Abstract
Many synapses, including parallel fiber synapses in the cerebellum, express presynaptic GABAA receptors. However, reports of the functional consequences of presynaptic GABAA receptor activation are variable across synapses, from inhibition to enhancement of transmitter release. We find that presynaptic GABAA receptor function is bidirectional at parallel fiber synapses depending on GABA concentration and modulation of GABAA receptors in mice. Activation of GABAA receptors by low GABA concentrations enhances glutamate release, whereas activation of receptors by higher GABA concentrations inhibits release. Furthermore, blocking GABAB receptors reduces GABAA receptor currents and shifts presynaptic responses toward greater enhancement of release across a wide range of GABA concentrations. Conversely, enhancing GABAA receptor currents with ethanol or neurosteroids shifts responses toward greater inhibition of release. The ability of presynaptic GABAA receptors to enhance or inhibit transmitter release at the same synapse depending on activity level provides a new mechanism for fine control of synaptic transmission by GABA and may explain conflicting reports of presynaptic GABAA receptor function across synapses.
NEW & NOTEWORTHY GABAA receptors are widely expressed at presynaptic terminals in the central nervous system. However, previous reports have produced conflicting results on the function of these receptors at different synapses. We show that presynaptic GABAA receptor function is strongly dependent on the level of receptor activation. Low levels of receptor activation enhance transmitter release, whereas higher levels of activation inhibit release at the same synapses. This provides a novel mechanism by which presynaptic GABAA receptors fine-tune synaptic transmission.
Keywords: cerebellum, GABAA, parallel fiber, presynaptic
INTRODUCTION
Neurotransmitter receptors expressed in the presynaptic membrane have proven to be widespread and important modulators of synaptic transmission. Activation of presynaptic GABAA receptors (GABAARs) is one of the earliest observed forms of synaptic modulation, beginning with primary afferent depolarization in the spinal cord (Andersen et al. 1964; Eccles et al. 1963). In recent years there has been a renewed interest in these receptors as they continue to be found at an expanding range of synapses throughout the central nervous system (Soiza-Reilly et al. 2013; Wang et al. 2019; Weisz et al. 2016; Zorrilla de San Martin et al. 2017). However, there continue to be widely divergent reports on the primary function of presynaptic GABAARs. While it is generally agreed that activation of these receptors depolarizes the presynaptic terminal (Price and Trussell 2006; Pugh and Jahr 2011; Ruiz et al. 2003; Zhang and Jackson 1995), studies are split on whether this depolarization reduces transmitter release (Eccles et al. 1963; Nicoll and Alger 1979; Wang et al. 2019; Zhang and Jackson 1995) or enhances transmitter release (Pouzat and Marty 1999; Pugh and Jahr 2011; Ruiz et al. 2010; Stell et al. 2007; Turecek and Trussell 2002; Zorrilla de San Martin et al. 2017). Currently, it is not clear why GABAARs inhibit transmitter release at some synapses and increase release at others. GABAAR function is regulated by a wide diversity of modulators, including GABAB receptors (Connelly et al. 2013; Tao et al. 2013), ethanol (Palmer and Hoffer 1990; Valenzuela and Jotty 2015), neurosteroids (Majewska et al. 1986), barbiturates (Nicoll et al. 1975), and benzodiazepines (Costa and Guidotti 1979), raising the possibility that modulation of presynaptic GABAARs may influence whether they enhance or inhibit synaptic transmission.
Parallel fiber synapses in the cerebellum are perhaps the most heavily studied site of presynaptic GABAAR function in recent years (Antflick and Hampson 2012; Berglund et al. 2016; Dellal et al. 2012; Howell and Pugh 2016; Pugh and Jahr 2011, 2013; Santhakumar et al. 2013; Stell 2011; Stell et al. 2007). These studies have resulted in broad agreement that presynaptic GABAARs enhance glutamate release at this synapse. However, even at this synapse there have been indications of inhibitory actions of GABAARs, particularly during periods of very high or prolonged GABA application (Berglund et al. 2016; Dellal et al. 2012; Stell 2011). This suggests that presynaptic GABAARs may have differing effects on transmitter release depending on the properties of the synapse or degree of activation (Trigo et al. 2008).
To understand how GABAARs inhibit transmitter release under some conditions but enhance release in others, we systematically investigated presynaptic GABAAR function at parallel fiber synapses in the cerebellum over a range of physiological conditions. We found that the direction of GABAAR effects on transmission is strongly dependent on GABA concentration and modulation of GABAARs. Contrary to previous studies, we found that transmitter release was enhanced over only a very narrow range of relatively low GABA concentrations, whereas higher (but likely still physiological) GABA concentrations inhibited release. We then tested whether modulators of presynaptic GABAARs could shift the balance of enhancement vs. inhibition of transmitter release. We found that increasing GABAAR activity through application of ethanol and the neurosteroid allopregnanolone shifts GABAAR responses toward greater inhibition of transmitter release across almost all GABA concentrations tested. In addition, GABABR antagonists reduce GABAAR activity, resulting in greater enhancement of transmitter release across a wide range of GABA concentrations.
METHODS
Animals.
All experimental procedures involving animals were approved by the Institutional Animal Care and Use Committee at the University of Texas Health Science Center San Antonio. Male and female C57BL/6 mice (Charles River, Wilmington, MA) 14–30 days old were used for all experiments. Animals were kept on a 12:12-h light-dark cycle with ad libitum access to food and water.
Slice preparation.
Male and female mice were deeply anesthetized with isoflurane, and the cerebellum was rapidly dissected and placed in ice-cold oxygenated artificial cerebrospinal fluid (ACSF) containing the following (in mM): 119 NaCl, 26.2 NaHCO3, 2.5 KCl, 1.0 NaH2PO4, 11 glucose, 2 CaCl2, and 1.3 MgCl2. Parasagittal slices (300 µm) were cut from the vermis of the cerebellum using a vibratome (Leica Biosystems, Buffalo Grove, IL) and then incubated at 34°C for 30 min before being transferred to the microscope slice chamber for recording.
Electrophysiology experiments.
During recording, slices were superfused with ACSF at a flow rate of ~2 ml/min. Electrophysiological currents were recorded with a Multiclamp 700B amplifier (Molecular Devices, Sunnyvale, CA), filtered at 5 kHz, and digitized at 50 kHz. Data were collected using pCLAMP software (Molecular Devices). For GABA uncaging experiments, 10 ml of ACSF containing 60 μM RuBi-GABA (Tocris, Bristol, UK) were recirculated using a fluid pump (Cole-Parmer, Vernon Hills, IL). This concentration of RuBi-GABA (60 μM) had no effect on the membrane resistance (P = 0.75, n = 4) or the amplitude of excitatory synaptic responses (P = 0.59, n = 4) in cerebellar stellate cells. GABA was uncaged by a brief (5 ms) illumination from a 473-nm laser light source (PSU-III-LED; Opto Engine, Midvale, UT) over a range of intensities (0.28–315 µW) at 30-s inter-sweep intervals to allow for clearance of GABA between sweeps. The uncaging laser intensity was measured using a photometer placed under the objective of the microscope. The theoretical maximum GABA concentration produced by photolytic uncaging is 60 μM (the concentration of RuBi-GABA in the bath solution). The uncaging laser intensity was adjusted over an ~1,000-fold range, suggesting the uncaged GABA concentrations varied roughly from tens or hundreds of nanomolar to tens of micromolar. Where indicated, ACSF contained one or more of the following: 3 µM CGP-55845 (CGP; Abcam, Cambridge, MA); 200 µM picrotoxin (PTX; Abcam); 1 μM tertiapin-Q (TPQ; Tocris); and 20 mM ethanol (Fisher, Hampton, NH).
For experiments on somatodendritic GABA currents, whole cell patch-clamp recordings were made from granule cells, with the following internal solution (in mM): 135 CsCl, 10 HEPES, 4 MgCl2, 5 EGTA, 4 Na-ATP, 0.5 Na-GTP, and 2 QX-314. The pH was adjusted to 7.2–7.4 using CsOH, and the osmolarity was 280–300 mosmol/l.
For experiments on presynaptic GABAA receptor function, putative stellate cells were identified in the outer third of the molecular layer and whole cell patch-clamp recordings were made with the following internal solution (in mM): 137 K-gluconate, 2 KCl, 4 MgCl2, 10 HEPES, 5 EGTA, 4 Na-ATP, and 0.5 Na-GTP. The pH was adjusted to 7.2–7.4 using KOH and, the osmolarity was 280–300 mosmol/l. The postsynaptic membrane potential was held near the chloride reversal potential to minimize postsynaptic GABA currents following GABA uncaging. However, small, long-lasting GABAAR-mediated currents were sometimes still observed immediately following the laser pulse as a slight change in the baseline holding current. Excitatory postsynaptic currents (EPSCs) were evoked by stimulation of parallel fibers with two pulses (100 μs, 10–60 V) 20 ms apart through a patch pipette filled with ACSF. GABA was uncaged by a laser pulse 50 ms before EPSCs were evoked, to activate presynaptic GABAARs as described previously (Howell and Pugh 2016). Uncaging and control sweeps were interleaved. Asynchronous events were identified by eye. Only synaptic events occurring more than 5 ms after the onset of the second stimulus artifact (in a pair of stimuli delivered at 50Hz) were counted as asynchronous events.
Fiber volleys from parallel fibers were measured using a standard patch-clamp electrode filled with ACSF inserted into the molecular layer of transverse cerebellar slices. Parallel fibers were stimulated by a second patch electrode filled with ACSF inserted in the molecular layer ~150–250 μm from the recording electrode. On alternating sweeps, GABA was uncaged 50 ms before parallel fiber stimulation by a 5-ms laser pulse centered on either the recording or stimulating electrode (results were similar at both uncaging locations). The stimulus intensity was kept relatively low to avoid saturating the fiber volley response. At the beginning of each experiment we confirmed that increasing the stimulus intensity increased the fiber volley amplitude.
Stimulus artifacts have been digitally removed in all figures.
Data analysis and statistical analysis.
Data were analyzed in IgorPro (Wavemetrics, Lake Oswego, OR) using the Neuromatic toolkit (Rothman and Silver 2018) and custom macros. Statistical significance was determined using two-tailed paired Student’s t-tests in Excel (Microsoft, Redmond, WA) and one- and two-way repeated-measures ANOVAs using SigmaStat 4.0 (Systat Software, San Jose, CA). Statistical values of P ≤ 0.05 were considered significant. Data are means ± SE.
RESULTS
To investigate presynaptic GABAAR function, we made whole cell patch-clamp recordings from stellate cells in acute cerebellar slices and evoked EPSCs by electrical stimulation of parallel fibers. On alternating sweeps, RuBi-GABA was uncaged by a brief (5 ms) laser flash in the molecular layer 50 ms before parallel fiber stimulation to activate presynaptic GABAARs (Fig. 1A). We have shown previously that this protocol activates GABAARs with minimal GABABR activation at the time of synaptic stimulation (Howell and Pugh 2016). To measure GABAAR effects on transmission across a range of activity levels, we systematically modulated the laser power used to uncage RuBi-GABA from 0.28 to 315 μW. We used a relatively low RuBi-GABA concentration (60 μM) in the bath solution to mimic the likely range of GABA concentrations resulting from spillover transmission (Barbour 2001; Barbour and Häusser 1997; Dzubay and Jahr 1999; Rusakov and Kullmann 1998). We first tested the effectiveness of modulating the laser power to control the uncaged GABA concentration by recording from cerebellar granule cells and focusing the laser on the soma. We found that the amplitude of the uncaging current varied directly with the power of the uncaging laser (Fig. 1B, inset). From these data we constructed a dose-response curve that is well fit by the Hill equation, indicating modulation of the laser intensity effectively alters the uncaged GABA concentration over the range of laser intensities used (0.28–850 μW).
Fig. 1.
Activation of presynaptic GABAA receptors has bidirectional effects on synaptic currents. A: diagram of GABA uncaging (laser) and parallel fiber stimulation (stim) protocol. Representative synaptic responses in control (black) and following GABA uncaging (red) are shown. B: dose-response relationship of GABAA receptor-mediated currents (I) at the granule cell body and GABA uncaging laser intensity. Inset: example traces of GABA uncaging-evoked currents at each laser intensity. C: representative average traces of excitatory postsynaptic current (EPSCs) from control sweeps (black) and GABA uncaging sweeps (red) for each uncaging laser power. D: change in evoked EPSC amplitudes recorded in stellate cells following GABA uncaging at each laser intensity compared with interleaved control sweeps without uncaging in control artificial cerebrospinal fluid (circles; n = 8–12) or in the presence of the GABAA receptor antagonist picrotoxin (PTX; triangles; n = 7). All traces are the average of 10 sweeps. *P ≤ 0.05.
Contrary to previous studies that showed parallel fiber GABAARs increase transmission (Pugh and Jahr 2011; Stell et al. 2007), we found that GABAARs increase EPSC amplitudes over only a very narrow range of GABA concentrations. Specifically, we observed enhancement of EPSCs at a laser power of 14.2 μW (148.1 ± 16.4% of control, n = 11, P = 0.004; Fig. 1, C and D), and all other laser powers tested either produced no change (power < 14.2 μW, n = 9–12) or decreased EPSC amplitudes (power > 14.2 μW, n = 8–11; for laser intensity of 315 μW: 56.1 ± 9.6% of control, n = 9, P = 0.005; Fig. 1, C and D). As expected from previous studies (Pugh and Jahr 2011; Stell et al. 2007), the enhancement of EPSCs at 14.2 μW was blocked by bath application of PTX (14.2 μW: 91.7 ± 3.3% of control, P = 0.05, n = 7). However, the inhibition of EPSCs observed at higher laser intensities was also blocked by PTX (315 μW: 100.2 ± 7.8% of control, P = 0.63, n = 7 for all laser intensities; Fig. 1D), indicating that this inhibition is also mediated by GABAARs and not by GABABRs or activation of G protein-coupled inwardly rectifying potassium (GIRK) channels. Interestingly, the uncaging laser power that produced an enhancement of EPSCs (14.2 μW) produced a current in granule cells only approximately one-third the maximal current (Fig. 1B), suggesting presynaptic GABAARs may only enhance transmitter release when partially activated. Collectively, these data suggest that presynaptic GABAARs may frequently inhibit release depending on the GABA concentration reaching the receptors.
Although it has been shown previously that presynaptic GABAARs increase EPSC amplitudes by increasing vesicle release probability (Pugh and Jahr 2011; Stell et al. 2007), it is not clear whether the inhibition of EPSCs seen at higher uncaging laser intensities is due to pre- or postsynaptic mechanisms. The increase of EPSC amplitudes following GABA uncaging at 14.2 μW power was associated with a decrease in the paired-pulse ratio (3.2 ± 0.4 vs. 2.1 ± 0.2, n = 10, P = 0.02; Fig. 2, A and B) and failure rate (0.16 ± 0.07 vs. 0.09 ± 0.04, n = 9, P = 0.07), consistent with an increase in vesicle release probability. At higher uncaging laser intensities (34–167μW), the paired-pulse ratio was significantly increased (P = 0.03; Fig. 2, A and B) following GABA uncaging, and there was a trend toward increased failure rate (P = 0.17), suggesting a decrease in vesicle release. However, this does not rule out additional inhibition of EPSCs due to postsynaptic mechanisms. To test this we measured the amplitude of asynchronous EPSCs that result from fusion of single vesicles following the synchronous EPSCs (Atluri and Regehr 1998; Diamond and Jahr 1995). We found no change in asynchronous EPSC amplitudes between control and GABA uncaging sweeps (33–167 μW: 58.9 ± 3.4 vs. 53.7 ± 3.3 pA, n = 39, P = 0.12), even though the synchronous EPSC amplitude was reduced (Figs. 1D and 2C). The lack of change in asynchronous EPSCs suggests the reduction in the synchronous EPSC is not due to postsynaptic shunting through GABAARs or other postsynaptic mechanisms. Together, these data indicate that both the enhancement of EPSCs at low uncaging powers and the inhibition of EPSCs at higher uncaging powers are due to changes in the presynaptic vesicle release.
Fig. 2.
Mechanism of presynaptic GABAA receptor inhibition of synaptic currents. A: representative traces of excitatory postsynaptic currents (EPSCs) from control sweeps (black) or GABA uncaging sweeps (red) using an uncaging laser power of 3.7 (top), 14 (middle), or 67 μW (bottom). GABA uncaging EPSCs have been normalized to the peak of the first control EPSC to show the change in paired-pulse ratio (PPR). B: average change in PPR (PPR GABA/PPR control) following GABA uncaging at low (3.7 μW), mid (14 μW), and high (34–167 μW) laser intensities. C: representative traces (10 overlaid sweeps) showing asynchronous synaptic events in control (black) and following GABA uncaging (red). D: average number (left) and amplitude (right) of asynchronous events during control or GABA uncaging sweeps for uncaging laser powers from 67 to 315 μW. E: representative traces of fiber volleys from control sweeps (black) or GABA uncaging sweeps (red). F: change in evoked EPSC amplitudes following GABA uncaging at each laser intensity in the presence of the G protein-coupled inwardly rectifying potassium (GIRK) channel blocker tertiapin Q (n = 6–7). The data from standard artificial cerebrospinal fluid are replotted from Fig. 1 (gray) for comparison. All traces are the average of 10 sweeps. *P ≤ 0.05; **P ≤ 0.01. asynch, Asynchronous; cnt, control.
Previous studies have found that presynaptic GABAARs at parallel fiber synapses enhance transmitter release by depolarizing the presynaptic terminal and activating voltage-gated calcium channels (Pugh and Jahr 2011; Stell et al. 2007). What is the mechanism producing inhibition of EPSCs at higher uncaging laser intensities? Is it also dependent on presynaptic depolarization? To address this question, we analyzed the frequency of asynchronous synaptic events in control and GABA uncaging sweeps. The frequency of asynchronous synaptic events is directly related to subthreshold depolarization (Christie et al. 2011) and bulk calcium (Atluri and Regehr 1998; Sun et al. 2007) in the presynaptic bouton. We found that the frequency of asynchronous events was increased following GABA uncaging compared with control sweeps at high uncaging laser powers (67–315 μW; P = 0.002, n = 28; Fig. 2, C and D), consistent with depolarization of parallel fiber boutons and elevated bulk calcium following GABA uncaging. This raises the question of how the synchronous EPSC amplitude is reduced but the number of asynchronous events is increased. A simple change in vesicular release probability generally modulates synchronous and asynchronous release together (Atluri and Regehr 1998; Christie et al. 2011; Howell and Pugh 2016). One possibility is that the synchronous EPSC is reduced by a decrease in parallel fiber excitability (possibly due to sodium channel inactivation) such that fewer axons are brought to threshold with each stimulus. To test this, we measured the fiber volley amplitude following stimulation of parallel fibers in the molecular layer during control and GABA uncaging sweeps. We observed a slight increase in fiber volley amplitude following GABA uncaging at high laser power (167 μW: 106.8 ± 3.2% of control, n = 7, P = 0.005; Fig. 2E), consistent with earlier results (Dellal et al. 2012). This suggests more, not fewer, parallel fibers are activated with each stimulus in the presence of GABA, ruling out reduced parallel fiber excitability as an explanation for the reduced synchronous EPSC amplitude. Instead, our data are more consistent with an excessive depolarization of the presynaptic bouton and inactivation of voltage-gated calcium (Catterall et al. 2013; Forsythe et al. 1998; Xu and Wu 2005) or sodium channels (Leão et al. 2005; Ohura and Kamiya 2018; Zhang and Jackson 1995). The observation that fiber volley amplitude is not reduced following GABA uncaging suggests that action potential initiation and propagation are normal; however, it remains possible that sodium channel inactivation reduces action potential amplitude (Jang et al. 2005; Trigo et al. 2008; Zhang and Jackson 1995). Inactivation of either calcium or sodium channels due to subthreshold depolarization could result in increased global calcium but reduced action potential-evoked calcium influx, a mechanism that would explain the decrease in the synchronous EPSC but increase in the frequency of asynchronous EPSCs.
We then tested if excessive depolarization of parallel fibers results in inhibition of release by artificially depolarizing the resting membrane potential in parallel fibers. We choose to do this by blocking GIRK channels with TPQ, which, unlike block of voltage-gated potassium channels, is expected to depolarize the axons with little direct effect on the action potential waveform. Previous work has shown that activation of GABAB or adenosine receptors reduces the excitability of parallel fibers (Dittman and Regehr 1996), suggesting GIRK channels are expressed in parallel fibers and can modulate their excitability. In the presence of TPQ, the frequency of spontaneous synaptic events was increased (control: 0.21 ± 0.04 Hz, n = 12; TPQ: 0.52 ± 0.19 Hz, n = 7, P = 0.05), consistent with a depolarization of parallel fibers by TPQ. With GIRK channels blocked we observed greater inhibition of EPSCs (34–167 μW: 49.5 ± 8.5% of control, P < 0.001) following GABA uncaging and did not observe enhancement of EPSCs at any laser power (Fig. 2F; data from Fig. 1D plotted for comparison). The increased inhibition observed in TPQ is consistent with excessive depolarization reducing transmitter release, possibly through inactivation of calcium channels.
These data suggest that presynaptic GABAARs may inhibit or enhance transmitter release depending on the magnitude of the GABAAR-mediated depolarization of the presynaptic bouton. This raises the possibility that positive or negative modulators of GABAARs can also modulate GABAAR effects on transmitter release. Previous studies have shown that GABABR activity enhances currents through extrasynaptic (likely δ-subunit containing) GABAARs in cerebellar granule cells and other cell types that exhibit a tonic GABAAR current (Connelly et al. 2013; Tao et al. 2013); however, the mechanism of this modulation is not currently known. Although the exact subunit expression of GABAARs expressed at parallel fibers is not known, they may also include δ-subunits and can be broadly classified as extrasynaptic receptors (Berglund et al. 2016; Dellal et al. 2012; Santhakumar et al. 2013). Furthermore, our previous studies found that blocking GABABRs reduces presynaptic GABAAR-mediated increase in transmitter release (Howell and Pugh 2016), suggesting parallel fiber GABAARs may also be enhanced by GABABR activity. Consistent with these earlier studies, in the presence of the GABABR antagonist CGP, we did not observe enhancement of EPSCs at an uncaging laser intensity of 14.2 μW as we did in control conditions (Fig. 3, A and C). Rather, the uncaging dose-response curve was shifted to the right, resulting in enhancement of EPSC amplitudes over a wide range of higher uncaging laser intensities (67–167 μW: 141.2 ± 12.4% of control, n = 37, P = 0.004; Fig. 3A; data from Fig. 1D plotted for comparison). Bath application of CGP alone did not alter the amplitude of control EPSC amplitudes (100.4 ± 15.8 vs. 114.7 ± 20.5 pA, n = 15, P = 0.2) or paired-pulse ratio (2.53 ± 0.28 vs. 2.29 ± 0.19, n = 15, P = 0.32). The range of laser intensities producing enhancement of EPSCs is likely broadened in CGP by saturation of GABAARs or saturation of RuBi-GABA uncaging (Fig. 1B). These data suggest that blocking GABABRs reduces the presynaptic GABAAR current and shifts the response to GABA uncaging toward enhancement of EPSCs.
Fig. 3.

Modulation of presynaptic GABAA receptor function. A: change in evoked excitatory postsynaptic current (EPSC) amplitudes recorded in stellate cells following GABA uncaging at each laser intensity in the presence of CGP-55845 (+CPG; black; n = 7–11). B: change in evoked EPSC amplitudes recorded in stellate cells following GABA uncaging at each laser intensity in the presence of ethanol (+EtOH; blue; n = 6) or allopregnanolone (+Allo; green; n = 4–6). The data from standard artificial cerebrospinal fluid are replotted from Fig. 1 (gray) in A and B for comparison. C: representative EPSCs recorded in control sweeps (black) or following GABA uncaging with laser power of 14.25 μW (red) for each pharmacological condition. All traces are the average of 10 sweeps.
Previous studies have shown that extrasynaptic GABAARs are particularly sensitive to enhancement by ethanol (Diaz and Valenzuela 2016; Hanchar et al. 2005; Santhakumar et al. 2007; but see Borghese and Harris 2007) or neurosteroids (Abramian et al. 2014; Stell et al. 2003). We therefore used ethanol and allopregnanolone to test how increasing presynaptic GABAAR activity alters effects on synaptic transmission. In the presence of 20 mM ethanol, GABA uncaging inhibited EPSCs at all uncaging laser powers except the lowest power tested (3.7–315 μW: n = 6, P = 0.009–0.057), shifting the dose-response curve to the left (Fig. 3B; data from Fig. 1D plotted for comparison). Ethanol alone did not increase fiber volley amplitudes (n = 6, P = 0.53), indicating the shift toward greater inhibition is not due to depolarization of parallel fibers by ethanol itself. Similarly, application of allopregnanolone resulted in significant inhibition of EPSC amplitudes following GABA uncaging across almost all laser powers tested (14.2–315 μW: n = 6, P = 0.0002–0.05; 167 μW: P = 0.36; Fig. 3B). We did not observe enhancement of EPSC amplitudes at any laser power in the presence of ethanol or allopregnanolone. This is likely due to the narrow range of GABA concentrations that produce enhancement and the relatively large (~10-fold) gaps in laser intensities tested at the low range (0.28 vs. 3.7 vs. 14.25 μW). These results demonstrate that the effects of GABAARs on transmitter release are highly sensitive to GABAAR modulators, shifting responses to primarily enhance or inhibit release across a range of GABA concentrations.
DISCUSSION
In this study we show that presynaptic GABAARs can enhance or inhibit transmitter release depending on the GABA concentration. In vivo, these receptors are likely exposed to GABA concentrations ranging from hundreds of nanomolar to tens of micromolar from either ambient GABA (Berglund et al. 2016; Bright et al. 2011; Santhakumar et al. 2006) or spillover transmission from neighboring inhibitory synapses (Barbour 2001; Barbour and Häusser 1997; Dzubay and Jahr 1999; Rusakov and Kullmann 1998). The GABA concentrations tested in this study likely span a similar range (see methods), suggesting both enhancement and inhibition of transmitter release could occur over physiological GABA concentrations.
Mechanisms of GABA-mediated enhancement and inhibition of vesicle release.
Previous studies have shown that activation of presynaptic GABAARs in parallel fibers depolarizes the bouton (Dellal et al. 2012; Stell 2011; Stell et al. 2007), activating voltage-gated calcium channels (Pugh and Jahr 2011) and enhancing vesicle release. In the present study we find that higher concentrations of GABA can also inhibit vesicle release. We argue that the switch from enhancement to inhibition is dependent on the amplitude of the GABAAR-mediated depolarization and inactivation of voltage-gated calcium and/or sodium channels (see Fig. 4) based on the following evidence: 1) The uncaging laser power producing enhancement of EPSCs in control solutions (14.25 μW) results in a GABA current only approximately one-third of the maximal current in granule cell body, suggesting higher laser powers can produce substantially larger GABAAR currents and depolarization in parallel fiber boutons. 2) The number of asynchronous synaptic events was significantly increased at high uncaging laser powers (Fig. 2, C and D), but not at lower powers (14.25 μW; 1.03 vs. 1.22 events/sweep, P = 0.12, n = 12), suggesting greater depolarization and bulk calcium in the boutons at uncaging laser intensities that produce inhibition of release. 3) Tonic depolarization of parallel fibers by blockade of GIRK channels results in greater inhibition of vesicle release (Fig. 2F). One possible caveat to this mechanism is that it requires depolarization and/or elevated calcium in the bouton for at least 50 ms (the interval between GABA uncaging and synaptic stimulation). In our experimental system this is not an issue because the GABA concentration likely remains elevated in the slice for at least 50 ms, maintaining GABAAR activation and presynaptic depolarization. This is evident in the relatively slow-decaying GABAAR-mediated currents in Fig. 1B. In a previous study we found that enhancement of EPSCs by GABA was the same with the use of a 20- or 50-ms interval between uncaging and synaptic stimulation (Howell and Pugh 2016), suggesting a prolonged depolarization is not necessary.
Fig. 4.
Model of presynaptic GABAA receptor (GABAAR) action at parallel fiber synapses. Proposed action of presynaptic GABAARs when exposed to low (left) or high (right) GABA concentrations is shown. GABAARs depolarize the presynaptic terminal, activating voltage-gated calcium channels, which allow calcium influx (Ca2+) and potentiation of action potential (AP)-evoked transmitter release. When the GABAAR-mediated current is increased (through a higher GABA concentration or exposure to positive modulators), the bouton experiences greater depolarization and calcium influx, resulting in reduced AP-evoked transmitter release, likely due to inactivation of calcium or sodium channels. Blocking GABABRs with CGP-55845 (+CGP) decreases the current through GABAARs, resulting in little or no GABAAR effect at low GABA and enhancement of AP-evoked transmitter release at high GABA. +EtOH/Allo, presence of ethanol or allopregnanolone; ΔVm, change in membrane potential.
Modulation by GABAB receptors.
We find that blocking GABAB receptors shifts the effects of presynaptic GABAA receptors such that we observed enhancement of release across a wide range of higher GABA concentrations. Several mechanisms could explain this shift in GABAA receptor behavior. Previous studies have shown that GABAB receptor activity enhances currents through extrasynaptic (likely δ-subunit containing) GABAA receptors in cerebellar granule cells and other neurons that exhibit a tonic GABAA receptor current (Connelly et al. 2013; Tao et al. 2013). The mechanism of this enhancement is currently not understood but likely involves activation of PKA and phosphorylation of extrasynaptic GABAA receptors (Connelly et al. 2013). If this is also true at parallel fiber boutons, then inhibition of GABAB receptors would reduce the GABAA receptor-mediated currents at each GABA concentration, resulting in smaller depolarizations of the bouton and favoring enhancement of release over inhibition. Another possibility is that blocking GABABRs shifts the chloride reversal potential (ECl). Wright et al. (2017) recently showed that activation of GABABRs in hippocampal neurons reduces the expression of the potassium-chloride cotransporter KCC2, which is largely responsible for extruding chloride from the cell. This raises the possibility that blockade of GABABRs at parallel fiber boutons increases KCC2 expression (and chloride extrusion from the bouton), resulting in a hyperpolarizing shift in ECl. This mechanism would also limit depolarization of the presynaptic bouton by GABAAR activation and, hence, favor enhancement of release over inhibition. However, this would not explain the results using allopregnanolone or ethanol, because these compounds have been shown to have little or no effect on ECl (Rodgers-Neame et al. 1992; Shimura et al. 1996; Soldo et al. 1994; Ye et al. 2001). Either of these potential mechanisms would be consistent with our hypothesis that small GABAA receptor-mediated depolarizations of the presynaptic membrane result in enhancement of release, whereas larger depolarizations result in inhibition. Alternatively, it is well established that GABABRs activate GIRK channels and inhibit voltage-gated calcium channels (Mintz and Bean 1993; Misgeld et al. 1995; Slesinger et al. 1997), either of which could possibly produce the shift in GABA enhancement/inhibition of EPSCs when GABABRs are blocked. We do not favor either of these potential mechanisms, because blocking GABABRs should reduce GIRK channel activity, depolarizing the boutons and shifting responses to GABA toward greater inhibition, as shown in Fig. 2F, not greater enhancement of EPSCs as seen in the presence of CGP (Fig. 3A). Likewise, blocking GABABRs should increase the activity of voltage-gated calcium channels, which is expected to shift enhancement of EPSCs toward lower GABA concentrations, not higher concentrations as seen in Fig. 3A. Regardless of the precise mechanism involved, these results demonstrate that activation of GABABRs can have a powerful influence on how GABAARs effect synaptic transmission.
Synapse specificity of GABAA receptor modulation.
These data suggest that presynaptic GABAAR function may vary widely across synapses depending on the GABA concentration and modulation of receptors. For example, synapses which coexpress GABAA and GABABRs may be more prone to GABAAR-mediated inhibition of release compared with synapses with GABAARs alone. In addition, modulators of GABAAR function, such as neurosteroids, may be present at some synapses and not others, or change over time. For example, changes in progesterone-derived neuroactive steroids during the menstrual cycle and pregnancy are linked to changes in tonic GABAAR currents (Maguire et al. 2005; Stell et al. 2003). However, our results suggest that changes in these steroids may also shift or reverse the effects of presynaptic GABAAR on vesicle release. These factors suggest that presynaptic GABAAR function is very synapse and time specific, explaining the conflicting results of presynaptic GABAAR function across synapses. Furthermore, modulation of presynaptic GABAARs by ethanol may play a role in the behavioral effects of alcohol consumption (Santhakumar et al. 2013), and modulation by GABABRs could contribute to the effects of GABABR agonists used clinically.
At parallel fiber synapses, previous reports have suggested that presynaptic GABAARs enhance transmitter release (Dellal et al. 2012; Pugh and Jahr 2011; Stell et al. 2007). However, in the present work we find that GABAARs inhibit release over the majority of GABA concentrations tested, with only a very narrow range of GABA concentrations enhancing release (Fig. 1D). This discrepancy is likely due to the use of GABABR antagonists or GABAAR-specific agonists in previous studies to isolate GABAAR effects. In the absence of GABABR activity, GABAARs are shifted toward much greater enhancement of release. Our results suggest that under physiological conditions (GABABRs unblocked), the functional consequence of presynaptic GABAAR function may be inhibition of release more often than enhancement of release. This result highlights the need for careful attention to experimental conditions when studying the function of presynaptic GABAARs. It will be interesting in the future to examine whether other synapses with presynaptic GABAARs that are thought to enhance transmitter release, such as mossy fiber boutons (Ruiz et al. 2010), Shaffer collaterals (Jang et al. 2006), and calyx of Held (Turecek and Trussell 2002), also show inhibition of release when GABABRs are not blocked.
Physiological consequences.
Many theories of cerebellar function hypothesize that pauses in Purkinje cell firing are a critical signal for downstream plasticity (Pugh and Raman 2006) and behavioral output (Jirenhed and Hesslow 2016; Medina et al. 2000). Pauses in Purkinje cell activity are produced by both reduced excitatory parallel fiber input (Ito 1993; Steuber et al. 2007) and increased inhibition from molecular layer interneurons (Barmack and Yakhnitsa 2008; Mathews et al. 2012). Presynaptic GABAARs are well positioned to contribute to pauses in Purkinje cell firing. During periods of low interneuron activity, spillover of GABA from single synapses or ambient GABA (Berglund et al. 2016) may enhance glutamate release from parallel fibers, but during periods of elevated inhibition, spillover and/or pooling of GABA from multiple inhibitory synapses may instead inhibit glutamate release through increased activation of GABAARs. The net effect of this process is that periods of high inhibition are complimented with decreased glutamate release from parallel fibers, resulting in a more robust pause in Purkinje cell firing.
Parallel fiber synapses also express presynaptic GABABRs, which inhibit transmitter release by reducing the activity of voltage-gated calcium channels (Dittman and Regehr 1996 1997). We have shown that presynaptic GABABR expression is downregulated at parallel fiber-stellate cell synapses following 4-Hz parallel fiber stimulation (Orts-Del’Immagine and Pugh 2018). This change not only removes GABABR-mediated inhibition of transmitter release but also removes enhancement of GABAAR currents, shifting presynaptic GABAAR function toward increasing transmitter release across a wide range of GABA concentrations (Fig. 3B). The result of these complimentary actions will be significantly greater synaptic transmission at parallel fibers exposed to 4-Hz activity. This is a possible mechanism by which information flow through different parallel fiber pathways and susceptibility to long-term synaptic plasticity could be regulated.
GRANTS
This work was funded by National Institutes of Health Grant R01 NS092809.
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared by the authors.
AUTHOR CONTRIBUTIONS
J.R.P. conceived and designed research; S.N.K., W.-C.W., and Y.Y. performed experiments; S.N.K., W.-C.W., Y.Y., and J.R.P. analyzed data; S.N.K. and J.R.P. interpreted results of experiments; J.R.P. prepared figures; J.R.P. drafted manuscript; S.N.K., W.-C.W., and J.R.P. edited and revised manuscript; S.N.K., W.-C.W., Y.Y., and J.R.P. approved final version of manuscript.
ACKNOWLEDGMENTS
We are grateful to Michael Beckstead, Tabby Kreko-Pierce, and Sriity Melley Sadanandan for helpful discussion and comments on the manuscript.
REFERENCES
- Abramian AM, Comenencia-Ortiz E, Modgil A, Vien TN, Nakamura Y, Moore YE, Maguire JL, Terunuma M, Davies PA, Moss SJ. Neurosteroids promote phosphorylation and membrane insertion of extrasynaptic GABAA receptors. Proc Natl Acad Sci USA 111: 7132–7137, 2014. doi: 10.1073/pnas.1403285111 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Andersen P, Eccles JC, Sears TA. Cortically evoked depolarization of primary afferent fibers in the spinal cord. J Neurophysiol 27: 63–77, 1964. doi: 10.1152/jn.1964.27.1.63 . [DOI] [PubMed] [Google Scholar]
- Antflick JE, Hampson DR. Modulation of glutamate release from parallel fibers by mGlu4 and pre-synaptic GABAA receptors. J Neurochem 120: 552–563, 2012. doi: 10.1111/j.1471-4159.2011.07611.x . [DOI] [PubMed] [Google Scholar]
- Atluri PP, Regehr WG. Delayed release of neurotransmitter from cerebellar granule cells. J Neurosci 18: 8214–8227, 1998. doi: 10.1523/JNEUROSCI.18-20-08214.1998 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barbour B. An evaluation of synapse independence. J Neurosci 21: 7969–7984, 2001. doi: 10.1523/JNEUROSCI.21-20-07969.2001 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barbour B, Häusser M. Intersynaptic diffusion of neurotransmitter. Trends Neurosci 20: 377–384, 1997. doi: 10.1016/S0166-2236(96)20050-5 . [DOI] [PubMed] [Google Scholar]
- Barmack NH, Yakhnitsa V. Functions of interneurons in mouse cerebellum. J Neurosci 28: 1140–1152, 2008. doi: 10.1523/JNEUROSCI.3942-07.2008 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Berglund K, Wen L, Dunbar RL, Feng G, Augustine GJ. Optogenetic visualization of presynaptic tonic inhibition of cerebellar parallel fibers. J Neurosci 36: 5709–5723, 2016. doi: 10.1523/JNEUROSCI.4366-15.2016 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Borghese CM, Harris RA. Studies of ethanol actions on recombinant δ-containing γ-aminobutyric acid type A receptors yield contradictory results. Alcohol 41: 155–162, 2007. doi: 10.1016/j.alcohol.2007.03.006 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bright DP, Renzi M, Bartram J, McGee TP, MacKenzie G, Hosie AM, Farrant M, Brickley SG. Profound desensitization by ambient GABA limits activation of δ-containing GABAA receptors during spillover. J Neurosci 31: 753–763, 2011. doi: 10.1523/JNEUROSCI.2996-10.2011 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Catterall WA, Leal K, Nanou E. Calcium channels and short-term synaptic plasticity. J Biol Chem 288: 10742–10749, 2013. doi: 10.1074/jbc.R112.411645 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Christie JM, Chiu DN, Jahr CE. Ca2+-dependent enhancement of release by subthreshold somatic depolarization. Nat Neurosci 14: 62–68, 2011. doi: 10.1038/nn.2718 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Connelly WM, Fyson SJ, Errington AC, McCafferty CP, Cope DW, Di Giovanni G, Crunelli V. GABAB receptors regulate extrasynaptic GABAA receptors. J Neurosci 33: 3780–3785, 2013. doi: 10.1523/JNEUROSCI.4989-12.2013 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Costa E, Guidotti A. Molecular mechanisms in the receptor action of benzodiazepines. Annu Rev Pharmacol Toxicol 19: 531–545, 1979. doi: 10.1146/annurev.pa.19.040179.002531. [DOI] [PubMed] [Google Scholar]
- Dellal SS, Luo R, Otis TS. GABAA receptors increase excitability and conduction velocity of cerebellar parallel fiber axons. J Neurophysiol 107: 2958–2970, 2012. doi: 10.1152/jn.01028.2011 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Diamond JS, Jahr CE. Asynchronous release of synaptic vesicles determines the time course of the AMPA receptor-mediated EPSC. Neuron 15: 1097–1107, 1995. doi: 10.1016/0896-6273(95)90098-5 . [DOI] [PubMed] [Google Scholar]
- Diaz MR, Valenzuela CF. Sensitivity of GABAergic tonic currents to acute ethanol in cerebellar granule neurons is not age- or δ subunit-dependent in developing rats. Alcohol Clin Exp Res 40: 83–92, 2016. doi: 10.1111/acer.12940 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dittman JS, Regehr WG. Contributions of calcium-dependent and calcium-independent mechanisms to presynaptic inhibition at a cerebellar synapse. J Neurosci 16: 1623–1633, 1996. doi: 10.1523/JNEUROSCI.16-05-01623.1996 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dittman JS, Regehr WG. Mechanism and kinetics of heterosynaptic depression at a cerebellar synapse. J Neurosci 17: 9048–9059, 1997. doi: 10.1523/JNEUROSCI.17-23-09048.1997 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dzubay JA, Jahr CE. The concentration of synaptically released glutamate outside of the climbing fiber-Purkinje cell synaptic cleft. J Neurosci 19: 5265–5274, 1999. doi: 10.1523/JNEUROSCI.19-13-05265.1999 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eccles JC, Schmidt R, Willis WD. Pharmacological studies on presynaptic inhibition. J Physiol 168: 500–530, 1963. doi: 10.1113/jphysiol.1963.sp007205 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Forsythe ID, Tsujimoto T, Barnes-Davies M, Cuttle MF, Takahashi T. Inactivation of presynaptic calcium current contributes to synaptic depression at a fast central synapse. Neuron 20: 797–807, 1998. doi: 10.1016/S0896-6273(00)81017-X . [DOI] [PubMed] [Google Scholar]
- Hanchar HJ, Dodson PD, Olsen RW, Otis TS, Wallner M. Alcohol-induced motor impairment caused by increased extrasynaptic GABAA receptor activity. Nat Neurosci 8: 339–345, 2005. doi: 10.1038/nn1398 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Howell RD, Pugh JR. Biphasic modulation of parallel fibre synaptic transmission by co-activation of presynaptic GABAA and GABAB receptors in mice. J Physiol 594: 3651–3666, 2016. doi: 10.1113/JP272124 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ito M. Synaptic plasticity in the cerebellar cortex and its role in motor learning. Can J Neurol Sci 20, Suppl 3: S70–S74, 1993. doi: 10.1017/S0317167100048551 . [DOI] [PubMed] [Google Scholar]
- Jang IS, Ito Y, Akaike N. Feed-forward facilitation of glutamate release by presynaptic GABAA receptors. Neuroscience 135: 737–748, 2005. doi: 10.1016/j.neuroscience.2005.06.030. [DOI] [PubMed] [Google Scholar]
- Jang IS, Nakamura M, Ito Y, Akaike N. Presynaptic GABAA receptors facilitate spontaneous glutamate release from presynaptic terminals on mechanically dissociated rat CA3 pyramidal neurons. Neuroscience 138: 25–35, 2006. doi: 10.1016/j.neuroscience.2005.11.001 . [DOI] [PubMed] [Google Scholar]
- Jirenhed DA, Hesslow G. Are Purkinje cell pauses drivers of classically conditioned blink responses? Cerebellum 15: 526–534, 2016. doi: 10.1007/s12311-015-0722-4 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leão RM, Kushmerick C, Pinaud R, Renden R, Li GL, Taschenberger H, Spirou G, Levinson SR, von Gersdorff H. Presynaptic Na+ channels: locus, development, and recovery from inactivation at a high-fidelity synapse. J Neurosci 25: 3724–3738, 2005. doi: 10.1523/JNEUROSCI.3983-04.2005 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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: 10.1038/nn1469 . [DOI] [PubMed] [Google Scholar]
- Majewska MD, Harrison NL, Schwartz RD, Barker JL, Paul SM. Steroid hormone metabolites are barbiturate-like modulators of the GABA receptor. Science 232: 1004–1007, 1986. doi: 10.1126/science.2422758 . [DOI] [PubMed] [Google Scholar]
- Mathews PJ, Lee KH, Peng Z, Houser CR, Otis TS. Effects of climbing fiber driven inhibition on Purkinje neuron spiking. J Neurosci 32: 17988–17997, 2012. doi: 10.1523/JNEUROSCI.3916-12.2012 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Medina JF, Nores WL, Ohyama T, Mauk MD. Mechanisms of cerebellar learning suggested by eyelid conditioning. Curr Opin Neurobiol 10: 717–724, 2000. doi: 10.1016/S0959-4388(00)00154-9 . [DOI] [PubMed] [Google Scholar]
- Mintz IM, Bean BP. GABAB receptor inhibition of P-type Ca2+ channels in central neurons. Neuron 10: 889–898, 1993. doi: 10.1016/0896-6273(93)90204-5 . [DOI] [PubMed] [Google Scholar]
- Misgeld U, Bijak M, Jarolimek W. A physiological role for GABAB receptors and the effects of baclofen in the mammalian central nervous system. Prog Neurobiol 46: 423–462, 1995. doi: 10.1016/0301-0082(95)00012-K . [DOI] [PubMed] [Google Scholar]
- Nicoll RA, Alger BE. Presynaptic inhibition: transmitter and ionic mechanisms. Int Rev Neurobiol 21: 217–258, 1979. doi: 10.1016/S0074-7742(08)60639-X. [DOI] [PubMed] [Google Scholar]
- Nicoll RA, Eccles JC, Oshima T, Rubia F. Prolongation of hippocampal inhibitory postsynaptic potentials by barbiturates. Nature 258: 625–627, 1975. doi: 10.1038/258625a0 . [DOI] [PubMed] [Google Scholar]
- Ohura S, Kamiya H. Short-term depression of axonal spikes at the mouse hippocampal mossy fibers and sodium channel-dependent modulation. eNeuro 5: ENEURO.0415-17, 2018. doi: 10.1523/ENEURO.0415-17.2018 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Orts-Del’Immagine A, Pugh JR. Activity-dependent plasticity of presynaptic GABAB receptors at parallel fiber synapses. Synapse 72: e22027, 2018. doi: 10.1002/syn.22027 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Palmer MR, Hoffer BJ. GABAergic mechanisms in the electrophysiological actions of ethanol on cerebellar neurons. Neurochem Res 15: 145–151, 1990. doi: 10.1007/BF00972204 . [DOI] [PubMed] [Google Scholar]
- Pouzat C, Marty A. Somatic recording of GABAergic autoreceptor current in cerebellar stellate and basket cells. J Neurosci 19: 1675–1690, 1999. doi: 10.1523/JNEUROSCI.19-05-01675.1999 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Price GD, Trussell LO. Estimate of the chloride concentration in a central glutamatergic terminal: a gramicidin perforated-patch study on the calyx of Held. J Neurosci 26: 11432–11436, 2006. doi: 10.1523/JNEUROSCI.1660-06.2006 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pugh JR, Jahr CE. Axonal GABAA receptors increase cerebellar granule cell excitability and synaptic activity. J Neurosci 31: 565–574, 2011. doi: 10.1523/JNEUROSCI.4506-10.2011 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pugh JR, Jahr CE. Activation of axonal receptors by GABA spillover increases somatic firing. J Neurosci 33: 16924–16929, 2013. doi: 10.1523/JNEUROSCI.2796-13.2013 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pugh JR, Raman IM. Potentiation of mossy fiber EPSCs in the cerebellar nuclei by NMDA receptor activation followed by postinhibitory rebound current. Neuron 51: 113–123, 2006. doi: 10.1016/j.neuron.2006.05.021 . [DOI] [PubMed] [Google Scholar]
- Rodgers-Neame NT, Covey DF, Hu Y, Isenberg KE, Zorumski CF. Effects of a benz[e]indene on gamma-aminobutyric acid-gated chloride currents in cultured postnatal rat hippocampal neurons. Mol Pharmacol 42: 952–957, 1992. [PubMed] [Google Scholar]
- Rothman JS, Silver RA. NeuroMatic: an integrated open-source software toolkit for acquisition, analysis and simulation of electrophysiological data. Front Neuroinform 12: 14, 2018. doi: 10.3389/fninf.2018.00014 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ruiz A, Campanac E, Scott RS, Rusakov DA, Kullmann DM. Presynaptic GABAA receptors enhance transmission and LTP induction at hippocampal mossy fiber synapses. Nat Neurosci 13: 431–438, 2010. doi: 10.1038/nn.2512 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ruiz A, Fabian-Fine R, Scott R, Walker MC, Rusakov DA, Kullmann DM. GABAA receptors at hippocampal mossy fibers. Neuron 39: 961–973, 2003. doi: 10.1016/S0896-6273(03)00559-2 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rusakov DA, Kullmann DM. Extrasynaptic glutamate diffusion in the hippocampus: ultrastructural constraints, uptake, and receptor activation. J Neurosci 18: 3158–3170, 1998. doi: 10.1523/JNEUROSCI.18-09-03158.1998 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Santhakumar V, Hanchar HJ, Wallner M, Olsen RW, Otis TS. Contributions of the GABAA receptor alpha6 subunit to phasic and tonic inhibition revealed by a naturally occurring polymorphism in the alpha6 gene. J Neurosci 26: 3357–3364, 2006. doi: 10.1523/JNEUROSCI.4799-05.2006 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Santhakumar V, Meera P, Karakossian MH, Otis TS. A reinforcing circuit action of extrasynaptic GABAA receptor modulators on cerebellar granule cell inhibition. PLoS One 8: e72976, 2013. doi: 10.1371/journal.pone.0072976 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Santhakumar V, Wallner M, Otis TS. Ethanol acts directly on extrasynaptic subtypes of GABAA receptors to increase tonic inhibition. Alcohol 41: 211–221, 2007. doi: 10.1016/j.alcohol.2007.04.011 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shimura M, Harata N, Tamai M, Akaike N. Allosteric modulation of GABAA receptors in acutely dissociated neurons of the suprachiasmatic nucleus. Am J Physiol Cell Physiol 270: C1726–C1734, 1996. doi: 10.1152/ajpcell.1996.270.6.C1726 . [DOI] [PubMed] [Google Scholar]
- Slesinger PA, Stoffel M, Jan YN, Jan LY. Defective γ-aminobutyric acid type B receptor-activated inwardly rectifying K+ currents in cerebellar granule cells isolated from weaver and Girk2 null mutant mice. Proc Natl Acad Sci USA 94: 12210–12217, 1997. doi: 10.1073/pnas.94.22.12210 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Soiza-Reilly M, Anderson WB, Vaughan CW, Commons KG. Presynaptic gating of excitation in the dorsal raphe nucleus by GABA. Proc Natl Acad Sci USA 110: 15800–15805, 2013. doi: 10.1073/pnas.1304505110 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Soldo BL, Proctor WR, Dunwiddie TV. Ethanol differentially modulates GABAA receptor-mediated chloride currents in hippocampal, cortical, and septal neurons in rat brain slices. Synapse 18: 94–103, 1994. doi: 10.1002/syn.890180204 . [DOI] [PubMed] [Google Scholar]
- Stell BM. Biphasic action of axonal GABA-A receptors on presynaptic calcium influx. J Neurophysiol 105: 2931–2936, 2011. doi: 10.1152/jn.01125.2010 . [DOI] [PubMed] [Google Scholar]
- 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 USA 100: 14439–14444, 2003. doi: 10.1073/pnas.2435457100 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stell BM, Rostaing P, Triller A, Marty A. Activation of presynaptic GABAA receptors induces glutamate release from parallel fiber synapses. J Neurosci 27: 9022–9031, 2007. doi: 10.1523/JNEUROSCI.1954-07.2007 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Steuber V, Mittmann W, Hoebeek FE, Silver RA, De Zeeuw CI, Häusser M, De Schutter E. Cerebellar LTD and pattern recognition by Purkinje cells. Neuron 54: 121–136, 2007. doi: 10.1016/j.neuron.2007.03.015 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sun J, Pang ZP, Qin D, Fahim AT, Adachi R, Südhof TC. A dual-Ca2+-sensor model for neurotransmitter release in a central synapse. Nature 450: 676–682, 2007. doi: 10.1038/nature06308 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tao W, Higgs MH, Spain WJ, Ransom CB. Postsynaptic GABAB receptors enhance extrasynaptic GABAA receptor function in dentate gyrus granule cells. J Neurosci 33: 3738–3743, 2013. doi: 10.1523/JNEUROSCI.4829-12.2013 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Trigo FF, Marty A, Stell BM. Axonal GABAA receptors. Eur J Neurosci 28: 841–848, 2008. doi: 10.1111/j.1460-9568.2008.06404.x . [DOI] [PubMed] [Google Scholar]
- Turecek R, Trussell LO. Reciprocal developmental regulation of presynaptic ionotropic receptors. Proc Natl Acad Sci USA 99: 13884–13889, 2002. doi: 10.1073/pnas.212419699 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Valenzuela CF, Jotty K. Mini-Review: Effects of ethanol on GABAA receptor-mediated neurotransmission in the cerebellar cortex–recent advances. Cerebellum 14: 438–446, 2015. doi: 10.1007/s12311-014-0639-3 . [DOI] [PubMed] [Google Scholar]
- Wang L, Kloc M, Maher E, Erisir A, Maffei A. Presynaptic GABAA receptors modulate thalamocortical inputs in layer 4 of rat V1. Cereb Cortex 29: 921–936, 2018. doi: 10.1093/cercor/bhx364 . [DOI] [PubMed] [Google Scholar]
- Weisz CJ, Rubio ME, Givens RS, Kandler K. Excitation by axon terminal GABA Spillover in a Sound Localization Circuit. J Neurosci 36: 911–925, 2016. doi: 10.1523/JNEUROSCI.1132-15.2016 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wright R, Newey SE, Ilie A, Wefelmeyer W, Raimondo JV, Ginham R, Mcllhinney RA, Akerman CJ. Neuronal chloride regulation via KCC2 is modulated through a GABAB receptor protein complex. J Neurosci 37: 5447–5462, 2017. doi: 10.1523/JNEUROSCI.2164-16.2017 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu J, Wu LG. The decrease in the presynaptic calcium current is a major cause of short-term depression at a calyx-type synapse. Neuron 46: 633–645, 2005. doi: 10.1016/j.neuron.2005.03.024 . [DOI] [PubMed] [Google Scholar]
- Ye JH, Tao L, Ren J, Schaefer R, Krnjevic K, Liu PL, Schiller DA, McArdle JJ. Ethanol potentiation of glycine-induced responses in dissociated neurons of rat ventral tegmental area. J Pharmacol Exp Ther 296: 77–83, 2001. [PubMed] [Google Scholar]
- Zhang SJ, Jackson MB. GABAA receptor activation and the excitability of nerve terminals in the rat posterior pituitary. J Physiol 483: 583–595, 1995. doi: 10.1113/jphysiol.1995.sp020608 . [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zorrilla de San Martin J, Trigo FF, Kawaguchi SY. Axonal GABAA receptors depolarize presynaptic terminals and facilitate transmitter release in cerebellar Purkinje cells. J Physiol 595: 7477–7493, 2017. doi: 10.1113/JP275369 . [DOI] [PMC free article] [PubMed] [Google Scholar]



