Sie sind auf Seite 1von 20

Progress in Neurobiology 95 (2011) 68–87

Contents lists available at ScienceDirect

Progress in Neurobiology
journal homepage: www.elsevier.com/locate/pneurobio

Emerging themes in GABAergic synapse development


Marissa S. Kuzirian, Suzanne Paradis *
Brandeis Univeristy, Department of Biology, National Center for Behavioral Genomics, Volen Center for Complex Systems, Waltham, MA 02453, United States

A R T I C L E I N F O A B S T R A C T

Article history: Glutamatergic synapse development has been rigorously investigated for the past two decades at both
Received 1 March 2011 the molecular and cell biological level yet a comparable intensity of investigation into the cellular and
Received in revised form 30 June 2011 molecular mechanisms of GABAergic synapse development has been lacking until relatively recently.
Accepted 3 July 2011
This review will provide a detailed overview of the current understanding of GABAergic synapse
Available online 20 July 2011
development with a particular emphasis on assembly of synaptic components, molecular mechanisms of
synaptic development, and a subset of human disorders which manifest when GABAergic synapse
Keywords:
development is disrupted. An unexpected and emerging theme from these studies is that glutamatergic
Synapse development
and GABAergic synapse development share a number of overlapping molecular and cell biological
GABAergic
Glutamatergic mechanisms that will be emphasized in this review.
ß 2011 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
1.1. The mature GABAergic synapse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
1.1.1. Presynaptic terminal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
1.1.2. Postsynaptic terminal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
2. Assembly of GABAergic synapses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
2.1. First contact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
2.2. Arrival of synaptic components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
2.3. Localization of synaptic contacts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
2.4. Relative order of glutamatergic and GABAergic synapse development in hippocampus. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
2.5. Fidelity of synapse assembly . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
3. Molecular mechanisms of GABAergic synapse development . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
3.1. Gephyrin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
3.2. Collybistin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
3.3. Neurexin–neuroligin signaling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
3.4. Dystrophin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
3.5. Neuregulin 1/ErbB4 signaling. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
3.6. FGF7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
3.7. Npas4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
3.8. Semaphorins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
3.9. Brain-derived neurotrophic factor (BDNF) . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.10. Astrocyte-derived factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

Abbreviations: GABA-g, aminobutyric acid; VAMP2, vesicle-associated membrane protein 2; GFP, reen fluorescent protein; GABAR, ABA receptor; AMPAR, a-amino-3-
hydroxy-5-methyl-4-isoxazoleproprionic acid receptor; NMDARs, N-methyl D-aspartate receptor; PSD, postsynaptic density; GlyR, glycine receptor; GISP, G protein-coupled
receptor interacting scaffold protein; Mupp1, multi-PDZ domain protein 1; GAD65/67, glutamic acid decarboxylase 65 kDa/67 kDa; CA1/CA3, cornu ammonis 1/3; GKAP,
guanylate kinase-associated protein; mIPSC, miniature inhibitory postsynaptic current; HAP1, huntingtin-associated protein; P#, postnatal day #; E#, embryonic day #; DIV,
day in vitro; EM, electron microscopy; VIAAT, vescicular inhibitory amino acid transporter; GEF, guanine nucleotide exchange factor; DGC, dystrophin-associated
glycoprotein complex; NMJ, neuromuscular junction; CNS, central nervous system; NRG1, neuregulin 1; EGF, epidermal growth factor; FGF, fibroblast growth factors; Sema,
Semaphorin; Plxn, Plexin; BDNF, brain-derived neurotrophic factor; PNS, peripheral nervous system; CREB, cAMP response element-binding; CaMKII, calcium calmodulin
kinase II; ACMas, trocyte-conditioned media; TSP, thrombospondin; NKCC1, Na+–K+–Cl transporter NKCC1; KCC2, K+–Cl transporter; RNAi, RNA interference; ACSF,
artificial cerebral spinal fluid; GDP, giant depolarizing potential; TTX, tetrodotoxin; ASD, autism spectrum disorder; MECP2, methyl-CpG binding protein-2; SNP, single
nucleotide polymorphisms.
* Corresponding author.
E-mail address: paradis@brandeis.edu (S. Paradis).

0301-0082/$ – see front matter ß 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.pneurobio.2011.07.002
M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87 69

4. Role of GABAergic synaptic transmission in synapse development. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78


4.1. Depolarizing actions of GABA: relevance to glutamatergic synapse development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.2. Giant depolarizing potentials (GDPs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
4.3. A model for the synchronized development of GABAergic and glutamatergic synapses in hippocampus . . . . . . . . . . . . . . . . . . . . . . . 80
4.4. Role of activity in GABAergic synapse development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
5. GABAergic synapse development and neurological disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
5.1. Rett syndrome. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
5.2. Autism spectrum disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
5.3. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

1. Introduction compare and contrast the development of these synapses to those


that release glutamate. Generally, mature GABAergic synapses are
The complex circuitry of the mammalian central nervous inhibitory, as they act to hyperpolarize the postsynaptic neuron,
system (CNS) enables the execution of fundamental cognitive while glutamatergic synapses are excitatory and depolarize the
processes such as learning, speech, and memory. A given neuron, postsynaptic neuron (Ben-Ari et al., 2007). While glutamatergic
through its synaptic connections, either excites or inhibits other neurotransmission is the primary mode of excitatory neurotrans-
neurons in the circuit. If the synaptic connections of this circuit are mission in the hippocampus and is essential for fast neuronal
disrupted, there are devastating consequences for nervous system communication, inhibitory synapses are thought provide a brake
function, as evidenced by the variety of neurological disorders that to neural firing and are important for a number of purposes which
are thought to be the manifestation of inappropriate circuit include, but are not limited to, preventing postsynaptic neurons
formation. In fact, aberrant development of either excitatory or from reaching threshold (Andersen et al., 1963; Buhl et al., 1994;
inhibitory synapses is now widely accepted as a contributing factor Kandel et al., 1961; Miles and Wong, 1987), modulating the pattern
to many neurological impairments such as mental retardation, of action potential firing (Andersen et al., 1964; Cobb et al., 1995;
autism spectrum disorders, and epilepsy (Bear et al., 2004; Miles and Wong, 1984), and modifying synaptic strength (Davies
Fernandez and Garner, 2007; Hong et al., 2005, 2009; Rubenstein et al., 1991).
and Merzenich, 2003; Tabuchi et al., 2007; Zoghbi, 2003) The vast majority of research on synapse development to date
In the central nervous system, the majority of synapses are has focused on glutamatergic synapses. The deficit in knowledge
chemical: they release neurotransmitter into the synaptic cleft, the with respect to GABAergic synapse development is based in large
space between the presynaptic neuron and the postsynaptic part on the difficulties inherent in a biochemical purification of
neuron. Neurotransmitters then bind to receptors in the postsyn- material from sparse GABAergic synapses, and only relatively
aptic membrane, and ion flow through these receptors can recently has there been significant interest in this topic. Therefore,
depolarize or hyperpolarize the postsynaptic neuron. Chemical in this review we will aim to discuss the current understanding of
synapses are defined by the neurotransmitter that they release. how GABAergic synapse development proceeds in the mammalian
Here, we will focus our discussion on synapses in the mammalian hippocampus, and compare and contrast to glutamatergic synapse
hippocampus that release GABA (g-aminobutyric acid), and development when it is relevant or informative.

Fig. 1. (Top A) An electron micrograph (EM) of a symmetric, inhibitory synapse in cortex. Here, both the presynaptic membrane and the postsynaptic membrane have
approximately equal electron densities. (Top B) An EM image of an asymmetric, excitatory synapse in cortex. Here, you can see the presence of a postsynaptic density (PSD).
(Top A and B) Synaptic vesicles are visible in the presynaptic terminal of both synapses. (Bottom A and B) Schematic diagram of the EM images above, depicting the major
structural entities observed.
EM images reproduced from Colonnier (1968).
70 M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87

1.1. The mature GABAergic synapse active zone in glutamatergic synapses (Richter et al., 1999;
Sanmarti-Vila et al., 2000). However, the precise role of Bassoon
In this section we will discuss the properties of a mature at GABAergic synapses remains to be determined. Presynaptic
GABAergic synapse, to help guide later discussion of GABAergic adhesion molecules, such as neurexins, play a role in the formation
synaptogenesis and maturation. and stabilization of GABAergic synapses and will be discussed in
detail in a later section of this review (Section 3.3).
1.1.1. Presynaptic terminal
Ultrastructural studies have revealed that the presynaptic 1.1.2. Postsynaptic terminal
terminal of GABAergic synapses contains synaptic vesicles poised GABAergic synapses are most frequently found on the dendritic
at the active zone, a presynaptic structure that is the region where shaft or soma of the postsynaptic neuron (Gray, 1959; Peters and
synaptic vesicles cluster, dock, and release neurotransmitter into Palay, 1996). In contrast to a glutamatergic synapse, which has a
the synaptic cleft (Fig. 1). Studies using aldehyde fixation thick, electron-dense postsynaptic density, GABAergic synapses
techniques demonstrate that GABAergic synapses contain synaptic have a much thinner postsynaptic density as visualized by electron
vesicles that are pleiomorphic in shape, whereas glutamatergic microscopy (EM) (Colonnier, 1968; Peters and Palay, 1996), and are
synapses contain vesicles that are larger and more uniformly thus frequently referred to as ‘‘symmetric synapses’’ (Fig. 1). Pre-
spherically structured (Peters and Harriman, 1990; Peters and embedding immunogold labeling of GABAA receptors (GABARs)
Palay, 1996). However, recent structural analysis of glutamatergic using high pressure freezing in the monkey subthalamic nucleus
synapses indicates that classic aldehyde fixation of tissue samples revealed a postsynaptic structure of GABAergic synapses that
leads to structural artifacts that may not accurately depict living contains an electron dense band along the plasma membrane with
tissue (Frotscher et al., 2007; Rostaing et al., 2006; Siksou et al., a less dense filamentous region beneath (Galvan et al., 2004).
2007, 2009). An alternative method to aldehyde fixation, high- While a large number of proteins have been identified at
pressure freezing, has recently been used to study the structure of glutamatergic synapses, relatively few have been identified at
glutamatergic synapses, and shows that the synaptic terminals are GABAergic synapses. This is in part due to the lack of a
larger than previously thought (Rostaing et al., 2006; Siksou et al., postsynaptic density (PSD) in GABAergic synapses, as synaptic
2007). In addition, high-pressure freezing has revealed that proteins localized to glutamatergic synapses were largely
synaptic vesicles are less tightly packed and small filaments identified by isolation of the PSD and subsequent biochemical
linking vesicles into subgroups within the presynaptic terminal are characterization (Kim and Sheng, 2004). This, combined with the
present (Rostaing et al., 2006; Siksou et al., 2009). These results fact that GABAergic synapses are less abundant than glutama-
suggest that there may be structural differences in GABAergic tergic synapses in cortex and hippocampus, has made it
synapses after aldehyde fixation compared to methods that prohibitively difficult to isolate and characterize GABAergic
maintain structural integrity more accurately. Thus far, a synapses using a biochemical approach.
comprehensive analysis of GABAergic synapse ultrastructure using Despite these technical obstacles, a variety of proteins present
high-pressure freezing techniques has not been performed. and functioning at GABAergic synapses continue to be discovered
Glutamatergic synapses contain two different functional pools and described. These include, but are not limited to, scaffolding
of synaptic vesicles, the readily releasable pool and the reserve proteins, cell adhesion molecules, and signal transduction proteins,
pool. Studies using the pH sensitive fluorophore synaptopHluorin all of which will be discussed in detail below (see Section 3). Of
have found that the readily releasable pool is less variable in size at course, neurotransmitter receptors are the hallmark of the
GABAergic synapses compared to glutamatergic synapses postsynaptic terminal in both GABAergic and glutamatergic
(Moulder et al., 2007). This was determined by measuring the synapses. In the hippocampus, ionotropic GABAA receptors
fluorescence intensity of synaptopHluorin (i.e. VAMP2 tagged with (GABAARs) predominate in the GABAergic postsynaptic density
a pH sensitive GFP (Burrone et al., 2006)) puncta after a train of (Fig. 2A). GABAARs are ligand-gated ion channels that are permeable
stimulation was applied to cultured rat hippocampal neurons to chloride and in mature neurons, activation of these receptors
(Moulder et al., 2007). Although this experiment allows a glimpse leads to hyperpolarization of the cell (Ben-Ari et al., 2007).
of the size of vesicle pools at GABAergic synapses, it is not clear if GABAARs are heteromeric pentamers and the molecular
there are multiple pools as in glutamatergic synapses, and if so, makeup of each GABAR determines its functional properties. To
how these pools are utilized in GABAergic synapses. date, 19 GABAAR subunits have been identified in the mammalian
There are a number of different proteins that are present at both CNS. However, each of these 19 subunits can be alternatively
GABAergic and glutamatergic synapses that aid in the organization spliced, thereby further increasing the heterogeneity of GABAARs
and release of synaptic vesicles from the presynaptic terminal. One (Cossart et al., 2006; Dumitriu et al., 2007; Freund and Buzsaki,
such molecule is Synapsin I, which binds presynaptic molecules 1996; Parra et al., 1998). A functional GABAAR contains a, b, and g
such as F-actin (Ceccaldi et al., 1993) and also binds to synaptic subunits (Pritchett et al., 1989). The inter- and intracellular
vesicles (Cheetham et al., 2003). Synapsins are important in distribution of different GABAAR subunits helps to confer different
regulating synaptic vesicle clustering at glutamatergic synapses functions and specificity to a GABAergic synapse. For example, the
(Cesca et al., 2010). The Synapsin I knockout mouse displays a a1 subunit is found in the postsynaptic membrane of almost all
decrease in the amplitude of evoked inhibitory postsynaptic GABAergic synapses throughout the soma and dendrites, while the
currents (IPSCs) in cultured hippocampal neurons (Chiappalone a2 subunit is only located at a subset of synapses, primarily those
et al., 2009). Using multiple probability fluctuation analysis in the axon initial segment (Baumann et al., 2001; Chang et al.,
(MPFA) of evoked IPSCs, the authors conclude that this phenotype 1996; Farrar et al., 1999; Nusser et al., 1996). The g2 subunit of the
is due to a reduction in the size of the readily releasable pool at GABAAR is synaptically localized (Essrich et al., 1998). Functional
GABAergic synapses (Baldelli et al., 2007, 2005; Chiappalone et al., GABAARs form in the absence of the g2 subunit, however these
2009). By contrast, the lack of Synapsin I leads to an increase in the synapses are primarily localized to extrasynaptic sites, indicating
amplitude of evoked excitatory postsynaptic currents (EPSCs) as a that the g2 subunit is required for the proper localization of
result of an increase in the size of the readily releasable pool GABAARs at synapses (Essrich et al., 1998) Mice genetically devoid
(Chiappalone et al., 2009). GABAergic synapses also contain other of the g2 subunit of GABAARs proceed normally through
active zone associated proteins, such as Bassoon, which is embryonic development. However, postnatal development of g2
important in regulating the loading of synaptic vesicles to the / mice is marked by sensorimotor dysfunction and retarded
M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87 71

Fig. 2. (A) Schematic of molecules that have been implicated in GABAergic synapse development. See Section 3 for details. (B) Schematic of cell adhesion molecules that have
been implicated in glutamatergic synapse development. See Section 3 and Dalva et al. (2007) for details. Note that a subset of molecules (e.g. Semaphorins, Neuroligins) are
implicated in both glutamatergic and GABAergic synapse development.

growth, and most only live for a few hours after birth (Essrich et al., hippocampal pyramidal neurons, the activation of GABABRs can
1998; Gunther et al., 1995). lead to hyperpolarization of the postsynaptic cell (Dutar and Nicoll,
Another type of GABAR, the metabotropic GABAB receptor 1988; Nicoll et al., 1990; Scanziani, 2000). This hyperpolarization is
(GABABR), plays an important role in the postsynaptic terminal of distinct from the hyperpolarization elicited by GABAARs, as it is
GABAergic synapses. Metabotropic GABABRs function through Go slower and prolonged (Dutar and Nicoll, 1988), and is necessary for
and Gi proteins that can lead to a number of different downstream rhythmic hippocampal activity (Scanziani, 2000). GABABRs are also
consequences, such as activation of potassium-permeable ion found in the presynaptic terminal of glutamatergic synapses and in
channels (Gahwiler and Brown, 1985) or inhibition of adenylyl GABAergic axons (Chalifoux and Carter, 2011; Kulik et al., 2003;
cyclase, leading to a decrease in protein kinase A (PKA) activity Vigot et al., 2006). There are a number of different roles of
(Chalifoux and Carter, 2011; Xu and Wojcik, 1986). Unlike presynaptic GABABRs; they can activate K+ channels and indirectly
GABAARs, where 19 different subunits have been identified, only limit calcium release (Thompson and Gahwiler, 1992), as well as
two subunits have been discovered for the GABABRs, GABAB1/B2 suppress multivesicular release (Chalifoux and Carter, 2010).
(Bowery and Brown, 1997; Isomoto et al., 1998; Kaupmann et al., Despite our limited knowledge of the ultrastructural and
1997; Pfaff et al., 1999; Schwarz et al., 2000). GABABRs elicit a molecular composition of GABAergic synapses at the present time,
number of different downstream signals, presumably in part by thus far it seems that the overall organization of GABAergic
signaling through auxiliary binding proteins such as GISP and postsynaptic terminals is similar to glutamatergic postsynaptic
Mupp1 (Balasubramanian et al., 2007; Chalifoux and Carter, 2011; sites, but with different molecular players. Currently, there is
Kantamneni et al., 2007). renewed interest in GABAergic synapse composition and develop-
Postsynaptic GABABRs are located extrasynaptically and are ment and it will be interesting to see how this vision changes as
activated by GABA spillover (Kulik et al., 2003; Scanziani, 2000). In more focus is directed towards GABAergic synapses.
72 M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87

2. Assembly of GABAergic synapses presynaptic molecules. These transport vesicles can accumulate at
a specific site to begin formation of the presynaptic active zone
Although identifying the major events in glutamatergic synapse (Ahmari et al., 2000; Shapira et al., 2003; Zhai et al., 2001). This pre-
assembly has been a focus in the field of synapse development for fabrication and organization of synaptic components into modular
many years (Ahmari et al., 2000; Dai and Peng, 1996; Gerrow et al., packets could facilitate the rapid rate of synaptogenesis, which can
2006; Sabo et al., 2006; Shapira et al., 2003; Zakharenko et al., occur in 30–60 min (Friedman et al., 2000).
1999; Zhai et al., 2001), assembly of GABAergic synapses is To date, it is unclear if GABAergic synapses use similar transport
considerably less well-understood. Here, we will discuss what is vesicles to deliver components to the synapse. However, on the
known about the order of events that comprise GABAergic synapse postsynaptic side, time-lapse imaging has been used to investigate
assembly. the trafficking of GABAARs in and out of mature synapses (for
comprehensive reviews see (Jacob et al., 2008; Luscher et al., 2011).
2.1. First contact To summarize a large and ongoing body of work, GABAARs
assemble in the endoplasmic reticulum, are transported through
An initial step in synapse development is contact between the the secretory pathway, and are then inserted into the plasma
presynaptic axon and the postsynaptic dendrite and subsequent membrane. Some mechanisms by which GABAARs are trafficked to
stabilization of this contact. The Bonhoeffer group has provided an and retained at synaptic sites will be discussed below (Section 3).
intriguing look at this issue using two-photon laser scanning For example, binding of GABAARs to components of the GABAergic
microscopy in hippocampal slice cultures derived from GAD65- postsynaptic scaffold such as gephyrin and collybistin could serve
GFP expressing mice, which express GFP in a subset of to recruit or retain GABAARs at synaptic sites (Bannai et al., 2009;
interneurons in the hippocampus (Wierenga et al., 2008). By Bogdanov et al., 2006).
visualizing GABAergic presynaptic boutons in this manner, and Kinesin family motor proteins have been implicated in the
postsynaptic CA1 pyramidal cell dendrites by Alexa Fluor 594 transport of GABAARs from internal compartments to extrasynap-
injection, the authors observed dendritic and axonal protrusions tic and synaptic sites (Twelvetrees et al., 2010), as demonstrated by
that came in contact with each other; however, these contacts immunostaining, live imaging, and electrophysiological recordings
were always transient (Wierenga et al., 2008). Instead, mature of synaptic currents from cultured neurons. For example, when the
GABAergic synapses form at pre-existing axo-dendritic crossings function of the kinesin KIF5 is blocked with a kinesin-specific
and are marked by the appearance of new, stable GABAergic antibody, the amplitude of miniature inhibitory postsynaptic
boutons (Wierenga et al., 2008). In addition, retrospective currents (mIPSCs) decreases, while the frequency of mIPSC events
immunostaining after time-lapse imaging indicates that new remains the same as control conditions (Twelvetrees et al., 2010).
synapses form on the order of a few hours (Wierenga et al., 2008), In addition, coimmunoprecipitation using GABAAR b subunits
similar to the time reported for glutamatergic synapse formation pulls down KIF5 as well as the adaptor protein huntingtin-
(Friedman et al., 2000). This study represents the first report that associated protein HAP1 (Twelvetrees et al., 2010). These and other
visualizes the initial contact between a GABAergic axon and its experiments from this study suggest that disruption of GABAAR
postsynaptic partner, and provides exciting evidence that initia- trafficking may underlie some of the etiology of Huntington’s
tion of contact in GABAergic synapse development has both Disease.
marked similarities and differences when compared to glutama-
tergic synaptogenesis. Future studies likely will aim to elucidate 2.3. Localization of synaptic contacts
which proteins are recruited to young GABAergic synapses, and
how these proteins arrive at the synapse. During development, synapses form within specific regions on
the postsynaptic neuron. For example, the excitatory synapses on
2.2. Arrival of synaptic components pyramidal neurons and cerebellar Purkinje neurons form primarily
onto dendritic spines, while GABAergic synapses are localized to
Unlike glutamatergic synapses which have been extensively the dendritic shaft or cell soma (Lardi-Studler and Fritschy, 2007;
studied by time-lapse imaging of fluorescently tagged proteins Tretter and Moss, 2008; Wierenga et al., 2008). Recent studies have
such as the AMPA receptor (ionotrophic glutamate receptor a- suggested that a ‘‘molecular code’’ regulates the subcellular
amino-3-hydroxy-5-methyl-4-isoxazoleproprionic acid receptor; organization of GABAergic synapses. For example, in the cerebellar
AMPAR) subunits and PSD-95 (Barrow et al., 2009; Bresler et al., cortex, synapses from inhibitory stellate cells form onto the
2001; Correia et al., 2008; Friedman et al., 2000; Gerrow et al., dendrites of Purkinje cells, while inhibitory basket cells localize
2006; Washbourne et al., 2002, 2004), the order of assembly of the their synapses onto the Purkinje cell soma and axon initial segment
components of the pre- and postsynaptic specializations of (Ango et al., 2004; Lardi-Studler and Fritschy, 2007). Neurofas-
GABAergic synapses has not been elucidated. However, it is cin186, a cell adhesion molecule, is expressed as a gradient along
instructive to summarize what has been learned from such studies the axon initial segment of Purkinje cells (Ango et al., 2004). This
of glutamatergic synapse development to provide a framework gradient is necessary for proper localization of basket cell
with which to think about GABAergic synapse development. innervation and is demarcated by the scaffolding protein
Significant evidence supports vesicular transport of both pre- ankyrin-G (Ango et al., 2004).
and postsynaptic proteins to a nascent synapse (McAllister, 2007). In addition to cell adhesion molecules, proper synapse
For example, an emerging view of postsynaptic assembly reveals localization can be dictated through postsynaptic receptors, such
that mobile transport packets deliver preformed scaffold com- as the GABAAR a1 subunit. Mice lacking the GABAAR subunit a1 are
plexes containing PSD-95, GKAP, and Shank to nascent synapses not able to maintain GABAergic synapses from stellate cells onto
(Gerrow et al., 2006; Prange and Murphy, 2001). Scaffolding Purkinje cells in the cerebellum. However, basket cells are
proteins accumulate in the postsynaptic terminal concurrently unaffected by the absence of this postsynaptic receptor in the
with ionotrophic glutamate receptors, AMPARs and NMDARs (N- innervated Purkinje cell soma (Fritschy et al., 2006). It is likely that
methyl D-aspartate receptor) (Barrow et al., 2009; Bresler et al., the differential expression of receptor subunits throughout the
2001; Friedman et al., 2000; Gerrow et al., 2006; Washbourne et al., neuron helps to orchestrate the complex pattern of connections
2002). Presynaptic assembly is also characterized by the accumu- between different types of neurons. This may be accomplished by
lation of discrete transport vesicles that carry a characteristic set of receptor subunits selectively binding to specific cell adhesion
M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87 73

molecules within the postsynaptic site, which are in turn studies do not recapitulate the studies of acute neonatal
recognized by their binding partners expressed at presynaptic hippocampal slice discussed above. However, dissociation of
terminals in axons of specific subtypes of neurons. Examples of neurons before plating involves the shearing of the neuronal
such molecules that are known to mediate GABAergic synapse processes, which must regrow in culture. Therefore, any synaptic
formation are discussed below (Section 3). contacts that had already been established in vivo must be re-
established in vitro, and this appears to take a few days. Thus, it
2.4. Relative order of glutamatergic and GABAergic synapse could be that this ‘‘re-setting’’ of synapse development overlaps
development in hippocampus with the period of time when robust glutamatergic and GABAergic
synapse development is occurring simultaneously, and thus the
The weight of the evidence suggests that GABAergic synapse relative order is lost. Alternatively, it could be that the molecular
development is initiated prior to glutamatergic synapse develop- mechanisms that govern glutamatergic synapse development are
ment in the hippocampus (Hennou et al., 2002; Tyzio et al., 1999). more promiscuous or inappropriately activated in the context of
Although nowhere near as well-studied as in the hippocampus, dissociated cultures of neurons.
this order seems to be recapitulated during development of other
areas of the CNS such as the neocortex and spinal cord (for a 2.5. Fidelity of synapse assembly
comprehensive discussion of this issue see Ben-Ari et al., 2007).
Electrophysiological recordings from CA1 neurons in acute Using time-lapse confocal imaging of glutamatergic synapses in
hippocampal slices taken from postnatal day 0 (P0) rats demon- the CA3 region of hippocampus in organotypic slice culture, the
strate that the vast majority of pyramidal cells are synaptically Bonhoeffer group demonstrated that dendritic filopodia differen-
silent (Tyzio et al., 1999). Of the non-silent neurons, approximately tiate between a glutamatergic axon and GABAergic axon (Lohmann
equal numbers of neurons contain GABAergic synapses only or and Bonhoeffer, 2008). They show that a local calcium transient
GABAergic and glutamatergic synapses, and synapse number occurs in a dendritic filipodium within seconds after making initial
increases as the neurons mature (Tyzio et al., 1999). In addition, contact with a presynaptic axon, and is therefore unlikely to be
electrophysiological recordings from interneurons in acute hippo- synaptic in origin (Lohmann and Bonhoeffer, 2008). Interestingly,
campal slices taken from rats at P0 reveals that approximately 20% these dendritic filopodia selectively establish contact with
of interneurons have only GABAergic synapses while approxi- glutamatergic axons and not with GABAergic axons, and these
mately 75% of interneurons have both glutamatergic and calcium transients occur at higher frequencies when the filopodia
GABAergic synapses (the remaining 5% are silent) (Hennou contacts glutamatergic axons (Lohmann and Bonhoeffer, 2008).
et al., 2002). Thus, the order of synaptogenesis is similar between This elegant study demonstrates not only that dendritic filopodia
pyramidal and interneurons. Further, on the basis of this study and are capable of selectively choosing their targets, but that they do so
others, the presence of an interneuron circuit has been proposed, via a contact-dependent calcium transient.
where interneurons first innervate other interneurons before they Despite this intriguing study, there is an accumulating body of
innervate pyramidal cells (Ben-Ari et al., 2004; Gozlan and Ben-Ari, evidence that demonstrates ‘‘mismatching’’ of glutamatergic and
2003; Gulyas et al., 1996). This is perhaps not surprising, as GABAergic synaptic elements during the early stages of synapse
interneurons are more mature than pyramidal cells as they migrate development. The group of Deanna Benson performed a compre-
into the hippocampus. hensive study, using immunostaining for various glutamatergic
In contrast, the first glutamatergic synapses are established and GABAergic synaptic proteins, of the protein composition of
around birth to postnatal day two in hippocampal CA1 pyramidal synapses while they were forming in cultured hippocampal
neurons as assayed by whole-cell voltage clamp recordings neurons over more than two weeks in vitro (Anderson et al.,
(Durand et al., 1996; Tyzio et al., 1999). However, most of these 2004). They found that early in the development of the cultures, for
synapses are ‘silent,’ meaning that they contain only NMDARs and example at 7 days in vitro (DIV), a significant percentage of GAD65
not AMPARs, and therefore are not able to participate in rapid positive presynaptic terminals co-stain for both gephyrin and PSD-
synaptic communication (Durand et al., 1996; Wu et al., 1996). The 95, postsynaptic GABAergic and glutamatergic synaptic markers,
percentage of silent synapses steadily decreases around postnatal respectively (Anderson et al., 2004). By 17 DIV however, the
day five until they represent only about 20% of the synapse percentage of mismatching significantly decreases. Additionally,
population after postnatal week 1 (Ben-Ari et al., 1997). the authors made similar observations at glutamatergic presyn-
However, even though GABAergic synapse development begins a aptic terminals: a subset of PSD-95 clusters are observed apposed
few days before glutamatergic synapse development, robust to GABAergic presynaptic markers such as GAD65 in young
synapse development continues throughout the first few postnatal cultures (Anderson et al., 2004). Interestingly, blockade of GABAAR
weeks of life, and of course proceeds throughout the life of the or NMDAR activity from 7 to 17 DIV leads to a decrease in fidelity of
animal. Therefore, much of GABAergic and glutamatergic synapse matching at GABAergic terminals, indicating that activity contrib-
development occurs concurrently (Chen et al., 1995; Hennou et al., utes to matching of proper components at GABAergic synapses
2002; Koller et al., 1990; Tyzio et al., 1999; Walton et al., 1993). This (Anderson et al., 2004). In contrast, NMDAR blockade leads to an
leads to an interesting set of problems for say, a developing CA1 increase in fidelity of matching of glutamatergic synapses
pyramidal neuron onto which glutamatergic and GABAergic (Anderson et al., 2004).
synapses are simultaneously forming. For example, what is the Other groups have made similar observations in traditional
signaling mechanism by which glutamate receptors get trafficked to hippocampal cultures (Brunig et al., 2002), autaptic cultures of
glutamatergic postsynaptic terminals and not GABAergic terminals? hippocampal neurons (Rao et al., 2000), cultured cerebellar granule
Some insight into these issues comes from studies performed cells (Studler et al., 2002) and Purkinje cells (Fritschy et al., 2006).
using retrospective immunostaining of dissociated cultures of Importantly, mismatched synapses have also been observed in vivo
hippocampal neurons isolated from embryonic day (E18) rat pups. in adult rodents, where immuno-EM of mossy fiber to granule cell
These results of these studies indicate that glutamatergic synapse synapses in cerebellum revealed the presence of both GABAAR and
development precedes GABAergic synapse development (Danglot GluA subunit of AMPARs opposed to these glutamatergic terminals
et al., 2006). This finding is also recapitulated by the work of (Nusser et al., 1998). Also, double-innervation of dendritic spines in
Deanna Bensen and colleagues discussed below in Section 2.5 cortex by glutamatergic and GABAergic inputs has been reported
(Anderson et al., 2004). It is not clear why these neuronal culture (Micheva and Beaulieu, 1995; Micheva et al., 2010).
74 M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87

Taken together, these studies suggest that glutamatergic and lead to a hexagonal lattice of gephyrin molecules (Fritschy et al.,
GABAergic synapse development are neither spatially nor tempo- 2008). Surprisingly, cloning and characterization of gephyrin
rally separable processes. It seems reasonable to postulate that as revealed that it is a mediator of molybdenum cofactor (Moco)
transient contacts between axons and dendrites are formed, biosynthesis (Feng et al., 1998), which is required for oxidation of
components of the postsynaptic specializations of both glutama- nitrate to nitrite and sulfate to sulfite (Fritschy et al., 2008). This
tergic and GABAergic synapses, such as scaffolding proteins and family of enzymes is well-conserved from plants to bacteria to
neurotransmitter receptors, are shuttled to these nascent synaptic animals, and perhaps gephyrin is the result of an ancient gene fusion
sites in a stochastic manner. As synapse development proceeds, the event, which resulted in a multimeric protein with a novel function
identity of the synapse as either glutamatergic or GABAergic is (Feng et al., 1998). Interestingly, it remains to be determined if the
established by retention of the appropriate synaptic proteins, two known functions of gephyrin (i.e. clustering of neurotransmitter
perhaps in an activity-dependent manner. It could be that receptors and Moco biosynthesis) are related (Fritschy et al., 2008).
examples of ‘‘mismatched’’ synapses from intact tissue prepara- An analysis of the spinal cord and hypothalamus of mutant mice
tions are rare simply because researchers have not yet examined in which the entire gephyrin gene was constitutively deleted
this question in a systematic manner. Along those lines, recent revealed an absolute requirement for gephyrin in clustering of GlyRs
work has provided evidence for co-release of GABA and glutamate at synapses (Feng et al., 1998). In contrast, a requirement for
at synapses in hippocampus, cortex, and cerebellum (Boulland gephyrin to localize GABAARs to the synapse is less clear. The
et al., 2004; Noh et al., 2010; Somogyi et al., 2004; Zander et al., gephyrin knockout mice exhibit a decrease in the number of
2010). Thus, it could be that ‘‘mismatched’’ synapses are actually synaptically localized GABAAR puncta in retina, spinal cord, and
synapses that are designed to respond to co-release of two hippocampus (Fischer et al., 2000; Kneussel et al., 2001, 1999; Levi
different neurotransmitters from the same presynaptic terminal. et al., 2004). Whole-cell voltage clamp recordings of mIPSCs from
cultured hippocampal neurons obtained from gephyrin knockout
3. Molecular mechanisms of GABAergic synapse development mice reveal a modest decrease in the amplitude of these events and
no change in the frequency (Levi et al., 2004). In addition,
Biochemical and candidate gene approaches by a large number demonstration of a direct interaction between gephyrin and
of labs over the past thirty years have led to the identification of GABAARs has been elusive (Meyer et al., 1995), although recently,
many molecules that function at excitatory, glutamatergic an interaction between the a2 subunit of the GABAAR and gephyrin
synapses in processes such as neurotransmitter release and has been reported (Saiepour et al., 2010; Tretter et al., 2008).
reception. However, despite this knowledge, we have only just Additional work from the laboratory of Stephen Moss has also
begun to define the molecules that mediate the actual assembly of demonstrated that cell-surface GABAARs are more mobile and less
the glutamatergic synapse. And, in contrast to glutamatergic likely to form synaptic clusters in the absence of gephyrin (Jacob
synapse formation, even less is known about inhibitory, GABAergic et al., 2005).
synapse formation in mammals. Notably, the only directed screen The fact that at least a subset of GABAAR subunits are able to
to identify genes that are specifically required for GABAergic localize at synaptic sites and form functional synapses in the
synapse formation and transmission published to date was absence of gephyrin suggests that other, non-gephyrin dependent
performed in C. elegans (Vashlishan et al., 2008). This study mechanisms exist to localize GABAARs to synapses. Interestingly,
implicated a number of molecules in regulating GABAergic post-synaptic clustering of gephyrin is disrupted in the absence of
synaptic transmission, a subset of which may be involved in certain GABAARs subunits (Essrich et al., 1998; Kralic et al., 2006;
assembling the appropriate pre- and postsynaptic specializations. Studer et al., 2006) suggesting a dynamic relationship between
In this section, we will focus our discussion on molecules that gephyrin and GABAARs. Further, an immuno-EM study of gephyrin
have been shown to play a role specifically in mammalian knockout mice examining synapses in the brainstem of P0 mice
inhibitory synapse formation, with a focus on GABAergic synapses demonstrated that GABA and VIAAT-positive synapses are
(Fig. 2A). Interestingly, a number of these molecules/families of morphologically normal in the absence of gephyrin (O’Sullivan
molecules have been shown to play a role in both glutamatergic et al., 2009). Thus, although gephyrin clearly plays an important
and GABAergic synapse formation. We will discuss the role of these role in modulating GABAAR subunit abundance at symmetric
proteins in GABAergic synapse formation specifically. As a number synapses, the precise nature of this modulation and the molecular
of recent reviews have comprehensively detailed the identity and mechanism by which it is achieved are not well understood.
mechanism of action of molecules that are thought to be involved
in mediating glutamatergic synapse development (Dalva et al., 3.2. Collybistin
2007; McAllister, 2007; Scheiffele, 2003), we will not discuss these
molecules further. Collybistin is a brain-specific Rho-family guanine nucleotide
exchange factor (GEF) that was originally identified in a yeast two-
3.1. Gephyrin hybrid screen as protein that interacts with gephyrin (Kins et al.,
2000). Work from a number of labs has implicated collybistin as a
Gephyrin was originally identified as a polypeptide that co- factor required for clustering gephyrin and GABAARs, and mouse
purifies with glycine receptor (GlyR) subunits (Pfeiffer et al., 1982). models in which the collybistin gene has been deleted support this
Gephyrin is expressed throughout the nervous system and in non- hypothesis (Harvey et al., 2004; Kins et al., 2000; Papadopoulos
neuronal tissues (Fritschy et al., 2008), is a prominent component et al., 2008, 2007). In support of the crucial role of collybistin in
of the postsynaptic specialization of glycinergic and GABAergic mediating GABAergic synapse development, a human patient
synapses, and is exclusively localized to these synapses (Sassoe- presenting with a variety of symptoms including mental retarda-
Pognetto et al., 1995). tion, seizures, and increased anxiety was found to have a mutation
For many years, gephyrin has been hypothesized to function in the collybistin gene (Kalscheuer et al., 2009).
analogously to the MAGUK scaffolding proteins such as PSD-95 at Interestingly, the presence of the collybistin N-terminal SH3
excitatory synapses that bind to and localize neurotransmitter domain interferes with the ability of collybistin to form gephyrin
receptors at the postsynaptic density (Kim and Sheng, 2004). aggregates in non-neuronal cells (Harvey et al., 2004; Kins et al.,
Support for this model arises in part from the fact that gephyrin can 2000), presumably through an auto-inhibitory mechanism, and a
form both dimers and trimers and further, these interactions can recent study has uncovered a possible mechanistic explanation for
M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87 75

this observation (Poulopoulos et al., 2009). Binding of the synaptic transmission and development (Varoqueaux et al., 2006).
collybistin SH3 domain to the intracellular domain of the These mice die at birth from respiratory failure, and an analysis of
Neuroligin 2 protein allows a collybistin/gephyrin complex to brainstem circuitry reveals a significant deficit in GABAergic,
nucleate assembly of a gephyrin scaffold specifically at the glycinergic, and glutamatergic synaptic transmission, but only
GABAergic postsynaptic membrane and recruit GABAARs (Poulo- modest changes in synapse number (Varoqueaux et al., 2006, see
poulos et al., 2009) (see further discussion of Neuroligin 2 below). particularly Figs. 4–6). Similarly, a mouse in which all three a-
However, another study demonstrated that overexpression of an neurexin isoforms are deleted displays a similar phenotype and
isoform of collybistin which lacks the SH3 domain in cultured also has a decrease in GABAergic synapse density in brainstem
hippocampal neurons is able to induce synaptic clustering of (Missler et al., 2003). A careful analysis of the phenotype of single
overexpressed gephyrin without a concomitant increase in Nlgn1 / or Nlgn2 / mice revealed that Neuroligin 1 acts
Neuroligin 2 clustering (Chiou et al., 2011). Thus, the precise specifically to promote the functional maturation of glutamatergic
relationship and timing of interactions between Neuroligin 2, synapses while Neuroligin 2 acts specifically to promote the
collybisitin, and gephyrin clustering remains to be determined. functional maturation of GABAergic synapses (Chubykin et al.,
Along those lines, a group investigating the relationship between 2007). Taken together, the in vitro and in vivo data suggest a model
gephyrin, collybistin, and various subunits of the GABAAR where the neuroligin/neurexin adhesion complex acts at multiple
demonstrated that collybistin itself binds directly to the a2 steps in synapse development, perhaps stabilizing nascent
subunit of the GABAAR in yeast and HEK293 cells (Saiepour et al., contacts and recruiting molecules such as neurotransmitter
2010), suggesting a more direct role for collybistin in GABAAR receptors to synapses as they are developing. In addition, the
clustering than was previously assumed. In addition, the authors of rather modest reductions in both synapse number and synaptic
this study provide evidence that GABAARa2, gephyrin, and transmission in the various neurexin and neuroligin knockout
collybistin are able to simultaneously co-associate (Saiepour mouse models suggest that these molecules certainly act
et al., 2010), suggesting a complex interaction between these redundantly with other proteins to promote synapse development.
proteins to regulate GABAAR synaptic clustering. Although the neurexin/neuroligin proteins were initially
studied as molecules which promote excitatory synapse develop-
3.3. Neurexin–neuroligin signaling ment, the specificity of Neuroligin 2 localization to GABAergic
synapses and the prominent deficit in GABAergic synapse
The presynaptic, transmembrane neurexin proteins and post- development in mice in which neurexin/neuroligin signaling is
synaptic, transmembrane neuroligin proteins comprise a large perturbed has led to an increased interest in the role of these
family of trans-synaptic, calcium-dependent cell adhesion mole- proteins in GABAergic synapse development. To this end, two
cules that function to regulate synapse development (Craig and relatively recent studies deserve specific discussion. As mentioned
Kang, 2007; Dalva et al., 2007; Dean and Dresbach, 2006; Huang above, the cytoplasmic domain of Neuroligin 2 binds both gephyrin
and Scheiffele, 2008). In mammals, the neurexins are encoded by and collybistin, resulting in nucleation of a gephyrin/collybistin
three genes, and transcription from two different promoters scaffold at GABAergic synapses (Poulopoulos et al., 2009). Further,
produces an a or b neurexin transcript from each gene (Dalva et al., the neuroligin 2/gephyrin/collybistin complex mediates clustering
2007). Additional alternative splicing generates more than 1000 of GABAARs at synapses (Poulopoulos et al., 2009). Intriguingly,
neurexin isoforms (Dalva et al., 2007). The neuroligins are encoded neurexins directly interact with and impair the function of
by at least 4 genes in rodents (Nlgn1-4); Neuroligin 1 is localized synaptic GABAARs, presumably through a trans-synaptic mecha-
primarily to excitatory synapses (Song et al., 1999) while nism (Zhang et al., 2010a). Thus, it seems as though neuroscientists
Neuroligin 2 is localized predominantly to GABAergic synapses are only beginning to uncover the numerous and varied functions
(Varoqueaux et al., 2004) (Fig. 2). Interestingly, alternative splicing of these cell adhesion complexes in synapse development.
in the extracellular domain of both neurexins and neuroligins
regulates binding selectivity and influences function at the synapse 3.4. Dystrophin
(Boucard et al., 2005; Chih et al., 2006; Graf et al., 2006). Both
neuroligins and neurexins are transmembrane proteins, contain Dystrophin was initially cloned and characterized in 1987;
PDZ domains in their intracellular, C-terminal regions, and have mutations in this gene are the underlying cause of Duchenne
been shown to interact with putative scaffolding proteins such as muscular dystrophy, a disorder that is characterized by muscle
PSD-95 (see references within Craig and Kang, 2007; Dalva et al., wasting and mental retardation (Hoffman et al., 1987). Dystrophin
2007). is expressed in the muscle and brain and is now known to be a
The current interest in these proteins as synaptogenic major component of a molecularly heterogenous complex called
molecules stems from the work of Peter Scheiffele and colleagues the dystrophin-associated glycoprotein complex (DGC) (Pilgram
who discovered that expression of neuroligins in heterologous et al., 2010). Despite decades of intensive investigation, the
cells drives assembly of presynaptic structures in co-cultured function of the DGC is not well understood, although it is largely
neurons (Scheiffele et al., 2000). Using a similar heterologous cell/ believed to form a scaffold which spans the plasma membrane and
neuron co-culture system (Craig et al., 2006), a subsequent study is thought to connect the extracellular matrix to the intracellular
demonstrated that expression of b-neurexins is sufficient to drive actin cytoskeleton (Pilgram et al., 2010). The DGC has been shown
assembly of both glutamatergic and GABAergic postsynaptic to play a role in neurotransmitter receptor clustering and synaptic
specializations in neurons (Graf et al., 2004). Since that time, transmission at both neuromuscular junctions (NMJ) and CNS
results from a large number of in vitro experiments suggested a synapses from worms to mammals (Pilgram et al., 2010).
model where an interaction between a postsynaptic neuroligin and Specifically, immunohistochemistry experiments reveal that
a presynaptic b-neurexin induces the formation of glutamatergic dystrophin clusters co-localize with clusters of both a1 and a2
and GABAergic synapses (Chih et al., 2005; Dean et al., 2003; Graf subunits of GABAAR in hippocampus and cerebellum (Brunig et al.,
et al., 2004; Scheiffele et al., 2000). 2002) (Knuesel et al., 1999). Further, GABAARa1 and a2 clusters
Interestingly, in vivo work has both supported and extended are decreased in the hippocampus and cerebellum of mice lacking
this relatively straightforward model. Analysis of mice in which the the full-length dystrophin isoform (Knuesel et al., 1999). In
Nlgn1,2,3 genes have been constitutively deleted reveals that the addition, these same mice also show a reduction in the number of
predominant function of these proteins is to mediate GABAergic GABAAR a2 clusters and the frequency of norepinephrine-induced
76 M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87

GABAergic IPSCs in amygdala (Sekiguchi et al., 2009). Taken signaling has been implicated includes patterning the early
together, these results, in conjunction with similar analyses of embryo, organogenesis, and cell proliferation and survival (Turner
other components of the DGC (see refs in Pilgram et al., 2010) point and Grose, 2010). FGFs were initially shown to regulate
to an important role for the DGC in clustering GABAARs at synapses. presynaptic differentiation of motoneurons and mossy fibers from
the cerebellum both in vitro and in vivo (Umemori et al., 2004). To
3.5. Neuregulin 1/ErbB4 signaling follow up on this finding, the Umemori group chose to examine
two particular family members, FGF7 and FGF22, in detail in the
Neuregulin 1 (NRG1) and its receptor ErbB4 have been in the hippocampus. This group examined both the intact brain of mice in
spotlight recently due to the identification of both genes as which either FGF7 or FGF22 was deleted, in combination with
schizophrenia susceptibility loci (Mei and Xiong, 2008). The NRG1 studies of neuronal cultures obtained from these animals, to define
locus generates at least 31 different isoforms which are either a role for FGF signaling in the pyramidal neurons of the CA3 region
cleaved or membrane-bound; all isoforms contain an epidermal of the hippocampus. Specifically, it was shown that FGF7 is
growth factor (EGF)-like domain (Mei and Xiong, 2008). The ErbB4 localized to the postsynaptic specialization of GABAergic synapses,
receptor tyrosine kinase is one of four potential NRG1 receptors and is required to promote the clustering of synaptic vesicles at the
(ErbB1-4) and is the only one in which NRG1 binding activates presynaptic termini of GABAergic synapses while having no effect
ErbB tyrosine kinase activity (Mei and Xiong, 2008). NRG1/ErbB4 on glutamatergic synapses (Terauchi et al., 2010). Interestingly,
signaling has been implicated in a variety of developmental FGF22 is localized to the postsynaptic density of glutamatergic
processes in the nervous system including the migration of both synapses, and is required in the same cells as FGF7 to promote
pyramidal and interneurons, axon guidance, axon myelination, and clustering of synaptic vesicles at the presynaptic termini of
synapse formation at the NMJ (Mei and Xiong, 2008). glutamatergic synapses (Terauchi et al., 2010). Thus, FGF7
Previous work from a number of groups suggests a role for represents the only molecule thus far implicated in the presynaptic
NRG1 and ErbB4 in mediating glutamatergic synapse develop- differentiation of GABAergic synapses specifically.
ment. For example, ErbB signaling regulates glutamatergic
synaptic transmission and dendritic spine number and size in 3.7. Npas4
pyramidal neurons in vitro and in vivo (Barros et al., 2009; Li et al.,
2007). However, it is entirely unclear if this is due to a cell Npas4 is a transcription factor whose expression is strongly
automonous effect of ErbB4 function in pyramidal neurons, or is upregulated in neurons by calcium influx in response to neuronal
perhaps due to a local circuit effect, a hypothesis supported by new depolarization (Lin et al., 2008). Knockdown of Npas4 expression
experimental evidence outlined below (Wen et al., 2010). In fact, by RNAi causes a decrease in the density of GABAergic synapses
although the expression pattern of ErbB4 in the rodent brain formed onto pyramidal neurons in the hippocampus while
remains controversial (Garcia et al., 2000; Huang et al., 2000; Woo overexpression of Npas4 increases GABAergic synapse density
et al., 2007), the weight of the evidence now points in favor of (Lin et al., 2008). Among many identified Npas4 targets, it was
ErbB4 expression in a subset of GABAergic interneurons in cortex shown that Npas4 regulates BDNF transcription in response to
and not in glutamatergic, pyramidal cells (Fazzari et al., 2010; neuronal depolarization, and that the effect of Npas4 on GABAergic
Vullhorst et al., 2009). Specifically, both immunostaining and synapse density is partially dependent on BDNF (Lin et al., 2008).
electron microscopy indicate that ErbB4 is predominantly Further studies of Npas4 targets should help to elucidate the
expressed by parvalbumin-positive cortical interneurons, where molecular mechanisms by which neuronal activity regulates the
it can be found both in the axon and in the postsynaptic density of density of GABAergic synapses and in turn, the balance of
asymmetric synapses formed onto these interneurons (Fazzari excitation and inhibition in the nervous system.
et al., 2010).
After making this discovery, the Marin and Rico groups went on 3.8. Semaphorins
to perform a genetic analysis to demonstrate that ErbB4 is cell-
autonomously required in interneurons to mediate GABAergic The Semaphorin family of proteins consists of over 25 members
synapse formation. To reach this conclusion, the authors grouped into eight different classes based largely on their sequence
performed both a morphological and functional analysis of homology and protein domain structures. The hallmark of a
parvalbumin-positive chandelier cells in which ErbB4 expression Semaphorin family member is the extracellular Semaphorin
was eliminated using Cre-mediated recombination in two (Sema) domain: a conserved, cysteine-rich region of 500 amino
independent strains of mice and found that the number of acids at the N-terminus of the protein (Yazdani and Terman, 2006).
GABAergic synapses made by these cells was reduced (Fazzari Semaphorins are perhaps best known by the function of the
et al., 2010). Performing similar experiments, the authors did not secreted, Class 3 Semaphorins as axon guidance molecules in the
observe a requirement for ErbB4 in pyramidal neurons to form developing nervous system, although the majority of Semaphorins,
glutamatergic synapses onto other pyramidal cells (Fazzari et al., including Class 4 Semaphorins, are in fact membrane-associated
2010). In addition, these effects are presumably mediated by a proteins. Semaphorin function has been implicated in numerous
Nrg1/ErbB4 ligand-receptor interaction, as overexpression of Nrg1 developmental processes outside of the nervous system, including
by pyramidal neurons increases the number of GABAergic lymphocyte specification and signaling, angiogenesis, cell migra-
synapses formed onto these cells (Fazzari et al., 2010). Thus, the tion, and vascular and heart morphogenesis (Kruger et al., 2005;
loss-of-function and gain-of-function data agree and support a Yazdani and Terman, 2006). Currently, there are numerous
requirement for ErbB4 signaling in interneurons to mediate examples of Semaphorin family members mediating synapse
GABAergic synapse formation. development in both vertebrate and invertebrate systems (Bou-
zioukh et al., 2006; Godenschwege et al., 2002; Lin et al., 2007;
3.6. FGF7 Morita et al., 2006; Murphey et al., 2003; O’Connor et al., 2009;
Paradis et al., 2007; Pecho-Vrieseling et al., 2009; Tran et al., 2009).
Fibroblast Growth Factors (FGFs) are secreted glycoproteins Our work has functionally implicated two members of the Class
which mediate their diverse biological functions in large part 4 Semaphorins in glutamatergic and/or GABAergic synapse
through binding to and activating FGF receptor tyrosine kinases formation: Sema4B and Sema4D. Importantly, our work has
(Turner and Grose, 2010). The range of processes in which FGF identified Class 4 Semaphorins as novel regulators of GABAergic
M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87 77

synapse formation. Specifically, RNAi-mediated knockdown of part by the calcium-responsive CREB transcription factor (Cohen
Sema4B in the postsynaptic neuron leads to a decrease in the and Greenberg, 2008). Investigation into the function of BDNF in
density of both glutamatergic and GABAergic synapses in cultured the CNS has focused on modulatory effects of secreted BDNF on
hippocampal neurons, as assessed by immunostaining and glutamatergic and GABAergic synaptic transmission, synaptic and
electrophysiological measurements (Paradis et al., 2007) (Fig. 2). structural plasticity, and synapse development (Gottmann et al.,
In addition, RNAi-mediated knockdown of Sema4D in the 2009; Greenberg et al., 2009; Yoshii and Constantine-Paton, 2010).
postsynaptic neuron leads to a decrease in the density of A role for BDNF in GABAergic synapse formation has been
GABAergic synapses without an apparent effect on glutamatergic elucidated by a number of beautiful studies (see references
synapses (Paradis et al., 2007). Interestingly, knockdown of contained in Gottmann et al., 2009), a subset of which will be
Sema4B and Sema4D in the postsynaptic neuron leads to a highlighted here. To circumvent the gross morphological defects
decrease in the density of postsynaptic specializations of and lethality associated with deletion of the BDNF gene, Huang and
GABAergic synapses while having no effect on the density of colleagues took a gain-of-function approach to studying BDNF in
presynaptic specializations (Paradis et al., 2007). This result the CNS. Specifically, they expressed BDNF in forebrain under the
implies that Sema4B and Sema4D might be acting in the control of the a-CaMKII promoter and found that maturation of
postsynaptic neuron to recruit neurotransmitter receptors or in GABAergic interneurons and presynaptic terminals is accelerated
other ways organize the postsynaptic density. in visual cortex (Huang et al., 1999). Using a related approach,
To date, the focus of most studies of Semaphorin signaling has Aguado et al. overexpressed BDNF in CNS progenitors and assayed
been on the function of Semaphorins as ligands for various GABAergic maturation both by in situ hybridization and electron
receptors (i.e. ‘‘forward signaling’’). However, as Class 4 Sema- microscopy. They found that GAD65/67 mRNA is dramatically
phorins are membrane-associated proteins with a C-terminal upregulated in hippocampi in which BDNF is overexpressed and
intracellular domain, the possibility of a role for Semaphorins as further, that the density of GABA-positive terminals is increased
bidirectional signaling proteins exists (i.e. both ‘‘forward’’ and (Aguado et al., 2003). Similarly, exogenous treatment of cultured
‘‘reverse signaling’’). Currently, the best candidate for a Sema4D hippocampal neurons with BDNF increases both the number of
receptor to mediate its effects on synapse development is the GABAAR clusters and the percentage of synaptically localized
transmembrane receptor PlexinB1 (PlxnB1). Sema4D has been GABAAR clusters (Elmariah et al., 2004).
shown to bind to PlxnB1 in heterologous cells with high affinity The Reichardt group took a loss-of-function approach to
(Tamagnone et al., 1999). In addition, PlxnB1 has been reported to studying BDNF signaling in synaptogenesis. To circumvent the
be expressed in hippocampal neurons by antibody staining lethality issues associated with constitutive knockout of genes in
(Swiercz et al., 2002), and by in situ hybridization (Brain Atlas the BDNF signaling pathway, Rico and colleagues generated a
of Gene Expression; Magdaleno et al., 2006). Thus, in the case of conditional allele of the trkB gene to produce mice in which trkB
Sema4D ‘‘reverse signaling,’’ Sema4D might be activated by was deleted in the cerebellum (Rico et al., 2002). The cerebellum in
binding to PlxnB1. Alternatively, Sema4D could ‘‘forward signal’’ these mice is normal with no change in cell number and
either as a soluble factor or membrane-bound factor to activate morphology, and the overall cerebellar architecture is grossly
PlxnB1. As our work demonstrates that loss of Sema4D in the normal as well (Rico et al., 2002). Both immunohistochemistry and
postsynaptic neuron leads to a decrease in GABAergic synapse electron microscopy revealed a deficit in the number of GABAergic
density in this same neuron (Paradis et al., 2007), the most synapses formed onto granule cells, with no change in the number
parsimonious model suggests that Sema4D might act in an of glutamatergic synapses formed onto these cells (Chen et al.,
autocrine manner by binding to PlxnB1 expressed on the same 2011; Rico et al., 2002). The Reichardt group went on to
cell which releases Sema4D. Current work is focused on testing demonstrate a cell-autonomous requirement for TrkB in both
these models. the presynaptic Golgi cell to cluster presynaptic GABAergic
proteins such as GAD67, and the postsynaptic granule cell to
3.9. Brain-derived neurotrophic factor (BDNF) cluster postsynaptic GABAergic proteins such as gephyrin (Chen
et al., 2011). Further, use of a TrkB ‘‘knock in’’ in which the kinase
BDNF is a member of the neurotrophin family of peptide growth activity of TrkB can be inhibited by feeding the animals 1NMPP1
factors that were originally isolated and described based on their (Chen et al., 2005) revealed that TrkB is required both in the intial
ability to promote survival and differentiation of peripheral stages of GABAergic synapse assembly (P0–28) and also later in
nervous system (PNS) neurons. Since that time, a multitude of development for synapse maintenance (P30–50) (Chen et al.,
functions for BDNF in the developing and mature PNS and CNS 2011).
have been described (Ernfors et al., 1994; Gottmann et al., 2009; Similarly, Kohara and colleagues injected virus expressing Cre
Huang and Reichardt, 2001; Jones et al., 1994; Riddle et al., 1996; recombinase into the cortex of mice containing a floxed allele of
Russo et al., 2009; Yoshii and Constantine-Paton, 2010). BDNF is a the BDNF gene. Immunohistochemical analysis of neurons in which
secreted protein that exerts its signaling effects through binding to the BDNF gene had been deleted by Cre-mediated recombination
its high-affinity receptor TrkB, a receptor tyrosine kinase, and a low revealed fewer GAD65-positive boutons onto the soma of
affinity receptor p75NTR (Huang and Reichardt, 2001). Interesting- excitatory neurons and a decrease in the frequency of mIPSCs
ly, BDNF is thought to be predominantly secreted from excitatory, (Kohara et al., 2007), consistent with a decrease in GABAergic
pyramidal neurons in cortex and hippocampus (Kokaia et al., 1993; synapse number. Again, BDNF is expressed by the excitatory
Miranda et al., 1993) while its receptor, TrkB, is expressed by both neurons and not the interneurons, suggesting that local secretion
pyramidal cells and GABAergic interneurons (Cabelli et al., 1996; of BDNF by the target cell influences whether a GABAergic synapse
Cellerino et al., 1996). Thus, secreted BDNF is poised to act in both gets made onto that cell. Taken together, both overexpression and
an autocrine and paracrine manner. loss-of-function analyses of BDNF signaling indicate a requirement
BDNF has an unusual gene structure: transcription is initiated for BDNF signaling to mediate formation of GABAergic synapses in
at eight distinct promoters resulting in, via alternative splicing and multiple brain regions.
polyadenylation, at least eighteen different transcripts that Until recently, it has been difficult to tease apart the activity and
remarkably all code for an identical BDNF protein. One of the non-activity-dependent functions of BDNF using straight loss of
interesting aspects of this gene structure is that BDNF transcription function or overexpression approaches. However, an elegant set of
from promoter IV is dependent on neuronal activity, mediated in experiments from the laboratory of Michael Greenberg directly
78 M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87

and definitively addressed this issue. Hong and colleagues unaltered ACM (Hughes et al., 2010). Thus, the identity of this
generated a ‘‘knock in’’ mouse in which the CREB binding site in secreted factor or factors remains elusive.
the activity-dependent promoter IV of the BDNF gene was In summary, astrocytes play an important role in both
mutated; they went on to show that transcription from promoter GABAergic and glutamatergic synapse development, but so far
IV is no longer responsive to activity in this animal (Hong et al., at least, they exert their effects on these two types of synapses
2008). Importantly, the authors found that GABAergic synapse through distinct molecular mechanisms.
development is perturbed in the cortex of mice that are unable to
express BDNF in response to neuronal activity. Specifically, both 4. Role of GABAergic synaptic transmission in synapse
immunohistochemistry and an analysis of mIPSC frequency development
revealed a decrease in the density of GABAergic synapses formed
onto layer II/III pyramidal cells in primary visual cortex (Hong There is no doubt that neuronal activity is an important
et al., 2008). Thus, the in vivo function of BDNF that is produced in modulator of synapse development (Anderson et al., 2004;
response to activity is to promote GABAergic synapse develop- Chattopadhyaya et al., 2004; Fritschy and Brunig, 2003; Katz
ment. This finding has far-reaching implications for cortical and Shatz, 1996; Rao et al., 2000). However, activity is not
development in general as it suggests that circuits respond to absolutely required as synapses still form in the absence of activity
sensory experience by dampening excitation through increased (Craig et al., 1994; Verhage et al., 2000; Studler et al., 2002;
inhibition. Varaqueaux et al., 2002; Harms and Craig, 2005). For example,
deletion of the protein Munc-18, which is required for synaptic
3.10. Astrocyte-derived factors vesicle exocytosis, leads to a complete loss of neurotransmitter
release from the presynaptic terminal (Verhage et al., 2000).
Astrocytes were long thought to function as merely ‘‘support’’ Despite this profound deficit in synaptic transmission, synapses
cells in the CNS, playing an important role in maintaining a healthy form proper morphological structures as determined by electron
metabolic state of neurons (Coles and Abbott, 1996; Tsacopoulos microscopy and immunohistochemistry (Verhage et al., 2000).
and Magistretti, 1996). However, they are now recognized as Similarly, simultaneous deletion of the Munc13-1 and Munc13-2
having crucial cell signaling roles in a structure that has been proteins, which are essential for vesicle priming, causes the
coined a ‘‘tripartite’’ synapse (Haydon and Carmignoto, 2006). The complete absence of neurotransmitter release (Varaqueaux et al.,
tripartite synapse contains astrocytic projections that closely 2002). However, morphologically normal synapses form in
appose the presynaptic and postsynaptic neuronal termini at neuronal cultures derived from the hippocampi of these animals
classical chemical synapses (Araque et al., 1999; Haydon and (Varaqueaux et al., 2002).
Carmignoto, 2006). Despite the finding that activity is not an absolute requirement
With respect to synapse development, astrocyte-derived for synapse development, the fact is that synapses do develop in
factors play an important role in both glutamatergic and the presence of activity in the intact nervous system. Furthermore,
GABAergic synapse formation and development (Boehler et al., there exists a growing body of literature that supports the idea that
2007; Elmariah et al., 2005; Hughes et al., 2010; Nagler et al., 2001; GABAergic synaptic transmission contributes to both glutamater-
Pfrieger and Barres, 1997; Ullian et al., 2004, 2001). Co-culture of gic and GABAergic synapse development, as outlined below.
neurons and glia, in conjunction with neuronal growth in
astrocyte-conditioned media (ACM), demonstrates that astrocytes 4.1. Depolarizing actions of GABA: relevance to glutamatergic synapse
release molecules that are necessary to maintain and promote the development
structure and function of early synapses (Beattie et al., 2002;
Faissner et al., 2010; Nagler et al., 2001; Pfrieger and Barres, 1997; Unlike GABA’s action in mature neurons, where the cell is
Slezak and Pfrieger, 2003; Ullian et al., 2004, 2001). In the past hyperpolarized after GABAAR channel opening, binding of GABA to
decade, much effort has been devoted to biochemical purifications GABAARs in immature neurons leads to depolarization (Fig. 3)
of astrocyte-conditioned media in order to identify the factors that (Ben-Ari, 2002; Ben-Ari et al., 2007; Khazipov et al., 2004; Tyzio
mediate these effects with great success (Christopherson et al., et al., 2008). This is the result of a change in the GABAA reversal
2005; Dowell et al., 2009; Hughes et al., 2010; Jeon et al., 2010). For potential due to changes in the Cl gradient across the cell
example, the extracellular matrix proteins Thrombospondins membrane, which is established by K+–Cl transporters. The
(TSPs) are both necessary and sufficient for aspects of glutama- expression of these transporters is dynamically regulated during
tergic synapse formation in cultured retinal ganglion cells development (Fig. 3C) (Inoue et al., 1991; Takebayashi et al., 1996).
(Christopherson et al., 2005). For example, during embryonic and early postnatal development
Research on GABAergic synapse formation and astrocytes is still the Na+–K+–Cl transporter NKCC1 is preferentially expressed in
in its infancy. Early work suggested that astrocytes play an neurons. This protein actively pumps Na+, K+ and Cl across the
important role at GABAergic synapses because hippocampal membrane into the cell, causing an increase in intracellular Cl
neurons grown either on astrocytes or with ACM had larger concentration (Fig. 3B). Postnatally, the expression of another K+–
GABA-induced Cl currents relative to neurons that were grown in Cl transporter, KCC2, dramatically increases between P0 and P9 in
the absence of astrocytes (Elmariah et al., 2005; Liu et al., 1997, the rat hippocampus (Rivera et al., 1999). KCC2 pumps Cl across
1996). In addition, ACM increases the number of GABAergic the membrane and out of the cell, thereby reversing the Cl
presynaptic terminals, the frequency of mIPSCs, and the number equilibrium potential, thus rendering GABA as hyperpolarizing
GABAAR clusters during the first 10 DIV of hippocampal neuron (Fig. 3B). In the hippocampus, the expression of NKCC1 peaks at P7
cultures (Elmariah et al., 2005). This increase in GABAergic synapse and decreases substantially by P20 (Wang et al., 2002). Likewise,
development is due to a secreted protein(s) from astrocytes, as the expression of KCC2 in the neocortex begins to increase in
conditioned media from fibroblasts was unable to alter synapse expression around P10 and peaks in the adult rat (Dzhala et al.,
density in neuron cultures, and trypsizination of ACM blocks 2005). This change in expression from NKCC1 to KCC2 marks the
increases in GABAergic synapse density (Hughes et al., 2010). The developmental switch from GABAergic depolarization to hyper-
authors ruled out TSPs because ACM that is immuno-depleted of polarization. Interestingly, GABA-mediated depolarization and
endogenous TSPs using antisera against TSP-1 and TSP-2 is still subsequent Ca2+ influx into the neuron drives this developmental
able to induce GABAergic synapse formation when compared to switch (Ganguly et al., 2001).
M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87 79

Fig. 3. (A) GABAA receptors are permeable to Cl ions. The direction of Cl ion flow depends on the electrochemical driving force (larger Cl indicates higher Cl
concentration). (B) The Cl concentration gradient is established by the expression of the chloride co-transporters NKCC1 and KCC2. (C) Expression of KCC2 and NKCC1 is
developmentally regulated, with NKCC1 predominating early in development and KCC2 later in development. (D) GABAA receptor activation after release of GABA from
interneurons causes an outward flow of Cl ions and subsequent membrane depolarization during early development. This efflux of negative ions across the membrane is
equivalent to an influx of positive ions and thus is depicted as an inward current. In later development, a hyperpolarization of the GABAA reversal potential occurs due to a
change in the Cl gradient across the cell membrane, where the extracellular Cl concentration is now higher than the intracellular Cl concentration. The change in the Cl
gradient causes an inward flow of Cl ions through GABAA receptors which is depicted as an outward current.

The importance of the timing of this shift from depolarization to perturbing GABAergic synaptogenesis (Wang and Kriegstein,
hyperpolarization and its relationship to glutamatergic synapse 2008). The authors propose that this NKCC1-dependent increase
development is exemplified by perturbations in the expression of in glutamatergic synaptogenesis acts through NMDAR activation
either KCC2 or NKCC1. In one study, premature KCC2 expression to recruit AMPARs to silent synapses (Wang and Kriegstein, 2008).
prompted an increase in GABAergic synapse density, without However, it should be noted that relatively recent work
altering the density of glutamatergic synapses, thus tipping the (Holmgren et al., 2010; Rheims et al., 2009; Tyzio et al., 2011;
delicate equilibrium between excitation and inhibition (Chudot- Zilberter et al., 2010) but see (Ben-Ari et al., 2007; Tyzio et al.,
vorova et al., 2005; Hubner et al., 2001). Similarly, an electrophys- 2011) has called into question the depolarizing effects of GABA
iological study from Xenopus laevis tectal neurons demonstrated during early postnatal development. The controversy, summarized
that manipulating EGABA in vivo by prematurely expressing KCC2 nicely in Khakhalin (2011), stems from the use of glucose as the
leads to a strengthening of GABAergic inputs onto tectal neurons sole energy source in artificial cerebral spinal fluid (ACSF) for in
(Akerman and Cline, 2006). However, in contrast to the previous vitro neuronal preparations and electrophysiological recordings.
work, the KCC2 pertubation in this study prevented the maturation Supplementing ACSF with other energy sources such as pyruvate
of functional glutamatergic synapses (Akerman and Cline, 2006). In caused a shift in the reversal potential of GABA towards more
addition, RNAi-mediated knockdown of NKCC1 expression in early hyperpolarized potentials and further, an amelioration of GDPs
development inhibits glutamatergic synaptogenesis, without (see Section 4.2 for description of GDPs). This suggests that the
80 M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87

depolarizing effects of GABA are an experimental artifact and from the same group demonstrated that either intraocular
therefore not biologically relevant. The resolution of this contro- injection of tetrodotoxin (TTX) or activity blockade by TTX
versy will not come quickly for sure, but hopefully will be aided by application to cortical organotypic cultures at discrete time points
an analysis of transgenic mouse models in which intracellular Cl caused decreased perisomatic synapse density of pyramidal
concentration is changed by deletion or overexpression of Cl neurons (Chattopadhyaya et al., 2004). However, in this study it
transporters and further, by supplementation of ACSF with is unclear whether the TTX-dependent effects on synapse
alternative energy sources for future experiments (Khakhalin, development reflect a change in GABAergic or glutamatergic
2011). synaptic transmission or both. While these intriguing results
suggest that GABA signaling is required for GABAergic innervation
4.2. Giant depolarizing potentials (GDPs) of pyramidal cells, it is difficult to separate the effects of decreased
axon branching from synapse development per se in these studies.
The depolarizing capability of GABA release in early postnatal Further, a study using organotypic hippocampal culture found that
development is important for regulating Ca2+ currents, and could treatment with the GABAAR antagonist bicuculline increased the
be used to facilitate circuit formation and synaptogenesis through density of GABAergic presynaptic terminals (Marty et al., 2000).
synchronous activity (Ben-Ari et al., 2007). GDPs are polysynaptic, While it is difficult to explain the disparate results from these two
slowly propagating network oscillations that have been observed studies, possibilities include the area of the brain studied and the
both in vivo and in vitro in the developing hippocampus in rodents timing and duration of pharmacological manipulations.
and in primates (Ben-Ari et al., 2007). GDPs are the first Evidence from Purkinje cells of the cerebellum suggests that
synaptically generated oscillations in the hippocampus (Allene different synapses (i.e. axo-dendritic vs. axo-somatic) are differen-
and Cossart, 2010) and are present during the first postnatal week, tially affected by loss of GABAergic transmission. Electrophysiology
during a time when the action of GABA is excitatory (Ben-Ari et al., and immunoelectron microscopy analysis of a mouse in which the
1989, 2007). They subside by the second postnatal week (Ben-Ari a1 subunit of the GABAAR is deleted and therefore lacks GABAergic
et al., 2007); correlating with the shift from predominantly synaptic transmission reveals that GABAergic synapses from basket
GABAergic synapse development to the maturation of glutama- cells form normally onto the soma but not the dendrites of Purkinje
tergic synapses (Durand and Konnerth, 1996; Hennou et al., 2002; cells (Fritschy et al., 2006). In addition, GABAergic axons from basket
Tyzio et al., 1999). cells go on to form inappropriate synapes onto Purkinje cell dendritic
spines (Fritschy et al., 2006). Another study engineered a mutation
4.3. A model for the synchronized development of GABAergic and into the b3 subunit of GABAARs, which prevents endocytosis at the
glutamatergic synapses in hippocampus synapse from occurring, and found an increase in both the size and
the number of GABAergic synapses (Jacob et al., 2009). Neurons with
The above data, combined with the fact that GABAergic synapse increased number of synaptic GABAARs exhibit an increase in the
development initiates before glutamatergic synapse development ratio of immature filopodial-like spines to mature, mushroom-like
(see Section 2.4, above) has led to the following model for synapse spines as well as a decrease in PSD-95 expression (Jacov et al., 2009).
development in the hippocampus, largely based upon work by Dr. Interestingly, the immature spine morphology and decrease in PSD-
Ben-Ari and colleagues (see Ben-Ari, 2002; Ben-Ari et al., 2007 and 95 expression is rescued by pharmacological blockade of GABAARs
references therein). To begin, the development of depolarizing, (Jacob et al., 2009), indicating that increased GABAergic transmis-
GABAergic synapses drives glutamatergic synapse development. In sion is responsible for the delay in excitatory synapse development.
particular, the depolarizing nature of early GABAergic synaptic
transmission is sufficient to relieve the NMDAR Mg2+ block, thus 5. GABAergic synapse development and neurological disorders
allowing for Ca2+ entry through NMDARs in response to glutamate
release. This in turn drives AMPAR insertion into NMDAR-only It is now widely believed that aberrant development of
containing ‘‘silent’’ synapses. In addition, the depolarizing nature excitatory and/or inhibitory synapses contributes to neurological
of early GABA release drives the GDPs and subsequent Ca2+ influx impairments such as mental retardation, autism spectrum
into neurons, which is required for the formation of proper disorders (ASDs) and epilepsy (Bear et al., 2004; Fernandez and
hippocampal networks. Garner, 2007; Hong et al., 2005; Rubenstein and Merzenich, 2003;
Tabuchi et al., 2007; Zoghbi, 2003). Numerous genetic studies on
4.4. Role of activity in GABAergic synapse development human populations with ASDs have found linkage to genes that
regulate synapse development, axon pathfinding, and neuronal
Not only does GABAergic activity initiate glutamatergic morphology (Gilman et al., 2011; Jamain et al., 2003; Levy et al.,
synaptogenesis as discussed above, it may strengthen existing 2011; Sanders et al., 2011; Szatmari et al., 2007). Thus, one
GABAergic synapses. The 65 kDa GAD65 and 67 kDa GAD67 hypothesis to explain ASDs posits that changes in circuit
proteins are isoforms of the enzyme that produces GABA, glutamic connectivity alter information processing in affected individuals.
acid decarboxylase (GAD) (Soghomonian and Martin, 1998). The Below, we have chosen to discuss in depth Rett syndrome and ASDs
Huang group created a conditional allele of the gene encoding in which changes in inhibitory drive contribute to the disorder, as
GAD67 (i.e. Gad1). Next, they simultaneously knocked out an emerging model for a connection between synaptogenic
expression of GAD67 in interneurons while labeling them with proteins and these disorder.
GFP by transfection of an interneuron-specific GFP–IRES–Cre
recombinase construct in organotypic cultures derived from 5.1. Rett syndrome
cortex of these mice (an analysis of the knockout animal is
complicated by lethality) (Chattopadhyaya et al., 2007). Using Rett syndrome is an X-linked developmental disorder that
immunostaining and imaging, the group found that knockdown of results from mutations in a single gene, Mecp2 (Amir et al., 1999).
GAD67 in basket cells leads to a profound defect in axon branching Mecp2 encodes methyl-CpG binding protein-2, which binds
and perisomatic synapse formation onto pyramidal neurons in methylated CpG sequences in vertebrate genomes (Tate et al.,
cortex, and these defects can be rescued by increasing GABA 1996). It is believed that methylated CpG islands recruit other
lifetime in the synaptic cleft by pharmacological blockade of GABA proteins that can lead to downstream transcriptional regulation.
transporters (Chattopadhyaya et al., 2004, 2007). A previous study Loss of MeCP2 can lead to both activation and repression of
M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87 81

transcription across the genome. However, 85% of genes assayed this analysis that loss of Mecp2 from GABAergic neurons
in Mecp2 null mice by Zoghbi and colleagues are downregulated, recapitulates features of Rett syndrome (Chao et al., 2010). In
suggesting a bias towards transcriptional activation for MeCP2 addition, the authors go on to show that deletion of MeCP2 in
function (Chahrour et al., 2008). GABAergic neurons decreases the amount of GABA by altering
Rett syndrome affects approximately 1:10,000 girls worldwide Gad1/2 expression, leading to a decrease in mIPSC quantal size
(Percy and Lane, 2005). Females with Rett syndrome develop (Chao et al., 2010). This observation differs from studies using mice
normally for the first 6–18 months of life at which point there is a in which Mecp2 was deleted in all neurons that show a slight
rapid regression in development. Interestingly, this age is after increase in mIPSC quantal size (Dani et al., 2005).
neurogenesis and migration have occurred, and corresponds with Clearly, much work remains to sort out the differences between
the period of synaptic development that includes later stages of the various mouse models in terms of behavior, cellular physiolo-
synaptogenesis and synaptic pruning (Zoghbi, 2003). In humans, gy, and circuit development to uncover the key sites of action of
onset of symptoms typically involves a deceleration in head and MeCP2 function. In addition, as MeCP2 functions as a global
body growth along with a regression in motor and speech abilities regulator of gene expression, further studies are needed to
and abnormal breathing patterns. Cognitive impairments are also elucidate the specific molecular mechanisms which lead to
present and manifest themselves similarly to ASDs (Percy and alterations in excitatory/inhibitory ratio in mouse models of Rett
Lane, 2005). syndrome. Taken together, these findings, coupled with the fact
There are a number of mouse models of Rett syndrome. Two of that the appearance of Rett symptoms in humans occurs during a
these models are deletions in the Mecp2 gene: the Jaenisch mouse period of rapid synaptogenesis, suggest that a defect in synapse
lacks exon 3 of Mecp2 (Chen et al., 2001), and the Bird mouse lacks development or maintentnace may be the underlying cause of Rett
both exons 3 and 4 (Guy et al., 2001). Mecp2 null mice (Jaenisch and syndrome. In addition, these studies point to the importance of
Bird models) appear normal until about 5 weeks postnatally. At maintaining the appropriate balance of excitation and inhibition
this point, mice begin to demonstrate abnormal behaviors for proper cortical circuit function.
resembling Rett syndrome, such as hindlimb clasping and irregular
breathing (please see reviews for further discussion of symptoms 5.2. Autism spectrum disorders
(Cobb et al., 2010; Zoghbi, 2003).
Classically, Rett syndrome has been thought of as a disorder of ASDs are diagnosed based on the presence of three behavioral
neural development. However recent work has extended this features: a deficit in social communication, stereotypies, and a
intepretation, as mice in which Mecp2 is deleted in adulthood delay or absence of language skills (van Spronsen and Hoogenraad,
display phenotypes similar to mice in which Mecp2 is constitu- 2010). ASDs are typically detected and diagnosed before the
tively deleted (McGraw et al., 2011). This suggests that Mecp2 second or third year of life. Again, this time frame is potentially
expression is required throughout the life of the animal for normal important for understanding the underlying pathology of the
nervous system function. Interestingly, a recent study from the C. disorder because, as with Rett syndrome, the onset and detection
Chen lab examining development of the retinogeniculate synapse of the disorder corresponds with a period of robust synapse
in the Bird Mecp2 null mouse found aberrant remodeling of this formation and maturation in humans (Lord et al., 2000). There is a
synapse, specifically during the phase of development which strong genetic basis to these disorders as they are highly heritable
requires visual experience (Noutel et al., 2011). The implication of (80–90%: Bailey et al., 1995; van Spronsen and Hoogenraad,
this finding is that defects in activity-dependent synapse 2010) with a large degree of genetic heterogeneity. Interestingly,
development occur in the absence of MeCP2. linkage analyses on families with ASDs have revealed mutations or
Although the behavioral symptoms of Rett syndrome are single nucleotide polymorphisms (SNPs) in many synaptic
prominent in human patients as well as in the mouse models, the proteins. These synaptic proteins include the cell adhesion
neuropathological phenotypes are relatively subtle. CA2 pyramidal molecules Neurexin and Neuroligin (Abrahams and Geschwind,
neurons in Mecp2 null mice are 15–25% smaller than the 2008; Jamain et al., 2003; Marshall et al., 2008; Tabuchi et al.,
corresponding neurons in wild type littermate control mice (Chen 2007), Semaphorins (Weiss et al., 2009), and Shank family
et al., 2001). In addition, the neocortical projections in Mecp2 null members (Moessner et al., 2007; Pinto et al., 2010) as well as
mice have simplified dendritic arbors (Kishi and Macklis, 2004). In many GTPases, which play a role in regulating cytoskeletal
layer5 of the somatosensory cortex of the Jaenisch Mecp2 null mice, dynamics in dendrites and dendritic spines (Pinto et al., 2010).
spontaneous activity of pyramidal neurons is reduced (Dani et al., A small percentage of ASD patients have mutations in
2005). While there is no change in the intrinsic properties of the Neuroligin 3. One of these is a change from an arginine to cysteine
cortical neurons studied, there is a shift in the balance between at amino acid 451 of the coding sequence (R451C) (Jamain et al.,
excitation and inhibition, which favors inhibition (Dani et al., 2003). To ask if this mutation could mimic ASDs in a rodent model,
2005). The increase in inhibition is due to a reduced connectivity a genetic knock-in strategy was utilized to introduce the R451C
and weaker excitatory synapses in the Mecp2 null mice compared mutation into the Nlgn3 gene in mice. Employing this experimental
to wild-type littermates (Chao et al., 2007; Dani and Nelson, 2009). approach, one group found that the Neuroligin 3 R451C knock-in
Similarly to the somatosensory cortex, a decrease in the excitatory mouse displays increases in both GABAergic synaptic number and
drive and an increase of inhibitory drive was detected in the strength and autism-like behavioral deficits such as impaired
thalamus of the Bird Mecp2 null mice (Zhang et al., 2010b). social interactions (Tabuchi et al., 2007), suggesting a role for
A recent study again points to an important role for MeCP2 in GABAergic synaptic function and development in ASDs. However,
regulating inhibition in the nervous system (Chao et al., 2010), another group using this same strategy did not find any behavioral
although the conclusions from this study differ from those deficits in their Neuroligin 3 R451C mouse model (Chadman et al.,
discussed above. In this study, Zoghbi and colleagues generate a 2008); electrophysiological recordings were not reported in this
knockout of Mecp2 that is confined to GABA-expressing neurons study. Thus, a conclusion with respect to the role of the R451C
and demonstrate that in this mouse, many of the same aspects of mutation in synapse development and ultimately, ASD-like
nervous system function are impaired as has been observed with phenotypes, awaits further investigation.
the Mecp2 null animals, such as motor control, breathing, and As mentioned previously, the Semaphorins have recently
weight regulation (Chao et al., 2010). The authors also find a broad emerged as important mediators of both glutamatergic and
spectrum of behavioral deficits in this mouse and conclude from GABAergic synapse development (Bouzioukh et al., 2006; God-
82 M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87

enschwege et al., 2002; Lin et al., 2007; Morita et al., 2006; their own right. However, much remains to be discovered. For
Murphey et al., 2003; O’Connor et al., 2009; Paradis et al., 2007; example, live-imaging studies that observe components of
Pecho-Vrieseling et al., 2009; Tran et al., 2009). Interestingly, in the GABAergic synapses assembling at nascent synaptic sites are
largest genome-wide association study of ASDs to date, a study of needed to inform our understanding of the steps in GABAergic
1,553 patients led by the Daly and Chakravarti labs, the strongest synapse development. More gene discovery efforts, in any
genome-wide-significant association involved markers adjacent to organism, geared at identifying molecules that regulate all steps
SEMA5A on chromosome 5p15 (Weiss et al., 2009). In addition, two in GABAergic synapse development (contact stabilization, recruit-
studies have reported altered regulation of SEMA5A mRNA in ing pre- and post-synaptic components to contact sites) must be a
patients with ASDs (Melin et al., 2006; Weiss et al., 2009). significant area of focus in the near future.
Interestingly, Sema5a is strongly expressed in the dentate gyrus of Much of the data presented in this review points to the idea that
the hippocampus and in layers 5 and 6 of the cortex in rodents similar mechanisms regulate glutamatergic and GABAergic syn-
(Lein et al., 2007), two areas where morphological changes apse development, as observed at the molecular level. No longer
correlate with ASDs (Hutsler and Zhang, 2010; Schumann et al., can we be content to think of these synaptic subtypes developing
2004). The prevalence of SEMA5A SNPs in autistic patients, coupled via independent mechanisms. It is abundantly clear that gluta-
with studies that demonstrate that Semaphorin family members matergic and GABAergic synapses co-develop in time and in space,
mediate both glutamatergic and GABAergic synapse development, utilize the same or homologous signal transduction pathways, and
suggests the intriguing possibility that the underlying cause of ASD have a complicated, activity-based relationship which influences
symptoms in patients with Sema5A mutations may be defects in their co-development. In addition, the field of synapse develop-
synapse development. ment must confront two interesting observations which promise
Scaffolding proteins and cytoskeletal regulation are important to further influence how we think about shared mechanisms of
in synapse development. Genetic mapping has implicated the synapse development: co-release of glutamate and GABA from
scaffolding proteins of the Shank family in ASDs (Durand et al., central synapses, and co-innervation of dendritic spines by both
2007; Marshall et al., 2008; Moessner et al., 2007; Pinto et al., glutamatergic and GABAergic synapses.
2010). Shank proteins bind many synaptic proteins and together
with Homer can synergistically increase dendritic spine size Acknowledgements
through formation of a scaffolding network in the dendritic spine
(Hayashi et al., 2009; Sala et al., 2001). It will be interesting to Work in the Paradis laboratory is supported by the Richard and
determine if the mutations in Shank family members that are Susan Smith Family Foundation, Chestnut Hill, MA, an Alfred P.
linked to ASDs cause an increase or a decrease in the amount of Sloan Research Fellowship, and NIH grants R01NS065856 and
Shank expression and likewise, how this change in expression R21DA026977. We thank Amy Ghiretti, Dr. Sonia Cohen, Dr.
affects synaptic development. In addition, small, Ras-like GTPases Elizabeth Hong, Dr. Anna Moore and anonymous reviewers for
are important regulators of the actin cytoskeleton and as such, helpful comments on this manuscript.
control both dendritic spine development and dendritic branching
(Ghiretti and Paradis, 2011; McNair et al., 2010; Paradis et al.,
References
2007; Tashiro et al., 2000). Interestingly, a recent linkage analysis
demonstrated that copy number variants (CNVs) of Rho family Abrahams, B.S., Geschwind, D.H., 2008. Advances in autism genetics: on the
GTPases are found with increased frequency in individuals with threshold of a new neurobiology. Nat. Rev. Genet. 9, 341–355.
ASDs compared to non-ASD individuals (Pinto et al., 2010). Aguado, F., Carmona, M.A., Pozas, E., Aguilo, A., Martinez-Guijarro, F.J., Alcantara, S.,
Borrell, V., Yuste, R., Ibanez, C.F., Soriano, E., 2003. BDNF regulates spontaneous
For obvious reasons, alterations in synapses are difficult to correlated activity at early developmental stages by increasing synaptogenesis
assess in autistic patients and further, the causal relationship and expression of the K+/Cl co-transporter KCC2. Development 130, 1267–
between any morphological changes and the ASD symptoms is 1280.
Ahmari, S.E., Buchanan, J., Smith, S.J., 2000. Assembly of presynaptic active zones
unclear. Changes in the shape of dendritic spines may affect from cytoplasmic transport packets. Nat. Neurosci. 3, 445–451.
synaptic connections by modulating the chemical microenviron- Akerman, C.J., Cline, H.T., 2006. Depolarizing GABAergic conductances regulate the
ment of the synapse (Kasai et al., 2003; Robinson and Kolb, 2004; balance of excitation to inhibition in the developing retinotectal circuit in vivo.
J. Neurosci. 26, 5117–5130.
Tsay and Yuste, 2004). A recent study demonstrates that post-
Allene, C., Cossart, R., 2010. Early NMDA receptor-driven waves of activity in the
mortem brains of autistic patients have significant increases in developing neocortex: physiological or pathological network oscillations? J.
spine density in the superficial layers of cortical pyramidal Physiol. 588, 83–91.
neurons, and this change was found in every brain from an Amir, R.E., Van den Veyver, I.B., Wan, M., Tran, C.Q., Francke, U., Zoghbi, H.Y., 1999.
Rett syndrome is caused by mutations in X-linked MECP2, encoding methyl-
autisitic individual surveyed in the study (Hutsler and Zhang, CpG-binding protein 2. Nat. Genet. 23, 185–188.
2010). This study is the first of its kind to present a clear correlation Andersen, P., Eccles, J.C., Loyning, Y., 1963. Recurrent inhibition in the hippocampus
between morphological synapses and ASDs. Because spines are with identification of the inhibitory cell and its synapses. Nature 198, 540–542.
Andersen, P., Eccles, J.C., Loyning, Y., 1964. Pathway of postsynaptic inhibition in the
easier experimentally to investigate in post-mortem tissue hippocampus. J. Neurophysiol. 27, 608–619.
samples from humans than other types of synapses, it may be Anderson, T.R., Shah, P.A., Benson, D.L., 2004. Maturation of glutamatergic and
more challenging to uncover the role of inhibitory synapse GABAergic synapse composition in hippocampal neurons. Neuropharmacology
47, 694–705.
development in ASDs. However, one inroad into this problem lies Ango, F., di Cristo, G., Higashiyama, H., Bennett, V., Wu, P., Huang, Z.J., 2004.
in further linkage analysis to identify autism susceptibility loci, Ankyrin-based subcellular gradient of neurofascin, an immunoglobulin family
coupled with increased understanding of the molecular mecha- protein, directs GABAergic innervation at purkinje axon initial segment. Cell
119, 257–272.
nisms of GABAergic synapse development.
Araque, A., Parpura, V., Sanzgiri, R.P., Haydon, P.G., 1999. Tripartite synapses: glia,
the unacknowledged partner. Trends Neurosci. 22, 208–215.
5.3. Conclusions Bailey, A., Le Couteur, A., Gottesman, I., Bolton, P., Simonoff, E., Yuzda, E., Rutter, M.,
1995. Autism as a strongly genetic disorder: evidence from a British twin study.
Psychol. Med. 25, 63–77.
The past decade has brought astounding progress to the field of Balasubramanian, S., Fam, S.R., Hall, R.A., 2007. GABAB receptor association with the
GABAergic synapse development. No longer the forgotten younger PDZ scaffold Mupp1 alters receptor stability and function. J. Biol. Chem. 282,
sibling, only appreciated for its ability to reign-in glutamatergic 4162–4171.
Baldelli, P., Fassio, A., Valtorta, F., Benfenati, F., 2007. Lack of synapsin I reduces the
synaptic transmission, GABAergic synapses are now being studied readily releasable pool of synaptic vesicles at central inhibitory synapses. J.
and appreciated by numerous researchers as signal transducers in Neurosci. 27, 13520–13531.
M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87 83

Baldelli, P., Hernandez-Guijo, J.M., Carabelli, V., Carbone, E., 2005. Brain-derived Chang, Y., Wang, R., Barot, S., Weiss, D.S., 1996. Stoichiometry of a recombinant
neurotrophic factor enhances GABA release probability and nonuniform distri- GABAA receptor. J. Neurosci. 16, 5415–5424.
bution of N- and P/Q-type channels on release sites of hippocampal inhibitory Chao, H.T., Chen, H., Samaco, R.C., Xue, M., Chahrour, M., Yoo, J., Neul, J.L., Gong, S.,
synapses. J. Neurosci. 25, 3358–3368. Lu, H.C., Heintz, N., et al., 2010. Dysfunction in GABA signalling mediates
Bannai, H., Levi, S., Schweizer, C., Inoue, T., Launey, T., Racine, V., Sibarita, J.B., autism-like stereotypies and Rett syndrome phenotypes. Nature 468, 263–269.
Mikoshiba, K., Triller, A., 2009. Activity-dependent tuning of inhibitory neuro- Chao, H.T., Zoghbi, H.Y., Rosenmund, C., 2007. MeCP2 controls excitatory synaptic
transmission based on GABAAR diffusion dynamics. Neuron 62, 670–682. strength by regulating glutamatergic synapse number. Neuron 56, 58–65.
Barros, C.S., Calabrese, B., Chamero, P., Roberts, A.J., Korzus, E., Lloyd, K., Stowers, L., Chattopadhyaya, B., Di Cristo, G., Higashiyama, H., Knott, G.W., Kuhlman, S.J.,
Mayford, M., Halpain, S., Muller, U., 2009. Impaired maturation of dendritic Welker, E., Huang, Z.J., 2004. Experience and activity-dependent maturation
spines without disorganization of cortical cell layers in mice lacking NRG1/ErbB of perisomatic GABAergic innervation in primary visual cortex during a post-
signaling in the central nervous system. Proc. Natl. Acad. Sci. U.S.A. 106, 4507– natal critical period. J. Neurosci. 24, 9598–9611.
4512. Chattopadhyaya, B., Di Cristo, G., Wu, C.Z., Knott, G., Kuhlman, S., Fu, Y., Palmiter,
Barrow, S.L., Constable, J.R., Clark, E., El-Sabeawy, F., McAllister, A.K., Washbourne, R.D., Huang, Z.J., 2007. GAD67-mediated GABA synthesis and signaling regulate
P., 2009. Neuroligin1: a cell adhesion molecule that recruits PSD-95 and NMDA inhibitory synaptic innervation in the visual cortex. Neuron 54, 889–903.
receptors by distinct mechanisms during synaptogenesis. Neural Dev. 4, 17. Cheetham, J.J., Murray, J., Ruhkalova, M., Cuccia, L., McAloney, R., Ingold, K.U.,
Baumann, S.W., Baur, R., Sigel, E., 2001. Subunit arrangement of gamma-aminobu- Johnston, L.J., 2003. Interaction of synapsin I with membranes. Biochem.
tyric acid type A receptors. J. Biol. Chem. 276, 36275–36280. Biophys. Res. Commun. 309, 823–829.
Bear, M.F., Huber, K.M., Warren, S.T., 2004. The mGluR theory of fragile X mental Chen, A.I., Nguyen, C.N., Copenhagen, D.R., Badurek, S., Minichiello, L., Ranscht, B.,
retardation. Trends Neurosci. 27, 370–377. Reichardt, L.F., 2011. TrkB (tropomyosin-related kinase B) controls the assem-
Beattie, E.C., Stellwagen, D., Morishita, W., Bresnahan, J.C., Ha, B.K., Von Zastrow, M., bly and maintenance of GABAergic synapses in the cerebellar cortex. J. Neurosci.
Beattie, M.S., Malenka, R.C., 2002. Control of synaptic strength by glial TNFalpha. 31, 2769–2780.
Science 295, 2282–2285. Chen, G., Trombley, P.Q., van den Pol, A.N., 1995. GABA receptors precede glutamate
Ben-Ari, Y., 2002. Excitatory actions of gaba during development: the nature of the receptors in hypothalamic development; differential regulation by astrocytes. J.
nurture. Nat. Rev. Neurosci. 3, 728–739. Neurophysiol. 74, 1473–1484.
Ben-Ari, Y., Cherubini, E., Corradetti, R., Gaiarsa, J.L., 1989. Giant synaptic potentials Chen, R.Z., Akbarian, S., Tudor, M., Jaenisch, R., 2001. Deficiency of methyl-CpG
in immature rat CA3 hippocampal neurones. J. Physiol. 416, 303–325. binding protein-2 in CNS neurons results in a Rett-like phenotype in mice. Nat.
Ben-Ari, Y., Gaiarsa, J.L., Tyzio, R., Khazipov, R., 2007. GABA: a pioneer transmitter Genet. 27, 327–331.
that excites immature neurons and generates primitive oscillations. Physiol. Chen, X., Ye, H., Kuruvilla, R., Ramanan, N., Scangos, K.W., Zhang, C., Johnson, N.M.,
Rev. 87, 1215–1284. England, P.M., Shokat, K.M., Ginty, D.D., 2005. A chemical-genetic approach to
Ben-Ari, Y., Khalilov, I., Represa, A., Gozlan, H., 2004. Interneurons set the tune of studying neurotrophin signaling. Neuron 46, 13–21.
developing networks. Trends Neurosci. 27, 422–427. Chiappalone, M., Casagrande, S., Tedesco, M., Valtorta, F., Baldelli, P., Martinoia, S.,
Ben-Ari, Y., Khazipov, R., Leinekugel, X., Caillard, O., Gaiarsa, J.L., 1997. GABAA, Benfenati, F., 2009. Opposite changes in glutamatergic and GABAergic trans-
NMDA and AMPA receptors: a developmentally regulated ‘menage a trois’. mission underlie the diffuse hyperexcitability of synapsin I-deficient cortical
Trends Neurosci. 20, 523–529. networks. Cereb. Cortex 19, 1422–1439.
Boehler, M.D., Wheeler, B.C., Brewer, G.J., 2007. Added astroglia promote greater Chih, B., Engelman, H., Scheiffele, P., 2005. Control of excitatory and inhibitory
synapse density and higher activity in neuronal networks. Neuron Glia Biol. 3, synapse formation by neuroligins. Science 307, 1324–1328.
127–140. Chih, B., Gollan, L., Scheiffele, P., 2006. Alternative splicing controls selective trans-
Bogdanov, Y., Michels, G., Armstrong-Gold, C., Haydon, P.G., Lindstrom, J., Pangalos, synaptic interactions of the neuroligin–neurexin complex. Neuron 51, 171–178.
M., Moss, S.J., 2006. Synaptic GABAA receptors are directly recruited from their Chiou, T.T., Bonhomme, B., Jin, H., Miralles, C.P., Xiao, H., Fu, Z., Harvey, R.J., Harvey,
extrasynaptic counterparts. EMBO J. 25, 4381–4389. K., Vicini, S., De Blas, A.L., 2011. Differential regulation of the postsynaptic
Boucard, A.A., Chubykin, A.A., Comoletti, D., Taylor, P., Sudhof, T.C., 2005. A splice clustering of {gamma}-aminobutyric acid type A (GABAA) receptors by colly-
code for trans-synaptic cell adhesion mediated by binding of neuroligin 1 to bistin isoforms. J. Biol. Chem. 286, 22456–22468.
alpha- and beta-neurexins. Neuron 48, 229–236. Christopherson, K.S., Ullian, E.M., Stokes, C.C., Mullowney, C.E., Hell, J.W., Agah, A.,
Boulland, J.L., Qureshi, T., Seal, R.P., Rafiki, A., Gundersen, V., Bergersen, L.H., Lawler, J., Mosher, D.F., Bornstein, P., Barres, B.A., 2005. Thrombospondins are
Fremeau Jr., R.T., Edwards, R.H., Storm-Mathisen, J., Chaudhry, F.A., 2004. astrocyte-secreted proteins that promote CNS synaptogenesis. Cell 120, 421–
Expression of the vesicular glutamate transporters during development indi- 433.
cates the widespread corelease of multiple neurotransmitters. J. Comp. Neurol. Chubykin, A.A., Atasoy, D., Etherton, M.R., Brose, N., Kavalali, E.T., Gibson, J.R.,
480, 264–280. Sudhof, T.C., 2007. Activity-dependent validation of excitatory versus inhibitory
Bouzioukh, F., Daoudal, G., Falk, J., Debanne, D., Rougon, G., Castellani, V., 2006. synapses by neuroligin-1 versus neuroligin-2. Neuron 54, 919–931.
Semaphorin3A regulates synaptic function of differentiated hippocampal neu- Chudotvorova, I., Ivanov, A., Rama, S., Hubner, C.A., Pellegrino, C., Ben-Ari, Y.,
rons. Eur. J. Neurosci. 23, 2247–2254. Medina, I., 2005. Early expression of KCC2 in rat hippocampal cultures aug-
Bowery, N.G., Brown, D.A., 1997. The cloning of GABA(B) receptors. Nature 386, ments expression of functional GABA synapses. J. Physiol. 566, 671–679.
223–224. Cobb, S., Guy, J., Bird, A., 2010. Reversibility of functional deficits in experimental
Bresler, T., Ramati, Y., Zamorano, P.L., Zhai, R., Garner, C.C., Ziv, N.E., 2001. The models of Rett syndrome. Biochem. Soc. Trans. 38, 498–506.
dynamics of SAP90/PSD-95 recruitment to new synaptic junctions. Mol. Cell. Cobb, S.R., Buhl, E.H., Halasy, K., Paulsen, O., Somogyi, P., 1995. Synchronization of
Neurosci. 18, 149–167. neuronal activity in hippocampus by individual GABAergic interneurons. Na-
Brunig, I., Suter, A., Knuesel, I., Luscher, B., Fritschy, J.M., 2002. GABAergic terminals ture 378, 75–78.
are required for postsynaptic clustering of dystrophin but not of GABA(A) Cohen, S., Greenberg, M.E., 2008. Communication between the synapse and the
receptors and gephyrin. J. Neurosci. 22, 4805–4813. nucleus in neuronal development, plasticity, and disease. Annu. Rev. Cell Dev.
Buhl, E.H., Halasy, K., Somogyi, P., 1994. Diverse sources of hippocampal unitary Biol. 24, 183–209.
inhibitory postsynaptic potentials and the number of synaptic release sites. Coles, J.A., Abbott, N.J., 1996. Signalling from neurones to glial cells in invertebrates.
Nature 368, 823–828. Trends Neurosci. 19, 358–362.
Burrone, J., Li, Z., Murthy, V.N., 2006. Studying vesicle cycling in presynaptic Colonnier, M., 1968. Synaptic patterns on different cell types in the different
terminals using the genetically encoded probe synaptopHluorin. Nat. Protoc. laminae of the cat visual cortex. An electron microscope study. Brain Res. 9,
1, 2970–2978. 268–287.
Cabelli, R.J., Allendoerfer, K.L., Radeke, M.J., Welcher, A.A., Feinstein, S.C., Shatz, C.J., Correia, S.S., Bassani, S., Brown, T.C., Lise, M.F., Backos, D.S., El-Husseini, A., Passa-
1996. Changing patterns of expression and subcellular localization of TrkB in faro, M., Esteban, J.A., 2008. Motor protein-dependent transport of AMPA
the developing visual system. J. Neurosci. 16, 7965–7980. receptors into spines during long-term potentiation. Nat. Neurosci. 11, 457–
Ceccaldi, P.E., Benfenati, F., Chieregatti, E., Greengard, P., Valtorta, F., 1993. Rapid 466.
binding of synapsin I to F- and G-actin. A study using fluorescence resonance Cossart, R., Petanjek, Z., Dumitriu, D., Hirsch, J.C., Ben-Ari, Y., Esclapez, M., Bernard,
energy transfer. FEBS Lett. 329, 301–305. C., 2006. Interneurons targeting similar layers receive synaptic inputs with
Cellerino, A., Maffei, L., Domenici, L., 1996. The distribution of brain-derived similar kinetics. Hippocampus 16, 408–420.
neurotrophic factor and its receptor trkB in parvalbumin-containing neurons Craig, A.M., Blackstone, C.D., Huganir, R.L., Banker, G., 1994. Selective clustering of
of the rat visual cortex. Eur. J. Neurosci. 8, 1190–1197. glutamate and gamma-aminobutyric acid receptors opposite terminals releas-
Cesca, F., Baldelli, P., Valtorta, F., Benfenati, F., 2010. The synapsins: key actors of ing the corresponding neurotransmitters. Proc. Natl. Acad. Sci. U.S.A. 91, 12373–
synapse function and plasticity. Prog. Neurobiol. 91, 313–348. 12377.
Chadman, K.K., Gong, S., Scattoni, M.L., Boltuck, S.E., Gandhy, S.U., Heintz, N., Craig, A.M., Graf, E.R., Linhoff, M.W., 2006. How to build a central synapse: clues
Crawley, J.N., 2008. Minimal aberrant behavioral phenotypes of neuroligin-3 from cell culture. Trends Neurosci. 29, 8–20.
R451C knockin mice. Autism Res. 1, 147–158. Craig, A.M., Kang, Y., 2007. Neurexin–neuroligin signaling in synapse development.
Chahrour, M., Jung, S.Y., Shaw, C., Zhou, X., Wong, S.T., Qin, J., Zoghbi, H.Y., 2008. Curr. Opin. Neurobiol. 17, 43–52.
MeCP2, a key contributor to neurological disease, activates and represses Dai, Z., Peng, H.B., 1996. Dynamics of synaptic vesicles in cultured spinal cord
transcription. Science 320, 1224–1229. neurons in relationship to synaptogenesis. Mol. Cell. Neurosci. 7, 443–452.
Chalifoux, J.R., Carter, A.G., 2010. GABAB receptors modulate NMDA receptor Dalva, M.B., McClelland, A.C., Kayser, M.S., 2007. Cell adhesion molecules: signalling
calcium signals in dendritic spines. Neuron 66, 101–113. functions at the synapse. Nat. Rev. Neurosci. 8, 206–220.
Chalifoux, J.R., Carter, A.G., 2011. GABA(B) receptor modulation of synaptic function. Danglot, L., Triller, A., Marty, S., 2006. The development of hippocampal interneur-
Curr. Opin. Neurobiol. 21, 339–344. ons in rodents. Hippocampus 16, 1032–1060.
84 M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87

Dani, V.S., Chang, Q., Maffei, A., Turrigiano, G.G., Jaenisch, R., Nelson, S.B., 2005. Ganguly, K., Schinder, A.F., Wong, S.T., Poo, M., 2001. GABA itself promotes the
Reduced cortical activity due to a shift in the balance between excitation and developmental switch of neuronal GABAergic responses from excitation to
inhibition in a mouse model of Rett syndrome. Proc. Natl. Acad. Sci. U.S.A. 102, inhibition. Cell 105, 521–532.
12560–12565. Garcia, R.A., Vasudevan, K., Buonanno, A., 2000. The neuregulin receptor ErbB-4
Dani, V.S., Nelson, S.B., 2009. Intact long-term potentiation but reduced connectivi- interacts with PDZ-containing proteins at neuronal synapses. Proc. Natl. Acad.
ty between neocortical layer 5 pyramidal neurons in a mouse model of Rett Sci. U.S.A. 97, 3596–3601.
syndrome. J. Neurosci. 29, 11263–11270. Gerrow, K., Romorini, S., Nabi, S.M., Colicos, M.A., Sala, C., El-Husseini, A., 2006. A
Davies, C.H., Starkey, S.J., Pozza, M.F., Collingridge, G.L., 1991. GABA autoreceptors preformed complex of postsynaptic proteins is involved in excitatory synapse
regulate the induction of LTP. Nature 349, 609–611. development. Neuron 49, 547–562.
Dean, C., Dresbach, T., 2006. Neuroligins and neurexins: linking cell adhesion, Ghiretti, A.E., Paradis, S., 2011. The GTPase Rem2 regulates synapse development
synapse formation and cognitive function. Trends Neurosci. 29, 21–29. and dendritic morphology. Dev. Neurobiol. 71, 374–389.
Dean, C., Scholl, F.G., Choih, J., DeMaria, S., Berger, J., Isacoff, E., Scheiffele, P., 2003. Gilman, S.R., Iossifov, I., Levy, D., Ronemus, M., Wigler, M., Vitkup, D., 2011. Rare de
Neurexin mediates the assembly of presynaptic terminals. Nat. Neurosci. 6, 708– novo variants associated with autism implicate a large functional network of
716. genes involved in formation and function of synapses. Neuron 70, 898–907.
Dowell, J.A., Johnson, J.A., Li, L., 2009. Identification of astrocyte secreted proteins Godenschwege, T.A., Hu, H., Shan-Crofts, X., Goodman, C.S., Murphey, R.K., 2002. Bi-
with a combination of shotgun proteomics and bioinformatics. J. Proteome Res. directional signaling by Semaphorin 1a during central synapse formation in
8, 4135–4143. Drosophila. Nat. Neurosci. 5, 1294–1301.
Dumitriu, D., Cossart, R., Huang, J., Yuste, R., 2007. Correlation between axonal Gottmann, K., Mittmann, T., Lessmann, V., 2009. BDNF signaling in the formation,
morphologies and synaptic input kinetics of interneurons from mouse visual maturation and plasticity of glutamatergic and GABAergic synapses. Exp. Brain
cortex. Cereb. Cortex 17, 81–91. Res. 199, 203–234.
Durand, C.M., Betancur, C., Boeckers, T.M., Bockmann, J., Chaste, P., Fauchereau, F., Gozlan, H., Ben-Ari, Y., 2003. Interneurons are the source and the targets of the first
Nygren, G., Rastam, M., Gillberg, I.C., Anckarsater, H., et al., 2007. Mutations in synapses formed in the rat developing hippocampal circuit. Cereb. Cortex 13,
the gene encoding the synaptic scaffolding protein SHANK3 are associated with 684–692.
autism spectrum disorders. Nat. Genet. 39, 25–27. Graf, E.R., Kang, Y., Hauner, A.M., Craig, A.M., 2006. Structure function and splice site
Durand, G.M., Konnerth, A., 1996. Long-term potentiation as a mechanism of analysis of the synaptogenic activity of the neurexin-1 beta LNS domain. J.
functional synapse induction in the developing hippocampus. J. Physiol. Paris Neurosci. 26, 4256–4265.
90, 313–315. Graf, E.R., Zhang, X., Jin, S.X., Linhoff, M.W., Craig, A.M., 2004. Neurexins induce
Durand, G.M., Kovalchuk, Y., Konnerth, A., 1996. Long-term potentiation and differentiation of GABA and glutamate postsynaptic specializations via neuro-
functional synapse induction in developing hippocampus. Nature 381, 71–75. ligins. Cell 119, 1013–1026.
Dutar, P., Nicoll, R.A., 1988. A physiological role for GABAB receptors in the central Gray, E.G., 1959. Axo-somatic and axo-dendritic synapses of the cerebral cortex: an
nervous system. Nature 332, 156–158. electron microscope study. J. Anat. 93, 420–433.
Dzhala, V.I., Talos, D.M., Sdrulla, D.A., Brumback, A.C., Mathews, G.C., Benke, T.A., Greenberg, M.E., Xu, B., Lu, B., Hempstead, B.L., 2009. New insights in the biology of
Delpire, E., Jensen, F.E., Staley, K.J., 2005. NKCC1 transporter facilitates seizures BDNF synthesis and release: implications in CNS function. J. Neurosci. 29,
in the developing brain. Nat. Med. 11, 1205–1213. 12764–12767.
Elmariah, S.B., Crumling, M.A., Parsons, T.D., Balice-Gordon, R.J., 2004. Postsyn- Gulyas, A.I., Hajos, N., Freund, T.F., 1996. Interneurons containing calretinin are
aptic TrkB-mediated signaling modulates excitatory and inhibitory neuro- specialized to control other interneurons in the rat hippocampus. J. Neurosci.
transmitter receptor clustering at hippocampal synapses. J. Neurosci. 24, 16, 3397–3411.
2380–2393. Gunther, U., Benson, J., Benke, D., Fritschy, J.M., Reyes, G., Knoflach, F., Crestani,
Elmariah, S.B., Oh, E.J., Hughes, E.G., Balice-Gordon, R.J., 2005. Astrocytes regulate F., Aguzzi, A., Arigoni, M., Lang, Y., et al., 1995. Benzodiazepine-insensitive
inhibitory synapse formation via Trk-mediated modulation of postsynaptic mice generated by targeted disruption of the gamma 2 subunit gene of
GABAA receptors. J. Neurosci. 25, 3638–3650. gamma-aminobutyric acid type A receptors. Proc. Natl. Acad. Sci. U.S.A. 92,
Ernfors, P., Lee, K.F., Jaenisch, R., 1994. Mice lacking brain-derived neurotrophic 7749–7753.
factor develop with sensory deficits. Nature 368, 147–150. Guy, J., Hendrich, B., Holmes, M., Martin, J.E., Bird, A., 2001. A mouse Mecp2-null
Essrich, C., Lorez, M., Benson, J.A., Fritschy, J.M., Luscher, B., 1998. Postsynaptic mutation causes neurological symptoms that mimic Rett syndrome. Nat. Genet.
clustering of major GABAA receptor subtypes requires the gamma 2 subunit and 27, 322–326.
gephyrin. Nat. Neurosci. 1, 563–571. Harms, K.J., Craig, A.M., 2005. Synapse composition and organization following
Faissner, A., Pyka, M., Geissler, M., Sobik, T., Frischknecht, R., Gundelfinger, E.D., chronic activity blockade in cultured hippocampal neurons. J. Comp. Neurol.
Seidenbecher, C., 2010. Contributions of astrocytes to synapse formation and 490, 72–84.
maturation – potential functions of the perisynaptic extracellular matrix. Brain Harvey, K., Duguid, I.C., Alldred, M.J., Beatty, S.E., Ward, H., Keep, N.H., Lingenfelter,
Res. Rev. 63, 26–38. S.E., Pearce, B.R., Lundgren, J., Owen, M.J., et al., 2004. The GDP-GTP exchange
Farrar, S.J., Whiting, P.J., Bonnert, T.P., McKernan, R.M., 1999. Stoichiometry of a factor collybistin: an essential determinant of neuronal gephyrin clustering. J.
ligand-gated ion channel determined by fluorescence energy transfer. J. Biol. Neurosci. 24, 5816–5826.
Chem. 274, 10100–10104. Hayashi, M.K., Tang, C., Verpelli, C., Narayanan, R., Stearns, M.H., Xu, R.M., Li, H., Sala,
Fazzari, P., Paternain, A.V., Valiente, M., Pla, R., Lujan, R., Lloyd, K., Lerma, J., Marin, O., C., Hayashi, Y., 2009. The postsynaptic density proteins Homer and Shank form a
Rico, B., 2010. Control of cortical GABA circuitry development by Nrg1 and polymeric network structure. Cell 137, 159–171.
ErbB4 signalling. Nature 464, 1376–1380. Haydon, P.G., Carmignoto, G., 2006. Astrocyte control of synaptic transmission and
Feng, G., Tintrup, H., Kirsch, J., Nichol, M.C., Kuhse, J., Betz, H., Sanes, J.R., 1998. Dual neurovascular coupling. Physiol. Rev. 86, 1009–1031.
requirement for gephyrin in glycine receptor clustering and molybdoenzyme Hennou, S., Khalilov, I., Diabira, D., Ben-Ari, Y., Gozlan, H., 2002. Early sequential
activity. Science 282, 1321–1324. formation of functional GABA(A) and glutamatergic synapses on CA1 inter-
Fernandez, F., Garner, C.C., 2007. Over-inhibition: a model for developmental neurons of the rat foetal hippocampus. Eur. J. Neurosci. 16, 197–208.
intellectual disability. Trends Neurosci. 30, 497–503. Hoffman, E.P., Brown Jr., R.H., Kunkel, L.M., 1987. Dystrophin: the protein product of
Fischer, F., Kneussel, M., Tintrup, H., Haverkamp, S., Rauen, T., Betz, H., Wassle, H., the Duchenne muscular dystrophy locus. Cell 51, 919–928.
2000. Reduced synaptic clustering of GABA and glycine receptors in the retina of Holmgren, C.D., Mukhtarov, M., Malkov, A.E., Popova, I.Y., Bregestovski, P., Zilberter,
the gephyrin null mutant mouse. J. Comp. Neurol. 427, 634–648. Y., 2010. Energy substrate availability as a determinant of neuronal resting
Freund, T.F., Buzsaki, G., 1996. Interneurons of the hippocampus. Hippocampus 6, potential, GABA signaling and spontaneous network activity in the neonatal
347–470. cortex in vitro. J. Neurochem. 112, 900–912.
Friedman, H.V., Bresler, T., Garner, C.C., Ziv, N.E., 2000. Assembly of new individual Hong, E.J., McCord, A.E., Greenberg, M.E., 2008. A biological function for the
excitatory synapses: time course and temporal order of synaptic molecule neuronal activity-dependent component of Bdnf transcription in the develop-
recruitment. Neuron 27, 57–69. ment of cortical inhibition. Neuron 60, 610–624.
Fritschy, J.M., Brunig, I., 2003. Formation and plasticity of GABAergic synapses: Hong, E.J., West, A.E., Greenberg, M.E., 2005. Transcriptional control of cognitive
physiological mechanisms and pathophysiological implications. Pharmacol. development. Curr. Opin. Neurobiol. 15, 21–28.
Ther. 98, 299–323. Hong, Q., Zhang, M., Pan, X.Q., Guo, M., Li, F., Tong, M.L., Chen, R.H., Guo, X.R., Chi, X.,
Fritschy, J.M., Harvey, R.J., Schwarz, G., 2008. Gephyrin: where do we stand, where 2009. Prefrontal cortex Homer expression in an animal model of attention-
do we go? Trends Neurosci. 31, 257–264. deficit/hyperactivity disorder. J. Neurol. Sci. 287, 205–211.
Fritschy, J.M., Panzanelli, P., Kralic, J.E., Vogt, K.E., Sassoe-Pognetto, M., 2006. Huang, E.J., Reichardt, L.F., 2001. Neurotrophins: roles in neuronal development and
Differential dependence of axo-dendritic and axo-somatic GABAergic synapses function. Annu. Rev. Neurosci. 24, 677–736.
on GABAA receptors containing the alpha1 subunit in Purkinje cells. J. Neurosci. Huang, Y.Z., Won, S., Ali, D.W., Wang, Q., Tanowitz, M., Du, Q.S., Pelkey, K.A.,
26, 3245–3255. Yang, D.J., Xiong, W.C., Salter, M.W., et al., 2000. Regulation of neuregulin
Frotscher, M., Zhao, S., Graber, W., Drakew, A., Studer, D., 2007. New ways of looking signaling by PSD-95 interacting with ErbB4 at CNS synapses. Neuron 26,
at synapses. Histochem. Cell Biol. 128, 91–96. 443–455.
Gahwiler, B.H., Brown, D.A., 1985. GABAB-receptor-activated K+ current in voltage- Huang, Z.J., Kirkwood, A., Pizzorusso, T., Porciatti, V., Morales, B., Bear, M.F., Maffei,
clamped CA3 pyramidal cells in hippocampal cultures. Proc. Natl. Acad. Sci. L., Tonegawa, S., 1999. BDNF regulates the maturation of inhibition and the
U.S.A. 82, 1558–1562. critical period of plasticity in mouse visual cortex. Cell 98, 739–755.
Galvan, A., Charara, A., Pare, J.F., Levey, A.I., Smith, Y., 2004. Differential subcellular Huang, Z.J., Scheiffele, P., 2008. GABA and neuroligin signaling: linking synaptic
and subsynaptic distribution of GABA(A) and GABA(B) receptors in the monkey activity and adhesion in inhibitory synapse development. Curr. Opin. Neurobiol.
subthalamic nucleus. Neuroscience 127, 709–721. 18, 77–83.
M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87 85

Hubner, C.A., Stein, V., Hermans-Borgmeyer, I., Meyer, T., Ballanyi, K., Jentsch, T.J., Kruger, R.P., Aurandt, J., Guan, K.L., 2005. Semaphorins command cells to move. Nat.
2001. Disruption of KCC2 reveals an essential role of K-Cl cotransport already in Rev. Mol. Cell Biol. 6, 789–800.
early synaptic inhibition. Neuron 30, 515–524. Kulik, A., Vida, I., Lujan, R., Haas, C.A., Lopez-Bendito, G., Shigemoto, R., Frotscher, M.,
Hughes, E.G., Elmariah, S.B., Balice-Gordon, R.J., 2010. Astrocyte secreted proteins 2003. Subcellular localization of metabotropic GABA(B) receptor subunits
selectively increase hippocampal GABAergic axon length, branching, and GABA(B1a/b) and GABA(B2) in the rat hippocampus. J. Neurosci. 23, 11026–
synaptogenesis. Mol. Cell. Neurosci. 43, 136–145. 11035.
Hutsler, J.J., Zhang, H., 2010. Increased dendritic spine densities on cortical projec- Lardi-Studler, B., Fritschy, J.M., 2007. Matching of pre- and postsynaptic specializa-
tion neurons in autism spectrum disorders. Brain Res. 1309, 83–94. tions during synaptogenesis. Neuroscientist 13, 115–126.
Inoue, M., Hara, M., Zeng, X.T., Hirose, T., Ohnishi, S., Yasukura, T., Uriu, T., Omori, K., Lein, E.S., Hawrylycz, M.J., Ao, N., Ayres, M., Bensinger, A., Bernard, A., Boe, A.F.,
Minato, A., Inagaki, C., 1991. An ATP-driven Cl pump regulates Cl concentra- Boguski, M.S., Brockway, K.S., Byrnes, E.J., et al., 2007. Genome-wide atlas of
tions in rat hippocampal neurons. Neurosci. Lett. 134, 75–78. gene expression in the adult mouse brain. Nature 445, 168–176.
Isomoto, S., Kaibara, M., Sakurai-Yamashita, Y., Nagayama, Y., Uezono, Y., Yano, K., Levi, S., Logan, S.M., Tovar, K.R., Craig, A.M., 2004. Gephyrin is critical for glycine
Taniyama, K., 1998. Cloning and tissue distribution of novel splice variants of receptor clustering but not for the formation of functional GABAergic synapses
the rat GABAB receptor. Biochem. Biophys. Res. Commun. 253, 10–15. in hippocampal neurons. J. Neurosci. 24, 207–217.
Jacob, T.C., Bogdanov, Y.D., Magnus, C., Saliba, R.S., Kittler, J.T., Haydon, P.G., Moss, Levy, D., Ronemus, M., Yamrom, B., Lee, Y.H., Leotta, A., Kendall, J., Marks, S.,
S.J., 2005. Gephyrin regulates the cell surface dynamics of synaptic GABAA Lakshmi, B., Pai, D., Ye, K., et al., 2011. Rare de novo and transmitted copy-
receptors. J. Neurosci. 25, 10469–10478. number variation in autistic spectrum disorders. Neuron 70, 886–897.
Jacob, T.C., Moss, S.J., Jurd, R., 2008. GABA(A) receptor trafficking and its role in the Li, B., Woo, R.S., Mei, L., Malinow, R., 2007. The neuregulin-1 receptor erbB4 controls
dynamic modulation of neuronal inhibition. Nat. Rev. Neurosci. 9, 331–343. glutamatergic synapse maturation and plasticity. Neuron 54, 583–597.
Jacob, T.C., Wan, Q., Vithlani, M., Saliba, R.S., Succol, F., Pangalos, M.N., Moss, S.J., Lin, X., Ogiya, M., Takahara, M., Yamaguchi, W., Furuyama, T., Tanaka, H., Tohyama,
2009. GABA(A) receptor membrane trafficking regulates spine maturity. Proc. M., Inagaki, S., 2007. Sema4D-plexin-B1 implicated in regulation of dendritic
Natl. Acad. Sci. U.S.A. 106, 12500–12505. spine density through RhoA/ROCK pathway. Neurosci. Lett. 428, 1–6.
Jamain, S., Quach, H., Betancur, C., Rastam, M., Colineaux, C., Gillberg, I.C., Soder- Lin, Y., Bloodgood, B.L., Hauser, J.L., Lapan, A.D., Koon, A.C., Kim, T.K., Hu, L.S., Malik,
strom, H., Giros, B., Leboyer, M., Gillberg, C., et al., 2003. Mutations of the X- A.N., Greenberg, M.E., 2008. Activity-dependent regulation of inhibitory syn-
linked genes encoding neuroligins NLGN3 and NLGN4 are associated with apse development by Npas4. Nature 455, 1198–1204.
autism. Nat. Genet. 34, 27–29. Liu, Q.Y., Schaffner, A.E., Chang, Y.H., Vaszil, K., Barker, J.L., 1997. Astrocytes regulate
Jeon, H., Lee, S., Lee, W.H., Suk, K., 2010. Analysis of glial secretome: the long amino acid receptor current densities in embryonic rat hippocampal neurons. J.
pentraxin PTX3 modulates phagocytic activity of microglia. J. Neuroimmunol. Neurobiol. 33, 848–864.
229, 63–72. Liu, Q.Y., Schaffner, A.E., Li, Y.X., Dunlap, V., Barker, J.L., 1996. Upregulation of GABAA
Jones, K.R., Farinas, I., Backus, C., Reichardt, L.F., 1994. Targeted disruption of the current by astrocytes in cultured embryonic rat hippocampal neurons. J.
BDNF gene perturbs brain and sensory neuron development but not motor Neurosci. 16, 2912–2923.
neuron development. Cell 76, 989–999. Lohmann, C., Bonhoeffer, T., 2008. A role for local calcium signaling in rapid synaptic
Kalscheuer, V.M., Musante, L., Fang, C., Hoffmann, K., Fuchs, C., Carta, E., Deas, E., partner selection by dendritic filopodia. Neuron 59, 253–260.
Venkateswarlu, K., Menzel, C., Ullmann, R., et al., 2009. A balanced chromo- Lord, C., Cook, E.H., Leventhal, B.L., Amaral, D.G., 2000. Autism spectrum disorders.
somal translocation disrupting ARHGEF9 is associated with epilepsy, anxiety, Neuron 28, 355–363.
aggression, and mental retardation. Hum. Mutat. 30, 61–68. Luscher, B., Fuchs, T., Kilpatrick, C.L., 2011. GABAA receptor trafficking-mediated
Kandel, E.R., Spencer, W.A., Brinley, J.R., 1961. Electrophysiology of hippocampal plasticity of inhibitory synapses. Neuron 70, 385–409.
neurons. I. Sequential invasion and synaptic organization. J. Neurophysiol. 24, Magdaleno, S., Jensen, P., Brumwell, C.L., Seal, A., Lehman, K., Asbury, A., Cheung, T.,
225–242. Cornelius, T., Batten, D.M., Eden, C., et al., 2006. BGEM: an in situ hybridization
Kantamneni, S., Correa, S.A., Hodgkinson, G.K., Meyer, G., Vinh, N.N., Henley, J.M., database of gene expression in the embryonic and adult mouse nervous system.
Nishimune, A., 2007. GISP: a novel brain-specific protein that promotes surface PLoS Biol. 4, e86.
expression and function of GABA(B) receptors. J. Neurochem. 100, 1003–1017. Marshall, C.R., Noor, A., Vincent, J.B., Lionel, A.C., Feuk, L., Skaug, J., Shago, M.,
Kasai, H., Matsuzaki, M., Noguchi, J., Yasumatsu, N., Nakahara, H., 2003. Structure- Moessner, R., Pinto, D., Ren, Y., et al., 2008. Structural variation of chromosomes
stability-function relationships of dendritic spines. Trends Neurosci. 26, 360– in autism spectrum disorder. Am. J. Hum. Genet. 82, 477–488.
368. Marty, S., Wehrle, R., Sotelo, C., 2000. Neuronal activity and brain-derived neuro-
Katz, L.C., Shatz, C.J., 1996. Synaptic activity and the construction of cortical circuits. trophic factor regulate the density of inhibitory synapses in organotypic slice
Science 274, 1133–1138. cultures of postnatal hippocampus. J. Neurosci. 20, 8087–8095.
Kaupmann, K., Huggel, K., Heid, J., Flor, P.J., Bischoff, S., Mickel, S.J., McMaster, G., McAllister, A.K., 2007. Dynamic aspects of CNS synapse formation. Annu. Rev.
Angst, C., Bittiger, H., Froestl, W., et al., 1997. Expression cloning of GABA(B) Neurosci. 30, 425–450.
receptors uncovers similarity to metabotropic glutamate receptors. Nature 386, McGraw, C.M., Samaco, R.C., Zoghbi, H.Y., 2011. Adult Neural Function Requires
239–246. MeCP2. Science.
Khakhalin, A.S., 2011. Questioning the depolarizing effects of GABA during early McNair, K., Spike, R., Guilding, C., Prendergast, G.C., Stone, T.W., Cobb, S.R., Morris,
brain development. J. Neurophysiol. (May, 18) doi:10.1152/jn.00293.2011. B.J., 2010. A role for RhoB in synaptic plasticity and the regulation of neuronal
Khazipov, R., Khalilov, I., Tyzio, R., Morozova, E., Ben-Ari, Y., Holmes, G.L., 2004. morphology. J. Neurosci. 30, 3508–3517.
Developmental changes in GABAergic actions and seizure susceptibility in the Mei, L., Xiong, W.C., 2008. Neuregulin 1 in neural development, synaptic plasticity
rat hippocampus. Eur. J. Neurosci. 19, 590–600. and schizophrenia. Nat. Rev. Neurosci. 9, 437–452.
Kim, E., Sheng, M., 2004. PDZ domain proteins of synapses. Nat. Rev. Neurosci. 5, Melin, M., Carlsson, B., Anckarsater, H., Rastam, M., Betancur, C., Isaksson, A.,
771–781. Gillberg, C., Dahl, N., 2006. Constitutional downregulation of SEMA5A expres-
Kins, S., Betz, H., Kirsch, J., 2000. Collybistin, a newly identified brain-specific GEF, sion in autism. Neuropsychobiology 54, 64–69.
induces submembrane clustering of gephyrin. Nat. Neurosci. 3, 22–29. Meyer, G., Kirsch, J., Betz, H., Langosch, D., 1995. Identification of a gephyrin binding
Kishi, N., Macklis, J.D., 2004. MECP2 is progressively expressed in post-migratory motif on the glycine receptor beta subunit. Neuron 15, 563–572.
neurons and is involved in neuronal maturation rather than cell fate decisions. Micheva, K.D., Beaulieu, C., 1995. An anatomical substrate for experience-depen-
Mol. Cell. Neurosci. 27, 306–321. dent plasticity of the rat barrel field cortex. Proc. Natl. Acad. Sci. U.S.A. 92,
Kneussel, M., Brandstatter, J.H., Gasnier, B., Feng, G., Sanes, J.R., Betz, H., 2001. 11834–11838.
Gephyrin-independent clustering of postsynaptic GABA(A) receptor subtypes. Micheva, K.D., Busse, B., Weiler, N.C., O’Rourke, N., Smith, S.J., 2010. Single-synapse
Mol. Cell. Neurosci. 17, 973–982. analysis of a diverse synapse population: proteomic imaging methods and
Kneussel, M., Brandstatter, J.H., Laube, B., Stahl, S., Muller, U., Betz, H., 1999. Loss of markers. Neuron 68, 639–653.
postsynaptic GABA(A) receptor clustering in gephyrin-deficient mice. J. Neu- Miles, R., Wong, R.K., 1984. Unitary inhibitory synaptic potentials in the guinea-pig
rosci. 19, 9289–9297. hippocampus in vitro. J. Physiol. 356, 97–113.
Knuesel, I., Mastrocola, M., Zuellig, R.A., Bornhauser, B., Schaub, M.C., Fritschy, J.M., Miles, R., Wong, R.K., 1987. Inhibitory control of local excitatory circuits in the
1999. Short communication: altered synaptic clustering of GABAA receptors in guinea-pig hippocampus. J. Physiol. 388, 611–629.
mice lacking dystrophin (mdx mice). Eur. J. Neurosci. 11, 4457–4462. Miranda, R.C., Sohrabji, F., Toran-Allerand, C.D., 1993. Neuronal colocalization of
Kohara, K., Yasuda, H., Huang, Y., Adachi, N., Sohya, K., Tsumoto, T., 2007. A local mRNAs for neurotrophins and their receptors in the developing central nervous
reduction in cortical GABAergic synapses after a loss of endogenous brain- system suggests a potential for autocrine interactions. Proc. Natl. Acad. Sci.
derived neurotrophic factor, as revealed by single-cell gene knock-out method. U.S.A. 90, 6439–6443.
J. Neurosci. 27, 7234–7244. Missler, M., Zhang, W., Rohlmann, A., Kattenstroth, G., Hammer, R.E., Gottmann, K.,
Kokaia, Z., Bengzon, J., Metsis, M., Kokaia, M., Persson, H., Lindvall, O., 1993. Sudhof, T.C., 2003. Alpha-neurexins couple Ca2+ channels to synaptic vesicle
Coexpression of neurotrophins and their receptors in neurons of the central exocytosis. Nature 423, 939–948.
nervous system. Proc. Natl. Acad. Sci. U.S.A. 90, 6711–6715. Moessner, R., Marshall, C.R., Sutcliffe, J.S., Skaug, J., Pinto, D., Vincent, J., Zwaigen-
Koller, H., Siebler, M., Schmalenbach, C., Muller, H.W., 1990. GABA and glutamate baum, L., Fernandez, B., Roberts, W., Szatmari, P., et al., 2007. Contribution of
receptor development of cultured neurons from rat hippocampus, septal region, SHANK3 mutations to autism spectrum disorder. Am. J. Hum. Genet. 81, 1289–
and neocortex. Synapse 5, 59–64. 1297.
Kralic, J.E., Sidler, C., Parpan, F., Homanics, G.E., Morrow, A.L., Fritschy, J.M., 2006. Morita, A., Yamashita, N., Sasaki, Y., Uchida, Y., Nakajima, O., Nakamura, F., Yagi, T.,
Compensatory alteration of inhibitory synaptic circuits in cerebellum and Taniguchi, M., Usui, H., Katoh-Semba, R., et al., 2006. Regulation of dendritic
thalamus of gamma-aminobutyric acid type A receptor alpha1 subunit knock- branching and spine maturation by semaphorin3A-Fyn signaling. J. Neurosci.
out mice. J. Comp. Neurol. 495, 408–421. 26, 2971–2980.
86 M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87

Moulder, K.L., Jiang, X., Taylor, A.A., Shin, W., Gillis, K.D., Mennerick, S., 2007. Vesicle both excitatory and inhibitory synapses of rat brain. J. Comp. Neurol. 408, 437–
pool heterogeneity at hippocampal glutamate and GABA synapses. J. Neurosci. 448.
27, 9846–9854. Rico, B., Xu, B., Reichardt, L.F., 2002. TrkB receptor signaling is required for
Murphey, R.K., Froggett, S.J., Caruccio, P., Shan-Crofts, X., Kitamoto, T., Godensch- establishment of GABAergic synapses in the cerebellum. Nat. Neurosci. 5,
wege, T.A., 2003. Targeted expression of shibire ts and semaphorin 1a reveals 225–233.
critical periods for synapse formation in the giant fiber of Drosophila. Develop- Riddle, D.R., McAllister, A.K., Lo, D.C., Katz, L.C., 1996. Neurotrophins in cortical
ment 130, 3671–3682. development. Cold Spring Harb. Symp. Quant. Biol. 61, 85–93.
Nagler, K., Mauch, D.H., Pfrieger, F.W., 2001. Glia-derived signals induce synapse Rivera, C., Voipio, J., Payne, J.A., Ruusuvuori, E., Lahtinen, H., Lamsa, K., Pirvola, U.,
formation in neurones of the rat central nervous system. J. Physiol. 533, 665– Saarma, M., Kaila, K., 1999. The K+/Cl co-transporter KCC2 renders GABA
679. hyperpolarizing during neuronal maturation. Nature 397, 251–255.
Nicoll, R.A., Malenka, R.C., Kauer, J.A., 1990. Functional comparison of neurotrans- Robinson, T.E., Kolb, B., 2004. Structural plasticity associated with exposure to drugs
mitter receptor subtypes in mammalian central nervous system. Physiol. Rev. of abuse. Neuropharmacology 47 (Suppl. 1), 33–46.
70, 513–565. Rostaing, P., Real, E., Siksou, L., Lechaire, J.P., Boudier, T., Boeckers, T.M., Gertler, F.,
Noh, J., Seal, R.P., Garver, J.A., Edwards, R.H., Kandler, K., 2010. Glutamate co-release Gundelfinger, E.D., Triller, A., Marty, S., 2006. Analysis of synaptic ultrastructure
at GABA/glycinergic synapses is crucial for the refinement of an inhibitory map. without fixative using high-pressure freezing and tomography. Eur. J. Neurosci.
Nat. Neurosci. 13, 232–238. 24, 3463–3474.
Noutel, J., Hong, Y.K., Leu, B., Kang, E., Chen, C., 2011. Experience-dependent Rubenstein, J.L., Merzenich, M.M., 2003. Model of autism: increased ratio of excita-
retinogeniculate synapse remodeling is abnormal in MeCP2-deficient mice. tion/inhibition in key neural systems. Genes Brain Behav. 2, 255–267.
Neuron 70, 35–42. Russo, S.J., Mazei-Robison, M.S., Ables, J.L., Nestler, E.J., 2009. Neurotrophic factors
Nusser, Z., Sieghart, W., Benke, D., Fritschy, J.M., Somogyi, P., 1996. Differential and structural plasticity in addiction. Neuropharmacology 56, 73–82.
synaptic localization of two major gamma-aminobutyric acid type A receptor Sabo, S.L., Gomes, R.A., McAllister, A.K., 2006. Formation of presynaptic terminals at
alpha subunits on hippocampal pyramidal cells. Proc. Natl. Acad. Sci. U.S.A. 93, predefined sites along axons. J. Neurosci. 26, 10813–10825.
11939–11944. Saiepour, L., Fuchs, C., Patrizi, A., Sassoe-Pognetto, M., Harvey, R.J., Harvey, K., 2010.
Nusser, Z., Sieghart, W., Somogyi, P., 1998. Segregation of different GABAA receptors Complex role of collybistin and gephyrin in GABAA receptor clustering. J. Biol.
to synaptic and extrasynaptic membranes of cerebellar granule cells. J. Neu- Chem. 285, 29623–29631.
rosci. 18, 1693–1703. Sala, C., Piech, V., Wilson, N.R., Passafaro, M., Liu, G., Sheng, M., 2001. Regulation of
O’Connor, T.P., Cockburn, K., Wang, W., Tapia, L., Currie, E., Bamji, S.X., 2009. dendritic spine morphology and synaptic function by Shank and Homer. Neuron
Semaphorin 5B mediates synapse elimination in hippocampal neurons. Neural. 31, 115–130.
Dev. 4, 18. Sanders, S.J., Ercan-Sencicek, A.G., Hus, V., Luo, R., Murtha, M.T., Moreno-De-
O’Sullivan, G.A., Hofer, W., Betz, H., 2009. Inhibitory postsynaptic membrane Luca, D., Chu, S.H., Moreau, M.P., Gupta, A.R., Thomson, S.A., et al., 2011.
specializations are formed in gephyrin-deficient mice. Neurosci. Lett. 458, Multiple recurrent de novo CNVs, including duplications of the 7q11.23
106–110. williams syndrome region, are strongly associated with autism. Neuron 70,
Papadopoulos, T., Eulenburg, V., Reddy-Alla, S., Mansuy, I.M., Li, Y., Betz, H., 2008. 863–885.
Collybistin is required for both the formation and maintenance of GABAergic Sanmarti-Vila, L., tom Dieck, S., Richter, K., Altrock, W., Zhang, L., Volknandt, W.,
postsynapses in the hippocampus. Mol. Cell. Neurosci. 39, 161–169. Zimmermann, H., Garner, C.C., Gundelfinger, E.D., Dresbach, T., 2000. Membrane
Papadopoulos, T., Korte, M., Eulenburg, V., Kubota, H., Retiounskaia, M., Harvey, R.J., association of presynaptic cytomatrix protein bassoon. Biochem. Biophys. Res.
Harvey, K., O’Sullivan, G.A., Laube, B., Hulsmann, S., et al., 2007. Impaired Commun. 275, 43–46.
GABAergic transmission and altered hippocampal synaptic plasticity in colly- Sassoe-Pognetto, M., Kirsch, J., Grunert, U., Greferath, U., Fritschy, J.M., Mohler, H.,
bistin-deficient mice. EMBO J. 26, 3888–3899. Betz, H., Wassle, H., 1995. Colocalization of gephyrin and GABAA-receptor
Paradis, S., Harrar, D.B., Lin, Y., Koon, A.C., Hauser, J.L., Griffith, E.C., Zhu, L., Brass, L.F., subunits in the rat retina. J. Comp. Neurol. 357, 1–14.
Chen, C., Greenberg, M.E., 2007. An RNAi-based approach identifies molecules Scanziani, M., 2000. GABA spillover activates postsynaptic GABA(B) receptors to
required for glutamatergic and GABAergic synapse development. Neuron 53, control rhythmic hippocampal activity. Neuron 25, 673–681.
217–232. Scheiffele, P., 2003. Cell-cell signaling during synapse formation in the CNS. Annu.
Parra, P., Gulyas, A.I., Miles, R., 1998. How many subtypes of inhibitory cells in the Rev. Neurosci. 26, 485–508.
hippocampus? Neuron 20, 983–993. Scheiffele, P., Fan, J., Choih, J., Fetter, R., Serafini, T., 2000. Neuroligin expressed in
Pecho-Vrieseling, E., Sigrist, M., Yoshida, Y., Jessell, T.M., Arber, S., 2009. Specificity nonneuronal cells triggers presynaptic development in contacting axons. Cell
of sensory-motor connections encoded by Sema3e-Plxnd1 recognition. Nature 101, 657–669.
459, 842–846. Schumann, C.M., Hamstra, J., Goodlin-Jones, B.L., Lotspeich, L.J., Kwon, H., Buono-
Percy, A.K., Lane, J.B., 2005. Rett syndrome: model of neurodevelopmental disor- core, M.H., Lammers, C.R., Reiss, A.L., Amaral, D.G., 2004. The amygdala is
ders. J. Child Neurol. 20, 718–721. enlarged in children but not adolescents with autism; the hippocampus is
Peters, A., Harriman, K.M., 1990. Different kinds of axon terminals forming sym- enlarged at all ages. J. Neurosci. 24, 6392–6401.
metric synapses with the cell bodies and initial axon segments of layer II/III Schwarz, D.A., Barry, G., Eliasof, S.D., Petroski, R.E., Conlon, P.J., Maki, R.A., 2000.
pyramidal cells. I. Morphometric analysis. J. Neurocytol. 19, 154–174. Characterization of gamma-aminobutyric acid receptor GABAB(1e), a
Peters, A., Palay, S.L., 1996. The morphology of synapses. J. Neurocytol. 25, 687–700. GABAB(1) splice variant encoding a truncated receptor. J. Biol. Chem. 275,
Pfaff, T., Malitschek, B., Kaupmann, K., Prezeau, L., Pin, J.P., Bettler, B., Karschin, A., 32174–32181.
1999. Alternative splicing generates a novel isoform of the rat metabotropic Sekiguchi, M., Zushida, K., Yoshida, M., Maekawa, M., Kamichi, S., Sahara, Y., Yuasa,
GABA(B)R1 receptor. Eur. J. Neurosci. 11, 2874–2882. S., Takeda, S., Wada, K., 2009. A deficit of brain dystrophin impairs specific
Pfeiffer, F., Graham, D., Betz, H., 1982. Purification by affinity chromatography of the amygdala GABAergic transmission and enhances defensive behaviour in mice.
glycine receptor of rat spinal cord. J. Biol. Chem. 257, 9389–9393. Brain 132, 124–135.
Pfrieger, F.W., Barres, B.A., 1997. Synaptic efficacy enhanced by glial cells in vitro. Shapira, M., Zhai, R.G., Dresbach, T., Bresler, T., Torres, V.I., Gundelfinger, E.D., Ziv,
Science 277, 1684–1687. N.E., Garner, C.C., 2003. Unitary assembly of presynaptic active zones from
Pilgram, G.S., Potikanond, S., Baines, R.A., Fradkin, L.G., Noordermeer, J.N., 2010. The Piccolo-Bassoon transport vesicles. Neuron 38, 237–252.
roles of the dystrophin-associated glycoprotein complex at the synapse. Mol. Siksou, L., Rostaing, P., Lechaire, J.P., Boudier, T., Ohtsuka, T., Fejtova, A., Kao, H.T.,
Neurobiol. 41, 1–21. Greengard, P., Gundelfinger, E.D., Triller, A., et al., 2007. Three-dimensional
Pinto, D., Pagnamenta, A.T., Klei, L., Anney, R., Merico, D., Regan, R., Conroy, J., architecture of presynaptic terminal cytomatrix. J. Neurosci. 27, 6868–6877.
Magalhaes, T.R., Correia, C., Abrahams, B.S., et al., 2010. Functional impact of Siksou, L., Triller, A., Marty, S., 2009. An emerging view of presynaptic structure from
global rare copy number variation in autism spectrum disorders. Nature 466, electron microscopic studies. J. Neurochem. 108, 1336–1342.
368–372. Slezak, M., Pfrieger, F.W., 2003. New roles for astrocytes: regulation of CNS synap-
Poulopoulos, A., Aramuni, G., Meyer, G., Soykan, T., Hoon, M., Papadopoulos, T., togenesis. Trends Neurosci. 26, 531–535.
Zhang, M., Paarmann, I., Fuchs, C., Harvey, K., et al., 2009. Neuroligin 2 drives Soghomonian, J.J., Martin, D.L., 1998. Two isoforms of glutamate decarboxylase:
postsynaptic assembly at perisomatic inhibitory synapses through gephyrin why? Trends Pharmacol. Sci. 19, 500–505.
and collybistin. Neuron 63, 628–642. Somogyi, J., Baude, A., Omori, Y., Shimizu, H., El Mestikawy, S., Fukaya, M., Shige-
Prange, O., Murphy, T.H., 2001. Modular transport of postsynaptic density-95 moto, R., Watanabe, M., Somogyi, P., 2004. GABAergic basket cells expressing
clusters and association with stable spine precursors during early development cholecystokinin contain vesicular glutamate transporter type 3 (VGLUT3) in
of cortical neurons. J. Neurosci. 21, 9325–9333. their synaptic terminals in hippocampus and isocortex of the rat. Eur. J.
Pritchett, D.B., Sontheimer, H., Shivers, B.D., Ymer, S., Kettenmann, H., Schofield, P.R., Neurosci. 19, 552–569.
Seeburg, P.H., 1989. Importance of a novel GABAA receptor subunit for benzo- Song, J.Y., Ichtchenko, K., Sudhof, T.C., Brose, N., 1999. Neuroligin 1 is a postsynaptic
diazepine pharmacology. Nature 338, 582–585. cell-adhesion molecule of excitatory synapses. Proc. Natl. Acad. Sci. U.S.A. 96,
Rao, A., Cha, E.M., Craig, A.M., 2000. Mismatched appositions of presynaptic and 1100–1105.
postsynaptic components in isolated hippocampal neurons. J. Neurosci. 20, Studer, R., von Boehmer, L., Haenggi, T., Schweizer, C., Benke, D., Rudolph, U.,
8344–8353. Fritschy, J.M., 2006. Alteration of GABAergic synapses and gephyrin clusters
Rheims, S., Holmgren, C.D., Chazal, G., Mulder, J., Harkany, T., Zilberter, T., Zilberter, in the thalamic reticular nucleus of GABAA receptor alpha3 subunit-null mice.
Y., 2009. GABA action in immature neocortical neurons directly depends on the Eur. J. Neurosci. 24, 1307–1315.
availability of ketone bodies. J. Neurochem. 110, 1330–1338. Studler, B., Fritschy, J.M., Brunig, I., 2002. GABAergic and glutamatergic terminals
Richter, K., Langnaese, K., Kreutz, M.R., Olias, G., Zhai, R., Scheich, H., Garner, C.C., differentially influence the organization of GABAergic synapses in rat cerebellar
Gundelfinger, E.D., 1999. Presynaptic cytomatrix protein bassoon is localized at granule cells in vitro. Neuroscience 114, 123–133.
M.S. Kuzirian, S. Paradis / Progress in Neurobiology 95 (2011) 68–87 87

Swiercz, J.M., Kuner, R., Behrens, J., Offermanns, S., 2002. Plexin-B1 directly interacts Varaqueaux, R., Sigler, A., Rhee, J.S., Brose, N., Enk, C., Reim, K., Rosenmund, C., 2002.
with PDZ-RhoGEF/LARG to regulate RhoA and growth cone morphology. Neu- Total arrest of spontaneous and evoked synaptic transmission by normal
ron 35, 51–63. synaptogenesis in the absence of Munc13-mediated vesicle priming. Proc. Natl.
Szatmari, P., Paterson, A.D., Zwaigenbaum, L., Roberts, W., Brian, J., Liu, X.Q., Vincent, Acad. U.S.A. 99, 9037–9042.
J.B., Skaug, J.L., Thompson, A.P., Senman, L., et al., 2007. Mapping autism risk loci Vashlishan, A.B., Madison, J.M., Dybbs, M., Bai, J., Sieburth, D., Ch’ng, Q., Tavazoie, M.,
using genetic linkage and chromosomal rearrangements. Nat. Genet. 39, 319– Kaplan, J.M., 2008. An RNAi screen identifies genes that regulate GABA synap-
328. ses. Neuron 58, 346–361.
Tabuchi, K., Blundell, J., Etherton, M.R., Hammer, R.E., Liu, X., Powell, C.M., Sudhof, Verhage, M., Maia, A.S., Plomp, J.J., Brussaard, A.B., Heeroma, J.H., Vermeer, H.,
T.C., 2007. A neuroligin-3 mutation implicated in autism increases inhibitory Toonen, R.F., Hammer, R.E., van den Berg, T.K., Missler, M., et al., 2000. Synaptic
synaptic transmission in mice. Science 318, 71–76. assembly of the brain in the absence of neurotransmitter secretion. Science 287,
Takebayashi, M., Kagaya, A., Hayashi, T., Motohashi, N., Yamawaki, S., 1996. gamma- 864–869.
Aminobutyric acid increases intracellular Ca2+ concentration in cultured corti- Vigot, R., Barbieri, S., Brauner-Osborne, H., Turecek, R., Shigemoto, R., Zhang, Y.P.,
cal neurons: role of Cl transport. Eur. J. Pharmacol. 297, 137–143. Lujan, R., Jacobson, L.H., Biermann, B., Fritschy, J.M., et al., 2006. Differential
Tamagnone, L., Artigiani, S., Chen, H., He, Z., Ming, G.I., Song, H., Chedotal, A., compartmentalization and distinct functions of GABAB receptor variants. Neu-
Winberg, M.L., Goodman, C.S., Poo, M., et al., 1999. Plexins are a large family ron 50, 589–601.
of receptors for transmembrane, secreted, and GPI-anchored semaphorins in Vullhorst, D., Neddens, J., Karavanova, I., Tricoire, L., Petralia, R.S., McBain, C.J.,
vertebrates. Cell 99, 71–80. Buonanno, A., 2009. Selective expression of ErbB4 in interneurons, but not
Tashiro, A., Minden, A., Yuste, R., 2000. Regulation of dendritic spine morphology by pyramidal cells, of the rodent hippocampus. J. Neurosci. 29, 12255–12264.
the rho family of small GTPases: antagonistic roles of Rac and Rho. Cereb. Cortex Walton, M.K., Schaffner, A.E., Barker, J.L., 1993. Sodium channels, GABAA receptors,
10, 927–938. and glutamate receptors develop sequentially on embryonic rat spinal cord
Tate, P., Skarnes, W., Bird, A., 1996. The methyl-CpG binding protein MeCP2 is cells. J. Neurosci. 13, 2068–2084.
essential for embryonic development in the mouse. Nat. Genet. 12, 205–208. Wang, C., Shimizu-Okabe, C., Watanabe, K., Okabe, A., Matsuzaki, H., Ogawa, T.,
Terauchi, A., Johnson-Venkatesh, E.M., Toth, A.B., Javed, D., Sutton, M.A., Umemori, Mori, N., Fukuda, A., Sato, K., 2002. Developmental changes in KCC1, KCC2, and
H., 2010. Distinct FGFs promote differentiation of excitatory and inhibitory NKCC1 mRNA expressions in the rat brain. Brain Res. Develop. Brain Res. 139,
synapses. Nature 465, 783–787. 59–66.
Thompson, S.M., Gahwiler, B.H., 1992. Comparison of the actions of baclofen at pre- Wang, D.D., Kriegstein, A.R., 2008. GABA regulates excitatory synapse formation in
and postsynaptic receptors in the rat hippocampus in vitro. J. Physiol. 451, 329– the neocortex via NMDA receptor activation. J. Neurosci. 28, 5547–5558.
345. Washbourne, P., Bennett, J.E., McAllister, A.K., 2002. Rapid recruitment of NMDA
Tran, T.S., Rubio, M.E., Clem, R.L., Johnson, D., Case, L., Tessier-Lavigne, M., Huganir, receptor transport packets to nascent synapses. Nat. Neurosci. 5, 751–759.
R.L., Ginty, D.D., Kolodkin, A.L., 2009. Secreted semaphorins control spine Washbourne, P., Liu, X.B., Jones, E.G., McAllister, A.K., 2004. Cycling of NMDA
distribution and morphogenesis in the postnatal CNS. Nature 462, 1065–1069. receptors during trafficking in neurons before synapse formation. J. Neurosci.
Tretter, V., Jacob, T.C., Mukherjee, J., Fritschy, J.M., Pangalos, M.N., Moss, S.J., 2008. 24, 8253–8264.
The clustering of GABA(A) receptor subtypes at inhibitory synapses is facilitated Weiss, L.A., Arking, D.E., Daly, M.J., Chakravarti, A., 2009. A genome-wide linkage
via the direct binding of receptor alpha 2 subunits to gephyrin. J. Neurosci. 28, and association scan reveals novel loci for autism. Nature 461, 802–808.
1356–1365. Wen, L., Lu, Y.S., Zhu, X.H., Li, X.M., Woo, R.S., Chen, Y.J., Yin, D.M., Lai, C., Terry Jr.,
Tretter, V., Moss, S.J., 2008. GABA(A) receptor dynamics and constructing GABAergic A.V., Vazdarjanova, A., et al., 2010. Neuregulin 1 regulates pyramidal neuron
synapses. Front. Mol. Neurosci. 1, 7. activity via ErbB4 in parvalbumin-positive interneurons. Proc. Natl. Acad. Sci.
Tsacopoulos, M., Magistretti, P.J., 1996. Metabolic coupling between glia and U.S.A. 107, 1211–1216.
neurons. J. Neurosci. 16, 877–885. Wierenga, C.J., Becker, N., Bonhoeffer, T., 2008. GABAergic synapses are formed
Tsay, D., Yuste, R., 2004. On the electrical function of dendritic spines. Trends without the involvement of dendritic protrusions. Nat. Neurosci. 11, 1044–
Neurosci. 27, 77–83. 1052.
Turner, N., Grose, R., 2010. Fibroblast growth factor signalling: from development to Woo, R.S., Li, X.M., Tao, Y., Carpenter-Hyland, E., Huang, Y.Z., Weber, J., Neiswender,
cancer. Nat. Rev. Cancer 10, 116–129. H., Dong, X.P., Wu, J., Gassmann, M., et al., 2007. Neuregulin-1 enhances
Twelvetrees, A.E., Yuen, E.Y., Arancibia-Carcamo, I.L., MacAskill, A.F., Rostaing, P., depolarization-induced GABA release. Neuron 54, 599–610.
Lumb, M.J., Humbert, S., Triller, A., Saudou, F., Yan, Z., et al., 2010. Delivery of Wu, G., Malinow, R., Cline, H.T., 1996. Maturation of a central glutamatergic
GABAARs to synapses is mediated by HAP1-KIF5 and disrupted by mutant synapse. Science 274, 972–976.
huntingtin. Neuron 65, 53–65. Xu, J., Wojcik, W.J., 1986. Gamma aminobutyric acid B receptor-mediated inhibition
Tyzio, R., Allene, C., Nardou, R., Picardo, M.A., Yamamoto, S., Sivakumaran, S., Caiati, of adenylate cyclase in cultured cerebellar granule cells: blockade by islet-
M.D., Rheims, S., Minlebaev, M., Milh, M., et al., 2011. Depolarizing actions of activating protein. J. Pharmacol. Exp. Ther. 239, 568–573.
GABA in immature neurons depend neither on ketone bodies nor on pyruvate. J. Yazdani, U., Terman, J.R., 2006. The semaphorins. Genome Biol. 7, 211.
Neurosci. 31, 34–45. Yoshii, A., Constantine-Paton, M., 2010. Postsynaptic BDNF-TrkB signaling in syn-
Tyzio, R., Minlebaev, M., Rheims, S., Ivanov, A., Jorquera, I., Holmes, G.L., Zilberter, Y., apse maturation, plasticity, and disease. Dev. Neurobiol. 70, 304–322.
Ben-Ari, Y., Khazipov, R., 2008. Postnatal changes in somatic gamma-amino- Zakharenko, S., Chang, S., O’Donoghue, M., Popov, S.V., 1999. Neurotransmitter
butyric acid signalling in the rat hippocampus. Eur. J. Neurosci. 27, 2515–2528. secretion along growing nerve processes: comparison with synaptic vesicle
Tyzio, R., Represa, A., Jorquera, I., Ben-Ari, Y., Gozlan, H., Aniksztejn, L., 1999. The exocytosis. J. Cell Biol. 144, 507–518.
establishment of GABAergic and glutamatergic synapses on CA1 pyramidal Zander, J.F., Munster-Wandowski, A., Brunk, I., Pahner, I., Gomez-Lira, G., Heine-
neurons is sequential and correlates with the development of the apical mann, U., Gutierrez, R., Laube, G., Ahnert-Hilger, G., 2010. Synaptic and vesicular
dendrite. J. Neurosci. 19, 10372–10382. coexistence of VGLUT and VGAT in selected excitatory and inhibitory synapses.
Ullian, E.M., Christopherson, K.S., Barres, B.A., 2004. Role for glia in synaptogenesis. J. Neurosci. 30, 7634–7645.
Glia 47, 209–216. Zhai, R.G., Vardinon-Friedman, H., Cases-Langhoff, C., Becker, B., Gundelfinger, E.D.,
Ullian, E.M., Sapperstein, S.K., Christopherson, K.S., Barres, B.A., 2001. Control of Ziv, N.E., Garner, C.C., 2001. Assembling the presynaptic active zone: a charac-
synapse number by glia. Science 291, 657–661. terization of an active one precursor vesicle. Neuron 29, 131–143.
Umemori, H., Linhoff, M.W., Ornitz, D.M., Sanes, J.R., 2004. FGF22 and its close Zhang, C., Atasoy, D., Arac, D., Yang, X., Fucillo, M.V., Robison, A.J., Ko, J., Brunger, A.T.,
relatives are presynaptic organizing molecules in the mammalian brain. Cell Sudhof, T.C., 2010a. Neurexins physically and functionally interact with
118, 257–270. GABA(A) receptors. Neuron 66, 403–416.
van Spronsen, M., Hoogenraad, C.C., 2010. Synapse pathology in psychiatric and Zhang, Z.W., Zak, J.D., Liu, H., 2010b. MeCP2 is required for normal development of
neurologic disease. Curr. Neurol. Neurosci. Rep. 10, 207–214. GABAergic circuits in the thalamus. J. Neurophysiol. 103, 2470–2481.
Varoqueaux, F., Aramuni, G., Rawson, R.L., Mohrmann, R., Missler, M., Gottmann, K., Zilberter, Y., Zilberter, T., Bregestovski, P., 2010. Neuronal activity in vitro and the in
Zhang, W., Sudhof, T.C., Brose, N., 2006. Neuroligins determine synapse matu- vivo reality: the role of energy homeostasis. Trends Pharmacol. Sci. 31, 394–
ration and function. Neuron 51, 741–754. 401.
Varoqueaux, F., Jamain, S., Brose, N., 2004. Neuroligin 2 is exclusively localized to Zoghbi, H.Y., 2003. Postnatal neurodevelopmental disorders: meeting at the
inhibitory synapses. Eur. J. Cell Biol. 83, 449–456. synapse? Science 302, 826–830.

Das könnte Ihnen auch gefallen