Skip to main content
The Journal of Neuroscience logoLink to The Journal of Neuroscience
. 2009 Oct 14;29(41):12757–12763. doi: 10.1523/JNEUROSCI.3340-09.2009

Extrasynaptic GABAA Receptors: Form, Pharmacology, and Function

Delia Belelli 1, Neil L Harrison 2, Jamie Maguire 3, Robert L Macdonald 4,5,6, Matthew C Walker 7, David W Cope 8,
PMCID: PMC2784229  NIHMSID: NIHMS158061  EMSID: EMS28079  PMID: 19828786

Abstract

GABA is the principal inhibitory neurotransmitter in the CNS and acts via GABAA and GABAB receptors. Recently, a novel form of GABAA receptor-mediated inhibition, termed “tonic” inhibition, has been described. Whereas synaptic GABAA receptors underlie classical “phasic” GABAA receptor-mediated inhibition (inhibitory postsynaptic currents), tonic GABAA receptor-mediated inhibition results from the activation of extrasynaptic receptors by low concentrations of ambient GABA. Extrasynaptic GABAA receptors are composed of receptor subunits that convey biophysical properties ideally suited to the generation of persistent inhibition and are pharmacologically and functionally distinct from their synaptic counterparts. This mini-symposium review highlights ongoing work examining the properties of recombinant and native extrasynaptic GABAA receptors and their preferential targeting by endogenous and clinically relevant agents. In addition, it emphasizes the important role of extrasynaptic GABAA receptors in GABAergic inhibition throughout the CNS and identifies them as a major player in both physiological and pathophysiological processes.

Introduction

It is only recently that two seemingly unrelated phenomena, the existence of a GABAA receptor (GABAAR)-dependent “tone” in some neurons (Otis et al., 1991; Salin and Prince, 1996) and the presence of GABAARs outside synaptic specializations (Somogyi et al., 1989; Soltesz et al., 1990), have been unified: GABA spillover from the synaptic cleft activates extrasynaptic or perisynaptic GABAARs to generate a persistent or tonic inhibition (for review, see Semyanov et al., 2004; Farrant and Nusser, 2005; Glykys and Mody, 2007). Tonic inhibition is distinct from the transient activation of synaptic GABAARs leading to classical inhibitory postsynaptic currents (phasic inhibition) and the slow, but still transient, response of the metabotropic GABABRs. The initial finding in cerebellar granule cells (Brickley et al., 1996; Wall and Usowicz, 1997; Nusser et al., 1998; Brickley et al., 2001; Hamann et al., 2002) was followed by subsequent discoveries in, among others, the dentate gyrus and hippocampus (Bai et al., 2001; Nusser and Mody, 2002; Semyanov et al., 2003; Wei et al., 2003; Caraiscos et al., 2004a,b; Scimemi et al., 2005; Glykys et al., 2007), neocortex (Drasbek and Jensen, 2006; Yamada et al., 2007; Krook-Magnuson et al., 2008), thalamus (Belelli et al., 2005; Cope et al., 2005; Jia et al., 2005), striatum (Ade et al., 2008; Janssen et al., 2009), hypothalamus (Park et al., 2006, 2007), and spinal cord (Takahashi et al., 2006; Wang et al., 2008), and also in humans (Scimemi et al., 2006). The occurrence of tonic GABAA inhibition coincides with the expression of relatively rare receptor subunits, particularly the α4, α6, and δ subunits, and as a general rule-of-thumb, δ subunit-containing receptors are extrasynaptic, but not all extrasynaptic GABAARs contain δ subunits. In comparison, the ubiquitous γ2 subunit is a major component of synaptic GABAARs and drives receptor clustering at the synapse (Essrich et al., 1998). The presence of the δ subunit in recombinant receptors conveys properties ideally suited to generating tonic inhibition, namely activation by low concentrations of GABA, such as may be found in the extracellular space and reduced desensitization (Saxena and Macdonald, 1994; Haas and Macdonald, 1999; Bianchi and Macdonald, 2002; Brown et al., 2002). The δ subunit can also govern receptor pharmacology, extrasynaptic GABAARs typically being insensitive to benzodiazepine agonists (Nusser et al., 2002; Cope et al., 2005) but highly sensitive to the GABAAR “super agonist” 4,5,6,7-tetrahydroisoxazolo[4,5-c]pyridine-3-ol (THIP/gaboxadol) (Brown et al., 2002; Wohlfarth and Macdonald, 2002).

More recently, studies have begun to identify extrasynaptic GABAARs as novel targets for a diverse array of endogenous and clinically relevant agents, including certain neuroactive steroids (Belelli et al., 2002; Wohlfarth et al., 2002; Stell et al., 2003; Cope et al., 2005) and the amino acid taurine (Jia et al., 2008a), as well as ethanol (Sundstrom-Poromaa et al., 2002; Wallner et al., 2003; Glykys et al., 2007; Jia et al., 2008b), several anesthetic and hypnotic agents (Bai et al., 2001; Caraiscos et al., 2004b; Belelli et al., 2005; Cheng et al., 2006a; Takahashi et al., 2006; Jia et al., 2008c), analgesics (Krogsgaard-Larsen et al., 2004), and some anticonvulsant drugs (Cheng et al., 2006b). What is more, the functional role of tonic inhibition is beginning to be elucidated, such as the dynamic regulation of neuronal output, firing mode, and gain-control of neurotransmission (Hamann et al., 2002; Mitchell and Silver, 2003; Semyanov et al., 2003; Chadderton et al., 2004; Cope et al., 2005; Park et al., 2006; Bright et al., 2007; Rothman et al., 2009). Last, aberrant tonic inhibition has been implicated in multiple pathophysiological conditions, including fragile X mental retardation (Curia et al., 2009), γ-hydroxybutyric acid (GHB)-uria (Drasbek et al., 2008), stress (Maguire and Mody, 2007), disorders associated with the menstrual cycle and puberty (Maguire et al., 2005; Shen et al., 2007), and idiopathic generalized and temporal lobe epilepsies (Dibbens et al., 2004; Naylor et al., 2005; Scimemi et al., 2005; Feng et al., 2006; Zhang et al., 2007). Thus, extrasynaptic GABAARs may be candidates for therapeutic treatment in a range of neurological disorders.

The mini-symposium described below, therefore, provided an overview of previous work on extrasynaptic GABAARs and tonic inhibition and documented ongoing studies in this exciting field of research. Findings ranging from molecular studies to behavioral experiments were discussed. Individual presentations focused on the biophysical properties and structure-function relationships of putative extrasynaptic GABAARs (Biophysical properties of recombinant δ subunit-containing GABAAR channels), their selective targeting by endogenous and clinically important agents (Neuron-selective actions of neurosteroids at synaptic and extrasynaptic GABAARs and Thalamic extrasynaptic GABAARs are a target for ethanol and volatile anesthetics), and the contribution of tonic GABAA inhibition to both physiological and pathophysiological processes (Steroid hormones regulate extrasynaptic GABAARs during the ovarian cycle, pregnancy, and associated disorders, Enhanced tonic GABAA inhibition in thalamic neurons is necessary and sufficient for absence seizures, and The contribution of tonic GABAA inhibition to physiological and pathological hippocampal excitability).

Biophysical properties of recombinant δ subunit-containing GABAAR channels

Molecular biological and biochemical techniques have been instrumental in determining the basic properties of ligand-gated ion channels, and by inference, the functional properties of native receptors in the CNS. Recombinant receptors have, therefore, been used to distinguish the stoichiometry, electrophysiological, and pharmacological properties of GABAARs. Macdonald and colleagues have examined the properties of recombinant αβδ and αβγ receptors, i.e., putative extrasynaptic and synaptic receptors, respectively, and the possible roles of δ subunit variants in the pathophysiology of idiopathic generalized epilepsies.

αβγ receptors likely have a stoichiometry of 2α:2β:1γ (Chang et al., 1996; Tretter et al., 1997), and while it is generally believed that the δ subunit can substitute for the γ subunit, the stoichiometry of αβδ receptors remains uncertain (Kaur et al., 2009). The functional properties between the two receptor subtypes differ. Compared with α1βγ2 receptors, α1βδ, α4βδ, and α6βδ receptors have smaller macroscopic current amplitudes, increased outward rectification, slower desensitization, and absence of fast desensitization (Saxena and Macdonald, 1994; Fisher and Macdonald, 1997; Haas and Macdonald, 1999; Bianchi et al., 2002). Single-channel recordings indicate that α1βδ receptors exhibit brief, isolated openings with two open states, whereas α4βδ and α1βγ2 receptors open to three states (Fisher and Macdonald, 1997; Feng et al., 2006). Furthermore, α1βδ receptors have a lower GABA EC50 than α1βγ2 receptors (Saxena and Macdonald, 1994; Fisher and Macdonald, 1997), and the presence of α4 or α6 subunit confers even higher GABA sensitivity (Saxena and Macdonald, 1996; Brown et al., 2002). The slower and less extensive desensitization, and high sensitivity to GABA, of αβδ receptors, and α4 and α6 containing receptors in particular, makes these receptors ideal candidates to generate tonic GABAA inhibition.

Pharmacologically, αβδ and αβγ receptors are distinct. Not only are αβδ receptors benzodiazepine insensitive, they have increased sensitivity, compared with αβγ receptors, to allosteric modulators including zinc and lanthanum (Saxena et al., 1994, 1996), neurosteroids (see Neuron-selective actions of neurosteroids at synaptic and extrasynaptic GABAARs), ethanol (see Thalamic extrasynaptic GABAARs are a target for ethanol and volatile anesthetics), barbiturates (Saxena and Macdonald, 1996; Feng et al., 2004), certain anesthetics (see Thalamic extrasynaptic GABAARs are a target for ethanol and volatile anesthetics), the nonbenzodiazepine anxiolytic tracazolate (Zheleznova et al., 2008), and protons (Feng and Macdonald, 2004). In addition, GABA exhibits high efficacy at αβγ receptors, whereas at αβδ receptors it has a low efficacy, suggesting GABA is a partial agonist at αβδ receptors (Bianchi and Macdonald, 2003). Changes in GABA efficacy of αβδ receptors may be a general mechanism by which allosteric modulators bring about their actions.

A role for dysfunction of αβδ receptors in the pathophysiology of idiopathic generalized epilepsies has been suggested. Two δ subunit variants (E177A and R220H) have been identified as susceptibility alleles for generalized epilepsy with febrile seizures plus and juvenile myoclonic epilepsy (Dibbens et al., 2004). In HEK293T cells, recombinant hα1β2δ(E177A) and hα1β2δ(R220H) receptors exhibited reduced receptor currents, although the GABA EC50 was no different from wild-type receptors (Dibbens et al., 2004). In recombinant hα4β2δ(E177A) and hα4β2δ(R220H) receptors, GABA EC50s were also similar to wild-type receptors, but reduced macroscopic currents were caused by reduced single-channel currents attributable to shorter mean open durations and to loss of cell-surface receptor expression (Feng et al., 2006). Thus, disruption of αβδ receptor function indicates a possible role for aberrant extrasynaptic GABAARs in epileptogenesis (see Enhanced tonic GABAA inhibition in thalamic neurons is necessary and sufficient for absence seizures and The contribution of tonic GABAA inhibition to physiological and pathological hippocampal excitability).

Neuron-selective actions of neurosteroids at synaptic and extrasynaptic GABAARs

Neurosteroids, typified by the progesterone metabolite allopregnanolone, potently modulate neuronal excitability through endocrine, paracrine, or autocrine actions at GABAARs (Belelli and Lambert, 2005). Estimated brain and plasma levels of neurosteroids (10–300 nm) are dynamically regulated during certain (patho)physiological conditions, including development, later stages of pregnancy, and episodes of stress. Thus, neurosteroid modulation of GABAAR function may play an important role in these conditions (see Steroid hormones regulate extrasynaptic GABAARs during the ovarian cycle, pregnancy, and associated disorders). However, given the ubiquitous expression of GABAARs, it might be predicted that neurosteroid actions would be widespread, causing a nonspecific enhancement of neuronal inhibition that would seem incompatible with a physiological role. Belelli and colleagues have investigated neurosteroid actions and shown that they are neuron selective.

Neuronal selectivity may be the product of a range of molecular mechanisms, including subunit composition (Herd et al., 2007) so that different populations of receptors within a given neuron may exhibit different neurosteroid sensitivity. Recombinant αβγ receptors are sensitive to neurosteroids, but the identity of the α or β subunit isoform has little influence on receptor responses (Belelli et al., 2002). In contrast, in native neurons, synaptic GABAAR responses are highly heterogeneous (Cooper et al., 1999; Belelli and Herd, 2003; Harney et al., 2003). For instance, synaptic GABAARs of thalamocortical (TC) neurons of the ventrobasal (VB) thalamus are sensitive to only high concentrations of allopregnanolone (Mitchell et al., 2007), whereas even low concentrations (10 and 100 nm) enhance the synaptic inhibition of CRF-releasing parvocellular neurons of the paraventricular nucleus of the hypothalamus and inhibit their output (Gunn et al., 2009). Compared with synaptic GABAARs, δ subunit-containing extrasynaptic receptors have been proposed to be highly sensitive to low, physiologically relevant concentrations of neurosteroids, a suggestion supported both by experiments on recombinant receptors (Belelli et al., 2002; Wohlfarth et al., 2002) and the reduced behavioral sensitivity of δ subunit knock-out mice to endogenous and synthetic neuroactive steroids (Mihalek et al., 1999). Furthermore, some native δ subunit-containing receptors are indeed sensitive to low concentrations of allopregnanolone, for instance cerebellar granule cells (Stell et al., 2003). However, extrasynaptic GABAARs in TC neurons are relatively insensitive to even high concentrations of allopregnanolone (Brown et al., 2009) or 5α-THDOC (5α-tetrahydrodeoxycorticosterone) (Porcello et al., 2003). Moreover, the modest 5α-THDOC-dependent effects seen in TC neurons are still present in δ subunit knock-out mice (Porcello et al., 2003). Additional mechanisms have been shown to contribute to the neuronal-selectivity of neurosteroid actions, including the phosphorylation state of native GABAARs (Harney et al., 2003; Koksma et al., 2003) and local steroid metabolism (Belelli and Herd, 2003), although their precise roles remain to be elucidated.

Thus, both synaptic and extrasynaptic GABAARs represent targets for the actions of neurosteroids. The imminent generation of transgenic mice harboring neurosteroid-insensitive receptor isoforms will greatly aid the exploration of the relative contribution of distinct synaptic and extrasynaptic GABAARS to the putative (patho)physiological roles of neurosteroids.

Thalamic extrasynaptic GABAARs are a target for ethanol and volatile anesthetics

It has been well documented that TC neurons of the VB thalamus exhibit tonic GABAA inhibition (Belelli et al., 2005; Cope et al., 2005; Jia et al., 2005). Coimmunoprecipitation studies show that antibodies to the δ subunit precipitate the α4 subunit (Jia et al., 2005), and it has been estimated that as much as 30% of the total TC neuron GABAAR population may contain the α4 subunit (Sur et al., 1999). Furthermore, α4 and δ subunits colocalize with each other and are found predominantly at extrasynaptic sites (Jia et al., 2005), a feature confirmed by the loss of tonic inhibition in TC neurons of α4 subunit knock-out mice (Chandra et al., 2006). Because the thalamus plays a crucial role in sleep regulation (Steriade, 2000), and GABA levels have been shown to fluctuate over the sleep–wake cycle (Kékesi et al., 1997), agents that target extrasynaptic GABAARs in the thalamus may play a significant role in governing sedation, hypnosis, and consciousness.

Harrison and colleagues showed that this is the case for two clinically important agents, ethanol and the volatile anesthetic isoflurane. In control mice, low concentrations of ethanol, such as may cause intoxication, elicited a sustained current in TC neurons of the VB, that was associated with a decrease in neuronal excitability and firing rate (Jia et al., 2008b). The steady current was completely abolished by the GABAAR antagonist gabazine, and ethanol had no effect on TC neurons from α4 subunit knock-out mice. In a similar vein, volatile anesthetics are used clinically to produce analgesia, amnesia, unconsciousness, blunted autonomic responsiveness, and immobility (Campagna et al., 2003), and, at lower doses, sedation (Dwyer et al., 1992). Harrison and colleagues demonstrated that even low concentrations of isoflurane, such as may be sufficient to cause sedation, elicited a sustained current in TC neurons of the VB, associated with a conductance increase (Jia et al., 2008c). The reversal potential of the isoflurane-evoked current was close to the Cl reversal potential, was blocked by gabazine, and, as for ethanol, there was no effect of isoflurane in α4 subunit knock-out mice, even at doses that produce unconsciousness (Jia et al., 2008c). Thus, in the thalamus, as elsewhere (Sundstrom-Poromaa et al., 2002; Wallner et al., 2003; Caraiscos et al., 2004b; Glykys et al., 2007), extrasynaptic GABAARs appear to be a preferential target for both ethanol and certain volatile anesthetics.

Although the effects of ethanol and volatile anesthetics may be mediated by multiple mechanisms in the brain, the work by Harrison and colleagues, and in agreement with other studies, implicates extrasynaptic GABAARs as a major player in their principal modes of action. Furthermore, the pivotal role played by the thalamus in controlling sleep–wake states suggest that modulation of extrasynaptic GABAARs in TC neurons may contribute to global states of arousal and be a candidate target for the treatment of thalamocortical related disorders such as sleep disturbances (Belelli et al. 2005; Herd et al., 2009) and epilepsy (see Enhanced tonic GABAA inhibition in thalamic neurons is necessary and sufficient for absence seizures). For instance, disruption of normal sleep patterns by alcohol has been well documented and can be both economically important (Stoller, 1994) as well as a factor in alcoholic relapse.

Steroid hormones regulate extrasynaptic GABAARs during the ovarian cycle, pregnancy, and associated disorders

Altered levels of neurosteroids in the CNS are associated with numerous psychiatric and neurological disorders, including premenstrual syndrome (PMS), premenstrual dysphoric disorder (PMDD), catamenial epilepsy, and postpartum depression. It has long been assumed that the pathophysiology of these disorders was attributable to an adverse reaction to changing steroid hormone levels. Recent evidence, however, indicates that exogenous administration of some neurosteroids alters the expression of certain GABAAR subunits, for example the α1, α4, and γ2 subunits (Smith et al., 1998; Follesa et al., 2000, 2002; Smith, 2002), whereas γ2 subunit expression is inversely correlated to neurosteroid levels (Follesa et al., 2004). In addition, certain neurosteroids are known to potentiate the effects of GABA at selective GABAARs, particularly those containing the δ subunit (see Neuron-selective actions of neurosteroids at synaptic and extrasynaptic GABAARs). Thus, the targeting of specific GABAAR populations by neurosteroids may underlie some of their actions. Jamie Maguire and colleagues have examined the contribution of extrasynaptic GABAARs to the ovarian cycle and throughout pregnancy and the role of dysfunctional neurosteroid regulation of these receptors in associated pathophysiological disorders.

In the hippocampus, changes in GABAAR subunit expression occur over the estrous cycle, in particular a reciprocal increase in δ and decrease in γ2 subunit expression at periods of the cycle associated with elevated levels of the steroid hormone progesterone (Maguire et al., 2005; Maguire and Mody, 2007). In dentate gyrus granule cells, elevated progesterone and δ subunit levels correlates with an increase in tonic inhibition and a decrease in both anxiety levels and seizure susceptibility (Maguire et al., 2005; Maguire and Mody, 2007). Postpartum is also a particularly vulnerable period for mood disorders. In the hippocampus, GABAAR subunit expression changes during pregnancy, in particular a reduction in the expression of δ and γ2 subunits 18 d after conception, which rebounds to virgin levels by 48 h postpartum. The loss of δ and γ2 subunits is accompanied by a decrease in both tonic and phasic inhibition in the dentate gyrus (Maguire and Mody, 2008) and may represent a homeostatic compensatory mechanism to maintain excitability levels during pregnancy. In δ subunit knock-out mice, failure to regulate GABAARs during pregnancy and postpartum is reflected in an abnormal phenotype including depression-like behavior, failure to build a nest, and increased pup mortality attributable to neglect and/or increased cannibalism (Maguire and Mody, 2008). Thus, not only do extrasynaptic GABAARs represent a target for certain steroid hormones but the subsequent regulation of neuronal excitability may likely play an important role in the normal function of these hormones during the ovarian cycle, pregnancy, and postpartum. However, dysfunction in GABAAR–steroid hormone interactions may underlie multiple associated neuropsychiatric disorders including PMS, PMDD, and postpartum depression.

Enhanced tonic GABAA inhibition in thalamic neurons is necessary and sufficient for absence seizures

Absence seizures are a feature of many idiopathic generalized epilepsies and are characterized by bilaterally synchronous spike-and-wave discharges (SWDs) generated in reciprocal corticothalamo–cortical networks (Crunelli and Leresche, 2002). Despite not being directly involved in seizure initiation, the thalamus is required for both the full electrographic and behavioral expression of seizures (Polack et al., 2007), and GABAARs are clearly important. For instance, selective application of GABAmimetics to sensory thalamic nuclei can exacerbate or initiate seizures (Danober et al., 1998). However, only modest alterations in phasic GABAA inhibition have been documented in TC neurons (Bessaïh et al., 2006), and the role of tonic inhibition is unknown, despite systemic THIP administration having previously been described as a model of absence seizures (Fariello and Golden, 1987). Cope and colleagues investigated tonic inhibition in TC neurons of a prototypical sensory thalamic nucleus, the VB, from a polygenic model of SWDs, the genetic absence epilepsy rats from Strasbourg (GAERS), and showed that it was larger compared with nonepileptic controls. Importantly, this increase was seen before the onset of seizures and may therefore contribute to seizure genesis. Furthermore, tonic inhibition was also increased in other genetic models, the monogenic mutant mice stargazer (stg) and lethargic (lh), and after the application of the SWD-inducing agents GHB and THIP. In GAERS, stg and lh animals, enhanced tonic inhibition is caused not by overexpression of extrasynaptic GABAARs but by compromised GABA uptake by the GABA transporter GAT-1, leading to an increase in ambient GABA concentration. The critical importance of thalamic GAT-1 in seizure genesis was highlighted by the presence of seizures in GAT-1 knock-out mice and their induction in normal Wistar rats after the intrathalamic microinjection of a selective GAT-1 blocker.

Cope and colleagues then examined whether enhanced tonic inhibition in thalamic neurons was important in seizure genesis, or just an interesting phenomenon. GHB failed to induce SWDs in δ subunit knock-out mice that show dramatically reduced, albeit not abolished, tonic inhibition compared with wild-type littermates (Herd et al., 2009). By comparison, GHB-induced seizures were readily apparent in the littermates. In addition, spontaneous seizures in GAERS were susceptible to the intrathalamic microinfusion of an antisense oligodeoxynucleotide (ODN) to the δ subunit (Maguire et al., 2005), whereas there was no affect after microinfusion of a missense ODN. Importantly, the antisense, but not the missense, ODN reduces tonic inhibition. Last, intrathalamic microinfusion of THIP in normal Wistar rats induces both SWDs and the behavioral correlates of seizures, i.e., the full electrographic and behavioral repertoire of absence seizures was replicated. Collectively, these findings demonstrate that enhanced tonic inhibition in TC neurons is common to multiple and diverse models of absence seizures and is both necessary and sufficient for the full, i.e., electrographic and behavioral, expression of absence seizures. Furthermore, extrasynaptic GABAARs and GABA transporters in the thalamus may represent a novel therapeutic target for the treatment of absence epilepsy.

The contribution of tonic GABAA inhibition to physiological and pathological hippocampal excitability

Activation of extrasynaptic GABAARs in the hippocampus can have a profound effect on neuronal excitability (Semyanov et al., 2003). Although neuronal gain (the slope of the relationship between excitatory input and firing rate) can be altered by subthreshold synaptic “noise” (Wolfart et al., 2007; Rothman et al., 2009), it has been suggested that tonic inhibition may also play a role in gain control (Mitchell and Silver, 2003; Rothman et al., 2009). In CA1 hippocampal neurons, Walker and colleagues have demonstrated that tonic GABAA inhibition exhibits a strong outward rectification (Pavlov et al., 2008) and thus has a greater modulatory effect on excitatory inputs at or close to threshold, compared with a much reduced effect on subthreshold noise. Furthermore, using a dynamic clamp system, tonic inhibition predominantly affects the offset (left–right position) of the relationship between input and firing while having only a minimal effect on gain. Thus, at least in the hippocampus, extrasynaptic GABAA receptors modulate network excitability without altering the sensitivity of neurons to changing inputs.

Because extracellular GABA can vary during different physiological and pathological conditions, tonic inhibition may be expected to change according to brain state. Temporal lobe epilepsy frequently results from a brain insult leading to the emergence of spontaneous seizures after alterations in cellular and network properties during the subsequent latent period. This can be mimicked in animal models by chemically or electrically inducing status epilepticus as the initiating insult. In poststatus epilepticus models, there is a loss of δ and α5 GABAAR subunits (Schwarzer et al., 1997; Peng et al., 2004) that usually mediate tonic inhibition under control conditions (Caraiscos et al., 2004a; Scimemi et al., 2005). This has lead to the hypothesis that epileptogenesis is accompanied by a loss of, or reduction in, tonic inhibition. However, this is not the case, and in temporal lobe epilepsy models either during induced status epilepticus or after seizure onset, tonic inhibition can be either unaltered (Zhang et al., 2007) or indeed increased (Naylor et al., 2005; Scimemi et al., 2005). The lack of effect of the loss of subunits that normally generate tonic inhibition appears to be attributable to the upregulation of other receptor subunits and/or the translocation of receptors that are typically found at synaptic specializations into the extrasynaptic membrane (Peng et al., 2004; Scimemi et al., 2005; Zhang et al., 2007). Furthermore, Walker and colleagues showed that the preservation of tonic inhibition also occurs in a posttraumatic epilepsy model (Kharatishvili et al., 2006), where a reduction in phasic GABAA inhibition is not accompanied by a loss of tonic inhibition. These findings are in agreement with human epileptic tissue where tonic inhibition is preserved (Scimemi et al., 2006) and support a common paradigm in which tonic inhibition is maintained or enhanced during temporal lobe epilepsy, perhaps as a homeostatic mechanism to counter the concomitant loss of phasic inhibition.

Concluding remarks

In conclusion, the field of GABAAR research is excitingly poised to make significant advances in our understanding of the distinct contributions of both synaptic and extrasynaptic GABAARs to physiological and pathophysiological CNS function. The advent of global GABAAR subunit-specific knock-out and knock-in mice has greatly aided the identification of synaptic and extrasynaptic GABAAR isoforms in discrete neuronal populations. However, notwithstanding the fact that synaptic and extrasynaptic GABAARs can be distinguished by classical GABAAR antagonists in some neurons (Park et al., 2006), the specific roles of different subtypes of GABAARs will only be determined by the development of neuron-specific and/or conditional transgenic mice (Gavériaux-Ruff and Kieffer, 2007; Wulff et al., 2007) or receptor subtype-specific antagonists and inverse agonists.

Footnotes

This work was supported by National Institutes of Health Grants GM61925, GM45129, and AA16393 (N.L.H.), MH076994 (J.M.), and NS33300 and NS51590 (R.L.M.); Wellcome Trust Grants WT083163MF (M.C.W.) and 71436 (D.W.C.); Medical Research Council Grant G0400136 (M.C.W.); Biotechnology and Biological Research Council Grant C509923, and Case and Strategic Studentships 11426 and 12019 (D.B.); European Union Grant FP6 LSHM-CT-2006-037315 (M.C.W.); Tenovus Scotland (D.B.); AJ Clarck Studentship 2007 (D.B.); and Epilepsy Research UK Grants 0404 (M.C.W.) and F0802 (D.W.C.).

References

  1. Ade KK, Janssen MJ, Ortinski PI, Vicini S. Differential tonic GABA conductances in striatal medium spiny neurons. J Neurosci. 2008;28:1185–1197. doi: 10.1523/JNEUROSCI.3908-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Bai D, Zhu G, Pennefather P, Jackson MF, MacDonald JF, Orser BA. Distinct functional and pharmacological properties of tonic and quantal inhibitory postsynaptic currents mediated by γ-aminobutyric acidA receptors in hippocampal neurons. Mol Pharmacol. 2001;59:814–824. doi: 10.1124/mol.59.4.814. [DOI] [PubMed] [Google Scholar]
  3. Belelli D, Herd MB. The contraceptive agent Provera enhanced GABAA receptor-mediated inhibitory neurotransmission in the rat hippocampus: evidence for endogenous neurosteroids? J Neurosci. 2003;23:10013–10020. doi: 10.1523/JNEUROSCI.23-31-10013.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Belelli D, Lambert JJ. Neurosteroids: endogenous regulators of the GABAA receptor. Nat Rev Neurosci. 2005;6:565–575. doi: 10.1038/nrn1703. [DOI] [PubMed] [Google Scholar]
  5. Belelli D, Casula A, Ling A, Lambert JJ. The influence of subunit composition on the interaction of neurosteroids with GABAA receptors. Neuropharmacology. 2002;43:651–661. doi: 10.1016/s0028-3908(02)00172-7. [DOI] [PubMed] [Google Scholar]
  6. Belelli D, Peden DR, Rosahl TW, Wafford KA, Lambert JJ. Extrasynaptic GABAA receptors of thalamocortical neurons: a molecular target for hypnotics. J Neurosci. 2005;25:11513–11520. doi: 10.1523/JNEUROSCI.2679-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Bessaïh T, Bourgeais L, Badiu CI, Carter DA, Toth TI, Ruano D, Lambolez B, Crunelli V, Leresche N. Nucleus-specific abnormalities of GABAergic synaptic transmission in a genetic model of absence seizures. J Neurophysiol. 2006;96:3074–3081. doi: 10.1152/jn.00682.2006. [DOI] [PubMed] [Google Scholar]
  8. Bianchi MT, Macdonald RL. Slow phases of GABAA receptor desensitization: structural determinants and possible relevance for synaptic function. J Physiol. 2002;544:3–18. doi: 10.1113/jphysiol.2002.020255. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Bianchi MT, Macdonald RL. Neurosteroids shift partial agonist activation of GABAA receptor channels from low- to high-efficacy gating patterns. J Neurosci. 2003;23:10934–10943. doi: 10.1523/JNEUROSCI.23-34-10934.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Bianchi MT, Haas KF, Macdonald RL. α1 and α6 subunits specify distinct desensitization, deactivation and neurosteroid modulation of GABAA receptors containing the δ subunit. Neuropharmacology. 2002;43:492–502. doi: 10.1016/s0028-3908(02)00163-6. [DOI] [PubMed] [Google Scholar]
  11. Brickley SG, Cull-Candy SG, Farrant M. Development of a tonic form of synaptic inhibition in rat cerebellar granule cells resulting from persistent activation of GABAA receptors. J Physiol. 1996;497:753–759. doi: 10.1113/jphysiol.1996.sp021806. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Brickley SG, Revilla V, Cull-Candy SG, Wisden W, Farrant M. Adaptive regulation of neuronal excitability by a voltage-independent potassium conductance. Nature. 2001;409:88–92. doi: 10.1038/35051086. [DOI] [PubMed] [Google Scholar]
  13. Bright DP, Aller MI, Brickley SG. Synaptic release generates a tonic GABAA receptor-mediated conductance that modulates burst precision in thalamic relay neurons. J Neurosci. 2007;27:2560–2569. doi: 10.1523/JNEUROSCI.5100-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Brown AR, Herd MB, Belelli D, Lambert JJ. The actions of neurosteroids and THIP upon thalamic inhibitory transmission: the role of extrasynaptic α4β2δ GABAA receptors. In: Panzica GC, Gotti S, editors. Trabajos del Instituto Cajal International Meeting: Steroids and nervous system. Madrid: Instituto Cajal; 2009. p. 182. [Google Scholar]
  15. Brown N, Kerby J, Bonnert TP, Whiting PJ, Wafford KA. Pharmacological characterization of a novel cell line expressing human α4β3δ GABAA receptors. Br J Pharmacol. 2002;136:965–974. doi: 10.1038/sj.bjp.0704795. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Campagna JA, Miller KW, Forman SA. Mechanisms of actions of inhaled anesthetics. N Engl J Med. 2003;348:2110–2124. doi: 10.1056/NEJMra021261. [DOI] [PubMed] [Google Scholar]
  17. 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 α5 subunit-containing γ-aminobutyric acid type A receptors. Proc Natl Acad Sci U S A. 2004a;101:3662–3667. doi: 10.1073/pnas.0307231101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Caraiscos VB, Newell JG, You-Ten KE, Elliott EM, Rosahl TW, Wafford KA, MacDonald JF, Orser BA. Selective enhancement of tonic GABAergic inhibition in murine hippocampal neurons by low concentrations of the volatile anesthetic isoflurane. J Neurosci. 2004b;24:8454–8458. doi: 10.1523/JNEUROSCI.2063-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Chadderton P, Margrie TW, Häusser M. Integration of quanta in cerebellar granule cells during sensory processing. Nature. 2004;428:856–860. doi: 10.1038/nature02442. [DOI] [PubMed] [Google Scholar]
  20. Chandra D, Jia F, Liang J, Peng Z, Suryanarayanan A, Werner DF, Spigelman I, Houser CR, Olsen RW, Harrison NL, Homanics GE. GABAA receptor α4 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]
  21. Chang Y, Wang R, Barot S, Weiss DS. Stoichiometry of a recombinant GABAA receptor. J Neurosci. 1996;16:5415–5424. doi: 10.1523/JNEUROSCI.16-17-05415.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Cheng VY, Martin LJ, Elliott EM, Kim JH, Mount HT, Taverna FA, Roder JC, Macdonald JF, Bhambri A, Collinson N, Wafford KA, Orser BA. α5GABAA receptors mediate the amnestic but not sedative-hypnotic effects of the general anesthetic etomidate. J Neurosci. 2006a;26:3713–3720. doi: 10.1523/JNEUROSCI.5024-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Cheng VY, Bonin RP, Chiu MW, Newell JG, MacDonald JF, Orser BA. Gabapentin increases a tonic inhibitory conductance in hippocampal pyramidal neurons. Anesthesiology. 2006b;105:325–333. doi: 10.1097/00000542-200608000-00015. [DOI] [PubMed] [Google Scholar]
  24. Cooper EJ, Johnston GA, Edwards FA. Effects of a naturally occurring neurosteroid on GABAA IPSCs during development in rat hippocampal slices. J Physiol. 1999;521:437–449. doi: 10.1111/j.1469-7793.1999.00437.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Cope DW, Hughes SW, Crunelli V. GABAA receptor-mediated tonic inhibition in thalamic neurons. J Neurosci. 2005;25:11553–11563. doi: 10.1523/JNEUROSCI.3362-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Crunelli V, Leresche N. Childhood absence epilepsy: genes, channels, neurons and networks. Nat Rev Neurosci. 2002;3:371–382. doi: 10.1038/nrn811. [DOI] [PubMed] [Google Scholar]
  27. Curia G, Papouin T, Séguéla P, Avoli M. Downregulation of tonic GABAergic inhibition in a mouse model of Fragile X syndrome. Cereb Cortex. 2009;19:1515–1520. doi: 10.1093/cercor/bhn159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Danober L, Deransart C, Depaulis A, Vergnes M, Marescaux C. Pathophysiological mechanisms of genetic absence epilepsy in the rat. Prog Neurobiol. 1998;55:27–57. doi: 10.1016/s0301-0082(97)00091-9. [DOI] [PubMed] [Google Scholar]
  29. Dibbens LM, Feng HJ, Richards MC, Harkin LA, Hodgson BL, Scott D, Jenkins M, Petrou S, Sutherland GR, Scheffer IE, Berkovic SF, Macdonald RL, Mulley JC. GABRD encoding a protein for extra- or peri-synaptic GABAA receptors is a susceptibility locus for generalized epilepsies. Hum Mol Genet. 2004;13:1315–1319. doi: 10.1093/hmg/ddh146. [DOI] [PubMed] [Google Scholar]
  30. Drasbek KR, Jensen K. THIP, a hypnotic and antinociceptive drug, enhances an extrasynaptic GABAA receptor-mediated conductance in mouse neocortex. Cereb Cortex. 2006;16:1134–1141. doi: 10.1093/cercor/bhj055. [DOI] [PubMed] [Google Scholar]
  31. Drasbek KR, Vardya I, Delenclos M, Gibson KM, Jensen K. SSADH deficiency leads to elevated extracellular GABA levels and increased GABAergic neurotransmission in the mouse barrel cortex. J Inherit Metab Dis. 2008;31:662–668. doi: 10.1007/s10545-008-0941-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Dwyer R, Bennett HL, Eger EI, 2nd, Heilbron D. Effects of isoflurane and nitrous oxide in subanesthetic concentrations on memory and responsiveness in volunteers. Anesthesiology. 1992;77:888–898. doi: 10.1097/00000542-199211000-00009. [DOI] [PubMed] [Google Scholar]
  33. Essrich C, Lorez M, Benson JA, Fritschy JM, Lüscher B. Postsynaptic clustering of major GABAA receptor subtypes requires the γ2 subunit and gephyrin. Nat Neurosci. 1998;1:563–571. doi: 10.1038/2798. [DOI] [PubMed] [Google Scholar]
  34. Fariello RG, Golden GT. The THIP-induced model of bilateral synchronous spike and wave in rodents. Neuropharmacology. 1987;26:161–165. doi: 10.1016/0028-3908(87)90204-8. [DOI] [PubMed] [Google Scholar]
  35. Farrant M, Nusser Z. Variations on an inhibitory theme: phasic and tonic activation of GABAA receptors. Nat Rev Neurosci. 2005;6:215–229. doi: 10.1038/nrn1625. [DOI] [PubMed] [Google Scholar]
  36. Feng HJ, Macdonald RL. Proton modulation of α1β3δ GABAA receptor channel gating and desensitization. J Neurophysiol. 2004;92:1577–1585. doi: 10.1152/jn.00285.2004. [DOI] [PubMed] [Google Scholar]
  37. Feng HJ, Bianchi MT, Macdonald RL. Pentobarbital differentially modulates α1β3δ and α1β3γ2L GABAA receptor currents. Mol Pharmacol. 2004;66:988–1003. doi: 10.1124/mol.104.002543. [DOI] [PubMed] [Google Scholar]
  38. 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. J Neurosci. 2006;26:1499–1506. doi: 10.1523/JNEUROSCI.2913-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Fisher JL, Macdonald RL. Single channel properties of recombinant GABAA receptors containing γ2 or δ subtypes expressed with α1 and β3 subtypes in mouse L929 cells. J Physiol. 1997;505:283–297. doi: 10.1111/j.1469-7793.1997.283bb.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Follesa P, Serra M, Cagetti E, Pisu MG, Porta S, Floris S, Massa F, Sanna E, Biggio G. Allopregnanolone synthesis in cerebellar granule cells: roles in regulation of GABAA receptor expression and function during progesterone treatment and withdrawal. Mol Pharmacol. 2000;57:1262–1270. [PubMed] [Google Scholar]
  41. Follesa P, Porcu P, Sogliano C, Cinus M, Biggio F, Mancuso L, Mostallino MC, Paoletti AM, Purdy RH, Biggio G, Concas A. Changes in GABAA receptor γ2 subunit gene expression induced by long-term administration of oral contraceptives in rats. Neuropharmacology. 2002;42:325–336. doi: 10.1016/s0028-3908(01)00187-3. [DOI] [PubMed] [Google Scholar]
  42. Follesa P, Biggio F, Caria S, Gorini G, Biggio G. Modulation of GABAA receptor gene expression by allopregnanolone and ethanol. Eur J Pharmacol. 2004;500:413–425. doi: 10.1016/j.ejphar.2004.07.041. [DOI] [PubMed] [Google Scholar]
  43. Gavériaux-Ruff C, Kieffer BL. Conditional gene targeting in the mouse nervous system: insights into brain function and diseases. Pharmacol Ther. 2007;113:619–634. doi: 10.1016/j.pharmthera.2006.12.003. [DOI] [PubMed] [Google Scholar]
  44. Glykys J, Mody I. Activation of GABAA receptors: views from outside the synaptic cleft. Neuron. 2007;56:763–770. doi: 10.1016/j.neuron.2007.11.002. [DOI] [PubMed] [Google Scholar]
  45. 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. Nat Neurosci. 2007;10:40–48. doi: 10.1038/nn1813. [DOI] [PubMed] [Google Scholar]
  46. Gunn BG, Lambert JJ, Belelli D. GABAA receptors and neurosteroid actions upon hypothalamic parvocellular neurones. In: Panzica GC, Gotti S, editors. Trabajos del Instituto Cajal International Meeting: Steroids and nervous system. Madrid: Instituto Cajal; 2009. p. 182. [Google Scholar]
  47. Haas KF, Macdonald RL. GABAA receptor subunit γ2 and δ subtypes confer unique kinetic properties on recombinant GABAA receptor currents in mouse fibroblasts. J Physiol. 1999;514:27–45. doi: 10.1111/j.1469-7793.1999.027af.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Hamann M, Rossi DJ, Attwell D. Tonic and spillover inhibition of granule cells control information flow through cerebellar cortex. Neuron. 2002;33:625–633. doi: 10.1016/s0896-6273(02)00593-7. [DOI] [PubMed] [Google Scholar]
  49. Harney SC, Frenguelli BG, Lambert JJ. Phosphorylation influences neurosteroid modulation of synaptic GABAA receptors in rat CA1 and dentate gyrus neurones. Neuropharmacology. 2003;45:873–883. doi: 10.1016/s0028-3908(03)00251-x. [DOI] [PubMed] [Google Scholar]
  50. Herd MB, Belelli D, Lambert JJ. Neurosteroid modulation of synaptic and extrasynaptic GABAA receptors. Pharmacol Ther. 2007;116:20–34. doi: 10.1016/j.pharmthera.2007.03.007. [DOI] [PubMed] [Google Scholar]
  51. Herd MB, Foister N, Chandra D, Peden DR, Homanics GE, Brown VJ, Balfour DJ, Lambert JJ, Belelli D. Inhibition of thalamic excitability by 4,5,6,7-tetrahydroisoxazolo[4,5-c]pyridine-3-ol: a selective role for δ-GABAA receptors. Eur J Neurosci. 2009;29:1177–1187. doi: 10.1111/j.1460-9568.2009.06680.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Janssen MJ, Ade KK, Fu Z, Vicini S. Dopamine modulation of GABA tonic conductance in striatal output neurons. J Neurosci. 2009;29:5116–5126. doi: 10.1523/JNEUROSCI.4737-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Jia F, Pignataro L, Schofield CM, Yue M, Harrison NL, Goldstein PA. An extrasynaptic GABAA receptor mediates tonic inhibition in thalamic VB neurons. J Neurophysiol. 2005;94:4491–4501. doi: 10.1152/jn.00421.2005. [DOI] [PubMed] [Google Scholar]
  54. Jia F, Yue M, Chandra D, Keramidas A, Goldstein PA, Homanics GE, Harrison NL. Taurine is a potent activator of extrasynaptic GABAA receptors in the thalamus. J Neurosci. 2008a;28:106–115. doi: 10.1523/JNEUROSCI.3996-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Jia F, Chandra D, Homanics GE, Harrison NL. Ethanol modulates synaptic and extrasynaptic GABAA receptors in the thalamus. J Pharmacol Exp Ther. 2008b;326:475–482. doi: 10.1124/jpet.108.139303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Jia F, Yue M, Chandra D, Homanics GE, Goldstein PA, Harrison NL. Isoflurane is a potent modulator of extrasynaptic GABAA receptors in the thalamus. J Pharmacol Exp Ther. 2008c;324:1127–1135. doi: 10.1124/jpet.107.134569. [DOI] [PubMed] [Google Scholar]
  57. Kaur KH, Baur R, Sigel E. Unanticipated structural and functional properties of δ-subunit-containing GABAA receptors. J Biol Chem. 2009;284:7889–7896. doi: 10.1074/jbc.M806484200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Kékesi KA, Dobolyi A, Salfay O, Nyitrai G, Juhász G. Slow wave sleep is accompanied by release of certain amino acids in the thalamus of cats. Neuroreport. 1997;8:1183–1186. doi: 10.1097/00001756-199703240-00025. [DOI] [PubMed] [Google Scholar]
  59. Kharatishvili I, Nissinen JP, McIntosh TK, Pitkänen A. A model of posttraumatic epilepsy induced by lateral fluid-percussion brain injury in rats. Neuroscience. 2006;140:685–697. doi: 10.1016/j.neuroscience.2006.03.012. [DOI] [PubMed] [Google Scholar]
  60. Koksma JJ, van Kesteren RE, Rosahl TW, Zwart R, Smit AB, Lüddens H, Brussaard AB. Oxytocin regulates neurosteroid modulation of GABAA receptors in supraoptic nucleus around partition. J Neurosci. 2003;23:788–797. doi: 10.1523/JNEUROSCI.23-03-00788.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Krogsgaard-Larsen P, Frølund B, Liljefors T, Ebert B. GABAA agonists and partial agonists: THIP (Gaboxadol) as a non-opioid analgesic and a novel type of hypnotic. Biochem Pharmacol. 2004;68:1573–1580. doi: 10.1016/j.bcp.2004.06.040. [DOI] [PubMed] [Google Scholar]
  62. Krook-Magnuson EI, Li P, Paluszkiewicz SM, Huntsman MM. Tonically active inhibition selectively controls feedforward circuits in mouse barrel cortex. J Neurophysiol. 2008;100:932–944. doi: 10.1152/jn.01360.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Maguire J, Mody I. Neurosteroid synthesis-mediated regulation of GABAA receptors: relevance to the ovarian cycle and stress. J Neurosci. 2007;27:2155–2162. doi: 10.1523/JNEUROSCI.4945-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Maguire J, Mody I. GABAAR plasticity during pregnancy: relevance to postpartum depression. Neuron. 2008;59:207–213. doi: 10.1016/j.neuron.2008.06.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Maguire JL, Stell BM, Rafizadeh M, Mody I. Ovarian cycle-linked changes in GABAA receptors mediating tonic inhibition alter seizure susceptibility and anxiety. Nat Neurosci. 2005;8:797–804. doi: 10.1038/nn1469. [DOI] [PubMed] [Google Scholar]
  66. Mihalek RM, Banerjee PK, Korpi ER, Quinlan JJ, Firestone LL, Mi ZP, Lagenaur C, Tretter V, Sieghart W, Anagnostaras SG, Sage JR, Fanselow MS, Guidotti A, Spigelman I, Li Z, DeLorey TM, Olsen RW, Homanics GE. Attenuated sensitivity to neuroactive steroids in γ-aminobutyrate type A receptor delta subunit knockout mice. Proc Natl Acad Sci U S A. 1999;96:12905–12910. doi: 10.1073/pnas.96.22.12905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Mitchell EA, Gentet LJ, Dempster J, Belelli D. GABAA and glycine receptor-mediated transmission in rat lamina II neurons: relevance to the analgesic actions of neuroactive steroids. J Physiol. 2007;583:1021–1040. doi: 10.1113/jphysiol.2007.134445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Mitchell SJ, Silver RA. Shunting inhibition modulates neuronal gain during synaptic excitation. Neuron. 2003;38:433–445. doi: 10.1016/s0896-6273(03)00200-9. [DOI] [PubMed] [Google Scholar]
  69. Naylor DE, Liu H, Wasterlain CG. Trafficking of GABAA receptors, loss of inhibition, and a mechanism for pharmacoresistance in status epilepticus. J Neurosci. 2005;25:7724–7733. doi: 10.1523/JNEUROSCI.4944-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Nusser Z, Mody I. Selective modulation of tonic and phasic inhibition in dentate gyrus granule cells. J Neurophysiol. 2002;87:2624–2628. doi: 10.1152/jn.2002.87.5.2624. [DOI] [PubMed] [Google Scholar]
  71. 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]
  72. Otis TS, Staley KJ, Mody I. Perpetual inhibitory activity in mammalian brain slices generated by spontaneous GABA release. Brain Res. 1991;545:142–150. doi: 10.1016/0006-8993(91)91280-e. [DOI] [PubMed] [Google Scholar]
  73. Park JB, Skalska S, Stern JE. Characterization of a novel tonic γ-aminobutyric acidA receptor-mediated inhibition in magnocellular neurosecretory neurons and its modulation by glia. Endocrinology. 2006;147:3746–3760. doi: 10.1210/en.2006-0218. [DOI] [PubMed] [Google Scholar]
  74. Park JB, Skalska S, Son S, Stern JE. Dual GABAA receptor-mediated inhibition in rat presympathetic paraventricular nucleus neurons. J Physiol. 2007;582:539–551. doi: 10.1113/jphysiol.2007.133223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Pavlov I, Semyanov A, Kullmann DM, Walker MC. Tonic GABAA receptor-mediated currents in hippocampal CA1 pyramidal cells exhibit strong outward rectification. Soc Neurosci Abstr. 2008 34.531.17. [Google Scholar]
  76. 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. J Neurosci. 2004;24:8629–8639. doi: 10.1523/JNEUROSCI.2877-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Polack PO, Guillemain I, Hu E, Deransart C, Depaulis A, Charpier S. Deep layer somatosensory cortical neurons initiate spike-and-wave discharges in a genetic model of absence seizures. J Neurosci. 2007;27:6590–6599. doi: 10.1523/JNEUROSCI.0753-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. 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 δ subunit. J Neurophysiol. 2003;89:1378–1386. doi: 10.1152/jn.00899.2002. [DOI] [PubMed] [Google Scholar]
  79. Rothman JS, Cathala L, Steuber V, Silver RA. Synaptic depression enables neuronal gain control. Nature. 2009;457:1015–1018. doi: 10.1038/nature07604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Salin PA, Prince DA. Spontaneous GABAA receptor-mediated inhibitory currents in adult rat somatosensory cortex. J Neurophysiol. 1996;75:1573–1588. doi: 10.1152/jn.1996.75.4.1573. [DOI] [PubMed] [Google Scholar]
  81. Saxena NC, Macdonald RL. Assembly of GABAA receptor subunits: role of the δ subunit. J Neurosci. 1994;14:7077–7086. doi: 10.1523/JNEUROSCI.14-11-07077.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Saxena NC, Macdonald RL. Properties of putative cerebellar γ-aminobutyric acidA receptor isoforms. Mol Pharmacol. 1996;49:567–579. [PubMed] [Google Scholar]
  83. Schwarzer C, Tsunashima K, Wanzenböck C, Fuchs K, Sieghart W, Sperk G. GABAA receptor subunits in the rat hippocampus II: altered distribution in kainic acid-induced temporal lobe epilepsy. Neuroscience. 1997;80:1001–1017. doi: 10.1016/s0306-4522(97)00145-0. [DOI] [PubMed] [Google Scholar]
  84. Scimemi A, Semyanov A, Sperk G, Kullmann DM, Walker MC. Multiple and plastic receptors mediate tonic GABAA receptor currents in the hippocampus. J Neurosci. 2005;25:10016–10024. doi: 10.1523/JNEUROSCI.2520-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Scimemi A, Andersson A, Heeroma JH, Strandberg J, Rydenhag B, McEvoy AW, Thom M, Asztely F, Walker MC. Tonic GABAA receptor-mediated currents in human brain. Eur J Neurosci. 2006;24:1157–1160. doi: 10.1111/j.1460-9568.2006.04989.x. [DOI] [PubMed] [Google Scholar]
  86. Semyanov A, Walker MC, Kullmann DM. GABA uptake regulates cortical excitability via cell type-specific tonic inhibition. Nat Neurosci. 2003;6:484–490. doi: 10.1038/nn1043. [DOI] [PubMed] [Google Scholar]
  87. 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]
  88. Shen H, Gong QH, Aoki C, Yuan M, Ruderman Y, Dattilo M, Williams K, Smith SS. Reversal of neurosteroid effects at α4β2δ GABAA receptors triggers anxiety at puberty. Nat Neurosci. 2007;10:469–477. doi: 10.1038/nn1868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Smith SS. Withdrawal properties of a neuroactive steroid: implication for GABAA receptor gene regulation in the brain and anxiety behavior. Steroids. 2002;67:519–528. doi: 10.1016/s0039-128x(01)00170-2. [DOI] [PubMed] [Google Scholar]
  90. Smith SS, Gong QH, Hsu FC, Markowitz RS, ffrench-Mullen JM, Li X. GABAA receptor α4 subunit suppression prevents withdrawal properties of an endogenous steroid. Nature. 1998;392:926–930. doi: 10.1038/31948. [DOI] [PubMed] [Google Scholar]
  91. Soltesz I, Roberts JD, Takagi H, Richards JG, Mohler H, Somogyi P. Synaptic and nonsynaptic localization of benzodiazepine/GABAA receptor/Cl- channel complex using monoclonal antibodies in the dorsal lateral geniculate nucleus of the cat. Eur J Neurosci. 1990;2:414–429. doi: 10.1111/j.1460-9568.1990.tb00434.x. [DOI] [PubMed] [Google Scholar]
  92. Somogyi P, Takagi H, Richards JG, Mohler H. Subcellular localization of benzodiazepine/GABAA receptors in the cerebellum of rat, cat and monkey using monoclonal antibodies. J Neurosci. 1989;9:2197–2209. doi: 10.1523/JNEUROSCI.09-06-02197.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. 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. Proc Natl Acad Sci U S A. 2003;100:14439–14444. doi: 10.1073/pnas.2435457100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Steriade M. Corticothalamic resonance, states of vigilance and mentation. Neuroscience. 2000;101:243–276. doi: 10.1016/s0306-4522(00)00353-5. [DOI] [PubMed] [Google Scholar]
  95. Stoller MK. Economic effects of insomnia. Clin Ther. 1994;16:873–897. discussion 854. [PubMed] [Google Scholar]
  96. Sundstrom-Poromaa I, Smith DH, Gong QH, Sabado TN, Li X, Light A, Wiedmann M, Williams K, Smith SS. Hormonally regulated α4β2δ GABAA receptors are a target for alcohol. Nat Neurosci. 2002;5:721–722. doi: 10.1038/nn888. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Sur C, Farrar SJ, Kerby J, Whiting PJ, Atack JR, McKernan RM. Preferential coassembly of α4 and δ subunits of the γ-aminobutyric acidA receptor in rat thalamus. Mol Pharmacol. 1999;56:110–115. doi: 10.1124/mol.56.1.110. [DOI] [PubMed] [Google Scholar]
  98. Takahashi A, Mashimo T, Uchida I. GABAergic tonic inhibition of substantia gelatinosa neurons in mouse spinal cord. Neuroreport. 2006;17:1331–1335. doi: 10.1097/01.wnr.0000230515.86090.bc. [DOI] [PubMed] [Google Scholar]
  99. 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]
  100. Wall MJ, Usowicz MM. Development of action potential-dependent and independent spontaneous GABAA receptor-mediated currents in granule cells of postnatal rat cerebellum. Eur J Neurosci. 1997;9:533–548. doi: 10.1111/j.1460-9568.1997.tb01630.x. [DOI] [PubMed] [Google Scholar]
  101. Wallner M, Hanchar HJ, Olsen RW. Ethanol enhances α4β3δ and α6β3δ γ-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]
  102. Wang L, Spary E, Deuchars J, Deuchars SA. Tonic GABAergic inhibition of sympathetic preganglionic neurons: a novel substrate for sympathetic control. J Neurosci. 2008;28:12445–12452. doi: 10.1523/JNEUROSCI.2951-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Wei W, Zhang N, Peng Z, Houser CR, Mody I. Perisynaptic localization of δ subunit-containing GABAA receptors and their activation by GABA spillover in the mouse dentate gyrus. J Neurosci. 2003;23:10650–10661. doi: 10.1523/JNEUROSCI.23-33-10650.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Wolfart J, Debay D, Le Masson G, Destexhe A, Bal T. Synaptic background activity controls spike transfer from thalamus to cortex. Nat Neurosci. 2005;8:1760–1767. doi: 10.1038/nn1591. [DOI] [PubMed] [Google Scholar]
  105. Wohlfarth KM, Bianchi MT, Macdonald RL. Enhanced neurosteroid potentiation of ternary GABAA receptors containing the δ subunit. J Neurosci. 2002;22:1541–1549. doi: 10.1523/JNEUROSCI.22-05-01541.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Wulff P, Goetz T, Leppä E, Linden AM, Renzi M, Swinny JD, Vekovischeva OY, Sieghart W, Somogyi P, Korpi ER, Farrant M, Wisden W. From synapses to behaviour: rapid modulation of defined neuronal types with engineered GABAA receptors. Nat Neurosci. 2007;10:923–929. doi: 10.1038/nn1927. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Yamada J, Furukawa T, Ueno S, Yamamoto S, Fukuda A. Molecular basis for the GABAA receptor-mediated tonic inhibition in rat somatosensory cortex. Cereb Cortex. 2007;17:1782–1787. doi: 10.1093/cercor/bhl087. [DOI] [PubMed] [Google Scholar]
  108. Zhang N, Wei W, Mody I, Houser CR. Altered localization of GABAA receptor subunits on dentate granule cell dendrites influences tonic and phasic inhibition in a mouse model of epilepsy. J Neurosci. 2007;27:7520–7531. doi: 10.1523/JNEUROSCI.1555-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Zheleznova N, Sedelnikova A, Weiss DS. α1β2δ, a silent GABAA receptor: recruitment by tracazolate and neurosteroids. Br J Pharmacol. 2008;153:1062–1071. doi: 10.1038/sj.bjp.0707665. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from The Journal of Neuroscience are provided here courtesy of Society for Neuroscience

RESOURCES