Skip to main content
Journal of Neurophysiology logoLink to Journal of Neurophysiology
. 2015 Mar 18;113(9):3421–3431. doi: 10.1152/jn.00045.2015

Neurosteroids increase tonic GABAergic inhibition in the lateral section of the central amygdala in mice

H Romo-Parra 1,*,, P Blaesse 1,*, L Sosulina 1,*, H-C Pape 1
PMCID: PMC4455571  PMID: 25787948

Abstract

Neurosteroids are formed de novo in the brain and can modulate both inhibitory and excitatory neurotransmission. Recent evidence suggests that the anxiolytic effects of neurosteroids are mediated by the amygdala, a key structure for emotional and cognitive behaviors. Tonic inhibitory signaling via extrasynaptic type A γ-aminobutyric acid receptors (GABAARs) is known to be crucially involved in regulating network activity in various brain regions including subdivisions of the amygdala. Here we provide evidence for the existence of tonic GABAergic inhibition generated by the activation of δ-subunit-containing GABAARs in neurons of the lateral section of the mouse central amygdala (CeAl). Furthermore, we show that neurosteroids play an important role in the modulation of tonic GABAergic inhibition in the CeAl. Taken together, these findings provide new mechanistic insights into the effects of pharmacologically relevant neurosteroids in the amygdala and might be extrapolated to the regulation of anxiety.

Keywords: neurosteroids, central amygdala, tonic inhibition, extrasynaptic GABAA receptors, anxiety


neurosteroids, also known as “neuroactive steroids” (Paul and Purdy 1992), can be synthesized de novo in both neurons and glia from progesterone (Cheney et al. 1995) or deoxycorticosterone (Khisti et al. 2005; Purdy et al. 1991) by the action of the enzymes 5α-reductase and 3α-hydroxysteroid dehydrogenase, respectively (Karavolas and Hodges 1991). A key function of neurosteroids is the modulation of neuronal excitability through interaction with certain neurotransmitter receptors (Lambert et al. 1995; Majewska et al. 1989; Paul and Purdy 1992; Rupprecht 1997; Rupprecht and Holsboer 1999).

Progesterone metabolites such as 3α-hydroxy-5α-pregnan-20-one (allopregnanolone or 3α,5α-THPROG) and 3α,5α-tetrahydrodeoxycorticosterone (3α,5α-THDOC) are among the most potent positive allosteric modulators of type A γ-aminobutyric acid receptors (GABAARs) (Majewska et al. 1989), and their administration in pharmacological doses elicits anxiolytic, anticonvulsant, and sedative-hypnotic effects in rodents (Majewska 1992). Specifically, local infusion of allopregnanolone into the rat central nucleus of the amygdala (CeA) produced anxiolytic-like behavioral effects (Akwa et al. 1999). Thus neuroactive steroids might play a specific role in the modulation of GABAergic transmission in this region and in the regulation of anxiety.

The neurotransmitter GABA can operate in two modes: a “phasic” mode involving the activation of synaptically clustered GABAARs by quantal release of GABA and a “tonic” mode mediated by GABAARs located extrasynaptically (Farrant and Nusser 2005; Mody 2005). GABAARs are ligand-gated chloride (and bicarbonate) channels (Olsen and Sieghart 2009) consisting of five subunits that belong to various subunit classes (α1–α6, β1–β4, γ1–γ3, δ, ε, π, θ, ρ1–ρ3) (Barnard et al. 1998; Kumar et al. 2002; Whiting 1999). Synaptic and extrasynaptic receptors exhibit distinct structural and biophysical properties (Belelli et al. 2009). Most GABAARs, however, consist of two α-, two β-, and one γ- or δ-subunit (Sieghart and Sperk 2002). Synaptic GABAARs almost invariably incorporate the γ2-subunit in combination with α (α1, α2, and α3)- and β2/3-subunit isoforms, while the extrasynaptic receptors are distinct from their synaptic counterparts and are composed of the α4-, α5-, or α6-subunit in combination with β- and δ-subunits (Belelli et al. 2006).

GABAergic tonic inhibition has been described in many regions of the brain (Farrant and Nusser 2005; Glykys and Mody 2007; Semyanov et al. 2004) including the lateral amygdala (LA), the basolateral amygdala (BA; Marowsky et al. 2012), and the medial section of the CeA (CeAm; Herman et al. 2013). Besides, so far, three GABAAR subunits have been found to be expressed in mouse CeA: α1, δ, and α5, the first two of which were functionally related to the generation of tonic inhibition (Herman et al. 2013). Neurosteroids act preferentially but not exclusively on the GABAAR δ-subunit (Brown et al. 2002). Moreover, changes in the tonic GABAAR-mediated conductance are of high clinical relevance as they are related to schizophrenia, epilepsy, and Parkinson disease (for review see Brickley and Mody 2012).

The CeA is a primarily GABAergic nucleus and forms the main output station of the amygdala complex, thereby being a critical component of the neurocircuitry underlying stress, fear and anxiety, and pain (Ehrlich et al. 2009; Neugebauer et al. 2004; Pape and Pare 2010). Not only does the GABAergic system in the amygdala modulate fear and anxiety under physiological conditions, but alterations in this system may also predispose individuals to pathological anxiety traits (Parsons and Ressler 2013; Tasan et al. 2011).

It has been shown that the neurosteroid allopregnanolone affects phasic GABAergic transmission in CeA neurons by the negative modulation of evoked inhibitory postsynaptic currents (IPSCs), which might involve the activation of NMDA receptors since the effect was occluded by the NMDA receptor antagonist dl-2-amino-5-phosphonovalerate (AP-5) (Wang et al. 2007). However, no neurosteroid effects on tonic GABAergic inhibition in CeA have been reported yet.

Here we tested the hypotheses that 1) neurons of the lateral section of the CeA (CeAl) receive tonic GABAergic input mediated by GABAARs containing the δ-subunit and that 2) this tonic inhibition is modulated by the action of neurosteroids.

MATERIALS AND METHODS

All experiments were carried out in accordance with European Communities Council Directive 86/609/EEC. Protocols were approved by the Landesamt für Natur, Umwelt, und Verbraucherschutz Nordrhein-Westfalen (LANUV NRW; AZ 8.87-50.10.46.09.170 and 87-51-04.2010.A322).

Slice preparation.

As previously described (Kamprath et al. 2011; Sosulina et al. 2010), C57BL/6J male mice (8–12 wk old; n = 30) were deeply anesthetized with isoflurane and decapitated. Their brains were rapidly removed and put into oxygenated (95% O2-5% CO2) ice-cold artificial cerebrospinal fluid (ACSF) containing (in mM) 120 NaCl, 2.5 KCl, 1.25 NaH2PO4, 22 NaHCO3, 2 MgSO4, 2 CaCl2, and 10 glucose; pH was 7.4. Coronal slices (250–300 μm thick) were prepared on a vibratome (Leica Microsystems, Wetzlar, Germany) at 4°C. Slices were incubated at 34°C for 20 min and then allowed to equilibrate for at least 1 h at room temperature (RT). After recovery, slices were transferred to a recording chamber continuously superfused with ACSF (∼2 ml/min).

Electrophysiological recordings and data analyses.

Whole cell (current clamp and voltage clamp) recordings were made at 30°C from the CeAl (Fig. 1A, inset) with an EPC-10 single patch-clamp amplifier (HEKA Elektronik, Lambrecht/Pfalz, Germany). Neurons were visualized by infrared video microscopy (camera CF8/1, Kappa, Gleichen, Germany). Cells were discarded from analysis if the resting membrane potential was more positive than −50 mV immediately after breaking through the cell membrane. For current-clamp recordings patch pipettes (3–4 MΩ) were filled with an intracellular solution containing (in mM) 95 K+-gluconate, 20 K3-citrate, 10 NaCl, 10 HEPES, 1 MgCl2, 0.1 CaCl2, 1.1 EGTA, 3 Mg-ATP, and 0.5 Na-GTP; pH 7.2 adjusted with 1 M KOH. For each current-clamp experiment, the amplitude of the voltage deflection (V) evoked by the application of negative current injections (I) (−10 pA, 500 ms) at −65 mV was measured and the input resistance was calculated based on Ohm's law (R = V/I). To determine excitability action potentials were evoked by positive current injections with varying amplitudes and 500-ms duration. The amplitude of the depolarizing step was adjusted (range 50–100 pA) to evoke a minimum of two action potentials in an individual cell. Manual clamping was performed by the injection of direct positive current through the recording pipette in order to depolarize the neuron to the membrane potential prior to THDOC application. For voltage-clamp recordings the pipette solution was (in mM) 110 CsCl, 30 K+-gluconate, 1.1 EGTA, 10 HEPES, 0.1 CaCl2, 4 Mg-ATP, and 0.3 Na-GTP. Neurons were voltage clamped at −70 mV, and input and access resistance were determined by analyzing the steady state and transient upon a 10-ms-long 3-mV pulse applied every 20 s, respectively. Access resistance amounted to ∼4–9.5 MΩ. Cells were discarded from analysis if the access resistance changed by >20% over the course of the experiment. Cell capacitance was measured in voltage-clamp mode with a brief (500 ms) 5-mV biphasic voltage step applied immediately after cell membrane rupture. The mean cell capacitance was 75.51 ± 3.52 pF (n = 71), and no differences in cell capacitance were found between the different experimental groups [1-way ANOVA F(10,60) = 1.563; P = 0.1401; n = 11 different experimental conditions]. To correct changes in tonic currents for differences in cell size (cell capacitance), measured currents were converted to current densities [CD = current (pA)/capacitance (pF) (Marowsky and Vogt 2014; Martin et al. 2014), and ΔCD describes the difference between baseline CD and CD in the presence of drugs]. Correction for liquid junction potential changes (10 mV) was applied. Recordings were filtered at 3 kHz (8-pole Bessel filter), sampled at 10 kHz with a gain of 1.0–1.2 with Patchmaster software (HEKA Elektronik), and analyzed off-line with the programs Mini Analysis (Synaptosoft, Decatur, GA) and Clampfit 10.0 (Molecular Devices, Ismaning, Germany). To isolate the inhibitory currents mediated by GABAARs, recordings were made in the presence of the glutamate receptor blockers 6,7-dinitroquinoxaline-2,3-dione (DNQX) and AP-5 and the GABAB receptor antagonist {[(2S)-3-[(1S)-1-(3,4-dichlorophenyl)ethyl]amino-2-hydroxypropyl] (phenylmethyl)phosphinic acid} (CGP55845A).

Fig. 1.

Fig. 1.

Tonic GABAergic inhibition mediated by δ- and α5-subunit-containing GABA type A receptors (GABAARs) is present in neurons of the lateral section of the central amygdala (CeAl) and is regulated by GABA transport. A: original current trace showing spontaneous inhibitory postsynaptic currents (sIPSCs) in a CeAl neuron. Gabazine (GBZ) increased holding current (Ihold) and blocked sIPSCs. Picrotoxin (PTX) shifted Ihold to more positive values compared with baseline (indicated by dotted line). Inset: infrared DIC image illustrating the recording site. CeAm, medial section of central amygdala; LA, lateral amygdala; BLA, basolateral amygdala; itc, intercalated cell masses. B: original current trace from a CeAl neuron in the continuous presence of tetrodotoxin (TTX). TTX did not prevent the effects of GBZ and PTX. C: original current trace during application of the GABA transporter 1 (GAT-1) blocker NNC-711 (5 μM), followed by PTX (100 μM). Note that Ihold was increased by NNC-711 and this effect was reversed by PTX. D: application of the δ-subunit-specific agonist THIP (10 μM) induced a fast negative shift in Ihold followed by a small desensitization reaching a plateau. Subsequently, PTX reversed the THIP effect and caused a significant shift in Ihold. E: current trace showing the increase in Ihold by application of the GABAAR δ-subunit agonist DS2 (1 μM). Application of PTX shifted Ihold to more positive values exceeding baseline. The shifts induced by DS2 and PTX were significant compared with baseline. F: representative data traces showing a positive shift in Ihold after application of L-655,708 (500 nM, blocking α5-subunit-containing GABAARs) and PTX. Application of PTX abolished sIPSCs and caused an additional significant shift in Ihold. G: summary data for each pharmacological condition; data were normalized to baseline (last 2 min before drug application) of each single experiment. Significance was tested against control (control = 0). n.s., Not significant. *P < 0.05, ***P < 0.001.

Baseline was defined from a time frame of 2 min of stable recording under control conditions. Holding current (Ihold) was determined as the current required to maintain the neuron at a holding potential of −70 mV in the whole cell configuration. The tonic GABA current was calculated and distinguished from IPSCs with a routine written in MATLAB (The MathWorks) in which the mean Ihold was obtained by a Gaussian fit to an all-points histogram from 2 min under steady-state conditions. We quantified responses as the difference in Ihold between baseline and experimental conditions. Frequency, amplitude, and decay time of isolated spontaneous IPSCs (sIPSCs) were analyzed with a detection threshold of 15 pA with Mini Analysis. All events were visually confirmed, and detection threshold was readjusted in case events were not detected with standard settings. All detected events were used for event frequency analysis, but superimposed events were eliminated for amplitude and decay kinetic analysis as reported previously (Herman et al. 2013). We determined means of IPSC characteristics from baseline and experimental drug conditions containing a minimum of 60 events in a time period of 2 min under steady-state conditions. Decay kinetics was calculated with monoexponential curve fittings and is reported as half-width decay time (milliseconds).

Immunohistochemistry.

Immunohistochemical detection of the GABAAR δ-subunit was performed with an anti-GABA δ-subunit antibody (AB9752, Millipore). Staining procedures were similar to those described in a previous study where the specificity of the antibody was confirmed with a GABAAR δ-subunit-knockout mouse (Maguire et al. 2009). In brief, C57BL/6J mice (8–12 wk old; n = 3 mice) were deeply anesthetized with pentobarbital (50 mg/kg) and perfused transcardially with phosphate-buffered saline (PBS; pH 7.4) followed by 4% paraformaldehyde (in 0.15 M Na-phosphate buffer, pH 7.4). Brains were removed and stored in 4% paraformaldehyde overnight at 4°C. After incubation in 30% sucrose-PBS for cryoprotection, coronal sections (30 μm) were cut with a freezing microtome (Frigomobil 1205; Jung, Heidelberg, Germany). Sections were collected in 15% sucrose-PBS, followed by three washes in PBS. After 1-h incubation in 3% H2O2 diluted in methanol, followed by two washes in PBS, sections were blocked for 1 h in 3% bovine serum albumin (BSA), 10% goat serum, and 0.3% Triton in PBS. A guinea pig anti-MAP2 antibody against microtubule-associated protein 2 was obtained from Synaptic Systems (Göttingen, Germany). Antibodies were diluted 1:1,000 in blocking solution. Incubation with agitation was carried out at 4°C overnight. After three wash steps in PBS, sections were transferred to carrier solution (0.3% Triton, 1% BSA, 1% goat serum) and treated with the secondary antibodies goat anti-rabbit conjugated to Alexa Fluor 488 and goat anti-guinea pig conjugated to Alexa Fluor 594 (both diluted 1:1,000, Life Technologies) for 2 h at RT. The fluorescent nuclear stain 4′,6-diamidino-2-phenylindole (DAPI, Sigma-Aldrich; 1:1,000) was added to label cell nuclei. After four additional washes in PBS, sections were mounted with Vectashield HardSet mounting medium (Vector laboratories, Burlingame, CA) and imaged with a laser scanning confocal microscope (Nikon eC1plus) equipped with a CFI75 LWD ×16/0.8 NA objective (Nikon). Sequential three-channel imaging with 405-, 488-, and 543-nm excitation in combination with adequate emission filters (450/30 nm; 515/30 nm, and 605/75 nm, respectively) prevented bleed-through. Images (1,024 × 1,024 pixels) were further processed with ImageJ (http://rsb.info.nih.gov/ij/). Final images represent image mosaics generated with the ImageJ plug-in MosaicJ (Thevenaz and Unser 2007).

Drugs and chemicals.

Drugs were applied by bath perfusion. DNQX (10 μM), AP-5 (50 μM), CGP55845A (1 μM), gabazine (GBZ, 0.5 μM), NNC-711 (5 μM), and picrotoxin (PTX, 100 μM) were from Ascent Scientific, tetrodotoxin (TTX, 1 μM), DS2 (1 μM), finasteride (5 μM), and allopregnanolone (5 μM) were from Tocris, and THDOC (1 and 10 nM) was from Sigma-Aldrich.

Statistical analyses.

Because of the variability in Ihold among different cells and to clearly show the full drug effect on Ihold, we defined the baseline current as 0 for each single experiment and the drug effect was normalized to baseline. We used the one-sample t-test to compare the normalized current and the baseline value 0. A paired two-tailed t-test was used for comparisons made within the same recording with raw data (IPSC parameters), an unpaired two-tailed t-test for comparisons between two different groups, and a one-way ANOVA with Bonferroni's post hoc analysis for comparisons of three or more groups. All statistical analyses were performed with MATLAB, Prism 5.02 (GraphPad Software, San Diego, CA), and Statistica 8 (StatSoft Europe). Data are presented as means ± SE. Statistical significance was accepted if P < 0.05.

RESULTS

Tonic inhibitory transmission is present in CeA neurons.

GABAergic currents were recorded from visually identified neurons located in the CeAl (Fig. 1A, inset). The use of CsCl inside the pipette, however, prevented the characterization of the recorded neurons based on their biophysical properties. The CsCl intracellular solutions has a calculated chloride reversal potential of 0 mV; therefore, GABAAR-mediated sIPSCs appeared as fast inward currents at a holding potential of −70 mV in the presence of NMDA receptor, non-NMDA glutamate receptor, and GABABR antagonists (see materials and methods; Fig. 1A). The frequency of sIPSCs under control conditions was 6.9 ± 0.84 Hz. The GABAAR-mediated tonic conductance was estimated by the shift in the baseline Ihold obtained by blocking GABAARs with selective antagonists. To isolate the contribution of action potential-dependent vesicular release of GABA from neighboring synapses, we used low concentrations of GBZ, a competitive GABAAR antagonist displacing GABA from its binding site (Hamann et al. 1988). A concentration of 0.5 μM has been reported to selectively abolish spontaneous GABAAR-mediated synaptic currents but not tonic currents (Bai et al. 2001; Semyanov et al. 2003). In CeAl neurons, GBZ (0.5 μM) abolished spontaneous synaptic events (Fig. 1A) and produced a small inward shift of the baseline IholdIhold = −2.63 ± 2.3 pA). Normalization to cell size showed no significant change in current density (ΔCD = −0.035 ± 0.03 pA/pF; 1-sample t-test; P = 0.2628 vs. baseline; n = 11; Fig. 1, A and G). To subsequently and completely block GABAAR-mediated conductance in the same neurons, we used the GABAAR channel blocker PTX. PTX is an open channel blocker and acts as a noncompetitive GABAAR antagonist that has equivalent efficacy in blocking low-affinity synaptic and high-affinity extrasynaptic GABAARs (Stell and Mody 2002). Thus application of PTX (100 μM) on top of GBZ produced an outward shift in IholdIhold = 57.19 ± 17.38 pA, ΔCD = 0.9155 ± 0.32 pA/pF; 1-sample t-test, P = 0.0181 vs. baseline; Fig. 1, A and G), reflecting the blockade of a tonic GABAAR-mediated current. To determine whether or not the induced shift in Ihold was due to the generation of action potential-dependent GABA release under basal conditions, we performed the same type of experiment in the presence of TTX (1 μM). Under these conditions, action potential-independent miniature IPSCs (mIPSCs) occurred at a frequency of 3.9 ± 0.4 Hz. Application of GBZ blocked mIPSCs and induced an inward shift of IholdIhold = −4.2 ± 0.63 pA; ΔCD = −0.07108 ± 0.15 pA/pF; 1-sample t-test, P = 0.0031 vs. baseline, n = 7; Fig. 1, B and G). It should be noted that GBZ also increases the chloride conductance (Bai et al. 2001; Scimemi et al. 2005; Semyanov et al. 2004). This inverse antagonist mechanism at low concentrations (0.2–0.5 μM; Wlodarczyk et al. 2013) might contribute to the small inward shift in Ihold seen in parallel with the expected block of spontaneous synaptic events (Fig. 1A). In the same seven neurons, the successive application of PTX shifted Ihold to more positive values (ΔIhold = 48.6 ± 16.8 pA, ΔCD = 0.7877 ± 0.24 pA/pF; 1-sample t-test, P = 0.0173 vs. baseline; Fig. 1, B and G). The fact that the PTX-induced shift in Ihold persisted in the presence of TTX indicates that presynaptic GABA release during spontaneous action potential-dependent network activity is not a major source of ambient GABA mediating the tonic current.

Next, we tested for the effect of an increased ambient GABA concentration. To do so, we applied NNC-711 (5 μM), a specific inhibitor of GABA transporter 1 (GAT-1), to block GABA reuptake (Suzdak et al. 1992). Consistent with previous reports in hippocampus (Holter et al. 2010; Stell and Mody 2002), NNC-711 shifted Ihold to more negative values in six of six cells (ΔIhold = −19.05 ± 5.96 pA, ΔCD = −0.24 ± 0.056 pA/pF; 1-sample t-test, P = 0.0080 vs. baseline; Fig. 1, C and G). The addition of PTX in the continuous presence of NNC-711 shifted Ihold back to baseline (ΔIhold = 0.1 ± 4.4 pA, ΔCD = −0.01 ± 0.05 pA/pF; 1-sample t-test, P = 0.8975 vs. baseline; Fig. 1C). Hence, while the effect of NNC-711 was small, GAT-1-mediated GABA transport reduced tonic GABAergic inhibition.

GABAAR δ-subunit is functionally expressed in CeAl neurons.

Tonic GABAergic currents are usually mediated by extrasynaptic GABAARs containing α4-, α5-, or α6-subunits in combination with β- and δ-subunits (Belelli et al. 2006). Recent immunohistochemical and electrophysiological data demonstrated the expression of α1- and δ-subunits in the CeAm in mice (Herman et al. 2013). Because there is evidence that the expression of the α1-subunit is much lower in CeAl compared with CeAm (Hortnagl et al. 2013) and because the GABAAR δ-subunit is the main target of neurosteroid action, we focused on this subunit.

Application of the δ-subunit-specific agonist THIP hydrochloride (10 μM; Krogsgaard-Larsen et al. 2004) resulted in an inward shift of Ihold in CeAl neurons (ΔIhold = −125.6 ± 23.2 pA, ΔCD = −2.281 ± 0.58 pA/pF; 1-sample t-test, P = 0.0109 vs. baseline, n = 6; Fig. 1, D and G), providing further evidence that GABAARs containing δ-subunits are involved in the tonic current component. When PTX was added, Ihold shifted back to baseline without exceeding it (ΔIhold = 9.72 ± 20.55 pA, ΔCD = −0.046 ± 0.35 pA/pF; 1-sample t-test, P = 0.9 vs. baseline; Fig. 1D).

Furthermore, we tested the effect of the δ-subunit positive allosteric modulator DS2. At a low concentration (1 μM), DS2 shifted Ihold to more negative values (ΔIhold = −13.16 ± 4.7 pA, ΔCD = −0.30 ± 0.096 pA/pF; 1-sample t-test, P = 0.0366 vs. baseline, n = 5; Fig. 1, E and G). The addition of PTX shifted Ihold to more positive values not different from baseline (ΔIhold = 6.4 ± 5.4 pA, ΔCD = 0.2104 ± 0.1391 pA/pF; 1-sample t-test, P = 0.2048 vs. baseline, n = 4; Fig. 1E).

While the GABAAR δ-subunit is the most relevant subunit in terms of neurosteroid action, we also checked for the contribution of the α5-subunit to tonic inhibition in CeAl neurons, using the α5-subunit partial inverse agonist L-655,708 (500 nM) (Prenosil et al. 2006; Quirk et al. 1996). While L-655,708 had no effect in the CeAm (Herman et al. 2013), bath application of this drug induced a small but significant effect on Ihold in CeAl neurons (ΔIhold = 15.2 ± 1.6 pA, ΔCD = 0.3207 ± 0.82 pA/pF; 1-sample t-test, P = 0.030 vs. baseline, n = 4; Fig. 1, F and G). An only minor contribution of the α5-subunit is in line with a rather low expression level of this subunit in the CeA (Marowsky et al. 2012; Pirker et al. 2000; Prut et al. 2010; but see Hortnagl et al. 2013). A summary of the pharmacological effects is depicted in Fig. 1G.

In agreement with the functional data (Fig. 1) and previous studies, immunostaining confirmed the expression of the δ-subunit in the CeA, with a stronger signal in CeAl compared with CeAm (Fig. 2, A–D). Indicating specificity of the staining, strong GABAAR δ-subunit labeling was seen in the dentate gyrus (Fig. 2E; Maguire et al. 2009).

Fig. 2.

Fig. 2.

GABAAR δ-subunit is expressed in CeAl. A: within the amygdala complex, GABAAR δ-subunit immunoreactivity is highest in CeAl. B and C: cytoskeleton protein MAP2 (B) and nuclear stain DAPI (C) were used to identify anatomical structures. D: merged channels. E: hippocampal formation served as a reference structure to test for GABAAR δ-subunit antibody specificity. Dentate gyrus shows strong GABAAR δ-subunit labeling. BA, basolateral amygdala; d, dorsal; m, medial. Scale bars in D and E, 200 μm.

Neurosteroids increase Ihold in CeAl neurons.

The progesterone metabolites THDOC and allopregnanolone induce opening of the GABAA receptor channel at nanomolar concentrations in vitro (Lambert et al. 1995; Majewska 1992), acting preferentially, but not exclusively, on δ-containing subunit assemblies of tonic GABAARs (Brown et al. 2002; Stell et al. 2003). To determine the effect of neurosteroids on GABAergic tonic inhibition, we tested THDOC and allopregnanolone in the presence of ionotropic glutamate receptor antagonists and GABABR inhibition.

Superfusion of THDOC (10 nM) produced a significant negative shift in baseline IholdIhold = −68.76 ± 19.6 pA, ΔCD = −1.4 ± 0.52 pA/pF; 1-sample t-test, P = 0.0295 vs. baseline, n = 8; Fig. 3, A and O). This effect was reversed by the subsequent application of PTX in eight of eight cells recorded (−26.78 ± 20.38 pA, ΔCD = 0.75 ± 0.56 pA/pF; 1-sample t-test, P = 0.2271 vs. baseline; Fig. 3A). Interestingly, the THDOC-induced ΔCD was almost as big as the ΔCD induced by THIP (−2.281 ± 0.58 pA/pF; see Fig. 1G; unpaired t-test, P = 0.1066). At this concentration, THDOC affected sIPSCs by increasing the half-width decay time (control: 14.6 ± 2.0 ms, THDOC: 27.7 ± 5.8 ms; paired t-test, P = 0.0396; Fig. 3, B and C), without changing sIPSC instant frequency (control: 3.1 ± 0.6 Hz, THDOC: 2.4 ± 0.7 Hz; paired t-test, P = 0.1172; Fig. 3D) and sIPSC amplitude (control: 42.7 ± 11.4 pA, THDOC: 38.3 ± 6.3 pA; P = 0.5541; Fig. 3E). Similar effects have been previously observed in hippocampal neurons (Belelli et al. 2006; Belelli and Lambert 2005; Jo et al. 2011). The THDOC-induced changes in Ihold were not prevented by the preceding application of TTX (ΔIhold = −59.5 ± 18.6 pA, ΔCD = 0.9 ± 0.33 pA/pF; 1-sample t-test, P = 0.0412 vs. baseline, n = 6; Fig. 3, F and O), suggesting that the effects of THDOC are not mediated by the generation of action potential-dependent GABA release under basal conditions. This increase was reversed by the successive application of PTX (ΔIhold = 22.1 ± 4.9 pA, ΔCD = 0.5036 ± 0.2127 pA/pF; 1-sample t-test, P = 0.0987 vs. baseline; Fig. 3G). Furthermore, a low dose of THDOC (1 nM), which is considered to be below physiological plasma concentrations (∼8 nM) (Stell et al. 2003), was sufficient to increase IholdIhold = −26.5 ± 4.9 pA, ΔCD = −0.5213 ± 0.94 pA/pF; 1-sample t-test, P = 0.0117, n = 5; Fig. 3, G and O), but to a lesser extent than 10 nM (10 nM THDOC vs. 1 nM THDOC, unpaired t-test, P = 0.0065).

Fig. 3.

Fig. 3.

Neurosteroids modulate tonic GABAergic inhibition in CeAl neurons. A: current trace showing the increase of the tonic current induced by the neurosteroid tetrahydrodeoxycorticosterone (THDOC, 10 nM). The effect was reversed by PTX. B: sIPSCs (mean of 10 traces) recorded at −70 mV in control and after application of THDOC (10 nM) displaying the increase in decay time. C–E: decay time half-width (C), instant frequency (D), and amplitude (E) of sIPSCs before (black) and after (white) THDOC application. While decay time was increased by THDOC, instant frequency and amplitude remained unaffected. F: same as in A but in the continuous presence of TTX. Note that the effect of THDOC on the neuron recorded was a postsynaptic effect since TTX did not prevent the increase in Ihold. G: even at a low concentration of 1 nM, THDOC was effective in enhancing the tonic current. H: current traces showing the increase in the tonic current induced by the neurosteroid allopregnanolone (ALLO, 5 nM). This effect was reversed by the successive application of PTX. I: sIPSCs (mean of 10 traces) recorded at −70 mV in control and after the application of allopregnanolone. J–L: decay time half-width (J), instant frequency (K), and amplitude (L) of sIPSCs. Note that although allopregnanolone increased the decay time this was not significantly different from control (paired t-test, P = 0.6905). M: same as in H but under the blockade of action potentials by TTX. N: representative current traces showing a positive shift in Ihold after application of the 5α-reductase inhibitor finasteride (FIN) and PTX. Application of PTX caused an additional significant shift in Ihold. O: summary of the different pharmacological conditions: effects of THDOC, allopregnanolone, and finasteride on tonic inhibitory currents. Data were normalized to baseline (last 2 min before drug application) of each single experiment, and significance was tested against control (control = 0). *P < 0.05, **P < 0.01.

Next we tested the effect of allopregnanolone on Ihold in CeAl neurons. Bath application of allopregnanolone (5 nM) induced a negative shift in IholdIhold = −22.53 ± 9.1 pA, ΔCD = −0.2942 ± 0.088 pA/pF; 1-sample t-test, P = 0.0202, n = 6; Fig. 3H), without affecting sIPSC half-width decay time (control: 8.8 ± 0.7 ms, allopregnanolone: 13.3 ± 4.7 ms; paired t-test, P = 0.6905; Fig. 3, I and J), sIPSC instant frequency (control: 5.0 ± 1.1 Hz, allopregnanolone: 5.1 ± 0. 5 Hz; paired t-test, P = 0.6250; Fig. 3K), or sIPSC amplitude (control: −46.1 ± 12.0 pA, allopregnanolone: 40.2 ± 10.2 pA; paired t-test, P = 0.0625; Fig. 3L). The effect on Ihold was reversed by the application of PTX, which shifted Ihold back to baseline in all six neurons tested (ΔIhold = 1.633 ± 4.65 pA, ΔCD = 0.4236 ± 0.057 pA/pF; 1-sample t-test, P = 0.4940; Fig. 3H). The blockade of action potentials by TTX did not prevent the allopregnanolone-induced effects (ΔIhold = −42.7 ± 13.3 pA, ΔCD = −0.4567 ± 0.1496 pA/pF; 1-sample t-test, P = 0.0117, n = 5; Fig. 3M). In the same neurons, the successive application of PTX reduced Ihold and reversed the effect produced by allopregnanolone (ΔIhold = 58.78 ± 21.6 pA, ΔCD = 0.67 ± 0.28 pA/pF; 1-sample t-test, P = 0.0960, n = 5; Fig. 3M), suggesting that the allopregnanolone effect is due to the activation of nonsynaptic GABAARs. The amplitude of the allopregnanolone effect might be underestimated because of the rather low concentration used (5 nM). However, it has been shown that high doses of allopregnanolone (300 nM) have a direct effect on chloride channels (Twyman and Macdonald 1992), and such an effect would lead to changes in Ihold.

It is known that the biosynthesis of THDOC and allopregnanolone is limited by the 5α-reductase pathway (Sanna et al. 2004). To test whether or not the inhibition of biosynthesis of neurosteroids would affect tonic GABAergic inhibition, we bath applied finasteride, a specific inhibitor of the 5α-reductase (Finn et al. 2006; Scarduzio et al. 2013). Finasteride application induced a trend toward an increased IPSC decay half-width (control: 14.48 ± 2.33 ms, finasteride: 21 ± 1.84 ms; P = 0.0522; n = 5), which, however, was not significant. In addition, neither the frequency (P = 0.9932) nor the amplitude (P = 0.8133) of the sIPSCs was affected by finasteride. Furthermore, as expected, finasteride (5 μM) was able to partially reduce IholdIhold = 9.6 ± 1.3 pA, ΔCD = 0.1274 ± 0.03 pA/pF; 1-sample t-test, P = 0.0050; time of onset of finasteride effect: 7.57 ± 0.74 min; n = 7; Fig. 3N). The application of PTX on top of finasteride further decreased IholdIhold = 21.47 ± 3.76 pA, ΔCD = 0.29 ± 0.09 pA/pF; 1-sample t-test, P = 0.0159 vs. baseline). These data suggest that neurosteroids are locally synthesized. Figure 3O shows a quantitative summary of the neurosteroid and finasteride effects.

THDOC hyperpolarizes CeAl neurons.

To test the effect of THDOC on the membrane potential and action potential activity, we recorded CeAl neurons under current-clamp conditions. Neurons were held at approximately −65 mV through direct current injection to minimize variability in membrane resting potential and resulting driving forces. Action potentials were evoked by a depolarizing square pulse every 20 s (see materials and methods). THDOC (10 nM) induced a hyperpolarization of the membrane potential resulting in a block of action potentials and reaching steady state after ∼20 min (hyperpolarization from −63.21 ± 0.9 mV to −69.34 ± 1.9 mV; paired t-test, P = 0.0286, n = 4; Fig. 4, A–C). Bringing back the membrane potential to preapplication level through injection of positive current [manual clamping (MC) to −63.08 ± 1.0 mV] during near-maximal effects of THDOC confirmed that the injection of positive current was not able to induce action potentials (Fig. 4, A–C). The apparent input resistance was significantly decreased by THDOC [control: 477.3 ± 35.11 MΩ, THDOC: 237.2 ± 19.84 MΩ, MC: 247.2 ± 19.8 MΩ, post-MC: 267.1 ± 25.6 MΩ; 1-way ANOVA F(3,16) = 21.93, P < 0.0001, n = 4; Fig. 4D]. A Bonferroni's multiple-comparison test revealed differences between control and THDOC (t = 6.969, P < 0.05), MC (t = 6.979, P < 0.05), and post-MC (t = 6.102, P < 0.05), indicating that the membrane conductance was increased by THDOC application leading to a reduction in excitability.

Fig. 4.

Fig. 4.

THDOC effects on membrane potential and action potential firing. A: current-clamp recording from a CeAl neuron showing the hyperpolarizing effect of THDOC (10 nM) when applied at a membrane potential of −65 mV. Upward and downward deflections represent responses to repetitive injection of depolarizing and hyperpolarizing current steps (+80 pA, 500 ms; −10 pA, 500 ms), respectively, as depicted in B. Large positive deflections represent action potentials. Note that THDOC led to a hyperpolarization with a decrement in excitability and decrease in input resistance. MC, manual clamp. B: traces were taken at times indicated by numbers in A. C and D: summary data of membrane potential (C) and input resistance (D) from 4 CeAl neurons recorded in current-clamp mode. One-way ANOVA, Bonferroni's multiple-comparison test, *P < 0.05, ***P < 0.001.

DISCUSSION

In the present study, we show that a tonic GABAergic current mediated by GABAARs containing the δ-subunit is present in neurons of the CeAl and that this current is under the control of the neurosteroids THDOC and allopregnanolone. According to the converging immunohistochemical and pharmacological evidence provided here, the GABAAR δ-subunit contributes substantially to the tonic inhibition in the CeAl. In detail, the δ-subunit was detected at the protein level in the CeA when using a knockout-tested δ-subunit-specific antibody (Maguire et al. 2009). Furthermore, in vitro application of the specific GABAAR δ-subunit agonist (THIP) and the nonselective GABAAR δ-subunit modulator (DS2) significantly evoked a tonic GABAergic current in CeAl neurons.

It is known that two subunit-specific forms of tonic inhibition exist in CeAm neurons in mouse and rat (Herman et al. 2013; Herman and Roberto 2014). A sustained tonic conductance mediated by the GABAAR α1-subunit is present in mouse neurons expressing corticotropin-releasing factor (CRF) receptor-1 (CRF1R). CRF1R-negative neurons, however, possess no intrinsic tonic inhibition but demonstrate a δ-subunit-mediated tonic conductance in the presence of an increased local GABA concentration (Herman et al. 2013). In contrast, we did not find any cell type specificity in the CeAl since all the neurons recorded displayed an ongoing and δ-subunit-mediated tonic GABAergic current. GABAergic tonic inhibition has also been shown in LA and BA, where it is mediated by GABAARs containing the α3-subunit and where the expression of the δ-subunit is rather low (Marowsky et al. 2012). The evidence for δ-subunit-mediated tonic GABAergic inhibition in CeAl neurons is in agreement with the observed strong effect of the neurosteroids THDOC and allopregnanolone. Interestingly, in paracapsular cells of the amygdala, which are also known to express GABAARs containing the δ-subunit, THDOC (at a concentration of 300 nM) affected the tonic current only when applied in combination with 5 μM GABA (Marowsky and Vogt 2014), suggesting that CeAl neurons are a dominant target of neurosteroid action within the amygdala. PTX clearly unmasked an intrinsic tonic current in (e.g., Fig. 1, A and B) as indicated by the induced outward current. However, when PTX was applied successively to drugs that induced a strong increase in the tonic current (e.g., THIP and THDOC; Fig. 1D and Fig. 3A, respectively), PTX still induced a shift in Ihold back to baseline, but without exceeding it. While similar effects have been seen elsewhere (Herman et al. 2013; Jo et al. 2011; Kolbaev et al. 2012), a clear explanation for this phenomenon is missing. One possible explanation is that the activation of δ-subunit-containing GABAARs by drugs like THIP and THDOC directly or indirectly also leads to an activation of a PTX-insensitive non-GABAergic conductance.

Changes in neurosteroid levels (e.g., during the ovarian cycle) and the consequent changes in tonic GABAergic signaling via δ-subunit-containing GABAARs have been linked to an altered seizure susceptibility and to anxiety (Maguire et al. 2005). Similarly, a reduction of endogenous THDOC and allopregnanolone levels has been found to lower the potency of GABA in eliciting GABAAR-mediated IPSCs (Barbaccia 2004; Dong et al. 2001). In the present study, THDOC hyperpolarized CeAl neurons, thereby decreasing their excitability. The CeAl neurons are mutually interconnected via GABAergic synapses, and they connect to the CeAm. An enhancement of tonic GABAergic inhibition in the CeAl might thus result in either an inhibitory or a disinhibitory overall influence on CeAm output neurons. In addition, modulatory effects of THDOC on tonic GABAergic currents are context dependent. For the hypothalamic-pituitary-adrenal (HPA) axis it has been shown that THDOC potentiates the inhibitory action of GABA on CRF neurons under control conditions, whereas THDOC activates the HPA axis after stress (Sarkar et al. 2011). This change in the direction of the THDOC effect on network activity is due to a stress-induced collapse of the chloride gradient in CRF neurons that results in an excitatory action of GABA (Sarkar et al. 2011).

Behaviorally, a reduction of endogenous THDOC and allopregnanolone levels is related to anxiety-like responses, increased aggression, and a reduced sensitivity to the loss of righting reflex induced by GABAAR agonists or positive modulators, but also to an enhanced expression of fear conditioning responses in rodents. Conversely, enhanced neurosteroid action results in anxiolysis, sedation/hypnosis, and anticonvulsant and anesthetic action (Barbaccia 2004). Furthermore, increases in contextual fear can be abolished by treatment with drugs that normalize the corticolimbic levels of allopregnanolone (Agis-Balboa et al. 2009). Therefore the enhancement of tonic inhibition through neurosteroid action in the CeAl is likely to reduce activity in CeAm neurons mediating fear responses.

Inhibition of the synthesis of neurosteroids with finasteride, a specific inhibitor of the 5α-reductase (Finn et al. 2006; Scarduzio et al. 2013), led to a decrease in the tonic GABAergic current in CeAl neurons. Although the bath application of finasteride does not reveal the cellular source of the neurosteroids, the observed effect indicates that there is an ongoing local neurosteroid synthesis. While we cannot fully exclude nonspecific effects of finasteride on other ligands or conductances, additional evidence for the local synthesis is provided by the demonstration of immunoreactivity of neurosteroids (Saalmann et al. 2007; Tokuda et al. 2010) and 5α-reductase type II expression (Castelli et al. 2013) in neurons of the amygdala. Interestingly, mice with enhanced contextual fear responses and impaired fear extinction show decreased 5α-reductase type I mRNA expression and allopregnanolone levels in the amygdala and other brain regions (Pibiri et al. 2008).

The combination of rapid local neurosteroid synthesis and GABA potentiation makes the neurosteroid system a target for local modulation of GABAergic inhibition, thereby contributing to synaptic plasticity. In line with this, finasteride has an effect on fear learning (Disney and Calford 2001). In addition, midazolam (a benzodiazepine acting on GABAARs) enhances a specific form of network inhibition and inhibits long-term potentiation and contextual fear learning, an effect that is blocked when finasteride is administrated 1 day before midazolam (Tokuda et al. 2010). The present study provides for the first time evidence that finasteride is able to decrease tonic inhibition in CeA neurons.

GABAARs containing the α5-subunit contribute to tonic inhibition only when the ambient GABA concentration increases but are not activated by endogenous ambient GABA (Scimemi et al. 2005). The majority of CeAl neurons are GABAergic cells (Ehrlich et al. 2009; Haubensak et al. 2010), which could contribute to an enhanced ambient GABA concentration. A higher ambient GABA concentration in CeAl versus CeAm would explain the small α5-subunit-mediated tonic current in CeAl, which is completely absent in CeAm (Herman et al. 2013). Furthermore, Li et al. (2013) suggested that excitatory synapses onto somatostatin-positive neurons in CeAl are strengthened while the somatostatin-negative neurons are weakened after fear conditioning. Because the somatostatin-positive neurons are forming intra-CeAl GABAergic synapses, an increased activity of these neurons might result in higher extracellular GABA levels. Thus, while activation of δ-subunit-containing GABAARs predominates at low ambient GABA concentrations, α5 subunit-containing GABAARs would be activated after fear conditioning.

It has been suggested that synaptic spillover after vesicular release (Attwell et al. 1993), reversal of GABA transporters (Gaspary et al. 1998), and a GABA “leak” through bestrophin channels (Lee et al. 2010) are sources of ambient GABA. As shown recently, reactive astrocytes seem to produce and release GABA through bestrophin channels under pathological conditions (Jo et al. 2014). Active uptake maintains GABA at sufficiently low concentrations to prevent tonic GABABR activation (Isaacson et al. 1993). The enhanced tonic GABAergic current seen in CeAl neurons after the application of NNC-711, a selective inhibitor of GABA uptake by GAT-1, confirms the presence of active GABA uptake but excludes reverse GABA transport as a source of extracellular GABA under these conditions. It should be noted that the effect of NNC-711 was relatively small, which might be due to a relatively high ambient GABA level in the CeAl (see above).

In conclusion, here we provide evidence that CeAl neurons contain GABAARs that mediate inhibitory tonic currents that are likely attributable to the expression of the GABAAR δ-subunit. Furthermore, we show that these tonic currents are modulated by neurosteroids and play an important role in regulating the excitability of CeAl neurons. As a consequence, the tonic GABAergic current and the neurosteroid-mediated regulation may be potential targets for a pharmacological manipulation of the imbalance in the amygdala output underlying anxiety disorders.

GRANTS

This work was supported by the German Research Foundation (DFG; SFB-TR58, TPA03 to H.-C. Pape).

DISCLOSURES

No conflicts of interest, financial or otherwise, are declared by the author(s).

AUTHOR CONTRIBUTIONS

Author contributions: H.R.-P. and H.-C.P. conception and design of research; H.R.-P., P.B., and L.S. performed experiments; H.R.-P., P.B., and L.S. analyzed data; H.R.-P., P.B., L.S., and H.-C.P. interpreted results of experiments; H.R.-P. and P.B. prepared figures; H.R.-P., P.B., and H.-C.P. drafted manuscript; H.R.-P., P.B., L.S., and H.-C.P. edited and revised manuscript; H.R.-P., P.B., L.S., and H.-C.P. approved final version of manuscript.

ACKNOWLEDGMENTS

We thank A. Klinge, K. Foraita, E. Nass, and S. Kiesling for expert technical assistance, P. Berenbrock for excellent animal care, and Dr. K. Jüngling for helpful discussion and comments on the manuscript. Thanks are due to Dr. T. Kanyshkova for the generation of preliminary data.

Present address of L. Sosulina: Neuronal Networks Group, Deutsches Zentrum für Neurodegenerative Erkrankungen, Ludwig-Erhard-Allee 2, 53175 Bonn, Germany.

REFERENCES

  1. Agis-Balboa RC, Pibiri F, Nelson M, Pinna G. Enhanced fear responses in mice treated with anabolic androgenic steroids. Neuroreport 20: 617–621, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Akwa Y, Purdy RH, Koob GF, Britton KT. The amygdala mediates the anxiolytic-like effect of the neurosteroid allopregnanolone in rat. Behav Brain Res 106: 119–125, 1999. [DOI] [PubMed] [Google Scholar]
  3. Attwell D, Barbour B, Szatkowski M. Nonvesicular release of neurotransmitter. Neuron 11: 401–407, 1993. [DOI] [PubMed] [Google Scholar]
  4. 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 gamma-aminobutyric acidA receptors in hippocampal neurons. Mol Pharmacol 59: 814–824, 2001. [DOI] [PubMed] [Google Scholar]
  5. Barbaccia ML. Neurosteroidogenesis: relevance to neurosteroid actions in brain and modulation by psychotropic drugs. Crit Rev Neurobiol 16: 67–74, 2004. [DOI] [PubMed] [Google Scholar]
  6. Barnard EA, Skolnick P, Olsen RW, Mohler H, Sieghart W, Biggio G, Braestrup C, Bateson AN, Langer SZ. International Union of Pharmacology. XV. Subtypes of gamma-aminobutyric acidA receptors: classification on the basis of subunit structure and receptor function. Pharmacol Rev 50: 291–313, 1998. [PubMed] [Google Scholar]
  7. Belelli D, Harrison NL, Maguire J, Macdonald RL, Walker MC, Cope DW. Extrasynaptic GABAA receptors: form, pharmacology, and function. J Neurosci 29: 12757–12763, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Belelli D, Herd MB, Mitchell EA, Peden DR, Vardy AW, Gentet L, Lambert JJ. Neuroactive steroids and inhibitory neurotransmission: mechanisms of action and physiological relevance. Neuroscience 138: 821–829, 2006. [DOI] [PubMed] [Google Scholar]
  9. Belelli D, Lambert JJ. Neurosteroids: endogenous regulators of the GABAA receptor. Nat Rev Neurosci 6: 565–575, 2005. [DOI] [PubMed] [Google Scholar]
  10. Brickley SG, Mody I. Extrasynaptic GABAA receptors: their function in the CNS and implications for disease. Neuron 73: 23–34, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Brown N, Kerby J, Bonnert TP, Whiting PJ, Wafford KA. Pharmacological characterization of a novel cell line expressing human alpha4beta3delta GABAA receptors. Br J Pharmacol 136: 965–974, 2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Castelli MP, Casti A, Casu A, Frau R, Bortolato M, Spiga S, Ennas MG. Regional distribution of 5alpha-reductase type 2 in the adult rat brain: an immunohistochemical analysis. Psychoneuroendocrinology 38: 281–293, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Cheney DL, Uzunov D, Costa E, Guidotti A. Gas chromatographic-mass fragmentographic quantitation of 3alpha-hydroxy-5alpha-pregnan-20-one (allopregnanolone) and its precursors in blood and brain of adrenalectomized and castrated rats. J Neurosci 15: 4641–4650, 1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Disney A, Calford MB. Neurosteroids mediate habituation and tonic inhibition in the auditory midbrain. J Neurophysiol 86: 1052–1056, 2001. [DOI] [PubMed] [Google Scholar]
  15. Dong E, Matsumoto K, Uzunova V, Sugaya I, Takahata H, Nomura H, Watanabe H, Costa E, Guidotti A. Brain 5alpha-dihydroprogesterone and allopregnanolone synthesis in a mouse model of protracted social isolation. Proc Natl Acad Sci USA 98: 2849–2854, 2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Ehrlich I, Humeau Y, Grenier F, Ciocchi S, Herry C, Luthi A. Amygdala inhibitory circuits and the control of fear memory. Neuron 62: 757–771, 2009. [DOI] [PubMed] [Google Scholar]
  17. Farrant M, Nusser Z. Variations on an inhibitory theme: phasic and tonic activation of GABAA receptors. Nat Rev Neurosci 6: 215–229, 2005. [DOI] [PubMed] [Google Scholar]
  18. Finn DA, Beadles-Bohling AS, Beckley EH, Ford MM, Gililland KR, Gorin-Meyer RE, Wiren KM. A new look at the 5alpha-reductase inhibitor finasteride. CNS Drug Rev 12: 53–76, 2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Gaspary HL, Wang W, Richerson GB. Carrier-mediated GABA release activates GABA receptors on hippocampal neurons. J Neurophysiol 80: 270–281, 1998. [DOI] [PubMed] [Google Scholar]
  20. Glykys J, Mody I. The main source of ambient GABA responsible for tonic inhibition in the mouse hippocampus. J Physiol 582: 1163–1178, 2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Hamann M, Desarmenien M, Desaulles E, Bader MF, Feltz P. Quantitative evaluation of the properties of a pyridazinyl GABA derivative (SR 95531) as a GABAA competitive antagonist. An electrophysiological approach. Brain Res 442: 287–296, 1988. [DOI] [PubMed] [Google Scholar]
  22. Haubensak W, Kunwar PS, Cai H, Ciocchi S, Wall NR, Ponnusamy R, Biag J, Dong HW, Deisseroth K, Callaway EM, Fanselow MS, Luthi A, Anderson DJ. Genetic dissection of an amygdala microcircuit that gates conditioned fear. Nature 468: 270–276, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Herman MA, Contet C, Justice NJ, Vale W, Roberto M. Novel subunit-specific tonic GABA currents and differential effects of ethanol in the central amygdala of CRF receptor-1 reporter mice. J Neurosci 33: 3284–3298, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Herman MA, Roberto M. Cell-type-specific tonic GABA signaling in the rat central amygdala is selectively altered by acute and chronic ethanol. Addict Biol (August 29, 2014). doi: 10.1111/adb.12181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Holter NI, Zylla MM, Zuber N, Bruehl C, Draguhn A. Tonic GABAergic control of mouse dentate granule cells during postnatal development. Eur J Neurosci 32: 1300–1309, 2010. [DOI] [PubMed] [Google Scholar]
  26. Hortnagl 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 236: 345–372, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Isaacson JS, Solis JM, Nicoll RA. Local and diffuse synaptic actions of GABA in the hippocampus. Neuron 10: 165–175, 1993. [DOI] [PubMed] [Google Scholar]
  28. Jo JY, Jeong JA, Pandit S, Stern JE, Lee SK, Ryu PD, Lee SY, Han SK, Cho CH, Kim HW, Jeon BH, Park JB. Neurosteroid modulation of benzodiazepine-sensitive GABAA tonic inhibition in supraoptic magnocellular neurons. Am J Physiol Regul Integr Comp Physiol 300: R1578–R1587, 2011. [DOI] [PubMed] [Google Scholar]
  29. Jo S, Yarishkin O, Hwang YJ, Chun YE, Park M, Woo DH, Bae JY, Kim T, Lee J, Chun H, Park HJ, Lee DY, Hong J, Kim HY, Oh SJ, Park SJ, Lee H, Yoon BE, Kim Y, Jeong Y, Shim I, Bae YC, Cho J, Kowall NW, Ryu H, Hwang E, Kim D, Lee CJ. GABA from reactive astrocytes impairs memory in mouse models of Alzheimer's disease. Nat Med 20: 886–896, 2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Kamprath K, Romo-Parra H, Haring M, Gaburro S, Doengi M, Lutz B, Pape HC. Short-term adaptation of conditioned fear responses through endocannabinoid signaling in the central amygdala. Neuropsychopharmacology 36: 652–663, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Karavolas HJ, Hodges DR. Metabolism of progesterone and related steroids by neural and neuroendocrine structures. In: Neurosteroids and Brain Functions, edited by Costa E, Paul SM. New York: Theime, 1991, p. 135–145. [Google Scholar]
  32. Khisti RT, Boyd KN, Kumar S, Morrow AL. Systemic ethanol administration elevates deoxycorticosterone levels and chronic ethanol exposure attenuates this response. Brain Res 1049: 104–111, 2005. [DOI] [PubMed] [Google Scholar]
  33. Kolbaev SN, Sharopov S, Dierkes PW, Luhmann HJ, Kilb W. Phasic GABAA-receptor activation is required to suppress epileptiform activity in the CA3 region of the immature rat hippocampus. Epilepsia 53: 888–896, 2012. [DOI] [PubMed] [Google Scholar]
  34. Krogsgaard-Larsen P, Frolund 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 68: 1573–1580, 2004. [DOI] [PubMed] [Google Scholar]
  35. Kumar S, Sieghart W, Morrow AL. Association of protein kinase C with GABAA receptors containing alpha1 and alpha4 subunits in the cerebral cortex: selective effects of chronic ethanol consumption. J Neurochem 82: 110–117, 2002. [DOI] [PubMed] [Google Scholar]
  36. Lambert JJ, Belelli D, Hill-Venning C, Peters JA. Neurosteroids and GABAA receptor function. Trends Pharmacol Sci 16: 295–303, 1995. [DOI] [PubMed] [Google Scholar]
  37. Lee S, Yoon BE, Berglund K, Oh SJ, Park H, Shin HS, Augustine GJ, Lee CJ. Channel-mediated tonic GABA release from glia. Science 330: 790–796, 2010. [DOI] [PubMed] [Google Scholar]
  38. Li H, Penzo MA, Taniguchi H, Kopec CD, Huang ZJ, Li B. Experience-dependent modification of a central amygdala fear circuit. Nat Neurosci 16: 332–339, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Maguire J, Ferando I, Simonsen C, Mody I. Excitability changes related to GABAA receptor plasticity during pregnancy. J Neurosci 29: 9592–9601, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. 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 8: 797–804, 2005. [DOI] [PubMed] [Google Scholar]
  41. Majewska MD. Neurosteroids: endogenous bimodal modulators of the GABAA receptor. Mechanism of action and physiological significance. Prog Neurobiol 38: 379–395, 1992. [DOI] [PubMed] [Google Scholar]
  42. Majewska MD, Falkay G, Baulieu EE. Modulation of uterine GABAA receptors during gestation and by tetrahydroprogesterone. Eur J Pharmacol 174: 43–47, 1989. [DOI] [PubMed] [Google Scholar]
  43. Marowsky A, Rudolph U, Fritschy JM, Arand M. Tonic inhibition in principal cells of the amygdala: a central role for alpha3 subunit-containing GABAA receptors. J Neurosci 32: 8611–8619, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Marowsky A, Vogt KE. Delta-subunit-containing GABAA-receptors mediate tonic inhibition in paracapsular cells of the mouse amygdala. Front Neural Circuits 8: 27, 2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Martin BS, Corbin JG, Huntsman MM. Deficient tonic GABAergic conductance and synaptic balance in the fragile X syndrome amygdala. J Neurophysiol 112: 890–902, 2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Mody I. Aspects of the homeostatic plasticity of GABAA receptor-mediated inhibition. J Physiol 562: 37–46, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Neugebauer V, Li W, Bird GC, Han JS. The amygdala and persistent pain. Neuroscientist 10: 221–234, 2004. [DOI] [PubMed] [Google Scholar]
  48. Olsen RW, Sieghart W. GABAA receptors: subtypes provide diversity of function and pharmacology. Neuropharmacology 56: 141–148, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Pape HC, Pare D. Plastic synaptic networks of the amygdala for the acquisition, expression, and extinction of conditioned fear. Physiol Rev 90: 419–463, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Parsons RG, Ressler KJ. Implications of memory modulation for post-traumatic stress and fear disorders. Nat Neurosci 16: 146–153, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Paul SM, Purdy RH. Neuroactive steroids. FASEB J 6: 2311–2322, 1992. [PubMed] [Google Scholar]
  52. Pibiri F, Nelson M, Guidotti A, Costa E, Pinna G. Decreased corticolimbic allopregnanolone expression during social isolation enhances contextual fear: a model relevant for posttraumatic stress disorder. Proc Natl Acad Sci USA 105: 5567–5572, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Pirker S, Schwarzer C, Wieselthaler A, Sieghart W, Sperk G. GABAA receptors: immunocytochemical distribution of 13 subunits in the adult rat brain. Neuroscience 101: 815–850, 2000. [DOI] [PubMed] [Google Scholar]
  54. Prenosil GA, Schneider Gasser EM, Rudolph U, Keist R, Fritschy JM, Vogt KE. Specific subtypes of GABAA receptors mediate phasic and tonic forms of inhibition in hippocampal pyramidal neurons. J Neurophysiol 96: 846–857, 2006. [DOI] [PubMed] [Google Scholar]
  55. Prut L, Prenosil G, Willadt S, Vogt K, Fritschy JM, Crestani F. A reduction in hippocampal GABAA receptor alpha5 subunits disrupts the memory for location of objects in mice. Genes Brain Behav 9: 478–488, 2010. [DOI] [PubMed] [Google Scholar]
  56. Purdy RH, Morrow AL, Moore PH Jr, Paul SM. Stress-induced elevations of gamma-aminobutyric acid type A receptor-active steroids in the rat brain. Proc Natl Acad Sci USA 88: 4553–4557, 1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Quirk K, Blurton P, Fletcher S, Leeson P, Tang F, Mellilo D, Ragan CI, McKernan RM. [3H]L-655,708, a novel ligand selective for the benzodiazepine site of GABAA receptors which contain the alpha5 subunit. Neuropharmacology 35: 1331–1335, 1996. [DOI] [PubMed] [Google Scholar]
  58. Rupprecht R. The neuropsychopharmacological potential of neuroactive steroids. J Psychiatr Res 31: 297–314, 1997. [DOI] [PubMed] [Google Scholar]
  59. Rupprecht R, Holsboer F. Neuroactive steroids: mechanisms of action and neuropsychopharmacological perspectives. Trends Neurosci 22: 410–416, 1999. [DOI] [PubMed] [Google Scholar]
  60. Saalmann YB, Kirkcaldie MT, Waldron S, Calford MB. Cellular distribution of the GABAA receptor-modulating 3alpha-hydroxy, 5alpha-reduced pregnane steroids in the adult rat brain. J Neuroendocrinol 19: 272–284, 2007. [DOI] [PubMed] [Google Scholar]
  61. Sanna E, Talani G, Busonero F, Pisu MG, Purdy RH, Serra M, Biggio G. Brain steroidogenesis mediates ethanol modulation of GABAA receptor activity in rat hippocampus. J Neurosci 24: 6521–6530, 2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Sarkar J, Wakefield S, MacKenzie G, Moss SJ, Maguire J. Neurosteroidogenesis is required for the physiological response to stress: role of neurosteroid-sensitive GABAA receptors. J Neurosci 31: 18198–18210, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Scarduzio M, Panichi R, Pettorossi VE, Grassi S. Synaptic long-term potentiation and depression in the rat medial vestibular nuclei depend on neural activation of estrogenic and androgenic signals. PLoS One 8: e80792, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Scimemi A, Semyanov A, Sperk G, Kullmann DM, Walker MC. Multiple and plastic receptors mediate tonic GABAA receptor currents in the hippocampus. J Neurosci 25: 10016–10024, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Semyanov A, Walker MC, Kullmann DM. GABA uptake regulates cortical excitability via cell type-specific tonic inhibition. Nat Neurosci 6: 484–490, 2003. [DOI] [PubMed] [Google Scholar]
  66. Semyanov A, Walker MC, Kullmann DM, Silver RA. Tonically active GABA A receptors: modulating gain and maintaining the tone. Trends Neurosci 27: 262–269, 2004. [DOI] [PubMed] [Google Scholar]
  67. Sieghart W, Sperk G. Subunit composition, distribution and function of GABAA receptor subtypes. Curr Top Med Chem 2: 795–816, 2002. [DOI] [PubMed] [Google Scholar]
  68. Sosulina L, Graebenitz S, Pape HC. GABAergic interneurons in the mouse lateral amygdala: a classification study. J Neurophysiol 104: 617–626, 2010. [DOI] [PubMed] [Google Scholar]
  69. Stell BM, Brickley SG, Tang CY, Farrant M, Mody I. Neuroactive steroids reduce neuronal excitability by selectively enhancing tonic inhibition mediated by delta subunit-containing GABAA receptors. Proc Natl Acad Sci USA 100: 14439–14444, 2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Stell BM, Mody I. Receptors with different affinities mediate phasic and tonic GABAA conductances in hippocampal neurons. J Neurosci 22: RC223, 2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Suzdak PD, Frederiksen K, Andersen KE, Sorensen PO, Knutsen LJ, Nielsen EB. NNC-711, a novel potent and selective gamma-aminobutyric acid uptake inhibitor: pharmacological characterization. Eur J Pharmacol 224: 189–198, 1992. [DOI] [PubMed] [Google Scholar]
  72. Tasan RO, Bukovac A, Peterschmitt YN, Sartori SB, Landgraf R, Singewald N, Sperk G. Altered GABA transmission in a mouse model of increased trait anxiety. Neuroscience 183: 71–80, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Thevenaz P, Unser M. User-friendly semiautomated assembly of accurate image mosaics in microscopy. Microsc Res Tech 70: 135–146, 2007. [DOI] [PubMed] [Google Scholar]
  74. Tokuda K, O'Dell KA, Izumi Y, Zorumski CF. Midazolam inhibits hippocampal long-term potentiation and learning through dual central and peripheral benzodiazepine receptor activation and neurosteroidogenesis. J Neurosci 30: 16788–16795, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Twyman RE, Macdonald RL. Neurosteroid regulation of GABAA receptor single-channel kinetic properties of mouse spinal cord neurons in culture. J Physiol 456: 215–245, 1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Wang C, Marx CE, Morrow AL, Wilson WA, Moore SD. Neurosteroid modulation of GABAergic neurotransmission in the central amygdala: a role for NMDA receptors. Neurosci Lett 415: 118–123, 2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Whiting PJ. The GABA-A receptor gene family: new targets for therapeutic intervention. Neurochem Int 34: 387–390, 1999. [DOI] [PubMed] [Google Scholar]
  78. Wlodarczyk AI, Sylantyev S, Herd MB, Kersante F, Lambert JJ, Rusakov DA, Linthorst AC, Semyanov A, Belelli D, Pavlov I, Walker MC. GABA-independent GABAA receptor openings maintain tonic currents. J Neurosci 33: 3905–3914, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Journal of Neurophysiology are provided here courtesy of American Physiological Society

RESOURCES