Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Apr 16.
Published in final edited form as: Curr Opin Neurobiol. 2008 May 29;18(1):77–83. doi: 10.1016/j.conb.2008.05.008

GABA and neuroligin signaling: linking synaptic activity and adhesion in inhibitory synapse development

Z Josh Huang 1, Peter Scheiffele 2
PMCID: PMC3988756  NIHMSID: NIHMS56996  PMID: 18513949

Abstract

GABA-mediated synaptic inhibition is crucial in neural circuit operations. In mammalian brains, the development of inhibitory synapses and innervation patterns is often a prolonged postnatal process, regulated by neural activity. Emerging evidence indicates that GABA acts beyond inhibitory transmission and regulates inhibitory synapse development. Indeed, GABAA receptors not only function as chloride channels that regulate membrane voltage and conductance but also play structural roles in synapse maturation and stabilization. The link from GABAA receptors to post- and pre- synaptic adhesion is likely mediated, in part, by neuroligin-reurexin interactions, which are potent in promoting GABAergic synapse formation. Therefore, similar to glutamate signaling at excitatory synapse, GABA signaling may coordinate maturation of pre- and post- synaptic sites at inhibitory synapses. Defining the many steps from GABA signaling to receptor trafficking/stability and neuroligin function will provide further mechanistic insights into activity-dependent development and possibly plasticity of inhibitory synapses.

Introduction

In many areas of the vertebrate brain, neural circuits rely on inhibition mediated by γ-aminobutyric acid (GABA) from diverse cell types to control the spatiotemporal patterns of electrical signaling [1]. The inhibitory output is distributed in the network through GABAergic axons and synapses, which constitute elaborate and cell type-specific inhibitory innervation patterns [2]. GABAergic neurons are generated during mid-embryonic stages and influence many aspects of early neural development. The subsequent maturation of GABAergic synapses and innervation patterns into a potent inhibitory network is often a protracted process, extending well into postnatal life, and is regulated by neural activity and experience [36]. Such activity-dependent development of inhibitory synapses and innervation patterns is a major component of neural circuit assembly, yet the underlying cellular and molecular mechanisms are poorly understood.

As key mediators of neural activity, neurotransmitters are particularly well suited to couple synaptic transmission with synaptic growth and refinement [7,8]. Although the initial steps of nascent synapse formation are likely independent of neurotransmitter release [9,10], there is substantial evidence that neurotransmitters and synaptic activity regulate nearly all aspects of synapse formation [1113], plasticity [1417], and the development of axonal and dendritic arbors [7,18,19]. Such synaptotrophic functions have been most extensively studied for glutamate, the main excitatory neurotransmitter in the brain, and for acetylcholine at the neuromuscular junction [20,21]. Therefore, in addition to synaptic transmission, excitatory transmitters also function as trophic signals for regulating synapse formation, maturation, and plasticity.

Initially discovered as an inhibitory transmitter, GABA has since been implicated in multiple processes of neural development, from cell proliferation to circuit formation [22]. The trophic effects of GABA on neuronal migration and neurite growth during the embryonic and perinatal period are largely explained by its depolarizing action in immature neurons, which triggers calcium influx and signaling [23,24,25]. In fact, GABA signaling in the neonatal brain represents an early form of neural communication, which precedes and promotes the formation of glutamatergic synapses [2628]. GABA signals are transduced through two general classes of receptors, the metabotropic GABAB receptors and ionotropic GABAARs which are permeable to chloride ions. The early, depolarizing function of GABA results from chloride ion efflux through the GABAA receptor. During the postnatal period, the up-regulation of the chloride transporter KCC2 in neurons results in increased extrusion of intracellular chloride [29], and GABA assumes its classic role as an inhibitory transmitter [30]. Recently, several studies suggest that, in addition to mediating synaptic inhibition in the more mature circuits, GABA signaling may promote and coordinate pre- and post- synaptic maturation during activity-dependent development of inhibitory synapses. In this review, we will first discuss recent genetic studies that revealed critical roles for both GABA and GABAA receptor complexes in inhibitory synapse development. We will then explore the possibility that the neuroligin-neurexin trans-synaptic adhesion system participates in this activity-dependent inhibitory synapse formation and maturation processes.

GABA signaling: linking inhibitory transmission to synaptic wiring

The rodent visual and somatosensory cortices are excellent model systems for studying activity-dependent development of GABAergic connectivity. For example, the maturation of inhibitory innervation [3,4] and transmission [4,6] proceeds well into postnatal weeks and is regulated by sensory experience [3,6,31]. Interestingly, the maturation of many features of interneuron axonal arbors and inhibitory synapses can be recapitulated in cortical organotypic cultures [32] and is strongly regulated by neuronal activity [3,33]. By genetic knockdown of GABA synthesis in cortical organotypic cultures and in vivo, a recent study implicates GABA signaling itself in the development of inhibitory synapses [34]. Unlike glutamate which is both the precursor and product of many essential metabolic and signaling processes in the cell, GABA can only be synthesized by two glutamate decarboxylases (GAD67 and GAD65), and the main function of GABA is intercellular signaling [35]. GAD67 is the rate-limiting enzyme and influences cellular GABA contents in a dosage dependent manner [36,37]. Knockdown of GAD67 in single GABAergic interneurons, which should have minimum impact on circuit activity levels, results in profound cell autonomous deficits in synapse formation, axon branching, and innervation field; such deficits were partially rescued by blocking GABA re-uptake or enhancing GABAA or GABAB receptor function [34]. These results demonstrate that GABA acts beyond its classic role in inhibitory transmission in the adolescent brain and regulates the maturation of inhibitory synapses and innervation patterns (Figure 1), thus revealing a new facet of GABA function distinct from its early tropic action in neonatal brain.

Figure 1.

Figure 1

GAD67 and GABA act beyond inhibitory transmission and regulate inhibitory synapse development. (a) GABA signaling may regulate the morphogenesis (e.g. growth and stability) of inhibitory synapses. (b) Since synapse formation is an integral part of axon growth and branching, activity-dependent GABA signaling may further influence the development of GABAergic axon arbor and innervation pattern.

Structural role of GABAA receptors: coupling synaptic transmission to synapse maturation and stability

A second set of findings that highlights a role for GABA signaling in the development of inhibitory synapses come from the analysis of knock-out mice lacking individual subunits of the GABAA receptor (GABAAR). GABAARs are heteropentameric chloride channels composed of several classes of subunits [38]. Although over 19 subunits have been identified, giving rise to a large number of possible subunit combinations, the vast majority of GABAARs consist of α, β, and γ2 subunits in a 2:2:1 stoichiometry. In the mature brain, GABAARs are primarily localized at postsynaptic and extrasynaptic membranes where they mediate phasic and tonic inhibition, respectively.

Purkinje cells in the mouse cerebellum provide a unique opportunity to test the role of GABAAR in synapse formation as deletion of the α1 subunit gene results in a complete loss of functional GABAARs by postnatal day 18 [39]. Purkinje cells are themselves GABAergic neurons but also receive two types of GABAergic inputs: the axo-somatic synapses from basket interneurons and the axo-dendritic synapses from stellate interneurons, both with GABAARs containing the α1 subunit. In α1−/− mice, GABAergic terminals from stellate axons are initially formed normally onto the Purkinje dendritic shaft. However, starting from postnatal day 7, synapse formation and stabilization on the dendritc shaft is severely perturbed [40]. Instead, the stellate cell terminals form aberrant and mismatched contacts with postsynaptic specialization on the spines of Purkinje dendrites. These results suggest that initial steps of GABAergic synapse formation can proceed in the absence of α1, but synapse maturation and maintenance require postsynaptic GABAARs. Notably, basket cell synapses on the Purkinje cell soma are not altered in α1−/− mice. Therefore, the two presynaptic inputs differ in their molecular requirements for synapse formation and stabilization onto Purkinje cells, even with respect to the postsynaptic GABAAR that is common to both types of synapses.

Similar conclusions regarding the role of GABAARs in synapse formation were derived from studies examining the γ2 subunit. The γ2 subunit is essential for accumulation of cell surface GABAARs at postsynaptic sites [41,42]. Interestingly, acute suppression of γ2 expression in cultured hippocampal neurons not only disrupts GABAAR clustering but also results in a profound reduction of GABAergic innervation of γ2 deficient neurons [43,44]. Moreover, when palmitoylation of the γ2 subunit was suppressed by knockdown of the DHHC-family palmitoyltransferase GODZ, trafficking of GABAARs to postsynaptic sites was perturbed and GABAergic innervation was reduced [43]. Interestingly, no comparable presynaptic defect is evident in neuron cultures and brain sections from germ-line γ2−/− mice, which exhibit a uniform deficit in postsynaptic GABAARs in all neurons. These data suggest the presence of a retrograde signal that allows GABAergic axons to preferentially innervate target neurons expressing higher levels of GABAAR at the cell surface. Because both presynaptic GABA and postsynaptic GABAA receptors influence GABAergic synapse development, a simple hypothesis is that activity-dependent GABA signaling promotes the differentiation of pre- and post- synaptic sites, and coordinates the maturation and stabilization of inhibitory synapses. The failure to form and stabilize presynaptic terminals after postsynaptic loss of GABAARs strongly suggests the presence of a retrograde signal that is regulated by synaptic activity or by association with postsynaptic GABAARs. Amongst the molecular mechanisms that may contribute to such an activity-regulated trans-synaptic signal, the neuroligin and neurexin complex represents one of the plausible candidates.

Trans-synaptic signaling by the neuroligin-neurexin adhesion system

The neuroligin-neurexin complex is a heterophilic adhesion system broadly expressed in the central nervous system [45]. Both neuroligins and neurexins are encoded by multiple genes and alternative splicing generates further isoform diversity from each gene. Cell biological studies have revealed potent “synaptogenic” or synapse-organizing activities for these proteins (recently reviewed in [46,47]. Postsynaptic neuroligins promote assembly of functional presynaptic specializations in axons. Conversely, presynaptic neurexins - through interaction with neuroligins - recruit postsynaptic scaffolding proteins and neurotransmitter receptors in dendrites.

While neuroligin-neurexin complexes are common building blocks of glutamatergic and GABAergic synapses and are required for normal glutamatergic and GABAergic transmission, the morphological loss of function data available so far most strongly support their critical roles in the organization of GABAergic synapses. Triple knockout mice lacking the three alpha-neurexin transcripts show a 50% reduction in the density of GABAergic synapses in the brainstem [48]. While triple knock-out animals die at birth, an analysis of more mature neuronal circuits could be performed in double knockout mice, some of which reach adulthood. In these mice, GABAergic synapse density is reduced by 30% whereas glutamatergic synapse density is apparently unchanged, although principal neurons display shortened dendritic branches and lower spine density compared to wild-type mice [49]. As for the neuroligins, mice lacking the three major isoforms (NL1,2, and 3) show perinatal lethality. Although there is only a relatively small (15–20%) reduction in the number synapses in the brain stem of these mutant mice, they show a severe loss of GABAARs and the scaffolding protein gephyrin from postsynaptic sites [50]. These studies uncovered a critical function of neuroligin proteins in postsynaptic differentiation in vivo, especially at GABAergic synapses.

More recent studies began to dissect potential synapse-specific functions of individual neuroligin and neurexin isoforms. In cell culture assays, alpha- and beta-neurexin splice variants differ significantly in their biochemical interactions with neuroligins as well as in their ability to organize glutamatergic versus GABAergic postsynaptic structures [5154]. This selectivity might reflect a trans-synaptic adhesive code regulated by gene and splice variant choice of the pre- and postsynaptic partners. Further analysis in the neuroligin single knock-out mice strongly support an important function of the NL2 isoform at GABAergic synapses. Layer 2/3 neurons in acute cortical slices from NL2 −/− mice show a selective impairment of GABAergic transmission whereas glutamatergic transmission is normal. In addition, NL2−/− mice display a selective decrease in the number of inhibitory synapses in the postnatal neocortex [55]. Overexpression of NL2 in cultured neurons increases the density of glutamatergic and GABAergic terminals but has a higher activity towards GABAergic axons [56]. At a functional level, NL2 overexpression selectively increases the amplitude of inhibitory postsynaptic currents suggesting that the postsynaptic coupling of this neuroligin isoform is specific for inhibitory synapses [55]. Notably, this overexpression-induced increase in GABAergic transmission is blocked by pharmacological inhibition of network activity in the culture. Therefore, neuronal and synaptic activity might either regulate the presynaptic response to NL2 or postsynaptic stabilization induced by NL2.

Taken all together, there is substantial evidence to indicate that both neuroligins and GABAA receptors play critical roles in the maturation of postsynaptic specializations and the differentiation and stabilization of presynaptic terminals at inhibitory synapses. These observations raise an obvious question: how do GABA/GABAAR-mediated synaptic signaling and neuroligin/neurexin-mediated synaptic adhesion interact and cooperate to regulate activity-dependent development of inhibitory synapse?

From GABAA receptors to synaptic adhesion and activity-dependent retrograde signaling

It is currently unknown at what stage of the biosynthetic pathway GABAARs first interact with NLs, and how such interactions might be regulated. In the plasma membrane, GABAARs exist as dispersed populations and as synaptic or extrasynaptic clusters, depending on their subunit compositions; lateral diffusion in the plasma membrane allows continual exchange among these groups [57]. Pentameric GABAARs are assembled in the endoplasmic reticulum (ER) from appropriate subunits, and are delivered into the plasma membrane through a highly regulated trafficking process. Most GABAARs are first delivered to extrasynaptic locations, and their subsequent diffusion and trapping increase their accumulation at postsynaptic sites. The scaffolding protein gephyrin promotes the synaptic targeting of GABAARs in part by direct binding to the α2 subunit [58], and the synaptic localization and function of gephyrin can be regulated by neural activity [59].

In one model (Figure 2), NL2 and synaptic GABAARs would stabilize each other, either through intracellular reciprocal interactions aided by scaffolding proteins such as gephyrin or through extracellular cis interactions. In addition, GABA activation of GABAARs might further stabilize GABAARs at synapses through as yet unknown structural or signaling mechanisms. Such activity- and GABA-mediated stabilization of GABAARs might further increase the levels of NL2 at postsynaptic sites; this, in turn, would stabilize the presynaptic terminals through trans-synaptic interactions with neurexins. Evidence consistent with this model include: 1) in vitro studies demonstrated a co-aggregation of NL2 and the GABAAR γ2 subunit in heterologous cells [60]; 2) the residence time of GABAARs on the plasma membrane and their targeting to synapses is regulated by synaptic activity [61]; 3) pharmacological blockade of neuronal activity in cultured neurons diminish the synaptogenic activity of NL2 [55]; 4) reduced GABA synthesis and release result in a reduction of inhibitory synapses [34]. Moreover, there is precedent for such mechanisms in activity-dependent recruitment of transmitter receptor and trans-synaptic signaling at glutamatergic synapses. Local spontaneous activity and glutamate release reduce diffusion exchange of GluR1 between synaptic and extrasynaptic domains, resulting in postsynaptic accumulation of GluR1 [62]. In addition, PSD-95 and NL1 retrogradely modulate presynaptic release probability and may coordinate post- and pre- synaptic morphological changes [63,64]. It remains to be seen whether analogous mechanisms for GABA and NL2 signaling exist at inhibitory synapses.

Figure 2.

Figure 2

A hypothetical model depicting how GABA-GABAAR signaling and neuroligin-neurexin adhesion may interact and cooperate to regulate the development of inhibitory synapses. Pentameric GABAARs are assembled in the endoplasmic reticulum and are subject to activity-dependent proteasomal degradation. Most GABAARs are first delivered to extrasynaptic locations, they then either diffuse to and become trapped at postsynaptic sites or undergo endocytosis. NL2 and synaptic GABAAR stabilize each other, either through intracellular reciprocal interactions aided by scaffolding proteins such as gephyrin or through extracellular cis interaction. In addition, GABA activation of GABAARs might further stabilize synaptic GABAARs through structural changes or signaling mechanisms. Such activity- and GABA-mediated stabilization of GABAAR might further increase the levels of NL2 at cell-cell contacts and, in turn, stabilize presynaptic terminals through trans-synaptic interactions with neurexins.

In an alternative model, it is not the GABAAR levels but rather the expression and/or localization of NL2 itself that might be regulated by GABA signaling, either through regulating NL2 protein levels or NL2-interacting proteins involved in its synaptic localization. It is also possible that GABA binding to GABAARs might modulate their coupling to NL2, thereby increasing the potency and affinity of NL2 towards neurexin in the presynaptic terminals.

Conclusion

Evidence from several lines of experiments are converging to suggest that synaptic signaling by GABA-GABAA receptors may engage the neuroligin-neurexin cell adhesion system to regulate activity-dependent development of inhibitory synapses and innervation patterns. Clearly, the models described above are highly speculative at this time but they provide exciting opportunities to test the coupling of GABAARs to synaptic adhesion complexes. A key question in understanding this molecular coupling will be to determine how GABAARs interact with NLs. A conventional model would include the interaction of NL2 with GABAergic scaffolding molecules which in turn couple to the receptors. However, other models such as direct interactions between NL2 and GABAAR subunits can not be ruled out. Another major question is whether and how GABA binding to GABAARs regulates their conformation and/or biochemical interactions with other synaptic proteins, thereby influencing receptor stability, trafficking, and turnover. The significant advances made during the past years have implicated the neuroligin-neurexin system as well as GABAARs as key players in inhibitory synapse development. Exploring whether these two systems organize GABAergic synapses through direct interactions or parallel pathways will be a challenge for the future.

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 citable 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.

References

  • 1.Markram H, Toledo-Rodriguez M, Wang Y, Gupta A, Silberberg G, Wu C. Interneurons of the neocortical inhibitory system. Nat Rev Neurosci. 2004;5:793–807. doi: 10.1038/nrn1519. [DOI] [PubMed] [Google Scholar]
  • 2.Huang ZJ, Di Cristo G, Ango F. Development of GABA innervation in the cerebral and cerebellar cortices. Nat Rev Neurosci. 2007;8:673–686. doi: 10.1038/nrn2188. [DOI] [PubMed] [Google Scholar]
  • 3.Chattopadhyaya B, Di Cristo G, Higashiyama H, Knott GW, Kuhlman SJ, Welker E, Huang ZJ. Experience and activity-dependent maturation of perisomatic GABAergic innervation in primary visual cortex during a postnatal critical period. J Neurosci. 2004;24:9598–9611. doi: 10.1523/JNEUROSCI.1851-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Jiao Y, Zhang C, Yanagawa Y, Sun QQ. Major effects of sensory experiences on the neocortical inhibitory circuits. J Neurosci. 2006;26:8691–8701. doi: 10.1523/JNEUROSCI.2478-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Miller MW. Maturation of rat visual cortex. III. Postnatal morphogenesis and synaptogenesis of local circuit neurons. Brain Res. 1986;390:271–285. doi: 10.1016/s0006-8993(86)80236-0. [DOI] [PubMed] [Google Scholar]
  • 6.Morales B, Choi SY, Kirkwood A. Dark rearing alters the development of GABAergic transmission in visual cortex. J Neurosci. 2002;22:8084–8090. doi: 10.1523/JNEUROSCI.22-18-08084.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Hua JY, Smith SJ. Neural activity and the dynamics of central nervous system development. Nat Neurosci. 2004;7:327–332. doi: 10.1038/nn1218. [DOI] [PubMed] [Google Scholar]
  • 8.Zhang LI, Poo MM. Electrical activity and development of neural circuits. Nat Neurosci. 2001;4:1207–1214. doi: 10.1038/nn753. [DOI] [PubMed] [Google Scholar]
  • 9.Varoqueaux F, Sigler A, Rhee JS, Brose N, Enk C, Reim K, Rosenmund C. Total arrest of spontaneous and evoked synaptic transmission but normal synaptogenesis in the absence of Munc13-mediated vesicle priming. Proc Natl Acad Sci U S A. 2002;99:9037–9042. doi: 10.1073/pnas.122623799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Verhage M, Maia AS, Plomp JJ, Brussaard AB, Heeroma JH, Vermeer H, Toonen RF, Hammer RE, van den Berg TK, Missler M, et al. Synaptic assembly of the brain in the absence of neurotransmitter secretion. Science. 2000;287:864–869. doi: 10.1126/science.287.5454.864. [DOI] [PubMed] [Google Scholar]
  • 11.Tashiro A, Dunaevsky A, Blazeski R, Mason CA, Yuste R. Bidirectional regulation of hippocampal mossy fiber filopodial motility by kainate receptors: a two-step model of synaptogenesis. Neuron. 2003;38:773–784. doi: 10.1016/s0896-6273(03)00299-x. [DOI] [PubMed] [Google Scholar]
  • 12.Wong WT, Wong RO. Changing specificity of neurotransmitter regulation of rapid dendritic remodeling during synaptogenesis. Nat Neurosci. 2001;4:351–352. doi: 10.1038/85987. [DOI] [PubMed] [Google Scholar]
  • 13.Zheng JQ, Felder M, Connor JA, Poo MM. Turning of nerve growth cones induced by neurotransmitters. Nature. 1994;368:140–144. doi: 10.1038/368140a0. [DOI] [PubMed] [Google Scholar]
  • 14.Carroll RC, Lissin DV, von Zastrow M, Nicoll RA, Malenka RC. Rapid redistribution of glutamate receptors contributes to long-term depression in hippocampal cultures. Nat Neurosci. 1999;2:454–460. doi: 10.1038/8123. [DOI] [PubMed] [Google Scholar]
  • 15.Shi SH, Hayashi Y, Petralia RS, Zaman SH, Wenthold RJ, Svoboda K, Malinow R. Rapid spine delivery and redistribution of AMPA receptors after synaptic NMDA receptor activation. Science. 1999;284:1811–1816. doi: 10.1126/science.284.5421.1811. [DOI] [PubMed] [Google Scholar]
  • 16.Malinow R, Malenka RC. AMPA receptor trafficking and synaptic plasticity. Annu Rev Neurosci. 2002;25:103–126. doi: 10.1146/annurev.neuro.25.112701.142758. [DOI] [PubMed] [Google Scholar]
  • 17.Bonhoeffer T, Yuste R. Spine motility. Phenomenology, mechanisms, and function. Neuron. 2002;35:1019–1027. doi: 10.1016/s0896-6273(02)00906-6. [DOI] [PubMed] [Google Scholar]
  • 18.Hua JY, Smear MC, Baier H, Smith SJ. Regulation of axon growth in vivo by activity-based competition. Nature. 2005;434:1022–1026. doi: 10.1038/nature03409. [DOI] [PubMed] [Google Scholar]
  • 19.Ruthazer ES, Akerman CJ, Cline HT. Control of axon branch dynamics by correlated activity in vivo. Science. 2003;301:66–70. doi: 10.1126/science.1082545. [DOI] [PubMed] [Google Scholar]
  • 20.Brandon EP, Lin W, D’Amour KA, Pizzo DP, Dominguez B, Sugiura Y, Thode S, Ko CP, Thal LJ, Gage FH, et al. Aberrant patterning of neuromuscular synapses in choline acetyltransferase-deficient mice. J Neurosci. 2003;23:539–549. doi: 10.1523/JNEUROSCI.23-02-00539.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Misgeld T, Burgess RW, Lewis RM, Cunningham JM, Lichtman JW, Sanes JR, Greenberg ME. Roles of neurotransmitter in synapse formation: development of neuromuscular junctions lacking choline acetyltransferase. Neuron. 2002;36:635–648. doi: 10.1016/s0896-6273(02)01020-6. [DOI] [PubMed] [Google Scholar]
  • 22.Owens DF, Kriegstein AR. Is there more to GABA than synaptic inhibition? Nat Rev Neurosci. 2002;3:715–727. doi: 10.1038/nrn919. [DOI] [PubMed] [Google Scholar]
  • 23.Ben-Ari Y, Cherubini E, Corradetti R, Gaiarsa JL. Giant synaptic potentials in immature rat CA3 hippocampal neurones. J Physiol. 1989;416:303–325. doi: 10.1113/jphysiol.1989.sp017762. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Leinekugel X, Tseeb V, Ben-Ari Y, Bregestovski P. Synaptic GABAA activation induces Ca2+ rise in pyramidal cells and interneurons from rat neonatal hippocampal slices. J Physiol. 1995;487 ( Pt 2):319–329. doi: 10.1113/jphysiol.1995.sp020882. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Represa A, Ben-Ari Y. Trophic actions of GABA on neuronal development. Trends Neurosci. 2005;28:278–283. doi: 10.1016/j.tins.2005.03.010. [DOI] [PubMed] [Google Scholar]
  • 26.Hennou S, Khalilov I, Diabira D, Ben-Ari Y, Gozlan H. Early sequential formation of functional GABA(A) and glutamatergic synapses on CA1 interneurons of the rat foetal hippocampus. Eur J Neurosci. 2002;16:197–208. doi: 10.1046/j.1460-9568.2002.02073.x. [DOI] [PubMed] [Google Scholar]
  • 27.Tyzio R, Represa A, Jorquera I, Ben-Ari Y, Gozlan H, Aniksztejn L. The establishment of GABAergic and glutamatergic synapses on CA1 pyramidal neurons is sequential and correlates with the development of the apical dendrite. J Neurosci. 1999;19:10372–10382. doi: 10.1523/JNEUROSCI.19-23-10372.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Ben-Ari Y, Khalilov I, Represa A, Gozlan H. Interneurons set the tune of developing networks. Trends Neurosci. 2004;27:422–427. doi: 10.1016/j.tins.2004.05.002. [DOI] [PubMed] [Google Scholar]
  • 29.Rivera C, Voipio J, Payne JA, Ruusuvuori E, Lahtinen H, Lamsa K, Pirvola U, Saarma M, Kaila K. The K+/Cl− co-transporter KCC2 renders GABA hyperpolarizing during neuronal maturation. Nature. 1999;397:251–255. doi: 10.1038/16697. [DOI] [PubMed] [Google Scholar]
  • 30.Ben-Ari Y, Gaiarsa JL, Tyzio R, Khazipov R. GABA: A Pioneer Transmitter That Excites Immature Neurons and Generates Primitive Oscillations. Physiol Rev. 2007;87:1215–1284. doi: 10.1152/physrev.00017.2006. [DOI] [PubMed] [Google Scholar]
  • 31.Di Cristo G, Chattopadhyaya B, Kuhjman SJ, Fu Y, Belanger M-D, Wu C, Rutishauser U, Maffei L, Huang ZJ. Activity-dependent PSA expression regulates inhibitory maturation and onset of critical period plasticity. Nature Neuroscience. 2007;10(12):1569–77. doi: 10.1038/nn2008. [DOI] [PubMed] [Google Scholar]
  • 32.Di Cristo G, Wu C, Chattopadhyaya B, Ango F, Knott G, Welker E, Svoboda K, Huang ZJ. Subcellular domain-restricted GABAergic innervation in primary visual cortex in the absence of sensory and thalamic inputs. Nat Neurosci. 2004;7:1184–1186. doi: 10.1038/nn1334. [DOI] [PubMed] [Google Scholar]
  • 33.Klostermann O, Wahle P. Patterns of spontaneous activity and morphology of interneuron types in organotypic cortex and thalamus-cortex cultures. Neuroscience. 1999;92:1243–1259. doi: 10.1016/s0306-4522(99)00009-3. [DOI] [PubMed] [Google Scholar]
  • **34.Chattopadhyaya B, Di Cristo G, Wu CZ, Knott G, Kuhlman S, Fu Y, Palmiter RD, Huang ZJ. GAD67-mediated GABA synthesis and signaling regulate inhibitory synaptic innervation in the visual cortex. Neuron. 2007;54:889–903. doi: 10.1016/j.neuron.2007.05.015. This paper provides the first evidence that GABA and GAD67 act beyond inhibitory transmission and regulate the development of inhibitory synapses and innervation pattern in the adolescent visual cortex. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Soghomonian JJ, Martin DL. Two isoforms of glutamate decarboxylase: why? Trends Pharmacol Sci. 1998;19:500–505. doi: 10.1016/s0165-6147(98)01270-x. [DOI] [PubMed] [Google Scholar]
  • 36.Asada H, Kawamura Y, Maruyama K, Kume H, Ding RG, Kanbara N, Kuzume H, Sanbo M, Yagi T, Obata K. Cleft palate and decreased brain gamma-aminobutyric acid in mice lacking the 67-kDa isoform of glutamic acid decarboxylase. Proc Natl Acad Sci U S A. 1997;94:6496–6499. doi: 10.1073/pnas.94.12.6496. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Ji F, Kanbara N, Obata K. GABA and histogenesis in fetal and neonatal mouse brain lacking both the isoforms of glutamic acid decarboxylase. Neurosci Res. 1999;33:187–194. doi: 10.1016/s0168-0102(99)00011-5. [DOI] [PubMed] [Google Scholar]
  • 38.Michels G, Moss SJ. GABAA receptors: properties and trafficking. Crit Rev Biochem Mol Biol. 2007;42:3–14. doi: 10.1080/10409230601146219. [DOI] [PubMed] [Google Scholar]
  • 39.Fritschy JM, Panzanelli P. Molecular and synaptic organization of GABAA receptors in the cerebellum: Effects of targeted subunit gene deletions. Cerebellum. 2006;5:275–285. doi: 10.1080/14734220600962805. [DOI] [PubMed] [Google Scholar]
  • **40.Fritschy JM, Panzanelli P, Kralic JE, Vogt KE, Sassoe-Pognetto M. Differential dependence of axo-dendritic and axo-somatic GABAergic synapses on GABAA receptors containing the alpha1 subunit in Purkinje cells. J Neurosci. 2006;26:3245–3255. doi: 10.1523/JNEUROSCI.5118-05.2006. The authors examine synapse formation in the cerebellum in GABAR alpha1 −/− mice. This is the first study to demonstrate the formation of mismatched GABAergic terminals on glutamatergic spines in absence of functional postsynaptic GABA receptors. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Essrich C, Lorez M, Benson JA, Fritschy JM, Luscher B. Postsynaptic clustering of major GABAA receptor subtypes requires the gamma 2 subunit and gephyrin. Nat Neurosci. 1998;1:563–571. doi: 10.1038/2798. [DOI] [PubMed] [Google Scholar]
  • 42.Schweizer C, Balsiger S, Bluethmann H, Mansuy IM, Fritschy JM, Mohler H, Luscher B. The gamma 2 subunit of GABA(A) receptors is required for maintenance of receptors at mature synapses. Mol Cell Neurosci. 2003;24:442–450. doi: 10.1016/s1044-7431(03)00202-1. [DOI] [PubMed] [Google Scholar]
  • *43.Fang C, Deng L, Keller CA, Fukata M, Fukata Y, Chen G, Luscher B. GODZ-mediated palmitoylation of GABA(A) receptors is required for normal assembly and function of GABAergic inhibitory synapses. J Neurosci. 2006;26:12758–12768. doi: 10.1523/JNEUROSCI.4214-06.2006. This paper shows that knockdown of the DHHC-family palmitoyltransferase GODZ perturbs the trafficking of GABAARs to postsynaptic sites and also reduces GABAergic innervation. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Li RW, Yu W, Christie S, Miralles CP, Bai J, Loturco JJ, De Blas AL. Disruption of postsynaptic GABA receptor clusters leads to decreased GABAergic innervation of pyramidal neurons. J Neurochem. 2005;95:756–770. doi: 10.1111/j.1471-4159.2005.03426.x. [DOI] [PubMed] [Google Scholar]
  • 45.Brose N. Synaptic cell adhesion proteins and synaptogenesis in the mammalian central nervous system. Naturwissenschaften. 1999;86:516–524. doi: 10.1007/s001140050666. [DOI] [PubMed] [Google Scholar]
  • 46.Craig AM, Kang Y. Neurexin-neuroligin signaling in synapse development. Curr Opin Neurobiol. 2007;17:43–52. doi: 10.1016/j.conb.2007.01.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Levinson JN, El-Husseini A. Building excitatory and inhibitory synapses: balancing neuroligin partnerships. Neuron. 2005;48:171–174. doi: 10.1016/j.neuron.2005.09.017. [DOI] [PubMed] [Google Scholar]
  • 48.Missler M, Zhang W, Rohlmann A, Kattenstroth G, Hammer RE, Gottmann K, Sudhof TC. Alpha-neurexins couple Ca2+ channels to synaptic vesicle exocytosis. Nature. 2003;423:939–948. doi: 10.1038/nature01755. [DOI] [PubMed] [Google Scholar]
  • 49.Dudanova I, Tabuchi K, Rohlmann A, Sudhof TC, Missler M. Deletion of alpha-neurexins does not cause a major impairment of axonal pathfinding or synapse formation. J Comp Neurol. 2007;502:261–274. doi: 10.1002/cne.21305. [DOI] [PubMed] [Google Scholar]
  • 50.Varoqueaux F, Aramuni G, Rawson RL, Mohrmann R, Missler M, Gottmann K, Zhang W, Sudhof TC, Brose N. Neuroligins determine synapse maturation and function. Neuron. 2006;51:741–754. doi: 10.1016/j.neuron.2006.09.003. [DOI] [PubMed] [Google Scholar]
  • *51.Boucard AA, Chubykin AA, Comoletti D, Taylor P, Sudhof TC. A splice code for trans-synaptic cell adhesion mediated by binding of neuroligin 1 to alpha- and beta-neurexins. Neuron. 2005;48:229–236. doi: 10.1016/j.neuron.2005.08.026. [DOI] [PubMed] [Google Scholar]
  • *52.Chih B, Gollan L, Scheiffele P. Alternative splicing controls selective trans-synaptic interactions of the neuroligin-neurexin complex. Neuron. 2006;51:171–178. doi: 10.1016/j.neuron.2006.06.005. [DOI] [PubMed] [Google Scholar]
  • *53.Graf ER, Kang Y, Hauner AM, Craig AM. Structure function and splice site analysis of the synaptogenic activity of the neurexin-1 beta LNS domain. J Neurosci. 2006;26:4256–4265. doi: 10.1523/JNEUROSCI.1253-05.2006. These three papers are the first to dissect splice-isoform specific synaptic functions of neuroligins and neurexins. Jointly, these studies suggest that gene and splice isoform choice of neurexins and neuroligins underlies selective trans-synaptic interactions in the central nervous system. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Kang Y, Zhang X, Dobie F, Wu H, Craig AM. Induction of GABAergic Postsynaptic Differentiation by {alpha}-Neurexins. J Biol Chem. 2008;283:2323–2334. doi: 10.1074/jbc.M703957200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • **55.Chubykin AA, Atasoy D, Etherton MR, Brose N, Kavalali ET, Gibson JR, Sudhof TC. Activity-dependent validation of excitatory versus inhibitory synapses by neuroligin-1 versus neuroligin-2. Neuron. 2007;54:919–931. doi: 10.1016/j.neuron.2007.05.029. Using tissue from NL1 and NL2 −/− mice, the authors define neuroligin isoform-specific functions. The pharmacological experiments in this study are the first evidence linking neuroligins to activity-dependent synapse formation. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Chih B, Engelman H, Scheiffele P. Control of excitatory and inhibitory synapse formation by neuroligins. Science. 2005;307:1324–1328. doi: 10.1126/science.1107470. [DOI] [PubMed] [Google Scholar]
  • *57.Jacob TC, Moss SJ, Jurd R. GABA(A) receptor trafficking and its role in the dynamic modulation of neuronal inhibition. Nat Rev Neurosci. 2008 doi: 10.1038/nrn2370. An timely and comprehensive review on the cell biology and regulation of GABA(A) receptors. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • *58.Tretter V, Jacob TC, Mukherjee J, Fritschy JM, Pangalos MN, Moss SJ. The clustering of GABA(A) receptor subtypes at inhibitory synapses is facilitated via the direct binding of receptor alpha 2 subunits to gephyrin. J Neurosci. 2008;28:1356–1365. doi: 10.1523/JNEUROSCI.5050-07.2008. This paper describes the first evidence that GABA(A) receptors bind to gephyrin and that disruption of this binding alters the synaptic targeting of receptor subtypes containg α2 subunit. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • *59.Hanus C, Ehrensperger MV, Triller A. Activity-dependent movements of postsynaptic scaffolds at inhibitory synapses. J Neurosci. 2006;26:4586–4595. doi: 10.1523/JNEUROSCI.5123-05.2006. Using live imaging of fluorescence tagged gephyrins, this paper describes the first evidence that GABA(A) receptors bind to gephyrin to reveal its constant synaptic movement, this paper shows that gephyrin is a significant dynamic force at inhibitory synapses. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Dong N, Qi J, Chen G. Molecular reconstitution of functional GABAergic synapses with expression of neuroligin-2 and GABAA receptors. Mol Cell Neurosci. 2007;35:14–23. doi: 10.1016/j.mcn.2007.01.013. [DOI] [PubMed] [Google Scholar]
  • 61.Saliba RS, Michels G, Jacob TC, Pangalos MN, Moss SJ. Activity-dependent ubiquitination of GABA(A) receptors regulates their accumulation at synaptic sites. J Neurosci. 2007;27:13341–13351. doi: 10.1523/JNEUROSCI.3277-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • *62.Ehlers MD, Heine M, Groc L, Lee MC, Choquet D. Diffusional trapping of GluR1 AMPA receptors by input-specific synaptic activity. Neuron. 2007;54:447–460. doi: 10.1016/j.neuron.2007.04.010. This paper shows that local spontaneous activity and glutamate release reduce diffusion exchange of GluR1 between synaptic and extrasynaptic domains, resulting in postsynaptic accumulation of GluR1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Ehrlich I, Klein M, Rumpel S, Malinow R. PSD-95 is required for activity-driven synapse stabilization. Proc Natl Acad Sci U S A. 2007;104:4176–4181. doi: 10.1073/pnas.0609307104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Futai K, Kim MJ, Hashikawa T, Scheiffele P, Sheng M, Hayashi Y. Retrograde modulation of presynaptic release probability through signaling mediated by PSD-95-neuroligin. Nat Neurosci. 2007;10:186–195. doi: 10.1038/nn1837. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES