Skip to main content

Main menu

  • HOME
  • CONTENT
    • Early Release
    • Featured
    • Current Issue
    • Issue Archive
    • Blog
    • Collections
    • Podcast
  • TOPICS
    • Cognition and Behavior
    • Development
    • Disorders of the Nervous System
    • History, Teaching and Public Awareness
    • Integrative Systems
    • Neuronal Excitability
    • Novel Tools and Methods
    • Sensory and Motor Systems
  • ALERTS
  • FOR AUTHORS
  • ABOUT
    • Overview
    • Editorial Board
    • For the Media
    • Privacy Policy
    • Contact Us
    • Feedback
  • SUBMIT

User menu

Search

  • Advanced search
eNeuro

eNeuro

Advanced Search

 

  • HOME
  • CONTENT
    • Early Release
    • Featured
    • Current Issue
    • Issue Archive
    • Blog
    • Collections
    • Podcast
  • TOPICS
    • Cognition and Behavior
    • Development
    • Disorders of the Nervous System
    • History, Teaching and Public Awareness
    • Integrative Systems
    • Neuronal Excitability
    • Novel Tools and Methods
    • Sensory and Motor Systems
  • ALERTS
  • FOR AUTHORS
  • ABOUT
    • Overview
    • Editorial Board
    • For the Media
    • Privacy Policy
    • Contact Us
    • Feedback
  • SUBMIT
PreviousNext
Research ArticleNegative Results, Development

Development of GABAergic Inputs Is Not Altered in Early Maturation of Adult Born Dentate Granule Neurons in Fragile X Mice

Christine L. Remmers and Anis Contractor
eNeuro 19 November 2018, 5 (6) ENEURO.0137-18.2018; DOI: https://doi.org/10.1523/ENEURO.0137-18.2018
Christine L. Remmers
1Department of Physiology, Feinberg School of Medicine, Northwestern University, Chicago, IL 60611
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Anis Contractor
1Department of Physiology, Feinberg School of Medicine, Northwestern University, Chicago, IL 60611
2Department of Neurobiology, Weinberg College of Arts and Sciences, Northwestern University, Evanston, IL 60208
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • ORCID record for Anis Contractor
  • Article
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF
Loading

Abstract

Fragile X syndrome (FXS) is the most common form of inherited mental retardation and the most common known cause of autism. Loss of fragile X mental retardation protein (FMRP) in mice (Fmr1 KO) leads to altered synaptic and circuit maturation in the hippocampus that is correlated with alterations in hippocampal-dependent behaviors. Previous studies have demonstrated that loss of FMRP increased the rate of proliferation of progenitor cells and altered their fate specification in adult Fmr1 KO mice. While these studies clearly demonstrate a role for FMRP in adult neurogenesis in the hippocampus, it is not known whether the functional synaptic maturation and integration of adult-born dentate granule cells (abDGCs) into hippocampal circuits is affected in Fmr1 KO mice. Here, we used retroviral labeling to birthdate abDGCs in Fmr1 KO mice which allowed us to perform targeted patch clamp recording to measure the development of synaptic inputs to these neurons at precise time points after differentiation. The frequency and amplitude of spontaneous GABAergic events increased over the first three weeks after differentiation; however, this normal development of GABAergic synapses was not altered in Fmr1 KO mice. Furthermore, the relatively depolarized GABA reversal potential (EGABA) in immature abDGCs was unaffected by loss of FMRP as was the development of dendritic arbor of the adult generated neurons. These studies systematically characterized the functional development of abDGCs during the first four weeks after differentiation and demonstrate that the maturation of GABAergic synaptic inputs to these neurons is not grossly affected by the loss of FMRP.

  • adult born neurons
  • fragile X
  • GABA
  • neurogenesis
  • synapse

Significance Statement

Loss of fragile X mental retardation protein (FMRP) causes Fragile X syndrome (FXS), a devastating neurodevelopmental disorder that causes multiple alterations in the development of synapses and neurons. Previous studies have described a role for FMRP in neurogenesis and hippocampal-dependent conditioning tasks linked to neurogenesis. This study systematically assessed the functional development of GABAergic inputs to adult born dentate granule cells (abDGCs) during the first four weeks after differentiation of adult neural stem cells in Fmr1 KO mice.

Introduction

Fragile X syndrome (FXS) is the most common known genetic cause of autism and intellectual disability. FXS is caused by the expansion of the CGG repeat in the 5’ UTR of the Fmr1 gene, which leads to hypermethylation, transcriptional silencing, and loss of expression of the protein product, fragile X mental retardation protein (FMRP; Willemsen et al., 2011). FMRP is an RNA-binding protein that regulates a large number of mRNAs, many of which encode synaptic proteins. Dysregulated expression of synaptic proteins is thought to perturb synapse maturation and plasticity; however, the specific mechanisms underlying synaptic and cognitive deficits in FXS remain unclear. There are a growing number of studies demonstrating that FMRP plays important roles in stem cells, including adult neural stem cells in the neurogenic niche in the subgranular zone (SGZ) of the hippocampus (Li and Zhao, 2014). The importance of adult-born dentate granule cells (abDGCs) to hippocampal function, memory processes, and potentially to several neuropsychiatric disorders (Gonçalves et al., 2016) has raised the possibility that alterations in neural stem cell proliferation and maturation and integration of abDGCs contribute to the pathology of FXS. Proliferation of adult neural progenitors is enhanced in mice lacking FMRP (Fmr1 KO; Luo et al., 2010; but also see Eadie et al., 2009), and fate specification is altered with fewer neural progenitors differentiating into neurons (Luo et al., 2010). In addition, it has been reported that reduced survival of adult-born neurons leads to an overall reduction in the number of DGCs in Fmr1 KO mice (Lazarov et al., 2012). Therefore, the evidence suggests that the number of abDGCs is reduced in Fmr1 KO mice and that FMRP ablation specifically in adult neural stem cells results in cell autonomous effects on proliferation, fate specification, and hippocampal-dependent behaviors (Guo et al., 2011). Despite these known disruptions in neurogenesis in Fmr1 KO mice, an outstanding question is whether FMRP loss also affects how abDGCs mature and integrate into the hippocampal network after differentiation.

In many neuronal types, FMRP and related proteins play a role in neuronal development and synaptic function. There are alterations in the development of synapses particularly during early developmental critical periods in the cortex in Fmr1 KO mice (Nimchinsky et al., 2001; Cruz-Martín et al., 2010; Harlow et al., 2010). In several instances, these morphologic and functional changes normalize during development (Bureau et al., 2008); however, phenotypes associated with these circuits persist in adult Fmr1 KO mice (Arnett et al., 2014; He et al., 2017). abDGCs also undergo critical periods of development during the first few weeks after differentiation when they are undergoing synapse formation and dendritic remodeling (Bergami et al., 2015) and have elevated plasticity (Ge et al., 2007b). However, it remains unknown whether loss of FMRP results in disruptions in synapse formation on abDGCs during this early developmental postmitotic period when they are actively integrating into the hippocampal network.

GABA synapses are the first inputs to form onto abDGCs and these synapses are initially excitatory due to a depolarized GABA reversal potential (EGABA) that results from a relatively high intracellular chloride concentration in young neurons (Overstreet Wadiche et al., 2005; Ge et al., 2006). These earliest GABA inputs are important for survival, dendritic development, and subsequent formation and unsilencing of glutamatergic synapses (Ge et al., 2006; Chancey et al., 2013; Song et al., 2013; Alvarez et al., 2016). The distinct characteristics of immature abDGCs endow these cells with properties that create unique roles for them in information processing (Ge et al., 2007a). Ablation of adult neurogenesis leads to impaired trace and contextual fear conditioning and pattern separation, indicating that newly generated neurons play a distinct role in learning that cannot be replicated by mature or developmentally born DGCs (Shors et al., 2001; Clelland et al., 2009; Nakashiba et al., 2012). Interestingly, these same behaviors that have been shown to require intact adult hippocampal neurogenesis are impaired in the Fmr1 KO. Fmr1 KO mice have deficits in trace fear conditioning (Zhao et al., 2005), and this phenotype can be recapitulated by specific deletion of Fmr1 in adult neural stem cells (Guo et al., 2011). Therefore, while it is known that there are reductions in the number of abDGCs and correlated changes in behavior, it remains to be determined whether loss of FMRP results in changes in synaptic formation and maturation in abDGCs, which would alter their integration and activity in the hippocampal circuit.

Given the critical role of GABA signaling in early postmitotic maturation of abDGCs (Ge et al., 2006), and also the known alterations in GABA signaling during early postnatal development (He et al., 2014; Nomura et al., 2017) and in juvenile Fmr1 KO mice (Paluszkiewicz et al., 2011; Martin et al., 2014; Zhang et al., 2017), we investigated whether the time course of development of inhibition onto abDGCs is altered in Fmr1 KO mice.

Materials and Methods

Animals

All procedures related to the care and treatment of animals were performed in accordance with the Northwestern University Animal Care Committee’s regulations. Fmr1 KO mice (C57Bl/6) were maintained by breeding heterozygous females with WT or KO males. All experiments were performed blind to genotype in age-matched male littermates. Tail biopsies were used to perform post hoc genotyping of all mice used in the study.

Retroviral birthdating

A replication incompetent retrovirus based on Moloney murine leukemia virus (MMLV) expressing RFP was prepared as described (Tashiro et al., 2006, 2015). Briefly, HEK-293 cells stably expressing GP2 were cotransfected with RFP and VSVG using Lipofectamine 2000. The media were collected from transfected cells 3 and 6 d after transfection, filtered, and centrifuged at 25,000 rpm to precipitate the virus. Eight- to 10-week-old Fmr1 WT or KO males were anesthetized using ketamine/xylazine and 1 μl of virus was injected bilaterally into the SGZ of the dentate gyrus at a rate of ∼0.3 μl/min.

Slice preparation and electrophysiology

We prepared 250-μm coronal slices at 14, 21, and 28 (±1) days postinjection (dpi). Slices were prepared using a Leica Vibratome in ice-cold high-sucrose artificial CSF (ACSF) containing the following: 85 mM NaCl, 2.5 mM KCl, 1.25 mM NaH2PO4, 25 mM NaHCO3, 25 mM glucose, 75 mM sucrose, 0.5 mM CaCl2, and 4 mM MgCl2, equilibrated with 95% O2 and 5% CO2 and including 10 μM DL-APV and 100 μM kynurenate. Slices were heated to 28°C in the same sucrose-ACSF, then the sucrose solution was gradually exchanged for recovery ACSF containing the following: 125 mM NaCl, 2.4 mM KCl, 1.2 mM NaH2PO4, 25 mM NaHCO3, 25 mM glucose, 1 mM CaCl2, 2 mM MgCl2, 0.01 mM dL-APV, and 0.1 mM kynurenic acid.

After a 60-min recovery, individual slices were transferred to a recording chamber and continuously perfused with oxygenated ACSF containing 2 mM CaCl2 and 1 mM MgCl2 at an elevated temperature of 32°C. The dentate gyrus was visually identified and targeted recordings were made from RFP-expressing dentate granule cells. Recording electrodes were manufactured from borosilicate glass pipettes and had tip resistances of 4–6 MΩ when filled with internal recording solution. For whole-cell recordings, internal recording solution contained the following: 95 mM CsF, 25 mM CsCl, 10 mM Cs-HEPES, 10 mM Cs-EGTA, 2 mM NaCl, 2 mM Mg-ATP, 10 mM QX-314, 5 mM TEA-Cl, and 5 mM 4-AP, pH adjusted to 7.3 with CsOH. Data were collected and analyzed using pClamp 10 software (Molecular Devices). Neurons were voltage-clamped at –70 mV to record spontaneous IPSCs (sIPSCs) and miniature IPSCs (mIPSCs), which were isolated by inclusion of D-APV (50 μM), CNQX (10 μM), and TTX (1 μM) for mIPSCs. MiniAnalysis (Synaptosoft) was used to analyze sIPSCs and mIPSCs. For perforated patch recordings, the pipette solution contained the following: 150 mM KCl and 100 mM HEPES, pH adjusted to 7.2 with Tris-OH. The pipette tip was filled with gramicidin-free KCl solution and then backfilled with solution containing gramicidin (100 μg/ml). GABAergic currents were evoked using a picospritzer to deliver a 50-ms puff of 10 μM GABA in the presence of 50 μM D-APV and 10 μM CNQX. Responses were recorded at holding potentials between –100 and 0 mV. The GABA reversal potential was calculated as the x-axis intercept of the best-fit line of the current−voltage plot.

Two-photon laser scanning microscopy

Labeled abDGCs were patched in the whole cell configuration as described above with Alexa Fluor 488 dye (50 μM) added to the internal solution. Dye was allowed to perfuse through the cell for ∼20 min before image acquisition. Fluorescent images were acquired with picosecond pulsed excitation at 790 nM. Images of the soma and dendrites were acquired with 0.19-μm2 pixels with 10-μs pixel dwell time with 1.0-μm focal steps. Neuron Studio was used to create 3-D reconstructions of the dendrites and morphology analysis was performed in NeuronStudio (Wearne et al., 2005) and ImageJ.

Data analysis

Data analysis was performed using Microsoft Excel and OriginPro 2017 software. mIPSC and sIPSCs were analyzed using MiniAnalysis (Synaptosoft). Decay kinetics of mIPSC events was measured as the time to decay from 90% to 37% of the peak amplitude on the falling phase of the response. Comparisons were made with a Mann–Whitney U test, unless otherwise indicated. Differences were considered significant when p < 0.05. Data are shown as mean ± SEM.

Results

Development of spontaneous GABA currents in abDGCs in Fmr1 KO mice

To identify and birthdate newborn DGCs, we injected a modified retrovirus expressing RFP into the SGZ of 8- to 10-week-old Fmr1 KO mice and WT littermates (Tashiro et al., 2015). Retroviral injection clearly labeled neurons located in the SGZ of the dentate gyrus (Fig. 1A). Evoked IPSCs have been detected in abDGCs as early as 7 dpi (Ge et al., 2006) and in our recordings the earliest time point at which we consistently observed spontaneous inhibitory events was 14 dpi. Based on this we performed targeted patch-clamp recordings from RFP-expressing cells at 14, 21, and 28 dpi. We first measured the frequency of sIPSCs and mIPSCs at these time points. sIPSC frequency increased over time as abDGCs matured in both genotypes (Fig. 1B). Despite the known delays in maturation of properties of neurons in other cortical regions in Fmr1 KO mice, we found that there was no difference in the sIPSC frequency at any postdifferentiation age of abDGC tested spanning this early period of development of these neurons (14 dpi WT: 0.028 ± 0.004 Hz; 14 dpi KO: 0.032 ± 0.004 Hz, n = 6/3 (cells, animals respectively), p = 0.49 Mann–Whitney U test; 21 dpi WT: 0.480 ± 0.074 Hz, n = 14/7, 21 dpi KO: 0.468 ± 0.070 Hz, n = 15/9, p = 0.95 Mann–Whitney U test; 28 dpi WT: 1.16 ± 0.031 Hz, n = 12, 4 KO: 0.90 ± 0.086 Hz, 11/6, p = 0.70 Mann–Whitney U test; Fig. 1B,D–F). In addition, we measured action potential independent spontaneous inhibitory events (mIPSCs) in abDGCs, which can be a good indicator of the number of inhibitory connections, or the release probability of those synapses. Again, we did not find a difference in the frequency of these events in comparisons between recordings in Fmr1 WT and Fmr1 KO mice at any of the ages tested (Fig. 1C,G–I). A comparison of the average mIPSC frequencies in each recording over time demonstrated an equivalent increase across this developmental period for abDGCs in both genotypes (Fig. 1C). Together, the lack of a difference in frequency of sIPSCs or mIPSCs during maturation of abDGCs indicates that there is no difference in the number of inhibitory synaptic connections in these neurons in the Fmr1 KO mice during the first 4 weeks after differentiation.

Figure 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 1.

Frequency of sIPSCs and mIPSCs is not altered in abDGCs in Fmr1 KO. A, Representative images of 21 dpi abDGCs virally labeled with RFP. Average sIPSC (B) and mIPSC (C) frequencies across all time points measured in Fmr1 WT (blue) and Fmr1 KO mice (red). D, Schematic of dendritic morphology of abDGCs and representative traces (top panel), cumulative distribution of interevent-interval and average frequency of each recorded neuron (inset) of sIPSCs (bottom panel) at 14 dpi, 21 dpi (E) and 28 dpi (F). G, H, I, Representative traces, cumulative distribution of interevent-intervals and average frequency of mIPSCs in each recorded abDGC (inset) at 14, 21, and 28 dpi, respectively.

The amplitude of spontaneous events, particularly mIPSCs, is an indicator of the strength of individual synapses and is expected to increase as abDGCs undergo maturation (Ge et al., 2006). Comparison of the mIPSC amplitudes at each of the days postdifferentiation spanning this period again did not reveal any difference between recordings in each of the genotypes (WT 14 dpi: 10.4 ± 0.827 pA, n = 7/3; KO 14 dpi: 11.7 ± 1.81 pA, n =,5/5, p = 0.64, Mann–Whitney U test; WT 21 dpi: 15.9 ± 1.45 pA, n = 12/5, KO 21 dpi: 15.6 ± 1.78 pA, n = 12/5, p = 1.00, Mann–Whitney U test; WT 28 dpi: 15.4 ± 1.04 pA, n = 11/4, KO 28 dpi: 16.6 ± 1.13 pA, n = 8/3, p = 0.49, Mann–Whitney U test; Fig. 2A–C). In addition, and consistent with no genotype-mediated changes in synaptic strength during the development of abDGCs in Fmr1 KO mice, there were no differences in the amplitude of sIPSCs (Fig. 2E). In recordings of both mIPSCs and sIPSCs there was an equivalent increase of the average amplitude of events in older neurons in both genotypes (Fig. 2D,E). Interestingly, at 14 dpi, the average amplitude for both mIPSCs and sIPSCs was similar, suggesting that at this time point presynaptic inhibitory neurons make a single synaptic connection per axon. However, the amplitudes diverged at 21 dpi, with sIPSC amplitudes larger than those of mIPSCs, indicating that single presynaptic axons make multiple contacts onto abDGCs in these older neurons (Fig. 2D,E).

Figure 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 2.

mIPSC amplitude is unaffected by loss of FMRP, but mIPSC decay is slower in abDGCs in Fmr1 KO. A, Representative traces of individual and averaged mIPSC events (top) and cumulative distribution and average amplitudes mIPSCs in each recording measured at 14 dpi, 21 dpi (B) and 28 dpi (C) abDGCs. Average mIPSC amplitude (D) and sIPSC amplitude (E) across all time points measured in Fmr1 WT (blue) and KO (red). Average decay of mIPSCs in 14 dpi (F), 21 dpi (G), and 28 dpi (H) abDGCs. I, Average mIPSC decay with zolpidem normalized to decay pre-zolpidem (*p < 0.05, Mann–Whitney U test).

Measuring the decay of mIPSCs, we found that there was a significant increase in the decay time course as abDGCs matured in both genotypes that could be indicative of a change in the subunit composition of GABAARs (Overstreet Wadiche et al., 2005). At the youngest time measured mIPSC decay was not significantly different in Fmr1 KO (WT 14 dpi: 2.39 ± 0.612 ms, n = 7/3; KO 14 dpi: 2.15 ± 0.517 ms, n = 5/4, p = 1.00, Mann–Whitney U test; Fig. 2F). However, at older ages of abDGCs, the decay of mIPSCs in Fmr1 KO neurons was slower (WT 21 dpi: 7.45 ± 0.722 ms, n = 12/5; KO: 11.1 ± 0.987 ms, n = 12/5, p = 0.008, Mann–Whitney U test; WT 28 dpi: 9.36 ± 0.492 ms, n = 11/4; KO 28 dpi: 10.8 ± 0.410 ms, n = 8/3 cells, p = 0.041, Mann–Whitney U test; Fig. 2G,H). In addition, sIPSC decay was not significantly different in Fmr1 KO (WT 14 dpi: 1.66 ± 0.275 ms, n = 6/3; KO 14 dpi: 1.57 ± 0.127 ms, n = 6/3; p = 0.7, Mann–Whitney U test; WT 21 dpi: 12.6 ± 1.19 ms, n = 14/8; KO 21 dpi: 16.0 ± 1.39 ms, n = 15/9, p = 0.08, Mann–Whitney U test; WT 28 dpi: 14.2 ± 1.32 ms, n = 12/4; KO 28 dpi: 13.3 ± 0.80 ms, n = 11/6 cells, p = 0.651, Mann–Whitney U test).

Previous studies have found that the decay kinetics of sIPSCs are slower in abDGCs than in mature granule cells because of the incorporation of α1 subunit into postsynaptic GABAA receptors as neurons mature (Overstreet Wadiche et al., 2005). We tested whether the increased decay of the mIPSCs in abDGCs in Fmr1 KO mice might reflect a lower α1 subunit incorporation by measuring the effect of zolpidem, an α1-specific positive allosteric modulator, on the decay kinetics of mIPSCs. mIPSCs were recorded at 21 dpi and the decay measured before and after the addition of 0.5 µM zolpidem (Fig. 2I). We found that zolpidem lengthened mIPSC decay in 21 dpi abDGCs in both Fmr1 KO and WT to the same degree suggesting that α1 subunit content does not underlie the differences in decay observed in the Fmr1 KO mice (decay zolpidem/decay control WT: 1.40 ± 0.11, n = 17/7; KO: 1.25 ± 0.05, n = 15/7, p = 0.87, Mann–Whitney U test).

Maturation of the GABA reversal potential (EGABA) in abDGCs

As abDGCs mature the chloride reversal potential becomes progressively hyperpolarized in a similar manner to what occurs in other developing neurons (Ge et al., 2006; Chancey et al., 2013). The chloride equilibrium potential in large part determines the reversal potential for GABAA receptors (EGABA) and therefore affects the strength of inhibitory transmission. EGABA reaches its mature value 4 weeks after differentiation in abDGCs (Ge et al., 2006). Prior analysis of Fmr1 KO mice has demonstrated that EGABA maturation is delayed in the developing cortex and hippocampus of Fmr1 KO mice (He et al., 2014; Tyzio et al., 2014). Therefore, we measured EGABA using perforated patch recording from abDGCs 21 d after differentiation. In WT abDGCs the reversal potential was still depolarized from the mature value at this age (WT EGABA: –55.9 ± 0.91 mV, n = 7/3). However, this value did not differ from that recorded in abDGCs in Fmr1 KO animals (KO EGABA: –55.6 ± 2.17 mV, n = 12/6, p = 0.967, Mann–Whitney U test; Fig. 3A–C). Therefore, while the reversal potential for GABA is still relatively immature and depolarized in 21 dpi abDGCs, there is no effect of the loss of FMRP on EGABA, as has been reported in other developing neurons.

Figure 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 3.

EGABA in 21 dpi abDGCs in Fmr1 KO and WT. A, Representative traces of GABA responses from perforated-patch recordings in 21 dpi abDGCs in Fmr1 WT (top, blue) and Fmr1 KO (bottom, red) measured at three holding potentials (–30, –60, –90 mV). The response to a 50-ms puff of 10 μM GABA was measured at several holding potentials in the presence of 50 μM D-APV and 10 μM CNQX. B, Representative current−voltage curves constructed from responses in two cells (blue WT, red KO). EGABA was calculated as the x-axis intercept of the best-fit line of the current−voltage plot. C, Grouped EGABA data from all recordings.

Development of dendrites of abDGCs in Fmr1 KO mice

Dendritic morphology of abDGCs resembles that of mature DGCs as early as 21 dpi, when distal dendrites reach the outer molecular layer and significant arborization is observed (Zhao et al., 2006). Dendritic spines begin to form around 16 dpi, consistent with the fact that glutamatergic signaling is rarely observed before 14 dpi (Ge et al., 2006; Zhao et al., 2006). We thus sought to determine if loss of FMRP would lead to alterations in dendritic morphology in 21 dpi abDGCs during this critical period of their development. abDGCs were filled with a morphologic dye and imaged using a two-photon microscope (Fig. 4A). Quantification of their dendritic complexity at 21 dpi, as assessed by Sholl analysis did not uncover any significant difference between the genotypes (two-way ANOVA for unbalanced design, p = 0.999 for genotype × radius interaction; Fig. 4B). In addition, measurement of total dendritic length did not reveal any difference in 21 dpi abDGCs in Fmr1 KO (WT: 701 ± 94.2 μm, n = 8/5; KO: 729 ± 57.5 μm, n = 15/8, p = 0.781, Mann–Whitney U test), and there was no difference in the number of dendritic branch points in Fmr1 KO at 21 dpi (WT: 8.63 ± 1.92, n = 9/5, KO: 7.20 ± 0.66, n = 14/8, p = 0.917, Mann–Whitney U test; Fig. 4C,D).

Figure 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 4.

Morphology of 21 dpi abDGCs in Fmr1 KO and WT. A, Representative two-photon images of 21 dpi abDGCs filled with Alexa Fluor 488 in slices from Fmr1 WT (right) and Fmr1 KO (left) mice. B, Sholl analysis of dendritic complexity of 21 dpi abDGCs in Fmr1 WT (blue) and Fmr1 KO mice (red). Average total dendritic length (C) and total number of branch points (D) of 21 dpi abDGCs in Fmr1 WT (blue) and Fmr1 KO (red) mice.

Discussion

In this study, we set out to systematically describe the development of abDGCs in the Fragile X mouse model, focusing on the formation of GABAergic inputs to these neurons during the first few weeks after differentiation. There are multiple studies that have found that loss of FMRP can delay neuronal development in the Fragile X brain, therefore, a similar delay in development of abDGCs could have an impact on how these neurons become connected to hippocampal circuits, and how they contribute to circuit function. Studying GABA synapse development is particularly relevant as these are the first synapses to form onto abDGCs and GABA also produces a trophic effect on abDGCs (Ge et al., 2006; Song et al., 2013; Alvarez et al., 2016). Surprisingly, our data indicate that the development of inhibitory signaling onto abDGCs during the first four weeks is mostly unaffected by the loss of FMRP in the Fmr1 KO mice. Given the range of impairments in GABA signaling that have been described in the hippocampus of Fmr1 KO mice including altered expression of GABA receptor subunits and GAD 65/67, and the alterations in EGABA in Fmr1 KO mice (Braat and Kooy, 2015a), it is surprising that GABA signaling appears to be unaltered in developing abDGCs during the critical period of their maturation.

Critical period development and chloride homeostasis in Fragile X

At the cellular level Fragile X is a complex disorder because loss of FMRP leads to the dysregulation of expression of many neuronal proteins (Tang et al., 2015). The mouse model has been particularly useful in describing these complex alterations in neuronal development (Contractor et al., 2015). An important aspect that has emerged from some of these studies is the alterations in synaptic development of neurons during early critical periods in the cortex (Cruz-Martín et al., 2010; Harlow et al., 2010; Nomura et al., 2017). While these numerous delays in both excitatory and inhibitory neuron maturation have been well documented, up until now, it has not been established whether the development of abDGCs shows similar alteration in maturation to the developmentally generated neurons.

Among the known effects in developing neurons is the delay in the maturation of the GABA reversal potential in both the somatosensory cortex (He et al., 2014) and the hippocampus (Tyzio et al., 2014). In abDGCs, EGABA has been demonstrated to be depolarized during early maturation in the weeks following differentiation (Tozuka et al., 2005; Ge et al., 2006) in a manner similar to developing neurons in other regions of the brain in perinatal mice. This occurs because of the relatively elevated intracellular chloride concentrations in abDGCs established by the juvenile chloride transporter NKCC1 (Ge et al., 2006). Disrupting chloride homeostasis in abDGCs leads to a profound alteration in the formation of synapses and in the maturation of the dendritic arbor. As there is growing evidence of altered EGABA in developing neurons, we also measured EGABA in abDGCs 21 d after differentiation when this measure is still maturing and not at its adult value (Ge et al., 2006). We confirmed that in 21 dpi abDGCS in both WT and Fmr1 KO mice, the measured EGABA is still relatively depolarized, but the values of EGABA were not significantly different between the genotypes. Therefore, this crucial measure that regulates the strength of GABA signaling, and has a major impact on neuronal development, is not affected in abDGCs in Fmr1 KO mice.

GABA and Fragile X

The sequence of the formation of inputs and neurotransmitter signaling onto abDGCs broadly reflects that of developing neurons in other brain regions, including the early establishment of tonic and phasic GABAergic signaling followed by the establishment of excitatory connections (Ge et al., 2007a). GABA has an established role in brain development affecting proliferation (LoTurco et al., 1995; Haydar et al., 2000), migration, and maturation of progenitors and neurons (Represa and Ben-Ari, 2005; Wang and Kriegstein, 2009). After postnatal development, there are multiple known defects in GABA signaling associated with Fragile X, including age- and region-specific changes in GABAA receptor subunit expression (Braat and Kooy, 2015b), changes in tonic and phasic GABA currents onto excitatory neurons (Centonze et al., 2008; Curia et al., 2009; Olmos-Serrano et al., 2010; Zhang et al., 2017) as well as defects in interneurons themselves (Gibson et al., 2008; Nomura et al., 2017). Expression of several GABAA receptor subunits is reduced in the cortex, hippocampus, or forebrain of Fmr1 KO mice (El Idrissi et al., 2005; D'Hulst et al., 2006; Gantois et al., 2006; Adusei et al., 2010). Expression of the glutamate decarboxylase enzyme (GAD) responsible for converting glutamate to GABA is increased in the hippocampus, but decreased in the amygdala of Fmr1 KO mice (El Idrissi et al., 2005; Olmos-Serrano et al., 2010). Functionally, both the frequency and amplitude of mIPSCs and sIPSCs is reduced in the amygdala of juvenile Fmr1 KO mice (Olmos-Serrano et al., 2010); however, a similar alteration is not observed in the subiculum of older mice (Curia et al., 2009). In some cases, these alterations in GABAergic signaling occur only early in postnatal development (Adusei et al., 2010; Nomura et al., 2017). Thus, loss of FMRP clearly affects GABA signaling in the postnatal mouse brain, but whether this is also the case for abDGCs was not known. By mapping spontaneous events by recording both sIPSCs and mIPSCs at three time points after differentiation of abDGCs, we were able to establish how inhibitory synapses form onto these neurons. We found that in both genotypes the frequency of IPSCs increases over time as would be expected if new synapses were being formed over the post differentiation period we analyzed. Interestingly, the amplitude of both the sIPSC and mIPSC events increased between 14 and 21 dpi. The increase in the mean quantal size at this developmental time point suggests that individual synapses become stronger. At the earliest time points, the amplitude of both mIPSCs and sIPSCs is equivalent suggesting that presynaptic axons of the inhibitory cells make a single contact whereas by 21 dpi the mean amplitude of sIPSCs is double the mean amplitude of quantal events, suggesting that the sIPSCs represent the release at multiple synapses. While again we did not observe any genotype differences in these parameters, we did observe consistent lengthening of the decay kinetics of mIPSCs in Fmr1 KO mice which were significant in recordings from both 21 dpi and 28 dpi neurons. Prior work that compared the decay kinetics of inhibitory events in abDGCs and mature granule neurons found that the sIPSCs in abDGCs are slower because of the subunit composition of postsynaptic GABAA receptors (Overstreet Wadiche et al., 2005). This study found that the zolpidem sensitivity of inhibitory events was increased in mature neurons suggesting that the α1 subunit is increasingly incorporated into neurons as they mature (Overstreet Wadiche et al., 2005). Given this, we considered the possibility that a reduction in the incorporation of the α1 subunit in abDGCs in Fmr1 KO mice could underlie the prolonged decay of mIPSCs in Fmr1 KO. However, we did not observe a significant difference in the effect of zolpidem on mIPSC decay in 21 dpi abDGCs between Fmr1 KO and WT, indicating that a reduction in α1 expression is unlikely to underlie the changes in mIPSC decay. It is possible that alterations in expression of other GABA receptor subunits or even differences in the location of GABAergic synapses on developing neurons may underlie the observed change in mIPSC decays recorded by a somatic electrode. Interestingly, while prior work has reported that abDGCs lack expression of the GABAA α1 subunit (Overstreet Wadiche et al., 2005), we found zolpidem had a significant effect on most mIPSCs in the 21 dpi birth-dated neurons in both genotypes. This may reflect differences in the populations of neurons that were analyzed in the previous study which recorded from POMC-GFP labeled neurons which are a more heterogeneous aged group (Overstreet et al., 2004).

In addition to the measures of synaptic function, we also measured the dendritic arbor to quantify both the complexity as well as the total dendritic length in 21 dpi abDGCs. A prior study has reported that selective deletion of FMRP in adult neural stem cells isolated from the dentate gyrus, as well as in situ, caused a reduction in both the dendritic complexity and total dendritic length when measured in neurons 56 d after differentiation (Guo et al., 2011). Dendritic length and complexity are also reduced in mice with knockout of the FMRP paralog FXR2P and double knockout of FXR2P and FMRP induced an additive effect on dendrites of abDGCs (Guo et al., 2015). In our experiments, we patched and filled abDGCs with morphologic dyes in live slices and imaged the dendritic arbor at 21 dpi. In these younger neurons, we did not observe any genotype related differences in the total dendritic length or complexity in Fmr1 KO mice. It is possible that this difference in our results and those of the earlier study is due to a dendritic phenotype that emerges later on in the development of FMRP lacking abDGCs. While there have also been reports of region-specific or developmental age-specific alterations in dendritic spine density or immature spine morphology in Fmr1 KO mice, there is no consensus on the effect of loss of FMRP on dendritic spines (for review, see He and Portera-Cailliau, 2013). While we did not examine this measure in these immature abDGCs, prior work has reported normal dendritic spines in the mature DG in Fmr1 KO (Grossman et al., 2010).

In summary, we performed a systematic analysis of the functional properties of GABAergic synapses in abDGCs during the first four weeks after differentiation in Fmr1 KO mice. While previous studies have demonstrated that cell proliferation and fate specification of adult neural stem cells is altered by FMRP loss, our data suggest that neurons that develop from these stem cells do not have major alterations in the maturation of their GABAergic synaptic inputs and dendrites during the first four weeks of their development.

Acknowledgments

We thank Stephen Kraniotis for technical help and members of the Contractor and Swanson laboratories at Northwestern for helpful discussion.

Footnotes

  • The authors declare no competing financial interests.

  • This work was supported by the National Institutes of Health (NIH)/National Institute of Mental Health (NIMH) Grant R21 MH104808 and the Department of Defense Grant W81XWH-13-ARP-IDA (to A.C.). C.L.R. was funded by the NIH/NIMH Training Program in Neurobiology of Information Storage T32 MH067564 training grant and the NIH/National Center for Advancing Translational Sciences Training Program in Clinical and Translational Science TL1 8UL1TR000150 training grant.

This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International license, which permits unrestricted use, distribution and reproduction in any medium provided that the original work is properly attributed.

References

  1. ↵
    Adusei DC, Pacey LK, Chen D, Hampson DR (2010) Early developmental alterations in GABAergic protein expression in fragile X knockout mice. Neuropharmacology 59:167–171. doi:10.1016/j.neuropharm.2010.05.002 pmid:20470805
    OpenUrlCrossRefPubMed
  2. ↵
    Alvarez DD, Giacomini D, Yang SM, Trinchero MF, Temprana SG, Büttner KA, Beltramone N, Schinder AF (2016) A disynaptic feedback network activated by experience promotes the integration of new granule cells. Science 354:459–465. doi:10.1126/science.aaf2156 pmid:27789840
    OpenUrlAbstract/FREE Full Text
  3. ↵
    Arnett MT, Herman DH, McGee AW (2014) Deficits in tactile learning in a mouse model of fragile X syndrome. PLoS One 9:e109116. doi:10.1371/journal.pone.0109116 pmid:25296296
    OpenUrlCrossRefPubMed
  4. ↵
    Bergami M, Masserdotti G, Temprana SG, Motori E, Eriksson TM, Göbel J, Yang SM, Conzelmann KK, Schinder AF, Götz M, Berninger B (2015) A critical period for experience-dependent remodeling of adult-born neuron connectivity. Neuron 85:710–717. doi:10.1016/j.neuron.2015.01.001 pmid:25661179
    OpenUrlCrossRefPubMed
  5. ↵
    Braat S, Kooy RF (2015a) The GABAA receptor as a therapeutic target for neurodevelopmental disorders. Neuron 86:1119–1130. doi:10.1016/j.neuron.2015.03.042
    OpenUrlCrossRefPubMed
  6. ↵
    Braat S, Kooy RF (2015b) Insights into GABAAergic system deficits in fragile X syndrome lead to clinical trials. Neuropharmacology 88:48–54. doi:10.1016/j.neuropharm.2014.06.028
    OpenUrlCrossRefPubMed
  7. ↵
    Bureau I, Shepherd GM, Svoboda K (2008) Circuit and plasticity defects in the developing somatosensory cortex of FMR1 knock-out mice. J Neurosci 28:5178–5188. doi:10.1523/JNEUROSCI.1076-08.2008 pmid:18480274
    OpenUrlAbstract/FREE Full Text
  8. ↵
    Centonze D, Rossi S, Mercaldo V, Napoli I, Ciotti MT, De Chiara V, Musella A, Prosperetti C, Calabresi P, Bernardi G, Bagni C (2008) Abnormal striatal GABA transmission in the mouse model for the fragile X syndrome. Biol Psychiatry 63:963–973. doi:10.1016/j.biopsych.2007.09.008 pmid:18028882
    OpenUrlCrossRefPubMed
  9. ↵
    Chancey JH, Adlaf EW, Sapp MC, Pugh PC, Wadiche JI, Overstreet-Wadiche LS (2013) GABA depolarization is required for experience-dependent synapse unsilencing in adult-born neurons. J Neurosci 33:6614–6622. doi:10.1523/JNEUROSCI.0781-13.2013 pmid:23575858
    OpenUrlAbstract/FREE Full Text
  10. ↵
    Clelland CD, Choi M, Romberg C, Clemenson GD Jr., Fragniere A, Tyers P, Jessberger S, Saksida LM, Barker RA, Gage FH, Bussey TJ (2009) A functional role for adult hippocampal neurogenesis in spatial pattern separation. Science 325:210–213. doi:10.1126/science.1173215
    OpenUrlAbstract/FREE Full Text
  11. ↵
    Contractor A, Klyachko VA, Portera-Cailliau C (2015) Altered neuronal and circuit excitability in fragile X syndrome. Neuron 87:699–715. doi:10.1016/j.neuron.2015.06.017 pmid:26291156
    OpenUrlCrossRefPubMed
  12. ↵
    Cruz-Martín A, Crespo M, Portera-Cailliau C (2010) Delayed stabilization of dendritic spines in fragile X mice. J Neurosci 30:7793–7803. doi:10.1523/JNEUROSCI.0577-10.2010 pmid:20534828
    OpenUrlAbstract/FREE Full Text
  13. ↵
    Curia G, Papouin T, Séguéla P, Avoli M (2009) Downregulation of tonic GABAergic inhibition in a mouse model of fragile X syndrome. Cereb Cortex 19:1515–1520. doi:10.1093/cercor/bhn159
    OpenUrlCrossRefPubMed
  14. ↵
    D'Hulst C, De Geest N, Reeve SP, Van Dam D, De Deyn PP, Hassan BA, Kooy RF (2006) Decreased expression of the GABAA receptor in fragile X syndrome. Brain Res 1121:238–245. doi:10.1016/j.brainres.2006.08.115 pmid:17046729
    OpenUrlCrossRefPubMed
  15. ↵
    Eadie BD, Zhang WN, Boehme F, Gil-Mohapel J, Kainer L, Simpson JM, Christie BR (2009) Fmr1 knockout mice show reduced anxiety and alterations in neurogenesis that are specific to the ventral dentate gyrus. Neurobiol Dis 36:361–373. doi:10.1016/j.nbd.2009.08.001 pmid:19666116
    OpenUrlCrossRefPubMed
  16. ↵
    El Idrissi A, Ding XH, Scalia J, Trenkner E, Brown WT, Dobkin C (2005) Decreased GABA(A) receptor expression in the seizure-prone fragile X mouse. Neurosci Lett 377:141–146. doi:10.1016/j.neulet.2004.11.087 pmid:15755515
    OpenUrlCrossRefPubMed
  17. ↵
    Gantois I, Vandesompele J, Speleman F, Reyniers E, D'Hooge R, Severijnen LA, Willemsen R, Tassone F, Kooy RF (2006) Expression profiling suggests underexpression of the GABA(A) receptor subunit delta in the fragile X knockout mouse model. Neurobiol Dis 21:346–357. doi:10.1016/j.nbd.2005.07.017
    OpenUrlCrossRefPubMed
  18. ↵
    Ge S, Goh EL, Sailor KA, Kitabatake Y, Ming GL, Song H (2006) GABA regulates synaptic integration of newly generated neurons in the adult brain. Nature 439:589–593. doi:10.1038/nature04404 pmid:16341203
    OpenUrlCrossRefPubMed
  19. ↵
    Ge S, Pradhan DA, Ming GL, Song H (2007a) GABA sets the tempo for activity-dependent adult neurogenesis. Trends Neurosci 30:1–8. doi:10.1016/j.tins.2006.11.001
    OpenUrlCrossRefPubMed
  20. ↵
    Ge S, Yang CH, Hsu KS, Ming GL, Song H (2007b) A critical period for enhanced synaptic plasticity in newly generated neurons of the adult brain. Neuron 54:559–566. doi:10.1016/j.neuron.2007.05.002
    OpenUrlCrossRefPubMed
  21. ↵
    Gibson JR, Bartley AF, Hays SA, Huber KM (2008) Imbalance of neocortical excitation and inhibition and altered UP states reflect network hyperexcitability in the mouse model of fragile X syndrome. J Neurophysiol 100:2615–2626. doi:10.1152/jn.90752.2008 pmid:18784272
    OpenUrlCrossRefPubMed
  22. ↵
    Gonçalves JT, Schafer ST, Gage FH (2016) Adult neurogenesis in the hippocampus: from stem cells to behavior. Cell 167:897–914. doi:10.1016/j.cell.2016.10.021 pmid:27814520
    OpenUrlCrossRefPubMed
  23. ↵
    Grossman AW, Aldridge GM, Lee KJ, Zeman MK, Jun CS, Azam HS, Arii T, Imoto K, Greenough WT, Rhyu IJ (2010) Developmental characteristics of dendritic spines in the dentate gyrus of Fmr1 knockout mice. Brain Res 1355:221–227. doi:10.1016/j.brainres.2010.07.090 pmid:20682298
    OpenUrlCrossRefPubMed
  24. ↵
    Guo W, Allan AM, Zong R, Zhang L, Johnson EB, Schaller EG, Murthy AC, Goggin SL, Eisch AJ, Oostra BA, Nelson DL, Jin P, Zhao X (2011) Ablation of Fmrp in adult neural stem cells disrupts hippocampus-dependent learning. Nat Med 17:559–565. doi:10.1038/nm.2336 pmid:21516088
    OpenUrlCrossRefPubMed
  25. ↵
    Guo W, Polich ED, Su J, Gao Y, Christopher DM, Allan AM, Wang M, Wang F, Wang G, Zhao X (2015) Fragile X proteins FMRP and FXR2P control synaptic GluA1 expression and neuronal maturation via distinct mechanisms. Cell Rep 11:1651–1666. doi:10.1016/j.celrep.2015.05.013 pmid:26051932
    OpenUrlCrossRefPubMed
  26. ↵
    Harlow EG, Till SM, Russell TA, Wijetunge LS, Kind P, Contractor A (2010) Critical period plasticity is disrupted in the barrel cortex of FMR1 knockout mice. Neuron 65:385–398. doi:10.1016/j.neuron.2010.01.024 pmid:20159451
    OpenUrlCrossRefPubMed
  27. ↵
    Haydar TF, Wang F, Schwartz ML, Rakic P (2000) Differential modulation of proliferation in the neocortical ventricular and subventricular zones. J Neurosci 20:5764–5774. pmid:10908617
    OpenUrlAbstract/FREE Full Text
  28. ↵
    He CX, Portera-Cailliau C (2013) The trouble with spines in fragile X syndrome: density, maturity and plasticity. Neuroscience 251:120–128. doi:10.1016/j.neuroscience.2012.03.049 pmid:22522472
    OpenUrlCrossRefPubMed
  29. ↵
    He CX, Cantu DA, Mantri SS, Zeiger WA, Goel A, Portera-Cailliau C (2017) Tactile defensiveness and impaired adaptation of neuronal activity in the Fmr1 knock-out mouse model of autism. J Neurosci 37:6475–6487. doi:10.1523/JNEUROSCI.0651-17.2017 pmid:28607173
    OpenUrlAbstract/FREE Full Text
  30. ↵
    He Q, Nomura T, Xu J, Contractor A (2014) The developmental switch in GABA polarity is delayed in fragile X mice. J Neurosci 34:446–450. doi:10.1523/JNEUROSCI.4447-13.2014 pmid:24403144
    OpenUrlAbstract/FREE Full Text
  31. ↵
    Lazarov O, Demars MP, Zhao Kda T, Ali HM, Grauzas V, Kney A, Larson J (2012) Impaired survival of neural progenitor cells in dentate gyrus of adult mice lacking fMRP. Hippocampus 22:1220–1224. doi:10.1002/hipo.20989 pmid:22128095
    OpenUrlCrossRefPubMed
  32. ↵
    Li Y, Zhao X (2014) Concise review: fragile X proteins in stem cell maintenance and differentiation. Stem Cells 32:1724–1733. doi:10.1002/stem.1698 pmid:24648324
    OpenUrlCrossRefPubMed
  33. ↵
    LoTurco JJ, Owens DF, Heath MJ, Davis MB, Kriegstein AR (1995) GABA and glutamate depolarize cortical progenitor cells and inhibit DNA synthesis. Neuron 15:1287–1298. pmid:8845153
    OpenUrlCrossRefPubMed
  34. ↵
    Luo Y, Shan G, Guo W, Smrt RD, Johnson EB, Li X, Pfeiffer RL, Szulwach KE, Duan R, Barkho BZ, Li W, Liu C, Jin P, Zhao X (2010) Fragile x mental retardation protein regulates proliferation and differentiation of adult neural stem/progenitor cells. PLoS Genet 6:e1000898. doi:10.1371/journal.pgen.1000898 pmid:20386739
    OpenUrlCrossRefPubMed
  35. ↵
    Martin BS, Corbin JG, Huntsman MM (2014) Deficient tonic GABAergic conductance and synaptic balance in the fragile X syndrome amygdala. J Neurophysiol 112:890–902. doi:10.1152/jn.00597.2013 pmid:24848467
    OpenUrlCrossRefPubMed
  36. ↵
    Nakashiba T, Cushman JD, Pelkey KA, Renaudineau S, Buhl DL, McHugh TJ, Rodriguez Barrera V, Chittajallu R, Iwamoto KS, McBain CJ, Fanselow MS, Tonegawa S (2012) Young dentate granule cells mediate pattern separation, whereas old granule cells facilitate pattern completion. Cell 149:188–201. doi:10.1016/j.cell.2012.01.046 pmid:22365813
    OpenUrlCrossRefPubMed
  37. ↵
    Nimchinsky EA, Oberlander AM, Svoboda K (2001) Abnormal development of dendritic spines in FMR1 knock-out mice. J Neurosci 21:5139–5146. pmid:11438589
    OpenUrlAbstract/FREE Full Text
  38. ↵
    Nomura T, Musial TF, Marshall JJ, Zhu Y, Remmers CL, Xu J, Nicholson DA, Contractor A (2017) Delayed maturation of fast-spiking interneurons is rectified by activation of the TrkB receptor in the mouse model of fragile X syndrome. J Neurosci 37:11298–11310. doi:10.1523/JNEUROSCI.2893-16.2017 pmid:29038238
    OpenUrlAbstract/FREE Full Text
  39. ↵
    Olmos-Serrano JL, Paluszkiewicz SM, Martin BS, Kaufmann WE, Corbin JG, Huntsman MM (2010) Defective GABAergic neurotransmission and pharmacological rescue of neuronal hyperexcitability in the amygdala in a mouse model of fragile X syndrome. J Neurosci 30:9929–9938. doi:10.1523/JNEUROSCI.1714-10.2010
    OpenUrlAbstract/FREE Full Text
  40. ↵
    Overstreet LS, Hentges ST, Bumaschny VF, de Souza FS, Smart JL, Santangelo AM, Low MJ, Westbrook GL, Rubinstein M (2004) A transgenic marker for newly born granule cells in dentate gyrus. J Neurosci 24:3251–3259. doi:10.1523/JNEUROSCI.5173-03.2004 pmid:15056704
    OpenUrlAbstract/FREE Full Text
  41. ↵
    Overstreet Wadiche L, Bromberg DA, Bensen AL, Westbrook GL (2005) GABAergic signaling to newborn neurons in dentate gyrus. J Neurophysiol 94:4528–4532. doi:10.1152/jn.00633.2005 pmid:16033936
    OpenUrlCrossRefPubMed
  42. ↵
    Paluszkiewicz SM, Olmos-Serrano JL, Corbin JG, Huntsman MM (2011) Impaired inhibitory control of cortical synchronization in fragile X syndrome. J Neurophysiol 106:2264–2272. doi:10.1152/jn.00421.2011 pmid:21795626
    OpenUrlCrossRefPubMed
  43. ↵
    Represa A, Ben-Ari Y (2005) Trophic actions of GABA on neuronal development. Trends Neurosci 28:278–283. doi:10.1016/j.tins.2005.03.010 pmid:15927682
    OpenUrlCrossRefPubMed
  44. ↵
    Shors TJ, Miesegaes G, Beylin A, Zhao M, Rydel T, Gould E (2001) Neurogenesis in the adult is involved in the formation of trace memories. Nature 410:372–376. doi:10.1038/35066584 pmid:11268214
    OpenUrlCrossRefPubMed
  45. ↵
    Song J, Sun J, Moss J, Wen Z, Sun GJ, Hsu D, Zhong C, Davoudi H, Christian KM, Toni N, Ming GL, Song H (2013) Parvalbumin interneurons mediate neuronal circuitry-neurogenesis coupling in the adult hippocampus. Nat Neurosci 16:1728–1730. doi:10.1038/nn.3572 pmid:24212671
    OpenUrlCrossRefPubMed
  46. ↵
    Tang B, Wang T, Wan H, Han L, Qin X, Zhang Y, Wang J, Yu C, Berton F, Francesconi W, Yates JR 3rd., Vanderklish PW, Liao L (2015) Fmr1 deficiency promotes age-dependent alterations in the cortical synaptic proteome. Proc Natl Acad Sci USA 112:E4697–E4706. doi:10.1073/pnas.1502258112
    OpenUrlAbstract/FREE Full Text
  47. ↵
    Tashiro A, Zhao C, Gage FH (2006) Retrovirus-mediated single-cell gene knockout technique in adult newborn neurons in vivo. Nat Protoc 1:3049–3055. doi:10.1038/nprot.2006.473
    OpenUrlCrossRefPubMed
  48. ↵
    Tashiro A, Zhao C, Suh H, Gage FH (2015) Preparation and use of retroviral vectors for labeling, imaging, and genetically manipulating cells. Cold Spring Harb Protoc 2015:883–888. doi:10.1101/pdb.top086363 pmid:26430260
    OpenUrlCrossRefPubMed
  49. ↵
    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. doi:10.1016/j.neuron.2005.08.023 pmid:16157276
    OpenUrlCrossRefPubMed
  50. ↵
    Tyzio R, Nardou R, Ferrari DC, Tsintsadze T, Shahrokhi A, Eftekhari S, Khalilov I, Tsintsadze V, Brouchoud C, Chazal G, Lemonnier E, Lozovaya N, Burnashev N, Ben-Ari Y (2014) Oxytocin-mediated GABA inhibition during delivery attenuates autism pathogenesis in rodent offspring. Science 343:675–679. doi:10.1126/science.1247190 pmid:24503856
    OpenUrlAbstract/FREE Full Text
  51. ↵
    Wang DD, Kriegstein AR (2009) Defining the role of GABA in cortical development. J Physiol 587:1873–1879. doi:10.1113/jphysiol.2008.167635 pmid:19153158
    OpenUrlCrossRefPubMed
  52. ↵
    Wearne SL, Rodriguez A, Ehlenberger DB, Rocher AB, Henderson SC, Hof PR (2005) New techniques for imaging, digitization and analysis of three-dimensional neural morphology on multiple scales. Neuroscience 136:661–680. doi:10.1016/j.neuroscience.2005.05.053 pmid:16344143
    OpenUrlCrossRefPubMed
  53. ↵
    Willemsen R, Levenga J, Oostra BA (2011) CGG repeat in the FMR1 gene: size matters. Clin Genet 80:214–225. doi:10.1111/j.1399-0004.2011.01723.x pmid:21651511
    OpenUrlCrossRefPubMed
  54. ↵
    Zhang N, Peng Z, Tong X, Lindemeyer AK, Cetina Y, Huang CS, Olsen RW, Otis TS, Houser CR (2017) Decreased surface expression of the δ subunit of the GABAA receptor contributes to reduced tonic inhibition in dentate granule cells in a mouse model of fragile X syndrome. Exp Neurol 297:168–178. doi:10.1016/j.expneurol.2017.08.008
    OpenUrlCrossRef
  55. ↵
    Zhao C, Teng EM, Summers RG Jr., Ming GL, Gage FH (2006) Distinct morphological stages of dentate granule neuron maturation in the adult mouse hippocampus. J Neurosci 26:3–11. doi:10.1523/JNEUROSCI.3648-05.2006
    OpenUrlAbstract/FREE Full Text
  56. ↵
    Zhao MG, Toyoda H, Ko SW, Ding HK, Wu LJ, Zhuo M (2005) Deficits in trace fear memory and long-term potentiation in a mouse model for fragile X syndrome. J Neurosci 25:7385–7392. doi:10.1523/JNEUROSCI.1520-05.2005 pmid:16093389
    OpenUrlAbstract/FREE Full Text

Synthesis

Reviewing Editor: Christian Hansel, University of Chicago

Decisions are customarily a result of the Reviewing Editor and the peer reviewers coming together and discussing their recommendations until a consensus is reached. When revisions are invited, a fact-based synthesis statement explaining their decision and outlining what is needed to prepare a revision will be listed below. The following reviewer(s) agreed to reveal their identity: Molly Huntsman, Emanuela Santini.

The paper has the potential to advance the field, but more work needs to be done to increase its impact. The study shows that abDGC neurons in Fmr1 KO mice have an immature phenotype with respect to the lengthening of mIPSC decay during their early development (at 21dpi and 28dpi). This is possibly a consequence of their postsynaptic GABAa subunit composition. However, lack of FMRP does not change other inhibitory synaptic properties or the morphological phenotype of the abDGCs. This study suggests that the absence of FMRP does not affect the development of neurons derived from neuronal stem cells.

Statistics:

The Kolmorov-Smirnovtest may be the more appropriate statistical test here, as it is more sensitive to differences in distributions. Inclusion of the KS statistics and probability inter-event intervals and frequency may be warranted.

Comments

Overall the paper is well-written, the experiments well-designed and executed and the statistic is appropriate.

The impression is that the authors do not fully explain and exploit the difference in lengthening decay of abDGCs when Fmr1 is genetically ablated. So there is space to improvements-these are some suggestions:

In the discussion section of the paper, the authors speculate that this phenotype may be due to a reduction of the alpha1 subunit incorporation in abDGCs of Fmr1 KO mice.

The paper would be improved by performing an experiment with zolpiderm to show that this is indeed the case in Fmr1 KO mice. Editorial note: this experiment was strongly suggested by BOTH reviewers, and thus should have preference in the revision plans.

The authors should also expand the discussion of this point-for instance, it will be interesting to speculate on the reason(s) of the reduction of alpha1 subunit incorporation in relation to the known function of FMRP as inhibitor of protein synthesis (maybe the reduced incorporation of alpha1 subunit is caused by enhanced translation of the other subunits?)

The authors reference a paper (Overstreet Wadiche et al., 2005) where it is showed that sIPSCs kinetic decay in newborn neurons decay more slowly than in mature granule cells. The mIPSCs seem to have opposite trend in WT (fig2 F-G-H), do the authors have data on sIPSCs? if so, this should be included in the paper or at least, the authors should comment on this.

The graphs in Fig2 G and H should have the same scale.

Line 227. The authors recorded 7 neuron from 30 mice (n=7/30) !?!. Please, revise.

Fig 3. In panel B seems that WT and KO have different sensitivity to GABA although the Egaba is very similar. The authors should comment on this.

In panel A the representative traces seem to support this difference in sensitivity (very obvious at -90). The authors should find a more representative traces or comment on this.

The authors analyzed different morphological parameters of these neurons. It is known that a main morphological difference in neurons of FXS model mice (and patients) is the

difference in spine density and increase density of immature spines. Do the authors have data on the spine density of abDGCs? If not, they should at least comment on this.

The study aims to investigate the functional development of adult born Dentate Granule Cells (abDGCs) and the effect of loss of the fragile-x mental retardation protein (FMRP) on synaptic maturation and integration. To that end, the investigators identified abDGCs in Fmr1 by retroviral labeling in the subgranular zone and performed patch-clamp recordings to assess synaptic function of abDGCs across three time points post-injection. The investigators assess the frequency and peak amplitudes of spontaneous and mini IPSCs (sIPSCs, mIPSCs) and found no difference in inhibitory synaptic connections or strength between the groups. The decay kinetics, however, were found to be significantly increased in the KO animal. The investigators went on to evaluate the GABA reversal potentials and dendritic morphology in abDGCS of the wild-type and Fmr1 animal and, again, observed no differences.

The authors' findings imply that abDGCs in the Fmr1 mouse track with normal development in the wild-type animal. This study has the potential to extend the observations of GABAergic signaling in Fragile X Syndrome to other areas of the post-natal mouse brain. However, clarification and additional data are needed to reinforce the impact of this study - including more careful statistical treatment of data, and further studies. Major and minor comments are below.

MAJOR POINTS

It is reported that mean quantal size indicates increased strength. This study would be significantly strengthened by including experiments aimed at characterizing short and long-term potentiation in the dentate gyrus. Do abDGCs in the Fmr1 animal exhibit altered plasticity? Or is there a change in paired pulse inhibition?

The author's study deviates importantly from other studies in the literature regarding dendritic morphology changes and tonic and phasic inhibition in the Fmr1 KO animal. A more careful consideration of the literature and discussion about these interesting results is warranted.

Additionally, where there any significant changes in soma shape of abDGCs?

The authors state the earliest time point for observations of evoked inhibitory events reflected in the literature was 7 dpi. However, in their system, the earliest time point for spontaneous events is stated to be 14 dpi. To demonstrate this, a 7dpi time point for spontaneous events should be included in Figure 1. This may be a typo, but an evaluation of stimulus-evoked IPSCs may be warranted across these time points.

All Figures - scale and calibration bars should be clearly labeled on all images

MINOR POINTS

Manuscript should be reviewed for grammatical and syntactical errors.

Abbreviations should be defined at first appearance in manuscript (i.e the abbreviation “abDGCs” first appears in the abstract but is not defined until the significance statement.

For clarity, references to figures should be mentioned in the text in the order they presented in the figures.

Back to top

In this issue

eneuro: 5 (6)
eNeuro
Vol. 5, Issue 6
November/December 2018
  • Table of Contents
  • Index by author
Email

Thank you for sharing this eNeuro article.

NOTE: We request your email address only to inform the recipient that it was you who recommended this article, and that it is not junk mail. We do not retain these email addresses.

Enter multiple addresses on separate lines or separate them with commas.
Development of GABAergic Inputs Is Not Altered in Early Maturation of Adult Born Dentate Granule Neurons in Fragile X Mice
(Your Name) has forwarded a page to you from eNeuro
(Your Name) thought you would be interested in this article in eNeuro.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Print
View Full Page PDF
Citation Tools
Development of GABAergic Inputs Is Not Altered in Early Maturation of Adult Born Dentate Granule Neurons in Fragile X Mice
Christine L. Remmers, Anis Contractor
eNeuro 19 November 2018, 5 (6) ENEURO.0137-18.2018; DOI: 10.1523/ENEURO.0137-18.2018

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Respond to this article
Share
Development of GABAergic Inputs Is Not Altered in Early Maturation of Adult Born Dentate Granule Neurons in Fragile X Mice
Christine L. Remmers, Anis Contractor
eNeuro 19 November 2018, 5 (6) ENEURO.0137-18.2018; DOI: 10.1523/ENEURO.0137-18.2018
del.icio.us logo Digg logo Reddit logo Twitter logo Facebook logo Google logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Google Plus One

Jump to section

  • Article
    • Abstract
    • Significance Statement
    • Introduction
    • Materials and Methods
    • Results
    • Discussion
    • Acknowledgments
    • Footnotes
    • References
    • Synthesis
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF

Keywords

  • adult born neurons
  • fragile X
  • GABA
  • neurogenesis
  • synapse

Responses to this article

Respond to this article

Jump to comment:

No eLetters have been published for this article.

Related Articles

Cited By...

More in this TOC Section

Negative Results

  • Closed-Loop Acoustic Stimulation Enhances Sleep Oscillations But Not Memory Performance
  • Cyfip1 Haploinsufficiency Does Not Alter GABAA Receptor δ-Subunit Expression and Tonic Inhibition in Dentate Gyrus PV+ Interneurons and Granule Cells
  • Glucagon-Like Peptide-1 Receptor Agonist Treatment Does Not Reduce Abuse-Related Effects of Opioid Drugs
Show more Negative Results

Development

  • Spatiotemporal regulation of de novo and salvage purine synthesis during brain development
  • A Novel Interaction between MFN2/Marf and MARK4/PAR-1 Is Implicated in Synaptic Defects and Mitochondrial Dysfunction
  • Vascularization in mTOR Mouse Mutants: An Effort Not in Vein
Show more Development

Subjects

  • Development

  • Home
  • Alerts
  • Visit Society for Neuroscience on Facebook
  • Follow Society for Neuroscience on Twitter
  • Follow Society for Neuroscience on LinkedIn
  • Visit Society for Neuroscience on Youtube
  • Follow our RSS feeds

Content

  • Early Release
  • Current Issue
  • Latest Articles
  • Issue Archive
  • Blog
  • Browse by Topic

Information

  • For Authors
  • For the Media

About

  • About the Journal
  • Editorial Board
  • Privacy Policy
  • Contact
  • Feedback
(eNeuro logo)
(SfN logo)

Copyright © 2023 by the Society for Neuroscience.
eNeuro eISSN: 2373-2822

The ideas and opinions expressed in eNeuro do not necessarily reflect those of SfN or the eNeuro Editorial Board. Publication of an advertisement or other product mention in eNeuro should not be construed as an endorsement of the manufacturer’s claims. SfN does not assume any responsibility for any injury and/or damage to persons or property arising from or related to any use of any material contained in eNeuro.