U.S. flag

An official website of the United States government

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

Noebels JL, Avoli M, Rogawski MA, et al., editors. Jasper's Basic Mechanisms of the Epilepsies. 5th edition. New York: Oxford University Press; 2024. doi: 10.1093/med/9780197549469.003.0047

Cover of Jasper's Basic Mechanisms of the Epilepsies

Jasper's Basic Mechanisms of the Epilepsies. 5th edition.

Show details

Chapter 47Transcriptional Regulation of Cortical Interneuron Development

, , , and .

Abstract

GABAergic interneurons (INs) constitute 20%–30% of the pallial (neocortical and hippocampal) neurons. They are the main source of cortical and hippocampal synaptic inhibitory signals. Distinct IN subtypes inhibit different cellular and subcellular components of pallial circuits. INs are generated from different subdomains of the embryonic basal ganglia (ganglionic eminences, GEs). Transcription factors (TFs) through regulatory elements (REs) that they bind are integral in programming gene expression in a temporally and spatially diverse manner. This then programs cell fate-commitment, differentiation, migration, and maturation to generate different IN subtypes. Of note, the prevailing molecular and genetic knowledge about IN development has been established by decades of studies in rodent and is hypothesized to reflect many of the processes common to mammals. Research on human and primate brain development will test this hypothesis and also offers the opportunity to identify species-specific variations. In this chapter, we focus on discoveries derived from mouse research. This chapter reviews the histology of the GEs, fate-mapping tools, TFs that control regional patterning of the GEs, TFs that regulate IN maturation, and REs that are involved in IN development.

Introduction

GABAergic interneurons (INs) are the primary source of inhibition in the cortex and hippocampus and are critical in tuning the balance of excitation and inhibition (E/I). Cortical and hippocampal interneurons (CINs and HINs) are derived prenatally from subpallial progenitor domains. Currently, there are roughly 20 subtypes of CINs and 23 subtypes of HINs that are defined based on their morphology, molecular markers, synaptic connectivity, and electrophysiological properties (Klausberger et al., 2003; Huang, Di Cristo, and Ango, 2007; Kepecs and Fishell, 2014; Kessaris et al., 2014; Pelkey et al., 2017; Harris et al., 2018).

The research into IN fate determination is an active field, with evidence that the mechanisms driving their identities act in both progenitor cells and maturing neurons. Cascades of transcription factors (TFs) expressed in specific temporal and spatial windows are critical in establishing IN fates, migration, morphogenesis, synaptogenesis, and function. Disruptions to these transcriptional networks lead to IN pathologies in mice that are relevant to human neurological disorders. In this regard, IN pathologies are hypothesized to underlie some phenotypes of disorders such as autism spectrum disorder (ASD), epilepsy, and schizophrenia (Rubenstein and Merzenich, 2003; Chao et al., 2010; Yizhar et al., 2011; Han et al., 2012; Mi et al., 2018; Sohal and Rubenstein, 2019). This chapter will discuss the role of TFs and other proteins or regulatory elements that are crucial for IN specification and maturation.

Subpallial Progenitor Domain Subdivisions

The telencephalic subpallium is synonymous with embryonic basal ganglia (also termed ganglionic eminences [GEs]). These are divided into five regions with distinct dorso-ventral and rostro-caudal positions: medial ganglionic eminence (MGE), caudal ganglionic eminence (CGE), lateral ganglionic eminence (LGE), septum, and the pre-optic area (POA). The GEs generate distinct nuclei and cell types (Puelles et al., 2016; Rubenstein & Campbell, 2020).

Progenitor cells in these domains give rise to projection neurons and INs, which are largely GABAergic, although smaller numbers of cholinergic and dopaminergic neurons are also generated. Over development these progenitor domains become progressively gliogenic; by birth in rodents they become largely non-neurogenic ependymal zones. However, neurogenesis within these domains may continue into neonatal stages in primate and human (Paredes et al., 2016). In this review, we will mainly focus on the findings from rodent studies.

There is evidence that the MGE, CGE, LGE, POA, and septal progenitor domains are further subdivided into spatial regions that generate distinct cell types. These domains have distinct gene regulatory element (enhancer) activation patterns and express distinct TF combinations and signaling molecules (e.g., sonic hedgehog [SHH]) (Flames et al., 2007; Silberberg et al., 2016; Hu et al., 2017a; Su-Feher et al., 2021).

We expect that further studies will reinforce this concept and provide further granularity into the spatio-temporal mechanisms that contribute to cell fate specification from the subpallial progenitor subdomains.

The ventriculo-pial (mediolateral) axis of the central nervous system (CNS) is the axis of differentiation. Adjacent to the lateral ventricle lies the highly proliferative neuroepithelium, which consists of a pseudostratified layer of subpallial stem cells generally called the ventricular zone (VZ). Around embryonic day (E)9.5, these transition from neuroepithelial cells to radial glia stem cells (apical progenitors); the latter generate intermediate progenitors (basal progenitors), neurons, and glia. Superficial to the VZ are the intermediate progenitors that are in the subventricular zone (SVZ). There is evidence for two SVZs: SVZ1 lies adjacent to the VZ, and SVZ2 is adjacent to SVZ1 (Petryniak et al., 2007). SVZ1 cells appear to be more immature and SVZ2 cells are highly intermixed with immature postmitotic cells. Superficial to the SVZs is the mantle zone (MZ) which contains migrating neurons, glia, and post-migratory cells forming the nuclei of the basal ganglia, such as the striatum (St), globus pallidus (GP), and the central and medial amygdala.

Below, we will briefly introduce each subdivision of the mouse embryonic basal ganglia and their main functions (Fig. 47–1).

Figure 47–1.. Illustration of the subpallial anatomy, representative transcription factors (TFs), and neuronal subtype derivatives.

Figure 47–1.

Illustration of the subpallial anatomy, representative transcription factors (TFs), and neuronal subtype derivatives. Anatomical mouse developing basal ganglia subregions, for the LGE, MGE, Septum, CGE, and POA, are illustrated and coded by colors. LGE (more...)

Medial Ganglionic Eminence

The MGE generates pallidal projection neurons, glia, and INs; the latter tangentially migrate throughout the telencephalon, particularly into the striatum, cortex, hippocampus, and amygdala (Marín et al., 2000; Kessaris et al., 2006; Xu et al., 2008; Fragkouli et al., 2009; Flandin et al., 2010; Nóbrega-Pereira et al., 2010; Bupesh et al., 2011; Silberberg et al., 2016).

The MGE is the major source of CINs (~70%) and generates roughly equal groups that are somatostatin+ (SST+) or parvalbumin+ (PV+) CINs (Wonders and Anderson, 2006; Miyoshi et al., 2010; Batista-Brito et al., 2020). MGE-derived CINs migrate tangentially to the neocortex where they integrate into cortical lamina (in an inside-out manner based on when they are generated) (Guo and Anton, 2014). Immature CINs form synaptic connections with excitatory projection neurons and other CINs.

TF expression and enhancer activity provide evidence that the MGE VZ has six subdomains (MGE0-5) along the dorsoventral axis (MGE0 most dorsal and MGE5 most ventral) that are biased in producing different neuronal subtypes (Flames et al., 2007; Silberberg et al., 2016; Hu et al., 2017a). MGE0 and MGE1 express the TFs Nr2f2 and Nkx6.2, but not Nkx2.1; its cellular output is uncertain but can produce progenitors that are displaced ventrally where they intercalate with other progenitors (Hu and Rubenstein, unpublished). MGE1 and MGE2 express Nkx2.1, Nr2f2, and Nkx6.2. Nkx6.2 loss of function and fate-mapping show that these subregions generate mostly SST CINs (Fogarty et al., 2007; Sousa et al., 2009). MGE3–5 encompass the ventral MGE and border the POA. These regions express Otx2 and Nkx2-1, but not Nr2f2 and Nkx6.2. Fate mapping using an enhancer expressed in this domain shows a bias for PV CINs (Silberberg et al., 2016; Hu et al., 2017a). Shh-Cre (active in MGE5) fate mapping suggests this ventral MGE subregion generates striatal cholinergic and PV+ INs, few PV+ CINs, and GP PV+ projection neurons (Flandin et al., 2010).

Biased production of PV and SST CINs is also seen across the MGE’s rostral-caudal axis. WNT signaling and Nr2f2 expression provide evidence for two rostral-caudal subregions. Transplantation of rostral MGE (low WNT signaling and absent of Nr2f2) resulted in mostly PV CINs, whereas the caudal domain (high WNT signaling and Nr2f2+) generates mostly SST CINs. Findings were supported by loss and gain of Nr2f2 function and gain of RYK function, a noncanonical WNT receptor (Hu et al., 2017b; McKenzie et al., 2018).

Caudal Ganglionic Eminence

The CGE progenitor zone, caudal to the MGE and LGE, lacks Nkx2.1 and has molecular properties similar to the LGE. The CGE produces CINs with distinct laminar positions, electrical properties, and molecular/cellular features (Nery et al., 2002; Miyoshi et al., 2010). There is no anatomical structure that clearly defines where the CGE begins and ends. A small zone of Nkx2.1+ MGE extends caudally adjacent to the CGE (Hu et al., 2017a). CINs derived from the CGE are delayed in their generation by about 3 days compared to the MGE (Wonders and Anderson, 2006). However, unlike MGE-derived CINs that laminate in an inside-out fashion, most CGE-derived CINs occupy upper cortical lamina regardless of birthdate (Guo and Anton, 2014).

CGE-lineage CINs comprise two broad groups based on mostly nonoverlapping molecular markers: (1) vasoactive intestinal peptide (VIP+) and (2) NDNF+; REELIN+; SST- (Miyoshi et al., 2010; Schuman et al., 2019). The latter group consists of neurogliaform CINs (Taniguchi et al., 2011; Silberberg et al., 2016; Abs et al., 2018). Notably, each of these groups expresses the CGE-enriched TFs, Sp8 and Prox1. The Serotonin-receptor 3a (5HT3aR) is also a CGE-lineage marker (Miyoshi et al., 2010; Lee et al., 2010; Rudy et al., 2011; Ma et al., 2012; Rubin and Kessaris, 2013; Miyoshi et al., 2015; Touzot et al., 2016; Wei et al., 2019).

There is no known TF that specifies the CGE, as Nkx2.1 does for the MGE. Perhaps the LGE/CGE regions are patterned by a default mechanism, while Nkx2.1 is required to drive MGE identity. This is supported by data showing that in the absence of Nkx2.1, LGE molecular identity is present in the space where the putative MGE would be (Sussel et al., 1999). However, some notable TFs have a role in CGE cell fate and IN development, including Gsx1, Gsx2, Nr2f1, Nr2f2, and Prox1. We will discuss these TFs below in the regional fate specification section (Wang et al., 2009, 2013; Pei et al., 2011; Chapman et al., 2018).

Lateral Ganglionic Eminence

The lateral ganglionic eminence (LGE) is dorsal to the MGE and rostral to the CGE. Unlike the sulcus that separates the MGE and LGE, the LGE and CGE do not have a clear border, and they share many molecular properties (Flames et al., 2007; Long et al., 2009a,b). The LGE produces at least three major groups of neurons: striatal projection neurons, olfactory bulb (OB) INs (Deacon et al., 1994; Olsson et al., 1997; Olsson, Björklund, and Campbell, 1998; Wichterle et al., 2001; Ma et al., 2012), and perhaps a small subset of CINs (Torigoe et al., 2016).

Evidence suggests the existence of different LGE progenitors in the SVZ is responsible for LGE-lineage neuronal diversity. For example, the Dlx+; Isl1+ progenitors in the SVZ predominantly generate striatal projection neurons, while the Dlx+; Isl1–; ER81+ progenitors that mainly reside in the dorsal LGE give rise to OB INs. As OB INs are born, they take the rostral migratory stream (RMS) to populate the granular/periglomerular layers in the OB (Stenman et al., 2003). The most dorsal LGE has strong expression of Sp8, which with Sp9, regulates the development of OB and CGE INs (Li et al., 2018; Wei et al., 2019).

Pre-Optic Area

The embryonic POA is interposed between the MGE and the hypothalamus (Puelles et al., 2000; Gelman et al., 2009). While both the MGE and the POA express Nkx2.1, they have molecular differences (Flames et al., 2007). The MGE expresses Lhx6, Lhx8, and Olig2, whereas the POA expresses Dbx1 and Nkx5.1.

The POA generates ~10% of CINs, most of which are NPY+/SST–, similar to CGE INs (Gelman et al., 2009, 2011). The POA also generates neurogliaform INs, neurons that migrate to GP and amygdala, and oligodendrocytes (Hirata et al., 2009; Flandin et al., 2010; Torigoe et al., 2016; Niquille et al., 2018). Nkx5.1 (Hmx3) demarcates a POA progenitor subpopulation that gives rise to neurogliaform INs (Niquille et al., 2018).

Septum

The septum is a rostral extension of the LGE and the MGE, but with distinct properties governed in part by FGF signaling and expression of the Zic TFs (Inoue et al., 2007; Hoch et al., 2015b). Patterning of the lateral septum is regulated by the Dlx and Gsx genes (Long et al., 2007, 2009a; Wang et al., 2009; Qin et al., 2017), whereas patterning of the medial septum is controlled by Lhx6, Lhx8, and Nkx2-1 (Zhao et al., 2003, 2008; Flandin et al., 2011; Magno et al., 2017).

The septum and adjacent regions of the MGE generate GABAergic and cholinergic neurons that populate the septum diagonal band, striatum, GP, and some pre-optic nuclei based on Fgf8-Cre and Fgf17-Cre fate mapping (Hoch et al., 2015b). Shh-Cre fate mapping showed that specific regions of the MGE generate neurons of the medial septum and diagonal band (Flandin et al., 2010). Combining Nkx2.1 and Zic4 in vivo genetic labeling, the lineages of septal progenitors can be traced temporally (Magno et al., 2017). Many of these neurons form projections to the hippocampus and are crucial for modulating synaptic plasticity, rhythm generation, and learning/memory (Colom, 2006; Drever, Riedel, and Platt, 2011; Magno et al., 2017).

Regional Specification of IN-Generating Progenitor Zones

MGE Specification and LGE/CGE Identity Repression

Nkx2.1 and Otx2 play crucial roles in MGE specification through their VZ expression. Nkx2.1 expression starts at ~E9.5 and is responsible for the induction of Lhx6 and Lhx8 in the MGE SVZ to specify IN and GP progenitors (Sussel et al., 1999; Flandin et al., 2011; Sandberg et al., 2016). As CINs are specified, they lose the expression of Nkx2.1 and Lhx8 but maintain Lhx6, which drives CIN differentiation, migration, and the subsequent expression of SST and PV (Liodis et al., 2007; Zhao et al., 2008; Vogt et al., 2014). In addition to promoting IN production, Nkx2.1 represses CGE and LGE fate. Constitutive loss of Nkx2.1 leads to loss of MGE CIN production and ectopic generation of striatal medium spiny neurons (Sussel et al., 1999); later removal (E12.5) of Nkx2.1 leads to CIN subclass switch from MGE- to CGE- lineage properties (Butt et al., 2005, 2008).

Although broadly expressed in the MGE VZ, the Otx2 homeodomain TF preferentially guides the fate specification of the ventral MGE (vMGE), a subregion that mainly produces projection neurons that migrate to the globus pallidus (GP) (Flandin et al., 2010). Conditional deletion of Otx2 in the Nkx2.1-cre defining MGE leads to diminished GP neuron production. Otx2 also interacts with the Fgf-signaling pathway and promotes the expression of Sall3, Tal2, and Tll2 that delineate vMGE from the POA and other parts of MGE (Hoch et al., 2015b).

Roles of Gsx1 and Gsx2 in Establishing LGE and CGE

Gsx1 and Gsx2 are homeodomain TFs that specify LGE and CGE identity. Gsx2 is expressed in the VZ and SVZ, whereas Gsx1 is expressed in the SVZ. Loss of Gsx2 and combined deletion of Gsx1/Gsx2 leads to progressively severe phenotypes in the LGE and CGE (Toresson and Campbell, 2001; Yun et al., 2003). No clear phenotypes have yet been seen in the Gsx1 mutant. In addition, Gsx1/Gsx2 may not be critical for MGE development (Xu et al., 2010).

Gsx2 is particularly important for dorsal LGE specification, a region that produces many OB INs and parts of the striatum. Gsx2 also represses expression of cortical TFs Ngn2, Tbr2, and Pax6 in the LGE (Fode et al., 2000; Toresson, Potter, and Campbell, 2000; Toresson and Campbell, 2001; Yun et al., 2003). Gsx1 functions with Gsx2 for LGE specification, the production of striatal neurons, and guidance of corticofugal/thalamocortical axons (Toresson and Campbell, 2001; Yun et al., 2003). Loss of Gsx1/Gsx2 leads to reductions in the expression of Dlx genes in the LGE SVZ, TFs required in maintaining the SVZ progenitor activity and the differentiation of striatal neurons (Anderson et al., 1997a; Toresson and Campbell, 2001; Yun, Fischman et al., 2002; Yun et al., 2003; Pei et al., 2011).

Gsx2 also facilitates CGE patterning and neurogenesis (Corbin et al., 2000, 2003; Xu et al., 2010; Wang et al., 2013). Constitutive Gsx2 mutants have a hypoplastic CGE (Xu et al., 2010). Gsx2 gain-of-function in the MGE induces a CGE-like identity to produce bipolar Calretinin+/SST– INs (Butt et al., 2008; Xu et al., 2010; Wang et al., 2013).

Ventral and Rostral Patterning Centers Together Induce Nkx2.1 and the MGE through SHH and FGF8 Signaling

Sonic hedgehog (SHH) is a secreted morphogen that promotes ventral regional identity throughout the developing CNS. Through its expression in the rostral hypothalamus and preoptic area, it promotes ventral telencephalic fate by inducing Nkx2.1, Dlx2, and Gsx2 expression (Xu et al., 2005, 2010). Reduced SHH signaling leads to a hypoplastic ventral telencephalon and ectopic expression of dorsal telencephalic markers, such as Pax6 and Emx2 (Ohkubo et al., 2002; Fuccillo et al., 2004). SHH signaling functions in part by transcriptionally activating Gli1 and Gli2. On the other hand, Gli3 can serve as a transcriptional repressor and antagonize Shh function to promote telencephalic dorsal identity (Litingtung and Chiang, 2000; Motoyama and Aoto, 2007; Zaki, Martynoga, and Price, 2008; Ruiz i Altaba, 1998; Jiang and Hui, 2008; Hui and Angers, 2011). During early MGE development, Lhx6 and Lhx8 have compensatory roles to induce the expression of Shh in the MGE mantle zone via regulation of a Shh enhancer (Flandin et al., 2011). In turn, the induction of Shh provides a feed-forward loop to maintain the expression of Lhx6, Lhx8, and Nkx2.1, which facilitate MGE-derived pallial IN and subpallial neuron generation (Xu et al., 2005; Flandin et al., 2011). Depletion of Shh in the Dlx1/2-Cre lineage leads to increased apoptosis, which results in a similar reduction of SST and PV CINs in adult cortices (Flandin et al., 2011).

Despite the antagonism between Shh and Gli3, Shh/Gli3 double deletion renders no phenotype in dorsoventral telencephalic patterning. This suggested that other morphogens may be involved in telencephalic patterning (Rallu et al., 2002; Fuccillo et al., 2004). Fibroblastic growth factors (FGFs) are expressed in the rostral midline of the telencephalon, also known as the rostral patterning center (RPC). One such FGF is FGF8, which is implicated in the induction of telencephalic identify by promoting Foxg1 (BF1) expression (Shimamura and Rubenstein, 1997; Storm et al., 2006). Fgf8 and Otx2 reciprocally repress each other in the RPC and the midbrain/ hindbrain boundary that instructs proper telencephalic morphogenesis (Crossley et al., 2001; Hoch et al., 2015a). Furthermore, Fgf8 is required for the induction of Shh in the ventral telencephalon (Storm et al., 2006). It is proposed that positive regulation between FGF and SHH signaling, and the convergence of the two pathways, is required for induction of Nkx2.1 (Gutin et al., 2006; Storm et al., 2006; Hoch, Rubenstein, and Pleasure, 2009). FGF function, downstream of Shh and Gli3, is required for ventral progenitor generation and differentiation (Fuccillo et al., 2004; Gutin et al., 2006; Storm et al., 2006). Although the rostral patterning center expresses several Fgf genes, FGF signaling in the ventral telencephalon appears to be dominated by Fgf8 expression and its interaction with FGFR1/FGFR2 (Meyers et al., 1998; Gutin et al., 2006; Storm et al., 2006). In addition to promoting ventral progenitor fate, FGF signaling is also required to maintain the ventral progenitor pool by promoting proliferation and preventing apoptosis (Storm et al., 2006).

IN Fate Mapping

Transcriptomic analyses (Porteus et al., 1992; Long et al., 2009a,b) and mutational analysis identified TFs and other genes that are crucial for subpallial subdomain fate specification. However, the neurogenesis in the GE is far more complicated, as each of the germinal domains can at different developmental stages produce distinct types of cells that populate unique brain regions.

To investigate this complexity, various molecular and genetic tools are used. Nucleotide labeling during DNA replication (EdU/BrdU) is used to determine when neuron progenitors had last divided (birthdate). In vivo transplantation assays can assess intrinsic cell fates of progenitors in temporal and spatial manners. Finally, the use of Cre or Flip recombinases and recombinase-dependent reporters is a powerful means to follow the descendants of progenitors that have regionally and temporally restricted expression of genes. This can be coupled with further temporal regulation by fusing part of the estrogen protein (ER) to the recombinase to make it only functional in the presence of an estrogen-like ligand (e.g., tamoxifen) (Taniguchi et al., 2011; Silberberg et al., 2016). The application of recombinases (sometimes multiple in combination), and recombinase-ER fusions, allows cell-type and temporal specificity to permanently label developing lineages to enable the anatomical, molecular, and physiological characterization of their descendants (Nagy et al., 2000; Kessaris et al., 2006; Xu et al., 2009). Fate mapping using transgenic Cre lines can have complications (Luo et al., 2020). Genetic background and germline recombination can strongly influence the results from many loci. For a given locus, Cre recombinase expression can vary across development, as can the degree of tamoxifen-induced recombination. In this section, we will discuss several approaches.

Identification of MGE-Derived Descendants

As noted above, Nkx2.1 expression marks MGE progenitors and the GP, but not CINs. A transgenic line carrying Cre-recombinase driven by Nkx2.1 regulatory elements was generated to determine the descendants of Nkx2.1-expressing cells (Xu et al., 2008). Fate mapping revealed that Nkx2-1-Cre activity labeled 70%–80% of CINs and HINs, with PV- and SST-expressing IN subgroups being evenly targeted (Xu et al., 2008). Nkx2.1-Cre marks other neurons produced by the MGE, such as striatal GABAergic (SST+ and PV+), cholinergic INs, and GP projection neurons (Xu et al., 2008; Flandin et al., 2010). Unlike pallial INs, many of these neurons express Nkx2.1 and Lhx8 (Marín et al., 2001; Zhao et al., 2008; Flandin et al., 2011).

Nkx2.1 facilitates MGE-derived CIN fate determination through promoting the expression of Lhx6 (Du et al., 2008; Sandberg et al., 2016). A similar Cre-recombinase transgenic line was generated for Lhx6. Lhx6-Cre is also an excellent tool for studying MGE-lineage CINs/HINs (Fogarty et al., 2007). Almost 100% of the PV and SST CINs/HINS are labeled by the Lhx6-Cre, consistent with the Nkx2.1-Cre;Nkx6.2-Cre combined fate-mapping result (Fogarty et al., 2007; Batista-Brito et al., 2009). In addition, Lhx6-Cre demonstrated high (~100%) efficiency in labeling SST/CR-expressing Martinotti CINs and NPY- expressing CA1 HINs (Fogarty et al., 2007).

Olig2 expression is restricted primarily to GE progenitors and later oligodendrocytes. A tamoxifen-regulated Cre (Cre-ER) was inserted in the Olig2 locus and allowed a temporal specificity to fate-mapping early-born and late-born INs, followed by morphological and electrophysiological analyses (Kimmel et al., 2000; Miyoshi et al., 2007, Takebayashi et al., 2000; Nery et al., 2001; Lu et al., 2002). Consistent with the Nkx2.1-Cre finding, SST-expressing IN production was enriched at earlier ages (E9.5–E12.5) (Xu et al., 2008). Finer IN subtyping was feasible when combining molecular marker, morphology, laminar location, and intrinsic firing patterns. The diversity of IN subtypes in the Olig2-iCre lineage correlated with their birthdates. Like the nucleotide-birthdating analyses, they found that MGE-derived CINs incorporated into the circuit in an inside-out manner (Valcanis and Tan, 2003; Pla et al., 2006; Rymar and Sadikot, 2007; Xu et al., 2008; Lim et al., 2018).

Identification of CGE-Derived Descendants

To determine birthdates, subtypes, and physiological properties of the CGE lineage, several Cre lines have been used. Mash1BAC-CreER, although widely active in the GEs, can be used to follow CGE-derived INs by taking advantage of temporal differences in IN production (Battiste et al., 2007; Miyoshi et al., 2010). Tamoxifen-induced recombination suggested that CGE neurogenesis begins ~E12.5, ~3 days after the MGE. ~75% of CGE-derived INs occupied upper cortical lamina (layer I–III) (Miyoshi et al., 2010), consistent with the proposal of Rymar and Sadikot (2007) that some INs do not undergo inside-out lamination. Taking into account the morphology and physiology diversity, at least nine subtypes of CGE-derived CINs were documented (Miyoshi et al., 2007; G. Miyoshi et al., 2010), including VIP+ and REELIN+ (SST–) CGE-derived INs. At least half of the REELIN CINs were late-spiking cells, while the majority of the VIP CINs were fast-adapting cells.

Silberberg et al. (2016) generated stable transgenic mouse lines expressing CreERT2 and GFP driven by two different enhancer elements (enhancer-CreERT2-IRES-GFP) (Silberberg et al., 2016). These enhancers, 953 and 1060, map close to Sp8 and Nr2f1, respectively. The 953 CreERT2 and 1060 CreERT2 enhancer lines exhibited specific activity in the CGE MZ. Tamoxifen induction at E10.5, E12.5, or E14.5 revealed CGE-derived (SP8+) amygdala neurons labeling in both lines (Silberberg et al., 2016). CGE-derived CINs were labeled by both lines when tamoxifen was administered post E14.5. ~80% of the 1060 CreERT2 fate-mapped CINs were REELIN+ and/ or SP8+ but not VIP+, likely due to preferential labeling of neurogliaform INs (Miyoshi et al., 2010; Silberberg et al., 2016). Approximately 30% of the 953 CreERT2 fate-mapped CINs were SP8+, suitable for experiments that require sparse CGE CIN targeting.

POA-Derived INs

Although the majority of CINs and HINs are derived from the MGE and CGE, the diversity and complexity of IN subtypes and properties are constantly being discovered, suggesting that there might be more regions in the brain that contribute to diverse IN populations. POA and MGE share high levels of Nkx2.1 expression (Sussel et al., 1999; Nery et al., 2002; Flames et al., 2007; Xu et al., 2008). When a Cre-dependent GFP reporter was injected in utero into the POA of Nkx2.1-cre animals, some of the POA-originated GFP+ cells tangentially migrated into the neocortex, yet lacked Lhx6 expression (Gelman et al., 2009). Further analyses, using Nkx5.1-Cre showed that the POA generated INs for the neocortex, piriform cortex, and the hippocampus (Fogarty et al., 2005; Gelman et al., 2009). These POA-originated INs are GABAergic and express NPY but not traditional IN markers, such as PV, Calretinin, SST, VIP, and nNOS; their morphology and electrophysiological properties are also distinct from MGE/CGE INs (Fogarty et al., 2005; Fogarty et al., 2007; Gelman et al., 2009).

Pallial Progenitors Contribute to Olfactory Bulb IN Diversity

The dorsal LGE is the major prenatal source of OB INs (Wichterle et al., 1999; Toresson and Campbell, 2001; Stenman et al., 2003; Yun et al., 2003). However, the pallium also generates OB INs, and this continues postnatally (Wichterle et al., 2001; Kohwi et al., 2007; Merkle, Mirzadeh, and Alvarez-Buylla, 2007; Kriegstein and Alvarez-Buylla, 2009; Obernier and Alvarez-Buylla, 2019; Zhang et al., 2020). Emx1-Cre lineage tracing at P0 and adulthood identified a unique subset of SVZ progenitors that are LGE-like (Dlx2/Gsx2-expressing) and can produce calretinin+ OB INs (Gorski et al., 2002; Kohwi et al., 2007; Zhang et al., 2020). It is proposed that Dlx2/Gsx2/Sp8 expression in the SVZ intermediate progenitors (mostly Emx1+) promotes OB IN fate specification and represses the default oligogenic potential (Gorski et al., 2002; Chapman et al., 2012, 2018; Li et al., 2018; Zhang et al., 2020).

Shh Fate Mapping

Shh expression in the MGE is enriched in the ventral MGE and dorsal POA (Flames et al., 2007). Shh-Cre fate mapping showed that ventral MGE generated (1) very few CINs; these were PV+; (2) neurons that primarily express PV and migrate to GP, septal/preoptic complexes, and a subset of amygdaloid nuclei; and (3) INs that were predominantly PV+ that migrate to the striatum (Flandin et al., 2010). In the GP, Shh-Cre preferentially captures the NKX2-1+/NPAS1+ subpopulation (~70%–80%), rather than the NKX2-1–/NPAS1+ cells (~35%) (Flandin et al., 2010). When Nkx2.1 is depleted in the Shh domain using Shh-Cre, expression of Er81, Npas1, and Lhx8 were greatly reduced in the GP, further supporting the roles of Shh and Nkx2.1 in specifying GP neuron progenitors (Sussel et al., 1999; Flandin et al., 2010).

Fgf8 and Fgf17 Fate Mapping

FGFs are produced by the rostral patterning center (RPC) to guide regional patterning and progenitor fate commitment. RPC lineages were assessed using tamoxifen-inducible Cre under the control of either Fgf8 or Fgf17 (Hoch et al., 2015b). Induction of Cre recombination at E7.5/E8.5/E9.5 and assessment at E18.5 to P40 fate mapping revealed that Fgf8-CreER and Fgf17-CreER lineages give rise to distinct groups of neurons that populate different parts of the brain. In the OB, sparse Fgf17+ neurons are observed, while more neurons belong to the Fgf8 lineage. The Fgf8+ OB neurons can be GABAergic INs (periglomerular and granule cells) and glutamatergic projection neurons (mitral cells). The Fgf8-lineage contribution to OB neurons tapers around E10–E11.5. Projection neurons in the rostro-dorsal cortex also preferentially come from the Fgf8 lineage. The basal ganglia and CINs are also the destination for cells in the Fgf8- and Fgf17 lineages. The RPC largely generates cells of the septum. A surprising result is that both Fgf8-CreER and Fgf17-CreER generate subpallial cholinergic projection neurons and INs. In the striatum, Fgf8 lineages label medium spiny neurons (CTIP2+) and INs (PV+/SST+; ChAT); in the GP, the Fgf8 lineage co-localizes with different subtype markers (Nkx2.1, Ctip2, and Npas1).

Progenitor Zones/Stem Cell Biology in the Ganglionic Eminences

The Dlx TFs were identified in the early 1990s, and some of the initial Dlx studies focused on the expression of the Dlx genes during craniofacial and forebrain development (Porteus et al., 1991, 1992; Price et al., 1991; Robinson et al., 1991; Depew et al., 2002). In the brachial arches, differential expression of Dlx1/2/5/6 controls regional fate specification (Depew et al., 2002; Panganiban and Rubenstein, 2002; Jeong et al., 2008, 2012). Dlx1/2 mutants had skeletal defects specifically in the upper jaw (Qiu et al., 1997; Depew et al., 2005), whereas deletion of Dlx5 resulted in lower jaw defects (Depew et al., 1999); in Dlx5/Dlx6 mutants the lower jaw transformed into an upper jaw (Depew et al., 2002).

Dlx genes were the first described TFs whose expression is largely restricted to forebrain GABAergic neurons (Panganiban and Rubenstein, 2002; Stühmer et al., 2002a; Stühmer et al., 2002)b. The Dlx TFs control GE cell fate specification and differentiation through their regulation of enhancer activity and epigenetic state (Anderson et al., 1997a, 1997b; Long et al., 2007, 2009a, 2009b; Lindtner et al., 2019). In the ganglionic eminences, Dlx TFs are expressed in a sequential manner (Dlx2Dlx1Dlx5Dlx6) starting from E9.5. Their enriched and overlapping expression in the SVZ roughly separates GE progenitor zones into VZ, early progenitors, SVZ1 and SVZ2, intermediate progenitors, and MZ, immature neurons (Eisenstat et al., 1999; Yun et al., 2002; Petryniak et al., 2007). In this section, we will focus on the roles of Dlx and other key TFs on GE progenitor cell biology.

GE Ventricular Zone

The GE ventricular zone (VZ) consists of proliferating stem cells that have the potential to generate glia and several neuronal subtypes. The stemness and the proliferative capacity of the progenitors are promoted by Notch signaling. Mash1 (Ascl1) enhances Notch activity by driving the expression of Dll1, Dll3, and Hes5 (Casarosa et al.,1999). The Hes TF genes are critical in specifying and maintaining the stem cell state (Kobayashi and Kageyama, 2014). Loss of Mash1 results in reduced Notch signaling and premature expression of Dlx2, Dlx5, and Gad67 in the VZ (Yun et al., 2002).

As VZ progenitors mature and migrate into SVZ, Dlx1 and Dlx2 expression is initiated, which represses Hes1/5, Gsx1/2 (Gsh), and Mash1 to facilitate differentiation. In line with this, Dlx1/2 constitutive mutants exhibit increased expression of Dll1, Gsx1/2, Hes5, Mash1, and other VZ markers in the SVZ, and show prolonged proliferation and impeded differentiation (Anderson et al., 1997a, 1997b; Yun, Fischman et al., 2002; Long et al., 2009a, 2009b). Note that Dlx1/2 constitutive mutants fail to induce expression of Dlx5&6; thus, these mutants lack expression of any Dlx gene (Anderson et al., 1997a). In addition to the role of the Mash1/Dlx1/Dlx2 transcriptional cascade in promoting neuronal differentiation, they also repress gliogenic differentiation (Yun et al., 2002; Petryniak et al., 2007). In the Dlx1/Dlx2 mutants, while GE neurogenesis was nearly abolished, Olig2 expression was increased (Petryniak et al., 2007). Dlx1/2 have a modest role in GE regional patterning; in Dlx1/2 mutants, the dorsal LGE is partially transformed toward a cortical fate (Long et al., 2009a).

What is the mechanism that balances gliogenesis versus neurogenesis in the VZ? Petryniak et al. and Silbereis et al. demonstrated that Olig2 and Olig1 are key TFs that control the neuron-glial fate switch in the GE VZ (Petryniak et al., 2007; Silbereis et al., 2014). Olig1 expression preserves progenitors’ gliogenic potential, and it represses Dlx1/2 through the DlxI12b enhancers (Ghanem et al., 2007a; Silbereis et al., 2014). Constitutive deletion of Olig1 results in (1) excessive expression of Lhx6/Dlx1/Dlx2 in the ventral MGE, (2) a decrease in oligodendrocyte progenitors in the GE, and (3) ~30% excess production of INs that are primarily PV+ and CR+. Of note, PV+ and CR+ INs are produced by late progenitors in the GE that also have the potential to generate oligodendrocytes. Olig1 null mutants had no changes in NPY+ and SST+ INs, perhaps because they are produced before oligodendrogliosis (Petros et al., 2015). It is likely that Olig1 controls the neuronal-glial fate switch only within a restricted time window. Olig2 shares overlapping expression pattern and redundant functions with Olig1.

There are several VZ-expressing TFs that play critical roles in guiding GE regional specification. For instance, Nkx2.1 and Otx2 guide MGE specification through their VZ expression to repress LGE/CGE fate (Sussel et al., 1999; Hoch et al., 2015a). Nkx2.1 loss-of-function impedes MGE neurogenesis and MGE-lineage reduction, and expansion of LGE/CGE-derived neuronal population (Sussel et al., 1999; Butt et al., 2005, 2008). Gsx1 and Gsx2, on the other hand, drive the identity of the LGE and CGE. Gsx1/Gsx2 double deletion resulted in hypoplastic CGE, loss of Dlx expression in the LGE SVZ, and a subsequent loss of LGE-derived striatal and OB neurons (Toresson and Campbell, 2001; Corbin et al., 2003; Yun et al., 2003).

GE Subventricular Zones

As VZ progenitors mature, they migrate superficially into the SVZ to become intermediate progenitors or neurons. In the mouse GEs, there are at least two SVZ layers. SVZ1 is next to the VZ; it is compact and largely contains proliferating cells. SVZ2 is superficial to SVZ1, is less compact, and is a mixture of some proliferating cells and many postmitotic neurons, and later glia (Yun et al., 2002; Petryniak et al., 2007). Intermediate progenitors have limited proliferative capacity and are more fate limited than the VZ. Note that the SVZ harbors early postmitotic neurons generated in the VZ and SVZ. As noted later, the Dlx genes have key roles in the cell biology of intermediate progenitors, including promoting GABAergic neurogenesis and differentiation through promoting the expression of Gad1, Gad2, and Vgat, enzymes, and a transporter required for GABA synthesis and loading into synaptic vesicles (Stühmer et al., 2002a; Stühmer et al., 2002b; Long et al., 2007; Yu et al., 2011; Le et al., 2017; Pla et al., 2018).

Functions of TFs in the MGE SVZ

The MGE generates a variety of cell types, including cortical and hippocampal INs, GABAergic GP projection neurons, cholinergic striatal neurons, and glia (Petryniak et al., 2007; Xu et al., 2008; Flandin et al., 2010). Two major transcriptional pathways are proposed to regulate MGE neurogenesis: (1) Nkx2.1– and (2) Dlx-pathways. Nkx2.1 is predominantly expressed in the MGE VZ and SVZ, where it drives SVZ expression of Lhx6 and Lhx8, TFs required for the specification of pallial and striatal INs (Zhao et al., 2003, 2008; Liodis et al., 2007; Fragkouli et al., 2009). Lhx6 subsequently induces the expression of Mafb, Shh, Sox6, and other genes that are required by PV+ and SST+ INs to acquire their physiological properties and cell fates (Sussel et al., 1999; Liodis et al., 2007; Zhao et al., 2008; Batista-Brito et al., 2009). Similar to Nkx2.1 loss of function, Lhx6 deletion results in a partial cell fate switch of MGE INs to CGE-like INs (Vogt et al., 2014). Importantly, Lhx6 mutant MGE-derived cells are still GABAergic but lack later defined CIN properties. This evidence supports the hypothesis that subsequent TF cascades define more developmentally refined aspects of these cells. For instance, loss of Dlx genes leads to a neuron versus glial versus cell fate switch and a failure to repress Nkx2.1 (see below), while loss of Lhx6 in the SVZ leads to a loss of CIN cell identity.

While Nkx2.1 and Lhx8 expression is repressed in Lhx6+ MGE-derived INs, a subset of MGE-derived neurons maintain the expression of Nkx2.1 and Lhx8; the latter is required for the differentiation of striatal cholinergic neurons (Zhao et al., 2003; Fragkouli et al., 2009; Nóbrega-Pereira et al., 2010).

Dlx1/2 drive SVZ expression of Zfhx1b that guides the CIN versus striatal IN fate decision (McKinsey et al., 2013). Zfhx1b promotes CIN differentiation by repressing Nkx2.1, which prevented CINs from migrating to the striatum, and maintaining pallidal projection neuron production by sustaining the expression of Lhx8 (McKinsey et al., 2013). It is worth noting that while Lhx6 and Lhx8 guide CIN versus striatal IN differentiation, respectively, they also share functions in driving Shh expression in the MGE MZ (Zhao et al., 2008; Flandin et al., 2011).

Functions of TFs in the CGE VZ and SVZ

The CGE is the caudal extension of LGE and MGE, and it produces ~20%–30% of CINs. Although there is no known specific VZ marker for the CGE, it does have high levels of Gsx2 and Nr2f1/2 expression (Kanatani et al., 2008; Lodato et al., 2011; Wang et al., 2013). Its SVZ shows strong expression of Dlx TFs (Long et al., 2009b). Dlx1/2 play crucial roles in CGE CIN differentiation. For instance, Dlx1/2 repress Hes1/5, Nr2f1, Mash1, and Olig2 expression (Long et al., 2009b). In the absence of Dlx1/2, ectopic expression of MGE (Gbx1), LGE (Islet1), and diencephalon (Otp) markers is observed in the CGE (Long et al., 2009b).

Sp8 is required in the survival, migration, and fate specification of OB interneurons (Waclaw et al., 2006). Sp8 is also present in the SVZ of the CGE. Sp8 labels a small subset of CGE-derived CINs that are not NR2F2+ or VIP+ (Ma et al., 2013). The Sp8+ CINs maintain Sp8 expression as they integrate in the neocortex following an inside-out manner (SVZ/IZ → cortical plate) (Ma et al., 2013; Wei et al., 2019). Sp8 and Sp9 together regulate CGE-derived cortical IN maturation by repressing Pak3, Robo1, and Slit1 expression (Wei et al., 2019).

Prox1 expression is enriched in the CGE SVZ. It is proposed that Prox1 might function independently of Dlx1/2 to repress Notch1 expression, promote cell cycle exit, and to facilitate neuronal differentiation (Kaltezioti et al., 2010; Rubin and Kessaris, 2013). Prox1 expression in postmitotic CGE IN precursors is important, as it differentially preprograms Reelin+ and VIP+ INs migratory routes, morphology maturation, and their firing properties. Deletion of Prox1 in the CGE results in ~40% reduction of CGE-derived CINs due to defective SVZ/IZ to cortical plate radial migration (Miyoshi et al., 2015).

CGE-derived neurons populate the pallium and subpallium through distinct migratory routes that are orchestrated in part by TFs expressed in the SVZ and immature neurons. For instance, Prox1/Sp8+ INs migrate toward the neocortex through the lateral migratory stream. Nr2f1 functions antagonistically with Sp8 and guides Sp8– CGE-derived INs along a medial migration toward subcortical nuclei like the amygdala (Touzot et al., 2016). Nr2f2, although it shares the same TF family with Nr2f1, marks CGE derivatives that are negative for Sp8. The Nr2f2+ INs preferably integrate into the neocortex and hippocampus through the caudal migratory stream (Miyoshi et al., 2015; Touzot et al., 2016).

Functions of TFs in the LGE SVZ for the Generation of OB INs

The LGE has dorsal and ventral parts (dLGE and vLGE). dLGE predominantly produces OB INs and striatal medium spiny neurons (MSNs), while vLGE largely generates MSNs (Yun et al., 2001; Corbin et al., 2000; Toresson and Campbell, 2001; Stenman et al.,2003; Yun et al., 2003). Dlx1/2 expression is high in the dLGE and low in the vLGE. In the absence of Dlx1/2, dLGE intermediate progenitors and immature neurons fail to express Gad67 and vGat; ventral pallial, MGE, and hypothalamic markers (Id2, Gbx1/2, Otp) are ectopically expressed in the LGE (Long et al., 2007; Long et al., 2009a; Anderson et al., 1997a), illustrating Dlx1/2’s roles in regional and cell-type specification.

Dlx1/2 have key roles in the dLGE in guiding OB IN fate specification and migration by promoting Arx, Etv1, Pbx3, Prokr2, Sp8, Sp9, Tshz1, and Vax1 (Soria et al., 2004; Yoshihara et al., 2005; Waclaw et al., 2006; Long et al., 2007; Guo et al., 2019).

Mechanisms Proposed for the Generation of SST and PV CIN

The contribution of VZ and SVZ progenitors to IN fate specification has been best studied in the MGE, with progress in understanding the mechanisms of PV and SST CIN specification. Several TFs have been shown to differentially mediate IN subtype specification along the MGE VZ-SVZ1-SVZ2 axis. For instance, Mafb and c-Maf in the SVZ repress SST IN neurogenesis and promote PV IN generation (Pai et al., 2019). In contrast, Nr2f2 in the VZ and SVZ promotes SST and represses PV IN production (Hu et al., 2017a).

Four MGE cell fate specification models have been proposed (Hu et al., 2017b). In model 1, different dorso-ventral and rostro-caudal regions of the MGE VZ/SVZ are biased but may not be specific for generating distinct MGE-derived INs and projection neurons (Flames et al., 2007; Flandin et al., 2010). The subsequent models proposed mechanisms within a given regional subdomain. In model 2, the “mosaic model,” distinct progenitors in the VZ are committed (or strongly biased) to generate either SST+ or PV+ CINs. Currently there is no strong evidence for this. In model 3, the “homogenous VZ model,” VZ progenitors can generate both subgroups. Retroviral labeling and lineage tracing of VZ progenitors has shown that individual MGE progenitors can generate both SST+ and PV+ CINs (Brown et al., 2011; Harwell et al., 2015; Mayer et al., 2015). In model 4, the “direct versus indirect neurogenesis model,” VZ progenitors generate SST+ CINs (direct neurogenesis), whereas SVZ progenitors generate PV+ CINs (indirect neurogenesis). The support for model 4 comes from evidence that the timing of progenitor cell cycle exit can independently dictate neuronal fates (Glickstein et al., 2007; Lodato et al., 2011; Petros et al., 2015). Petros et al. showed that SST INs are mainly produced by MGE VZ progenitors that exit the cell cycle and initiate differentiation early, while PV INs are mainly produced by MGE progenitors that proliferate and cycle longer (from VZ to SVZ) (Petros et al., 2015). By repressing Notch signaling and pushing progenitors to exit the cell cycle in the SVZ, more SST IN production was observed; when the cell cycle was extended in the SVZ, more PV IN production was noticed (Glickstein et al., 2009; Petros et al., 2015). Finally, the observation of ventrally displaced progenitors from MGE0/1 adds another mechanism that may contribute to the specification of neuronal diversity (Hu and Rubenstein, unpublished). Thus, multiple mechanisms could contribute to cell fate specification in the GEs. Future studies using single cell in vivo lineage analysis, spatial transcriptomics, single-cell RNA-seq coupled with RNA velocity, and cell cycle status analyses are needed to better understand these mechanisms (Kowalczyk et al., 2015; La Manno et al., 2018; Hsiao et al., 2020).

Functions of TFs Expressed in Migratory and Post-Migratory Immature INs

After their roles in progenitors are complete, many TFs also mediate CIN migration, differentiation, maturation, and function. Here, we will summarize some of the TF functions in the order of when these TFs start to be expressed in the GEs (VZ/SVZ: Dlx, Nr2f2; SVZ: Arx, Npas1/Npas3, Mafb/c-Maf; MZ and immature INs: Satb1, Mef2c).

Postmitotic Roles of Dlx TFs in IN Survival, Morphogenesis, and Synapse Formation and Function

Dlx expression is restricted to forebrain GABAergic progenitors (GEs) and several subtypes of GABAergic neurons, including CINs/HINs/OB INs. In the stem cell biology section, we addressed that Dlx1/2 initiates the expression of Dlx5/6 in the GE progenitors to drive neurogenesis. Dlx1/2/5/6 expression and their downstream transcriptional cascades also provide guidance for INs to properly migrate and mature (Anderson et al., 1997a, 1997b; Eisenstat et al., 1999; Stühmer et al., 2002a, 2002b; Cobos et al., 2005b; Petryniak et al., 2007; Long et al., 2009a, 2009b; McKinsey et al., 2013). For instance, Dlx1/2 promote the expression of cytokine receptors (Cxcr4 and Cxcr7), which allow INs to migrate along pathways secreting the ligand Cxcl12 (Wang et al., 2011; Wu et al., 2017; Cobos et al., 2007). To facilitate tangential migration, Dlx1/2 also delay the initiation of IN axon and dendritic growth (Cobos et al., 2007).

To further understand the postnatal function of Dlx1/2, postmitotic and postnatal Dlx1/2 conditional deletion mouse models were generated using Calretinin (CR)-Cre, Sst-Cre, PV-Cre, and DlxI12b-Cre (Hippenmeyer et al., 2005; Potter et al., 2009; Taniguchi et al., 2011; Silbereis et al., 2014; Pla et al., 2018; Sohal and Rubenstein, 2019). Deletion of Dlx1/2 in postmitotic neurons does not disrupt the expression of Dlx5/6. As a result, IN generation and migration are not affected in these mutants. However, several changes were observed after Dlx1/2 loss, including IN survival, dendritic morphogenesis, and synaptic number/function. These phenotypes derive likely in part from the decreased expression of Gad1, Gad2, Vgat, Cxcr4, and Grin2b (Cobos et al., 2007; Long et al., 2009; Le et al., 2017; Pla et al., 2018). Specifically, the conditional Dlx1/2 mutants had extensive CIN cell death affecting all classes of CINs by P7; at later stages, CINs failed to receive proper excitatory inputs due to hypoplastic dendritic arbors and formed fewer inhibitory synapses onto excitatory neurons, resulting in reduced mIPSC frequency (Pla et al., 2018). HINs with constitutive loss of Dlx1 also exhibit similar hypoexcitability and reduced glutamatergic excitation, supporting a role for Dlx1 in IN maturation (Jones et al., 2011; Howard et al., 2014). Of note, while Dlx genes are generally imporant for postnatal IN survival, PV INs maybe be less affected as Dlx1 expression is downregulated earlier in these progenitors. In addition. Dlx1 constitutive mutants exhibited ~30% pan CIN/HIN loss except for the PV subpopulation (Cobos et al., 2005a).

Dlx5/6 homozygous mutants have exencephaly, which makes it difficult to study Dlx5/6 in CIN development (Schüle et al., 2007). Using an MGE transplantation assay, Dlx5/6 loss of function was found to preferentially result in reduced PV CINs (Wang et al., 2010). This impact on PV CINs was corroborated in Dlx5/6 heterozygous mutants, where electrophysiologcal firing defects were observed in PV INs, leading to alterations in behavior and epilepsy (Cho et al., 2015). The transcriptional cascades downstream of Dlx5/6 are currently an active pursuit in the field.

Nr2f1/Nr2f2 Promotes CGE Fate Specification and MGE Generation of SST CIN

Nr2f1 and Nr2f2 (CoupTFI and CoupTFII) nuclear receptor TFs have regional expression within the MGE and pan-CGE expression (Kanatani et al., 2008; Lodato et al., 2011; Cai et al., 2013; Hu et al., 2017a). Conditional deletion of Nr2f1/Nr2f2 in GE intermediate progenitors, using Dlx5/6-Cre, reduced CGE-derived INs (VIP+; CR+) and increased MGE proliferation, including an excessive production of PV-expressing INs (Lodato et al., 2011). In addition, Nr2f1/Nr2f2 have instructive roles in guiding CGE-derived CINs to integrate into the neocortical circuit through the caudal migratory stream (Tripodi et al., 2004; Kanatani et al., 2008; Touzot et al., 2016). Forced expression of Nr2f2 is sufficient to convert the MGE-derived INs to migrate in a manner similar to the CGE-derived INs (Kanatani et al., 2008), suggesting the regulation of potential guidance cues by these genes during development.

In the MGE, Nr2f1/Nr2f2 expression is in the dorsal domain with a caudo-rostral gradient and is complementary to the Otx2-expressing ventral rosto-caudal gradient (Hoch et al., 2015a; Hu et al., 2017a). Nr2f1/Nr2f2 represses Ccnd2 and promotes Sox6 to drive SST+ IN fate specification (Glickstein et al., 2007; Azim et al., 2009; Batista-Brito et al., 2009; Hu et al., 2017a). Nr2f1/Nr2f2 conditional nulls also exhibited lengthened cell cycle and a shifted SST/PV ratio in the MGE that led to an increase in PV+ CINs (Glickstein et al., 2007; Hu et al., 2017a). Finally, forced expression of Nr2f2 in MGE cells increased Sox6 expression, suppressed PV fate, and promoted SST fate; expression of Sox6 in MGE cells also resulted in increased SST+ fate at the expense of PV (Vogt et al., 2014; Hu et al., 2017a).

Role of Arx in MGE-Derived IN Migration and PV IN Fate Specification

Arx encodes an X-linked homeodomain protein; when mutated, humans have severe epilepsy and cognitive impairment (Kitamura et al., 2002). In the developing telencephalon, Arx is expressed in pallial VZ progenitors, subpallial SVZ progenitors, and immature neurons, including INs tangentially migrating to the pallium (Friocourt et al., 2006; Colasante et al., 2008, 2009; Friocourt and Parnavelas, 2010). Constitutive mouse Arx mutants, deletions, or triplet repeat expansions modeling human mutations have molecular defects in the MGE and decreased numbers of CINs, as well as epilepsy (Kitamura et al., 2002; Colasante et al., 2009; Price et al., 2009). Deletion of Arx in newborm subpallial neurons also results in CIN defects that appear to result in part from migration defects (Marsh et al., 2009, 2016); some of the abnormal migration is linked to reduced expression of Cxcr4 and Cxcr7 (Fulp et al., 2008).

Arx expression is promoted and maintained by the Dlx and Lhx6 genes (Cobos et al., 2005; Colombo et al., 2007; Colasante et al., 2008; Zhao et al., 2008; Vogt et al., 2014). Furthermore, Arx re-expression in the MGE of telencephalic slices of E14.5 Dlx1/2 mutants enables immature INs to tangentially migrate to the cortex (Colasante et al., 2008). In the Lhx6 null mutant, MGE-derived CINs fail to express PV and SST. Arx complementation of Lhx6 loss-of-function INs, using an MGE transplantation assay, was sufficient to induce PV expression, suggesting Arx has a role in fate promotion in postmitotic MGE-derived INs (Vogt et al., 2014). Arx’s role in postmitotic CGE-derived INs is uncertain.

Opposing Roles of Npas1 and Npas3 in the MGE/CGE-Derived IN Generation and Differentiation

Npas1 and Npas3 bHLH-PAS TFs are expressed in MGE/CGE progenitor zones and in immature and mature CINs (Zhou et al., 1997; Erbel-Sieler et al., 2004; Stanco et al., 2014). While coexpression of Npas1 and Npas3 had been identified in GAD67-expressing CINs, these two TFs exhibit distinct functions to guide IN progenitor proliferation and differentiation (Erbel-Sieler et al., 2004; Stanco et al., 2014). In the MGE and CGE, Npas1–/– had increased proliferation, ERK signaling, and Arx expression. They also generated an excess of SST and VIP CINs. In contrast, Npas3–/– showed decreased progenitor proliferation and ERK signaling and generated fewer SST+ and VIP+ INs. The deletion of Npas1 and Npas3, however, did not affect the generation of PV CINs.

In addition, Npas1 and Npas3, or Npas3 alone may be crucial in guiding Reelin+ CIN production and differentiation. A drastic reduction of Reelin+ CINs is observed both in the Npas3 null and the Npas1/Npas3 double mutants (Erbel-Sieler et al., 2004; Stanco et al., 2014). Of note, Npas1 also mediates IN spontaneous activity. In the Npas1 null, upper-layer pyramidal neurons receive higher frequency inhibitory inputs from CINs, which implies a role of Npas1 in regulating IN excitability (Stanco et al., 2014).

Roles of the Mafb and c-Maf TFs in PV/SST IN Fate Specification, Migration, and Maturation

Mafb and c-Maf are BZIP TFs that are expressed in the MGE SVZ and in immature/mature INs generated from the MGE. The roles of Mafb and c-Maf were explored in single and double mutants (Pai et al., 2019, 2020). Nkx2.1-cre-depleted single Mafb or c-Maf conditional mutants did not show severe phenotypes, whereas double deletion mice (Maf cDKO) exhibited strong phenotypes, suggesting compensatory roles. First, there is ~70% reduction of total MGE-derived INs in the adult neocortex, indicating that the Mafb and c-Maf work together to produce the proper amount of mature INs. This cell loss did not start until P0. Second, of the INs that remained, the proportion expressing PV was decreased by about half, while there was no change in the proportion of SST INs. Looking at earlier developmental stages, ~50% increase in SST INs was observed at neonatal ages, just before IN programmed apoptosis would occur (Southwell et al., 2012). While the increase of the SST CIN phenotype is transient, and is diminished by ~P7, it highlights a potential mechanism whereby an early increase in one subtype of IN could potentially outcompete another during early stages of development resulting in skewed ratios by the time the brain reaches adulthood. Third, cells were becoming postmitotic/neuron fated at a higher rate in the mutants in the MGE progenitor domain, potentially in line with one hypothesis of how SST CINs could be preferentially generated (Petros et al., 2015). Finally, INs prematurely migrated into the cortical plate layer of the developing neocortex, suggesting further migratory regulation by these TFs. This could be partially explained by the decrease in Cxcr4 expression, as Cxcr4 constitutive null mice exhibit similar migration defects (Stumm et al., 2003; Wang et al., 2011; Pai et al., 2019, 2020).

In a later study, single-cell RNA sequencing showed that Mef2c was downregulated in the Maf cDKO neonatal pallial INs. Mef2c conditional ablation in postmitotic INs showed preferential loss of PV CINs, similar to what was observed in the Maf cDKO (Mayer et al., 2018; Pai et al., 2020). By restoring the expression of Mef2c in Maf cDKOs, PV expression was rescued (Pai et al., 2020). A transcriptional pathway was proposed in which Mafb and c-Maf promote the expression of Mef2c to drive PV IN specification (Mayer et al., 2018; Pai et al., 2020).

When Maf single and double mutant CINs were evaluated through whole-cell patch-clamp recordings, Mafb cKO and c-Maf cKO showed opposite spontaneous firing patterns, while Maf cDKO remained relatively normal. This evidence suggests that Mafb and c-Maf have distinct roles in CIN differentiation and physiology (Pai et al., 2019). A separate study focused on the function of Mafb in migratory SST+ Martinotti CINs. Their result suggested that Mafb can serve as an intrinsic factor that preprograms the migratory route for Martinotti cells and their layer I dendritic arborization (Lim et al., 2018). After Mafb depletion (using SST-Cre), Martinotti CINs failed to migrate to the neocortex through the marginal zone. These mutant CINs also formed poor layer I dendritic trees and failed to inhibit layer II/III pyramidal cells properly (Lim et al., 2018).

Activity-Dependent Expression of Satb1 Drives SST CIN Maturation

The expression of Satb1 homeobox TF initiates several days after the induction of Lhx6; it becomes more prominent when MGE-derived INs terminate their tangential migration and undergo radial migration into the cortical plate (Denaxa et al., 2012). Loss of Satb1 expression in postmitotic MGE-derived INs leads to reduced survival of postnatal PV and SST CINs and more severe deficit in the SST subtype (Close et al., 2012).

IN morphology and connectivity diversification occurs as immature INs integrate into the cortical circuitry (Moody and Bosma, 2005; Cossart, 2011; Denaxa et al., 2012). Satb1 may translate environmental cues during SST CIN differentiation (Denaxa et al., 2012). At perinatal stages, Satb1 expression is downstream of Lhx6 and is also activity-dependent; potassium chloride-induced depolarization in the INs is sufficient to promote Satb1 expression (Denaxa et al., 2012; Neves et al., 2013). Satb1 expression then promotes (1) Sst and other SST IN subtype marker (Npy, Calretinin) coexpression and (2) the spontaneous excitability of SST INs (Close et al., 2012; Wonders and Anderson, 2006; Denaxa et al., 2012).

Mef2c Controls PV CIN Differentiation

The Mef2c MAD box TF is expressed in both excitatory and inhibitory cortical neurons. It is also a risk gene for autism spectrum disorder and schizophrenia. Haploinsufficiency of MEF2C in humans can lead to a range of symptoms, including but not limited to autistic behaviors and intellectual disability (Li et al., 2008; Le Meur et al., 2010; Novara et al., 2010; Paciorkowski et al., 2013; Harrington et al., 2016; Tu et al., 2017). In excitatory neurons, Mef2c is a transcriptional repressor that regulates expression of synaptosomal genes. Mef2c depletion in the Emx1-Cre lineage (pyramidal neurons) results in an inverted excitation/inhibition circuit balance in the somatosensory cortex. Mutant excitatory neurons exhibited increased inhibitory inputs and a decrease in the formation of excitatory spine density (Harrington et al., 2016).

The role of Mef2c was surveyed in the mouse GE and CINs across different developmental time points (E13.5, E14.5, E18.5, P0, and P56) using single-cell transcriptomic and computational techniques (Mayer et al., 2018). Based on the presumption that each GE progenitor and their progenies share a group of developmentally conserved markers, transcriptomic features of mature PV CINs were traced back to MGE progenitors. In this way they identified that Mef2c is a gene that is enriched in immature MGE cells that will become mature PV CINs (Mayer et al., 2018). To test the function of Mef2c in PV IN development, Mef2c expression was deleted in GEs using the Dlx6a-Cre (Mayer et al., 2018). These mutants showed PV CIN reductions, similar to what was observed in the Nkx2.1-Cre generated Mafb/c-Maf double mutants (Mayer et al., 2018; Pai et al., 2019). A MGE transplantation rescue assay, together with MAFB ChIP-seq, suggested that Mafb/c-Maf could be the upstream TFs driving Mef2c expression that promotes PV IN differentiation (Pai et al., 2020).

Genomic Approaches to Understand Transcriptional Control of IN Development

Genomic-level understanding about the transcriptional control of CINs has come from single-cell RNA sequencing (scRNA-seq), epigenomic assessment on chromatin state, and the identification of noncoding DNA regulatory elements (REs; also named enhancers) (Ghanem et al., 2007; Visel et al., 2013; Sandberg et al., 2016; Silberberg et al., 2016; Lim et al., 2018; Mayer et al., 2018; Mi et al., 2018; Lindtner et al., 2019; Rubin et al., 2020). In this last section, we will review several recent scientific discoveries that refine GE progenitor regulation and CIN subtyping using these novel technologies.

Forebrain Enhancer Identification and Functional Characterization

The sequential and overlapping expression of Dlx1/2/5/6 drives basal ganglia development. Several intergenic regulatory elements (enhancers) of Dlx genes have been identified, including Dlx1/2 I12b and URE2, and Dlx5/6 I56i and I56ii. Their regional activities and function have been documented (Zerucha et al., 2000; Ghanem et al., 2007a; Fazel Darbandi et al., 2016). URE2 was the only Dlx RE active in the VZ between E11.5 and E15.5. The URE2 active progenitors formed radial clusters, suggesting that these VZ stem cells might be clonally related. I12b/I56i exhibited strong and overlapping expression in the SVZ, complementary to the weak URE2 activity in the same domain. This differential RE expression pattern could be one mechanism that delineates the border between VZ/SVZ/MZ. In addition, the majority of migrating immature INs that insert into the neocortex in multiple zones were equally labeled by I12b and I56i, although a small subset of INs migrating along the marginal zone showed single labeling for URE2 but not I12b/I56i. The activity of these three REs continues to be observed in mature CINs, and I12b/I56i broadly labels derivatives of the LGE, MGE, and CGE lineages; on the other hand, URE2 may be preferentially active in PV+, CR+, and NPY+ (but not SST+) INs (Ghanem et al., 2007a). Thus, distinct REs, within either the Dlx1/2 or Dlx5/6 loci, drive expression in distinct distributions of cells at different developmental stages.

Higher throughput approaches have greatly expanded our understanding of the complexity and specificity of gene regulatory elements that are active in the developing forebrain. In 2013, Visel et al. identified nearly 5000 candidate embryonic forebrain enhancers using DNA conservation, proximity to genes that regulate forebrain development, and p300 ChIP-seq followed by an in vivo transgenic assay of transcriptional activity (Visel et al., 2013). Through RE-driven transient transgene genetic tools, the activity and expression pattern of 145 evolutionarily conserved forebrain enhancers were studied in vivo, which further lead to a digital atlas that remains one of the most comprehensive resources for RE studies (Visel et al., 2013; https://enhancer.lbl.gov/). Some interesting discoveries are listed here: (1) a subset of enhancers whose expression pattern mirrors their associated genes, which makes RE-based transgenic tool a potential surrogate for studying gene functions (Frankel et al., 2010; Visel et al., 2013). (2) Some enhancers show spatial and temporal redundancy. (3) Approximately two-thirds of the candidate enhancers identified from human fetal brain tissues have orthologous sites in rodents, some of which may be implicated in neurodevelopmental or neuropsychiatric disorders (Bejerano et al., 2004; Sebat et al., 2007; Walsh et al., 2008; Cooper et al., 2011; Visel et al., 2013).

Hundreds of ultraconserved enhancers have been conserved over millions of years of evolution (Bejerano et al., 2004). While initial mutagenesis analyses found no overt phenotypes of deletion of some enhancers (Ahituv et al., 2007), it was likely that functional redundancy was the cause, particularly for the Arx locus because it has two pairs of enhancers that are active in the GEs or in the cortex (Ahituv et al., 2007; Colasante et al., 2008; Chiang et al., 2008; Visel et al., 2013). Dickel et al. (2018) explored this further by generating single or multiple enhancer deletion mouse lines, specifically focusing on the region containing the Arx locus and eight nearby enhancers (Bejerano et al., 2004; Dickel et al., 2018). CRISPR/Cas9 was utilized to delete individual and pairs of enhancers. While Arx null mice are lethal (Kitamura et al., 2002; Colombo et al., 2007), these enhancer double mutants were viable and fertile (Dickel et al., 2018). However, there was significant decreased Arx expression in the corresponding embryonic brain regions. In line with Arx’s function in GE-derived IN and striatal neuron migration and differentiation (Kitamura et al., 2002; Colombo et al., 2007; Colasante et al., 2009; Vogt et al., 2014), one double mutant (hs119 & hs121 null) had a near total loss of cholinergic striatal neurons, and a ~20%–30% decrease in PV and VIP CINs (Dickel et al., 2018).

Epigenetic Functions of Dlx1/2/5, Gsx2, Lhx6, Nkx2.1, and Otx2 during GE Development

Recent genomic studies have identified noncoding REs that possess dynamic regional and temporal-specific activity across forebrain development (Visel et al., 2009, 2013; Nord et al., 2013, 2015; Silberberg et al., 2016). How TFs interact with REs that drive cell fate and differentiation is a burgeoning field. Two recent studies revealed sets of Dlx and Nkx2.1-bound REs, and how Dlx and Nkx2.1 transcriptionally interact with REs to drive GE subregion fate specification by combining bulk RNAseq, TF chromatin-immunoprecipitation followed by sequencing (ChIP-seq), and histone ChIP-seq (Sandberg et al., 2019).

To understand how Dlx TFs coordinately regulate the transcriptional programs that drive GE development, Lindtner et al. (2019) utilized a functional genomic approach by conducting integrated analyses on (1) DLX1/2/5 ChIP-seq and histone ChIP-seq from E11.5, E13.5, and E16.5 Dlx1/2 wild-type and null embryonic GEs, and (2) RNA seq from E13.5 Dlx1/2 wild-type and null GEs (Lindtner et al., 2019). Sequential expression of Dlx2/1/5 in GE VZ and SVZ represses TFs Hes5 and Otx2 to enable GE stem cells to undergo neurogenesis (Eisenstat et al., 1999; Yun et al., 2002; Petryniak et al., 2007; Long et al., 2009; Hoch et al., 2015a; Lindtner et al., 2019). Their data suggested that Dlx TFs bind at enhancers and transcriptional starting sites (TSS, promoters) to repress the genes that drive proliferative cell states (Lindtner et al., 2019). Dlx TFs also bind distal regulator elements, which promote gene expression that drives GABAeric neuron fate specification and differentiation. For instance, Dlx promotes Sp8 expression, which drives LGE fate. Through DLX and H3K27ac ChIP-seq, they identified several Sp8 distal regulatory sequences bound by DLX that correspond to characterized GE enhancers. Transcription assays validated that DLX TFs promote RE-dependent transcription, with some enhancers having stronger effects. When the RE binding affinity was interfered (using CRISPRi), E13.5 cultured LGE neurons exhibited down-regulated Sp8 RNA expression. Furthermore, TF binding motif analyses revealed multiple active REs (a.RE) and repressive REs (r.RE) that modulate gene expression to facilitate Dlx function in GE biology. These include (1) a.REs for Gad1/2, Arx, Cxcr4/ Cxcr7, and Nrxn3 genes that promotes IN fate specification and maturation, and (2) r.REs for Pax7, Lhx2, and Otp, whose repression prevents GE progenitors from adopting hypothalamus and midbrain neuronal fates (Long et al., 2009; Lindtner et al., 2019). Together these results are elucidating the transcriptional circuitry that drives the development of forebrain GABAergic neurons.

A similar functional genomic approach was adopted to explore the genome-wide binding activity of NKX2-1 in the developing basal ganglia (Sandberg et al., 2016). Their data suggested that NKX2-1 has both activating and repressive roles on RE regulation to mediate MGE development. NKX proteins recruit Gro/TLE proteins that facilitate chromatin condensation, which leads to reduced expression of TFs and regulators in the SHH/WNT/BMP signaling pathways to pattern the ventral CNS (Muhr et al., 2001; Patel et al., 2012).

NKX2-1 acts as a repressor in the MGE VZ, where it binds to r.REs to quell the expression of LGE genes (Sandberg et al., 2016). However, NKX2-1 adopts an activator role in the MGE SVZ/MZ and binds to a.REs to promote the expression of Lhx6/Lhx8 and other genes that promote MGE-neuronal differentiation. In the MGE-derived CIN lineage, Nkx2.1 expression becomes repressed, in part via Zfhx1b (McKinsey et al., 2013), while Lhx6 is necessary and sufficient for the MGE-derived CIN maturation (Du et al., 2008; Vogt et al., 2014). Indeed, several genes expressed in CINs, such as Arx, Cxcr7, Maf, and Nxph1, are activated through LHX6-bound REs that lack NKX2-1 binding. On the other hand, two-thirds of the LHX6-bound REs have the NKX2-1 DNA-binding motif. These dual-bound distal REs are hypothesized to drive the expression of genes that are either specific to MGE-derived pallidal projection neurons, or with broad expression in cells within the MGE lineage.

These DLX1/2/5, NKX2-1, and LHX6 studies are unveiling the epigenetic complexity that coordinates gene regulation in GE development. They also support the hypothesis that REs serve as competitive binding hubs for cell-type-specific TFs, and it is the RE-TF interaction-driven combinatorial gene expression that facilitates neuronal subtype fate specification and development. Indeed, computational analyses of 10 different TFs binding to MGE-specific and LGE-specific enhancers showed that NKX2-1 and LHX6 were among the most highly correlated TFs binding to enhancers with MGE-specific activities (Silberberg et al., 2016).

MGE regional specification is driven by parallel functions of Nkx2.1 and Otx2 (Hoch et al., 2015a). OTX2 genomic binding provided evidence for its direct roles in regulating enhancers that control MGE regional specification and in promoting neurogenesis and oligodendrogenesis (Hoch et al., 2015a). As described earlier, LGE specification is driven by Gsx2. This TF has been shown to bind to as monomer to a monomer motif, and dimer to a motif; the monomeric motif is associated with repression and the dimer motif with activation (Salomone et al., 2021). The site where GSX2 and DLX co-bind is associated with gene activation in the LGE.

Molecular and Genetic Tools Utilizing Enhancers

Transgenic approaches to study cell lineages and genetic modifications in specific cell types have been widely adopted for rodent research, with a plethora of developed models (e.g., Taniguchi et al., 2011). It is difficult to translate these approaches to primate, human, or other organisms that are not readily accessible to gene manipulations. However, the use of viral vectors, such as lentiviruses and adeno-associated viruses (AAVs), has opened the door to such approaches. Recent progress with viral vectors that utilize conserved DNA enhancers to bias expression in GABAergic cells has drawn attention for cell type-targeted labeling. This has been recently highlighted and easily applied to different model organisms, with the potential for developing gene therapies (Nathanson et al., 2009; Betley and Sternson, 2011; Lee et al., 2014; Dimidschstein et al., 2016; Sun et al., 2016; Rubin et al., 2020; Vormstein-Schneider et al., 2020). While each virus type has advantages, some limitations exist: (1) there are multiple AAV serotypes with different tropisms, some with poor neuron transduction; (2) AAVs have a more limited genome packaging capacity (Wu et al., 2010; Choi et al., 2014); (3) lentiviruses have greater genomic packaging load but lack the variety of serotypes that have been worked out for AAVs; (4) unlike AAVs, lentiviruses integrate randomly into the host genome, potentially disrupting gene function in a subset of cells.

Regulatory elements (REs, also named enhancers) are noncoding DNA fragments that regulate gene transcription through DNA interactions with TFs (Long et al., 2016; Visel et al., 2013). Transgenic in vivo studies have identified evolutionarily conserved REs that have activity patterns similar to nearby genes (Visel et al., 2013; Silberberg et al., 2016). Coupled with a fluorescent reporter and/or Cre expression, REs have become a useful tool for lineage tracing. For instance, I12b, a Dlx1/2 RE was utilized to generate a Cre line. This Dlx I12b-Cre is active during the development of most MGE and CGE-derived INs in the neocortex and hippocampus, recapitulating the expression pattern of Dlx genes (Potter et al., 2009). In addition, the 1056- CreERT2 and 1538-CreERT2 transgenic alleles are active in ventral and rostro-dorsal MGE progenitors, respectively (Silberberg et al., 2016). 799 is another MGE-specific RE that is proximal to the neurexophilin1 gene. The 799-CreERT2 line targets both PV and SST CINs (Silberberg et al., 2016). Finally, the line 1060-CreERT2 has restricted expression in the CGE MZ where postmitotic neurons accumulate and targets ~80% of the CGE-derived CINs as well as a subset of CGE-originated amygdala neurons (Silberberg et al., 2016).

Building from the successes of the enhancer-driven transgenic mouse lines, subsequent studies asked whether these REs could drive biased expression in GABAergic neurons using viral transduction. One of the first set of studies used the mouse DlxI1/2b enhancer to specifically drive GFP in GABAergic CINs using lentiviral delivery (Arguello et al., 2013; Vogt et al., 2014). Using a number of conserved enhancers, including DlxI12b, Chen et al. showed that specific progenitor and neuronal cells types could be labeled using enhancer-driven lentiviruses in mouse embryonic stem cells programmed to be ventral telencephalic cells (Chen et al., 2013). Moreover, the DlxI12b enhancer has been used in human stem cells differentiated into mixed cultures including GABAergic neurons to label these cells in vitro for manipulation (Sun et al., 2016). Due to their capacity to carry large DNA cargoes, lentiviruses have also been used to combine biased expression of a reporter using an enhancer such as DlxI12b with the delivery of other genes. These approaches were able to drive Cre and/or a fluorescent reporter and specific genes under the control of an enhancer. These studies included assays to complement/rescue genetic loss of function in CINs, as well as screen how human genomic variants impact CIN development at the molecular, cellular, and electrophysiological levels (Vogt et al., 2014, 2015, 2018; Elbert et al., 2019; Malik et al., 2019; Wundrach et al., 2020). Overall, lentiviruses have shown promise, but in part due to their biosafety concerns, other viral types have been explored.

AAVs are safer than lentiviruses but carry less DNA. To this end, many AAV studies have focused on solely driving reporters/single genes to manipulate specific cell types using distinct enhancers to drive expression in GABAergic and other cell types. Several studies have taken advantage of the increased knowledge about RE activity patterns, TF interactions with REs, and the small size of REs. These studies utilized REs to give cell type and temporal specificity for targeting expression to cortical pyramidal and INs (Visel et al., 2013; Pattabiraman et al., 2014; Dimidschstein et al., 2016; Silberberg et al., 2016; Graybuck et al., 2021). For instance, REs around Dlx1/2 (I12b) and Dlx5/6 (I56i and I56ii) have been successfully utilized to generate AAVs, which reliably bias expression in a pan-GABAergic fashion across different species, and can be applied in vitro and in vivo (Lee et al., 2014; Cho et al., 2015; Dimidschstein et al., 2016; Mehta et al., 2019; Rubin et al., 2020).

When these enhancers are utilized to drive chemogenetic receptor DREADD and opsin expression in conjunction with AAV shuttling, it allows in vivo manipulation and activity to be monitored in INs (Dimidschstein et al., 2016; Vormstein-Schneider et al., 2020). Furthermore, REs close to genes Arl4d, Dlgap1, and Scn1a were recently identified that demonstrate robust expression in AAV vectors and a strong tropism toward fast-spiking PV CINs (Rubin et al., 2020; Vormstein-Schneider et al., 2020). When coupling Arld4d/Dlgap1 RE-driven AAV vector with channelrhodopsin expression, behavioral deficits derived from the lack of PV IN activity can be optogenetically rescued in the Dlx5/6 constitutive mutant, further extending the RE-AAV application (Cho et al., 2015; Rubin et al., 2020). Intriguingly, aside from IN enhancers, Graybuck et al. validated several enhancers—mscRE4 and mscRE16 provide specific labeling toward deep layer (layer V/VI) pyramidal neurons through single-cell chromatin accessibility analysis (Graybuck et al., 2021). They further showcased that multiple RE copies can be cloned into AAV vectors to enhance gene expression, and that the compatibility of the RE vector with the PHP.eB serotype allows better blood–brain barrier penetration and broader application when used with different transgenic lines (Graybuck et al., 2021).

All the above examples showcase the broad utility of enhancer elements (Fig. 47–2). Continuing efforts strive to better understand RE-TF temporal-spatial dynamic interactions and to polish RE-based molecular approaches that can facilitate cell-type-specific manipulation and therapeutics.

Figure 47–2.. Enhancer applications.

Figure 47–2.

Enhancer applications. Enhancers, also known as DNA regulatory elements (REs), regulate gene expression through their interactions with transcription factors (TFs). Complex TF-RE networks regulate temporal and spatial aspects of lineage specification. (more...)

Interneuron Classification and New Marker Discovery through Single-Cell RNA Sequencing

Historically, cell morphology, histochemical properties, intrinsic firing properties, and inter-neuronal connectivity have been utilized to classify neuronal subtypes (Rudy et al., 2011; DeFelipe et al., 2013; Greig et al., 2013). More recently, RNA expression has been employed to characterize cell subtypes. Tissue-based RNA sequencing facilitated identification of genes that specify pallial and subpallial subdomains and lineages across development. The temporal and spatial resolution from this method comes from researcher-driven microdissection of different parts of the brain and/or fluorescence- activated sorting (FACS). However, this can be labor-intensive and is subject to the accuracy of the dissection. The throughput and depth of the analyses are also limited.

Droplet and microfluidic-based high-throughput single-cell transcriptomic platforms that enable the profiling of neuronal and non-neuronal subtypes at higher temporal and spatial resolution have emerged in the past decade (Tasic et al., 2016, 2018; Paul et al., 2017; Mayer et al., 2018; Mi et al., 2018; Saunders et al., 2018). Taking advantage of the laminar-specific and IN-specific transgenic lines for targeted cell population enrichment, Tasic et al. used scRNA-seq to identify 49 clusters of cells from adult mouse visual cortex. Eighteen clusters belong to three broad GABAergic IN families (PV, SST, and VIP), and several “unique” markers were validated that can better subclassify INs (Tasic et al., 2016). For instance, cerebellum 4 (Cbln4) coexpression with Sst and Calretinin is predominantly present in upper-layer INs, suggesting the specific role of Cbln4 in labeling Martinotti cells. Similarly, Cpne5 was identified to preferentially label upper-layer PV+ chandelier INs, which was corroborated in a chandelier cell-enriched RNA-seq dataset (Tasic et al., 2016, 2018). Ndnf coexpresses with Reelin in the superficial layer in the absence of MGE-lineage markers, suggesting that it marks neurogliaform cells. Indeed, Ndnf-cre-lineage INs are mostly present in layer I and have morphological and physiological features of neurogliaform CINs (Tasic et al., 2016, 2018).

Several studies aimed to uncover heterogeneity and diversity in GEs and along IN developmental trajectories using scRNA-seq (Chen et al., 2017; Mayer et al., 2018; Mi et al., 2018). One study profiled the transcriptomes from dMGE, vMGE, and CGE at E12.5 and E14.5 (Mi et al., 2018). Through semi-supervised analysis (clustering using researcher annotated gene list), temporal factor appeared to be the main driving force that segregates E12.5 and E14.5 VZ/SVZ progenitor cells. This finding implies that GE progenitors, especially in the SVZ, may have high dynamic turnovers that allow generation of different types of neurons. When cross-referencing conserved variable genes identified from adult IN subtypes (Tasic et al., 2016), at least 11 (out of 23) IN subclasses were identified at E12.5 and E14.5. Based upon the hypothesis that progenitors and newborn neurons share cardinal genes that determine lineages, potential PV and SST IN progenitors were identified and compared. Several markers were found enriched in the potential SST IN progenitors, including Epha5, Cdk14, and c-Maf (Mi et al., 2018). Nkx2.1-Cre generated c-Maf conditional deletion and c-Maf overexpression through MGE transplantation were adopted for validation. Loss of c-Maf in the Nkx2.1-Cre lineage resulted in SST+ CIN reduction in adulthood, while overt c-Maf expression led to a slight increase in the ratio of SST/PV INs, supporting the hypothesis that c-Maf may have roles in promoting SST fate (Mi et al., 2018). Based on a similar assumption that certain genes are conserved from precursor to mature INs, Mayer et al. conducted an integrated study of MGE/CGE cells (progenitors and immature neurons) from E13.5, E14.5, and Dlx6a-Cre enriched pan-GABAergic INs harvested at E18.5 and P10 (Mayer et al., 2018). Several candidate early markers were identified for SST progenitors (Sst, Tspan7, and Satb1) and PV progenitors (Mef2c, Erbb4, and Plcxd3). Encouragingly, Mef2c stands as a marker enriched in PV IN subtype across evolution, in rodents and in human (Habib et al., 2017; Mayer et al., 2018; Pai et al., 2020). Preferential loss of PV CINs was also observed in Dlx6a-Cre-generated conditional Mef2c mutant neocortices, reinforcing the essential role of Mef2c in PV IN production (Mayer et al., 2018).

The advancement of scRNA-seq technology has enhanced our knowledge of the temporal and spatial specification and classification of INs. However, many factors can confound the interpretation of these studies, which include but are not limited to sample handling, sequencing method, sequencing depth, and computational analysis algorithms. Therefore, in-depth molecular and cell biology studies should be pursued in parallel to validate these sequencing results.

The Mef2c finding of Mayer et al. (2018) was corroborated and extended by a study that surveyed the roles of Mafb and c-Maf in CIN development and maturation (Pai et al., 2020). Double deletion of Mafb and c-Maf led to reduced Mef2c expression in Nkx2.1-Cre- lineage CINs. Furthermore, both the Mafb/c-Maf double and Mef2c conditional mutants show preferential loss of PV CINs. When Mef2c was transduced into the Mafb/c-Maf double mutant MGE cells, PV expression was rescued, suggesting that Mafb/c-Maf and Mef2c may be in the same core transcriptional pathway that promotes PV IN fate (Mayer et al., 2018; Pai et al., 2019, 2020).

Contrary to the conclusion that c-Maf alone promotes SST fate (Mi et al., 2018), Pai et al. (2019) came to a different conclusion. c-Maf mutants did not show a reduction of SST CINs at E12.5, E15.5, and P0. Second, both c-Maf single and double deletion in the Nkx2.1-Cre lineage resulted in loss of MGE INs (both PV and SST) starting at P7. Thus, c-Maf function may be implicated in CIN survival and/or maturation, and it may not necessarily be required to promote specification of SST CINs (Mi et al., 2018; Pai et al., 2019, 2020).

Finally, studies using other species are required to study the transcriptomic features of CINs during vertebrate evolution (Ma et al., 2013; Boldog et al., 2018; Liu et al., 2021). This information will likely be of import in humans to help elucidate genetic contributions to neuropsychiatric disorders.

Conclusion

Since the discovery of the mammalian Dlx genes in 1990s (Porteus et al., 1991; Price et al., 1991; Robinson et al., 1991), major progress has been made in the genetic analysis of GE and IN development. This includes elucidating the roles of TFs, TF-RE interactions, and transcriptomic profiling in regional/cell-type specification and in distinct IN classes differentiation.

As technology continues to be developed, researchers are combining single-cell epigenomic analyses (e.g., ATAC seq), with single- cell RNA-seq, to link open-chromatin landscapes to cell type specificity and to identify enhancers for cell-type-specific targeting (Graybuck et al., 2021). However, many gaps remain. For instance, if TF-RE machinery that programs genetic expression is really the mechanism that drives progenitor fates, it will be crucial to visualize the TF-RE interaction in a lineage-specific manner, if not at single-cell resolution (currently, single-cell TF ChIP-seq is not feasible). The definition of lineage relies on proper cell type classification and marker discovery. However, a wide range of discrepancies still exists among high-throughput sequencing analyses and interpretation. Consensus on big data processing needs to be determined within the research community to make data interpretation more unified and easy to extrapolate. Furthermore, with many scRNA-seq studies performed on embryonic GEs, we still cannot fully solve the puzzle of whether GE progenitors use cell-cycle exit timing to diversify their descendants (Petros et al., 2015). Cell cycling status needs be integrated with temporal and spatial molecular mechanisms. Finally, a long-term goal of rodent studies is to understand human biology. While neural development (neurogenesis and cell type classification) of the mouse and human brain is highly conserved, there are some important differences (Paredes et al., 2016; Boldog et al., 2018; Pollen et al., 2019; Velmeshev et al., 2021). With the help of burgeoning technologies, such as iPSCs and organoids, progress will be made to understand the genetic and other mechanisms that control the development of human cortical interneurons in health and in disease, as exemplified by some prescient papers (Sun et al., 2016; Birey et al., 2017).

Disclosure Statement

J. L. R. is the cofounder, stockholder, and a current scientific board member of Neurona Therapeutics, a company that studies the potential translational application of interneuron transplantation.

Acknowledgments

This work was supported by the research grants to: JLR from NIMH R01-MH081880 and NIMH R01-MH049428; DV from Spectrum Health-Michigan State Alliance Corporation. Both figures were created with BioRender.com; Figure 47–2 was adapted from the “Regulation of Transcription in Eukaryotic Cells” template (2021). Author contribution: ELLP, DV, and JLR conceptualized, drafted, and finalized the article; JSH had partial contribution to the second and fifth sections of this chapter.

References

  1. Abs, E. et al. (2018) ‘Learning-Related Plasticity in Dendrite-Targeting Layer 1 Interneurons’, Neuron, 100(3), pp. 684–699. doi: 10.1016/j.neuron.2018.09.001. [PMC free article: PMC6226614] [PubMed: 30269988]
  2. Ahituv, N. et al. (2007) ‘Deletion of ultraconserved elements yields viable mice’, PLoS Biology, 5(9), p. e234. doi: 10.1371/journal.pbio.0050234. [PMC free article: PMC1964772] [PubMed: 17803355]
  3. Anderson, S. A. (1997a) ‘Interneuron Migration from Basal Forebrain to Neocortex: Dependence on Dlx Genes’, Science, 278(5337), pp. 474–476. doi: 10.1126/science.278.5337.474. [PubMed: 9334308]
  4. Anderson, Stewart A. et al. (1997b) ‘Mutations of the homeobox genes Dlx-1 and Dlx-2 disrupt the striatal subventricular zone and differentiation of late born striatal neurons’, Neuron, 19(1), pp. 27–37. doi: 10.1016/S0896-6273(00)80345-1. [PubMed: 9247261]
  5. Arguello, A. et al. (2013) ‘Dapper Antagonist of Catenin-1 Cooperates with Dishevelled-1 during Postsynaptic Development in Mouse Forebrain GABAergic Interneurons’, PLoS ONE, 8(6), p. e67679. doi: 10.1371/journal.pone.0067679. [PMC free article: PMC3691262] [PubMed: 23826333]
  6. Azim, E. et al. (2009) ‘SOX6 controls dorsal progenitor identity and interneuron diversity during neocortical development.’, Nature neuroscience, 12(10), pp. 1238–47. doi: 10.1038/nn.2387. [PMC free article: PMC2903203] [PubMed: 19657336]
  7. Batista-Brito, Renata, Ward Claire, F. G. (2020) Patterning and Cell Type specification in the Developing CNS and PNS (2nd Edition). Chapter 20- The generation of cortical interneurons. Available at: https://doi​.org/10.1016​/B978-0-12-814405-3.00020-5.
  8. Batista-Brito, R. et al. (2009) ‘The Cell-Intrinsic Requirement of Sox6 for Cortical Interneuron Development’, Neuron, 63(4), pp. 466–481. doi: 10.1016/j.neuron.2009.08.005. [PMC free article: PMC2773208] [PubMed: 19709629]
  9. Battiste, J. et al. (2007) ‘Ascl 1 defines sequentially generated lineage-resricted neuronal and oligodendrocyte precursor cells in the spinal cord’, Development, 134(2), pp. 283–293. doi: 10.1242/dev.02727. [PubMed: 17166924]
  10. Bejerano, G. et al. (2004) ‘Ultraconserved elements in the human genome’, Science, 304(5675), pp. 1321–1325. doi: 10.1126/science.1098119. [PubMed: 15131266]
  11. Betley, J. N. and Sternson, S. M. (2011) ‘Adeno-associated viral vectors for mapping, monitoring, and manipulating neural circuits’, Human Gene Therapy, 22(6), pp. 669–677. doi: 10.1089/hum.2010.204. [PMC free article: PMC3107581] [PubMed: 21319997]
  12. Birey, F. et al. (2017) ‘Assembly of functionally integrated human forebrain spheroids’, Nature, 545(7652), pp. 54–59. doi: 10.1038/nature22330. [PMC free article: PMC5805137] [PubMed: 28445465]
  13. Boldog, E. et al. (2018) ‘Transcriptomic and morphophysiological evidence for a specialized human cortical GABAergic cell type’, Nature Neuroscience, 21(9), pp. 1185–1195. doi: 10.1038/s41593-018-0205-2. [PMC free article: PMC6130849] [PubMed: 30150662]
  14. Brown, K. N. et al. (2011) ‘Clonal production and organization of inhibitory interneurons in the neocortex’, Science, 334(6055), pp. 480–486. doi: 10.1126/science.1208884. [PMC free article: PMC3304494] [PubMed: 22034427]
  15. Bupesh, M., Abellán, A. and Medina, L. (2011) ‘Genetic and experimental evidence supports the continuum of the central extended amygdala and a mutiple embryonic origin of its principal neurons’, Journal of Comparative Neurology, 519(17), pp. 3507–3531. doi: 10.1002/cne.22719. [PubMed: 21800302]
  16. Butt, S. J. B. et al. (2005) ‘The temporal and spatial origins of cortical interneurons predict their physiological subtype’, Neuron, 48(4), pp. 591–604. doi: 10.1016/j.neuron.2005.09.034. [PubMed: 16301176]
  17. Butt, S. J. B. et al. (2008) ‘The Requirement of Nkx2-1 in the Temporal Specification of Cortical Interneuron Subtypes’, Neuron, 59(5), pp. 722–732. doi: 10.1016/j.neuron.2008.07.031. [PMC free article: PMC2562525] [PubMed: 18786356]
  18. Cai, Y. et al. (2013) ‘Nuclear receptor COUP-TFII-expressing neocortical interneurons are derived from the medial and lateral/caudal ganglionic eminence and define specific subsets of mature interneurons’, Journal of Comparative Neurology, 521(2), pp. 479–497. doi: 10.1002/cne.23186. [PubMed: 22791192]
  19. Chao, H.-T. et al. (2010) ‘Dysfunction in GABA signalling mediates autism-like stereotypies and Rett syndrome phenotypes’, Nature, 468(7321), pp. 263–269. doi: 10.1038/nature09582. [PMC free article: PMC3057962] [PubMed: 21068835]
  20. Chapman, H. et al. (2012) ‘The homeobox gene Gsx2 controls the timing of oligodendroglial fate specification in mouse lateral ganglionic eminence progenitors’, Development (Cambridge), 140(11), pp. 2289–2298. doi: 10.1242/dev.091090. [PMC free article: PMC3653554] [PubMed: 23637331]
  21. Chapman, H. et al. (2018) ‘Gsx transcription factors control neuronal versus glial specification in ventricular zone progenitors of the mouse lateral ganglionic eminence’, Developmental Biology, 442(1), pp. 115–126. doi: 10.1016/j.ydbio.2018.07.005. [PMC free article: PMC6158017] [PubMed: 29990475]
  22. Chen, Y.-J. J. et al. (2017) ‘Single-cell RNA sequencing identifies distinct mouse medial ganglionic eminence cell types’, Scientific Reports, 7, p. 45656. doi: 10.1038/srep45656. [PMC free article: PMC5374502] [PubMed: 28361918]
  23. Chen, Y. J. J. et al. (2013) ‘Use of “MGE Enhancers” for Labeling and Selection of Embryonic Stem Cell-Derived Medial Ganglionic Eminence (MGE) Progenitors and Neurons’,  PLoS ONE, 8(5), p. e61956. doi: 10.1371/journal.pone.0061956. [PMC free article: PMC3641041] [PubMed: 23658702]
  24. Chiang, C. W. K. et al. (2008) ‘Ultraconserved elements: Analyses of dosage sensitivity, motifs and boundaries’, Genetics, 180(4), pp. 2277–2293. doi: 10.1534/genetics.108.096537. [PMC free article: PMC2600958] [PubMed: 18957701]
  25. Cho, K. K. A. et al. (2015) ‘Gamma rhythms link prefrontal interneuron dysfunction with cognitive inflexibility in dlx5/6+/- mice’, Neuron, 85(6), pp. 1332–1343. doi: 10.1016/j.neuron.2015.02.019. [PMC free article: PMC4503262] [PubMed: 25754826]
  26. Choi, J. H. et al. (2014) ‘Optimization of AAV expression cassettes to improve packaging capacity and transgene expression in neurons’, Molecular Brain, 7(1), p. 17. doi: 10.1186/1756-6606-7-17. [PMC free article: PMC3975461] [PubMed: 24618276]
  27. Close, J. et al. (2012) ‘Satb1 is an activity-modulated transcription factor required for the terminal differentiation and connectivity of medial ganglionic eminence-derived cortical interneurons.’, The Journal of neuroscience : the official journal of the Society for Neuroscience, 32(49), pp. 17690–705. [PMC free article: PMC3654406] [PubMed: 23223290]
  28. Cobos, I. et al. (2005) ‘Mice lacking Dlx1 show subtype-specific loss of interneurons, reduced inhibition and epilepsy.’, Nature neuroscience, 8(8), pp. 1059–68. doi: 10.1038/nn1499. [PubMed: 16007083]
  29. Cobos, I., Borello, U. and Rubenstein, J. L R (2007) ‘Dlx Transcription Factors Promote Migration through Repression of Axon and Dendrite Growth’, Neuron, 54(6), pp. 873–888. doi: 10.1016/j.neuron.2007.05.024. [PMC free article: PMC4921237] [PubMed: 17582329]
  30. Cobos, I., Broccoli, V. and Rubenstein, J. L. R. (2005) ‘The vertebrate ortholog of Aristaless is regulated by Dlx genes in the developing forebrain’, Journal of Comparative Neurology, 483(3), pp. 292–303. doi: 10.1002/cne.20405. [PubMed: 15682394]
  31. Colasante, G. et al. (2008) ‘Arx is a direct target of Dlx2 and thereby contributes to the tangential migration of GABAergic interneurons’, Journal of Neuroscience, 28(42), pp. 10674–10686. doi: 10.1523/JNEUROSCI.1283-08.2008. [PMC free article: PMC3844830] [PubMed: 18923043]
  32. Colasante, G. et al. (2009) ‘Arx acts as a regional key selector gene in the ventral telencephalon mainly through its transcriptional repression activity’, Developmental Biology, 334(1), pp. 59–71. doi: 10.1016/j.ydbio.2009.07.014. [PubMed: 19627984]
  33. Colom, L. V. (2006) ‘Septal networks: Relevance to theta rhythm, epilepsy and Alzheimer’s disease’, Journal of Neurochemistry, 96(3), pp. 609–623 doi: 10.1111/j.1471-4159.2005.03630.x. [PubMed: 16405497]
  34. Colombo, E. et al. (2007) ‘Inactivation of Arx, the Murine Ortholog of the X-Linked Lissencephaly with Ambiguous Genitalia Gene, Leads to Severe Disorganization of the Ventral Telencephalon with Impaired Neuronal Migration and Differentiation’, Journal of Neuroscience, 27(17), pp. 4786–4798. doi: 10.1523/JNEUROSCI.0417-07.2007. [PMC free article: PMC4916654] [PubMed: 17460091]
  35. Cooper, G. M. et al. (2011) ‘A copy number variation morbidity map of developmental delay’, Nature Genetics, 43(9), pp. 838–846. doi: 10.1038/ng.909. [PMC free article: PMC3171215] [PubMed: 21841781]
  36. Corbin, J. G. et al. (2000) ‘The Gsh2 homeodomain gene controls multiple aspects of telencephalic development’, Development, 127(23), pp. 5007–5020. doi: 10.1242/dev.127.23.5007. [PubMed: 11060228]
  37. Corbin, J. G. et al. (2003) ‘Combinatorial function of the homeodomain proteins Nkx2.1 and Gsh2 in ventral telencephalic patterning’, Development, 130(20), pp. 4898–4906. doi: 10.1242/dev.00717. [PubMed: 12930780]
  38. Cossart, R. (2011) ‘The maturation of cortical interneuron diversity: How multiple developmental journeys shape the emergence of proper network function’, Current Opinion in Neurobiology, 21(1), pp. 160–168. doi: 10.1016/j.conb.2010.10.003. [PubMed: 21074988]
  39. Crossley, P. H. et al. (2001) ‘Coordinate expression of Fgf8, Otx2, Bmp4, and Shh in the rostral prosencephalon during development of the telencephalic and optic vesicles’, Neuroscience, 108(2), pp. 183–206. doi: 10.1016/S0306-4522(01)00411-0. [PubMed: 11734354]
  40. Deacon, T. W., Pakzaban, P. and Isacson, O. (1994) ‘The lateral ganglionic eminence is the origin of cells committed to striatal phenotypes: neural transplantation and developmental evidence’, Brain Research, 668(1–2), pp. 211–219. doi: 10.1016/0006-8993(94)90526-6. [PubMed: 7704606]
  41. DeFelipe, J. et al. (2013) ‘New insights into the classification and nomenclature of cortical GABAergic interneurons.’, Nature reviews. Neuroscience, 14(3), pp. 202–16. doi: 10.1038/nrn3444. [PMC free article: PMC3619199] [PubMed: 23385869]
  42. Denaxa, M. et al. (2012) ‘Maturation-Promoting Activity of SATB1 in MGE-Derived Cortical Interneurons’, Cell Reports, 2(5), pp. 1351–1362. doi: 10.1016/j.celrep.2012.10.003. [PMC free article: PMC3607226] [PubMed: 23142661]
  43. Depew, M. J. et al. (1999) ‘Dlx5 regulates regional development of the branchial arches and sensory capsules’, Development, 126(17), pp. 3831–3846. doi: 10.1242/dev.126.17.3831. [PubMed: 10433912]
  44. Depew, M. J. et al. (2005) ‘Reassessing the Dlx code: The genetic regulation of branchial arch skeletal pattern and development’, Journal of Anatomy, 207(5), pp. 501–561 doi: 10.1111/j.1469-7580.2005.00487.x. [PMC free article: PMC1571560] [PubMed: 16313391]
  45. Depew, M. J., Lufkin, T. and Rubenstein, J. L. R. (2002) ‘Specification of jaw subdivisions by Dlx genes’, Science, 298(5592), pp. 381–385. doi: 10.1126/science.1075703. [PubMed: 12193642]
  46. Dickel, D. E. et al. (2018) ‘Ultraconserved Enhancers Are Required for Normal Development’, Cell, 172(3), pp. 491–499.e15. doi: 10.1016/j.cell.2017.12.017. [PMC free article: PMC5786478] [PubMed: 29358049]
  47. Dimidschstein, J. et al. (2016) ‘A viral strategy for targeting and manipulating interneurons across vertebrate species’, Nature Neuroscience, 19(12), pp. 1743–1749 doi: 10.1038/nn.4430. [PMC free article: PMC5348112] [PubMed: 27798629]
  48. Drever, B. D., Riedel, G. and Platt, B. (2011) ‘The cholinergic system and hippocampal plasticity’, Behavioural Brain Research, 221(2), pp. 505–514. doi: 10.1016/j.bbr.2010.11.037. [PubMed: 21130117]
  49. Du, T. et al. (2008) ‘NKX2.1 specifies cortical interneuron fate by activating Lhx6’, Development, 135(8), pp. 1559–1567. doi: 10.1242/dev.015123. [PubMed: 18339674]
  50. Eisenstat, D. D. et al. (1999) ‘DLX-1, DLX-2, and DLX-5 expression define distinct stages of basal forebrain differentiation’, Journal of Comparative Neurology, 414(2), pp. 217–237. doi: 10.1002/(SICI)1096-9861(19991115)414:2<217::AID-CNE6>3.0.CO;2-I. [PubMed: 10516593]
  51. Elbert, A. et al. (2019) ‘CTCF governs the identity and migration of MGE-derived cortical interneurons’, Journal of Neuroscience, 39(1), pp. 177–192. doi: 10.1523/JNEUROSCI.3496-17.2018. [PMC free article: PMC6331652] [PubMed: 30377227]
  52. Erbel-Sieler, C. et al. (2004) ‘Behavioral and regulatory abnormalities in mice deficient in the NPAS1 and NPAS3 transcription factors’, Proceedings of the National Academy of Sciences of the United States of America, 101(37), pp. 13648–13653. doi: 10.1073/pnas.0405310101. [PMC free article: PMC518807] [PubMed: 15347806]
  53. Fazel Darbandi, S. et al. (2016) ‘Functional consequences of I56ii Dlx enhancer deletion in the developing mouse forebrain’, Developmental Biology, 420(1), S0012–1606(16), pp. 30263–30269. doi: 10.1016/j.ydbio.2016.10.015. [PubMed: 27983964]
  54. Flames, N. et al. (2007) ‘Delineation of multiple subpallial progenitor domains by the combinatorial expression of transcriptional codes’, Journal of Neuroscience, 27(36), pp. 9682–9695. doi: 10.1523/JNEUROSCI.2750-07.2007. [PMC free article: PMC4916652] [PubMed: 17804629]
  55. Flandin, P., Zhao, Y., Vogt, D., Jeong, J., Long, J., Potter, G., Westphal, H. and Rubenstein, John L.R. (2011) ‘Lhx6 and Lhx8 Coordinately Induce Neuronal Expression of Shh that Controls the Generation of Interneuron Progenitors’, Neuron, 70(5), pp. 939–950. doi: 10.1016/j.neuron.2011.04.020. [PMC free article: PMC3153409] [PubMed: 21658586]
  56. Flandin, P., Kimura, S. and Rubenstein, J. L. R. (2010) ‘The progenitor zone of the ventral medial ganglionic eminence requires Nkx2-1 to generate most of the globus pallidus but few neocortical interneurons’, Journal of Neuroscience, 30(8), pp. 2018–2823. doi: 10.1523/JNEUROSCI.4228-09.2010. [PMC free article: PMC2865856] [PubMed: 20181579]
  57. Fode, C. et al. (2000) ‘A role for neural determination genes in specifying the dorsoventral identity of telencephalic neurons’, Genes and Development, 14(1), pp. 67–80. doi: 10.1101/gad.14.1.67. [PMC free article: PMC316337] [PubMed: 10640277]
  58. Fogarty, M. et al. (2007) ‘Spatial genetic patterning of the embryonic neuroepithelium generates GABAergic interneuron diversity in the adult cortex’, Journal of Neuroscience, 27(41), pp. 10935–10946. doi: 10.1523/JNEUROSCI.1629-07.2007. [PMC free article: PMC6672847] [PubMed: 17928435]
  59. Fogarty, M., Richardson, W. D. and Kessaris, N. (2005) ‘A subset of oligodendrocytes generated from radial glia in the dorsal spinal cord’, Development, 132(8), pp. 1951–1959. doi: 10.1242/dev.01777. [PubMed: 15790969]
  60. Fragkouli, A. et al. (2009) ‘LIM homeodomain transcription factor-dependent specification of bipotential MGE progenitors into cholinergic and GABAergic striatal interneurons’, Development, 136(22), pp. 3841–3851. doi: 10.1242/dev.038083. [PMC free article: PMC2766344] [PubMed: 19855026]
  61. Frankel, N. et al. (2010) ‘Phenotypic robustness conferred by apparently redundant transcriptional enhancers’, Nature, 466(7305), pp. 490–493. doi: 10.1038/nature09158. [PMC free article: PMC2909378] [PubMed: 20512118]
  62. Friocourt, G. et al. (2006) ‘The role of ARX in cortical development’, European Journal of Neuroscience, 23(4), pp. 869–876. doi: 10.1111/j.1460-9568.2006.04629.x. [PubMed: 16519652]
  63. Friocourt, G. and Parnavelas, J. G. (2010) ‘Mutations in ARX result in several defects involving GABAergic neurons’, Frontiers in Cellular Neuroscience, 4. P. 4. doi: 10.3389/fncel.2010.00004. [PMC free article: PMC2841486] [PubMed: 20300201]
  64. Fuccillo, M. et al. (2004) ‘Temporal requirement for hedgehog signaling in ventral telencephalic patterning’, Development, 131(20), pp. 5031–5040. doi: 10.1242/dev.01349. [PubMed: 15371303]
  65. Fulp, C. T. et al. (2008) ‘Identification of Arx transcriptional targets in the developing basal forebrain’, Human Molecular Genetics, 17(23), pp. 3740–3760. doi: 10.1093/hmg/ddn271. [PMC free article: PMC2581427] [PubMed: 18799476]
  66. Gelman, D. et al. (2011) ‘A wide diversity of cortical GABAergic interneurons derives from the embryonic preoptic area’, J Neurosci, 31(46), pp. 16570–16580. doi: 10.1523/JNEUROSCI.4068-11.2011. [PMC free article: PMC6633309] [PubMed: 22090484]
  67. Gelman, D. M. et al. (2009) ‘The embryonic preoptic area is a novel source of cortical GABAergic interneurons’, Journal of Neuroscience, 29(29), pp. 9380–9389. doi: 10.1523/JNEUROSCI.0604-09.2009. [PMC free article: PMC6665570] [PubMed: 19625528]
  68. Ghanem, N. et al. (2007) ‘Distinct cis-regulatory elements from the Dlx1/Dlx2 locus mark different progenitor cell populations in the ganglionic eminences and different subtypes of adult cortical interneurons’, Journal of Neuroscience, 27(19), pp. 5012–5022. doi: 10.1523/JNEUROSCI.4725-06.2007. [PMC free article: PMC4917363] [PubMed: 17494687]
  69. Glickstein, S. B. et al. (2007) ‘Selective cortical interneuron and GABA deficits in cyclin D2-null mice’, Development, 134(22), pp. 4083–4093. doi: 10.1242/dev.008524. [PMC free article: PMC3396210] [PubMed: 17965053]
  70. Glickstein, S. B. et al. (2009) ‘Cyclin D2 is critical for intermediate progenitor cell proliferation in the embryonic cortex.’, The Journal of neuroscience : the official journal of the Society for Neuroscience, 29(30), pp. 9614–24. doi: 10.1523/JNEUROSCI.2284-09.2009. [PMC free article: PMC2811167] [PubMed: 19641124]
  71. Gorski, J. A. et al. (2002) ‘Cortical excitatory neurons and glia, but not GABAergic neurons, are produced in the Emx1-expressing lineage’, Journal of Neuroscience, 22(15), pp. 6309–6314. doi: 10.1523/jneurosci.22-15-06309.2002. [PMC free article: PMC6758181] [PubMed: 12151506]
  72. Graybuck, L. T. et al. (2021) ‘Enhancer viruses for combinatorial cell-subclass-specific labeling’, Neuron, 109(9), pp. 1449–1464.e13. doi: 10.1016/j.neuron.2021.03.011. [PMC free article: PMC8610077] [PubMed: 33789083]
  73. Greig, L. C. et al. (2013) ‘Molecular logic of neocortical projection neuron specification, development and diversity’, Nature Reviews Neuroscience, 14(11), pp. 755–769. doi: 10.1038/nrn3586. [PMC free article: PMC3876965] [PubMed: 24105342]
  74. Guo, J. and Anton, E. S. (2014) ‘Decision making during interneuron migration in the developing cerebral cortex’, Trends in Cell Biology, 24(6), pp. 342–351. doi: 10.1016/j.tcb.2013.12.001. [PMC free article: PMC4299592] [PubMed: 24388877]
  75. Guo, T. et al. (2019) ‘Dlx1/2 are Central and Essential Components in the Transcriptional Code for Generating Olfactory Bulb Interneurons’, Cerebral Cortex, 29(11), pp. 4831–4849. doi: 10.1093/cercor/bhz018. [PMC free article: PMC6917526] [PubMed: 30796806]
  76. Gutin, G. et al. (2006) ‘FGF signalling generates ventral telencephalic cells independently of SHH’, Development, 133(15), pp. 2937–2946. doi: 10.1242/dev.02465. [PubMed: 16818446]
  77. Habib, N. et al. (2017) ‘Massively parallel single-nucleus RNA-seq with DroNc-seq’, Nature Methods, 14(10), pp. 955–958. doi: 10.1038/nmeth.4407. [PMC free article: PMC5623139] [PubMed: 28846088]
  78. Han, S. et al. (2012) ‘Autistic-like behaviour in Scn1a+/− mice and rescue by enhanced GABA-mediated neurotransmission’, Nature, 489(7416), pp. 385–390. doi: 10.1038/nature11356. [PMC free article: PMC3448848] [PubMed: 22914087]
  79. Harrington, A. J. et al. (2016) ‘MEF2C regulates cortical inhibitory and excitatory synapses and behaviors relevant to neurodevelopmental disorders’, eLife, 5, p. e20059. doi: 10.7554/elife.20059. [PMC free article: PMC5094851] [PubMed: 27779093]
  80. Harris, K. D. et al. (2018) ‘Classes and continua of hippocampal CA1 inhibitory neurons revealed by single-cell transcriptomics’, PLoS Biology, 16(6), p. e2006787. doi: 10.1371/journal.pbio.2006387. [PMC free article: PMC6029811] [PubMed: 29912866]
  81. Harwell, C. C. et al. (2015) ‘Wide Dispersion and Diversity of Clonally Related Inhibitory Interneurons’, Neuron, 87(5), pp. 999–1007. doi: 10.1016/j.neuron.2015.07.030. [PMC free article: PMC4581718] [PubMed: 26299474]
  82. Hippenmeyer, S. et al. (2005) ‘A developmental switch in the response of DRG neurons to ETS transcription factor signaling’, PLoS Biology, 3(5), p. e159. doi: 10.1371/journal.pbio.0030159. [PMC free article: PMC1084331] [PubMed: 15836427]
  83. Hirata, T. et al. (2009) ‘Identification of distinct telencephalic progenitor pools for neuronal diversity in the amygdala’, Nature Neuroscience, 12(2), pp. 141–149. doi: 10.1038/nn.2241. [PMC free article: PMC2747779] [PubMed: 19136974]
  84. Hoch, R. V. et al. (2015a) ‘OTX2 Transcription Factor Controls Regional Patterning within the Medial Ganglionic Eminence and Regional Identity of the Septum’, Cell Reports, 12(3), pp. 482–494. doi: 10.1016/j.celrep.2015.06.043. [PMC free article: PMC4512845] [PubMed: 26166575]
  85. Hoch, R. V., Clarke, J. A. and Rubenstein, J. L. R. (2015b) ‘Fgf signaling controls the telencephalic distribution of Fgf-expressing progenitors generated in the rostral patterning center’, Neural Development, 10(1), p. 8. doi: 10.1186/s13064-015-0037-7. [PMC free article: PMC4416298] [PubMed: 25889070]
  86. Hoch, R. V., Rubenstein, J. L. R. and Pleasure, S. (2009) ‘Genes and signaling events that establish regional patterning of the mammalian forebrain’, Seminars in Cell and Developmental Biology, 20(4), pp. 378–386. doi: 10.1016/j.semcdb.2009.02.005. [PubMed: 19560042]
  87. Howard, M. A., Rubenstein, J. L. R. and Baraban, S. C. (2014) ‘Bidirectional homeostatic plasticity induced by interneuron cell death and transplantation in vivo’, Proceedings of the National Academy of Sciences of the United States of America, 111(1), pp. 492–497. doi: 10.1073/pnas.1307784111. [PMC free article: PMC3890856] [PubMed: 24344303]
  88. Hsiao, C. J. et al. (2020) ‘Characterizing and inferring quantitative cell cycle phase in single-cell RNA-seq data analysis’, Genome Research, 30(4), pp. 611–621. doi: 10.1101/gr.247759.118. [PMC free article: PMC7197478] [PubMed: 32312741]
  89. Hu, J. S., Vogt, D., Lindtner, S., Sandberg, M., Silberbergand, S. N., et al. (2017a) ‘Coup-TF1&2 (Nr2f1 and Nr2f2) control subtype and laminar identity of MGE- derived neocortical interneurons’,  Development, 144(15), pp. 2837–2851. doi: 10.1242/dev.150664. [PMC free article: PMC5560044] [PubMed: 28694260]
  90. Hu, J. S., Vogt, D., Sandberg, M., et al. (2017b) ‘Cortical interneuron development: a tale of time and space’, Development, 144(21), pp. 3867–3878. doi: 10.1242/dev.132852. [PMC free article: PMC5702067] [PubMed: 29089360]
  91. Huang, Z. J., Di Cristo, G. and Ango, F. (2007) ‘Development of GABA innervation in the cerebral and cerebellar cortices’, Nature Reviews Neuroscience, 8(9), pp. 673–686. doi: 10.1038/nrn2188. [PubMed: 17704810]
  92. Inoue, T. et al. (2007) ‘Zic1 and Zic3 regulate medial forebrain development through expansion of neuronal progenitors’, Journal of Neuroscience, 27(20), pp. 5461–5473. doi: 10.1523/JNEUROSCI.4046-06.2007. [PMC free article: PMC6672357] [PubMed: 17507568]
  93. Jeong, J. et al. (2008) ‘Dlx genes pattern mammalian jaw primordium by regulating both lower jaw-specific and upper jaw-specific genetic programs’, Development, 135(17), pp. 2905–2916. doi: 10.1242/dev.019778. [PMC free article: PMC4913551] [PubMed: 18697905]
  94. Jeong, J. et al. (2012) ‘Cleft palate defect of Dlx1/2-/- mutant mice is caused by lack of vertical outgrowth in the posterior palate’, Developmental Dynamics, 241(11), pp. 1757–1769. doi: 10.1002/dvdy.23867. [PMC free article: PMC3988582] [PubMed: 22972697]
  95. Jones, D. L. et al. (2011) ‘Deletion of Dlx1 results in reduced glutamatergic input to hippocampal interneurons’, Journal of Neurophysiology, 105(5), pp. 1984–1991. doi: 10.1152/jn.00056.2011. [PMC free article: PMC3094166] [PubMed: 21325686]
  96. Joyner, A. L. and Zervas, M. (2006) ‘Genetic inducible fate mapping in mouse: Establishing genetic lineages and defining genetic neuroanatomy in the nervous system’, Developmental Dynamics, 235(9), pp. 2376–2385. doi: 10.1002/dvdy.20884. [PubMed: 16871622]
  97. Kaltezioti, V. et al. (2010) ‘Prox1 regulates the Notch1-mediated inhibition of neurogenesis’, PLoS Biology, 8(12), p. e1000565. doi: 10.1371/journal.pbio.1000565. [PMC free article: PMC3006385] [PubMed: 21203589]
  98. Kanatani, S. et al. (2008) ‘COUP-TFII is preferentially expressed in the caudal ganglionic eminence and is involved in the caudal migratory stream’, Journal of Neuroscience, 28(50), pp. 13582–13591. doi: 10.1523/JNEUROSCI.2132-08.2008. [PMC free article: PMC6671763] [PubMed: 19074032]
  99. Kepecs, A. and Fishell, G. (2014) ‘Interneuron cell types are fit to function’, Nature, 505(7483), pp. 318–326. [PMC free article: PMC4349583] [PubMed: 24429630]
  100. Kessaris, N. et al. (2006) ‘Competing waves of oligodendrocytes in the forebrain and postnatal elimination of an embryonic lineage’, Nature Neuroscience, 9(2), pp. 173–179. doi: 10.1038/nn1620. [PMC free article: PMC6328015] [PubMed: 16388308]
  101. Kessaris, N. et al. (2014) ‘Genetic programs controlling cortical interneuron fate’, Current Opinion in Neurobiology, 26(100), pp. 79–87. doi: 10.1016/j.conb.2013.12.012. [PMC free article: PMC4082532] [PubMed: 24440413]
  102. Kimmel, R. A. et al. (2000) ‘Two lineage boundaries coordinate vertebrate apical ectodermal ridge formation’, Genes and Development, 14(11), pp. 1377–1389. doi: 10.1101/gad.14.11.1377. [PMC free article: PMC316660] [PubMed: 10837030]
  103. Kitamura, K. et al. (2002) ‘Mutation of ARX causes abnormal development of forebrain and testes in mice and X-linked lissencephaly with abnormal genitalia in humans’, Nature Genetics, 32(3), pp. 359–369. doi: 10.1038/ng1009. [PubMed: 12379852]
  104. Klausberger, T. et al. (2003) ‘Brain-state- and cell-type-specific firing of hippocampal interneurons in vivo’, Nature, 421(6925), pp. 844–848. doi: 10.1038/nature01374. [PubMed: 12594513]
  105. Kobayashi, T. and Kageyama, R. (2014) ‘Expression dynamics and functions of hes factors in development and diseases’,  Current Topics in Developmental Biology, 110, pp. 263–283. doi: 10.1016/B978-0-12-405943-6.00007-5. [PubMed: 25248479]
  106. Kohwi, M. et al. (2007) ‘A subpopulation of olfactory bulb GABAergic interneurons is derived from Emx1- and Dlx5/6-expressing progenitors’, Journal of Neuroscience, 27(26), pp. 6878–6891. doi: 10.1523/JNEUROSCI.0254-07.2007. [PMC free article: PMC4917362] [PubMed: 17596436]
  107. Kowalczyk, M. S. et al. (2015) ‘Single-cell RNA-seq reveals changes in cell cycle and differentiation programs upon aging of hematopoietic stem cells’, Genome Research, 25(12), pp. 1860–1872. doi: 10.1101/gr.192237.115. [PMC free article: PMC4665007] [PubMed: 26430063]
  108. Kriegstein, A. and Alvarez-Buylla, A. (2009) ‘The glial nature of embryonic and adult neural stem cells’, Annual Review of Neuroscience, 32, pp. 149–184. doi: 10.1146/annurev.neuro.051508.135600. [PMC free article: PMC3086722] [PubMed: 19555289]
  109. Le, T. N. et al. (2017) ‘GABAergic interneuron differentiation in the basal forebrain is mediated through direct regulation of glutamic acid decarboxylase isoforms by Dlx homeobox transcription factors’, Journal of Neuroscience, 37(36), pp. 8816–8829. doi: 10.1523/JNEUROSCI.2125-16.2017. [PMC free article: PMC6596671] [PubMed: 28821666]
  110. Lee, A. T., Vogt, D., et al. (2014) ‘A class of GABAergic neurons in the prefrontal cortex sends long-range projections to the nucleus accumbens and elicits acute avoidance behavior’, Journal of Neuroscience, 34(35), pp. 11519–11525. doi: 10.1523/JNEUROSCI.1157-14.2014. [PMC free article: PMC4145166] [PubMed: 25164650]
  111. Lee, A. T., Gee, S. M., et al. (2014) ‘Pyramidal neurons in prefrontal cortex receive subtype-specific forms of excitation and inhibition’, Neuron, 81(1), pp. 61–68. doi: 10.1016/j.neuron.2013.10.031. [PMC free article: PMC3947199] [PubMed: 24361076]
  112. Lee, S. H. et al. (2010) ‘The largest group of superficial neocortical GABAergic interneurons expresses ionotropic serotonin receptors’, Journal of Neuroscience, 30(50), pp. 16796–16808. doi: 10.1523/JNEUROSCI.1869-10.2010. [PMC free article: PMC3025500] [PubMed: 21159951]
  113. Li, H. et al. (2008) ‘Transcription factor MEF2C influences neural stem/progenitor cell differentiation and maturation in vivo’, Proceedings of the National Academy of Sciences of the United States of America, 105(27), pp. 9197–9402. doi: 10.1073/pnas.0802876105. [PMC free article: PMC2453715] [PubMed: 18599437]
  114. Li, J. et al. (2018) ‘Transcription factors Sp8 and Sp9 coordinately regulate olfactory bulb interneuron development’, Cerebral Cortex, 28(9), pp. 3278–3294. doi: 10.1093/cercor/bhx199. [PMC free article: PMC6095218] [PubMed: 28981617]
  115. Lim, L., Mi, D., et al. (2018) ‘Development and Functional Diversification of Cortical Interneurons’, Neuron, 100(2), pp. 294–313. doi: 10.1016/j.neuron.2018.10.009. [PMC free article: PMC6290988] [PubMed: 30359598]
  116. Lim, L., Pakan, J. M. P., et al. (2018) ‘Optimization of interneuron function by direct coupling of cell migration and axonal targeting’, Nature Neuroscience, 21(7), pp. 920–931. doi: 10.1038/s41593-018-0162-9. [PMC free article: PMC6061935] [PubMed: 29915195]
  117. Lindtner, S. et al. (2019) ‘Genomic Resolution of DLX-Orchestrated Transcriptional Circuits Driving Development of Forebrain GABAergic Neurons’, Cell Reports, 28(8), pp. 2048–2063.e8. doi: 10.1016/j.celrep.2019.07.022. [PMC free article: PMC6750766] [PubMed: 31433982]
  118. Liodis, P. et al. (2007) ‘Lhx6 Activity Is Required for the Normal Migration and Specification of Cortical Interneuron Subtypes’, Journal of Neuroscience, 27(12), pp. 3078–3089. doi: 10.1523/JNEUROSCI.3055-06.2007. [PMC free article: PMC6672459] [PubMed: 17376969]
  119. Litingtung, Y. and Chiang, C. (2000) ‘Specification of ventral neuron types is mediated by an antagonistic interaction between Shh and Gli3’, Nature Neuroscience, 3(10), pp. 979–985. doi: 10.1038/79916. [PubMed: 11017169]
  120. Liu, T. et al. (2021) ‘Cross-species single-cell transcriptomic analysis reveals pre-gastrulation developmental differences among pigs, monkeys, and humans’, Cell Discovery, 7(1), p. 8. doi: 10.1038/s41421-020-00238-x. [PMC free article: PMC7854681] [PubMed: 33531465]
  121. Lodato, S. et al. (2011) ‘Loss of COUP-TFI alters the balance between caudal ganglionic eminence- and medial ganglionic eminence-derived cortical interneurons and results in resistance to epilepsy’, Journal of Neuroscience, 31(12), pp. 4650–4662. doi: 10.1523/JNEUROSCI.6580-10.2011. [PMC free article: PMC6622915] [PubMed: 21430164]
  122. Long, H. K., Prescott, S. L., & Wysocka, J. (2016).‘Ever-Changing Landscapes: Transcriptional Enhancers in Development and Evolution’, Cell, 167 (5), pp. 1170–1187. doi: 10.1016/j.cell.2016.09.018 [PMC free article: PMC5123704] [PubMed: 27863239]
  123. Long, J. E. et al. (2007) ‘Dlx-dependent and -independent regulation of olfactory bulb interneuron differentiation’, Journal of Neuroscience, 27(12), pp. 3230–3243. doi: 10.1523/JNEUROSCI.5265-06.2007. [PMC free article: PMC4922751] [PubMed: 17376983]
  124. Long, Jason E., Swan, C., et al. (2009a) ‘Dlx1&2 and Mash1 transcription factors control striatal patterning and differentiation through parallel and overlapping pathways’, Journal of Comparative Neurology, 512(4), pp. 556–572. doi: 10.1002/cne.21854. [PMC free article: PMC2761428] [PubMed: 19030180]
  125. Long, Jason E., Cobos, I., et al. (2009b) ‘Dlx1&2 and Mash1 transcription factors control MGE and CGE patterning and differentiation through parallel and overlapping pathways’, Cerebral Cortex, suppl 1(suppl1), i96–106. doi: 10.1093/cercor/bhp045. [PMC free article: PMC2693539] [PubMed: 19386638]
  126. Luo, L. et al. (2020) ‘Optimizing Nervous System-Specific Gene Targeting with Cre Driver Lines: Prevalence of Germline Recombination and Influencing Factors’, Neuron, 106(1), pp. 36–65.e5. doi: 10.1016/j.neuron.2020.01.008. [PMC free article: PMC7377387] [PubMed: 32027825]
  127. Ma, T. et al. (2012) ‘A subpopulation of dorsal lateral/caudal ganglionic eminence-derived neocortical interneurons expresses the transcription factor Sp8’, Cerebral Cortex, 22(9), pp. 2120–2130. doi: 10.1093/cercor/bhr296. [PubMed: 22021915]
  128. Ma, T. et al. (2013) ‘Subcortical origins of human and monkey neocortical interneurons’, Nature Neuroscience, 16(11), pp. 1588–1597. doi: 10.1038/nn.3536. [PubMed: 24097041]
  129. Magno, L. et al. (2017) ‘NKX2-1 Is Required in the Embryonic Septum for Cholinergic System Development, Learning, and Memory’, Cell Reports, 20(7), pp. 1572–1584. doi: 10.1016/j.celrep.2017.07.053. [PMC free article: PMC5565637] [PubMed: 28813670]
  130. Malik, R. et al. (2019) ‘Tsc1 represses parvalbumin expression and fast-spiking properties in somatostatin lineage cortical interneurons’, Nature Communications, 10(1), p. 4994. doi: 10.1038/s41467-019-12962-4. [PMC free article: PMC6825152] [PubMed: 31676823]
  131. La Manno, G. et al. (2018) ‘RNA velocity of single cells’, Nature, 560(7719), pp. 494–498. doi: 10.1038/s41586-018-0414-6. [PMC free article: PMC6130801] [PubMed: 30089906]
  132. Marín, O. et al. (2001) ‘Sorting of striatal and cortical interneurons regulated by semaphorin-neuropilin interactions’, Science, 293(5531), pp. 872–875. doi: 10.1126/science.1061891. [PubMed: 11486090]
  133. Marín, O., Anderson, S. A. and Rubenstein, J. L. R. (2000) ‘Origin and molecular specification of striatal interneurons’, Journal of Neuroscience, 20(16), pp. 6063–6076. doi: 10.1523/jneurosci.20-16-06063.2000. [PMC free article: PMC6772576] [PubMed: 10934256]
  134. Marsh, E. et al. (2009) ‘Targeted loss of Arx results in a developmental epilepsy mouse model and recapitulates the human phenotype in heterozygous females’, Brain, 132(6), pp. 1563–1576. doi: 10.1093/brain/awp107. [PMC free article: PMC2685924] [PubMed: 19439424]
  135. Marsh, E. D. et al. (2016) ‘Developmental interneuron subtype deficits after targeted loss of Arx’, BMC Neuroscience, 17(1), p. 35. doi: 10.1186/s12868-016-0265-8. [PMC free article: PMC4902966] [PubMed: 27287386]
  136. Mayer, C. et al. (2015) ‘Clonally Related Forebrain Interneurons Disperse Broadly across Both Functional Areas and Structural Boundaries’, Neuron, 87(5), pp. 989–998. doi: 10.1016/j.neuron.2015.07.011. [PMC free article: PMC4560602] [PubMed: 26299473]
  137. Mayer, C. et al. (2018) ‘Developmental diversification of cortical inhibitory interneurons’, Nature, 555(7697), pp. 457–462. doi: 10.1038/nature25999. [PMC free article: PMC6052457] [PubMed: 29513653]
  138. McKenzie, M. et al. (2018) ‘Non-Canonical Wnt-Signaling through Ryk Regulates the Generation of Somatostatin- and Parvalbumin-Expressing Cortical Interneurons’, SSRN Electronic Journal, 103(5), pp. 853–864.e4. doi: 10.2139/ssrn.3155597. [PMC free article: PMC6728181] [PubMed: 31257105]
  139. McKinsey, G. L. et al. (2013) ‘Dlx1&amp;2-Dependent Expression of Zfhx1b (Sip1, Zeb2) Regulates the Fate Switch between Cortical and Striatal Interneurons’,  Neuron, 77(1), pp. 83–98. doi: 10.1016/j.neuron.2012.11.035. [PMC free article: PMC3547499] [PubMed: 23312518]
  140. Mehta, P. et al. (2019) ‘Functional Access to Neuron Subclasses in Rodent and Primate Forebrain’, Cell Reports, 26(10), pp. 2818–2832.e8. doi: 10.1016/j.celrep.2019.02.011. [PMC free article: PMC6509701] [PubMed: 30840900]
  141. Merkle, F. T., Mirzadeh, Z. and Alvarez-Buylla, A. (2007) ‘Mosaic organization of neural stem cells in the adult brain’, Science, 317(5836), pp. 381–384. doi: 10.1126/science.1144914. [PubMed: 17615304]
  142. Le Meur, N. et al. (2010) ‘MEF2C haploinsufficiency caused by either microdeletion of the 5q14.3 region or mutation is responsible for severe mental retardation with stereotypic movements, epilepsy and/or cerebral malformations’, Journal of Medical Genetics, 47(1), pp. 22–29. doi: 10.1136/jmg.2009.069732. [PMC free article: PMC2848840] [PubMed: 19592390]
  143. Meyers, E. N., Lewandoski, M. and Martin, G. R. (1998) ‘An Fgf8 mutant allelic series generated by Cre-and Flp-mediated recombination’, Nature Genetics, 18(2), pp. 136–141. doi: 10.1038/ng0298-136. [PubMed: 9462741]
  144. Mi, D. et al. (2018) ‘Early emergence of cortical interneuron diversity in the mouse embryo’, Science, 360(6384), pp. 81–85. doi: 10.1126/science.aar6821. [PMC free article: PMC6195193] [PubMed: 29472441]
  145. Miyoshi, G. et al. (2007) ‘Physiologically distinct temporal cohorts of cortical interneurons arise from telencephalic Olig2-expressing precursors.’, The Journal of neuroscience : the official journal of the Society for Neuroscience, 27(29), pp. 7786–7798. doi: 10.1523/JNEUROSCI.1807-07.2007. [PMC free article: PMC6672881] [PubMed: 17634372]
  146. Miyoshi, Goichi et al. (2010) ‘Genetic fate mapping reveals that the caudal ganglionic eminence produces a large and diverse population of superficial cortical interneurons.’, Journal of Neuroscience, 30(5), pp. 1582–1594. [PMC free article: PMC2826846] [PubMed: 20130169]
  147. Miyoshi, G. et al. (2010) ‘Genetic Fate Mapping Reveals That the Caudal Ganglionic Eminence Produces a Large and Diverse Population of Superficial Cortical Interneurons’, Journal of Neuroscience, 30(5), pp. 1582–1594. doi: 10.1523/JNEUROSCI.4515-09.2010. [PMC free article: PMC2826846] [PubMed: 20130169]
  148. Miyoshi, G. et al. (2015) ‘Prox1 regulates the subtype-specific development of caudal ganglionic eminence-derived GABAergic cortical interneurons’, Journal of Neuroscience, 35(37), pp. 12869–12889. doi: 10.1523/JNEUROSCI.1164-15.2015. [PMC free article: PMC4571608] [PubMed: 26377473]
  149. Moody, W. J. and Bosma, M. M. (2005) ‘Ion channel development, spontaneous activity, and activity-dependent development in nerve and muscle cells’, Physiological Reviews, 85(3), pp. 883–941. doi: 10.1152/physrev.00017.2004. [PubMed: 15987798]
  150. Motoyama, J. and Aoto, K. (2007) ‘Important Role of Shh Controlling Gli3 Functions during the Dorsal-Ventral Patterning of the Telencephalon’, Hedgehog-Gli Signaling in Human Disease, chapter 511, pp. 177–186. doi: 10.1007/0-387-33777-6_14.
  151. Muhr, J. et al. (2001) ‘Groucho-mediated transcriptional repression establishes progenitor cell pattern and neuronal fate in the ventral neural tube’, Cell, 104(6), pp. 861–873. doi: 10.1016/S0092-8674(01)00283-5. [PubMed: 11290324]
  152. Nagy, A. (2000) ‘Cre recombinase: The universal reagent for genome tailoring’, Genesis, 26(2), pp. 99–109. doi: 10.1002/(SICI)1526-968X(200002)26:2<99::AID-GENE1>3.0.CO;2-B. [PubMed: 10686599]
  153. Nathanson, J. L. et al. (2009) ‘Preferential labeling of inhibitory and excitatory cortical neurons by endogenous tropism of adeno-associated virus and lentivirus vectors’, Neuroscience, 161(2), pp. 441–450. doi: 10.1016/j.neuroscience.2009.03.032. [PMC free article: PMC2728494] [PubMed: 19318117]
  154. Nery, S., Fishell, G. and Corbin, J. G. (2002) ‘The caudal ganglionic eminence is a source of distinct cortical and subcortical cell populations’, Nature Neuroscience, 5(12), pp. 1279–1887. doi: 10.1038/nn971. [PubMed: 12411960]
  155. Neves, G. et al. (2013) ‘The LIM homeodomain protein Lhx6 regulates maturation of interneurons and network excitability in the mammalian cortex’, Cerebral Cortex, 23(8), pp. 1811–1823. doi: 10.1093/cercor/bhs159. [PMC free article: PMC3698365] [PubMed: 22710612]
  156. Niquille, M. et al. (2018) ‘Neurogliaform cortical interneurons derive from cells in the preoptic area’, eLife, 7, p. e32017. doi: 10.7554/eLife.32017. [PMC free article: PMC5860868] [PubMed: 29557780]
  157. Nóbrega-Pereira, S. et al. (2010) ‘Origin and molecular specification of globus pallidus neurons’, Journal of Neuroscience, 30(8), pp. 2824–2834. doi: 10.1523/JNEUROSCI.4023-09.2010. [PMC free article: PMC6633933] [PubMed: 20181580]
  158. Nord, A. S. et al. (2013) ‘Rapid and pervasive changes in genome-wide enhancer usage during mammalian development’, Cell, 155(7), pp. 1521–1531. doi: 10.1016/j.cell.2013.11.033. [PMC free article: PMC3989111] [PubMed: 24360275]
  159. Nord, A. S. et al. (2015) ‘Genomic Perspectives of Transcriptional Regulation in Forebrain Development’, Neuron, 85(1), pp. 27–47. doi: 10.1016/j.neuron.2014.11.011. [PMC free article: PMC4438709] [PubMed: 25569346]
  160. Novara, F. et al. (2010) ‘Refining the phenotype associated with MEF2C haploinsufficiency’, Clinical Genetics, 78(5), pp. 471–477. doi: 10.1111/j.1399-0004.2010.01413.x. [PubMed: 20412115]
  161. Obernier, K. and Alvarez-Buylla, A. (2019) ‘Neural stem cells: Origin, heterogeneity and regulation in the adult mammalian brain’, Development (Cambridge), 146(4), dev156059. doi: 10.1242/dev.156059. [PMC free article: PMC6398449] [PubMed: 30777863]
  162. Ohkubo, Y., Chiang, C. and Rubenstein, J. L. R. (2002) ‘Coordinate regulation and synergistic actions of BMP4, SHH and FGF8 in the rostral prosencephalon regulate morphogenesis of the telencephalic and optic vesicles’, Neuroscience, 111(1), pp. 1–17. doi: 10.1016/S0306-4522(01)00616-9. [PubMed: 11955708]
  163. Olsson, M. et al. (1997) ‘Extensive migration and target innervation by striatal precursors after grafting into the neonatal striatum’, Neuroscience, 79(1), pp. 57–78. doi: 10.1016/S0306-4522(96)00606-9. [PubMed: 9178865]
  164. Olsson, M., Björklund, A. and Campbell, K. (1998) ‘Early specification of striatal projection neurons and interneuronal subtypes in the lateral and medial ganglionic eminence’, Neuroscience, 84(3), pp. 867–876. doi: 10.1016/S0306-4522(97)00532-0. [PubMed: 9579790]
  165. Paciorkowski, A. R. et al. (2013) ‘MEF2C Haploinsufficiency features consistent hyperkinesis, variable epilepsy, and has a role in dorsal and ventral neuronal developmental pathways’, Neurogenetics, 14(2), pp. 99–111. doi: 10.1007/s10048-013-0356-y. [PMC free article: PMC3773516] [PubMed: 23389741]
  166. Pai, E. L.-L. et al. (2019) ‘Mafb and c-Maf Have Prenatal Compensatory and Postnatal Antagonistic Roles in Cortical Interneuron Fate and Function’, Cell Reports, 26(5), pp. 1157–1173.e5. doi: 10.1016/j.celrep.2019.01.031. [PMC free article: PMC6602795] [PubMed: 30699346]
  167. Pai, E. L.-L. et al. (2020) Maf and Mafb control mouse pallial interneuron fate and maturation through neuropsychiatric disease gene regulation’,  eLife. Edited by J. G. Gleeson, M. E. Bronner, and S. Anderson, 9, p. e54903. doi: 10.7554/eLife.54903. [PMC free article: PMC7282818] [PubMed: 32452758]
  168. Panganiban, G. and Rubenstein, J. L. R. (2002) ‘Developmental functions of the Distal-less/Dlx homeobox genes’, Development, 129(19), pp. 4371–4386. [PubMed: 12223397]
  169. Paredes, M. F. et al. (2016) ‘Extensive migration of young neurons into the infant human frontal lobe’, Science, 354(6308), p. aaf7073. doi: 10.1126/science.aaf7073. [PMC free article: PMC5436574] [PubMed: 27846470]
  170. Patel, S. R. et al. (2012) ‘Epigenetic Mechanisms of Groucho/Grg/TLE Mediated Transcriptional Repression’, Molecular Cell, 45(2), pp. 185–195. doi: 10.1016/j.molcel.2011.11.007. [PMC free article: PMC3268830] [PubMed: 22169276]
  171. Pattabiraman, K. et al. (2014) ‘Transcriptional regulation of enhancers active in protodomains of the developing cerebral cortex’, Neuron, 82(5), pp. 989-1003. doi: 10.1016/j.neuron.2014.04.014. [PMC free article: PMC4104757] [PubMed: 24814534]
  172. Patterning and Cell Type Specification in the Developing CNS and PNS (2020) Patterning and Cell Type Specification in the Developing CNS and PNS. doi: 10.1016/c2017-0-00860-1.
  173. Paul, A. et al. (2017) ‘Transcriptional Architecture of Synaptic Communication Delineates GABAergic Neuron Identity’, Cell, 171(3), pp. 522–539.e20. doi: 10.1016/j.cell.2017.08.032. [PMC free article: PMC5772785] [PubMed: 28942923]
  174. Pei, Z. et al. (2011) ‘Homeobox genes Gsx1 and Gsx2 differentially regulate telencephalic progenitor maturation’, Proceedings of the National Academy of Sciences of the United States of America, 108(4), pp. 1675–1680. doi: 10.1073/pnas.1008824108. [PMC free article: PMC3029701] [PubMed: 21205889]
  175. Pelkey, K. A. et al. (2017) ‘Hippocampal GABAergic Inhibitory Interneurons’, Physiological Reviews, 97(4), pp. 1619–1747. doi: 10.1152/physrev.00007.2017. [PMC free article: PMC6151493] [PubMed: 28954853]
  176. Petros, T. J. et al. (2015) ‘Apical versus Basal Neurogenesis Directs Cortical Interneuron Subclass Fate’, Cell Reports, 13(6), pp. 1090–1095. doi: 10.1016/j.celrep.2015.09.079. [PMC free article: PMC4704102] [PubMed: 26526999]
  177. Petryniak, Magdalena A., Potter, G. B., Rowitch, D. H. and Rubenstein, John L.R. (2007) ‘Dlx1 and Dlx2 Control Neuronal versus Oligodendroglial Cell Fate Acquisition in the Developing Forebrain’, Neuron, 55(3), pp. 417–433. doi: 10.1016/j.neuron.2007.06.036. [PMC free article: PMC2039927] [PubMed: 17678855]
  178. Pla, R. et al. (2006) ‘Layer acquisition by cortical GABAergic interneurons is independent of Reelin signaling.’, The Journal of neuroscience : the official journal of the Society for Neuroscience, 26(26), pp. 6924–6934. doi: 10.1523/JNEUROSCI.0245-06.2006. [PMC free article: PMC6673924] [PubMed: 16807322]
  179. Pla, R. et al. (2018) ‘Dlx1 and Dlx2 Promote Interneuron GABA Synthesis, Synaptogenesis, and Dendritogenesis’, Cerebral Cortex, 28(11), pp. 3797–3815. doi: 10.1093/cercor/bhx241. [PMC free article: PMC6188538] [PubMed: 29028947]
  180. Pollen, A. A. et al. (2019) ‘Establishing Cerebral Organoids as Models of Human-Specific Brain Evolution’, Cell, 176(4), pp. 743–756.e17. doi: 10.1016/j.cell.2019.01.017. [PMC free article: PMC6544371] [PubMed: 30735633]
  181. Porteus, M. H. et al. (1991) ‘Isolation and characterization of a novel cDNA clone encoding a homeodomain that is developmentally regulated in the ventral forebrain [published erratum appears in Neuron 1992 Jul;9(1):187]’, Neuron, 7(2), pp. 221–229. [PubMed: 1678612]
  182. Porteus, M. H. et al. (1992) ‘Isolation and characterization of a library of cDNA clones that are preferentially expressed in the embryonic telencephalon’, Molecular Brain Research, 12(1–3), pp. 7–22. doi: 10.1016/0169-328X(92)90063-H. [PubMed: 1372074]
  183. Potter, G. B. et al. (2009) ‘Generation of Cre-transgenic mice using Dlx1/Dlx2 enhancers and their characterization in GABAergic interneurons’, Molecular and Cellular Neuroscience, 40(2), pp. 167–186. doi: 10.1016/j.mcn.2008.10.003. [PMC free article: PMC2693920] [PubMed: 19026749]
  184. Price, M. et al. (1991) ‘A mouse gene related to Distal-less shows a restricted expression in the developing forebrain’, Nature, 351(6329), pp. 748–751. doi: 10.1038/351748a0. [PubMed: 1676488]
  185. Price, M. G. et al. (2009) ‘A triplet repeat expansion genetic mouse model of infantile spasms syndrome, Arx(GCG)10+7, with interneuronopathy, spasms in infancy, persistent seizures, and adult cognitive and behavioral impairment’, Journal of Neuroscience, 29(27), pp. 8752–8763. doi: 10.1523/JNEUROSCI.0915-09.2009. [PMC free article: PMC2782569] [PubMed: 19587282]
  186. Puelles, L. et al. (2000) ‘Pallial and subpallial derivatives in the embryonic chick and mouse telencephalon, traced by the expression of the genes Dlx-2, Emx-1, Nkx-2.1, Pax-6, and Tbr-1’, Journal of Comparative Neurology, 424(3), pp. 409–438. doi: 10.1002/1096-9861(20000828)424:3<409::AID-CNE3>3.0.CO;2-7. [PubMed: 10906711]
  187. Puelles, L. et al. (2016) ‘Radial and tangential migration of telencephalic somatostatin neurons originated from the mouse diagonal area’, Brain Structure and Function, 221(6), pp. 3027–3065. doi: 10.1007/s00429-015-1086-8. [PMC free article: PMC4920861] [PubMed: 26189100]
  188. Qin, S. et al. (2017) ‘Septal contributions to olfactory bulb interneuron diversity in the embryonic mouse telencephalon: Role of the homeobox gene Gsx2’, Neural Development, 12(1), p. 13. doi: 10.1186/s13064-017-0090-5. [PMC free article: PMC5559835] [PubMed: 28814342]
  189. Qiu, M. et al. (1997) ‘Role of the Dlx homeobox genes in proximodistal patterning of the branchial arches: Mutations of Dlx-1, Dlx-2, and Dlx-1 and -2 alter morphogenesis of proximal skeletal and soft tissue structures derived from the first and second arches’, Developmental Biology, 185(2), pp. 165–184. doi: 10.1006/dbio.1997.8556. [PubMed: 9187081]
  190. Rallu, M. et al. (2002) ‘Dorsoventral patterning is established in the telencephalon of mutants lacking both Gli3 and hedgehog signaling’, Development, 129(21), pp. 4963–4974. [PubMed: 12397105]
  191. Robinson, G. W., Wray, S. and Mahon, K. A. (1991) ‘Spatially restricted expression of a member of a new family of murine Distal-less homeobox genes in the developing forebrain’, New Biologist, 3(12), pp. 1183–1194. [PubMed: 1687503]
  192. Rubenstein, J.L.R. and Campbell, K. (2020). Neurogenesis in the basal ganglia. Comprehensive Developmental Neuroscience. Eds: J. Rubenstein and P. Rakic. Vol 1. Patterning and Cell Type Specification in the Developing CNS and PNS, 2nd Edition. Chapter 18. 399-403. Elsevier, Academic Press.
  193. Rubenstein, J. L. R. and Merzenich, M. M. (2003) ‘Model of autism: increased ratio of excitation/inhibition in key neural systems.’, Genes, brain, and behavior, 2(5), pp. 255–267. doi: 10.1046/j.1601-183X.2003.00037.x. [PMC free article: PMC6748642] [PubMed: 14606691]
  194. Rubin, A. N. et al. (2020) ‘Regulatory elements inserted into aavs confer preferential activity in cortical interneurons’, eNeuro, 7(6), eneuro.0211-0220.2020. doi: 10.1523/ENEURO.0211-20.2020. [PMC free article: PMC7768279] [PubMed: 33199411]
  195. Rubin, A. N. and Kessaris, N. (2013) ‘PROX1: A Lineage Tracer for Cortical Interneurons Originating in the Lateral/Caudal Ganglionic Eminence and Preoptic Area’, PLoS ONE, 8(10), p. e77339. doi: 10.1371/journal.pone.0077339. [PMC free article: PMC3796451] [PubMed: 24155945]
  196. Rudy, B. et al. (2011) ‘Three groups of interneurons account for nearly 100% of neocortical GABAergic neurons’, Developmental Neurobiology, 71(1), pp. 45–61. doi: 10.1002/dneu.20853. [PMC free article: PMC3556905] [PubMed: 21154909]
  197. Rymar, V. V. and Sadikot, A. F. (2007) ‘Laminar fate of cortical GABAergic interneurons is dependent on both birthdate and phenotype’, Journal of Comparative Neurology, 501(3), pp. 369–380. doi: 10.1002/cne.21250. [PubMed: 17245711]
  198. Salomone, J. et al. (2021) ‘Conserved Gsx2/Ind homeodomain monomer versus homodimer DNA binding defines regulatory outcomes in flies and mice’, Genes and Development, 35(1), pp. 157–174. doi: 10.1101/GAD.343053.120. [PMC free article: PMC7778271] [PubMed: 33334823]
  199. Sandberg, M., Flandin, P., Silberberg, S., Su-Feher, L., Price, J. D., Hu, J. S., Kim, C., Visel, A., Nord, A. S. and Rubenstein, John L R (2016) ‘Transcriptional Networks Controlled by NKX2-1 in the Development of Forebrain GABAergic Neurons’, Neuron, 91(6), pp. 1260–1275. doi: 10.1016/j.neuron.2016.08.020. [PMC free article: PMC5319854] [PubMed: 27657450]
  200. Saunders, A. et al. (2018) ‘Molecular Diversity and Specializations among the Cells of the Adult Mouse Brain’, Cell, 174(4), pp. 1015–1030.e16. doi: 10.1016/j.cell.2018.07.028. [PMC free article: PMC6447408] [PubMed: 30096299]
  201. Schüle, B. et al. (2007) ‘DLX5 and DLX6 expression is biallelic and not modulated by MeCP2 deficiency’, American Journal of Human Genetics, 81(3), pp. 492–506. doi: 10.1086/520063. [PMC free article: PMC1950824] [PubMed: 17701895]
  202. Schuman, B. et al. (2019) ‘Four unique interneuron populations reside in neocortical layer 1’, Journal of Neuroscience, 39(1), pp. 125–139. doi: 10.1523/JNEUROSCI.1613-18.2018. [PMC free article: PMC6325270] [PubMed: 30413647]
  203. Sebat, J. et al. (2007) ‘Strong association of de novo copy number mutations with autism’, Science, 316(5823), pp. 445–449. doi: 10.1126/science.1138659. [PMC free article: PMC2993504] [PubMed: 17363630]
  204. Shimamura, K. and Rubenstein, J. L. R. (1997) ‘Inductive interactions direct early regionalization of the mouse forebrain’, Development, 124(14), pp. 2709–2718. [PubMed: 9226442]
  205. Silberberg, S. N., Taher, L., Lindtner, S., Sandberg, M., Nord, A. S., Vogt, D., Mckinsey, G. L., Hoch, R., Pattabiraman, K., Zhang, D., Ferran, J. L., Rajkovic, A., Golonzhka, O., Kim, C., Zeng, H., Puelles, L., Visel, A. and Rubenstein, John L R (2016) ‘Subpallial Enhancer Transgenic Lines: a Data and Tool Resource to Study Transcriptional Regulation of GABAergic Cell Fate’, Neuron, 92(1), pp. 59–74. doi: 10.1016/j.neuron.2016.09.027. [PMC free article: PMC5063253] [PubMed: 27710791]
  206. Silbereis, J. C. et al. (2014) ‘Olig1 Function Is Required to Repress Dlx1/2 and Interneuron Production in Mammalian Brain’, Neuron, 81(3), pp. 574–587. doi: 10.1016/j.neuron.2013.11.024. [PMC free article: PMC3979971] [PubMed: 24507192]
  207. Sohal, V. S. and Rubenstein, J. L. R. (2019) ‘Excitation- inhibition balance as a framework for investigating mechanisms in neuropsychiatric disorders’, Molecular Psychiatry, 24(9), pp. 1248–1257. doi: 10.1038/s41380-019-0426-0. [PMC free article: PMC6742424] [PubMed: 31089192]
  208. Soria, J. M. et al. (2004) ‘Defective postnatal neurogenesis and disorganization of the rostral migratory stream in absence of the Vax1 homeobox gene’, Journal of Neuroscience, 24(49), pp. 111717–11181. doi: 10.1523/JNEUROSCI.3248-04.2004. [PMC free article: PMC6730283] [PubMed: 15590934]
  209. Sousa, V. H. et al. (2009) ‘Characterization of Nkx6-2-derived neocortical interneuron lineages’, Cerebral Cortex, Suppl1(Suppl1):i1–10. doi: 10.1093/cercor/bhp038. [PMC free article: PMC2693535] [PubMed: 19363146]
  210. Southwell, D. G. et al. (2012) ‘Intrinsically determined cell death of developing cortical interneurons’, Nature, 491(7422), pp. 109–113. doi: 10.1038/nature11523. [PMC free article: PMC3726009] [PubMed: 23041929]
  211. Stanco, A. et al. (2014) ‘NPAS1 Represses the Generation of Specific Subtypes of Cortical Interneurons’, Neuron, 84(5), pp. 940–953. doi: 10.1016/j.neuron.2014.10.040. [PMC free article: PMC4258152] [PubMed: 25467980]
  212. Stenman, J., Toresson, H. and Campbell, K. (2003) ‘Identification of two distinct progenitor populations in the lateral ganglionic eminence: Implications for striatal and olfactory bulb neurogenesis’, Journal of Neuroscience, 23(1), pp. 167–174. doi: 10.1523/jneurosci.23-01-00167.2003. [PMC free article: PMC6742158] [PubMed: 12514213]
  213. Storm, E. E. et al. (2006) ‘Dose-dependent functions fo Fgf8 in regulating telencephalic patterning centers’, Development, 133(9), pp. 1831–1844. doi: 10.1242/dev.02324. [PubMed: 16613831]
  214. Stühmer, T., Anderson, S. A., et al. (2002a) ‘Ectopic expression of the Dlx genes induces glutamic acid decarboxylase and Dlx expression’, Development, 129(1), pp. 245–252. [PubMed: 11782417]
  215. Stühmer, T., Puelles, L., et al. (2002b) ‘Expression from a Dlx gene enhancer marks adult mouse cortical GABAergic neurons’, Cerebral Cortex, 12(1), pp. 75–85. doi: 10.1093/cercor/12.1.75. [PubMed: 11734534]
  216. Stumm, R. K. et al. (2003) ‘CXCR4 regulates interneuron migration in the developing neocortex’, The Journal of Neuroscience, 23(12), pp. 5123–5130. doi: 23/12/5123 [pii]. [PMC free article: PMC6741192] [PubMed: 12832536]
  217. Su-Feher, L. et al. (2021) ‘Single Cell Enhancer Activity Distinguishes GABAergic and Cholinergic Lineages in Embryonic Mouse Basal Ganglia’, Proc Natil Acad Sci, 119(15), p. e2108760119. doi: 10.1073/pnas.2108760119. [PMC free article: PMC9169651] [PubMed: 35377797]
  218. Sun, Y. et al. (2016) ‘A deleterious Nav1.1 mutation selectively impairs telencephalic inhibitory neurons derived from Dravet Syndrome patients’, eLife, 5, p. e13073. doi: 10.7554/eLife.13073. [PMC free article: PMC4961470] [PubMed: 27458797]
  219. Sussel, L. et al. (1999) ‘Loss of Nkx2.1 homeobox gene function results in a ventral to dorsal molecular respecification within the basal telencephalon: evidence for a transformation of the pallidum into the striatum’, Development, 126(15), pp. 3359–3370. doi: 10393115. [PubMed: 10393115]
  220. Taniguchi, H. et al. (2011) ‘A Resource of Cre Driver Lines for Genetic Targeting of GABAergic Neurons in Cerebral Cortex’, Neuron, 71(6), pp. 995–1013. doi: 10.1016/j.neuron.2011.07.026. [PMC free article: PMC3779648] [PubMed: 21943598]
  221. Tasic, B. et al. (2016) ‘Adult mouse cortical cell taxonomy revealed by single cell transcriptomics’, Nature Neuroscience, 19(2), pp. 335–346). doi: 10.1038/nn.4216. [PMC free article: PMC4985242] [PubMed: 26727548]
  222. Tasic, B. et al. (2018) ‘Shared and distinct transcriptomic cell types across neocortical areas.’, Nature, 563(7729), pp. 72–78. doi: 10.1038/s41586-018-0654-5. [PMC free article: PMC6456269] [PubMed: 30382198]
  223. Toresson, H. and Campbell, K. (2001) ‘A role Gsh1 in the developing striatum and olfactory bulb of Gsh2 mutant mice’, Development, 128(23), pp. 4769–4780. [PubMed: 11731457]
  224. Toresson, H., Potter, S. S. and Campbell, K. (2000) ‘Genetic control of dorsal-ventral identity in the telencephalon: Opposing roles for Pax6 and Gsh2’, Development, 127(20), pp. 4361–4371. [PubMed: 11003836]
  225. Torigoe, M. et al. (2016) ‘Evidence that the laminar fate of LGE/CGE-derived neocortical interneurons is dependent on their progenitor domains’, Journal of Neuroscience, 36(6), pp. 2044–2056. doi: 10.1523/JNEUROSCI.3550-15.2016. [PMC free article: PMC6602017] [PubMed: 26865626]
  226. Touzot, Audrey et al. (2016) ‘Molecular control of two novel migratory paths for CGE-derived interneurons in the developing mouse brain’, Development, 143(10), pp. 1753–1765. doi: 10.1242/dev.131102. [PubMed: 27034423]
  227. Tripodi, M. et al. (2004) ‘The COUP-TF nuclear receptors regulate cell migration in the mammalian basal forebrain’, Development, 131(24), pp. 6119–6129. doi: 10.1242/dev.01530. [PubMed: 15548577]
  228. Tu, S. et al. (2017) ‘NitroSynapsin therapy for a mouse MEF2C haploinsufficiency model of human autism’, Nature Communications, 8(1), p. 1488. doi: 10.1038/s41467-017-01563-8. [PMC free article: PMC5684358] [PubMed: 29133852]
  229. Valcanis, H. and Tan, S. S. (2003) ‘Layer specification of transplanted interneurons in developing mouse neocortex’, Journal of Neuroscience, 23(12), pp. 5113–5122. doi: 10.1523/jneurosci.23-12-05113.2003. [PMC free article: PMC6741168] [PubMed: 12832535]
  230. Velmeshev, D.  et al. (2021) ‘Molecular Diversity and Lineage Commitment of Human Interneuron Progenitors’, Biorxiv. doi: 10.1101/2021.05/13.444045
  231. Visel, A. et al. (2009) ‘ChIP-seq accurately predicts tissue-specific activity of enhancers’, Nature, 457(7231), pp. 854–858. doi: 10.1038/nature07730. [PMC free article: PMC2745234] [PubMed: 19212405]
  232. Visel, A. et al. (2013) ‘A high-resolution enhancer atlas of the developing telencephalon’, Cell, 152(4), pp. 895–908 doi: 10.1016/j.cell.2012.12.041. [PMC free article: PMC3660042] [PubMed: 23375746]
  233. Vogt, D., Hunt, Robert F., et al. (2014) ‘Lhx6 Directly Regulates Arx and CXCR7 to Determine Cortical Interneuron Fate and Laminar Position’, Neuron, 82(2), pp. 350–364. doi: 10.1016/j.neuron.2014.02.030. [PMC free article: PMC4261952] [PubMed: 24742460]
  234. Vogt, D. et al. (2015) ‘The Parvalbumin/Somatostatin Ratio Is Increased in Pten Mutant Mice and by Human PTEN ASD Alleles’, Cell Reports, 11(6), pp. 944–956. doi: 10.1016/j.celrep.2015.04.019. [PMC free article: PMC4431948] [PubMed: 25937288]
  235. Vogt, D. et al. (2018) ‘Mouse Cntnap2 and Human CNTNAP2 ASD Alleles Cell Autonomously Regulate PV + Cortical Interneurons’, Cerebral Cortex, 25(11), pp. 3868–3879. doi: 10.1093/cercor/bhx248. [PMC free article: PMC6455910] [PubMed: 29028946]
  236. Vormstein-Schneider, D. et al. (2020) ‘Viral manipulation of functionally distinct interneurons in mice, non-human primates and humans’, Nature Neuroscience, 23(12), pp. 1629–1636. doi: 10.1038/s41593-020-0692-9. [PMC free article: PMC8015416] [PubMed: 32807948]
  237. Waclaw, R. R. et al. (2006) ‘The zinc finger transcription factor Sp8 regulates the generation and diversity of olfactory bulb interneurons’, Neuron, 49(4), pp. 503–516. doi: 10.1016/j.neuron.2006.01.018. [PubMed: 16476661]
  238. Walsh, T. et al. (2008) ‘Rare structural variants disrupt multiple genes in neurodevelopmental pathways in schizophrenia’, Science, 320(5875), pp. 539–543. doi: 10.1126/science.1155174. [PubMed: 18369103]
  239. Wang, B. et al. (2009) ‘Ascl1 is a required downstream effector of Gsx gene function in the embryonic mouse telencephalon’, Neural Development, 4(1), p. 5. doi: 10.1186/1749-8104-4-5. [PMC free article: PMC2644683] [PubMed: 19208224]
  240. Wang, B. et al. (2013) ‘Loss of Gsx1 and Gsx2 function rescues distinct phenotypes in Dlx1/2 mutants’, Journal of Comparative Neurology, 521(7), pp. 1561–1584. doi: 10.1002/cne.23242. [PMC free article: PMC3615175] [PubMed: 23042297]
  241. Wang, Y. et al. (2010) ‘Dlx5 and Dlx6 regulate the development of parvalbumin-expressing cortical interneurons.’, Journal of Neuroscience, 30(15), pp. 5334–5345. [PMC free article: PMC2919857] [PubMed: 20392955]
  242. Wang, Y., Li, G., Stanco, A., Long, J. E., Crawford, D., Potter, G. B., Pleasure, S. J., Behrens, T. and Rubenstein, John L R (2011) ‘CXCR4 and CXCR7 Have Distinct Functions in Regulating Interneuron Migration’, Neuron, 69(1), pp. 61–76. doi: 10.1016/j.neuron.2010.12.005. [PMC free article: PMC3025760] [PubMed: 21220099]
  243. Wei, S. et al. (2019) ‘Transcription factors Sp8 and Sp9 regulate the development of caudal ganglionic eminence-derived cortical interneurons’, Journal of Comparative Neurology, 527(17), pp. 2860–2874. doi: 10.1002/cne.24712. [PubMed: 31070778]
  244. Wichterle, H. et al. (1999) ‘Young neurons from medial ganglionic eminence disperse in adult and embryonic brain’, Nature Neuroscience, 2(5), pp. 461–466. doi: 10.1038/8131. [PubMed: 10321251]
  245. Wichterle, H. et al. (2001) ‘In utero fate mapping reveals distinct migratory pathways and fates of neurons born in the mammalian basal forebrain’, Development, 128(19), pp. 3759–3771. [PubMed: 11585802]
  246. Wonders, Carl P and Anderson, S. A. (2006) ‘The origin and specification of cortical interneurons.’, Nature reviews. Neuroscience, 7(9), pp. 687–696. doi: 10.1038/nrn1954. [PubMed: 16883309]
  247. Wu, P. R. et al. (2017) ‘The cytokine CXCL12 promotes basket interneuron inhibitory synapses in the medial prefrontal cortex’, Cerebral Cortex, 27(9), pp. 4303–4313. doi: 10.1093/cercor/bhw230. [PMC free article: PMC6410508] [PubMed: 27497284]
  248. Wu, Z., Yang, H. and Colosi, P. (2010) ‘Effect of genome size on AAV vector packaging’, Molecular Therapy, 18(1), pp. 80–86. doi: 10.1038/mt.2009.255. [PMC free article: PMC2839202] [PubMed: 19904234]
  249. Wundrach, D. et al. (2020) ‘A Human TSC1 Variant Screening Platform in Gabaergic Cortical Interneurons for Genotype to Phenotype Assessments’, Frontiers in Molecular Neuroscience, 13, p. e573409. doi: 10.3389/fnmol.2020.573409. [PMC free article: PMC7539171] [PubMed: 33071758]
  250. Xu, Q. et al. (2010) ‘Sonic Hedgehog Signaling Confers Ventral Telencephalic Progenitors with Distinct Cortical Interneuron Fates’, Neuron, 65(3), pp. 328–340. doi: 10.1016/j.neuron.2010.01.004. [PMC free article: PMC2868511] [PubMed: 20159447]
  251. Xu, Q., Tam, M. and Anderson, S. A. (2008) ‘Fate mapping Nkx2.1-lineage cells in the mouse telencephalon’, Journal of Comparative Neurology, 506(1), pp. 16–29. doi: 10.1002/cne.21529. [PubMed: 17990269]
  252. Xu, Q., Wonders, C. P. and Anderson, S. A. (2005) ‘Sonic hedgehog maintains the identity of cortical interneuron progenitors in the ventral telencephalon’, Development, 132(22), pp. 4987–4998. doi: 10.1242/dev.02090. [PubMed: 16221724]
  253. Yizhar, O. et al. (2011) ‘Neocortical excitation/inhibition balance in information processing and social dysfunction’, Nature, 477(7363), pp. 171–178. doi: 10.1038/nature10360. [PMC free article: PMC4155501] [PubMed: 21796121]
  254. Yoshihara, S. I. et al. (2005) ‘Arx homeobox gene is essential for development of mouse olfactory system’, Development (2005) 132 (4): 751–762. doi: 10.1242/dev.01619. [PubMed: 15677725]
  255. Yu, M. et al. (2011) ‘Activity of dlx5a/dlx6a regulatory elements during zebrafish GABAergic neuron development’, International Journal of Developmental Neuroscience, 29(7), pp. 681–691. doi: 10.1016/j.ijdevneu.2011.06.005. [PubMed: 21723936]
  256. Yun, K., Fischman, S., Johnson, J., Hrabe de Angelis, M., Weinmaster, G., et al. (2002) ‘Modulation of the notch signaling by Mash1 and Dlx1/2 regulates sequential specification and differentiation of progenitor cell types in the subcortical telencephalon.’, Development (Cambridge, England), 129(21), pp. 5029–5040. [PubMed: 12397111]
  257. Yun, K. et al. (2003) ‘Patterning of the lateral ganglionic eminence by the Gsh1 and Gsh2 homeobox genes regulates striatal and olfactory bulb histogenesis and the growth of axons through the basal ganglia’, Journal of Comparative Neurology, 461(2), pp. 151–165. doi: 10.1002/cne.10685. [PubMed: 12724834]
  258. Yun, K., Potter, S. and Rubenstein, J. L. R. (2001) ‘Gsh2 and Pax6 play complementary roles in dorsoventral patterning of the mammalian telencephalon’, Development, 128(2), pp. 193–205. [PubMed: 11124115]
  259. Zaki, P. A., Martynoga, B. and Price, D. J. (2008) ‘Role of Hedgehog and Gli Signalling in Telencephalic Development’, in Shh and Gli Signalling and Development. doi: 10.1007/978-0-387-39957-7_3.
  260. Zerucha, T. et al. (2000) ‘A highly conserved enhancer in the Dlx5/Dlx6 intergenic region is the site of cross- regulatory interactions between Dlx genes in the embryonic forebrain’, Journal of Neuroscience, 20(2), pp. 709–721. doi: 10.1523/jneurosci.20-02-00709.2000. [PMC free article: PMC6772408] [PubMed: 10632600]
  261. Zhang, Y. et al. (2020) ‘Cortical Neural Stem Cell Lineage Progression Is Regulated by Extrinsic Signaling Molecule Sonic Hedgehog’, Cell Reports, 30(13), pp. 4490–4504.e4. doi: 10.1016/j.celrep.2020.03.027. [PMC free article: PMC7197103] [PubMed: 32234482]
  262. Zhao, Y. et al. (2003) ‘The LIM-homeobox gene Lhx8 is required for the development of many cholinergic neurons in the mouse forebrain’, Proceedings of the National Academy of Sciences of the United States of America, 100(15), pp. 9005–9010. doi: 10.1073/pnas.1537759100. [PMC free article: PMC166428] [PubMed: 12855770]
  263. Zhao, Y. et al. (2008) ‘Distinct molecular pathways of development of telencephalic interneuron subtypes revealed through analysis of Lhx6 mutants’, Journal of Comparative Neurology, 510(1), pp. 79–99. doi: 10.1002/cne.21772. [PMC free article: PMC2547494] [PubMed: 18613121]
  264. Zhou, Y. D. et al. (1997) ‘Molecular characterization of two mammalian bHLH-PAS domain proteins selectively expressed in the central nervous system’, Proceedings of the National Academy of Sciences of the United States of America, 94(2), pp. 713–718. doi: 10.1073/pnas.94.2.713. [PMC free article: PMC19579] [PubMed: 9012850]
Copyright Notice

This is an open access publication, available online and distributed under the terms of a Creative Commons Attribution-Non Commercial-No Derivatives 4.0 International licence (CC BY-NC-ND 4.0), a copy of which is available at https://creativecommons.org/licenses/by-nc-nd/4.0/. Subject to this license, all rights are reserved.

Bookshelf ID: NBK609832PMID: 39637134DOI: 10.1093/med/9780197549469.003.0047

Views

  • PubReader
  • Print View
  • Cite this Page
  • PDF version of this title (283M)

Related information

  • PMC
    PubMed Central citations
  • PubMed
    Links to PubMed

Similar articles in PubMed

See reviews...See all...

Recent Activity

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

See more...