SUMMARY
Developing neural circuits, including GABAergic circuits, switch receptor types. But the role of early GABA receptor expression for establishment of functional inhibitory circuits remains unclear. Tracking the development of GABAergic synapses across axon terminals of retinal bipolar cells (BCs) we uncovered a crucial role of early GABAA receptor expression for the formation and function of presynaptic inhibitory synapses. Specifically, early α3-subunit-containing GABAA (GABAAα3) receptors are a key developmental organizer. Before eye-opening, GABAAα3 gives way to GABAAα1 at individual BC presynaptic inhibitory synapses. The developmental downregulation of GABAAα3 is independent of GABAAα1 expression. Importantly, lack of early GABAAα3 impairs clustering of GABAAα1 and formation of functional GABAA synapses across mature BC terminals. This impacts the sensitivity of visual responses transmitted through the circuit. Lack of early GABAAα3 also perturbs aggregation of LRRTM4, the organizing protein at GABAergic synapses of rod BC terminals, and their arrangement of output ribbon synapses.
Keywords: inhibitory circuits, synapse formation, GABA receptor, retina, development
eTOC Blurb
Sinha et al. show that GABAA synapses on axon terminals of retinal bipolar cells alter receptor types before eye-opening. GABAAα3 receptors present during early development regulate the assembly and function of these synapses by promoting the clustering of GABAAα1 receptors and the organizing protein LRRTM4.
INTRODUCTION
Formation of accurate and efficient synapses relies on a collaborative effort of molecular and activity-dependent processes1. A common feature of developing neural circuits is a change in pre- and postsynaptic functional properties as circuits mature. For instance, developing inhibitory brainstem circuits switch neurotransmitter types (GABA->Glycine) and developing excitatory NMDA postsynapses switch receptor composition (GluN2B->2A) during maturation2, 3. Although these developmental alterations have been documented for brain regions and excitatory and inhibitory neurons across the CNS2, 4-8, the organizational role of these alterations have been more extensively studied for excitatory than inhibitory circuits.
Inhibitory circuits in the developing CNS are known to switch receptor composition and distinct receptor-types dominate in the immature vs mature CNS9, 10. Of note, α3-subunit containing GABAA receptors (GABAAα3Rs) are widely expressed in the CNS at the time of birth6, 11, 12, with a much restricted expression in the mature CNS13-15. But the role of this early receptor expression for the formation and function of individual inhibitory synapses remains unclear. Here, we focused on presynaptic inhibitory circuits of the mammalian retina to determine the role of early GABA receptors (GABARs) for the formation of functional inhibitory synapses.
Presynaptic inhibition onto axon terminals of neurons is a common circuit-motif that regulates neurotransmitter release. It is found in the spinal motor neuronal circuit16, the olfactory glomeruli circuit17 and the mammalian retinal circuit18. In the retina, amacrine cell (AC) interneurons provide presynaptic inhibition onto axon terminals of bipolar cells (BCs) to regulate glutamate release from BCs19, 20. This regulation of glutamate release serves several functions, including adjusting the dynamic range of operation of these neurons21, 22 and regulating the threshold of retinal visual responses23. BCs are glutamatergic second-order neurons that relay photoreceptor input from the outer retina to ACs and retinal output neurons (ganglion cells) in the inner retina (Figure 1A). Specific rod and cone BCs transfer dim-light (rod photoreceptor) and bright-light (cone photoreceptor) information to inner retinal neurons. BCs are classified as ‘ON’ if they depolarize to light increments, or ‘OFF’ if they depolarize to light decrements18, 24, 25. Rod BCs (RBCs) are ON-BCs, whereas cone BCs are either ON or OFF-BCs. BCs receive GABAergic presynaptic inhibition at their terminals, with both GABAA and GABACRs mediating inhibition onto ON-BC boutons18, 26-29. Presynaptic inhibition is generated in two ways: (1), an AC receives excitatory input from the same BC that it provides feedback inhibition onto (reciprocal connection; Figure 1A’), and (2), an AC receives excitatory input from a different BC (non-reciprocal inhibition). Both reciprocal and non-reciprocal synapses can be GABAergic and utilize ionotropic GABAARs30-32, composed of 2α-2β-1γ subunits33. In the retina, specific α-subunit (α1-α3) containing GABAARs are localized at distinct non-overlapping postsynapses34, 35.
Figure 1: Retinal BC terminals replace GABAAα3 for GABAAα1 during development.
(A) Schematic of the vertebrate retina, showing rod and cone photoreceptors (Pr) providing input to rod and cone bipolar cells (BCs). Cone BCs transfer visual signals to output ganglion cells (GC). BC terminals receive inhibitory input from amacrine cells (ACs) at the inner plexiform layer (IPL) of the retina (shaded grey box).
(A’) Glutamate (Glu) release at BC ribbon synapses is modulated by presynaptic inhibition from ACs. Enlarged view of a reciprocal GABAergic AC (rAC) which receives BC input and provides feedback inhibition, mediated by pentameric GABAARs composed of α, β and γ subunits.
(B) Single confocal image planes at the level of the IPL where rod BC axons stratify, co-labeled for GABAAα3 and GABAAα1Rs across development (P:postnatal day).
(C) Expression of GABAAα3 and GABAAα1Rs (pixels above background shown in white) within terminals (red) of developing rod and T6 BCs.
(D) % GABAAα3 and GABAAα1 occupancy within developing BC axons. For each time-point n>4 cells and N>3 animals; GABAAα1 P12-P30 data from Hoon et al.28. Data plotted as mean±SEM for all Figures.
Here, we determined the role of early GABAARs for the developmental organization and function of GABAergic synapses at ON-BC terminals. We chose the mouse retina for our study due to the wide availability of genetic tools that target specific BC and AC-types and retinal GABAAR populations. Combining these genetic tools with electrophysiology, high-resolution light and electron microscopy (EM) we identified a GABAAR rearrangement at developing BC presynaptic inhibitory synapses and uncovered a critical role of early GABAAR expression for the formation of functional presynaptic inhibitory synapses.
RESULTS
Early GABAAα3 receptors are replaced with GABAAα1Rs at retinal BC axon terminals
We first determined which GABAAR-types are expressed at developing BC terminals to identify the receptor-types that provide early presynaptic inhibition. We immunostained for two GABAAR-types in the developing mouse retina: GABAAα3R and GABAAα1R, because these two types have been found at adult BC terminals26, 28, 36. In postnatal day 7 (P7) retina, GABAAα3R labeling was abundant at the outer portion of the retinal inner plexiform layer (IPL), where ON-BC axons stratify. GABAAα1R clusters were, however, hardly noticeable in this layer at this early age (Figure 1B). Just before eye-opening, at P12, GABAAα1 clustering increased in this region of the IPL, increasing further by P30. In contrast, GABAAα3R labeling in the ON-lamina of the IPL decreased with age (Figure 1B).
The developmental downregulation of retinal GABAAα3R expression is also reflected in the promoter accessibility of retinal GABAARs. Promoter accessibilities of GABAAα3 and GABAAα1 were compared using a previous retinal ATAC-seq (Assay for Transposase-Accessible Chromatin using sequencing) dataset37 that spanned different time-points (embryonic day 17.5, P0/day of birth, P3-10, P14/eye-opening, and P21). Whole retina analyses showed high promoter accessibility for GABAARs at early time-points with diminished accessibility after eye-opening (Figure S1A). To compare GABAA promoter accessibility specifically across retinal BCs, we analyzed the Jorstad et al.38 ATAC-seq dataset from adult ON-BCs. This dataset utilized FACS-purified BCs from the adult Grm6-tdTomato transgenic line where ON-BCs selectively express the fluorescent protein tdTomato39. Comparing GABAAα3 and GABAAα1 promoter accessiblity from adult ON-BCs revealed that whereas the GABAAα1 promoter was still accessible in adult ON-BCs, the GABAAα3 promoter accessiblity was undetectable (Figure S1A). We also compared GABAA mRNA levels between P6 vs P50 retinas and found an almost 2.5 fold increase in the mRNA levels of GABAAα1 between these time-points, whereas the relative GABAAα3 mRNA levels at P50 reduced to about half their P6 level (Figure S1B).
To visualize GABAAα3 and GABAAα1R clusters specifically on developing BC terminals, we used Grm6-tdTomato mice to visualize individual ON-BCs, and immunolabeled these retinas for GABAAα3R and GABAAα1R at different ages (Figure 1C). We focused on two ON-BC types: RBCs that convey dim-light signals and Type 6 ON-cone BCs (T6) that stratify in the same plexus as RBCs but convey high-luminance information. We quantified the percent receptor occupancy of each GABAAR-type relative to axon terminal volume, as a measure of receptor levels. We performed this analysis on P9, P12, P16 and P30 BC terminals (Figure 1C-D). For both RBCs and T6s, GABAAα3 levels were more than 10 fold higher at P9 compared to P30 (RBC P9 GABAAα3= 6.21±0.65; RBC P30 GABAAα3= 0.28±0.07; p-value9vs30 = 0.000083: T6 P9 GABAAα3= 5.67±0.16; T6 P30 GABAAα3= 0.48±0.14; p-value9vs30 = 0.0075). A rapid decline in GABAAα3R levels was observed around P12 just before eye-opening (Figure 1D). Conversely, axonal GABAAα1R levels of RBCs and T6s increased from P9 until P16 (two days after eye-opening) and thereafter reduced relative to the mature/P30 BC terminal volume (Figure 1D). Together, our observations reveal a developmental downregulation of GABAAα3 with a concomitant increase in GABAAα1 expression during retinal circuit maturation, at both the mRNA and protein level.
To determine whether or not the developmental transition from GABAAα3Rs to GABAAα1Rs on BC axons occurs at the same synapse during maturation, we focused on the A17->RBC GABAergic synapse because of the stereotyped arrangement of this connection across species and the availability of transgenic tools to directly visualize this contact. GABAergic A17 ACs provide reciprocal feedback inhibition onto RBC terminals23, 40. Using the Ai9/slc6a5-cre double transgenic line in which A17 ACs express tdTomato (Figure S2A), individual A17s were targeted and filled with the dye, lucifer yellow, at two developmental time-points: P11 (before eye-opening) and >P30 (adult/mature). A17 dye-filled retinas were immunolabeled for GABAAα3R or GABAAα1R together with the RBC marker, protein kinase C (PKC)41, enabling identification of GABAAR-types at A17->RBC synapses (Figure 2A). To quantify GABAAR levels at A17->RBC synapses, the volume overlap between A17 varicosities and RBC terminals (i.e. synaptic contact volume) was digitally isolated, and thereafter the pixel volume of GABAAR immunoreactivity relative to the synaptic overlap volume was expressed as ‘percent occupancy’ (Figure 2B). Before eye-opening (P11), A17->RBC synapses were enriched with GABAAα3Rs. Mature A17->RBC synapses instead largely contained GABAAα1Rs (Figure 2B), demonstrating a developmental alteration in the GABAAR-type enriched at A17->RBC synapses.
Figure 2: Individual A17-rod BC inhibitory synapses alter their GABAAR-type before eye-opening.
(A) Top-down view of lucifer yellow (LY/red) filled A17s in adult and P11 retinas co-immunolabeled with PKC (blue) and GABAAα3 or GABAAα1Rs (yellow). The GABAA signal above background and within the A17-RBC volume-overlap is depicted in yellow pixels. Higher-magnification views of three regions per A17, and side-view of the A17-fills are provided.
(B) % GABAA occupancy at A17-rod BC appositions in P11 (top) and mature retina (>P30; bottom plot). n=number of A17s analyzed. Each retina-piece contained a single A17-fill; N=3 animals.
GABAAα3 receptor clustering at BC terminals show a developmental decline even in the absence of GABAAα1
We next determined whether elimination of GABAAα1R from ON-BCs would favor maintenance of high GABAAα3R clusters across adult ON-BC terminals. Crossing the GABAAα1 floxed line42 with the ON BC-specific Grm6-Cre28 line and the Ai9 reporter line (Ai9/Grm6-Cre/GABAAα1 cKO) enabled specific deletion of GABAAα1 from ON BCs (GABAAα1 conditional knockout or cKO) and visualization of the GABAAα1-deficient BCs (Figure S2B). To confirm a lack of GABAA currents from adult RBCs in the GABAAα1cKO, we performed single-cell patch-clamp recordings from RBCs and recorded their response to GABA-puffs. Previous studies using this technique demonstrated that the GABA-evoked response from RBCs is mediated by both GABAC and GABAARs19, 28, 36, 43. We thus measured the RBC response to GABA-puffs at axon terminals before and after application of the GABACR antagonist, TPMPA (Figure S3A). The difference between these responses, i.e., the GABAA-specific response component, was strikingly lower in GABAAα1cKO RBCs compared to controls (Figure S3A-B), confirming the absence of GABAAα1Rs across adult GABAAα1cKO RBC terminals.
We next analyzed GABAAα3R clustering on adult RBCs and T6s in the GABAAα1cKO. GABAAα3R levels across GABAAα1cKO RBC and T6s were comparable (Figure 3A) to control BCs of the same type. Quantification of GABAAα1R clusters in GABAAα1cKO RBCs and T6s confirmed the downregulation of GABAAα1Rs across both axons and dendrites of KO BCs compared to control (Figure 3B). Quantification of GABAAα3R clusters within GABAAα1cKO RBCs and T6s and control BCs of the same type did not reveal any differences (Figure 3C). Of note, GABAAα3 immunolabeling is restricted to BC axons (Hoon et al.28 and Figure 3C) with negligible GABAAα3R clustering in BC dendrites across genotypes (Figure 3C). Thus, elimination of GABAAα1 from ON-BCs does not alter mature levels of GABAAα3 at BC terminals, underscoring that increasing GABAAα1 expression does not serve as a developmental cue to downregulate GABAAα3 at maturing BC terminals. GABAAα3R levels within GABAAα1cKO RBC terminals were also comparable to controls at P12 (Figure S3C), ruling out compensatory regulation in the GABAAα1cKO early in development, and highlighting a GABAAα1R-independent regulation of GABAAα3R clustering at inhibitory synapses onto ON-BC terminals.
Figure 3: Absence of GABAAα1 on BC axons does not impact GABAAα3R clustering within adult terminals.
(A) Examples of GABAAα3 signal (white pixels represent signal above background) within adult rod BC (top) and T6 (bottom panel: red) terminals in GABAAα1cKO-littermate control (Ctrl) retina.
% occupancy of GABAAα1 (B) and of GABAAα3 (C) within adult α1cKO-Ctrl BC axons and dendrites (top plots: rod BC; bottom plots:T6). N≥4 GABAAα1 cKO-Ctrl pairs. For all Figures: Number in parenthesis indicates cells analyzed and p-value listed for two-tailed unpaired T-test.
Early presence of GABAAα3 is necessary for accruing GABAAα1Rs at BC terminals
Although GABAAα3R clustering on ON-BCs does not rely on the presence on GABAAα1, the developmental increase in GABAAα1R clusters may depend on the early presence of GABAAα3. To test this hypothesis, we utilized mice that lack expression of GABAAα3 subunits globally44. We confirmed that GABAAα3R clustering was abolished in the retinal IPL of the GABAAα3KO (Figure S4A). We next crossed GABAAα3KO mice with Grm6-tdTomato mice to enable visualization of individual RBCs and T6s in the KO background. We first quantified the occupancy of GABAAα3 within RBC and T6 terminals in the GABAAα3KO (Figure S4B). GABAAα3 was significantly diminished across terminals of both BCs in the GABAAα3KO compared to control (Figure S4B). RBC and T6 dendrites showed negligible GABAAα3 clustering across genotypes (Figure S4B).
We next evaluated GABAAα1R clustering in GABAAα3KO retinas (Figure 4). Figure 4A shows a retinal volume of an adult GABAAα3KO and control, immunolabeled for GABAAα1. The outer plexiform layer (OPL) where BC dendrites stratify had comparable GABAAα1 levels across genotypes (Figure 4A). However, the IPL where BC terminals stratify, displayed a striking reduction in GABAAα1 immunolabeling in the GABAAα3KO, compared to control (Figure 4A). The drastic reduction in axonal GABAAα1R clustering was confirmed when we quantified GABAAα1 occupancy at axon terminals of individual RBCs and T6s (Figure 4B-C; GABAAα1 occupancy at BC dendrites was unchanged). Thus, GABAAα1R clustering on RBC and T6 terminals requires the early presence of GABAAα3. GABAAα1R clusters were also significantly downregulated at inhibitory synapses onto another cone BC-type (Type1 OFF cone BC) in the GABAAα3KO (Figure S5A). This BC-type laminates at the opposite border of the IPL and the GABAAα3KO line was crossed to the Vsx1-cerulean line28 to visualize Type1s in the KO. In contrast, glycinergic inhibitory synapses on Type1 terminals, as evaluated by presence of α1-subunit containing glycine receptors at these synapses28, remained unaltered in the GABAAα3KO (Figure S5B). Thus, the role of early GABAAα3 in establishing GABAAα1R presynaptic inhibitory synapses appears to be conserved across diverse BC-types.
Figure 4: GABAAα1R clustering across BC terminals requires early GABAAα3 expression.
(A) Volume views of GABAAα1 immunofluorescence in the outer and inner plexiform layers (OPL; IPL) of an adult GABAAα3KO and littermate control retina.
(B) GABAAα1 expression (white pixels) within adult rod (left) and T6 (right) BC axons and dendrites in GABAAα3KO-littermate control (Ctrl) retina.
(C) % GABAAα1 occupancy within adult GABAAα3KO-Ctrl BC axons and dendrites (left: rod BC; right plot:T6). N>5 GABAAα3KO-Ctrl pairs.
(D) % occupancy of GABAAα1 within P12 GABAAα3KO-Ctrl BC axons and dendrites (left: rod BC; right plot:T6). N>4 GABAAα3KO-Ctrl pairs.
See also Figures S4-7.
Presynaptic inhibition at RBC and T6 terminals is mediated by both GABAA and GABACRs28. Blocking vesicular release of inhibitory neurotransmitters by elimination of the retinal vesicular inhibitory amino acid transporter (VIAAT) impairs maintenance of both GABAA and GABACRs at RBC and T6 terminals28. We thus immunolabeled GABACRs in GABAAα3KO retina to determine whether or not early expression of GABAAα3 impacts GABACR clustering at RBC and T6 axons (Figure S6A). GABACR occupancy remained unperturbed in both RBC and T6 terminals (Figure S6A), suggesting that early GABAAα3 expression specifically regulates maturational changes at GABAA synapses. GABACR immunoreactivity on dendrites of RBC and T6s was negligible in both GABAAα3KO and control (Figure S6A), in keeping with previous findings that GABACR are sparse on ON-BC dendrites28.
To test whether early GABAAα3 plays a role in the establishment or maintenance of GABAAα1Rs across BC terminals, we quantified GABAAα1R occupancy across P12 GABAAα3KO ON-BC terminals. We observed a substantial reduction of GABAAα1R clustering in P12 GABAAα3KO RBC and T6 terminals relative to control (Figure 4D). Early GABAAα3 expression is thus necessary for the developmental increase of GABAAα1Rs on ON-BC terminals, unlike a lack of VIAAT, which impairs maintenance but not the initial accruing of GABAAα1Rs across RBC and T6 terminals before eye-opening28. As an important internal control, GABAAα1R levels across dendrites of the same BCs were not altered in the GABAAα3KO (Figure 4D). We confirmed that GABAAα3 expression was abolished across both axons and dendrites of P12 GABAAα3KO BCs (Figure S4C), and also determined that GABACR levels at ON-BC axons were unchanged in the P12 GABAAα3KO (Figure S6B). Together, our observations reveal a critical role for GABAAα3 in establishing GABAAα1 inhibitory synapses at developing BC axon terminals.
GABAARs do not aggregate at BC axons in GABAAα3KO and failure of GABAAα1 to cluster may be due to impaired receptor trafficking
The lack of GABAAα1R clustering on RBC terminals prompted us to evaluate whether other GABAAR subunits are also affected in the GABAAα3KO. GABAAR pentamers are comprised of 2α-2β-1γ subunits33, and our previous work has shown that α1-subunits come together with γ2-subunits at GABAAR synapses across ON-BC terminals28; see also Wässle et al.34 and Greferath et al.45. GABAAβ2/3 expression is enriched at the outermost lamina of the IPL, where RBC terminals stratify34, 45, 46 and specifically mRNA for GABAAβ3 is found in the layer where BC somata reside45. We immunolabeled for GABAAγ2 and GABAAβ3-subunits in retinal slices together with the RBC marker, PKC (Figure 5A) and found both receptor subunits to be robustly expressed on wildtype RBC terminals (control panel in Figure 5A). However, in the GABAAα3KO, levels of both GABAAγ2 and GABAAβ3-subunits on adult RBC terminals were drastically reduced compared to controls (Figure 5A-B), suggesting a lack of clustering of all GABAAR subunits on adult BC terminals in the absence of early GABAAα3.
Figure 5: GABAAR subunits are downregulated in GABAAα3KO but GABAAα1 total protein expression and promoter accessibility are unchanged in α3KO retina.
(A) Slices from adult GABAAα3KO-littermate control retinas co-immunolabeled for PKC (red) and GABAAγ2 (left) or GABAAβ3 (right panel: yellow). Bottom panels show higher-magnification views of the selected regions (white box) showing receptor immunostaining within PKC-labeled boutons.
(B) % occupancy of GABAAγ2 and GABAAβ3 within adult rod BC boutons across genotypes (N>4 GABAAα3KO-Ctrl pairs; ≥6 sections analyzed per condition).
(C) Western blot showing total protein levels of GABAAα1 compared to the reference Actin, of retinas from two pairs of adult GABAAα3KO-littermate control animals (KO-WT respectively).
(D) Normalized ATAC-Seq tracks from adult GABAAα3KO and littermate control (Ctrl) retinas showing promoter accessibility for the Gabra1 locus. Red box highlights similar Gabra1 promoter accessibility between genotypes.
See also Figures S6 and S7.
Immunolabeling for GABAAα1 reveals the amount of receptor clustered at the postsynaptic surface. Receptor clustering depends not only on total receptor protein expression but also on efficient trafficking from the protein synthesis machinery to the postsynaptic membrane. To determine which of these two possibilities account for decreased GABAAα1R clustering in the GABAAα3KO, we measured the total retinal GABAAα1 protein levels by western blot analyses. Total protein levels of GABAAα1 were equivalent between GABAAα3KO and control retinas (Figure 5C), indicating that GABAAα1 net protein levels are unaltered in the GABAAα3KO. Although we could not measure protein levels specifically in RBCs, the pan-reduction in GABAAα1 clustering in the IPL of the α3KO suggests that it is unlikely that GABAAα1 protein levels were differentially affected across cell-types. We also compared GABAAα1 mRNA levels in GABAAα3KO retinas compared to control and did not observe any change (Figure S6C). As a control, we determined mRNA levels of the RBC specific gene PCP247 and GABAC in GABAAα3KO retinas compared to control (Figure S6D). The mRNA levels for both these genes remained comparable across genotypes (Figure S6C-D). As a final assay on the integrity of the GABAAα1 transcriptional machinery in GABAAα3KO retinas, we performed ATAC-seq analyses on adult GABAAα3KO-control retina pairs. We did not observe any noticeable differences in the GABAAα1 promoter accessibility (Gabra1 locus, Figure 5D) in the GABAAα3KO retina compared to control. Together, our observations suggest that transcription and translation of retinal GABAAα1 is unlikely to be perturbed in the GABAAα3KO, raising the possibility that a deficit in the trafficking of GABAAα1 to the postsynaptic membrane underlies the lack of GABAAα1R clusters on BC axons in the GABAAα3KO.
Lack of GABAAergic presynaptic inhibition at GABAAα3-deficient rod BC terminals alters rod BC output
To determine how the absence of GABAAα1R clusters in GABAAα3KO retina impacts GABA-evoked currents of RBCs, we performed whole-cell patch-clamp recordings from these BCs in slice preparations of adult retina, and recorded currents in response to GABA-puffs at their axon terminals. GABA-evoked responses in RBCs are mediated by GABAA and GABACRs19, 28, 36, 43, with near equal contributions of both receptor-types to the total evoked current amplitude28, 36, 43. Comparison of evoked responses of RBCs in GABAAα3KO-control retinas suggests a ~50% reduction in the total GABA-evoked response in the KO (Figure 6A). Application of a GABACR antagonist (TPMPA) isolated the GABAA-component of the response and revealed that GABAA-mediated currents were substantially reduced in GABAAα3KO RBCs compared to control (Figure 6A). The response remaining in TPMPA was eliminated upon application of the GABAAR antagonist, GABAzine (Figure 6A). The amplitudes of both the total GABA-evoked current and the GABAAR-mediated component were significantly reduced in GABAAα3KO RBCs compared to control (Figure 6B). These recordings corroborate the lack of GABAAR clusters on axon terminals of adult GABAAα3KO RBCs.
Figure 6: GABAA-mediated response from GABAAα3KO rod BC terminals is attenuated leading to increased AII sensitivity.
(A) Responses from adult rod BCs after GABA-puff application at their terminals. Exemplar trace depicting net GABA-evoked responses from GABAAα3KO and littermate control (Control) rod BCs, and responses after application of the GABACR antagonist, TPMPA, and GABAAR antagonist, GABAzine.
(B) Quantification of rod BC current amplitude before and after TPMPA application from adult GABAAα3KO-Control retinas. Numbers within the histograms represent recorded cells in each condition.
(C) Exemplar recordings of responses to brief light flashes from an AII in adult GABAAα3KO (right) and Control (middle) whole-mount retinas. Thick lines represent the average response (shaded region=SEM) to ten repeats of a dim flash (grey arrow/trace; 0.025 R*/Rod/flash) and a flash that was ~100 times brighter (purple arrow/trace; 3.2 R*/Rod/flash).
(D) Normalized flash responses across adult AIIs for Control (closed markers/Ctrl, N=5 animals) and GABAAα3KO (open markers/α3KO, N=4 animals) retinas. Insets: Example responses to a range of flash strengths.
(E) Quantification of the flash strength at which the amplitude of the AII response reaches 50% of its maximum value (R50%) across adult GABAAα3KO-Ctrl.
A lack of GABAA inhibition at RBC terminals could affect the efficacy of visual information transfer from RBCs. RBC terminals release glutamate at ribbon synaptic sites apposed to two postsynaptic AC partners. One of the AC partners, the AII, is responsible for conveying visual information from RBCs to ganglion cells18, 24, 25. To determine how lack of GABAA presynaptic inhibition impacts RBC->AII transmission, we performed voltage clamp recordings from AII ACs in GABAAα3KO-control retinas in a wholemount preparation and evaluated responses of AII cells across genotypes. We isolated the excitatory RBC-driven synaptic input across the dim-light regime (Figure 6C). Across flash strengths we observed exaggerated responses from GABAAα3KO AIIs compared to control (Figure 6C). Normalization of the AII response across the flash strengths probed showed a leftward shift of the AII response amplitude vs flash intensity curve in GABAAα3KO compared to control (Figure 6D), indicating an increased light sensitivity of GABAAα3KO AIIs (Figure 6D). We quantified the sensitivity of AIIs across genotypes by fitting the stimulus-response data of individual AIIs with a sigmoid and extracting the flash strength at which the response amplitude reached 50% of its maximum value (R50%). The R50% value was significantly reduced for GABAAα3KO AIIs compared to control (Figure 6E), confirming an increased light sensitivity of GABAAα3KO AIIs. Our observations thus reveal an abnormal RBC->AII transmission when GABAA presynaptic inhibition is impaired in the GABAAα3KO.
Synaptic dyad assembly at rod BC terminals is perturbed in the GABAAα3KO
Our observations thus far led us to ask whether clustering of LRRTM4, the organizing protein at GABAergic synapses of RBC axons43 could also be affected in GABAAα3KOs. We immunolabeled for LRRTM4 in adult GABAAα3KO-control retinas (Figure 7A) and found that LRRTM4 occupancy on RBC terminals was significantly reduced by almost half in GABAAα3KO compared to control (Figure 7B). As GABAergic presynaptic inhibition onto RBC axons is mediated almost equally by GABAA and GABACRs28, 36, 43, and because GABACR occupancy is unchanged in the GABAAα3KO (Figure S6A-B), our observations imply that LRRTM4 associated with GABAA synapses could be selectively altered in GABAAα3KO RBC terminals.
Figure 7: LRRTM4 is downregulated in GABAAα3KO rod BC terminals with perturbed synaptic dyad arrangements.
(A) LRRTM4 immunofluorescence (white) localized within adult rod BC terminals (red) in a littermate control (Ctrl) and GABAAα3KO retina. Insets show side-views.
(B) % occupancy of LRRTM4 within adult GABAAα3KO-Ctrl rod BC boutons. N=4 animal pairs.
(C) Single-plane electron micrograph of an adult rod BC terminal in the GABAAα3KO retina with two ribbon sites apposed to an A17 and an AII process characteristic of a normal dyad arrangement. The right panel is pseudo-colored for visualization of synaptic elements/partners. A17->RBC inhibitory synapse denoted as Inh Syn.
(D) 3D view of an entire rod BC terminal (gray) from an adult GABAAα3KO showing ribbons (yellow), inhibitory sites (red), AII (magenta) and A17 (green) partners at each ribbon. Inset shows magnified view of a ribbon synapse with dyad components.
(E-E”) Single-plane electron micrograph of an adult GABAAα3KO rod BC with ribbons mis-localized opposite a single AII process or present at a triadic arrangement with an AII and two different A17 processes. Bottom panels are pseudo-colored. E” is the 3D rendering of GABAAα3KO triad arrangement shown in E’.
(F) Fraction of total rod BC ribbon sites with indicaed postsynaptic arrangements in adult wildtype-GABAAα3KOs. (3 wildtype, 3 GABAAα3KO reconstructions from 2 animal pairs). For AII/A17 arrangements, the values for GABAAα3KO include AII/A17 dyads with lateral/horizontal and vertical ribbons. AII/A17 dyads with lateral ribbons were never observed in wildtype. Wildtype data from Sinha et al.43.
See also Figure S6.
Presynaptic inhibition at RBC terminals regulates assembly of output synapses at these terminals which occur at specialized ribbon sites apposed to two postsynaptic ‘dyad’ partners43. Ribbon synapses are sites of glutamate release stereotypically organized such that a single RBC ribbon is apposed to an A17 and AII AC process with the A17 partner providing reciprocal inhibition (Figure 7C). Both loss of inhibitory neurotransmission and LRRTM4 expression were previously found to disrupt the assembly of the RBC dyad, seen as a reduction in the number of ribbons correctly apposed to one AII and one A17 process43. We thus performed serial blockface scanning EM and reconstructed RBC terminals in GABAAα3KO retina to determine the ribbon synapse arrangements of adult GABAAα3KO RBC terminals (Figure 7D). On average, we observed 53.67±4.26 ribbons in the GABAAα3KO RBC terminals (n=3) we reconstructed, which was similar to the ribbon numbers in wildtype RBC terminals (50.0±2.08; p-value = 0.48). We then determined the postsynaptic partners at each ribbon site and found several RBC ribbons in the KO erroneously localized across a single AII AC or localized at a three-partner ‘triad’ junction (Figure 7E-E”); such erroneous contacts were not observed in wildtype. These anomalies led to a significant reduction in the number of GABAAα3KO RBC terminal ribbons correctly apposed across a pair of AII-A17 partners (Figure 7F; correct ribbon assembly rate reduced to 76±3% in the α3KO compared to 91±1% in wildtype). This impairment was less severe than that produced by reduction of both GABAA and GABAC-mediated inhibition across RBC terminals (e.g. in the LRRTM4KO where the correct dyad assembly is reduced to ~52%43). These observations suggest a ‘dose-dependent’ effect of GABAergic inhibition in influencing RBC dyad assembly. The lack of GABAA-mediated inhibition in GABAAα3KO, however, did not impair the formation of inhibitory synapses across RBC terminals because the number of both reciprocal (A17-mediated) and non-reciprocal inhibitory synapses in the KO was comparable to wildtype (reciprocal synapses at wildtype terminals = 34.00±1.15; reciprocal synapses at GABAAα3KO terminals = 31.33±1.33; p-value = 0.21; nonreciprocal synapses across wildtype terminals = 42.33±1.20; nonreciprocal synapses across GABAAα3KO terminals = 43.00±1.53; p-value = 0.75; n= 3 BCs per genotype).
We also quantified inhibitory synapse number across P12 RBC terminals in GABAAα3KO-control retinas by determining the number of VIAAT-immunoreactive boutons that were spatially apposed to P12 RBC terminals (Figure S6E). We did not observe any difference across genotypes. We verified that the number of inhibitory synapses obtained through our analyses of VIAAT-immunolabeled appositions in wildtype retina (44.13±0.696; n=15 RBC terminals, 7 animals) are comparable to the total number of inhibitory synapses at RBC axonal arbors (reciprocal + non-reciprocal), determined by a serial EM dataset from P12 wildtype retina (45±2; n= 2 RBC terminals reconstructed; P12-EM dataset from Sinha et al.43). Thus, the total number of inhibitory synapses onto a RBC terminal is not altered in the GABAAα3KO.
DISCUSSION
Role of early GABARs in the development of inhibitory circuits
Our study unveiled an essential role of early GABAAα3 for inhibitory synapse formation and function at axo-axonic synapses. GABAAR clustering was reduced across BC axons in the GABAAα3KO (Figure S7); clustering of several GABAAR subunits, GABAAα1, GABAAβ3 and GABAAγ2, were also diminished. LRRTM4, an organizing protein at GABAergic synapses of RBC axons43, was also significantly reduced in GABAAα3KOs. These changes are specific to GABAAR synapses; GABACR synapses adjacent to GABAAR synapses on BC axons23 were not affected by a loss of GABAAα3. The absence of GABAC upregulation in GABAAα3KO BCs also underscores a lack of engagement of homeostatic mechanisms to compensate for the loss of GABAAα1 and reduced inhibition. This contrasts with findings in retinas with impaired GABA synthesis where RBCs homeostatically adjust their output36. The absence of compensation in GABAAα3KO BC terminals is similar to observations from other brain regions where a loss of GABAAα3 does not cause an upregulation of other GABAAR subunits48. Our current observations also suggest that early GABAAα3Rs have a separate role in GABAergic synapse development, distinct from that of inhibitory neurotransmission, which influences clustering of both GABAA and GABACRs at ON-BC terminals28. In somatic and dendritic synapses of the thalamic reticular nucleus, GABAAγ2 subunit expression is disrupted in the absence of GABAAα3 and leads to the mis-localization of the inhibitory scaffolding protein, gephyrin, to non-synaptic sites48. Thus, early GABAAα3 expression may have common organizational role(s) for inhibitory synapse development at both axons and dendrites and across diverse CNS circuits.
Our estimates of total GABAAα1 protein expression, mRNA levels and promoter accessibility did not yield significant differences between GABAAα3KO-control retinas. The reduced clustering of GABAAα1 at BC axons may therefore be due to perturbed trafficking of GABAAα1 to postsynaptic sites when GABAAα3 is not expressed earlier in development. However, future experiments assessing protein levels specifically from BCs will be needed to test this hypothesis. Whether the physical presence of GABAAα3 is needed to guide GABAAα1 and the scaffolding protein LRRTM4 to GABAA postsynapses of BC axons, or whether early GABAAα3-mediated transmission is key for recruiting GABAAα1 and LRRTM4 to these synapses is unknown. But our previous observations from retinas deficient in inhibitory neurotransmission (VIAAT or Glutamic acid decarboxylase KO) may help distinguish between these possibilities. We had found that inhibitory neurotransmission regulates the maintenance but not the initial accumulation of GABAAα1Rs on BC terminals28, 36. Because GABAAα3 is critical for the initial accruing of GABAAα1Rs at BC axons, it is likely that GABAAα3 regulates the formation of inhibitory synapses onto BC axons through an activity-independent mechanism.
We demonstrate that early GABAAα3Rs are critical for the functional development of GABAA synapses at RBC terminals. The absence of this receptor-type during development leads to hypersensitivity of AII AC visual responses, and perturbed RBC->AII transmission. The deficiency of GABAA presynaptic inhibition at GABAAα3KO terminals could underlie the hypersensitivity of AII visual responses. Furthermore, the ultrastructural misarrangements of RBC output ribbon synapses could also contribute towards AIIs receiving abnormal RBC-mediated input. The number of RBC ribbons correctly apposed to an A17-AII pair in the GABAAα3KO was reduced by ~16%. This is, however, not as severe as the disruption to dyad assembly (~37% reduction in AII/A17 dyads) at RBC axons when both GABAA and GABAC-mediated presynaptic inhibition are impaired43. Because GABAA and GABACRs are expressed by RBCs in near equal proportions and contribute equally to the net GABA-evoked responses28, 36, 43, the extent of error in RBC dyad assembly may correlate with the ‘amount’ of presynaptic inhibition onto BC axons during development. Because GABAAα3KO RBCs still have GABAC-mediated inhibition, the ultrastructural mis-arrangements in the assembly of their output synapses may be less severe.
Taken together, our observations underscore a critical role of early GABAAα3 for the formation and function of inhibitory synapses at retinal BC axons. Given the extensive expression of GABAAα3 across developing brain regions, our findings raise the possibility that this receptor-type may have a more widespread developmental role in shaping inhibitory circuits than previously appreciated.
Receptor composition changes at synapses during development
Subunit composition of receptors at the postsynapse is a key determinant of response type and response kinetics. Thus, receptor composition has a direct functional correlate. It is well known that glutamate receptor composition in the CNS changes with circuit maturation. The GluN2B->GluN2A receptor subunit switch results in faster kinetics of NMDA receptor-mediated responses3. GluN2A-containing receptors exhibit faster rising and decaying currents compared to GluN2B-containing receptors3, 7, 49. The GluN2B->2A switch has also been suggested to support the acquisition or enhancement of learning capabilities3, 50, with the shift from GluN2B->2A controlling the threshold for modifying synaptic strength51.
Developing GABAergic neurons have similarly been found to change their GABAR composition to support faster inhibitory responses with maturation. The α–subunit of the GABAAR is an important determinant of response kinetics6. In the thalamic reticular nucleus, GABAAα5-containing receptors mediate prolonged early postnatal tonic inhibitory current and a switch to GABAAα3 allows shortening of inhibitory postsynaptic currents and generation of faster rhythmic oscillations5. This receptor-type switch enables facilitation of the spontaneous network activity in the thalamic reticular nucleus, which is important for supporting increased external environment awareness of the organism5. GABAAR subunit changes also occur in the rat visual cortex during development and have been correlated with an accelerated decay of spontaneous inhibitory currents52. Like these past findings, our current observations in the retina also reveal a change in the expression of GABAAα subunits, transitioning from GABAAα3 to GABAAα1, during retinal circuit maturation. However, our findings further demonstrate that such GABAAα-subunit changes at inhibitory synapses occur not only on dendrites of CNS neurons but also on their axons. The decline in GABAAα3 and increase in GABAAα1 at BC axons would ensure faster inhibitory modulation of BC excitatory transmission in the adult. This is because GABAAα3Rs are known to confer inhibition with slower response kinetics compared to GABAAα1Rs6, 42, 53, 54. This is likely important for both rod and cone BCs because the GABAAα3->GABAAα1 transition occurs in both BC types. We demonstrate a functional impact when GABAAα3, and thus GABAAα1, are absent in the rod pathway, but the functional consequences of the GABAAα3->GABAAα1 dominance on the cone pathway remains to be determined.
Whether or not the change in GABAAα-subunit expression is also accompanied by modifications of other GABAergic postsynaptic components at BC terminals remains as yet unknown. Developmental alterations in the expression of both receptor subunits and associated scaffolding proteins are known to occur at glutamatergic postsynapses. NMDA receptors switch receptor-types (GluN2B->GluN2A) together with associated scaffolding proteins at a time coincident with synapse maturation3, 50. SAP102, a scaffold that forms complexes with GluN2B receptors is replaced with PSD95, a scaffold of GluN2A receptors3, 55. We previously found that LRRTM4, a postsynaptic organizing protein, is present at mature RBC GABAAα1 synapses43. In the absence of LRRTM4, GABAAα1 accumulation at these synapses is much reduced43. In the CNS, LRRTM4 expression is steeply upregulated from the first postnatal week until P3056 matching the timeline we observe for GABAAα1 expression across RBC terminals. As GABAAα3R clustering on ON-BC terminals is highest around the first postnatal week it is unlikely that LRRTM4 is a common organizing protein of GABAAα3 and GABAAα1-containing BC synapses. Future studies, probably relying on single-BC transcriptomics from early developmental time-points are needed to uncover the identity of organizing protein(s) at early GABAAα3 synapses.
We observed that GABAAα3->GABAAα1R expression transitioned around eye-opening, a correlation that has been observed previously in other CNS regions. The GABAAα5-GABAAα3 switch in thalamic reticular nucleus occurs around the time of eye-opening5 as does the GABAAR subunit switch in the visual cortex52. This common timeline in the expression changes of the α-subunits of GABAARs may suggest distinct roles for the early-expressing and late-expressing GABAAR-types in synapse formation versus plasticity. Indeed, in visual cortical circuits, GABAAα1 is pivotal for experience-dependent plasticity but not for circuit development; the reverse is true for GABAAα3Rs6, 57. For the retina, however, future experiments are needed to determine whether GABAAα1 and GABAAα3R-types could play distinct roles in supporting plasticity in the inner retina and to uncover the underlying developmental mechanisms regulating the GABAAα3->GABAAα1 dominance at inner retinal synapses.
By examining the stereotypically arranged RBC-A17 synapse, we were able to show that GABAAα1R clustering increases at this synapse during development as GABAAα3 clustering declines. This decline at most RBC-A17 synapses is unlikely to be due to a major loss of early connections, but rather reflects a change in receptor composition at the same synapses. Whether or not receptor subunit composition changes at the same synapse in other parts of the CNS and whether GABAR composition at the mature synapse always depends on the earlier presence of another receptor-type, have yet to be explored in detail. Such studies at the resolution of individual inhibitory synapses would augment our understanding of the process of inhibitory synapse maturation. Our findings here support a mechanism by which already established inhibitory synapses could alter receptor composition during circuit maturation to support functional demands of the emerging circuit.
STAR METHODS
RESOURCE AVAILABILITY
Lead contact
Further information and requests for resources and reagents should be directed to and will be fulfilled by the Lead Contact, Mrinalini Hoon (mhoon@wisc.edu).
Materials availability
This study did not generate any unique reagents.
Data and code availability
The ATAC-seq dataset from GABAAα3KO-littermate control retina has been deposited (GEO: GSE180163). The remaining datasets supporting the current study have not been deposited in a public repository because of extremely large file sizes but are available from the corresponding author on request. This study did not generate a unique code.
EXPERIMENTAL MODELS AND SUBJECT DETAILS
All experiments were performed in accordance with the guidelines of the Institutional Animal Care and Use Committees (IACUC) of the University of Washington and the University of Wisconsin-Madison and the National Institutes of Health. Developing rod BCs and Type 6 ON bipolar cells were visualized in the Grm6-tdTomato mouse line39 in which the metabotropic glutamate receptor-6 (mGluR6) promoter drives tdTomato expression in ON bipolar cells. Age matched littermate control and GABAAα3 KO44 mice were utilized for analyses. GABAAα3 KO-littermate mice were analyzed at the adult time-point (> 6 weeks) and at P12 (before eye-opening). To visualize rod BCs and Type 6 ON bipolar cells in the GABAAα3 KO, the GABAAα3 KO line was crossed into the Grm6-tdTomato background. To visualize Type 1 OFF cone bipolar cells in the GABAAα3 KO, the GABAAα3 KO line was crossed into the Vsx1-cerulean mouse line where OFF bipolar cells are specifically labeled by expression of cerulean fluorescent protein28. To target A17 amacrine cells, the Ai9 reporter line (Jackson Laboratory, B6.Cg-Gt(ROSA)26Sortm9(CAG-tdTomato)Hze/J) was crossed into the slc6a5-cre transgenic (GENSAT) to generate the Ai9/slc6a5-cre line. In this line amacrine cells including A17 amacrine cells are fluorescently labeled and can thus be visualized. The Ai9/slc6a5-cre line was used to target and fill A17 before eye-opening (P11) and at a mature time point (>P30). To eliminate GABAAα1 receptors from rod BCs (GABAAα1 cKO), the GABAAα1 floxed line (B6.129(FVB)-Gabra1tm1Geh/J) was crossed into the ON bipolar cell specific Grm6-Cre28 line and the Ai9 reporter line. Cre-expressing BCs in a wildtype background served as littermate controls. GABAAα1 cKO-littermate mice were analyzed at the adult time-point (> 6 weeks) and at P12 (before eye-opening). Mice of both sexes were utilized. For gene expression experiments P6 and P50 C57BL6/J wildtype animals (Jackson Laboratory) were utilized. 2 adult C57BL6/J wildtype animals were used in the Control data-set for AII electrophysiological experiments.
METHOD DETAILS
Immunohistochemistry
Immunolabeling was performed on whole-mount retinas isolated in cold oxygenated mouse artificial cerebrospinal fluid (mACSF, pH 7.4) containing (in mM): 119 NaCl, 2.5 KCl, 2.5 CaCl2, 1.3 MgCl2, 1 NaH2PO4, 11 glucose, and 20 HEPES. Retinas were flattened on a filter paper (Millipore, HABG01300), fixed for 15 mins in 4% paraformaldehyde prepared in mACSF, rinsed in phosphate buffer (PBS) and incubated with primary antibody in blocking solution containing 5% donkey serum and 0.5 Triton X-100 at 4°C for 3-4 days. Antibodies utilized were as follows: anti-PKC (1:1000, mouse, Sigma); anti-lucifer yellow (1:500, rabbit, Invitrogen), anti-LRRTM4 (BC-262) (1:500, rabbit56); anti-GABAAα1 (1:5000, guinea pig, J.M Fritschy13); anti-GABACρ (1:500, rabbit, R. Enz, H. Wassle and S. Haverkamp34); anti-GABAAα3 (1:3000, guinea pig, J.M Fritschy13), anti-GABAAγ2 (1:1000, rabbit, Synaptic Systems), anti-GABAAβ3 (1:500, guinea pig, Synaptic Systems), anti-GlyRα1 (1:500, mouse monoclonal mAb2b, Synaptic Systems), anti-RFP (mouse monoclonal, 1:1000, Abcam), anti-VIAAT (rabbit polyclonal, 1:1000, Synaptic Systems), anti-DsRed (rabbit polyclonal, 1:1000, Clontech). After incubation with primary antibodies, retinas were rinsed in PBS and incubated with anti-isotypic Alexa Fluor (1:1000, Invitrogen) or DyLight (1:1000, Jackson Immunoresearch) secondary antibodies overnight at 4°C. Thereafter retinas were rinsed in PBS and mounted on slides with Vectashield antifade mounting medium (Vector Labs). For generating retina slices, fixed retinas were embedded in agarose (Sigma, low gelling temperature) and sectioned (120 μm) at a Leica Vibratome (VT1000S). The slices were collected in PBS and subsequently processed for immunohistochemistry.
Confocal microscopy and Image analyses
Images were acquired with an Olympus FV 1000 laser scanning confocal microscope and a 1.35 NA 60X oil immersion objective or a Leica SP8 confocal microscope and a 1.4 NA 63X oil immersion objective. Voxel size for acquired images was around 0.05-0.05-0.3 μm (x-y-z). Image stacks were processed using Image J (NIH) and Amira (ThermoFisher Scientific) software. Individual bipolar cell processes were isolated in 3D using the LabelField function in Amira. To isolate the receptor signal within a BC process, the receptor channel was multiplied with the BC mask using the Arithmetic function in Amira. To quantify the amount of receptor expressed within the BC process, we determined the volume occupied by the receptor signal within the BC terminal and expressed it relative to the volume of the BC terminal (% volume occupancy) as previously described28, 43, 58. A threshold was applied to eliminate background pixels and the total volume of receptor signal/pixels above background was thereafter expressed as % occupancy relative to the BC volume28, 43, 58.
To determine contacts between PKC positive rod BCs and A17 amacrine cell varicosities, the A17 amacrine cell varicosities and rod BC axonal boutons were first masked in 3D using the LabelField function in Amira. Using the Arithmetic function, the A17 varicosities mask and the rod BC axon terminal mask were then multiplied to determine the synaptic overlap. The GABA receptor channel was thereafter multiplied with the A17- rod BC synaptic overlap channel to isolate the receptor signal specifically within synaptic overlap. % occupancy of receptor signal within the overlap was determined as described above.
To determine appositions of VIAAT-positive terminals (inhibitory presynapses) across PKC positive rod BC boutons, individual rod BC terminals were first masked in 3D (volume isolation) using the LabelField function in Amira. Thereafter regions of interest were created for each VIAAT positive puncta that had volume overlap (in 3D) with the rod BC mask. The number of such VIAAT positive regions of interest were thereafter summed for each complete rod BC terminal.
Electrophysiology recordings and A17 cell fills
Experiments were conducted on dark-adapted GABAAα3 KO and littermate control mice in a whole-mount preparation for AII amacrine cell recording or a slice (200 μm thick) preparation for determining rod BC puff responses from GABAAα3 KO-littermate control mice and GABAAα1 cKO-littermate control mice. For AII recordings 3 littermate control (7 AII cells) and 2 non-littermate wildtype (C57BL6/J) animals (2 AII cells) were used. For recording light responses from AII amacrine cells, isolated retinas were stored in oxygenated (95% O2/5% CO2) bicarbonate Ames medium (Sigma) at ~32°C. Retinas were mounted photoreceptor-side down on poly-L-lysine coated cover slips. The mounted retina was continuously super fused with warm oxygenated Ames (~8mL/min). Retinal dissections and mounting were conducted exclusively under infrared illumination (>900nm) to preserve visual sensitivity.
Voltage-clamp whole-cell recordings from AII amacrine cells were conducted with electrodes (5-6 MΩ) containing (in mM): 105 Cs methanesulfonate, 10 TEA-Cl, 20 HEPES, 10 EGTA, 2 QX-314, 5 Mg-ATP, 0.5 Tris-GTP and 0.1 Alexa (594) hydrazide (~280 mOsm; pH ~7.3 with CsOH). To isolate excitatory synaptic input, cells were held at the estimated reversal potential for inhibitory input (−68.5 mV). Absolute voltage values were corrected for liquid junction potentials (−8.5 mV). Full field illumination (diameter: 500-560 μm) was delivered to the photoreceptors through a customized condenser from short wavelength (peak power at 405 or 460 nm) LEDs. Light intensities (photons/μm2/s) were converted to photoisomerization rates (R*/photoreceptor/s) using the estimated collecting area of rods (0.5 μm2; Field and Rieke59, the LED emission spectra and the photoreceptor absorption spectra60). Flashes were 10 ms in duration. Electrophysiology example traces in the figures represent the average of ~10 raw responses to the same stimuli. Responses (peak) from individual AII amacrine cells were normalized to the response amplitude at the brightest flash strength (3.2 R*/Rod/flash) and fit with the sigmoid fit function in Igor Pro (Wavemetrics). The flash strength that generates 50% of the maximum AII response amplitude (R50%) was estimated from these fits.
Retinas were embedded in agarose and sliced as previously described28, 43 for rod BC recordings. For rod BC puff recordings 5 adult GABAAα3 KO animals and 6 littermate control animals were used and 3 adult GABAAα1 cKO and 3 littermate control animals were used. Voltage-clamp recordings from rod BCs used pipettes (10–14 MΩ) filled with an intracellular solution containing (in mM) the following: 105 Cs methanesulfonate, 10 TEA-Cl, 20 HEPES, 10 EGTA, 2 QX-314, 5 Mg-ATP, 0.5 Tris-GTP, and 0.1 Alexa-594 hydrazide (~280 mOsm; pH ~7.2 with KOH). To isolate the GABAA and GABAC mediated currents, (1,2,5,6-Tetrahydropyridin-4-yl) methylphosphinic acid (TPMPA, 50 μM; Tocris, Bristol, United Kingdom) and GABAzine (20 μM; Sigma) were added to the perfusion solution. To isolate the GABA mediated inhibitory currents cells were held at the estimated reversal potential for excitatory input ~+10 mV. Absolute voltage values were not corrected for liquid junction potentials (−8.5 mV). GABA was applied using a Picospritzer II (General Valve) connected to a patch pipette (resistance, ~5–7 MΩ). GABA (200 μM) was dissolved in Hepes-buffered Ames medium with 0.1 mM Alexa-488 hydrazide and applied with the puff pipette. The puffing direction and the duration of the 50 ms puff were selected such that the GABA puff completely covered the axon of the rod BC being recorded from. Our previous studies using this technique have confirmed that the GABA puff application at rod BC axon terminals only activates GABA receptors specifically at the axon terminal28, 43. To quantify GABA-evoked currents, peak amplitude was calculated by subtracting the pre-stimulus baseline current from the peak response and averaging across several trials.
A17 cells were targeted with 2-photon microscopy (980 nm) in whole mount retinas61 from the Ai9/slc6a5-cre transgenic line. Two different developmental time points, postnatal day 11 and adult (>4 weeks) were used. A17 cells were injected with 2% Lucifer yellow (LY) prior to fixation with 4% paraformaldehyde (prepared in mACSF) for 15 mins. The retinas were thereafter rinsed with PBS and processed for immunohistochemistry as described above.
3D serial block face serial scanning electron microscopy and rod BC reconstructions
Retinas were immersion fixed in 4% glutaraldehyde in 0.1 M sodium cacodylate buffer (pH 7.4) and were processed and embedded in Durcupan resin as previously described62. A Zeiss 3View Serial block face scanning electron microscope was used to image retinal regions comprising a 2x2 montage of tiles (8600x8600 pixels; ~48μmx48μm) at a section thickness of 50nm. Image stacks were aligned, and rod BCs reconstructed using TrakEM2 module of Image J (NIH). Amira (ThermoFisher Scientific) was used for display of reconstructed cell profiles and associated synaptic profiles.
Rod BC terminals were determined in the EM stack by their characteristic morphology18, 24, 43, 63. Ribbons at rod BC terminals are apposed to two postsynaptic amacrine interneurons, one of which (the A17) makes a reciprocal feedback inhibitory synapse back onto the same rod BC terminal it receives input from18, 24, 43, 63. This feature was used to distinguish A17 profiles at each rod BC ribbon site. The AII partner was recognized through ultrastructural features as previously described43, 64. Electron dense ribbons were easily recognized in EM images allowing demarcation of dyad sites and apposed AII and A17 synaptic partners. Inhibitory synapses were determined by the characteristic accumulation of synaptic vesicles along a defined release site and the presence of a thickening of the pre- and postsynaptic membranes as described previously65. 3 GABAAα3 KO rod BC terminals were reconstructed from 2 adult animal pairs. A P12 wildtype retina dataset from Sinha et al.43 was used to determine the total inhibitory synapse number across rod P12 BC terminals.
Real Time - Quantitative PCR (qPCR) from retina samples
Retinas were collected from 3 adult GABAAα3 KO and littermate control animals. For the developmental comparison, retinas were collected from 3 P6 and P50 C57BL/6J animal pairs. For each animal, both retinas were pooled into one sample. RNA was extracted using DNA/RNA/Protein extraction kit (IBI Scientific). cDNA was reverse-transcribed using M-MLV reverse transcriptase (Promega) and Oligo(dT) Primers (5′–AAGCAGTGGTATCAACGCAGAGTACT30VN-3)66. The same concentration of total RNA was used for each sample. Real-Time qPCRs were run in 10 μL fast reactions (PowerUp SYBR™ Green Master Mix, Applied Biosystems) in a Quant Studio 7 Flex machine (Life Technologies). PCR’s were run in pair-matched batches with GAPDH as a reference gene. Relative Gene Expression was calculated using the ΔΔ Ct method with the Pfaffl correction67. Primers were used from published literature or designed with the NCBI BLAST tool. New primer pairs were verified by band size and Sanger Sequencing of the PCR product. Product sequences obtained were analyzed with UGENE. Determined primer efficiencies are as follows; GAPDH 1.09, GABAAα1 0.91, GABAAα3 0.93, GABACρ 1.05, PCP2 1.03. Primer sequences are as follows: GAPDH68 Forward: GGCCGGTGCTGAGTATGTCG Reverse: TTCTGGGTGGCAGTGATGGC, GABAAα1 Forward: CACCATGAGGTTGACCGTGA Reverse: CTACAACCACTGAACGGGCT, GABAAα3 Forward: GTGACACTCGATCTCACAGGT Reverse: ATATCTGGGGCATGCTTGGG, GABACρ Forward: GAGTTTCCCTGGGGATCACG Reverse: GCCATGGCTTGAACAGCATC, PCP 2 Forward: CAGACCTTCTAGACAAGGCAGG Reverse: TCGTTTCTGCATTCCATCCTTG.
Assay for Transposase-Accessible Chromatin (ATAC)-Seq analyses
One pair of GABAAα3 KO-littermate control animal was used for the ATAC-Seq experiment. Flash-frozen retinas were thawed by addition of 700μl of lysis buffer (10 mM Tris-HCl, 10 mM NaCl, 3 mM MgCl2, 0.1% NP-40), homogenized by trituration 10x with a p1000 pipet set to 500 μl, dounced in an ice cold RNase-free 2mL glass dounce 10x with a loose pestle and 10x with tight pestle, and transferred back to original 1.5mL LoBind tube (Eppendorf). Dounce was washed with an additional 700μl ice cold lysis buffer and buffer was transferred to tube with sample for a final volume of 1.4mL. Sample was then centrifuged in a microcentrifuge at 4°C for 10min at 500 rcf. The pellet was re-eluted in 1.4mL ice cold lysis buffer, transferred to a pre-chilled dounce, homogenized again 12x with tight pestle, and transferred back to 1.5mL LoBind tube. Nuclei were counted on a hemocytometer and the volume of lysis buffer/nuclei suspension needed for 20,000 nuclei was aliquoted into a separate 1.5mL LoBind tube. Samples were centrifuged along with tube containing remaining nuclei (for nuclear localized RNA) at 4°C for 5 min at 500g (RCF). The supernatant was removed from samples and remaining nuclei. Subsequent ATAC libraries were generated with 20,000 nuclei per sample using the Nextera DNA library prep kit (FC-121-1030; Illumina, San Diego, CA USA) according to the protocol described in Buenrostro et al.69, 70. ATAC-Seq libraries were prepared and sequenced on the Illumina NextSeq 500 platform with 75 bp single-end reads.
For ATAC-Seq Read processing and alignment. Demultiplexed FASTQ files were trimmed with Trimmomatic (version 0.33) using the parameter SLIDINGWINDOW:5:3071. Trimmed reads were indexed and aligned to the mouse genome (Mouse GRCm38/mm10 assembly, December 2011) using the Burrows-Wheeler Aligner (bwa) tool72 with the parameters: bwa aln -q 0 -t 4 -n 2 -k 2 -l 32 -e −1 -o 0 and bwa samse -n 5. Tag directories of reads were created using HOMER (version 4.6) makeTagDirectory73. Bed files from processed and aligned sequence reads were extended to 200bp and normalized to 10M reads using BEDtools (version 2.23.0)74 genomeCoverageBED using the -scale parameter before being converted into bigWig format for display on the UCSC genome browser (https://genome.ucsc.edu/).
Western Blot
Retinae from GABAAα3 KO and littermate control mice were homogenized in lysis buffer containing 50mM Tris, 100mM NaCl, 1% Triton X-100, 5mM EDTA, 0.1% SDS, 2.5% glycerol and 1x protease inhibitor cocktail. Both retinas were pooled per animal. Equal amounts of protein samples were run in a 10% SDS gel (Bio-Rad Labs), blotted onto nitrocellulose membranes, incubated with antibodies and visualized by ECL. Primary antibodies utilized were anti-GABAAα1 (clone N95/35 from NeuroMab at 1:5000) and anti-Actin (mouse monoclonal 1:5000 Chemicon).
QUANTIFICATION AND STATISTICAL ANALYSIS
Statistical details of experiments including number of cells (denoted as n) and number of animals (denoted as N) analyzed is provided in the Figure legends. All data are presented as mean ± SEM (standard error of mean) and an unpaired two-tailed T-test was used to determine significance across genotypes.
Supplementary Material
| REAGENT or RESOURCE | SOURCE | IDENTIFIER |
|---|---|---|
| Antibodies | ||
| Mouse monoclonal anti-PKC clone MC5 | Sigma | Catalog # P5704; RRID:AB_477375 |
| Rabbit polyclonal anti-VIAAT | Synaptic Systems | Catalog # 131003; RRID:AB_887869 |
| Rabbit polyclonal anti-GABAAγ2 | Synaptic Systems | Catalog # 224003; RRID:AB_2263066 |
| Guinea pig polyclonal anti-GABAAβ3 | Synaptic Systems | Catalog # 224 404; RRID:AB_2619936 |
| Rabbit polyclonal anti-Dsred | Clontech/Takara Bio | Takara Bio Cat# 632496; RRID:AB_10013483 |
| Mouse monoclonal anti-RFP | Abcam | Catalog # ab65856 |
| Rabbit polyclonal anti-lucifer yellow | Invitrogen | Catalog # A5750; RRID:AB_2536190 |
| Mouse monoclonal anti-GlyRα1 | Synaptic Systems | Catalog # 146111; RRID: AB_887723 |
| Rabbit polyclonal anti-GABAC | 34 | Generated in Heinz Wässle and Joachim Bormann’s Lab. |
| Guinea pig polyclonal anti-GABAAα1 | 13 | Generated in Jean-Marc Fritschy’s Lab |
| Guinea pig polyclonal anti-GABAAα3 | 13 | Generated in Jean-Marc Fritschy’s Lab |
| Rabbit polyclonal anti-LRRTM4 (BC262) | 56 | Generated in Ann Marie Craig’s Lab |
| Mouse anti-GABAAα1 | Neuromab | Catalog # 75-136; RRID:AB_2108811 |
| Mouse anti-Actin | Chemicon/Millipore | Catalog # MAB1501; RRID:AB_2223041 |
| Chemicals, peptides, and recombinant proteins | ||
| Ames | Sigma | A1420 |
| Lucifer yellow | Sigma | L0259 |
| GABAzine (SR-95531) | Sigma | S106 |
| TPMPA | Tocris | 1040 |
| GABA | Sigma | A2129 |
| Alexa 594-hydrazide | Invitrogen | A10438 |
| Alexa 488-hydrazide | Invitrogen | A10436 |
| Vectashield antifade mounting medium | Vector Labs | Catalog# H-1000 |
| M-MLV reverse transcriptase | Promega | M1701 |
| Critical commercial assays | ||
| DNA/RNA/Protein extraction kit | IBI Scientific | IB47702 |
| PowerUp SYBR™ Green Master Mix | Applied Biosystems | A25741 |
| Nextera DNA library prep kit | Illumina | FC-121-1030 |
| Deposited data | ||
| ATAC-seq α3KO/control retina | This paper | GEO: GSE180163 |
| Experimental models: Organisms/strains | ||
| Mouse: GABAAα3 knockout | U. Rudolph44 | N/A |
| Mouse: C57BL/6J | Jackson Labs | JAX Stock No: 000664 |
| Mouse: Grm6-tdtomato | Rachel Wong39 | N/A |
| Mouse: Ai9 reporter; B6.Cg-Gt(ROSA)26Sortm9(CAG-tdTomato)Hze/J strain | Jackson Labs | JAX Stock No: 007909 |
| Mouse: slc6a5-Cre | Allen Brain Institute (GENSAT) | N/A |
| Mouse: Grm6-Cre | Rachel Wong28 | N/A |
| Mouse: Vsx1-cerulean | Rachel Wong28 | N/A |
| Mouse: GABAAα1 floxed; B6.129(FVB)-Gabra1tm1Geh/J | Jackson Labs | JAX Stock No: 004318 |
| Oligonucleotides | ||
| GAPDH forward: GGCCGGTGCTGAGTATGTCG | 68 | N/A |
| GAPDH reverse: TTCTGGGTGGCAGTGATGGC | 68 | N/A |
| GABAAα1 Forward: CACCATGAGGTTGACCGTGA | This paper | N/A |
| GABAAα1 Reverse: CTACAACCACTGAACGGGCT | This paper | N/A |
| GABAAα3 Forward: GTGACACTCGATCTCACAGGT | This paper | N/A |
| GABAAα3 Reverse: ATATCTGGGGCATGCTTGGG | This paper | N/A |
| GABACρ Forward: GAGTTTCCCTGGGGATCACG | This paper | N/A |
| GABACρ Reverse: GCCATGGCTTGAACAGCATC | This paper | N/A |
| PCP2 Forward: CAGACCTTCTAGACAAGGCAGG | This paper | N/A |
| PCP2 Reverse: TCGTTTCTGCATTCCATCCTTG | This paper | N/A |
| Software and algorithms | ||
| BEDtools (version 2.23.0) | 74 | https://github.com/arq5x/bedtools2 |
| HOMER (version 4.6) | 73 | http://homer.ucsd.edu/homer/ |
| Trimmomatic (version 0.33) | 71 | http://www.usadellab.org/cms/?page=trimmomatic |
| IGOR Pro | WaveMetrics | https://www.wavemetrics.com/ |
| MATLAB | Mathworks | https://ch.mathworks.com/products/matlab |
| Symphony | Symphony-DAS | https://github.com/symphony-das |
| ImageJ | NIH | https://imagej.nih.gov/ij/ |
| Amira | ThermoFisher Scientific | https://www.fei.com/software/amira/ |
HIGHLIGHTS.
GABAA synapses on bipolar cell axons change receptor types before eye-opening.
GABAAα3 recruits GABAAα1 and LRRTM4 to inhibitory synapses on bipolar cell axons.
Early GABAAα3 is required for functional GABAA synapses on mature bipolar cell axons.
Early GABAAα3 expression regulates organization of rod bipolar cell ribbon synapses.
ACKNOWLEDGEMENTS
This work was supported by NIH Grants EY031677 (to M.H.), EY10699 (to R.O.W.), EY026070 (to R.S.), EY028111 (to F.R.), Vision Core Grant EY01730 (to M. Neitz), Core grant for Vision Research (P30EY016665), the McPherson Eye Research Institute’s Rebecca Meyer Brown/Retina Research Foundation Professorship (to M.H), and an unrestricted Grant from Research to Prevent Blindness, Inc. to UW Madison Department of Ophthalmology. We thank J.M. Fritschy, R. Enz, S. Haverkamp and H. Wässle for generously providing GABAA and GABAC receptor antibodies, T.J. Siddiqui for generously providing the LRRTM4 antibody and E. Parker, M. Zhang and K. Oda for expert technical assistance.
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
DECLARATION OF INTERESTS
The authors declare no competing interests.
REFERENCES
- 1.Sudhof TC (2018). Towards an Understanding of Synapse Formation. Neuron 100, 276–293. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Gamlin CR, Yu WQ, Wong ROL, and Hoon M (2018). Assembly and maintenance of GABAergic and Glycinergic circuits in the mammalian nervous system. Neural Dev 13, 12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Paoletti P, Bellone C, and Zhou Q (2013). NMDA receptor subunit diversity: impact on receptor properties, synaptic plasticity and disease. Nat Rev Neurosci 14, 383–400. [DOI] [PubMed] [Google Scholar]
- 4.Lohmann C, and Kessels HW (2014). The developmental stages of synaptic plasticity. J Physiol 592, 13–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Pangratz-Fuehrer S, Sieghart W, Rudolph U, Parada I, and Huguenard JR (2016). Early postnatal switch in GABAA receptor alpha-subunits in the reticular thalamic nucleus. J Neurophysiol 115, 1183–1195. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Bosman LW, Rosahl TW, and Brussaard AB (2002). Neonatal development of the rat visual cortex: synaptic function of GABAA receptor alpha subunits. J Physiol 545, 169–181. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Flint AC, Maisch US, Weishaupt JH, Kriegstein AR, and Monyer H (1997). NR2A subunit expression shortens NMDA receptor synaptic currents in developing neocortex. J Neurosci 17, 2469–2476. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Sheng M, Cummings J, Roldan LA, Jan YN, and Jan LY (1994). Changing subunit composition of heteromeric NMDA receptors during development of rat cortex. Nature 368, 144–147. [DOI] [PubMed] [Google Scholar]
- 9.Malosio ML, Marqueze-Pouey B, Kuhse J, and Betz H (1991). Widespread expression of glycine receptor subunit mRNAs in the adult and developing rat brain. EMBO J 10, 2401–2409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Fritschy JM, Paysan J, Enna A, and Mohler H (1994). Switch in the expression of rat GABAA-receptor subtypes during postnatal development: an immunohistochemical study. J Neurosci 14, 5302–5324. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Laurie DJ, Wisden W, and Seeburg PH (1992). The distribution of thirteen GABAA receptor subunit mRNAs in the rat brain. III. Embryonic and postnatal development. J Neurosci 12, 4151–4172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Liu Q, and Wong-Riley MT (2004). Developmental changes in the expression of GABAA receptor subunits alpha1, alpha2, and alpha3 in the rat pre-Botzinger complex. J Appl Physiol (1985) 96, 1825–1831. [DOI] [PubMed] [Google Scholar]
- 13.Fritschy JM, and Mohler H (1995). GABAA-receptor heterogeneity in the adult rat brain: differential regional and cellular distribution of seven major subunits. J Comp Neurol 359, 154–194. [DOI] [PubMed] [Google Scholar]
- 14.McKernan RM, and Whiting PJ (1996). Which GABAA-receptor subtypes really occur in the brain? Trends Neurosci 19, 139–143. [DOI] [PubMed] [Google Scholar]
- 15.Pirker S, Schwarzer C, Wieselthaler A, Sieghart W, and Sperk G (2000). GABA(A) receptors: immunocytochemical distribution of 13 subunits in the adult rat brain. Neuroscience 101, 815–850. [DOI] [PubMed] [Google Scholar]
- 16.Fink AJ, Croce KR, Huang ZJ, Abbott LF, Jessell TM, and Azim E (2014). Presynaptic inhibition of spinal sensory feedback ensures smooth movement. Nature 509, 43–48. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.McGann JP (2013). Presynaptic inhibition of olfactory sensory neurons: new mechanisms and potential functions. Chem Senses 38, 459–474. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Hoon M, Okawa H, Della Santina L, and Wong RO (2014). Functional architecture of the retina: development and disease. Prog Retin Eye Res 42, 44–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Eggers ED, McCall MA, and Lukasiewicz PD (2007). Presynaptic inhibition differentially shapes transmission in distinct circuits in the mouse retina. J Physiol 582, 569–582. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Eggers ED, and Lukasiewicz PD (2011). Multiple pathways of inhibition shape bipolar cell responses in the retina. Vis Neurosci 28, 95–108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Sagdullaev BT, McCall MA, and Lukasiewicz PD (2006). Presynaptic inhibition modulates spillover, creating distinct dynamic response ranges of sensory output. Neuron 50, 923–935. [DOI] [PubMed] [Google Scholar]
- 22.Pan F, Toychiev A, Zhang Y, Atlasz T, Ramakrishnan H, Roy K, Volgyi B, Akopian A, and Bloomfield SA (2016). Inhibitory masking controls the threshold sensitivity of retinal ganglion cells. J Physiol 594, 6679–6699. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Grimes WN, Zhang J, Tian H, Graydon CW, Hoon M, Rieke F, and Diamond JS (2015). Complex inhibitory microcircuitry regulates retinal signaling near visual threshold. J Neurophysiol 114, 341–353. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Wassle H (2004). Parallel processing in the mammalian retina. Nat Rev Neurosci 5, 747–757. [DOI] [PubMed] [Google Scholar]
- 25.Demb JB, and Singer JH (2015). Functional Circuitry of the Retina. Annu Rev Vis Sci 1, 263–289. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Fletcher EL, Koulen P, and Wassle H (1998). GABAA and GABAC receptors on mammalian rod bipolar cells. J Comp Neurol 396, 351–365. [DOI] [PubMed] [Google Scholar]
- 27.Koulen P, Brandstatter JH, Enz R, Bormann J, and Wassle H (1998). Synaptic clustering of GABA(C) receptor rho-subunits in the rat retina. Eur J Neurosci 10, 115–127. [DOI] [PubMed] [Google Scholar]
- 28.Hoon M, Sinha R, Okawa H, Suzuki SC, Hirano AA, Brecha N, Rieke F, and Wong RO (2015). Neurotransmission plays contrasting roles in the maturation of inhibitory synapses on axons and dendrites of retinal bipolar cells. Proc Natl Acad Sci U S A 112, 12840–12845. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Eggers ED, and Lukasiewicz PD (2006). GABA(A), GABA(C) and glycine receptor-mediated inhibition differentially affects light-evoked signalling from mouse retinal rod bipolar cells. J Physiol 572, 215–225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Singer JH, and Diamond JS (2003). Sustained Ca2+ entry elicits transient postsynaptic currents at a retinal ribbon synapse. J Neurosci 23, 10923–10933. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Chavez AE, Singer JH, and Diamond JS (2006). Fast neurotransmitter release triggered by Ca influx through AMPA-type glutamate receptors. Nature 443, 705–708. [DOI] [PubMed] [Google Scholar]
- 32.Chavez AE, Grimes WN, and Diamond JS (2010). Mechanisms underlying lateral GABAergic feedback onto rod bipolar cells in rat retina. J Neurosci 30, 2330–2339. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Moss SJ, and Smart TG (2001). Constructing inhibitory synapses. Nat Rev Neurosci 2, 240–250. [DOI] [PubMed] [Google Scholar]
- 34.Wassle H, Koulen P, Brandstatter JH, Fletcher EL, and Becker CM (1998). Glycine and GABA receptors in the mammalian retina. Vision Res 38, 1411–1430. [DOI] [PubMed] [Google Scholar]
- 35.Koulen P, Sassoe-Pognetto M, Grunert U, and Wassle H (1996). Selective clustering of GABA(A) and glycine receptors in the mammalian retina. J Neurosci 16, 2127–2140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Schubert T, Hoon M, Euler T, Lukasiewicz PD, and Wong RO (2013). Developmental regulation and activity-dependent maintenance of GABAergic presynaptic inhibition onto rod bipolar cell axonal terminals. Neuron 78, 124–137. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Aldiri I, Xu B, Wang L, Chen X, Hiler D, Griffiths L, Valentine M, Shirinifard A, Thiagarajan S, Sablauer A, et al. (2017). The Dynamic Epigenetic Landscape of the Retina During Development, Reprogramming, and Tumorigenesis. Neuron 94, 550–568 e510. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Jorstad NL, Wilken MS, Grimes WN, Wohl SG, VandenBosch LS, Yoshimatsu T, Wong RO, Rieke F, and Reh TA (2017). Stimulation of functional neuronal regeneration from Muller glia in adult mice. Nature 548, 103–107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Kerschensteiner D, Morgan JL, Parker ED, Lewis RM, and Wong RO (2009). Neurotransmission selectively regulates synapse formation in parallel circuits in vivo. Nature 460, 1016–1020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Grimes WN, Zhang J, Graydon CW, Kachar B, and Diamond JS (2010). Retinal parallel processors: more than 100 independent microcircuits operate within a single interneuron. Neuron 65, 873–885. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Haverkamp S, and Wassle H (2000). Immunocytochemical analysis of the mouse retina. J Comp Neurol 424, 1–23. [PubMed] [Google Scholar]
- 42.Vicini S, Ferguson C, Prybylowski K, Kralic J, Morrow AL, and Homanics GE (2001). GABA(A) receptor alpha1 subunit deletion prevents developmental changes of inhibitory synaptic currents in cerebellar neurons. J Neurosci 21, 3009–3016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Sinha R, Siddiqui TJ, Padmanabhan N, Wallin J, Zhang C, Karimi B, Rieke F, Craig AM, Wong RO, and Hoon M (2020). LRRTM4: A Novel Regulator of Presynaptic Inhibition and Ribbon Synapse Arrangements of Retinal Bipolar Cells. Neuron 105, 1007–1017 e1005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Yee BK, Keist R, von Boehmer L, Studer R, Benke D, Hagenbuch N, Dong Y, Malenka RC, Fritschy JM, Bluethmann H, et al. (2005). A schizophrenia-related sensorimotor deficit links alpha 3-containing GABAA receptors to a dopamine hyperfunction. Proc Natl Acad Sci U S A 102, 17154–17159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Greferath U, Grunert U, Fritschy JM, Stephenson A, Mohler H, and Wassle H (1995). GABAA receptor subunits have differential distributions in the rat retina: in situ hybridization and immunohistochemistry. J Comp Neurol 353, 553–571. [DOI] [PubMed] [Google Scholar]
- 46.Greferath U, Muller F, Wassle H, Shivers B, and Seeburg P (1993). Localization of GABAA receptors in the rat retina. Vis Neurosci 10, 551–561. [DOI] [PubMed] [Google Scholar]
- 47.Woods SM, Mountjoy E, Muir D, Ross SE, and Atan D (2018). A comparative analysis of rod bipolar cell transcriptomes identifies novel genes implicated in night vision. Sci Rep 8, 5506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Studer R, von Boehmer L, Haenggi T, Schweizer C, Benke D, Rudolph U, and Fritschy JM (2006). Alteration of GABAergic synapses and gephyrin clusters in the thalamic reticular nucleus of GABAA receptor alpha3 subunit-null mice. Eur J Neurosci 24, 1307–1315. [DOI] [PubMed] [Google Scholar]
- 49.Vicini S, Wang JF, Li JH, Zhu WJ, Wang YH, Luo JH, Wolfe BB, and Grayson DR (1998). Functional and pharmacological differences between recombinant N-methyl-D-aspartate receptors. J Neurophysiol 79, 555–566. [DOI] [PubMed] [Google Scholar]
- 50.Dumas TC (2005). Developmental regulation of cognitive abilities: modified composition of a molecular switch turns on associative learning. Prog Neurobiol 76, 189–211. [DOI] [PubMed] [Google Scholar]
- 51.Yashiro K, and Philpot BD (2008). Regulation of NMDA receptor subunit expression and its implications for LTD, LTP, and metaplasticity. Neuropharmacology 55, 1081–1094. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Heinen K, Bosman LW, Spijker S, van Pelt J, Smit AB, Voorn P, Baker RE, and Brussaard AB (2004). GABAA receptor maturation in relation to eye opening in the rat visual cortex. Neuroscience 124, 161–171. [DOI] [PubMed] [Google Scholar]
- 53.Gingrich KJ, Roberts WA, and Kass RS (1995). Dependence of the GABAA receptor gating kinetics on the alpha-subunit isoform: implications for structure-function relations and synaptic transmission. J Physiol 489 (Pt 2), 529–543. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Barberis A, Mozrzymas JW, Ortinski PI, and Vicini S (2007). Desensitization and binding properties determine distinct alpha1beta2gamma2 and alpha3beta2gamma2 GABA(A) receptor-channel kinetic behavior. Eur J Neurosci 25, 2726–2740. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Sans N, Petralia RS, Wang YX, Blahos J 2nd, Hell JW, and Wenthold RJ (2000). A developmental change in NMDA receptor-associated proteins at hippocampal synapses. J Neurosci 20, 1260–1271. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Siddiqui TJ, Tari PK, Connor SA, Zhang P, Dobie FA, She K, Kawabe H, Wang YT, Brose N, and Craig AM (2013). An LRRTM4-HSPG complex mediates excitatory synapse development on dentate gyrus granule cells. Neuron 79, 680–695. [DOI] [PubMed] [Google Scholar]
- 57.Fagiolini M, Fritschy JM, Low K, Mohler H, Rudolph U, and Hensch TK (2004). Specific GABAA circuits for visual cortical plasticity. Science 303, 1681–1683. [DOI] [PubMed] [Google Scholar]
- 58.Hoon M, Sinha R, and Okawa H (2017). Using Fluorescent Markers to Estimate Synaptic Connectivity In Situ. Methods Mol Biol 1538, 293–320. [DOI] [PubMed] [Google Scholar]
- 59.Field GD, and Rieke F (2002). Nonlinear signal transfer from mouse rods to bipolar cells and implications for visual sensitivity. Neuron 34, 773–785. [DOI] [PubMed] [Google Scholar]
- 60.Govardovskii VI, Fyhrquist N, Reuter T, Kuzmin DG, and Donner K (2000). In search of the visual pigment template. Vis Neurosci 17, 509–528. [DOI] [PubMed] [Google Scholar]
- 61.Grimes WN, Hoon M, Briggman KL, Wong RO, and Rieke F (2014). Cross-synaptic synchrony and transmission of signal and noise across the mouse retina. Elife 3, e03892. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Della Santina L, Kuo SP, Yoshimatsu T, Okawa H, Suzuki SC, Hoon M, Tsuboyama K, Rieke F, and Wong ROL (2016). Glutamatergic Monopolar Interneurons Provide a Novel Pathway of Excitation in the Mouse Retina. Curr Biol 26, 2070–2077. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Tsukamoto Y, and Omi N (2017). Classification of Mouse Retinal Bipolar Cells: Type-Specific Connectivity with Special Reference to Rod-Driven AII Amacrine Pathways. Front Neuroanat 11, 92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Gamlin CR, Zhang C, Dyer MA, and Wong ROL (2020). Distinct Developmental Mechanisms Act Independently to Shape Biased Synaptic Divergence from an Inhibitory Neuron. Curr Biol 30, 1258–1268 e1252. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Gray EG (1969). Electron microscopy of excitatory and inhibitory synapses: a brief review. Prog Brain Res 31, 141–155. [DOI] [PubMed] [Google Scholar]
- 66.Picelli S, Faridani OR, Bjorklund AK, Winberg G, Sagasser S, and Sandberg R (2014). Full-length RNA-seq from single cells using Smart-seq2. Nat Protoc 9, 171–181. [DOI] [PubMed] [Google Scholar]
- 67.Pfaffl MW (2001). A new mathematical model for relative quantification in real-time RT-PCR. Nucleic Acids Res 29, e45. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Nickells RW, and Pelzel HR (2015). Tools and resources for analyzing gene expression changes in glaucomatous neurodegeneration. Exp Eye Res 141, 99–110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Buenrostro JD, Giresi PG, Zaba LC, Chang HY, and Greenleaf WJ (2013). Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nat Methods 10, 1213–1218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Buenrostro JD, Wu B, Chang HY, and Greenleaf WJ (2015). ATAC-seq: A Method for Assaying Chromatin Accessibility Genome-Wide. Curr Protoc Mol Biol 109, 21 29 21–21 29 29. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Bolger AM, Lohse M, and Usadel B (2014). Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Li H, and Durbin R (2010). Fast and accurate long-read alignment with Burrows-Wheeler transform. Bioinformatics 26, 589–595. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Heinz S, Benner C, Spann N, Bertolino E, Lin YC, Laslo P, Cheng JX, Murre C, Singh H, and Glass CK (2010). Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol Cell 38, 576–589. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Quinlan AR, and Hall IM (2010). BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The ATAC-seq dataset from GABAAα3KO-littermate control retina has been deposited (GEO: GSE180163). The remaining datasets supporting the current study have not been deposited in a public repository because of extremely large file sizes but are available from the corresponding author on request. This study did not generate a unique code.







