IMR Press / JIN / Volume 20 / Issue 1 / DOI: 10.31083/j.jin.2021.01.284
Open Access Review
Extrasynaptic δ-subunit containing GABAA receptors
Show Less
1 School of Advanced Studies, University of Tyumen, 9 Marta Street 2k1, 625000 Tyumen, Russia
*Correspondence: a.arslan@utmn.ru (Ayla Arslan)
J. Integr. Neurosci. 2021, 20(1), 173–184; https://doi.org/10.31083/j.jin.2021.01.284
Submitted: 17 September 2020 | Revised: 18 October 2020 | Accepted: 1 February 2021 | Published: 30 March 2021
Copyright: © 2021 The Authors. Published by IMR Press.
This is an open access article under the CC BY 4.0 license (https://creativecommons.org/licenses/by/4.0/).
Abstract

γ-Aminobutyric acid type A receptors (GABAARs) are GABA gated heteropentameric chloride channels responsible for the adult brain’s primary inhibition. In specific brain cells, such as in the hippocampus, one of the subtypes of GABAARs, the δ subunit containing GABAARs (δ-GABAARs), is predominantly expressed and located in extrasynaptic or perisynaptic positions. δ-GABAARs mediate a slow constant inhibitory current called tonic inhibition. While δ-GABAARs and tonic inhibition is critical for the excitability of single neurons, accumulating data suggest that the function of δ-GABAARs are broader and includes an integrative role in the network oscillations. While these open new horizons on the neurobiology of δ-GABAARs, the complexity continues to challenge the analysis of GABAARs and their subtypes. This review will summarize the current knowledge of molecular, cellular and physiological characteristics of δ-GABAARs during health and disease.

Keywords
GABA
Inhibition
Tonic
Phasic
Synaptic
Extrasynaptic
Cys-loop receptors
δ-subunit
GABA(A) receptor
GABA receptor
Ion channel
GABRD
Hippocampus
Dentate gyrus
GABAergic interneurons
Granule cells
Epilepsy
Anxiety
Postpartum
Depression
1. Introduction

Hippocampus is a unique structure. One aspect of this uniqueness is its special anatomy. Concealed between the mesencephalon and the medial temporal lobe, this deep cortical structure extends through the lateral ventricle’s inferior horn, where it lies at the posterior border of the amygdala [1]. Critical for learning and memory, the hippocampus is segmented into several regions [2], including the hippocampus proper, CA3, CA2, CA1 regions and the dentate gyrus (DG), a key region in hippocampal memory formation (reviewed in [3]).

A principal cell type in the DG is the dentate gyrus granule cells (DGGCs), characterized by unique anatomical features (reviewed in [4]). The cone-shaped spiny dendritic arbors of the DGGCs are innervated by different neuronal ensembles such as the input from the entorhinal cortex via the perforant path and contralateral hippocampus via the commissural path [5, 6, 7]. Diverse GABAergic interneurons synapse on the soma, axon initial segment, proximal and distal dendrites of DGGCs. For example, parvalbumin-positive interneurons (PPI) synapse on the axon- initial segment and the perisomatic domain [8]. These GABAergic interneuron inputs to the DGGCs are involved in the synchronization of the network activities during theta- frequency (4-10 Hz) and gamma-frequency (30-150 Hz) oscillations [9], sharp waves-ripples (SWRs) [10], and dentate spikes [11].

The critical network operations for neuronal synchronization require the presynaptic terminals of the GABAergic interneurons (such as PPIs) to precisely match their molecular counterparts at the postsynaptic sites of the DGGCs. Here, γ-Aminobutyric acid type A receptors (GABAARs), the GABA gated heteropentameric chloride channels, are massively clustered in the postsynaptic sites of the symmetric inhibitory synapses. GABAARs belong to the superfamily of ligand-gated ion channels (Cys-loop receptors) [12], which also includes the nicotinic acetylcholine receptors (nAChRs), the 5-hydroxytryptamine type 3 (5-HT3) receptors, the zinc-activated ion channel (ZAC) and the glycine receptors in vertebrates [13]. Upon GABA release, the postsynaptic GABAARs in the mature granule cells become active and elicit hyperpolarizing inhibitory postsynaptic currents (IPSCs), during which chloride and bicarbonate ions will travel through the receptor channel depending on their electrochemical gradient. Known to be benzodiazepine (BZ) sensitive [14, 15], these IPSCs are called phasic inhibition. The phasic signals are typically generated rapidly (often with sub-millisecond rise times), with the stimulus-evoked synaptic currents being in the range of less than 10 to 200 pA at a holding potential of -50 mV, which is known to vary in a typical quantal fashion [16].

In addition to the phasic synaptic inhibition, the DGGCs and PPIs have some other spots where a subset of high-affinity extrasynaptic GABAARs is strategically located in the hippocampus mediate a different tone of GABAergic inhibition than phasic inhibition. This type of GABAergic signal is called tonic inhibition. Tonic inhibition is characterized by constant, slow currents, high GABA affinity, slow desensitization and BZ insensitivity [17]. The tonic current is about four times larger than the total phasic current in the DGGCs [18]. Like cerebellar granule cells [19], where the tonic inhibition was first described, distinct subtypes of extrasynaptic GABAARs appear to mediate these relatively constant and slow inhibitory currents [18], which have also been shown in the neurons found almost all other major brain areas: neocortex, thalamus, hypothalamus and brain stem [20]. This distributed fashion of tonic inhibition, mediated by GABAARs, represents distinct subunit composition, which involves either α5 or δ subunits [21]. This review will focus on the δ subunit containing GABAARs (δ-GABAARs), which mediate a significant fraction of tonic current.

2. Molecular and cellular properties of δ-GABAARs

For GABAAR research, the late 80s and 90s were exciting years. Almost the entire GABAAR subunit family was cloned by Seeburg and his colleagues [22, 23, 24]. The cloning strategy was based on the classical approach: Screening the brain cDNA libraries by synthetic DNA probes derived from purified receptors’ peptides. Thus, eventually, it became clear that GABAARs were assembled from 19 subunit isoforms (α(1-6), β(1-3), γ(1-3), δ, ϵ, θ, π and ρ(1-3)) which correspond to 11 structurally and functionally distinct receptor subtypes [22, 23, 24]. In general, all these subunits share a common topological structure: a peptide sequence which is about 450 amino acids long, made up of a long extracellular N-terminal, a short C-terminal, four transmembrane domains, intracellular or cytoplasmic domain located between the third and the fourth transmembrane domains (Fig. 1). This organization was originally based on the structural studies of acetylcholine-binding protein and nAChRs [25]. In particular, the subunits of acetylcholine receptors and the human GABAAR β3 homopentamer’s crystal structure at 3Åresolution confirmed this prediction [25, 26]. In contrast to the subunits’ above-described properties, the hetero-pentameric receptor structure was not fully known until recently. In recent years, oligomerized heteropentameric receptor structure has also been resolved in detail [27, 28, 29, 30].

Fig. 1.

Basic structure of δ subunit. δ subunit shares a standard topological structure with other subunits of GABAARs: a long extracellular N-terminal, a short extracellular C-terminal, four transmembrane domains (TM1, TM2, TM3, TM4), cytoplasmic domain located between the third (TM3) and the fourth (TM4) transmembrane domains (Figure not to scale).

It is well known that the GABAAR subunit composition determines their differential distribution and functionality [31, 32, 33, 34, 35, 36, 37, 38]. Among the possible subunit combinations, typically, there is a combination of 2α and 2β subunits and a single γ2 or δ subunit (Fig. 2), the 2α, 2β and γ2 combinations being the most abundant. Indeed, about 90% of all GABAARs are made up of γ2-GABAARs [33]. Thus, the most GABAAR research is directed to α, β and γ subunits, which are found both in the postsynaptic and extrasynaptic locations [39]. Among the α subunits, the BZ insensitive α4 and α6 subunits form a unique partnership with the δ subunit (together with β subunit isoforms) in the forebrain and cerebellum, respectively [36]. Thus, in the arrangement of δ-GABAARs, δ subunit has been hypothesized as a replacement of the γ2 subunit in the receptor heteropentamer recruited exclusively to extrasynaptic or perisynaptic locations [40]. In the DGGCs, α4βδ receptors, the most common isoform of δ-GABAARs, are expressed. Also, α4βδ receptors have been identified in several other neuronal cell types (see also Table 1) [41, 42], and like other δ-GABAAR isoforms, localized in the extrasynaptic and perisynaptic positions but never in the postsynaptic sites [43, 44].

Fig. 2.

Heteropentameric structure of δ-GABA𝐀Rs. The diagram showing the heteropentameric structure of GABAARs, which are typically composed of two α, two β and one γ or δ subunit. Experimental studies suggest that the subunit arrangement of γ2-subunit containing GABAARs (γ2-GABAARs) is counterclockwise when viewed from the extracellular space. It is not known if this arrangement also applies to δ-subunit containing GABAARs (δ-GABAARs) (Figure not to scale).

Table 1.Summary of δ-GABA𝐀R isoforms found in specific cells in the forebrain and cerebellum.
Subunit composition Cell type Reference
α6bδ Cerebellar granule cells [46]
α1bδ Hippocampal interneurons, Neocortical interneurons [47, 48]
α4β2δ Thalamic relay neurons, Striatal spiny neurons, Hippocampal dentate granule cells, Neocortical pyramidal cells [41, 42, 49, 50, 51, 52, 53]

By in situ hybridization analysis, the regions of the adult rat brain in which δ subunits are expressed have been studied in detail [45]: The δ-subunit is expressed weakly or moderately in the regions of the olfactory bulb (granule cells and periglomerular), neocortex (layer II/III, layer IV, layer V/VI and pyriform cortex), hippocampus (DGGCs, stratum pyramidalis CA1 and stratum pyramidalis CA3), basal ganglia (caudate, putamen, nucleus accumbens, claustrum), thalamus (mediodorsal, ventral posterior nucleus, medial-, dorso- and ventrolateral geniculate nucleus). In Table 1, a summary of δ-GABAAR isoforms (e.g., α4βδ, α6βδ, or α1βδ) and their cell type specific distribution are shown. This specific distribution is well reflected with the tonic inhibition. For example, α4βδ receptors mediate the larger fraction (> 70%) of the tonic inhibition in the DGGCs [21].

3. Variety of δ-GABAAR mediated inhibition

It is known that GABA mediates multiple forms of postsynaptic inhibitory signals, such as fast and slow inhibitory postsynaptic currents [54, 55]. Additionally, δ-GABAARs, mediating tonic inhibition and characterized by relatively constant, slow IPSC, has been known for the last few decades. However, tonic inhibition with these characteristics has been considered uniform and the only inhibition associated with the δ-GABAARs. For example, the α4βδ receptors in DGGCs and thalamic relay neurons mediate such tonic currents [41, 42]. These BZ insensitive GABAARs have a high affinity for the GABA diffused from the synaptic cleft [56], besides the GABA released from GABA transporters [57, 58]. Interestingly, the literature has started to dissect the tonic inhibition: Depending on the subunit co-assembly, δ-GABARs have different GABA sensitivity, desensitization, and kinetics [59]. For example, it was shown that the α4βδ GABAARs are the most sensitive to GABA levels ranging from 100 nM to 800 nM. Whereas α1β2δ and α5β3γ2 (in addition to α1β2γ2) receptors detect GABA levels 1-10 μM range [59].

Accumulating data suggest that extrasynaptic GABAARs might mediate a significant part of tonic inhibition, independent of gating by GABA; thus, spontaneous activity could occur [60, 61]. Such spontaneous activity also applies to δ-GABAARs mediated tonic inhibition [62]. This phenomenon’s functional significance is not understood, and it is probably dependent on the specific cell types and isoforms of δ-GABAARs, such as α1βδ and α4βδ expressed in these cells.

In addition to studies focusing on δ-GABAARs, some studies dissect the physiological roles of GABAergic inhibition without explicitly indicating the associated subunit. So far, a few types of GABAARs such as the ones containing either the δ, α5 subunits or receptors containing only αβ subunits have been shown to mediate the tonic inhibition [63]. Thus, it is hard to predict the role of δ-GABAARs in these studies. For example, in mice, in the reticular thalamic neurons, a phasic inhibition with slowed-down kinetics is mediated by GABAARs [64]. This association is linked to α4 containing GABAARs, but the exact receptor co-assembly is not clear. Possibly δ-GABAARs might mediate this activity because, in the thalamus, most of the α4 containing receptors involve δ-subunit [41]. This is supported by some other findings, too. For example, the δ subunit is expressed explicitly in the thalamus [45], including the reticular thalamic nucleus [38]. However, this latter study represents the monkey brain, reflecting some differences compared to the rodent brain. It turns out that, in rat and mouse brains, δ-subunit is not expressed in the reticular thalamic nucleus [38, 65], whereas in the monkey, it is [38]. Thus, it is not clear if the α4 subunit linked phasic inhibition with slowed-down kinetics [64] is mediated by δ-GABAARs even though the specific partnership of δ subunit with α4 subunit in the forebrain, including the thalamus, is well known [31, 41, 42, 66, 67].

Nevertheless, there is a collection of data supporting an additional GABAergic inhibition representing an intermediate form between the classical phasic (GABA, fast) and tonic inhibition, which is called GABAA, slow [54, 55] some of which may be mediated by δ-GABAARs as experimental evidence supports that δ-GABAARs contribute to postsynaptic inhibition. Postsynaptic inhibition contributed by δ-GABAARs was observed in the cerebellum, thalamus and neocortex [68]; in DGGCs of the mouse hippocampus [69, 70]. Thus, Fig. 3 shows different types of inhibition mediated by δ-GABAARs as a proposition. For reference, postsynaptic γ2-GABAARs, which mediate phasic inhibition, are also shown (Fig. 3).

Fig. 3.

Variations of GABAergic inhibition. Fast, point to point, phasic inhibition is typically mediated by synaptic GABAARs, clustered in the postsynaptic membrane of the inhibitory synapses. These receptors evoke inhibitory postsynaptic current (IPSC) (phasic inhibition) in a millisecond range upon GABA binding. The GABA spillover from the synaptic region (black arrows) results in extrasynaptic receptors, which mediate a slow inhibitory conductance, the tonic inhibition. Phasic and tonic inhibition of synaptic and extrasynaptic GABAARs has led to a functional distinction of these receptor subtypes. On the other hand, subsets of GABAARs, including δ-GABAARs may have intermediate activation, desensitization, and deactivation rates determined by the receptor subunit isoforms between these two states. This leads to the idea that δ-GABAARs may contribute to postsynaptic inhibitory currents (IPSCs) (Figure not to scale).

4. Variety of functions

As mentioned above, at the neuronal level, δ-GABAARs mediated tonic inhibition, which is important for the threshold of action potential generation [36, 71, 72, 73]. It is generally hypothesized that tonic inhibition decreases neuronal excitability. Recent evidence-based computer models revealed that tonic inhibition might also increase excitability [74].

Increasing literature shows the critical role of the nonsynaptic GABAAR and/or tonic inhibition in various functions, including network oscillations [64, 75, 76], synaptic plasticity [77], synaptic pruning during adolescence [78], neurogenesis [79, 80], neuronal development [81], information processing, and cognition [81]. For example, in the dentate gyrus, δ-subunit is linked to enhanced memory and neurogenesis [82].

δ-GABAARs mediated tonic inhibition is indicated for modulation of γ oscillations in the mouse hippocampal CA3 interneurons [75]. Also, coupling presynaptic activity to postsynaptic Inhibition in the somatosensory thalamus involved a process that influenced the δ-selective allosteric modulator, DS2 [76]. These take δ-GABAARs from being the mediators of “shunting” inhibition involved in controlling neuronal excitability to additional roles in the network level activities, including but not limited to the thalamocortical system and neurogenesis in the hippocampus.

5. δ-GABAARs and associated pathophysiology

The δ subunit modulators such as sedative and hypnotic agents [83], anxiolytic and anticonvulsive agents [84, 85] suggest that δ subunit may play a role in the etiology of the relevant disorders. Alterations of δ subunit or their modulation as therapeutic targets have been linked to sex specific behavioral disruption [86], Alzheimer’s disease [87], stress induced deficiency in learning and memory [88], fragile X syndrome [89] schizophrenia [90], epilepsy [91], mood disorders [92, 93, 94], childhood mood disorders [95], anxiety in methamphetamine dependence [96], major depression [97]; post-partum depression, and post-partum psychosis [94, 98], consumption of opioids [99], menstrual cycle related problems [100, 101], stroke [102], Fragile X Syndrome [89, 103], traumatic brain injury [104, 105], Huntington’s disease [106], pain [107], insomnia [83, 108, 109, 110], alcohol use disorders [111].

In animal studies, alcohol use disorders or associated behavioral alterations have been linked to δ-GABAARs [112, 113] and sex-dependent [114] as well as developmental [115] factors seem to play a role in the underlying mechanisms. At the molecular level, ethanol impacts the modulation of the clathrin adaptor-mediated endocytosis of δ-GABAARs [116], and its withdrawal influences δ-GABAARs via PKCδ Activation [117]. Due to the estrous cycle-dependent plasticity of δ-GABAARs, which was previously shown as associated with seizure susceptibility and anxiety [100], one study, using the model of “Drinking-in-the-Dark binge-drinking”, showed that δ-GABAARs are a critical target for binge drinking in females, a phenomenon observed at higher rates among women and girls [118]. The methylation pattern of δ subunit was also suggested as a diagnostic biomarker for alcohol use disorders [111].

It is important to talk about the special link between the δ-subunit and epilepsy. Various mutations (missense, nonsense, and frameshift mutations in coding DNA sequences besides mutations in the intronic, 3’ downstream, or 5’ upstream mutations) in GABA receptor subunit encoding genes have been linked to consequences such as the distortion of protein structure, conformation, abundance, or localization. Some of these mutations, which are detected in α1, β3, γ2, and δ subunits, have been associated with idiopathic generalized epilepsies (IGEs). For example, mutations in the γ2 subunit are characterized by change of a single amino acid (γ2(Q351X [119], γ2(R43Q) [120], and premature translation-termination codon (PTC)-generating mutations γ2(Q351X) [121]) are associated with different IGEs. Two δ subunit missense mutations, namely δ(E177A) and δ(R220H), were reported [122, 123]. Due to the distortion in the coding sequence, missense mutations lead to an altered amino acid sequence in the signal peptide regions of mature peptide regions. Dibbens et al. [123] reported mutations in the genomic region (1p36.3) of the δ subunit, representing susceptibility locus for generalized epilepsies. The δ subunit missense mutations, located in the subunit’s extracellular N-terminus, are associated with generalized epilepsy with febrile seizures plus (GEFS+), a type of IGEs. These mutations alter the channel conductance [123], gating and surface expression of δ-GABAARs [122]. Thus, δ-GABAARs are considered as targets in the treatment of epilepsy.

Neurosteroids are endogenous substances synthesized from cholesterol into pregnenolone, which is then converted to compounds such as allopregnanolone and allotetrahydrodeoxycorticosterone [124]. It is suggested that fluctuations in neurosteroid interactions, such as those seen during stress or the ovarian cycle, determine the seizure threshold, a phenomenon that is partially mediated by δ-GABAARs [100]. This and other evidence [125, 126] suggest that neurosteroids are novel drug candidates for epileptic disorders [125, 128]. Consequently, due to their potent actions on δ-GABAARs [128, 129], δ-GABAARs are novel therapeutic targets for the treatment of epileptic disorders and maybe a future perspective to control epileptogenesis [91, 130]. Ganaxolone, the synthetic analog of endogenous neurosteroid, is used as an antiepileptic agent (catamenial epilepsy), although it is the modulator of all GABAARs, it shows a higher effect on δ-GABAARs [131, 132, 133, 134].

Interestingly, the modulation and pharmacology of δ-GABAARs have become more critical recently. In addition to their modulation by insulin [135] and oxytocin [136], recently in 2019, the allopregnanolone brexanolone (ZulressoTM, the brand of Sage Therapeutics, Inc.), one of the neurosteroids known as a potent modulator of δ-GABAARs has been approved by the Food and Drug Administration (FDA) for postpartum depression1 (1Drug Approval Package, FDA (https://www.accessdata.fda.gov/drugsatfda_docs/nda/2019/211371Orig1s000TOC.cfm), accessed in 28.01.2021.) as a result of successful clinical trials [137, 138, 139]. Brexanolone seems to be effective on other mood disorders, such as major unipolar depression and post-traumatic stress disorder [140]. A synthetic GABAAR modulator that shares a similar molecular pharmacological profile as brexanolone, the zuranolone (SGE-217), resulted in a reduction in depressive symptoms according to a recent phase 2 clinical trial [141].

Despite the progress, the field is dominated by many unknowns, which is a significant bottleneck. For example, the above-mentioned preferential modulation of δ-GABAARs by neurosteroids is controversial and requires further validation. Regarding this, some studies suggested that the neurosteroid sensitivity of α4/δ-containing extrasynaptic receptors may not be different than that of α/β/γ2-containing receptors [142, 143]. Along with the other inconsistencies, which will be summarized in the section “7. The basics of unknowns”, more research is needed for δ-GABAARs.

6. A circuit pharmacology for δ-GABAARs

The variations and specificities of δ-GABAARs in terms of their isoforms, inhibitory action, distribution, sensitivity, modulation and spontaneous activity, which have been described so far, lead to the question to ask whether these properties can be utilized for circuit pharmacology. The idea of GABAAR circuit pharmacology has probably gained momentum when the diversity of subunits and their specific pharmacology in the subunit assembly have started to be shown [144, 145]. However, the focus was mainly on the modulators of α subunit isoforms [23, 144, 145, 146] such as α5 inverse agonists RO4938581 [147]; S44819 [148], L-655,708 [149], Alpha5IA [150]. RO4938581 is under preclinical investigation for its potential to cure cognitive deficits in people with Down syndrome [151], for example.

Since the subunit-specific function and specific modulation are key to the strategy of circuit pharmacology, δ-GABAAR seem to fit into this strategy. Among the isoforms of δ-GABAARs, two population receptors are expressed in the hippocampus. α1βδ receptors are expressed predominantly in hippocampal interneurons, whereas α4βδ receptors are expressed predominantly in granule cells of the dentate gyrus (DGGCs) (Table 1). One study selectively silenced one population of these isoforms: α1βδ expressed in the Parvalbumin positive interneurons [152]. Thus, using the “PV/Cre-Gabrd/floxed system”, it was reported that in vitro γ oscillations in the CA3 region were altered in both PV-Gabrd(+/-) and PV-Gabrd(-/-) mice in these interneurons. Interestingly, the increased γ oscillations were lowered to control PV-Gabrd(+/-) levels when 100 nM allopregnanolone (3α,5α-tetrahydroprogesterone) was used. But when 10 μM synthetic δ-GABAAR positive allosteric modulator 4-Chloro-N-[2-(2-thienyl)imidazo[1,2-a]pyridin-3-yl] benzamide (DS-2) was used, this was not observed. DS-2 selectively targets α4βδ receptors but not the α1βδ receptors, which are expressed in the interneurons. These suggest the specific role of α1βδ isoform in the hippocampus’s integrative network operations, in a way that can be modulated by selective agents. In line with this, another study examined the paired whole-cell recordings from synaptically coupled reticular thalamus and thalamocortical neurons of the ventrobasal complex in brain slices of α4 knock-out (α4(0/0)) mice. Results suggest a dynamic and activity-dependent engagement of δ-GABAARs receptors for the coupling of presynaptic activity to postsynaptic excitability, a process sensitive to DS2, the specific modulator α4/β/δ receptors [76]. The resolution of the three-dimensional structure of GABAAR subtypes in recent years will trigger the design of novel drugs targeting specific δ-GABAAR isoforms, which will likely aid the treatment of network disorders by circuit pharmacology approach.

7. The remaining unknowns

GABAAR research has been challenged by the receptors’ unusual molecular and cellular diversity and thus a huge effort is required to fully understand the properties of GABAARs. Here, we will briefly mention the unknowns related to molecular and modulatory features of δ-GABAARs, only. The nonsynaptic localization δ-GABAARs is well established by electron microscopy studies [43, 44]. However, it is unknown if a passive or an active mechanism mediates this specific nonsynaptic localization pattern. Previously, it was suggested that the subunit’s intracellular domain might play a role in this process [153]. The intracellular domain, which is found in between the third and the fourth transmembrane domains, is a large cytoplasmic domain, highly conserved across the whole span of vertebrate evolution [153].

Despite new studies [154, 155, 156], the current knowledge about the assembly and stoichiometry δ-GABAARs is limited. Several studies have shown the stoichiometry of δ-GABAARs as 2α, 2β and δ [157, 158]. For example, one recent study suggested that recombinant α1β3δ receptors have the same stoichiometry and subunit arrangement with α1β3γ2 receptors. However, these results are not entirely conclusive [155]. Thus, the basics such as assembly rules, stoichiometry, and arrangement of δ-GABAARs, and their membrane trafficking, maintenance and modulation are not precisely known. For instance, in the in vitro live neuroblastoma cells, our group reported that recombinant δ subunits require both α and β subunits for membrane targeting [159], confirming the previously hypothesized analogy (γ2 subunit is replaced by δ in the δ-GABAAR arrangement) between δ subunit and γ2 subunit: it is known that γ2 cannot assemble into receptors inserted in the cell membrane without α and/or β subunits [160, 161]. In contrast to our findings [159], some other previous studies suggest that βγ and βδ containing receptors exist and show functionality in Xenopus oocytes [162, 163]. So, there is no consensus. This may arise from the methodological variations used during in vitro studies: use of different vectors, cell types, or subunit isoforms, experimental strategy (such as fluorescent protein tagging location) may impact on these results. For example, in HEK-293T cells, quantification of fluorescent alpha-bungarotoxin bound subunits on Western blots of surface immunopurified tagged GABAARs led to the conclusion that the cell surface expression of α42β2δ- GABAARs was regulated by the ratio of subunit cDNAs transfected [164].

The distribution of δ subunit has been shown in different species, which shows species-specific variations. For instance, in the reticular thalamus, caudate, putamen and globus pallidus, there is an expression of δ subunit in the monkey, while this expression is absent in the rat [38]. The human brain distribution of δ subunit is not known fully. At the same time, some studies reported the distribution of α1-α3, β2/β3, and γ2 subunits in the human striatum [165] and thalamus [37].

Sensitivity to neuroactive steroids has also been questioned. Neuroactive steroids such as allopregnanolone (3α5αP) and allotetrahydrodeoxycorticosterone (THDOC) are considered to selectively affect δ-GABAARs over γ2-GABAARs. δ-GABAARs sensitivity to neurosteroids in specific brain regions [166] is hypothesized to be very specific such that the endogenous neurosteroid THDOC at physiologically relevant concentrations (10-100 nM) selectively increases the tonic current, with almost no effect on the phasic current in mouse dentate gyrus granule cells and cortical granule cells [128, 167]. Thus, selective interaction of δ-GABAARs with neurosteroids has been hypothesized to have clinical significance due to tonic inhibition’s modulation, impacting excitability, seizure susceptibility, and behavior [100]. On the other hand, the neurosteroid binding site has been identified in the transmembrane domain of the α-subunit [168]. Moreover, a recent study suggests that neurosteroids act through both δ-containing and non-δ-containing receptors [143]. Thus, the degree of neurosteroid selectivity of δ-GABAARs is questionable [142, 143].

Similarly, the mechanism by which ethanol potentiates GABAARs is still not fully understood, and several publications have reported contradicting results. In general, γ2-GABAAR subtypes are sensitive to ethanol at amounts required for high intoxication, whereas the extra-synaptic δ-GABAARs are hypothesized to be most sensitive to ethanol at levels of social drinking, that is less than 30 mM [47, 70, 113, 169, 170]. However, this has been challanged by some publications [171, 172].

8. Conclusions

Increasing studies open new horizons on the δ-GABAAR’s neurobiology; however, the complexity continues to be a challenge. On the one hand, it could turn out that δ-GABAARs function may be broader than previously hypothesized. This is well reflected with studies showing the possible contribution of δ-GABAAR mediated inhibition to the control of major thalamocortical oscillations. Also, the possibility of some other forms of phasic inhibition, with roles in the integrative function and network oscillations, may underlie an even broader spectrum of physiological functions of δ-GABAAR during health and disease.

On the other hand, knowledge is deficient in the level of “basics”. There is uncertainty regarding the knowledge about the assembly [155], membrane targeting [159], clustering [153] and modulation of δ-GABAARs [142], for example. Without elucidation of the mechanisms involved in these basic receptor mechanisms, precisely, it will be challenging to unravel the δ-GABAA receptor physiological significance and plasticity during health and disease. Thus, there is a need for a focused establishment of these “basics” in a subtype-specific fashion. Such an effort requires novel methodologies and careful consideration of experimental subject design. Experimental parameters appear to have a critical impact on the GABAAR research illustrated by the lack of convergent findings obtained by the experimentation on the same subject by different methods.

Abbreviations

BZs, Benzodiazepines; CNS, Central nervous system; DGGC, Dentate gyrus granule cell; DS-2, 4-Chloro-N-[2-(2-thienyl)imidazo[1,2-a]pyridin-3-yl] benzamide; FDA, Food and Drug Administration; GABA, γ-Aminobutyric acid; GABAARs, γ-Aminobutyric acid type A receptors; GABA(A) receptor, γ-Aminobutyric acid type A receptor; GABA-d, GABA receptor δsubunit; δ-GABAARs, δ subunit containing GABAARs; γ2-GABAARs, γ2 subunit containing GABAARs; HEK293T cells, Human embryonic kidney 293T cells; IPSC, Inhibitory postsynaptic currents; μM, Micro molar; N, Asparagine; nM, Nano molar; nAChRs, Nicotinic acetylcholine receptors; NMDA, N-methyl-D-aspartate receptor; pA, pico Amper; PPI, Parvalbumin positive interneurons; Q, Glutamine; THDOC, Allotetrahydrodeoxycorticosterone; W, Tryptophan.

Author contributions

AA conceptualized the study, identified the purpose and the scope, analyzed the literature, synthesized the knowledge and wrote the paper.

Acknowledgment

The author thanks three anonymous reviewers for excellent criticism of the article.

Conflict of interest

The authors declare no conflict of interest. Given her role as the Editorial Board Member of JIN, Prof. Ayla Arslan had no involvement in the peer-review of this article and has no access to information regarding its peer-review.

References
[1]
Destrieux C, Bourry D, Velut S. Surgical anatomy of the hippocampus. Neuro-Chirurgie. 2013; 59: 149-158.
[2]
Dalton MA, Zeidman P, Barry DN, Williams E, Maguire EA. Segmenting subregions of the human hippocampus on structural magnetic resonance image scans: an illustrated tutorial. Brain and Neuroscience Advances. 2017; 1: 2398212817701448.
[3]
Hainmueller T, Bartos M. Dentate gyrus circuits for encoding, retrieval and discrimination of episodic memories. Nature Reviews Neuroscience. 2020; 21: 153-168.
[4]
Scharfman HE, Myers CE. Hilar mossy cells of the dentate gyrus: a historical perspective. Frontiers in Neural Circuits. 2013; 6: 106.
[5]
Witter MP, Amaral DG. Entorhinal cortex of the monkey: V. projections to the dentate gyrus, hippocampus, and subicular complex. Journal of Comparative Neurology. 1991; 307: 437-459.
[6]
Halasy K, Somogyi P. Subdivisions in the multiple GABAergic innervation of granule cells in the dentate gyrus of the rat hippocampus. European Journal of Neuroscience. 1993; 5: 411-429.
[7]
Sik A, Penttonen M, Buzsáki G. Interneurons in the hippocampal dentate gyrus: an in vivo intracellular study. European Journal of Neuroscience. 1997; 9: 573-588.
[8]
Freund, T. F., Buzsáki, G. Interneurons of the hippocampus. Hippocampus. 1996; 6: 347-470.
[9]
Pernía-Andrade A, Jonas P. Theta-gamma-modulated synaptic currents in hippocampal granule cells in vivo define a mechanism for network oscillations. Neuron. 2014; 81: 140-152.
[10]
Szabo GG, Du X, Oijala M, Varga C, Parent JM, Soltesz I. Extended interneuronal network of the dentate gyrus. Cell Reports. 2017; 20: 1262-1268.
[11]
Penttonen M, Kamondi A, Sik A, Acsády L, Buzsáki G. Feed-forward and feed-back activation of the dentate gyrus in vivo during dentate spikes and sharp wave bursts. Hippocampus. 1997; 7: 437-450.
[12]
Barnard EA, Darlison MG, Seeburg P. Molecular biology of the GABAA receptor: the receptor/channel superfamily. Trends in Neurosciences. 1987; 10: 502-509.
[13]
Goetz T, Arslan A, Wisden W, Wulff P. GABA(A) receptors: structure and function in the basal ganglia. Progress in Brain Research. 2007; 160: 21-41.
[14]
De Koninck Y, Mody I. Noise analysis of miniature IPSCs in adult rat brain slices: properties and modulation of synaptic GABAA receptor channels. Journal of Neurophysiology. 1994; 71: 1318-1335.
[15]
Hájos N, Nusser Z, Rancz EA, Freund TF, Mody I. Cell type- and synapse-specific variability in synaptic GABAA receptor occupancy. European Journal of Neuroscience. 2000; 12: 810-818.
[16]
Edwards FA, Konnerth A, Sakmann B. Quantal analysis of inhibitory synaptic transmission in the dentate gyrus of rat hippocampal slices: a patch-clamp study. Journal of Physiology. 1990; 430: 213-249.
[17]
Stell BM, Mody I. Receptors with different affinities mediate phasic and tonic GABA(A) conductances in hippocampal neurons. Journal of Neuroscience. 2002; 22: RC223.
[18]
Nusser Z, Mody I. Selective modulation of tonic and phasic inhibitions in dentate gyrus granule cells. Journal of Neurophysiology. 2002; 87: 2624-2628.
[19]
Chadderton P, Margrie TW, Häusser M. Integration of quanta in cerebellar granule cells during sensory processing. Nature. 2004; 428: 856-860.
[20]
Farrant M, Nusser Z. Variations on an inhibitory theme: phasic and tonic activation of GABAA receptors. Nature Reviews Neuroscience. 2005; 6: 215-229.
[21]
Glykys J, Mann EO, Mody I. Which GABA(A) receptor subunits are necessary for tonic inhibition in the hippocampus? Journal of Neuroscience. 2008; 28: 1421-1426.
[22]
Schofield PR, Darlison MG, Fujita N, Burt DR, Stephenson FA, Rodriguez H, et al. Sequence and functional expression of the GABAA receptor shows a ligand-gated receptor super-family. Nature. 1987; 328: 221-227.
[23]
Pritchett DB, Sontheimer H, Shivers BD, Ymer S, Kettenmann H, Schofield PR, et al. Importance of a novel GABAA receptor subunit for benzodiazepine pharmacology. Nature. 1989; 338: 582-585.
[24]
Seeburg PH, Wisden W, Verdoorn TA, Pritchett DB, Werner P, Herb A, et al. The GABAA receptor family: molecular and functional diversity. Cold Spring Harbor Symposia on Quantitative Biology. 1990; 55: 29-40.
[25]
Unwin N. Nicotinic acetylcholine receptor and the structural basis of neuromuscular transmission: insights from Torpedo postsynaptic membranes. Quarterly Reviews of Biophysics. 2013; 46: 283-322.
[26]
Miller PS, Aricescu AR. Crystal structure of a human GABAA receptor. Nature. 2014; 512: 270-275.
[27]
Phulera S, Zhu H, Yu J, Claxton DP, Yoder N, Yoshioka C, et al. Cryo-EM structure of the benzodiazepine-sensitive α1β1γ2S tri-heteromeric GABAA receptor in complex with GABA. ELife. 2018; 7: e39383.
[28]
Zhu S, Noviello CM, Teng J, Walsh RM, Kim JJ, Hibbs RE. Structure of a human synaptic GABAA receptor. Nature. 2018; 559: 67-72.
[29]
Laverty D, Desai R, Uchański T, Masiulis S, Stec WJ, Malinauskas T, et al. Cryo-EM structure of the human α1β3γ2 GABAA receptor in a lipid bilayer. Nature. 2019; 565: 516-520.
[30]
Masiulis S, Desai R, Uchański T, Serna Martin I, Laverty D, Karia D, et al. GABAA receptor signalling mechanisms revealed by structural pharmacology. Nature. 2019; 565: 454-459.
[31]
Pirker S, Schwarzer C, Wieselthaler A, Sieghart W, Sperk G. GABAA receptors: immunocytochemical distribution of 13 subunits in the adult rat brain. Neuroscience. 2000; 101: 815-850.
[32]
Korpi ER, Gründer G, Lüddens H. Drug interactions at GABA(A) receptors. Progress in Neurobiology. 2002; 67: 113-159.
[33]
Sieghart W, Sperk G. Subunit composition, distribution and function of GABA(A) receptor subtypes. Current Topics in Medicinal Chemistry. 2002; 2: 795-816.
[34]
Sun C, Sieghart W, Kapur J. Distribution of α1, α4, γ2, and δ subunits of GABAA receptors in hippocampal granule cells. Brain Research. 2004; 1029: 207-216.
[35]
Mortensen M, Smart TG. Extrasynaptic alpha beta subunit GABAA receptors on rat hippocampal pyramidal neurons. Journal of Physiology. 2006; 577: 841-856.
[36]
Arslan A. Distinct roles of gamma-aminobutyric acid type a receptor subtypes: a focus on phasic and tonic inhibition. Journal of Neurobehavioral Sciences. 2015; 2: 72-76.
[37]
Waldvogel HJ, Munkle M, van Roon-Mom W, Mohler H, Faull RLM. The immunohistochemical distribution of the GABAA receptor α1, α2, α3, β2/3 and γ2 subunits in the human thalamus. Journal of Chemical Neuroanatomy. 2017; 82: 39-55.
[38]
Sperk G, Kirchmair E, Bakker J, Sieghart W, Drexel M, Kondova I. Immunohistochemical distribution of 10 GABA(A) receptor subunits in the forebrain of the rhesus monkey Macaca mulatta. Journal of Comparative Neurology. 2020; 528: 2551-2568.
[39]
Kasugai Y, Swinny JD, Roberts JDB, Dalezios Y, Fukazawa Y, Sieghart W, et al. Quantitative localisation of synaptic and extrasynaptic GABAA receptor subunits on hippocampal pyramidal cells by freeze-fracture replica immunolabelling. European Journal of Neuroscience. 2010; 32: 1868-1888.
[40]
Arslan A. Clustering of gamma-aminobutyric acid type A receptors. Periodicals of Engineering and Natural Sciences. 2015; 3: 28-32.
[41]
Sur C, Farrar SJ, Kerby J, Whiting PJ, Atack JR, McKernan RM. Preferential coassembly of α4 and δ subunits of the γ-aminobutyric acid A receptor in rat thalamus. Molecular Pharmacology. 1999; 56: 110-115.
[42]
Peng Z, Hauer B, Mihalek RM, Homanics GE, Sieghart W, Olsen RW, et al. GABA(A) receptor changes in delta subunit-deficient mice: altered expression of alpha4 and gamma2 subunits in the forebrain. Journal of Comparative Neurology. 2002; 446: 179-197.
[43]
Nusser Z, Sieghart W, Somogyi P. Segregation of different GABAA receptors to synaptic and extrasynaptic membranes of cerebellar granule cells. Journal of Neuroscience. 1998; 18: 1693-1703.
[44]
Wei W, Zhang N, Peng Z, Houser CR, Mody I. Perisynaptic localization of delta subunit-containing GABA(A) receptors and their activation by GABA spillover in the mouse dentate gyrus. Journal of Neuroscience. 2003; 23: 10650-10661.
[45]
Wisden W, Laurie D, Monyer H, Seeburg P. The distribution of 13 GABAA receptor subunit mRNAs in the rat brain. I. Telencephalon, diencephalon, mesencephalon. Journal of Neuroscience. 1992; 12: 1040-1062.
[46]
Jones A, Korpi ER, McKernan RM, Pelz R, Nusser Z, Mäkelä R, et al. Ligand-gated ion channel subunit partnerships: GABAA receptor α6 subunit gene inactivation inhibits δ subunit expression. Journal of Neuroscience. 1997; 17: 1350-1362.
[47]
Glykys J, Peng Z, Chandra D, Homanics GE, Houser CR, Mody I. A new naturally occurring GABAA receptor subunit partnership with high sensitivity to ethanol. Nature Neuroscience. 2007; 10: 40-48.
[48]
Salin PA, Prince DA. Spontaneous GABAA receptor-mediated inhibitory currents in adult rat somatosensory cortex. Journal of Neurophysiology. 1996; 75: 1573-1588.
[49]
Porcello DM, Huntsman MM, Mihalek RM, Homanics GE, Huguenard JR. Intact synaptic GABAergic inhibition and altered neurosteroid modulation of thalamic relay neurons in mice lacking delta subunit. Journal of Neurophysiology. 2003; 89: 1378-1386.
[50]
Ade KK, Janssen MJ, Ortinski PI, Vicini S. Differential tonic GABA conductances in striatal medium spiny neurons. Journal of Neuroscience. 2008; 28: 1185-1197.
[51]
Santhakumar V, Jones RT, Mody I. Developmental regulation and neuroprotective effects of striatal tonic GABAA currents. Neuroscience. 2010; 167: 644-655.
[52]
Drasbek KR, Jensen K. THIP, a hypnotic and antinociceptive drug, enhances an extrasynaptic GABAA receptor-mediated conductance in mouse neocortex. Cerebral Cortex. 2006; 16: 1134-1141.
[53]
Kirmse K, Dvorzhak A, Kirischuk S, Grantyn R. GABA transporter 1 tunes GABAergic synaptic transmission at output neurons of the mouse neostriatum. Journal of Physiology. 2008; 586: 5665-5678.
[54]
Prenosil GA, Schneider Gasser EM, Rudolph U, Keist R, Fritschy J, Vogt KE. Specific subtypes of GABAA receptors mediate phasic and tonic forms of inhibition in hippocampal pyramidal neurons. Journal of Neurophysiology. 2006; 96: 846-857.
[55]
Szabadics J, Tamás G, Soltesz I. Different transmitter transients underlie presynaptic cell type specificity of GABAA, slow and GABAA, fast. Proceedings of the National Academy of Sciences. 2007; 104: 14831-14836.
[56]
Sun MY, Ziolkowski L, Mennerick S. δ subunit-containing GABA(A) IPSCs are driven by both synaptic and diffusional GABA in mouse dentate granule neurons. Journal of Physiology. 2020; 598: 1205-1221.
[57]
Héja L, Nyitrai G, Kékesi O, Dobolyi A, Szabó P, Fiáth R, et al. Astrocytes convert network excitation to tonic inhibition of neurons. BMC Biology. 2012; 10: 26.
[58]
Wójtowicz AM, Dvorzhak A, Semtner M, Grantyn R. Reduced tonic inhibition in striatal output neurons from Huntington mice due to loss of astrocytic GABA release through GAT-3. Frontiers in Neural Circuits. 2013; 7: 188.
[59]
Lagrange AH, Hu N, Macdonald RL. GABA beyond the synapse: defining the subtype-specific pharmacodynamics of non-synaptic GABAA receptors. Journal of Physiology. 2018; 596: 4475-4495.
[60]
McCartney MR, Deeb TZ, Henderson TN, Hales TG. Tonically active GABAA receptors in hippocampal pyramidal neurons exhibit constitutive GABA-independent gating. Molecular Pharmacology. 2007; 71: 539-548.
[61]
Wlodarczyk AI, Sylantyev S, Herd MB, Kersante F, Lambert JJ, Rusakov DA, et al. GABA-independent GABAA receptor openings maintain tonic currents. Journal of Neuroscience. 2013; 33: 3905-3914.
[62]
Dalby NO, Falk-Petersen CB, Leurs U, Scholze P, Krall J, Frølund B, et al. Silencing of spontaneous activity at α4β1/3δ GABA(A) receptors in hippocampal granule cells reveals different ligand pharmacology. British Journal of Pharmacology. 2020; 177: 3975-3990.
[63]
Glykys J, Mody I. Activation of GABAA receptors: views from outside the synaptic cleft. Neuron. 2007; 56: 763-770.
[64]
Rovó Z, Mátyás F, Barthó P, Slézia A, Lecci S, Pellegrini C, et al. Phasic, nonsynaptic GABA-A receptor-mediated inhibition entrains thalamocortical oscillations. Journal of Neuroscience. 2014; 34: 7137-7147.
[65]
Hörtnagl H, Tasan RO, Wieselthaler A, Kirchmair E, Sieghart W, Sperk G. Patterns of mRNA and protein expression for 12 GABAA receptor subunits in the mouse brain. Neuroscience. 2013; 236: 345-372.
[66]
Vithlani M, Terunuma M, Moss SJ. The dynamic modulation of GABA(A) receptor trafficking and its role in regulating the plasticity of inhibitory synapses. Physiological Reviews. 2011; 91: 1009-1022.
[67]
Nusser Z, Hájos N, Somogyi P, Mody I. Increased number of synaptic GABAA receptors underlies potentiation at hippocampal inhibitory synapses. Nature. 1998; 395: 172-177.
[68]
Ye Z, McGee TP, Houston CM, Brickley SG. The contribution of δ subunit-containing GABAA receptors to phasic and tonic conductance changes in cerebellum, thalamus and neocortex. Frontiers in Neural Circuits. 2013; 7: 203.
[69]
Sun M, Shu H, Benz A, Bracamontes J, Akk G, Zorumski CF, et al. Chemogenetic isolation reveals synaptic contribution of δ GABAA receptors in mouse dentate granule neurons. Journal of Neuroscience. 2018; 38: 8128-8145.
[70]
Sundstrom-Poromaa I, Smith DH, Gong QH, Sabado TN, Li X, Light A, et al. Hormonally regulated α4β2δ GABA A receptors are a target for alcohol. Nature Neuroscience. 2002; 5: 721-722.
[71]
Dembitskaya Y, Wu Y, Semyanov A. Tonic GABAA conductance favors spike-timing-dependent over theta-burst-induced long-term potentiation in the hippocampus. Journal of Neuroscience. 2020; 40: 4266-4276.
[72]
Farrant M, Nusser Z. Variations on an inhibitory theme: phasic and tonic activation of GABAA receptors. Nature Reviews Neuroscience. 2005; 6: 215-229.
[73]
Semyanov A, Walker MC, Kullmann DM. GABA uptake regulates cortical excitability via cell type-specific tonic inhibition. Nature Neuroscience. 2003; 6: 484-490.
[74]
Bryson A, Hatch RJ, Zandt B-J, Rossert C, Berkovic SF, Reid CA, et al. GABA-mediated tonic inhibition differentially modulates gain in functional subtypes of cortical interneurons. Proceedings of the National Academy of Sciences. 2020; 117: 3192–3202.
[75]
Mann EO, Mody I. Control of hippocampal gamma oscillation frequency by tonic inhibition and excitation of interneurons. Nature Neuroscience. 2010; 13: 205-212.
[76]
Herd MB, Brown AR, Lambert JJ, Belelli D. Extrasynaptic GABA(A) receptors couple presynaptic activity to postsynaptic inhibition in the somatosensory thalamus. Journal of Neuroscience. 2013; 33: 14850-14868.
[77]
Shen H, Kenney L, Smith SS. Increased dendritic branching of and reduced δ-GABA(A) receptor expression on parvalbumin-positive interneurons increase inhibitory currents and reduce synaptic plasticity at puberty in female mouse CA1 hippocampus. Frontiers in Cellular Neuroscience. 2020; 14: 203.
[78]
Parato J, Shen H, Smith SS. α4βδ GABAA receptors trigger synaptic pruning and reduce dendritic length of female mouse CA3 hippocampal pyramidal cells at puberty. Neuroscience. 2019; 398: 23-36.
[79]
Duveau V, Laustela S, Barth L, Gianolini F, Vogt KE, Keist R, et al. Spatiotemporal specificity of GABAA receptor-mediated regulation of adult hippocampal neurogenesis. European journal of neuroscience. 2011; 34: 362–373.
[80]
Holter NI, Zylla MM, Zuber N, Bruehl C, Draguhn A. Tonic GABAergic control of mouse dentate granule cells during postnatal development. European journal of neuroscience. 2010; 32: 1300–1309.
[81]
Martin LJ, Zurek AA, MacDonald JF, Roder JC, Jackson MF, Orser BA. α5GABAA receptor activity sets the threshold for long-term potentiation and constrains hippocampus-dependent memory. Journal of Neuroscience. 2010; 30: 5269–5282.
[82]
Whissell PD, Rosenzweig S, Lecker I, Wang D, Wojtowicz JM, Orser BA. γ-aminobutyric acid type a receptors that contain the δ subunit promote memory and neurogenesis in the dentate gyrus. Annals of Neurology. 2013; 74: 611-621.
[83]
Herd MB, Foister N, Chandra D, Peden DR, Homanics GE, Brown VJ, et al. Inhibition of thalamic excitability by 4,5,6,7-tetrahydroisoxazolo[4,5-c]pyridine-3-ol: a selective role for δ-GABA(A) receptors. European Journal of Neuroscience. 2009; 29: 1177-1187.
[84]
Young R, Urbancic A, Emrey TA, Hall PC, Metcalf G. Behavioral effects of several new anxiolytics and putative anxiolytics. European journal of pharmacology. 1987; 143: 361–371.
[85]
Thompson SA, Wingrove PB, Connelly L, Whiting PJ, Wafford KA. Tracazolate reveals a novel type of allosteric interaction with recombinant γ-aminobutyric acid(A) receptors. Molecular pharmacology. 2002; 61: 861–869.
[86]
Rudolph S, Guo C, Pashkovski SL, Osorno T, Gillis WF, Krauss JM, et al. Cerebellum-specific deletion of the GABAA receptor δ subunit leads to sex-specific disruption of behavior. Cell Reports. 2020; 33: 108338.
[87]
Govindpani K, Turner C, Waldvogel HJ, Faull RLM, Kwakowsky A. Impaired expression of GABA signaling components in the Alzheimer’s disease middle temporal gyrus. International Journal of Molecular Sciences. 2020; 21: 8704.
[88]
Lee V, MacKenzie G, Hooper A, Maguire J. Reduced tonic inhibition in the dentate gyrus contributes to chronic stress-induced impairments in learning and memory. Hippocampus. 2016; 26: 1276-1290.
[89]
Zhang N, Peng Z, Tong X, Lindemeyer AK, Cetina Y, Huang CS, et al. Decreased surface expression of the δ subunit of the GABAA receptor contributes to reduced tonic inhibition in dentate granule cells in a mouse model of fragile X syndrome. Experimental Neurology. 2017; 297: 168-178.
[90]
Maldonado-Avilés JG, Curley AA, Hashimoto T, Morrow AL, Ramsey AJ, O’Donnell P, et al. Altered markers of tonic inhibition in the dorsolateral prefrontal cortex of subjects with schizophrenia. American Journal of Psychiatry. 2009; 166: 450-459.
[91]
Peng Z, Huang CS, Stell BM, Mody I, Houser CR. Altered expression of the δ subunit of the GABAA receptor in a mouse model of temporal lobe epilepsy. Journal of Neuroscience. 2004; 24: 8629-8639.
[92]
Nin MS, Ferri MK, Couto-Pereira NS, Souza MF, Azeredo LA, Agnes G, et al. The effect of intra-nucleus accumbens administration of allopregnanolone on δ and γ2 GABAA receptor subunit mRNA expression in the hippocampus and on depressive-like and grooming behaviors in rats. Pharmacology Biochemistry and Behavior. 2012; 103: 359-366.
[93]
Almeida FB, Gomez R, Barros HMT, Nin MS. Hemisphere-dependent changes in mRNA expression of GABAA receptor subunits and BDNF after intra-prefrontal cortex allopregnanolone infusion in rats. Neuroscience. 2019; 397: 56-66.
[94]
Maguire J. Neuroactive steroids and GABAergic involvement in the neuroendocrine dysfunction associated with major depressive disorder and postpartum depression. Frontiers in Cellular Neuroscience. 2019; 13: 83.
[95]
Feng Y, Kapornai K, Kiss E, Tamás Z, Mayer L, Baji I, et al. Association of the GABRD gene and childhood-onset mood disorders. Genes, Brain, and Behavior. 2010; 9: 668-672.
[96]
Shen H, Mohammad A, Ramroop J, Smith SS. A stress steroid triggers anxiety via increased expression of α4βδ GABAA receptors in methamphetamine dependence. Neuroscience. 2013; 254: 452-475.
[97]
Merali Z, Du L, Hrdina P, Palkovits M, Faludi G, Poulter MO, et al. Dysregulation in the suicide brain: mRNA expression of corticotropin-releasing hormone receptors and GABAA receptor subunits in frontal cortical brain region. Journal of Neuroscience. 2004; 24: 1478-1485.
[98]
Ferando I, Mody I. Altered gamma oscillations during pregnancy through loss of δ subunit-containing GABA(A) receptors on parvalbumin interneurons. Frontiers in Neural Circuits. 2013; 7: 144.
[99]
Siivonen MS, de Miguel E, Aaltio J, Manner AK, Vahermo M, Yli-Kauhaluoma J, et al. Conditioned reward of opioids, but not psychostimulants, is impaired in GABA-A receptor δ subunit knockout mice. Basic & Clinical Pharmacology & Toxicology. 2018; 123: 558-566.
[100]
Maguire JL, Stell BM, Rafizadeh M, Mody I. Ovarian cycle-linked changes in GABAA receptors mediating tonic inhibition alter seizure susceptibility and anxiety. Nature Neuroscience. 2005; 8: 797-804.
[101]
Barth AMI, Ferando I, Mody I. Ovarian cycle-linked plasticity of δ-GABAA receptor subunits in hippocampal interneurons affects γ oscillations in vivo. Frontiers in Cellular Neuroscience. 2014; 8: 222.
[102]
Neumann S, Boothman-Burrell L, Gowing EK, Jacobsen TA, Ahring PK, Young SL, et al. The delta-subunit selective GABA A receptor modulator, DS2, improves stroke recovery via an anti-inflammatory mechanism. Frontiers in Neuroscience. 2019; 13: 1133.
[103]
Smith SS, Ruderman Y, Frye C, Homanics G, Yuan M. Steroid withdrawal in the mouse results in anxiogenic effects of 3α,5β-THP: a possible model of premenstrual dysphoric disorder. Psychopharmacology. 2006; 186: 323-333.
[104]
Raible DJ, Frey LC, Cruz Del Angel Y, Russek SJ, Brooks-Kayal AR. GABAA receptor regulation after experimental traumatic brain injury. Journal of Neurotrauma. 2012; 29: 2548-2554.
[105]
Drexel M, Puhakka N, Kirchmair E, Hörtnagl H, Pitkänen A, Sperk G. Expression of GABA receptor subunits in the hippocampus and thalamus after experimental traumatic brain injury. Neuropharmacology. 2015; 88: 122-133.
[106]
Rosas-Arellano A, Estrada-Mondragón A, Mantellero C, Tejeda-Guzmán C, Castro M. The adjustment of γ-aminobutyric acid A tonic subunits in Huntington’s disease: from transcription to translation to synaptic levels into the neostriatum. Neural Regeneration Research. 2018; 13: 584-590.
[107]
Bonin RP, Labrakakis C, Eng DG, Whissell PD, De Koninck Y, Orser BA. Pharmacological enhancement of δ-subunit-containing GABAA receptors that generate a tonic inhibitory conductance in spinal neurons attenuates acute nociception in mice. Pain. 2011; 152: 1317-1326.
[108]
Walsh JK, Mayleben D, Guico-Pabia C, Vandormael K, Martinez R, Deacon S. Efficacy of the selective extrasynaptic GABAA agonist, gaboxadol, in a model of transient insomnia: a randomized, controlled clinical trial. Sleep Medicine. 2008; 9: 393-402.
[109]
Winsky-Sommerer R, Vyazovskiy VV, Homanics GE, Tobler I. The EEG effects of THIP (Gaboxadol) on sleep and waking are mediated by the GABAA δ-subunit-containing receptors. European Journal of Neuroscience. 2007; 25: 1893-1899.
[110]
Lancel M, Langebartels A. γ-aminobutyric Acid(A) (GABAA) agonist 4,5,6,7-tetrahydroisoxazolo[4,5-c]pyridin-3-ol persistently increases sleep maintenance and intensity during chronic administration to rats. Journal of Pharmacology and Experimental Therapeutics. 2000; 293: 1084-1090.
[111]
Liu C, Marioni RE, Hedman Å K, Pfeiffer L, Tsai PC, Reynolds LM, et al. A DNA methylation biomarker of alcohol consumption. Molecular Psychiatry. 2018; 23: 422-433.
[112]
Melón LC, Nasman JT, John AS, Mbonu K, Maguire JL. Interneuronal δ-GABAA receptors regulate binge drinking and are necessary for the behavioral effects of early withdrawal. Neuropsychopharmacology. 2019; 44: 425-434.
[113]
Arslan A. Alcohol modulation of extra-synaptic gamma-aminobutyric acid type A receptors. Periodicals of Engineering and Natural Sciences. 2016; 4: 29-36.
[114]
Darnieder LM, Melón LC, Do T, Walton NL, Miczek KA, Maguire JL. Female-specific decreases in alcohol binge-like drinking resulting from GABAA receptor delta-subunit knockdown in the VTA. Scientific Reports. 2019; 9: 8102.
[115]
Centanni SW, Burnett EJ, Trantham-Davidson H, Chandler LJ. Loss of δ-GABAA receptor-mediated tonic currents in the adult prelimbic cortex following adolescent alcohol exposure. Addiction Biology. 2017; 22: 616-628.
[116]
Gonzalez C, Moss SJ, Olsen RW. Ethanol promotes clathrin adaptor-mediated endocytosis via the intracellular domain of δ-containing GABAA receptors. Journal of Neuroscience. 2012; 32: 17874-17881.
[117]
Chen J, He Y, Wu Y, Zhou H, Su L-D, Li W-N, et al. Single Ethanol withdrawal regulates extrasynaptic δ-GABAA receptors via PKCδ activation. Frontiers in Molecular Neuroscience. 2018; 11: 141-141.
[118]
Melón LC, Nolan ZT, Colar D, Moore EM, Boehm SL. Activation of extrasynaptic δ-GABAA receptors globally or within the posterior-VTA has estrous-dependent effects on consumption of alcohol and estrous-independent effects on locomotion. Hormones and Behavior. 2017; 95: 65-75.
[119]
Kang JQ, Shen W, Macdonald RL. The GABRG2 mutation, Q351X, associated with generalized epilepsy with febrile seizures plus, has both loss of function and dominant-negative suppression. Journal of Neuroscience. 2009; 29: 2845-2856.
[120]
Wallace RH, Marini C, Petrou S, Harkin LA, Bowser DN, Panchal RG, et al. Mutant GABAA receptor γ2-subunit in childhood absence epilepsy and febrile seizures. Nature Genetics. 2001; 28: 49-52.
[121]
Harkin LA, Bowser DN, Dibbens LM, Singh R, Phillips F, Wallace RH, et al. Truncation of the GABAA-receptor γ2 subunit in a family with generalized epilepsy with febrile seizures plus. American Journal of Human Genetics. 2002; 70: 530-536.
[122]
Feng HJ, Kang JQ, Song L, Dibbens L, Mulley J, Macdonald RL. δ subunit susceptibility variants E177A and R220H associated with complex epilepsy alter channel gating and surface expression of α4β2δ GABAA receptors. Journal of Neuroscience. 2006; 26: 1499-1506.
[123]
Dibbens LM, Feng H, Richards MC, Harkin LA, Hodgson BL, Scott D, et al. GABRD encoding a protein for extra- or peri-synaptic GABAA receptors is a susceptibility locus for generalized epilepsies. Human Molecular Genetics. 2004; 13: 1315-1319.
[124]
Biagini G, Panuccio G, Avoli M. Neurosteroids and epilepsy. Current Opinion in Neurology. 2010; 23: 170-176.
[125]
Meletti S, Lucchi C, Monti G, Giovannini G, Bedin R, Trenti T, et al. Decreased allopregnanolone levels in cerebrospinal fluid obtained during status epilepticus. Epilepsia. 2017; 58: e16-e20.
[126]
Meletti S, Lucchi C, Monti G, Giovannini G, Bedin R, Trenti T, et al. Low levels of progesterone and derivatives in cerebrospinal fluid of patients affected by status epilepticus. Journal of Neurochemistry. 2018; 147: 275-284.
[127]
Lévesque M, Biagini G, Avoli M. Neurosteroids and focal epileptic disorders. International Journal of Molecular Sciences. 2020; 21: 9391.
[128]
Stell BM, Brickley SG, Tang CY, Farrant M, Mody I. Neuroactive steroids reduce neuronal excitability by selectively enhancing tonic inhibition mediated by δ subunit-containing GABAA receptors. Proceedings of the National Academy of Sciences. 2003; 100: 14439-14444.
[129]
Spigelman I, Li Z, Liang J, Cagetti E, Samzadeh S, Mihalek RM, et al. Reduced inhibition and sensitivity to neurosteroids in hippocampus of mice lacking the GABAA receptor δ subunit. Journal of Neurophysiology. 2003; 90: 903-910.
[130]
Spigelman I, Li Z, Banerjee PK, Mihalek RM, Homanics GE, Olsen RW. Behavior and physiology of mice lacking the GABAA-receptor δ subunit. Epilepsia. 2002; 43: 3-8.
[131]
Stewart GO, Dobb GJ, Craib IA. Clinical trial of continuous infusion of alphaxalone/alphadolone in intensive care patients. Anaesthesia and Intensive Care. 1983; 11: 107-112.
[132]
Chiara DC, Jayakar SS, Zhou X, Zhang X, Savechenkov PY, Bruzik KS, et al. Specificity of intersubunit general anesthetic-binding sites in the transmembrane domain of the human α1β3γ2 γ-aminobutyric acid type A (GABAA) receptor. Journal of Biological Chemistry. 2013; 288: 19343-19357.
[133]
Vashchinkina E, Manner AK, Vekovischeva O, Hollander BD, Uusi-Oukari M, Aitta-aho T, et al. Neurosteroid agonist at GABAA receptor induces persistent neuroplasticity in VTA dopamine neurons. Neuropsychopharmacology. 2014; 39: 727-737.
[134]
Monagle J, Siu L, Worrell J, Goodchild CS, Serrao JM. A phase 1C trial comparing the efficacy and safety of a new aqueous formulation of alphaxalone with propofol. Anesthesia & Analgesia. 2015; 121: 914-924.
[135]
Trujeque-Ramos S, Castillo-Rolón D, Galarraga E, Tapia D, Arenas-López G, Mihailescu S, et al. Insulin regulates GABAA receptor-mediated tonic currents in the prefrontal cortex. Frontiers in Neuroscience. 2018; 12: 345.
[136]
Maniezzi C, Talpo F, Spaiardi P, Toselli M, Biella G. Oxytocin increases phasic and tonic GABAergic transmission in CA1 region of mouse hippocampus. Frontiers in Cellular Neuroscience. 2019; 13: 178.
[137]
Kanes S, Colquhoun H, Gunduz-Bruce H, Raines S, Arnold R, Schacterle A, et al. Brexanolone (SAGE-547 injection) in post-partum depression: a randomised controlled trial. The Lancet. 2017; 390: 480-489.
[138]
Kanes SJ, Colquhoun H, Doherty J, Raines S, Hoffmann E, Rubinow DR, et al. Open-label, proof-of-concept study of brexanolone in the treatment of severe postpartum depression. Human Psychopharmacology. 2017; 32: e2576.
[139]
Meltzer-Brody S, Colquhoun H, Riesenberg R, Epperson CN, Deligiannidis KM, Rubinow DR, et al. Brexanolone injection in post-partum depression: two multicentre, double-blind, randomised, placebo-controlled, phase 3 trials. The Lancet. 2018; 392: 1058-1070.
[140]
Pinna G. Allopregnanolone, the neuromodulator turned therapeutic agent: thank you, next? Frontiers in Endocrinology. 2020; 11: 236.
[141]
Gunduz-Bruce H, Silber C, Kaul I, Rothschild AJ, Riesenberg R, Sankoh AJ, et al. Trial of SAGE-217 in patients with major depressive disorder. New England Journal of Medicine. 2019; 381: 903-911.
[142]
Shu H-J, Bracamontes J, Taylor A, Wu K, Eaton MM, Akk G, et al. Characteristics of concatemeric GABAA receptors containing α4/δ subunits expressed in Xenopus oocytes. British Journal of Pharmacology. 2012; 165: 2228-2243.
[143]
Lu X, Zorumski CF, Mennerick S. Lack of neurosteroid selectivity at δ vs. γ2-containing GABAA receptors in dentate granule neurons. Frontiers in Molecular Neuroscience. 2020; 13: 6.
[144]
Pritchett DB, Seeburg PH. γ-aminobutyric acid A receptor α 5-subunit creates novel type II benzodiazepine receptor pharmacology. Journal of Neurochemistry. 1990; 54: 1802-1804.
[145]
Hadingham KL, Wingrove P, Le Bourdelles B, Palmer KJ, Ragan CI, Whiting PJ. Cloning of cDNA sequences encoding human alpha 2 and alpha 3 gamma-aminobutyric acid A receptor subunits and characterization of the benzodiazepine pharmacology of recombinant alpha 1-, alpha 2-, alpha 3-, and alpha 5-containing human gamma-aminobutyric acid A receptors. Molecular Pharmacology. 1993; 43: 970-975.
[146]
Hadingham KL, Garrett EM, Wafford KA, Bain C, Heavens RP, Sirinathsinghji DJ, et al. Cloning of cDNAs encoding the human gamma-aminobutyric acid type a receptor alpha 6 subunit and characterization of the pharmacology of alpha 6-containing receptors. Molecular Pharmacology. 1996; 49: 253-259.
[147]
Ballard TM, Knoflach F, Prinssen E, Borroni E, Vivian JA, Basile J, et al. RO4938581, a novel cognitive enhancer acting at GABAA alpha5 subunit-containing receptors. Psychopharmacology. 2009; 202: 207-223.
[148]
Etherington L, Mihalik B, Pálvölgyi A, Ling I, Pallagi K, Kertész S, et al. Selective inhibition of extra-synaptic α5-GABAA receptors by S44819, a new therapeutic agent. Neuropharmacology. 2017; 125: 353-364.
[149]
Atack JR, Bayley PJ, Seabrook GR, Wafford KA, McKernan RM, Dawson GR. L-655,708 enhances cognition in rats but is not proconvulsant at a dose selective for α5-containing GABAA receptors. Neuropharmacology. 2006; 51: 1023-1029.
[150]
Dawson GR, Maubach KA, Collinson N, Cobain M, Everitt BJ, MacLeod AM, et al. An inverse agonist selective for α5 subunit-containing GABAA receptors enhances cognition. Journal of Pharmacology and Experimental Therapeutics. 2006; 316: 1335-1345.
[151]
Martinez-Cue C, Martinez P, Rueda N, Vidal R, Garcia S, Vidal V, et al. Reducing GABAA 5 receptor-mediated inhibition rescues functional and neuromorphological deficits in a mouse model of down syndrome. Journal of Neuroscience. 2013; 33: 3953-3966.
[152]
Ferando I, Mody I. Altered gamma oscillations during pregnancy through loss of δ subunit-containing GABAA receptors on parvalbumin interneurons. Frontiers in Neural Circuits. 2013; 7: 144.
[153]
Arslan A, von Engelhardt J, Wisden W. Cytoplasmic domain of δ subunit is important for the extra-synaptic targeting of GABAA receptor subtypes. Journal of Integrative Neuroscience. 2014; 13: 617-631.
[154]
Botzolakis EJ, Gurba KN, Lagrange AH, Feng H, Stanic AK, Hu N, et al. Comparison of γ-aminobutyric acid, type A (GABAA), receptor αβγ and αβδ expression using flow cytometry and electrophysiology: evidence for alternative stoichiometries and arrangements. Journal of Biological Chemistry. 2016; 291: 20440-20461.
[155]
Feng H, Forman SA. Comparison of αβδ and αβγ GABAA receptors: Allosteric modulation and identification of subunit arrangement by site-selective general anesthetics. Pharmacological Research. 2018; 133: 289-300.
[156]
Martenson JS, Yamasaki T, Chaudhury NH, Albrecht D, Tomita S. Assembly rules for GABAA receptor complexes in the brain. ELife. 2017; 6: e27443.
[157]
Barrera NP, Betts J, You H, Henderson RM, Martin IL, Dunn SMJ, et al. Atomic force microscopy reveals the stoichiometry and subunit arrangement of the α4β3δ GABAA receptor. Molecular Pharmacology. 2008; 73: 960-967.
[158]
Patel B, Mortensen M, Smart TG. Stoichiometry of δ subunit containing GABAA receptors. British Journal of Pharmacology. 2014; 171: 985-994.
[159]
Oflaz FE, Son Ç D, Arslan A. Oligomerization and cell surface expression of recombinant GABAA receptors tagged in the δ subunit. Journal of Integrative Neuroscience. 2019; 18: 341-350.
[160]
Tretter V, Ehya N, Fuchs K, Sieghart W. Stoichiometry and assembly of a recombinant GABAA receptor subtype. Journal of Neuroscience. 1997; 17: 2728-2737.
[161]
Chang Y, Wang R, Barot S, Weiss DS. Stoichiometry of a recombinant GABAA Receptor. Journal of Neuroscience. 1996; 16: 5415-5424.
[162]
Chua HC, Absalom NL, Hanrahan JR, Viswas R, Chebib M. The direct actions of GABA, 2’-methoxy-6-methylflavone and general anaesthetics at β3γ2L GABAA receptors: evidence for receptors with different subunit stoichiometries. PLoS ONE. 2015; 10: e0141359.
[163]
Lee HJ, Absalom NL, Hanrahan JR, van Nieuwenhuijzen P, Ahring PK, Chebib M. A pharmacological characterization of GABA, THIP and DS2 at binary α4β3 and β3δ receptors: GABA activates β3δ receptors via the β3(+)δ(−) interface. Brain Research. 2016; 1644: 222-230.
[164]
Wagoner KR, Czajkowski C. Stoichiometry of expressed α4β2δγ-aminobutyric acid type A receptors depends on the ratio of subunit cDNA transfected. Journal of Biological Chemistry. 2010; 285: 14187-14194.
[165]
Waldvogel HJ, Baer K, Gai W-, Gilbert RT, Rees MI, Mohler H, et al. Differential localization of GABAA receptor subunits within the substantia nigra of the human brain: an immunohistochemical study. Journal of Comparative Neurology. 2008; 506: 912-929.
[166]
Wohlfarth KM, Bianchi MT, Macdonald RL. Enhanced neurosteroid potentiation of ternary GABAA receptors containing the δ subunit. Journal of Neuroscience. 2002; 22: 1541-1549.
[167]
Belelli D, Lambert JJ. Neurosteroids: endogenous regulators of the GABAA receptor. Nature Reviews Neuroscience. 2005; 6: 565-575.
[168]
Laverty D, Thomas P, Field M, Andersen OJ, Gold MG, Biggin PC, et al. Crystal structures of a GABAA-receptor chimera reveal new endogenous neurosteroid-binding sites. Nature Structural & Molecular Biology. 2017; 24: 977-985.
[169]
Wallner M, Hanchar HJ, Olsen RW. Ethanol enhances α4β3δ and α6β3δγ-aminobutyric acid type A receptors at low concentrations known to affect humans. Proceedings of the National Academy of Sciences. 2003; 100: 15218-15223.
[170]
Kumar S, Porcu P, Werner DF, Matthews DB, Diaz-Granados JL, Helfand RS, et al. The role of GABAA receptors in the acute and chronic effects of ethanol: a decade of progress. Psychopharmacology. 2009; 205: 529-564.
[171]
Borghese CM, Stórustovu SÍ, Ebert B, Herd MB, Belelli D, Lambert JJ, et al. The δ subunit of γ-aminobutyric acid type a receptors does not confer sensitivity to low concentrations of ethanol. Journal of Pharmacology and Experimental Therapeutics. 2006; 316: 1360-1368.
[172]
Yamashita M, Marszalec W, Yeh JZ, Narahashi T. Effects of ethanol on tonic GABA currents in cerebellar granule cells and mammalian cells recombinantly expressing GABAA receptors. Journal of Pharmacology and Experimental Therapeutics. 2006; 319: 431-438.
Share
Back to top