Adult neurogenesis is tightly regulated through the interaction of neural stem/progenitor cells (NSCs) with their niche. Neurotransmitters, including GABA activation of GABAA receptor ion channels, are important niche signals. We show that adult mouse hippocampal NSCs and their progeny express metabotropic GABAB receptors. Pharmacological inhibition of GABAB receptors stimulated NSC proliferation and genetic deletion of GABAB1 receptor subunits increased NSC proliferation and differentiation of neuroblasts in vivo. Cell-specific conditional deletion of GABAB receptors supports a cell-autonomous role in newly generated cells. Our data indicate that signaling through GABAB receptors is an inhibitor of adult neurogenesis.

The adult hippocampus contains neural stem/progenitor cells (NSCs) within a specialized subgranular zone (SGZ) niche of the dentate gyrus (DG) (Kempermann et al., 2004). Hippocampal NSCs depend on canonical Notch signaling for their maintenance and express the Notch target gene Hes5 (Breunig et al., 2007; Ables et al., 2010; Ehm et al., 2010; Lugert et al., 2010). Hes5+ NSCs produce intermediate progenitors that generate proliferating neuroblasts, which exit the cell cycle before differentiating into granule neurons (Lugert et al., 2012). Neurogenesis is tightly regulated through a balance of NSC maintenance and differentiation signals within the SGZ niche. Neurotransmitters may mediate crosstalk between newly generated cells and the surrounding neuronal network (Masiulis et al., 2011). Under physiological conditions, DG neurogenesis is modulated by neural excitation (Deisseroth et al., 2004; Tozuka et al., 2005; Parent, 2007) and accumulating evidence indicates that neurotransmitters can influence the proliferation and differentiation of newborn cells (Ge et al., 2006; Jagasia et al., 2009; Jhaveri et al., 2010; Duveau et al., 2011; Song et al., 2012). GABA is the major inhibitory neurotransmitter in the adult brain acting via two main receptor types: ionotropic GABAA and G-protein coupled metabotropic GABAB receptors. Adult neurogenesis is sensitive to GABAA receptor signaling (Masiulis et al., 2011; Song et al., 2012); however, a role for GABA signaling through GABAB receptors in the regulation of adult NSCs remains poorly defined.

GABAB receptors are heterodimers composed of GABAB1 and GABAB2 (Gabbr1 and Gabbr2 - Mouse Genome Informatics) subunits, both of which are required for normal receptor function (Ulrich and Bettler, 2007). Accordingly, mice lacking the GABAB1 subunit (Gabbr1-/-; hereafter GABAB1-/-) show a complete absence of GABAB responses (Schuler et al., 2001). Distinct isoforms of GABAB1 receptor subunits (GABAB1a and GABAB1b) are generated from the GABAB1 gene by differential promoter usage. Receptors containing GABAB1a and GABAB1b subunits exhibit a preferential axonal versus dendritic distribution, respectively, and accordingly they mediate distinct synaptic functions (Pérez-Garci et al., 2006; Vigot et al., 2006). GABAB receptors regulate neuronal excitability controlling the activity of voltage-gated calcium channels and inward-rectifying potassium channels (Ulrich and Bettler, 2007). GABAB receptors affect progenitor proliferation and migration in the developing brain (Fukui et al., 2008; Salazar et al., 2008; Wang and Kriegstein, 2009). However, whether GABAB receptors play a role in adult NSC biology in vivo is unclear.

Here we employed genetic and pharmacological approaches to investigate GABAB receptor function in regulating adult hippocampal neurogenesis. We show that GABAB receptors are expressed by many cell types in the adult DG. GABAB signaling is active in cells throughout the adult neurogenic lineage including the most primitive Hes5-expressing quiescent NSCs. Genetic and pharmacological inhibition of GABAB receptor signaling increases proliferation of Hes5+ NSCs, and increases the production of new neurons. Hence, our data indicate that GABAB signaling is an important inhibitor of adult neurogenesis and promotes the quiescence of NSCs in the DG though an ion-channel-independent mechanism.

GABAB receptors are expressed by cells in the adult neurogenic niche

GABAB receptors are expressed by most hippocampal neurons in mice (Fig. 1A-F) (Schuler et al., 2001); however, it is not known whether they are expressed by newly generated cells in the SGZ. To address whether newly generated adult granule neurons express GABAB receptors, we labeled proliferating cells in vivo with bromodeoxyuridine (BrdU) followed by a chase period of 30 days to allow for maturation of BrdU-labeled cells. Most BrdU-labeled neurons expressed GABAB1 and GABAB2 subunits (Fig. 1G,H) suggesting that GABAB signaling may have cell-autonomous functions in adult-generated granule cells. We also observed GABAB-expressing NeuN-negative cells in the SGZ (Fig. 1F). We analyzed mice expressing functional GABAB1-GFP fusion proteins under the control of GABAB1 regulatory elements (Fig. 1I) (Casanova et al., 2009). GABAB1-GFP colocalized with the neuroblast marker polysialylated neural cell adhesion molecule and brain lipid binding protein in progenitor cells (Fig. 1J,K). Hence, GABAB1 receptors were expressed by cells early within the neurogenic lineage and before neuronal maturation.

Fig. 1.

GABAB receptors are expressed and active in the mouse adult hippocampal neurogenic niche. (A-D) Immunostaining for GABAB1 (A,B) and GABAB2 (C,D) receptor subunits expression in the adult hippocampal DG of wild-type (A,C) and GABAB1-/- (B) or GABAB2-/- (D) mice. GABAB1 and GABAB2 subunits are expressed in a similar pattern in the granule cell layer (GrL), SGZ and hilus of the DG (A,C). Only weak residual background staining is visible in the mutant mice (B,D). (E,F) Most Calbindin- and NeuN-positive granule neurons express GABAB1 and GABAB2 subunits (arrows). Note that some NeuN-negative cells in the SGZ express GABAB receptors (F, arrowheads). The asterisk in F indicates a hilar neuron expressing GABAB2 subunits. (G,H) Newly generated hippocampal granule neurons express GABAB receptors. Thirty days after injection of bromodeoxyuridine (BrdU), newborn neurons were identified by BrdU and Calbindin or NeuN immunostaining. BrdU-positive neurons express both GABAB1 and GABAB2 receptor subunits (G,H, arrowheads). (I-K) Expression of the GABAB1-GFP transgene in the hippocampal neurogenic niche. GABAB1-GFP fusion is expressed in the GrL, SGZ and hilus similar to the endogenous GABAB1 subunit. PSA-NCAM positive neuroblasts (J, arrowheads) and BLBP positive progenitors (K, arrowheads) in the SGZ express GABAB1-GFP. (L,M) Representative images showing Hes5::GFP expressing NSCs in the SGZ of the adult hippocampus. GABAB1-positive cells (arrowheads) can be found among the Hes5-expressing population. Scale bars: A-D, 100 μm; E,F,I, 20 μm; G,H,J-M, 10 μm. GrL, granule cell layer; SGZ, subgranular zone.

Fig. 1.

GABAB receptors are expressed and active in the mouse adult hippocampal neurogenic niche. (A-D) Immunostaining for GABAB1 (A,B) and GABAB2 (C,D) receptor subunits expression in the adult hippocampal DG of wild-type (A,C) and GABAB1-/- (B) or GABAB2-/- (D) mice. GABAB1 and GABAB2 subunits are expressed in a similar pattern in the granule cell layer (GrL), SGZ and hilus of the DG (A,C). Only weak residual background staining is visible in the mutant mice (B,D). (E,F) Most Calbindin- and NeuN-positive granule neurons express GABAB1 and GABAB2 subunits (arrows). Note that some NeuN-negative cells in the SGZ express GABAB receptors (F, arrowheads). The asterisk in F indicates a hilar neuron expressing GABAB2 subunits. (G,H) Newly generated hippocampal granule neurons express GABAB receptors. Thirty days after injection of bromodeoxyuridine (BrdU), newborn neurons were identified by BrdU and Calbindin or NeuN immunostaining. BrdU-positive neurons express both GABAB1 and GABAB2 receptor subunits (G,H, arrowheads). (I-K) Expression of the GABAB1-GFP transgene in the hippocampal neurogenic niche. GABAB1-GFP fusion is expressed in the GrL, SGZ and hilus similar to the endogenous GABAB1 subunit. PSA-NCAM positive neuroblasts (J, arrowheads) and BLBP positive progenitors (K, arrowheads) in the SGZ express GABAB1-GFP. (L,M) Representative images showing Hes5::GFP expressing NSCs in the SGZ of the adult hippocampus. GABAB1-positive cells (arrowheads) can be found among the Hes5-expressing population. Scale bars: A-D, 100 μm; E,F,I, 20 μm; G,H,J-M, 10 μm. GrL, granule cell layer; SGZ, subgranular zone.

The Notch target Hes5 is expressed by NSCs in the adult DG segregating the most primitive Sox2+ progenitors from more committed cells (Lugert et al., 2010; Lugert et al., 2012). By analyzing Hes5::GFP mice we found that Hes5+ NSCs expressed GABAB1 subunits (Fig. 1L,M). Hes5+ cells with both radial and horizontal morphologies expressed GABAB1 subunits (radial, 86%; horizontal, 76%). These data indicate that adult hippocampal NSCs and their progeny express GABAB receptors.

Increased adult progenitor proliferation in GABA B1-/- mice

We addressed whether GABAB receptors play a role in adult hippocampal neurogenesis by analyzing GABAB1-/- mice (Schuler et al., 2001). The number of proliferating [proliferating cell nuclear antigen (PCNA) or phospho-histone-H3-expressing] cells in the SGZ and granule cell layer (GrL) in GABAB1-/- mice was significantly increased compared with wild-type controls (Fig. 2A-C; supplementary material Fig. S1A-C). PCNA+ cells in the adult DG include partially overlapping Sox2 progenitor and Doublecortin+ (Dcx) neuroblast populations (Fig. 2D-G) (Kempermann et al., 2004; Lugert et al., 2010). PCNA+ Sox2+ Dcx- progenitors but not PCNA+ Dcx+ neuroblasts were increased in GABAB1-/- mice, indicating that enhanced proliferation results from activation of the more undifferentiated progenitor populations (Fig. 2H). Moreover, the number of Dcx-expressing neuroblasts was increased, whereas Sox2+ progenitors were slightly decreased per mm2 in the DG of GABAB1-/- mice, suggesting augmented neurogenesis and enhanced progenitor differentiation (Fig. 2I). Taken together, these results indicate that GABAB receptor activity controls the number of proliferating progenitors in the adult hippocampus.

Fig. 2.

Increased proliferation in the hippocampus of GABAB1-deficient mice. (A,B) Representative images of proliferating PCNA+ cells in the adult hippocampal DG of GABAB1-/- and wild-type control mice. (C) The density of PCNA+ proliferating cells is increased in mutant mice (control 169±11; GABAB1-/- 311±33; n=6/5). (D-G) Phenotypic analysis of PCNA+ proliferating cells in control (D,F) and GABAB1-/- mice (E,G). Sox2 and Dcx were used to label progenitor cells and neuroblasts, respectively. The density of PCNA+Sox2+ cells (E) as well as the overall Dcx+ population (G) are increased in the DG of mutant mice. (H) Most PCNA+ cells express the progenitor markers Sox2 and not Dcx, and their density is increased in GABAB1 mutants (control 120±10; GABAB1-/- 233±40; n=6/5). The density of proliferating neuroblasts (PCNA+Dcx+) is unchanged in GABAB1 mutants (control 46±6.5; GABAB1-/- 69±13; n=6/5; P=0.16). (I) The density of Sox2+ cells is slightly reduced in mutant mice (control 1671±67; GABAB1-/- 1278±77; n=6/5) whereas that of Dcx+ cells is increased (control 858±69; GABAB1-/- 1477±220; n=6/5). t-test: *P<0.05. Error bars indicate s.e.m. Scale bars: A,B, 100 μm; D-G, 50 μm.

Fig. 2.

Increased proliferation in the hippocampus of GABAB1-deficient mice. (A,B) Representative images of proliferating PCNA+ cells in the adult hippocampal DG of GABAB1-/- and wild-type control mice. (C) The density of PCNA+ proliferating cells is increased in mutant mice (control 169±11; GABAB1-/- 311±33; n=6/5). (D-G) Phenotypic analysis of PCNA+ proliferating cells in control (D,F) and GABAB1-/- mice (E,G). Sox2 and Dcx were used to label progenitor cells and neuroblasts, respectively. The density of PCNA+Sox2+ cells (E) as well as the overall Dcx+ population (G) are increased in the DG of mutant mice. (H) Most PCNA+ cells express the progenitor markers Sox2 and not Dcx, and their density is increased in GABAB1 mutants (control 120±10; GABAB1-/- 233±40; n=6/5). The density of proliferating neuroblasts (PCNA+Dcx+) is unchanged in GABAB1 mutants (control 46±6.5; GABAB1-/- 69±13; n=6/5; P=0.16). (I) The density of Sox2+ cells is slightly reduced in mutant mice (control 1671±67; GABAB1-/- 1278±77; n=6/5) whereas that of Dcx+ cells is increased (control 858±69; GABAB1-/- 1477±220; n=6/5). t-test: *P<0.05. Error bars indicate s.e.m. Scale bars: A,B, 100 μm; D-G, 50 μm.

Accelerated neuronal differentiation in GABA B1-/- mice

We followed the differentiation of newborn cells in GABAB1-/- mice (Fig. 3A). Two weeks after BrdU labeling the number of newly generated cells was four times higher in GABAB1-/- mice than in control littermates, consistent with the increased proliferation seen in the SGZ (Fig. 3B,C,F). At this time point after BrdU labeling, neuronal differentiation of BrdU+ cells was apparent, with overlapping expression of Dcx and NeuN (Fig. 3D,E) (Kempermann et al., 2004). The proportion of newly generated BrdU+ NeuN+ mature granule cells was significantly increased in GABAB1-/- mice at the expense of BrdU+ neuroblasts and Dcx/NeuN double-positive immature neurons (Fig. 3G). Neurogenesis and differentiation were also enhanced in the GABAB1-/- mice after a 30-day chase (supplementary material Fig. S2A-E). Therefore, accelerated neuronal maturation, in addition to increased cell proliferation, contributes to enhanced neurogenesis in the DG of adult GABAB1-/- mice.

Fig. 3.

Increased differentiation in the hippocampus of GABAB1-deficient mice. (A) BrdU was injected intraperitoneally four times on day 0 (d0) to label newly generated cells and the mice were sacrificed (†) 15 days later (d15). (B,C) Representative images of BrdU+ cells in the hippocampus of GABAB1-/- and control mice. (D,E) Phenotypic analysis of BrdU+ cells. Dcx and NeuN were used to label neuroblasts and mature neurons, respectively. (F) The density of BrdU+ newborn cells is increased fourfold in mutant mice at d15 (control 111±25; GABAB1-/- 429±104; n=6/5). (G) Fifteen days after BrdU injection, most BrdU+ cells are differentiating neurons expressing both Dcx and NeuN. Some BrdU+ cells express only Dcx (neuroblasts) and others only NeuN (mature neurons). Note that the proportion of mature neurons is increased at the expense of neuroblasts and differentiating neurons in GABAB1 mutant mice (Dcx, control 13.6±1.3; GABAB1-/- 5.2±2; Dcx/NeuN, control 60±2.7; GABAB1-/- 45±4.9; NeuN, control 9.7±1.3; GABAB1-/- 30±4.7; n=6/5). t-test: *P<0.05. Error bars indicate s.e.m. Scale bars: B,C, 100 μm; D,E, 50 μm.

Fig. 3.

Increased differentiation in the hippocampus of GABAB1-deficient mice. (A) BrdU was injected intraperitoneally four times on day 0 (d0) to label newly generated cells and the mice were sacrificed (†) 15 days later (d15). (B,C) Representative images of BrdU+ cells in the hippocampus of GABAB1-/- and control mice. (D,E) Phenotypic analysis of BrdU+ cells. Dcx and NeuN were used to label neuroblasts and mature neurons, respectively. (F) The density of BrdU+ newborn cells is increased fourfold in mutant mice at d15 (control 111±25; GABAB1-/- 429±104; n=6/5). (G) Fifteen days after BrdU injection, most BrdU+ cells are differentiating neurons expressing both Dcx and NeuN. Some BrdU+ cells express only Dcx (neuroblasts) and others only NeuN (mature neurons). Note that the proportion of mature neurons is increased at the expense of neuroblasts and differentiating neurons in GABAB1 mutant mice (Dcx, control 13.6±1.3; GABAB1-/- 5.2±2; Dcx/NeuN, control 60±2.7; GABAB1-/- 45±4.9; NeuN, control 9.7±1.3; GABAB1-/- 30±4.7; n=6/5). t-test: *P<0.05. Error bars indicate s.e.m. Scale bars: B,C, 100 μm; D,E, 50 μm.

Unaltered cell survival in GABA B1-/- mice

We addressed whether enhanced survival, in addition to augmented proliferation and differentiation, is responsible for the increased number of newly generated granule neurons observed in GABAB1-/- mice. To analyze apoptosis, we performed TUNEL assays in GABAB1-/- and GABAB1+/+ mice and quantified TUNEL-labeled cells in the SGZ (Fig. 4A). Loss of GABAB1 receptor subunits did not affect apoptosis in the DG (Fig. 4B). As an independent measure of the survival of newly generated neurons, we calculated the fraction of BrdU+ cells at 30 days compared to 15 days after BrdU labeling. The percentage of BrdU+ cells surviving at 30 days was similar in GABAB1-/- mice and in control littermates (Fig. 4C). Thus, increased proliferation and differentiation are the main factors responsible for the neurogenic phenotype seen in GABAB1-/- mice, and conversely the survival of newborn neurons is not significantly affected.

Fig. 4.

Apoptosis and cell survival are not affected in the SGZ of GABAB1-deficient mice. (A) Representative images showing pyknotic Tunel+ cells in the SGZ and GrL of GABAB1-/- and wild-type control mice. (B) The density of Tunel+ cells in the adult hippocampus of GABAB1-/- mice is unchanged compared with control mice (control 1.4±0.3; GABAB1-/- 1.3±0.4; n=6). (C) Cell survival is not dramatically affected in GABAB1-/- mice. Survival is depicted as proportion of BrdU+ cells after a 30-day chase versus BrdU+ cells after a 15-day chase (control 41.5; GABAB1-/- 42.5; n=6/5). Error bars indicate s.e.m. Scale bar: 10 μm.

Fig. 4.

Apoptosis and cell survival are not affected in the SGZ of GABAB1-deficient mice. (A) Representative images showing pyknotic Tunel+ cells in the SGZ and GrL of GABAB1-/- and wild-type control mice. (B) The density of Tunel+ cells in the adult hippocampus of GABAB1-/- mice is unchanged compared with control mice (control 1.4±0.3; GABAB1-/- 1.3±0.4; n=6). (C) Cell survival is not dramatically affected in GABAB1-/- mice. Survival is depicted as proportion of BrdU+ cells after a 30-day chase versus BrdU+ cells after a 15-day chase (control 41.5; GABAB1-/- 42.5; n=6/5). Error bars indicate s.e.m. Scale bar: 10 μm.

Conditional deletion of GABAB1 subunit from adult neural progenitors affects neurogenesis

GABAB receptor subunits are expressed not only by NSCs and their progeny but also by other cells within the DG (Fig. 1). We inactivated GABAB1 in DG NSCs using conditional floxed GABAB1lox511/lox511 and GFAP::CreERT2 alleles (Haller et al., 2004; Hirrlinger et al., 2006) and visualized cells where Cre-recombinase had been active by following the recombination of the mR26CS-EGFP Cre-reporter allele (rGFP) (Tchorz et al., 2012). We traced newborn cells by BrdU labeling (Fig. 5A). The proportion of rGFP+ cells that incorporated BrdU was increased in GABAB1 conditional knockouts compared with controls, suggesting that loss of GABAB1 receptors induces cell proliferation (Fig. 5B-D). The proportion of rGFP+ cells that expressed neuronal markers (Dcx or NeuN) and incorporated BrdU was also increased, indicating enhanced neurogenesis (Fig. 5E).

Fig. 5.

GABAB1 deficiency cell-autonomously affects adult neurogenesis. (A) Tamoxifen (TAM) and BrdU induction regimes in GFAP::CreERT2, rGFP and GABAB1lox511/lox511 conditional knockout (cKO) mice or GABAB1lox511/+ control mice. TAM was injected once per day for five consecutive days before the mice were sacrificed (†) 35 days after the end of induction. BrdU was administered through the drinking water for 7 days starting from 2 weeks after the end of TAM induction to detect early changes after conditional deletion. (B,C) Conditional GABAB1 loss promotes proliferation (BrdU incorporation) in comparison to control mice. The majority of the BrdU+ rGFP+ cells acquired a neuronal phenotype (Dcx+ and/or NeuN+) 2 weeks after BrdU administration. (D) BrdU+ rGFP+ cells are significantly increased in GABAB1 conditional mutants (control 6.5±1.5; cKO 12±1.3; n=7/6). (E) The proportion of BrdU-labeled cells among rGFP+ neuronal cells (Dcx+ and NeuN+) also increases in GABAB1 cKO mice compared with controls (control 13±1.15; cKO 19.7±0.88; n=3). t-test: *P<0.05. Error bars indicate s.e.m. Scale bars: 10 μm.

Fig. 5.

GABAB1 deficiency cell-autonomously affects adult neurogenesis. (A) Tamoxifen (TAM) and BrdU induction regimes in GFAP::CreERT2, rGFP and GABAB1lox511/lox511 conditional knockout (cKO) mice or GABAB1lox511/+ control mice. TAM was injected once per day for five consecutive days before the mice were sacrificed (†) 35 days after the end of induction. BrdU was administered through the drinking water for 7 days starting from 2 weeks after the end of TAM induction to detect early changes after conditional deletion. (B,C) Conditional GABAB1 loss promotes proliferation (BrdU incorporation) in comparison to control mice. The majority of the BrdU+ rGFP+ cells acquired a neuronal phenotype (Dcx+ and/or NeuN+) 2 weeks after BrdU administration. (D) BrdU+ rGFP+ cells are significantly increased in GABAB1 conditional mutants (control 6.5±1.5; cKO 12±1.3; n=7/6). (E) The proportion of BrdU-labeled cells among rGFP+ neuronal cells (Dcx+ and NeuN+) also increases in GABAB1 cKO mice compared with controls (control 13±1.15; cKO 19.7±0.88; n=3). t-test: *P<0.05. Error bars indicate s.e.m. Scale bars: 10 μm.

GABAB receptor antagonist activates quiescent NSCs, whereas GABAB receptor agonist promotes NSC quiescence

Most adult hippocampal NSCs are quiescent (Kronenberg et al., 2003; Lugert et al., 2010; Bonaguidi et al., 2011; Dranovsky et al., 2011). NSC quiescence is reversible in response to a number of pathophysiological stimuli (Lugert et al., 2010; Bonaguidi et al., 2011; Dranovsky et al., 2011). Neurotransmitters can directly regulate hippocampal NSC quiescence (Jhaveri et al., 2010; Song et al., 2012). The increased progenitor proliferation in GABAB1-/- mice suggested that GABAB receptors may modulate NSCs quiescence. We inhibited GABAB receptor function by infusing the GABAB antagonist CGP54626A (CGP) intracranially for six consecutive days into Hes5::GFP mice (Fig. 6A). Proliferation (PCNA+ cells) increased dramatically in the SGZ of CGP-versus saline-treated mice (Fig. 6B,C,F) and the density of Hes5+ PCNA+ cells was increased, indicating that Hes5+ NSCs were affected (Fig. 6D,E,G). Interestingly, although the proportion of Hes5::GFP+ cells that expressed PCNA increased after CGP infusion, the density of Hes5::GFP+ cells was unchanged (Fig. 6H,I). This implied that although blocking GABAB function recruited quiescent cells to the active proliferative stem cell pool it did not induce an expansion of the stem cell population.

Fig. 6.

Infusion of GABAB antagonist activates adult hippocampal NSCs. (A) GABAB antagonist induction regime. CGP54626A (CGP) was infused for 6 days into the hippocampus of adult Hes5::GFP+ mice. The mice were sacrificed (†) at day 6 (d6). (B,C) Representative images of proliferating cells (PCNA+), neuroblasts (Dcx+) and NSCs (Hes5::GFP+) in the SGZ of CGP and control (saline) infused mice. (D,E) CGP induces Hes5::GFP+ cells to proliferate (arrows). (F) The density of PCNA+ proliferating cells is increased in CGP-treated mice (control 456±108; CGP 1268±299; n=5/7). (G) Proliferating (PCNA+) Hes5::GFP+ cells are increased in number in CGP-treated mice (control 94±8; CGP 195±32; n=5/7). (H) The Hes5::GFP+ population does not expand after CGP treatment (control 1267±74; CGP 1489±69; n=5/7; P=0.06). (I) The proportion of Hes5::GFP+ cells that proliferate (PCNA+) increases after CGP treatment (control 7.4±0.3; CGP 12.8±1.8; n=5/7). t-test: *P<0.05. Error bars indicate s.e.m. Scale bars: B,C, 50 μm; D,E, 10 μm.

Fig. 6.

Infusion of GABAB antagonist activates adult hippocampal NSCs. (A) GABAB antagonist induction regime. CGP54626A (CGP) was infused for 6 days into the hippocampus of adult Hes5::GFP+ mice. The mice were sacrificed (†) at day 6 (d6). (B,C) Representative images of proliferating cells (PCNA+), neuroblasts (Dcx+) and NSCs (Hes5::GFP+) in the SGZ of CGP and control (saline) infused mice. (D,E) CGP induces Hes5::GFP+ cells to proliferate (arrows). (F) The density of PCNA+ proliferating cells is increased in CGP-treated mice (control 456±108; CGP 1268±299; n=5/7). (G) Proliferating (PCNA+) Hes5::GFP+ cells are increased in number in CGP-treated mice (control 94±8; CGP 195±32; n=5/7). (H) The Hes5::GFP+ population does not expand after CGP treatment (control 1267±74; CGP 1489±69; n=5/7; P=0.06). (I) The proportion of Hes5::GFP+ cells that proliferate (PCNA+) increases after CGP treatment (control 7.4±0.3; CGP 12.8±1.8; n=5/7). t-test: *P<0.05. Error bars indicate s.e.m. Scale bars: B,C, 50 μm; D,E, 10 μm.

In a complementary approach, we activated GABAB receptors by intracranial infusion of the GABAB agonist baclofen for six consecutive days into Hes5::GFP mice (supplementary material Fig. S3A). The density of PCNA+ cells and Hes5+ cells did not decrease significantly in the SGZ of baclofen-versus saline-treated mice (supplementary material Fig. S3B-F). However, the proportion of Hes5::GFP+ cells that expressed PCNA decreased after baclofen infusion, suggesting that Hes5+ NSCs were preferentially affected and switched to a quiescent state (supplementary material Fig. S3G).

Much effort has been put into understanding the regulation of neurogenesis in the hippocampus of adult mammals and the functions of these newborn neurons in homeostasis and disease. However, our knowledge of how the brain coordinates network activity and the generation of new neurons is still limited. GABA released by local interneurons is a major extrinsic regulator that can profoundly affect adult hippocampal neurogenesis (Masiulis et al., 2011; Song et al., 2012). The action of GABA on neural stem and progenitor cell proliferation is complex and still controversial. GABA can promote or suppress proliferation depending on developmental stage, brain region and the fate of distinct progenitor populations (Haydar et al., 2000; Liu et al., 2005; Duveau et al., 2011). In the adult hippocampus, ionotropic GABAA receptors have been reported to decrease cell proliferation (Duveau et al., 2011; Song et al., 2012). It remains unclear whether differential regulation occurs at the level of intermediate progenitors and neuroblasts (Tozuka et al., 2005; Ge et al., 2006) versus NSCs (Wang et al., 2005; Song et al., 2012). Moreover, although ionotropic GABAA receptors mediate most of the GABA effects on adult neurogenesis described to date, little is known of the function of GABAB receptors in this context (Felice et al., 2012). We provide evidence that metabotropic GABAB receptors may directly suppress NSC proliferation and neuroblast differentiation in the adult hippocampus.

Our results show that GABA signaling through GABAB receptors inhibits DG NSC proliferation. We propose that this inhibition is, at least in part, a direct effect of GABAB signaling in NSCs. Neurotransmitters may mediate crosstalk between newly generated cells and the surrounding neuronal network, thereby matching neural activity with neurogenic output (Masiulis et al., 2011). Signaling via GABAB receptors is a novel regulator that may contribute to coordinate hippocampal network activity and NSC proliferation. Understanding the molecular mechanisms regulating proliferation versus quiescence of adult NSCs is crucial. NSCs become mostly quiescent during aging, and this correlates with a dramatic reduction in neurogenesis with age (Hattiangady and Shetty, 2008; Jessberger and Gage, 2008; Lugert et al., 2010). However, NSC quiescence is reversible, and this could be exploited to rejuvenate neurogenesis in the aged or damaged brain (Hattiangady and Shetty, 2008; Lugert et al., 2010). Importantly, excitation as well as specific neurotransmitters can activate the latent stem cell pool (Jhaveri et al., 2010; Lugert et al., 2010), and here we propose that GABAB receptors can contribute to this process. Therefore, manipulation of GABAB function may be a novel approach to modulate adult hippocampal neurogenesis in vivo and during aging. Recently, GABAB receptors have attracted attention as potentially being involved in the etiology of depression, and GABAB blockade causes antidepressant-like effects (Cryan and Slattery, 2010). Given that antidepressant drugs can promote adult neurogenesis and new hippocampal neurons have been implicated in mediating some effects of antidepressants (Petrik et al., 2012), our findings are relevant for human disease. Indeed, increased proliferation in the ventral hippocampus has been suggested as a plausible mechanism for the antidepressant-like effects of chronic treatment with GABAB receptor antagonists (Felice et al., 2012).

Together, our data suggest that metabotropic GABAB receptors are already active in the cells at the start of the adult neurogenic lineage. This may represent a novel mechanism to integrate hippocampal network activity, GABA release and NSC proliferation. Based on continued expression of GABAB subunits in more differentiated cell types, further regulation by GABAB may occur downstream of NSCs during adult hippocampal neurogenesis. Indeed, our results show that differentiation of neuroblasts is accelerated in mice lacking the GABAB1 receptor subunits, without there being a significant effect on newborn neuron survival. Notably, and in contrast to the action of the GABAB receptors, activation of GABAA receptors promotes differentiation along the neuronal lineage, survival of new neurons as well as asynaptic integration in the adult DG (Tozuka et al., 2005; Ge et al., 2006; Jagasia et al., 2009). Therefore, GABAB receptors can potentially synergize with GABAA receptors to inhibit NSC division (Song et al., 2012) and counteract the differentiation-promoting effects of GABAA receptors later within the neurogenic lineage (Tozuka et al., 2005; Ge et al., 2006).

Little is known about the molecular mechanisms and signaling pathways that mediate the effects of neurotransmitters on adult NSCs and their progeny. GABAB receptors can modulate ion channels opening at the plasma membrane (Ulrich and Bettler, 2007). Postsynaptic GABAB1b-containing receptors activate K+ channels. In contrast to the K-current effects of GABAB receptors on neurons, hippocampal NSCs showed leaky membrane currents and their K-currents were not dramatically affected by pharmacological activation of GABAB receptors (data not shown) (Filippov et al., 2003). Thus, we suggest that GABAB-induced hyperpolarization is unlikely to be the main mechanism that mediates the inhibitory action of GABAB receptors on progenitor proliferation, but this will require closer scrutiny in the future. In addition to modulating ion channels, GABAB receptors can inhibit adenylate-cyclase activity (Kaupmann et al., 1997; Kuner et al., 1999; Martin et al., 1999). Activation of Beta3-adrenergic receptors, which positively regulate the adenylate cyclase via G-protein coupling and are specifically expressed by Hes5+ NSCs in the SGZ, induces cell proliferation in the adult DG (Ursino et al., 2009; Jhaveri et al., 2010). The adenylate-cyclase-cAMP-CREB axis is also a key signal transduction pathway that promotes neuronal differentiation in the DG (Palmer et al., 1997; Fujioka et al., 2004) and is potentiated by GABAA-mediated depolarization in SGZ neuroblasts (Jagasia et al., 2009). Thus, released inhibition of the adenylate cyclase may contribute to increased neurogenesis in the GABAB1-deficient mouse hippocampus by counteracting the effects of Beta3-adrenergic receptors and GABAA receptors in NSCs and neuroblasts, respectively. Future studies will need to address a potential role for second-messenger regulation by GABAB receptors in adult neurogenesis.

Animals and husbandry

GABAB1-/-, GABAB2-/-, GABAB1lox511/lox511, GABAB1-GFP, GFAP::CreERT2, mR26CS-EGFP and Hes5::GFP mice have been described elsewhere (Schuler et al., 2001; Gassmann et al., 2004; Haller et al., 2004; Hirrlinger et al., 2006; Vigot et al., 2006; Casanova et al., 2009; Lugert et al., 2010; Tchorz et al., 2012). Mice were maintained on a 12-hour day/night cycle with adequate food and water under specific-pathogen-free (SPF) conditions according to institutional regulations and under license numbers 35/9185.81/G-09/19 (Ethical Commission Freiburg, Germany) and 2537 and 2538 (Kantonales Veterinäramt, Basel).

BrdU and tamoxifen administration

Young adult mice (7-8 weeks old) received four consecutive intraperitoneal injections (every 2 hours) of BrdU (Sigma; 50 mg/kg body weight). Alternatively, BrdU was given to the mice for seven consecutive days dissolved in the drinking water at 0.8 mg/ml. Stock solution of tamoxifen (TAM, Sigma) were prepared at a concentration of 20 mg/ml in corn oil (Sigma). Adult mice were injected intraperitoneally with TAM once per day for five consecutive days at a dose of 2 mg per day.

CGP and baclofen infusion

Adult (2 months old) Hes5::GFP mice were anesthetized by intraperitoneal injection of a ketamine/xylazine/flunitrazepam solution (100 mg, 5 and 0.4 mg/kg body weight, respectively) and positioned in a stereotaxic apparatus (David Kopf Instruments). The skull was exposed by an incision in the scalp and a small hole (1 mm) was drilled through. Cannulas (Brain Infusion Kit 3, Alzet) were implanted at -2 mm posterior, 1.5 mm lateral to the bregma and 2 mm below the surface of the cortex to target the dorsal aspect of the anterior DG. CGP54626A (CGP, Tocris Bioscience; 500 μM in 0.9% saline), baclofen (Tocris Bioscience; 1 mM in 0.9% saline) or vehicle alone was infused for 6 days into the brain with an osmotic pump (model 1007D, Alzet). After 6 days of infusion the animals were sacrificed and analyzed. Brains were processed for immunohistochemistry as described below.

Tissue preparation, immunohistochemistry and antibodies

Mice were deeply anesthetized by injection of a ketamine/xylazine/flunitrazepam solution (150 mg, 7.5 and 0.6 mg/kg body weight, respectively) and perfused with ice-cold 0.9% saline solution followed by ice-cold 4% paraformaldehyde (PFA) solution in 0.1 M phosphate buffer (PB). Brains were post-fixed with 4% PFA overnight, washed in PB, cryoprotected in a 30% sucrose solution in 0.1 M PB for 48 hours, frozen and sectioned at -20°C. Free-floating coronal sections (30 μm) were collected in multiwell dishes (Corning) and stored at -20°C in antifreeze solution until use.

For immunostaining, sections were incubated overnight at 4°C with the primary antibody diluted in blocking solution of 2% normal donkey serum (Jackson ImmunoResearch) 0.5% Triton X-100 in phosphate-buffered saline (PBS). Sections were washed three times in PBS and incubated at room temperature for 1 hour with the corresponding secondary antibodies in blocking solution. When necessary, sections were washed and incubated for 1 hour at room temperature in streptavidin fluorescein isothiocyanate (FITC; Jackson ImmunoResearch; 1:400). Sections were mounted on Superfrost glass slides (Thermo Scientific), embedded in mounting medium containing 1,4-diazabicyclo[2.2.2]octane (DABCO; Sigma) as an antifading agent and visualized using a Zeiss LSM510 confocal microscope. For the avidin-biotin-peroxidase method, sections were washed in PBS after incubation with secondary biotinylated antibody and then incubated for 1 hour at room temperature in peroxidase-conjugated streptavidin (Jackson ImmunoResearch; 1:1000). Sections were incubated with 0.015% 3,3′-diaminobenzidine, 0.0024% H2O2 in 0.05 M Tris-HCl, pH 7.6. Sections were mounted on glass slides (Thermo Scientific), dehydrated and embedded in DePeX mounting medium (SERVA, Heidelberg, Germany).

Antibodies were used against the following antigens: NeuN (mouse, Sigma; 1:800); Calbindin D28k (mouse, Swant; 1:2000); Sox2 (rabbit, Chemicon; 1:1000); Sox2 (goat, Santa Cruz; 1:200); BLBP (rabbit, Chemicon; 1:1500); PCNA (mouse, Dako; 1:1000); pH3 (rabbit, Millipore; 1:100); BrdU (rat, AbD Serotec; 1:2000); Doublecortin (goat, Santa Cruz; 1:500); PSA-NCAM (mouse, Chemicon; 1:2000); GFP (sheep, Biogenesis; 1:500); GFP (rabbit, Invitrogen; 1:500); GABAB1 (mouse, Abcam; 1:300); GABAB1 (rabbit 174.1; 1:300) (Malitschek et al., 1998); GABAB1 (rabbit AB25; 1:1000) (Engle et al., 2006); GABAB2 receptor (rabbit AB27; 1:1000; generated against a glutathione-S-transferase fusion protein containing carboxyterminal residues T746-L941 of rat GABAB2 protein); Cy3/Cy5/biotin conjugated anti-mouse, rabbit, rat and goat immunoglobulins (donkey, Jackson ImmunoResearch; 1:500-1000).

TUNEL staining

Sections were washed in PBS for 10 minutes and blocked for 1 hour with 10% goat serum, 1% Triton X-100, and 0.1% bovine serum albumin (BSA) in PBS. Terminal deoxynucleotidyl transferase mediated biotinylated UTP nick end labeling (TUNEL) assays were performed according to the manufacturer’s instructions (Roche).

Quantification and statistical analyses

Immunostained hippocampal sections were analyzed on a Zeiss LSM510 confocal microscope. Data are presented as average percentages of co-labeled cells. The number of marker-positive cells in the SGZ was estimated using a 63× magnification objective. The area of the GrL was measured using ImageJ software and used to calculate the number of labeled cells per mm2. Statistical comparisons were conducted by two-tailed unpaired Student’s t-test. Significance was established at P<0.05. In all graphs error bars represent standard error of the mean (s.e.m.).

We thank Dr Sebastian Lugert for comments and Frank Sager for technical assistance.

Author contributions

C.G. carried out most of the experiments and generated the figures. M.B., J.S.T., M.T. and M.G. contributed to the analysis of the GABAB mutant mice, performed electrophysiology and were involved in the preparation of the manuscript. V.T., C.G., J.B. and B.B. conceived the project, designed the experiments and wrote the manuscript.

Funding

This work was supported by the Deutsche Forschungsgemeinschaft [DFG SFB592; TA-310-1; TA-310-2] and the Max Planck Society. We acknowledge the support of the Swiss Science Foundation [31003A-133124 and CRSII3_136210] the National Center of Competences in Research (NCCR) ‘Synapsy, Synaptic Bases of Mental Diseases’ and the European Community’s 7th Framework Program [FP7/2007-2013] under Grant Agreement 201714 to B.B.

Ables
J. L.
,
Decarolis
N. A.
,
Johnson
M. A.
,
Rivera
P. D.
,
Gao
Z.
,
Cooper
D. C.
,
Radtke
F.
,
Hsieh
J.
,
Eisch
A. J.
(
2010
).
Notch1 is required for maintenance of the reservoir of adult hippocampal stem cells
.
J. Neurosci.
30
,
10484
10492
.
Bonaguidi
M. A.
,
Wheeler
M. A.
,
Shapiro
J. S.
,
Stadel
R. P.
,
Sun
G. J.
,
Ming
G. L.
,
Song
H.
(
2011
).
In vivo clonal analysis reveals self-renewing and multipotent adult neural stem cell characteristics
.
Cell
145
,
1142
1155
.
Breunig
J. J.
,
Silbereis
J.
,
Vaccarino
F. M.
,
Sestan
N.
,
Rakic
P.
(
2007
).
Notch regulates cell fate and dendrite morphology of newborn neurons in the postnatal dentate gyrus
.
Proc. Natl. Acad. Sci. USA
104
,
20558
20563
.
Casanova
E.
,
Guetg
N.
,
Vigot
R.
,
Seddik
R.
,
Julio-Pieper
M.
,
Hyland
N. P.
,
Cryan
J. F.
,
Gassmann
M.
,
Bettler
B.
(
2009
).
A mouse model for visualization of GABA(B) receptors
.
Genesis
47
,
595
602
.
Cryan
J. F.
,
Slattery
D. A.
(
2010
).
GABAB receptors and depression. Current status
.
Adv. Pharmacol.
58
,
427
451
.
Deisseroth
K.
,
Singla
S.
,
Toda
H.
,
Monje
M.
,
Palmer
T. D.
,
Malenka
R. C.
(
2004
).
Excitation-neurogenesis coupling in adult neural stem/progenitor cells
.
Neuron
42
,
535
552
.
Dranovsky
A.
,
Picchini
A. M.
,
Moadel
T.
,
Sisti
A. C.
,
Yamada
A.
,
Kimura
S.
,
Leonardo
E. D.
,
Hen
R.
(
2011
).
Experience dictates stem cell fate in the adult hippocampus
.
Neuron
70
,
908
923
.
Duveau
V.
,
Laustela
S.
,
Barth
L.
,
Gianolini
F.
,
Vogt
K. E.
,
Keist
R.
,
Chandra
D.
,
Homanics
G. E.
,
Rudolph
U.
,
Fritschy
J. M.
(
2011
).
Spatiotemporal specificity of GABAA receptor-mediated regulation of adult hippocampal neurogenesis
.
Eur. J. Neurosci.
34
,
362
373
.
Ehm
O.
,
Göritz
C.
,
Covic
M.
,
Schäffner
I.
,
Schwarz
T. J.
,
Karaca
E.
,
Kempkes
B.
,
Kremmer
E.
,
Pfrieger
F. W.
,
Espinosa
L.
, et al. 
. (
2010
).
RBPJkappa-dependent signaling is essential for long-term maintenance of neural stem cells in the adult hippocampus
.
J. Neurosci.
30
,
13794
13807
.
Engle
M. P.
,
Gassman
M.
,
Sykes
K. T.
,
Bettler
B.
,
Hammond
D. L.
(
2006
).
Spinal nerve ligation does not alter the expression or function of GABA(B) receptors in spinal cord and dorsal root ganglia of the rat
.
Neuroscience
138
,
1277
1287
.
Felice
D.
,
O’Leary
O. F.
,
Pizzo
R. C.
,
Cryan
J. F.
(
2012
).
Blockade of the GABA(B) receptor increases neurogenesis in the ventral but not dorsal adult hippocampus: relevance to antidepressant action
.
Neuropharmacology
63
,
1380
1388
.
Filippov
V.
,
Kronenberg
G.
,
Pivneva
T.
,
Reuter
K.
,
Steiner
B.
,
Wang
L. P.
,
Yamaguchi
M.
,
Kettenmann
H.
,
Kempermann
G.
(
2003
).
Subpopulation of nestin-expressing progenitor cells in the adult murine hippocampus shows electrophysiological and morphological characteristics of astrocytes
.
Mol. Cell. Neurosci.
23
,
373
382
.
Fujioka
T.
,
Fujioka
A.
,
Duman
R. S.
(
2004
).
Activation of cAMP signaling facilitates the morphological maturation of newborn neurons in adult hippocampus
.
Neuroscience
24
,
319
328
.
Fukui
M.
,
Nakamichi
N.
,
Yoneyama
M.
,
Ozawa
S.
,
Fujimori
S.
,
Takahata
Y.
,
Nakamura
N.
,
Taniura
H.
,
Yoneda
Y.
(
2008
).
Modulation of cellular proliferation and differentiation through GABA(B) receptors expressed by undifferentiated neural progenitor cells isolated from fetal mouse brain
.
J. Cell. Physiol.
216
,
507
519
.
Gassmann
M.
,
Shaban
H.
,
Vigot
R.
,
Sansig
G.
,
Haller
C.
,
Barbieri
S.
,
Humeau
Y.
,
Schuler
V.
,
Müller
M.
,
Kinzel
B.
, et al. 
. (
2004
).
Redistribution of GABAB(1) protein and atypical GABAB responses in GABAB(2)-deficient mice
.
J. Neurosci.
24
,
6086
6097
.
Ge
S.
,
Goh
E. L.
,
Sailor
K. A.
,
Kitabatake
Y.
,
Ming
G. L.
,
Song
H.
(
2006
).
GABA regulates synaptic integration of newly generated neurons in the adult brain
.
Nature
439
,
589
593
.
Haller
C.
,
Casanova
E.
,
Müller
M.
,
Vacher
C. M.
,
Vigot
R.
,
Doll
T.
,
Barbieri
S.
,
Gassmann
M.
,
Bettler
B.
(
2004
).
Floxed allele for conditional inactivation of the GABAB(1) gene
.
Genesis
40
,
125
130
.
Hattiangady
B.
,
Shetty
A. K.
(
2008
).
Aging does not alter the number or phenotype of putative stem/progenitor cells in the neurogenic region of the hippocampus
.
Neurobiol. Aging
29
,
129
147
.
Haydar
T. F.
,
Wang
F.
,
Schwartz
M. L.
,
Rakic
P.
(
2000
).
Differential modulation of proliferation in the neocortical ventricular and subventricular zones
.
J. Neurosci.
20
,
5764
5774
.
Hirrlinger
P. G.
,
Scheller
A.
,
Braun
C.
,
Hirrlinger
J.
,
Kirchhoff
F.
(
2006
).
Temporal control of gene recombination in astrocytes by transgenic expression of the tamoxifen-inducible DNA recombinase variant CreERT2
.
Glia
54
,
11
20
.
Jagasia
R.
,
Steib
K.
,
Englberger
E.
,
Herold
S.
,
Faus-Kessler
T.
,
Saxe
M.
,
Gage
F. H.
,
Song
H.
,
Lie
D. C.
(
2009
).
GABA-cAMP response element-binding protein signaling regulates maturation and survival of newly generated neurons in the adult hippocampus
.
J. Neurosci.
29
,
7966
7977
.
Jessberger
S.
,
Gage
F. H.
(
2008
).
Stem-cell-associated structural and functional plasticity in the aging hippocampus
.
Psychol. Aging
23
,
684
691
.
Jhaveri
D. J.
,
Mackay
E. W.
,
Hamlin
A. S.
,
Marathe
S. V.
,
Nandam
L. S.
,
Vaidya
V. A.
,
Bartlett
P. F.
(
2010
).
Norepinephrine directly activates adult hippocampal precursors via beta3-adrenergic receptors
.
J. Neurosci.
30
,
2795
2806
.
Kaupmann
K.
,
Huggel
K.
,
Heid
J.
,
Flor
P. J.
,
Bischoff
S.
,
Mickel
S. J.
,
McMaster
G.
,
Angst
C.
,
Bittiger
H.
,
Froestl
W.
, et al. 
. (
1997
).
Expression cloning of GABA(B) receptors uncovers similarity to metabotropic glutamate receptors
.
Nature
386
,
239
246
.
Kempermann
G.
,
Jessberger
S.
,
Steiner
B.
,
Kronenberg
G.
(
2004
).
Milestones of neuronal development in the adult hippocampus
.
Trends Neurosci.
27
,
447
452
.
Kronenberg
G.
,
Reuter
K.
,
Steiner
B.
,
Brandt
M. D.
,
Jessberger
S.
,
Yamaguchi
M.
,
Kempermann
G.
(
2003
).
Subpopulations of proliferating cells of the adult hippocampus respond differently to physiologic neurogenic stimuli
.
J. Comp. Neurol.
467
,
455
463
.
Kuner
R.
,
Köhr
G.
,
Grünewald
S.
,
Eisenhardt
G.
,
Bach
A.
,
Kornau
H. C.
(
1999
).
Role of heteromer formation in GABAB receptor function
.
Science
283
,
74
77
.
Liu
X.
,
Wang
Q.
,
Haydar
T. F.
,
Bordey
A.
(
2005
).
Nonsynaptic GABA signaling in postnatal subventricular zone controls proliferation of GFAP-expressing progenitors
.
Nat. Neurosci.
8
,
1179
1187
.
Lugert
S.
,
Basak
O.
,
Knuckles
P.
,
Haussler
U.
,
Fabel
K.
,
Götz
M.
,
Haas
C. A.
,
Kempermann
G.
,
Taylor
V.
,
Giachino
C.
(
2010
).
Quiescent and active hippocampal neural stem cells with distinct morphologies respond selectively to physiological and pathological stimuli and aging
.
Cell Stem Cell
6
,
445
456
.
Lugert
S.
,
Vogt
M.
,
Tchorz
J. S.
,
Müller
M.
,
Giachino
C.
,
Taylor
V.
(
2012
).
Homeostatic neurogenesis in the adult hippocampus does not involve amplification of Ascl1(high) intermediate progenitors
.
Nat. Commun.
3
,
670
.
Malitschek
B.
,
Rüegg
D.
,
Heid
J.
,
Kaupmann
K.
,
Bittiger
H.
,
Fröstl
W.
,
Bettler
B.
,
Kuhn
R.
(
1998
).
Developmental changes of agonist affinity at GABABR1 receptor variants in rat brain
.
Mol. Cell. Neurosci.
12
,
56
64
.
Martin
S. C.
,
Russek
S. J.
,
Farb
D. H.
(
1999
).
Molecular identification of the human GABABR2: cell surface expression and coupling to adenylyl cyclase in the absence of GABABR1
.
Mol. Cell. Neurosci.
13
,
180
191
.
Masiulis
I.
,
Yun
S.
,
Eisch
A. J.
(
2011
).
The interesting interplay between interneurons and adult hippocampal neurogenesis
.
Mol. Neurobiol.
44
,
287
302
.
Palmer
T. D.
,
Takahashi
J.
,
Gage
F. H.
(
1997
).
The adult rat hippocampus contains primordial neural stem cells
.
Mol. Cell. Neurosci.
8
,
389
404
.
Parent
J. M.
(
2007
).
Adult neurogenesis in the intact and epileptic dentate gyrus
.
Prog. Brain Res.
163
,
529
540
,
817
.
Pérez-Garci
E.
,
Gassmann
M.
,
Bettler
B.
,
Larkum
M. E.
(
2006
).
The GABAB1b isoform mediates long-lasting inhibition of dendritic Ca2+ spikes in layer 5 somatosensory pyramidal neurons
.
Neuron
50
,
603
616
.
Petrik
D.
,
Lagace
D. C.
,
Eisch
A. J.
(
2012
).
The neurogenesis hypothesis of affective and anxiety disorders: are we mistaking the scaffolding for the building?
Neuropharmacology
62
,
21
34
.
Salazar
P.
,
Velasco-Velázquez
M. A.
,
Velasco
I.
(
2008
).
GABA effects during neuronal differentiation of stem cells
.
Neurochem. Res.
33
,
1546
1557
.
Schuler
V.
,
Lüscher
C.
,
Blanchet
C.
,
Klix
N.
,
Sansig
G.
,
Klebs
K.
,
Schmutz
M.
,
Heid
J.
,
Gentry
C.
,
Urban
L.
, et al. 
. (
2001
).
Epilepsy, hyperalgesia, impaired memory, and loss of pre- and postsynaptic GABA(B) responses in mice lacking GABA(B(1))
.
Neuron
31
,
47
58
.
Song
J.
,
Zhong
C.
,
Bonaguidi
M. A.
,
Sun
G. J.
,
Hsu
D.
,
Gu
Y.
,
Meletis
K.
,
Huang
Z. J.
,
Ge
S.
,
Enikolopov
G.
, et al. 
. (
2012
).
Neuronal circuitry mechanism regulating adult quiescent neural stem-cell fate decision
.
Nature
489
,
150
154
.
Tchorz
J. S.
,
Suply
T.
,
Ksiazek
I.
,
Giachino
C.
,
Cloëtta
D.
,
Danzer
C. P.
,
Doll
T.
,
Isken
A.
,
Lemaistre
M.
,
Taylor
V.
, et al. 
. (
2012
).
A modified RMCE-compatible Rosa26 locus for the expression of transgenes from exogenous promoters
.
PLoS ONE
7
,
e30011
.
Tozuka
Y.
,
Fukuda
S.
,
Namba
T.
,
Seki
T.
,
Hisatsune
T.
(
2005
).
GABAergic excitation promotes neuronal differentiation in adult hippocampal progenitor cells
.
Neuron
47
,
803
815
.
Ulrich
D.
,
Bettler
B.
(
2007
).
GABA(B) receptors: synaptic functions and mechanisms of diversity
.
Curr. Opin. Neurobiol.
17
,
298
303
.
Ursino
M. G.
,
Vasina
V.
,
Raschi
E.
,
Crema
F.
,
De Ponti
F.
(
2009
).
The beta3-adrenoceptor as a therapeutic target: current perspectives
.
Pharmacol. Res.
59
,
221
234
.
Vigot
R.
,
Barbieri
S.
,
Bräuner-Osborne
H.
,
Turecek
R.
,
Shigemoto
R.
,
Zhang
Y. P.
,
Luján
R.
,
Jacobson
L. H.
,
Biermann
B.
,
Fritschy
J. M.
, et al. 
. (
2006
).
Differential compartmentalization and distinct functions of GABAB receptor variants
.
Neuron
50
,
589
601
.
Wang
D. D.
,
Kriegstein
A. R.
(
2009
).
Defining the role of GABA in cortical development
.
J. Physiol.
587
,
1873
1879
.
Wang
L. P.
,
Kempermann
G.
,
Kettenmann
H.
(
2005
).
A subpopulation of precursor cells in the mouse dentate gyrus receives synaptic GABAergic input
.
Mol. Cell. Neurosci.
29
,
181
189
.

Competing interests

The authors declare no competing financial interests.

Supplementary information