KK, Greenblatt EP, Harris RA, Harrison NL. Sites of alcohol and volatile anaesthetic action on GABA(A) and glycine receptors. Nature. 1997;389:385–389. doi: 10.1038/38738. [DOI] [PubMed] [Google Scholar]
  58. Milligan CJ, Buckley NJ, Garret M, Deuchars J, Deuchars SA. Evidence for inhibition mediated by coassembly of GABAA and GABAC receptor subunits in native central neurons. J Neurosci. 2004;24:7241–7250. doi: 10.1523/JNEUROSCI.1979-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Möhler H. GABAA receptor diversity and pharmacology. Cell Tissue Res. 2006;326:505–516. doi: 10.1007/s00441-006-0284-3. [DOI] [PubMed] [Google Scholar]
  60. Mody I, Pearce RA. Diversity of inhibitory neurotransmission through GABA(A) receptors. Trends Neurosci. 2004;27:569–575. doi: 10.1016/j.tins.2004.07.002. [DOI] [PubMed] [Google Scholar]
  61. Moragues N, Ciofi P, Lafon P, Odessa MF, Tramu G, Garret M. cDNA cloning and expression of a gamma-aminobutyric acid A receptor epsilon-subunit in rat brain. Eur J Neurosci. 2000;12:4318–4330. [PubMed] [Google Scholar]
  62. Moragues N, Ciofi P, Tramu G, Garret M. Localisation of GABA(A) receptor epsilon-subunit in cholinergic and aminergic neurones and evidence for co-distribution with the theta-subunit in rat brain. Neuroscience. 2002;111:657–669. doi: 10.1016/s0306-4522(02)00033-7. [DOI] [PubMed] [Google Scholar]
  63. Mortensen M, Smart TG. Extrasynaptic alphabeta subunit GABAA receptors on rat hippocampal pyramidal neurons. J Physiol. 2006;577:841–856. doi: 10.1113/jphysiol.2006.117952. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Nusser Z, Mody I. Selective modulation of tonic and phasic inhibitions in dentate gyrus granule cells. J Neurophysiol. 2002;87:2624–2628. doi: 10.1152/jn.2002.87.5.2624. [DOI] [PubMed] [Google Scholar]
  65. Nusser Z, Roberts JD, Baude A, Richards JG, Sieghart W, Somogyi P. Immunocytochemical localization of the alpha 1 and beta 2/3 subunits of the GABAA receptor in relation to specific GABAergic synapses in the dentate gyrus. Eur J Neurosci. 1995;7:630–646. doi: 10.1111/j.1460-9568.1995.tb00667.x. [DOI] [PubMed] [Google Scholar]
  66. Nusser Z, Sieghart W, Somogyi P. Segregation of different GABAA receptors to synaptic and extrasynaptic membranes of cerebellar granule cells. J Neurosci. 1998;18:1693–1703. doi: 10.1523/JNEUROSCI.18-05-01693.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Ogris W, Lehner R, Fuchs K, Furtmuller B, Hoger H, Homanics GE, Sieghart W. Investigation of the abundance and subunit composition of GABAA receptor subtypes in the cerebellum of alpha1-subunit-deficient mice. J Neurochem. 2006;96:136–147. doi: 10.1111/j.1471-4159.2005.03509.x. [DOI] [PubMed] [Google Scholar]
  68. Olsen RW, Homanics GE. Function of GABAA receptors: insights from mutant and knockout mice. In: Martin DL, Olsen RW, editors. In GABA in the Nervous System: The View at 50 Years. Philadelphia: Lippincott, Williams & Wilkins; 2000. pp. 81–96. [Google Scholar]
  69. Olsen RW, Sieghart W. International Union of Pharmacology. (***) Subtypes of γ-aminobutyric acidA receptors: classification on the basis of subunit composition, pharmacology, and function. 2008 Update. Pharmacol Rev in press. [Google Scholar]
  70. Pan ZH, Zhang D, Zhang X, Lipton SA. Evidence for coassembly of mutant GABAC rho1 with GABAA gamma2S, glycine alpha1 and glycine alpha2 receptor subunits in vitro. Eur J Neurosci. 2000;12:3137–3145. doi: 10.1046/j.1460-9568.2000.00198.x. [DOI] [PubMed] [Google Scholar]
  71. Persohn E, Malherbe P, Richards JG. Comparative molecular neuroanatomy of cloned GABAA receptor subunits in the rat CNS. J Comp Neurol. 1992;326:193–216. doi: 10.1002/cne.903260204. [DOI] [PubMed] [Google Scholar]
  72. Pirker S, Schwarzer C, Wieselthaler A, Sieghart W, Sperk G. GABA(A) receptors: immunocytochemical distribution of 13 subunits in the adult rat brain. Neuroscience. 2000;101:815–850. doi: 10.1016/s0306-4522(00)00442-5. [DOI] [PubMed] [Google Scholar]
  73. Poltl A, Hauer B, Fuchs K, Tretter V, Sieghart W. Subunit composition and quantitative importance of GABAA receptor subtypes in the cerebellum of mouse and rat. J Neurochem. 2003;87:1444–1455. doi: 10.1046/j.1471-4159.2003.02135.x. [DOI] [PubMed] [Google Scholar]
  74. Reynolds DS, Rosahl TW, Cirone J, O'Meara GF, Haythornthwaite A, Newman RJ, Myers J, Sur C, Howell O, Rutter AR, Atack J, Macaulay AJ, Hadingham KL, Hutson PH, Belelli D, Lambert JJ, Dawson GR, McKernan R, Whiting PJ, Wafford KA. Sedation and anesthesia mediated by distinct GABA(A) receptor isoforms. J Neurosci. 2003;23:8608–8617. doi: 10.1523/JNEUROSCI.23-24-08608.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Rossi DJ, Hamann M, Attwell D. Multiple modes of GABAergic inhibition of rat cerebellar granule cells. J Physiol. 2003;548:97–110. doi: 10.1113/jphysiol.2002.036459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Rudolph U, Antkowiak B. Molecular and neuronal substrates for general anaesthetics. Nat Rev Neurosci. 2004;5:709–720. doi: 10.1038/nrn1496. [DOI] [PubMed] [Google Scholar]
  77. Rudolph U, Crestani F, Benke D, Brunig I, Benson JA, Fritschy JM, Martin JR, Bluethmann H, Mohler H. Benzodiazepine actions mediated by specific gamma-aminobutyric acid(A) receptor subtypes. Nature. 1999;401:796–800. doi: 10.1038/44579. [DOI] [PubMed] [Google Scholar]
  78. Rudolph U, Mohler H. Analysis of GABAA receptor function and dissection of the pharmacology of benzodiazepines and general anesthetics through mouse genetics. Annu Rev Pharmacol Toxicol. 2004;44:475–498. doi: 10.1146/annurev.pharmtox.44.101802.121429. [DOI] [PubMed] [Google Scholar]
  79. Saxena NC, Macdonald RL. Assembly of GABAA receptor subunits: role of the delta subunit. J Neurosci. 1994;14:7077–7086. doi: 10.1523/JNEUROSCI.14-11-07077.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Semyanov A, Walker MC, Kullmann DM, Silver RA. Tonically active GABAA receptors: modulating gain and maintaining the tone. Trends Neurosci. 2004;27:262–269. doi: 10.1016/j.tins.2004.03.005. [DOI] [PubMed] [Google Scholar]
  81. Sieghart W. Multiplicity of GABAA--benzodiazepine receptors. Trends Pharmacol Sci. 1989;10:407–411. doi: 10.1016/0165-6147(89)90189-2. [DOI] [PubMed] [Google Scholar]
  82. Sieghart W. Structure and pharmacology of γ-aminobutyric acidA receptor subtypes. Pharmacol Rev. 1995;47:181–234. [PubMed] [Google Scholar]
  83. Sieghart W, Sperk G. Subunit composition, distribution and function of GABAA receptor subtypes. Curr Top Med Chem. 2002;2:795–816. doi: 10.2174/1568026023393507. [DOI] [PubMed] [Google Scholar]
  84. Sigel E, Buhr A. The benzodiazepine binding site of GABAA receptors. Trends Pharmacol Sci. 1997;18:425–429. doi: 10.1016/s0165-6147(97)01118-8. [DOI] [PubMed] [Google Scholar]
  85. Simon J, Wakimoto H, Fujita N, Lalande M, Barnard EA. Analysis of the set of GABA(A) receptor genes in the human genome. J Biol Chem. 2004;279:41422–41435. doi: 10.1074/jbc.M401354200. [DOI] [PubMed] [Google Scholar]
  86. Sine SM, Engel AG. Recent advances in Cys-loop receptor structure and function. Nature. 2006;440:448–455. doi: 10.1038/nature04708. [DOI] [PubMed] [Google Scholar]
  87. Smith GB, Olsen RW. Functional domains of GABAA receptors. Trends Pharmacol Sci. 1995;16:162–168. doi: 10.1016/s0165-6147(00)89009-4. [DOI] [PubMed] [Google Scholar]
  88. Sohal VS, Keist R, Rudolph U, Huguenard JR. Dynamic GABA(A) receptor subtype-specific modulation of the synchrony and duration of thalamic oscillations. J Neurosci. 2003;23:3649–3657. doi: 10.1523/JNEUROSCI.23-09-03649.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Somogyi P, Fritschy JM, Benke D, Roberts JD, Sieghart W. The gamma 2 subunit of the GABAA receptor is concentrated in synaptic junctions containing the alpha 1 and beta 2/3 subunits in hippocampus, cerebellum and globus pallidus. Neuropharmacology. 1996;35:1425–1444. doi: 10.1016/s0028-3908(96)00086-x. [DOI] [PubMed] [Google Scholar]
  90. Stell BM, Brickley SG, Tang CY, Farrant M, Mody I. Neuroactive steroids reduce neuronal excitability by selectively enhancing tonic inhibition mediated by delta subunit-containing GABAA receptors. Proc Natl Acad Sci U S A. 2003;100:14439–14444. doi: 10.1073/pnas.2435457100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Sternfeld F, Carling RW, Jelley RA, Ladduwahetty T, Merchant KJ, Moore KW, Reeve AJ, Street LJ, O'Connor D, Sohal B, Atack JR, Cook S, Seabrook G, Wafford K, Tattersall FD, Collinson N, Dawson GR, Castro JL, MacLeod AM. Selective, orally active gamma-aminobutyric acidA alpha5 receptor inverse agonists as cognition enhancers. J Med Chem. 2004;47:2176–2179. doi: 10.1021/jm031076j. [DOI] [PubMed] [Google Scholar]
  92. Sur C, Quirk K, Dewar D, Atack J, McKernan R. Rat and human hippocampal alpha5 subunit-containing gamma-aminobutyric AcidA receptors have alpha5 beta3 gamma2 pharmacological characteristics. Mol Pharmacol. 1998;54:928–933. doi: 10.1124/mol.54.5.928. [DOI] [PubMed] [Google Scholar]
  93. Tretter V, Ehya N, Fuchs K, Sieghart W. Stoichiometry and assembly of a recombinant GABAA receptor subtype. J Neurosci. 1997;17:2728–2737. doi: 10.1523/JNEUROSCI.17-08-02728.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Tretter V, Hauer B, Nusser Z, Mihalek RM, Hoger H, Homanics GE, Somogyi P, Sieghart W. Targeted disruption of the GABA(A) receptor delta subunit gene leads to an up-regulation of gamma 2 subunit-containing receptors in cerebellar granule cells. J Biol Chem. 2001;276:10532–10538. doi: 10.1074/jbc.M011054200. [DOI] [PubMed] [Google Scholar]
  95. Unwin N. Structure and action of the nicotinic acetylcholine receptor explored by electron microscopy. FEBS Lett. 2003;555:91–95. doi: 10.1016/s0014-5793(03)01084-6. [DOI] [PubMed] [Google Scholar]
  96. Unwin N. Refined structure of the nicotinic acetylcholine receptor at 4A resolution. J Mol Biol. 2005;346:967–989. doi: 10.1016/j.jmb.2004.12.031. [DOI] [PubMed] [Google Scholar]
  97. Wallner M, Hanchar HJ, Olsen RW. Ethanol enhances alpha 4 beta 3 delta and alpha 6 beta 3 delta gamma-aminobutyric acid type A receptors at low concentrations known to affect humans. Proc Natl Acad Sci U S A. 2003;100:15218–15223. doi: 10.1073/pnas.2435171100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Wegelius K, Pasternack M, Hiltunen JO, Rivera C, Kaila K, Saarma M, Reeben M. Distribution of GABA receptor rho subunit transcripts in the rat brain. Eur J Neurosci. 1998;10:350–357. doi: 10.1046/j.1460-9568.1998.00023.x. [DOI] [PubMed] [Google Scholar]
  99. Whiting PJ. GABA-A receptors: a viable target for novel anxiolytics? Curr Opin Pharmacol. 2006;6:24–29. doi: 10.1016/j.coph.2005.08.005. [DOI] [PubMed] [Google Scholar]
  100. Wisden W, Laurie DJ, Monyer H, Seeburg PH. The distribution of 13 GABAA receptor subunit mRNAs in the rat brain. I. Telencephalon, diencephalon, mesencephalon. J Neurosci. 1992;12:1040–1062. doi: 10.1523/JNEUROSCI.12-03-01040.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Wohlfarth KM, Bianchi MT, Macdonald RL. Enhanced neurosteroid potentiation of ternary GABA(A) receptors containing the delta subunit. J Neurosci. 2002;22:1541–1549. doi: 10.1523/JNEUROSCI.22-05-01541.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Yamakura T, Bertaccini E, Trudell JR, Harris RA. Anesthetics and ion channels: molecular models and sites of action. Annu Rev Pharmacol Toxicol. 2001;41:23–51. doi: 10.1146/annurev.pharmtox.41.1.23. [DOI] [PubMed] [Google Scholar]
  103. Yee BK, Keist R, von Boehmer L, Studer R, Benke D, Hagenbuch N, Dong Y, Malenka RC, Fritschy JM, Bluethmann H, Feldon J, Mohler H, Rudolph U. A schizophrenia-related sensorimotor deficit links alpha 3-containing GABAA receptors to a dopamine hyperfunction. Proc Natl Acad Sci U S A. 2005;102:17154–17159. doi: 10.1073/pnas.0508752102. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES