Next Article in Journal
Lamin A/C Is Dispensable to Mechanical Repression of Adipogenesis
Next Article in Special Issue
RGS14 Regulation of Post-Synaptic Signaling and Spine Plasticity in Brain
Previous Article in Journal
Effects of Incretin-Related Diabetes Drugs on Bone Formation and Bone Resorption
Previous Article in Special Issue
Role of Satb1 and Satb2 Transcription Factors in the Glutamate Receptors Expression and Ca2+ Signaling in the Cortical Neurons In Vitro
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

FMRP Interacts with RARα in Synaptic Retinoic Acid Signaling and Homeostatic Synaptic Plasticity

Departments of Neurosurgery, Neuropsychiatry and Behavioral Sciences, Stanford University School of Medicine, 265 Campus Drive, Stanford, CA 94305-5453, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally.
Int. J. Mol. Sci. 2021, 22(12), 6579; https://doi.org/10.3390/ijms22126579
Submission received: 31 May 2021 / Revised: 16 June 2021 / Accepted: 17 June 2021 / Published: 19 June 2021
(This article belongs to the Special Issue Synaptic Transmission and Protein Interaction)

Abstract

:
The fragile X syndrome (FXS) is an X-chromosome-linked neurodevelopmental disorder with severe intellectual disability caused by inactivation of the fragile X mental retardation 1 (FMR1) gene and subsequent loss of the fragile X mental retardation protein (FMRP). Among the various types of abnormal synaptic function and synaptic plasticity phenotypes reported in FXS animal models, defective synaptic retinoic acid (RA) signaling and subsequent defective homeostatic plasticity have emerged as a major synaptic dysfunction. However, the mechanism underlying the defective synaptic RA signaling in the absence of FMRP is unknown. Here, we show that RARα, the RA receptor critically involved in synaptic RA signaling, directly interacts with FMRP. This interaction is enhanced in the presence of RA. Blocking the interaction between FMRP and RARα with a small peptide corresponding to the critical binding site in RARα abolishes RA-induced increases in excitatory synaptic transmission, recapitulating the phenotype seen in the Fmr1 knockout mouse. Taken together, these data suggest that not only are functional FMRP and RARα necessary for RA-dependent homeostatic synaptic plasticity, but that the interaction between these two proteins is essential for proper transcription-independent RA signaling. Our results may provide further mechanistic understanding into FXS synaptic pathophysiology.

1. Introduction

The fragile X syndrome (FXS) is the most common genetic form of intellectual disability with impaired cognitive and social skills, anxiety and autistic-like behaviors [1,2]. FXS is caused by the loss of function mutations in the fragile X mental retardation 1 (FMR1) gene, which encodes the fragile X mental retardation protein (FMRP) [3]. The most common cause of FXS is the expansion of CGG repeats in the 5′ untranslated region of the FMR1 gene resulting in hypermethylation and silencing of the gene, although other loss-of-function mutations in the FMR1 coding regions have also been identified in rare cases of FXS [3,4,5,6]. Clinical studies in FXS patients combining the high-resolution FMR1 expression level analysis with neurobehavioral assessments indicate that FMRP expression in the brain is the ultimate factor determining the severity of the neurobehavioral phenotype in humans with FXS, including autism spectrum disorders (ASDs) [7]. Additionally, repeat-associated non-AUG-initiated (RAN) translation of expanded CGG repeats generates toxic proteins and contributes to neurodegeneration in fragile-X associated disorders [8]. FMRP is an RNA-binding protein that regulates the translation of specific mRNA targets. It also interacts with other RNA-binding proteins to regulate a variety of cellular functions such as translation, transport, and splicing of mRNAs [9,10,11,12,13,14]. FMRP is abundantly expressed in the brain and is involved in synaptic functions [15]. Indeed, many of the mRNA targets and interacting proteins of FMRP are known to regulate the synapse structure and function [10,12,16,17,18,19,20]. The altered synaptic function and plasticity have been reported in various regions of the Fmr1 KO mouse brain, including altered basal excitatory and inhibitory synaptic transmission, phasic and/or tonic GABAergic inhibition, certain forms of LTP and metabotropic glutamate receptor (mGluR)-mediated LTD [21,22,23,24,25,26,27,28]. Recent reports further establish impaired homeostatic synaptic plasticity as another major synaptic plasticity phenotype in FXS [29,30,31].
Homeostatic synaptic plasticity is a mechanism by which neurons maintain network stability in response to prolonged changes in network activity deviating from its optimal dynamic range [32,33,34]. For example, the prolonged blockade of neuronal activity leads to a general increase in excitatory and a decrease in inhibitory synaptic strengths. At excitatory synapses, changes in synaptic strength are often mediated by an altered synaptic abundance of a major type of glutamate receptor, the α-amino-3-hydroxy-5-methyl-4-isoxazoleprprionic acid receptors (AMPARs). Two major molecular mechanisms have been described to drive changes in AMPAR synaptic expressions in homeostatic synaptic plasticity—a retinoic acid (RA)-mediated translation-dependent mechanism responsible for an early homeostatic adaptation [35], and a CaMKIV-mediated transcription-dependent mechanism for a late phase homeostatic compensation [36]. Interestingly, the loss of FMRP in both mouse and human neurons results in the loss of RA-dependent homeostatic synaptic plasticity, while having no effect on the late phase transcription-dependent homeostatic plasticity [29,30,31].
RA is a well-known developmental morphogen known for its role in neurogenesis and neuronal development [37]. Beyond its function in development, RA is also the key molecule mediating translation-dependent homeostatic synaptic plasticity in mature neurons [29,38,39]. Suppressed under normal synaptic activity, the synthesis of RA is triggered by a significant reduction in intracellular calcium levels (i.e., during prolonged blockade of excitatory synaptic transmission) [40,41]. RA receptors (RARα, RARβ, and RARγ) are nuclear receptors that regulate RA-dependent transcriptional events during development [37]. In the context of homeostatic plasticity, RARα, which is broadly expressed in adult brains, acts as a translational repressor that mediates the non-genomic actions of RA. Specifically, the distribution of RARα shifts from a strictly nuclear-localized pattern during development towards a more cytosolic-localized pattern in mature neurons [38,42]. Cytosolic RARα functions as a RNA-binding protein that suppresses the translation of specific mRNAs recognized by RARα in a sequence-specific manner [38]. This translational repression can be removed once newly synthesized RA binds to RARα, leading to the activation of dendritic synthesis of specific proteins, such as the AMPAR subunit GluA1 [38]. Increased expression of GluA1 in dendrites supports the homeostatic increase in synaptic strength in response to prolonged synaptic inactivity. Previous EM studies show that RA stimulation recruits RARα to FMRP-containing granules, and subsequent GluR1 synthesis is triggered in these granules [42]. In Fmr1 KO neurons, the loss of FMRP abolishes the RA-induced increase in protein synthesis and subsequent homeostatic changes at both excitatory and inhibitory synapses [30], indicating that FMRP plays a major role in RA-dependent homeostatic synaptic plasticity. However, it remains unclear how FMRP and RARα interact functionally to regulate the protein synthesis process required for RA-induced homeostatic synaptic plasticity.
In this study, we investigated the physical interaction between FMRP and RARα and found that RARα directly binds to FMRP, and that the binding is enhanced in the presence of RA. Moreover, we established that this direct interaction is critical for RA-induced homeostatic synaptic plasticity.

2. Materials and Methods

2.1. Animals

All breeding colonies were maintained in the animal facility at Stanford Medical School following standard procedures approved by the Stanford University Administrative Panel on Laboratory Animal Care. Fmr1 knock-out mice (JAX: 004624 RRID:IMSR_JAX:004624) in the FVB background were obtained from The Jackson Laboratory. The RARα floxed mice were a gift from Drs. Pierre Chambon and Norbert Ghyselinck (IGBMC, Strasbourg, France) [43].
The Fmr1−/y::RARαfl/fl double mutant mice were generated by crossing the Fmr1 knock-out mice with the RARαfl/fl mice. Genotyping was performed with the following primer sets. For RARα WT, Primer 1 (fwd) 5′-GTGTGTGTGTGTATTCGCGTGC-3′, Primer 2 (rev) 5′-ACAAAGCAAGGCTTGTAGATGC-3′; for RARα flox, Primer 1 with Primer 3 (rev) 5′-TACACTAACTACCCTTGACC-3′; for FMRP WT, WT (fwd): 5′-GTGGTTAGCTAAAGTGAGGATGAT-3′, WT (rev): 5′-CAGGTTTGTTGGGATTAACAGATC-3′; and for FMRP KO, KO (fwd) 5′-CACGAGACTAGTGAGACGTG-3′ and KO (rev) 5′-CTTCTGGCACCTCCAGCTT-3′.

2.2. Primary Hippocampal Cultures

Primary hippocampal cultures were made from postnatal day 0 (P0) or P1 Fmr1+/y::RARαfl/fl and Fmr1−/y::RARαfl/fl mice of both sexes as previously described [30,44]. The primary cultured hippocampal neurons were maintained in Neurobasal media supplemented with B-27 and Glutamax (Invitrogen, Waltham, MA, USA). The neurons were transduced with lentivirus at DIV5 or 6 and harvested at DIV14–15.

2.3. Constructs

Recombinant protein expression constructs: All RARα constructs were cloned from rat sequences into pGEX-KG. Construction of full-length GST-RARα, GST-DBD, and GST-LBD was previously described [38]. The truncations of GST-DBD were PCR amplified from full-length rat RARα and cloned into pGEX-KG using BamHI and EcoRI: For GST-C/D, nucleotides 241–567; for GST-tC/D, nucleotides 361–567; and for GST-D, nucleotides 469–567. The FMRP-6XHis construct was amplified by PCR from FMRP-GFP and cloned into pET28a (Novagen, Madison, WI, USA) using EcoRI and NotI. The blocking peptide (nucleotides 361–567 of GST-RARα) and control peptide (nucleotides 4–210 of GST-RARα) were amplified by PCR from full-length rat RARα and cloned in pET28a (Novagen, Madison, WI, USA) using BamHI and EcoRI.
Lentiviral constructs: The FMRP-GFP, I304N-GFP, and ΔRGG-GFP in pJHUG were previously described [30]. The truncation mutants of FMRP in pJHUG were PCR amplified using FMRP-GFP as a template and the following nucleotides were cloned into pJHUG using EcoRI and NotI; for 128–614-GFP, nucleotides 382–1842; for 218–614-GFP, nucleotides 652–1842; for 438–614-GFP, nucleotides 1318–1842; and for 1–219-GFP, nucleotides 1–654.
The lentiviral plasmids expressing the blocking peptide or control peptide fused to the C-terminus of GFP were constructed by inserting the same nucleotides from the His-tagged recombinant protein expression plasmids described above into pFUGW using BsrGI and EcoRI.
Flag-RARα was cloned by EcoRI into the L309 already containing IRES-GFP-CRE (gift from Dr. Thomas Sudhof’s lab) for lentiviral expression of Flag-RARα and GFP-CRE under human synapsin promoter (SYN-Flag-RARα-IRES-GFP-CRE).
Mammalian expression constructs: The RARα-GFP in pEGFP-N1 was previously described [39]. The RARαΔtC/D-GFP was constructed by overlap PCR from rat RARα to delete nucleotides 361–567. The resulting fragment was cloned into pEGFP-N1 (Clontech, Mountain View, CA, USA) using XhoI and EcoRI.

2.4. Fusion Protein Production

GST purification: The GST-RARα and truncation fusion proteins were isolated as previously described [38]. Briefly, the GST constructs were transformed into BL21(DE3) cells (NEB, Ipswich, MA, USA). Cells were grown to OD 600 of 0.5 then induced with 1 mM IPTG (Goldbio, Saint Louis, MO, USA) for 2–4 h at 37 °C. The cells were pelleted and lysed in the lysis buffer (20 mM sodium phosphate 7.4, 150 mM NaCl, 1% TX-100, protease cocktail (Roche, Indianapolis, IN, USA) containing lysozyme and universal DNAse (Pierce, Waltham, MA, USA)) at 4 °C for 30 min. Debris was cleared by centrifuging at 15,000 rpm for 20 min. The supernatant was then incubated with glutathione sepharose beads (Amersham, Piscataway, NL, USA) end over end at 4 °C for 2 h. The beads were then washed 5 times with a lysis buffer. Protein expression and purification was confirmed by SDS-PAGE followed by Coomassie staining or Ponceau S staining.
His-tagged purification: Isolation of FMRP-His tagged recombinant protein was the same as above except the FMRP lysis buffer was composed of 10 mM HEPES pH 7.5, 300 mM LiCl, 5 mM β-mercaptoethanol, 10 mM imidazole, 5% glycerol and protease inhibitor cocktail pellets (Roche, Indianapolis, IN, USA), and the supernatant was incubated with Ni-NTA agarose (Qiagen, Redwood City, CA, USA) rather than glutathione-sepharose beads. The beads were washed 3 times with the FMRP-lysis buffer containing 100 mM imidazole (rather than 10 mM). Purified peptides were eluted from beads with 500 mM imidazole. Protein expression, purification, and elution were confirmed by SDS-PAGE followed by Coomassie staining or Ponceau S staining.
Isolation of the blocking and control His-tagged peptides was performed as above using the following His-lysis buffer (50 mM NaH2PO4 pH 8.0, 300 mM NaCl, 10 mM imidazole, and protease inhibitors (Roche, Indianapolis, IN, USA)) and incubated with Ni-NTA agarose (Qiagen, Redwood City, CA, USA) for purification. The beads were washed 3 times with the His-lysis buffer containing 100 mM imidazole (rather than 10 mM). Purified peptides were eluted from beads with 500 mM imidazole. Protein expression, purification, and elution were confirmed by SDS-PAGE followed by Coomassie staining.

2.5. GST Pull-Down

Equal amounts of GST-purified recombinant proteins on glutathione-sepharose beads were incubated with either HEK cell lysate, lysate from Fmr1 KO mouse embryonic fibroblast (MEF) cells or with equal amounts of purified T7-FMRP-His at 4 °C for 1 h end over end. The beads were then washed with the lysis buffer 5 times before eluting the protein in the 1X sample buffer for Western blot analysis.

2.6. Immunoprecipitation

Primary hippocampal neuron IP: Fmr1+/y::RARαfl/fl or Fmr−/y::RARαfl/fl mice were used to culture primary hippocampal neurons. At DIV 5–6, the neurons were infected with lentivirus expressing Flag-RARα and CRE-GFP to knockout endogenous RARα. The infected neurons were harvested on DIV 14-15. The neurons were first washed with 1X PBS and then lysed for 30 min at 4 °C in the IP lysis buffer (50 mM Tris pH 7.3, 150 mM NaCl, 1% Triton X-100, 1 mM NaF, 1 mM sodium orthovanadate (NEB, Ipswich, MA, USA), and protease inhibitor cocktail 2 (Sigma, Saint Louis, MO, USA)). After lysing, samples were centrifuged at 14,000 rpm for 15 min at 4 °C. Then, the supernatant was removed and incubated with dynabeads (Invitrogen, Waltham, MA, USA) to preclear the lysate for 1 h at 4 °C. The precleared lysate was incubated with a M2 Flag antibody (mouse, Sigma, Saint Louis, MO, USA) or FMRP antibody (rabbit, Abcam, Cambridge, MA, USA) for 1 h at 4 °C, to which dynabeads (Invitrogen, Waltham, MA, USA) were added and further incubated at 4 °C for 1 h. The beads were washed 5 times with the IP lysis buffer. The proteins were eluted from the beads with the 1X SDS loading buffer then run on SDS-PAGE for Western blot analysis.
HEK cell IP: HEK293T cells were cultured in 6-well dishes and transfected with 1 μg of pEGFPN-1 or RARα-pEGFPN-1 by calcium phosphate. The media was changed after 4 h. The following day the transfected HEK293T cells were treated with 1 μM of RA (Sigma, Saint Louis, MO, USA) for 1 h at 37 °C before harvesting. The cells were washed 2 times with PBS before harvesting in the IP-lysis buffer. In addition, 1 μM RA was added to the IP-lysis buffer for RA treated samples throughout the experiment to maintain a constant level of RA. After centrifuging, the lysate was incubated for 1 h with anti-GFP magnetic beads (MBL International, Woburn, MA, USA) at 4 °C. The beads were washed 5 times with the IP lysis buffer and proteins were eluded with a 1X SDS-loading buffer before the Western blot analysis.

2.7. Western Blot

Samples were run on 7%, 10%, 4–15% or 12.5% SDS-PAGE and transferred to either nitrocellulose or PVDF membranes. The membranes were blocked with 5% milk in TBS-T (1X Tris buffered saline containing 0.3% Tween-20) for 30 min at room temperature. Membranes were then incubated with a primary antibody diluted in 5% milk for 1 h at room temperature or 4 °C overnight. The primary antibody was washed 3 times for 10 min each with TBS-T, then incubated in the secondary diluted in 5% milk for 1 h at room temperature. The secondary was washed 3 times for 10 min each with TBS-T. The signal was detected by either ECL followed by development with film or scanning on Licor for the quantitative analysis.

2.8. Antibodies

The following monoclonal antibodies were used in this study: FMRP 1C3 (WB, 1:1000, Millipore, Hayward, CA, USA), FMRP 7G1-1 (WB, 1:500, DSHB), and Flag (IP, Sigma, Saint Louis, MO, USA). The following rabbit polyclonal antibodies were used in this study: Flag (WB, 1:1000, Sigma, Saint Louis, MO, USA), GFP (WB, 1:5000, Abcam, Cambridge, MA, USA), and FMRP (WB, IP, 1:1000, Abcam, Cambridge, MA, USA). A chicken polyclonal antibody for GFP (WB, 1:5000, Abcam, Cambridge, MA, USA) was also used. The antibody for T7 was directly conjugated to HRP (WB, 1:5000, Millipore, Hayward, CA, USA). Anti-mouse IgG-HRP, anti-rabbit IgG-HRP, and protein-HRP (1:5000, Jackson Labs, West Grove, PA, USA) were used for ECL detected Western blots. For Licor imaging, IRDye anti-mouse IgG-800CW and IRDye anti-rabbit IgG-680RD was used (1:10,000, Licor, Lincoln, NE, USA).

2.9. Organotypic Hippocampal Slice Cultures

Organotypic slice cultures were prepared from young RARα conditional knockout mice (P6 to P7) and placed on semi-porous membranes (Millipore, Hayward, CA, USA) for 10 to 12 days prior to recording [45,46]. Briefly, slices were maintained in a MEM based culture media comprised of 1 mM CaCl2, 2 mM MgSO4, 1 mM L-glutamine, 1 mg/L insulin, 0.0012% ascorbic acid, 30 mM HEPES, 13 mM D-glucose, and 5.2 mM NaHCO3. Culture media was a pH of 7.25 and the osmolarity was 320. Cultures were maintained in an incubator with 95% O2/5% CO2 at 34 °C.

2.10. Viral Vectors and Viral Infection

Recombinant lentivirus was produced as previously described [39,47]. Briefly, using calcium phosphate, HEK293T cells were transfected with four plasmids, the lentiviral shuttle vectors (pVSVg, pMDL, and pREV) and transfer vector. The HEK293 culture media was collected 40–44 h after transfection and filtered with the 0.45 μm PVDF filter (Millipore, Hayward, CA, USA) to remove cellular debris followed by centrifugation through a 20% sucrose cushion at 50,000x g to concentrate the virus. The concentrated virus was dissolved in a small volume of PBS and stored at −80 °C. Lentiviruses expressing either the blocking peptide or control peptide fused to GFP [48] were injected into the CA1 region of organotypic hippocampal slice cultures at DIV 0 and expressed for 10 days. The viral injection was carried out using a picospritzer with 18-20 PSI coupled to a micropipette with a resistance of 8–10 MOhms. The virus was delivered into the CA1 cell body layer at 4–5 injection sites with 4–5 pulses (5 ms duration) per injection site. Recordings from the infected neurons and their uninfected neighboring neurons were carried out at DIV 10–12.

2.11. Electrophysiology

Voltage-clamp whole-cell recordings were obtained from CA1 pyramidal neurons that were either uninfected or sparsely infected with constructs expressing control and blocking peptide fused to GFP, treated with either vehicle controls or 10 µM RA (for 4 h prior to recording), under visual guidance using transmitted light illumination. The vehicle control and RA treated cells were obtained from the same batches of slices on the same experimental day. A spontaneous miniature transmission was obtained in the presence of 1 μM of TTX and 100 μM picrotoxin in the external solution. For slices previously exposed to RA, the slices were washed out prior to recording.
The recording chamber was perfused with 119 mM NaCl, 2.5 mM KCl, 4 mM CaCl2, 4 mM MgCl2, 26 mM NaHCO3, 1 mM NaH2PO4, 11 mM glucose, 0.1 mM picrotoxin, and 1 μM TTX at pH 7.4, gassed with 5% CO2/95% O2 and held at 30 °C. Patch recording pipettes (3–6 MOhms) were filled with 115 mM cesium methanesulfonate, 20 mM CsCl, 10 mM HEPES, 2.5 mM MgCl2, 4 mM Na2ATP, 0.4 mM Na3GTP, 10 mM sodium phosphocreatine, and 0.6 mM EGTA at pH 7.25.
All electrophysiological recordings were carried out with Multiclamp 700B amplifiers (Axon Instruments, Perth, Western Australia) and the analysis was completed in Clampfit 10.2 (Molecular Devices, San Jose, CA, USA), MiniAnalysis version 6 (Synaptosoft, Decatur, GA, USA), Excel (Microsoft, Redmond, WA, USA), and Prism 10 (GraphPad, San Diego, CA, USA).

2.12. Statistical Analysis

All graphs represent average values ± SEM. Statistical differences were calculated according to parametric (for datasets with normal distribution) or nonparametric (for datasets that do not conform to normal distribution) tests using GraphPad Prism 9.0 software. Comparisons between multiple groups were performed with the Kruskal-Wallis ANOVA. When significant differences were observed, p-values for pairwise comparisons were calculated according to two-tailed Mann-Whitney tests (for unpaired data). For RA-mediated effects on interaction, p-values were calculated by the student’s t-test. All experiments were repeated at least 3 times.

3. Results

3.1. RARα Interacts Directly with FMRP

We have shown previously that synaptic inactivity-triggered RA synthesis is normal in Fmr1 KO mouse neurons [30], suggesting that the impaired RA-dependent homeostatic synaptic plasticity in these neurons is likely caused by alterations in the signaling pathway downstream of RA. Given that both RARα and FMRP are RNA-binding proteins that can be localized to the same RNA granules [42], we wondered whether they interact directly. First, in reduced systems, incubation of purified bacterial-expressed GST-RARα fusion protein with HEK cell lysates successfully pulled down endogenous FMRP from HEK cells (Figure 1A) and from P30 mouse hippocampal lysate (Figure 1B). Additionally, purified T7- and His-tagged FMRP was found associated with purified GST-RARα in an in vitro binding assay, suggesting that this interaction is indeed a direct binding between the two proteins (Figure 1C).
Next, we tested whether the FMRP-RARα interaction occurs in hippocampal neurons. To circumvent the problem of lacking a reliable and specific antibody against RARα, we took advantage of the conditional knockout RARα mice (RARαfl/fl), in which molecular and functional replacement of endogenous RARα with an epitope-tagged RARα can be achieved [49]. Moreover, we used the constitutive Fmr1 KO mice as a control for nonspecific binding that could occur in immunoprecipitation assays. We cultured primary hippocampal neurons from Fmr1+/y::RARαfl/fl and Fmr1−/y::RARαfl/fl mice. Transducing these cultured neurons with a lentivirus expressing both a Flag-tagged RARα and a GFP-tagged Cre recombinase driven by the human Synapsin promoter led to expression of Flag-RARα in the absence of endogenous RARα in hippocampal neurons with WT or FMRP KO background. Immunoprecipitation with anti-FMRP antibodies pulled down FMRP as well as RARα in the WT but not Fmr1 KO neurons (Figure 1D), indicating high specificity of the FMRP antibody. Likewise, immunoprecipitation with anti-Flag antibodies pulled down RARα as well as FMRP in only Fmr1 WT neurons (Figure 1E). Taken together, these data demonstrate that RARα interacts directly with FMRP in hippocampal neurons.

3.2. Identification of Domains Necessary for RARα-FMRP Binding

Among the six distinct domains within the RARα protein, the A/B and C domains are important for transcriptional activity and for DNA binding (Figure 2A) [37]. The D domain contains the nuclear localization signal (NLS) and part of the hinge region (Figure 2A). The E domain binds to retinoic acid and the F domain is critical for mRNA binding (Figure 2A) [38]. To determine which regions are important for RARα binding to FMRP, we first generated GST-fusion proteins of two RARα truncation mutants that include the DNA binding domain (DBD) or the ligand-binding domain (LBD) (Figure 2A). Incubation of these purified GST-RARα truncation mutants with P30 mouse hippocampal lysates demonstrated a clear binding between FMRP and RARα DBD but not the LBD domain (Figure 2B). Further dissection of subregions within DBD using additional RARα truncation mutants isolated the amino acids 120–155 of RARα as the critical region for its interaction with FMRP (Figure 2B).
To further verify the critical requirement of the identified region in RARα for binding, we expressed GFP-tagged wildtype RARα or RARα-ΔtC/D (Δ120–189 a.a.) in HEK cells and examined their co-immunoprecipitation with endogenously expressed FMRP. As expected, both GFP-RARα and RARα-ΔtC/D were abundantly present in the protein complex pulled down by a GFP antibody, but FMRP was only present in complexes pulled down from GFP-WT-RARα-expressing cells, indicating that deletion of this region blocked the interaction between RARα and FMRP (Figure 2C).
Next, we investigated which domains of FMRP are essential for its interaction with RARα. FMRP consists of five domains that are critical for its function. Both KH-domains and the RGG box have been shown to have RNA-binding activity. The other two domains, the N-terminal domain of FMRP (NDF) and the protein-protein interaction domain (PPID), are known to mediate the FMRP interactions with other proteins such as CYFIP1, NuFIP2, FXR1, etc. [18,50,51,52] (Figure 3A). Thus, we generated various FMRP truncation mutants and point mutations (Figure 3A) and expressed them in mouse embryonic fibroblasts (MEFs) generated from Fmr1 KO mouse (hence no interference from endogenously expressed WT FMRP). MEF lysates were incubated with purified GST-RARα in a pull-down assay. Previous work has shown that mutations that abolish the FMRP RNA binding activity and its association with polyribosomes (e.g., I304N mutation in KH2 RNA-binding domain or RGG box deletion) disrupt RA-mediated homeostatic synaptic plasticity [30,53,54,55,56,57]. Surprisingly, both FMRP-∆RGG and FMRP (I304N) were still able to bind to RARα. However, deletion of the PPID abolished the interaction of FMRP with RARα (Figure 3B). Moreover, a truncated form of FMRP containing only the NDF and PPID was able to interact with RARα (Figure 3B), suggesting that the PPID is necessary and sufficient for the FMRP interaction with RARα.

3.3. RA Enhances the Interaction between RARα and FMRP

Our previous EM study showed that RARα is recruited to FMRP-containing dendritic granules shortly after RA stimulation [42], raising the possibility that RA may influence the interaction between RARα and FMRP. We tested this in HEK cells expressing endogenous FMRP and a N-terminal GFP-tagged RARα. Immunoprecipitation with a GFP antibody pulled down RARα as well as FMRP in GFP-RARα expressing but not GFP expressing control HEK cells (Figure 4A). Treating the HEK cells with 1 μM RA for an hour significantly increased the pull-down efficiency (Figure 4B), indicating that RA enhances the interaction between FMRP and RARα.

3.4. Interaction between FMRP and RARα Is Critical for RA-Dependent Synaptic Plasticity

What might the FMRP-RARα interaction be important for? Previous studies showed that in Fmr1 KO mouse neurons, the lack of FMRP expression leads to a complete loss of RA-dependent homeostatic synaptic plasticity [29,30]. In these studies, impaired homeostatic synaptic plasticity in FMRP-∆RGG- and I304N-expressing FMRP KO neurons indicate that signaling pathways downstream of FMRP/RARα binding were required for RA-dependent synaptic plasticity since these two FMRP mutants have normal binding to RARα. However, this does not exclude the possibility that normal FMRP/RARα binding is also required for RA-dependent synaptic plasticity. Here, we sought to test whether blocking the FMRP-RARα interaction affects RA-dependent synaptic plasticity.
We generated a His-tagged peptide of the FMRP binding region in RARα to block the FMRP-RARα interaction (blocking peptide, Figure 5A). A His-tagged control peptide, generated with a sequence outside of the FMRP binding region and of similar molecular size, served as a control. The addition of purified bacterial-expressed blocking peptide to the mixture of purified GST-RARα and T7-FMRP-His significantly reduced the interaction between RARα and FMRP compared to the interaction in the presence of the control peptide (Figure 5B).
After verifying the efficacy of the blocking peptide, we infected CA1 pyramidal neurons in the cultured wild type hippocampal slices with lentiviruses expressing GFP-tagged (to facilitate visualization of infected neurons) control and blocking peptides and examined the RA-induced increase in excitatory synaptic transmission in these neurons. Consistent with previous reports [39,45], the acute RA treatment (10 μM, 4 h) induced a rapid increase in excitatory synaptic responses in uninfected neurons as well as those infected with the control peptide, evidenced by a significant increase in the amplitude of miniature excitatory postsynaptic responses (mEPSCs) (Figure 5C,D). By contrast, neurons expressing the blocking peptide failed to respond to the RA treatment (Figure 5C,D). Thus, we established that the FMRP-RARα interaction is essential for RA-induced synaptic plasticity.

4. Discussion

Synaptic RA signaling has been established as a major signaling pathway mediating synaptic inactivity-induced homeostatic synaptic plasticity [35,39]. Prolonged synaptic silencing triggers RA synthesis and RARα-mediated local translation of proteins essential for the expression of homeostatic synaptic plasticity. Previous studies suggest that FMRP may be critically involved in this form of homeostatic synaptic plasticity as in Fmr1 KO mouse neurons and fragile-X syndrome patient-derived human neurons, where RA-dependent homeostatic synaptic plasticity is absent [29,30,31]. In this study, we investigated the FMRP-RARα interaction as a potential molecular requirement for RA-dependent homeostatic synaptic plasticity in neurons. Using both in vitro and in situ biochemical binding assays, we demonstrated that FMRP and RARα directly bind to each other. Systematic mapping of the binding domains in both proteins allowed for us to identify regions critical for the direct interaction of RARα and FMRP. Subsequently, the generation of a peptide that blocks this interaction impaired homeostatic synaptic plasticity. Thus, we established that the FMRP-RARα interaction is required for RA-induced synaptic plasticity.
During RA-dependent homeostatic synaptic plasticity, a reduction in dendritic calcium levels due to reduced synaptic activity results in a decrease in calcineurin activity, which triggers RA synthesis [40,41]. Downstream of RA synthesis, complex molecular and cellular processes occur to realize multiple aspects of homeostatic adjustment at both pre- and postsynaptic sites. These processes include, but are not limited to, RA-induced disinhibition of protein synthesis from RARα-bound mRNAs, exocytosis of AMPA receptor (AMPAR)-containing vesicles and synaptic delivery of AMPARs, endocytosis of GABAA receptors (GABAAR), and enhancement of presynaptic release probability through retrograde signaling [29,38,39,41,45]. The discovery of the requirement of FMRP in RA-dependent homeostatic synaptic plasticity adds additional complexity to the signaling pathways [29,30,31]. The RA synthesis in response to the activity blockade has been shown to be intact in FMRP KO neurons [30], placing the FMRP involvement at a step downstream of RA synthesis. Given that the major role of FMRP is an mRNA-binding protein and a translational regulator [9,10,11,17], and synaptic RA signaling requires RARα-mediated local protein synthesis, it is perhaps not surprising that FMRP mutants that are impaired in translational regulation of target mRNAs (FMRP-∆RGG and I304N) fail to rescue RA-dependent homeostatic synaptic plasticity [30]. However, results from this study indicate that the FMRP involvement in synaptic plasticity may be multi-layered, and that a direct binding between FMRP and RARα is necessary (albeit not sufficient since both ∆RGG- and I304N mutants of FMRP have normal binding to RARα).
What might the FMRP-RARα interaction be required for? One possibility is protein trafficking and dendritic targeting. Similar to many nuclear receptors that act as transcription factors, RARα has two nuclear localization signals (NLS) that are critical for its nuclear targeting [38]. Although mostly confined within the nucleus of neurons during development, the RARα distribution shifts as neurons mature, displaying a gradual increase in cytoplasmic localization [38,58]. A nuclear export signal (NES) in RARα is required for its translocation into the cytosol as RARα bearing a NES mutation is strictly localized to the nucleus [38]. Interestingly, the two NLS are located in domains C and D of RARα, a region where FMRP binds. FMRP contains both NLS and NES signals and shuttles between the nucleus and cytoplasm [59,60]. It is conceivable that FMRP binding masks the NLS in RARα, facilitating its NES-mediated nuclear export.
Our previous study showed that FMRP is enriched in dendritic RNA granules in cultured hippocampal neurons [42], a prime location for dendritic protein synthesis. Importantly, although RARα is present only at low levels in RNA granules at basal condition, the treatment with RA or synaptic activity blocking reagents (e.g., TTX + APV) that induce de novo RA synthesis within neurons, triggers rapid translocation of RARα into FMRP-rich RNA granules [42]. Following RARα translocation, the synthesis of GluA1 protein ensues in these FMRP-RNA granules [42]. Interestingly, the results in this study show that the RA treatment enhances FMRP-RARα binding, suggesting a scenario that synaptic inactivity-induced RA synthesis facilitates translocation of RARα and its associated target mRNAs (e.g., GluA1 mRNA) into dendritic RNA granules by enhancing FMRP-RARα binding, permitting translational activation of specific mRNA species. This notion is consistent with the observation that although both FMRP-∆RGG- and FMRP-I304N mutants have normal binding with RARα, they do not rescue homeostatic synaptic plasticity in Fmr1 KO neurons since both mutants are defective in their incorporation into polyribosomes or RNA granules [13,53,54,61,62,63,64].
The connection between FMRP and RARα not only expands the functional repertoire of FMRP to the regulation of homeostatic synaptic plasticity, but also brings up additional synaptic mechanisms that should be taken into consideration when investigating neural mechanisms underlying cognitive defects in FXS. Neurological diseases associated with intellectual disability are often attributed to changes in Hebbian synaptic plasticity because LTP/LTD are considered cellular mechanisms of learning and memory. However, increasing evidence show defective homeostatic synaptic plasticity in various ASD models [29,30,65,66,67,68]. In addition to the disrupted excitation/inhibition (E/I) balance in various ASD models (which may be a consequence of defective homeostatic plasticity) [69,70], defective homeostatic synaptic plasticity has been directly examined and reported in Mecp2 KO (a Rett syndrome model) [66,67,71], Fmr1 KO mice [29,30], and FXS human neurons [31]. Might homeostatic synaptic plasticity contribute to learning and memory through interaction with Hebbian plasticity? A recent study using RARα conditional KO mice showed that homeostatic synaptic plasticity indeed significantly impacts Hebbian plasticity and behavioral learning [72]. In human subjects, seizure (resulting from E/I imbalance) and intellectual disability are common symptoms affecting a fraction of patients with ASDs, Schizophrenia, and neurodegeneration disorders. Indeed, some of the disease-relevant biological pathways are interconnected between FXS, ASDs, and Schizophrenia [73,74,75]. Thus, understanding the molecular details of impaired homeostatic synaptic plasticity in ASD models may provide insights into discovering potential targets for therapeutic treatment of ASDs, including ASDs in FXS, and other neurodevelopmental disorders [68]. Those studies should incorporate further refined methodologies [76] to assess whether parameters that may underlay variability in FXS, such as FMRP levels and comorbid diagnosis of ASD [7], have an impact on the response to those potential targeted therapeutics.

Author Contributions

Conceptualization, A.G.L. and L.C.; methodology, A.G.L., E.P., K.L.A. and L.C.; investigation, A.G.L., E.P. and K.L.A.; data curation, A.G.L. and E.P.; writing—original draft preparation, A.G.L. and L.C.; writing—review and editing, A.G.L., E.P., K.L.A. and L.C.; supervision, L.C.; project administration, L.C.; funding acquisition, L.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Institute of Health (NIH), MH086403; NIH, NS11566001; NIH, HD084215; and NIH, HD104458.

Institutional Review Board Statement

The study was conducted according to the guidelines of NIH and approved by the Insitutional Animal Care and Use Committee of Stanford University (Protocol ID 26428, Approval date May 13, 2021).

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no competing interest.

References

  1. Santoro, M.R.; Bray, S.M.; Warren, S.T. Molecular mechanisms of fragile X syndrome: A twenty-year perspective. Annu. Rev. Pathol. 2012, 7, 219–245. [Google Scholar] [CrossRef] [Green Version]
  2. Bagni, C.; Tassone, F.; Neri, G.; Hagerman, R. Fragile X syndrome: Causes, diagnosis, mechanisms, and therapeutics. J. Clin. Investig. 2012, 122, 4314–4322. [Google Scholar] [CrossRef] [Green Version]
  3. Verkerk, A.J.; Pieretti, M.; Sutcliffe, J.S.; Fu, Y.H.; Kuhl, D.P.; Pizzuti, A.; Reiner, O.; Richards, S.; Victoria, M.F.; Zhang, F.P.; et al. Identification of a gene (FMR-1) containing a CGG repeat coincident with a breakpoint cluster region exhibiting length variation in fragile X syndrome. Cell 1991, 65, 905–914. [Google Scholar] [CrossRef]
  4. Suhl, J.A.; Warren, S.T. Single-Nucleotide Mutations in FMR1 Reveal Novel Functions and Regulatory Mechanisms of the Fragile X Syndrome Protein FMRP. J. Exp. Neurosci. 2015, 9, 35–41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Heitz, D.; Devys, D.; Imbert, G.; Kretz, C.; Mandel, J.L. Inheritance of the fragile X syndrome: Size of the fragile X premutation is a major determinant of the transition to full mutation. J. Med. Genet. 1992, 29, 794–801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Fu, Y.H.; Kuhl, D.P.; Pizzuti, A.; Pieretti, M.; Sutcliffe, J.S.; Richards, S.; Verkerk, A.J.; Holden, J.J.; Fenwick, R.G., Jr.; Warren, S.T.; et al. Variation of the CGG repeat at the fragile X site results in genetic instability: Resolution of the Sherman paradox. Cell 1991, 67, 1047–1058. [Google Scholar] [CrossRef]
  7. Budimirovic, D.B.; Schlageter, A.; Filipovic-Sadic, S.; Protic, D.D.; Bram, E.; Mahone, E.M.; Nicholson, K.; Culp, K.; Javanmardi, K.; Kemppainen, J.; et al. A Genotype-Phenotype Study of High-Resolution FMR1 Nucleic Acid and Protein Analyses in Fragile X Patients with Neurobehavioral Assessments. Brain Sci. 2020, 10, 694. [Google Scholar] [CrossRef] [PubMed]
  8. Castren, M.; Lampinen, K.E.; Miettinen, R.; Koponen, E.; Sipola, I.; Bakker, C.E.; Oostra, B.A.; Castren, E. BDNF regulates the expression of fragile X mental retardation protein mRNA in the hippocampus. Neurobiol. Dis. 2002, 11, 221–229. [Google Scholar] [CrossRef] [Green Version]
  9. Chen, L.; Yun, S.W.; Seto, J.; Liu, W.; Toth, M. The fragile X mental retardation protein binds and regulates a novel class of mRNAs containing U rich target sequences. Neuroscience 2003, 120, 1005–1017. [Google Scholar] [CrossRef]
  10. Darnell, J.C.; Van Driesche, S.J.; Zhang, C.; Hung, K.Y.; Mele, A.; Fraser, C.E.; Stone, E.F.; Chen, C.; Fak, J.J.; Chi, S.W.; et al. FMRP stalls ribosomal translocation on mRNAs linked to synaptic function and autism. Cell 2011, 146, 247–261. [Google Scholar] [CrossRef] [Green Version]
  11. Miyashiro, K.Y.; Beckel-Mitchener, A.; Purk, T.P.; Becker, K.G.; Barret, T.; Liu, L.; Carbonetto, S.; Weiler, I.J.; Greenough, W.T.; Eberwine, J. RNA cargoes associating with FMRP reveal deficits in cellular functioning in Fmr1 null mice. Neuron 2003, 37, 417–431. [Google Scholar] [CrossRef] [Green Version]
  12. Dictenberg, J.B.; Swanger, S.A.; Antar, L.N.; Singer, R.H.; Bassell, G.J. A direct role for FMRP in activity-dependent dendritic mRNA transport links filopodial-spine morphogenesis to fragile X syndrome. Dev. Cell 2008, 14, 926–939. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Wang, H.; Dictenberg, J.B.; Ku, L.; Li, W.; Bassell, G.J.; Feng, Y. Dynamic association of the fragile X mental retardation protein as a messenger ribonucleoprotein between microtubules and polyribosomes. Mol. Biol. Cell 2008, 19, 105–114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Didiot, M.C.; Tian, Z.; Schaeffer, C.; Subramanian, M.; Mandel, J.L.; Moine, H. The G-quartet containing FMRP binding site in FMR1 mRNA is a potent exonic splicing enhancer. Nucleic Acids Res. 2008, 36, 4902–4912. [Google Scholar] [CrossRef] [Green Version]
  15. Hinds, H.L.; Ashley, C.T.; Sutcliffe, J.S.; Nelson, D.L.; Warren, S.T.; Housman, D.E.; Schalling, M. Tissue specific expression of FMR-1 provides evidence for a functional role in fragile X syndrome. Nat. Genet. 1993, 3, 36–43. [Google Scholar] [CrossRef] [PubMed]
  16. Antar, L.N.; Bassell, G.J. Sunrise at the synapse: The FMRP mRNP shaping the synaptic interface. Neuron 2003, 37, 555–558. [Google Scholar] [CrossRef] [Green Version]
  17. Bassell, G.J.; Warren, S.T. Fragile X syndrome: Loss of local mRNA regulation alters synaptic development and function. Neuron 2008, 60, 201–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Fernandez, E.; Rajan, N.; Bagni, C. The FMRP regulon: From targets to disease convergence. Front. Neurosci. 2013, 7, 191. [Google Scholar] [CrossRef] [Green Version]
  19. Greenough, W.T.; Klintsova, A.Y.; Irwin, S.A.; Galvez, R.; Bates, K.E.; Weiler, I.J. Synaptic regulation of protein synthesis and the fragile X protein. Proc. Natl. Acad. Sci. USA 2001, 98, 7101–7106. [Google Scholar] [CrossRef] [Green Version]
  20. Ifrim, M.F.; Williams, K.R.; Bassell, G.J. Single-Molecule Imaging of PSD-95 mRNA Translation in Dendrites and Its Dysregulation in a Mouse Model of Fragile X Syndrome. J. Neurosci. 2015, 35, 7116–7130. [Google Scholar] [CrossRef] [Green Version]
  21. Tyzio, R.; Nardou, R.; Ferrari, D.C.; Tsintsadze, T.; Shahrokhi, A.; Eftekhari, S.; Khalilov, I.; Tsintsadze, V.; Brouchoud, C.; Chazal, G.; et al. Oxytocin-mediated GABA inhibition during delivery attenuates autism pathogenesis in rodent offspring. Science 2014, 343, 675–679. [Google Scholar] [CrossRef]
  22. Olmos-Serrano, J.L.; Paluszkiewicz, S.M.; Martin, B.S.; Kaufmann, W.E.; Corbin, J.G.; Huntsman, M.M. Defective GABAergic neurotransmission and pharmacological rescue of neuronal hyperexcitability in the amygdala in a mouse model of fragile X syndrome. J. Neurosci. 2010, 30, 9929–9938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Gibson, J.R.; Bartley, A.F.; Hays, S.A.; Huber, K.M. Imbalance of neocortical excitation and inhibition and altered UP states reflect network hyperexcitability in the mouse model of fragile X syndrome. J. Neurophysiol. 2008, 100, 2615–2626. [Google Scholar] [CrossRef] [PubMed]
  24. Curia, G.; Papouin, T.; Seguela, P.; Avoli, M. Downregulation of tonic GABAergic inhibition in a mouse model of fragile X syndrome. Cereb. Cortex 2009, 19, 1515–1520. [Google Scholar] [CrossRef] [Green Version]
  25. Centonze, D.; Rossi, S.; Mercaldo, V.; Napoli, I.; Ciotti, M.T.; De Chiara, V.; Musella, A.; Prosperetti, C.; Calabresi, P.; Bernardi, G.; et al. Abnormal striatal GABA transmission in the mouse model for the fragile X syndrome. Biol. Psychiatry 2008, 63, 963–973. [Google Scholar] [CrossRef] [PubMed]
  26. Li, J.; Pelletier, M.R.; Perez Velazquez, J.L.; Carlen, P.L. Reduced cortical synaptic plasticity and GluR1 expression associated with fragile X mental retardation protein deficiency. Mol. Cell. Neurosci. 2002, 19, 138–151. [Google Scholar] [CrossRef] [PubMed]
  27. Larson, J.; Jessen, R.E.; Kim, D.; Fine, A.K.; du Hoffmann, J. Age-dependent and selective impairment of long-term potentiation in the anterior piriform cortex of mice lacking the fragile X mental retardation protein. J. Neurosci. 2005, 25, 9460–9469. [Google Scholar] [CrossRef] [Green Version]
  28. Huber, K.M.; Gallagher, S.M.; Warren, S.T.; Bear, M.F. Altered synaptic plasticity in a mouse model of fragile X mental retardation. Proc. Natl. Acad. Sci. USA 2002, 99, 7746–7750. [Google Scholar] [CrossRef] [Green Version]
  29. Sarti, F.; Zhang, Z.; Schroeder, J.; Chen, L. Rapid suppression of inhibitory synaptic transmission by retinoic acid. J. Neurosci. 2013, 33, 11440–11450. [Google Scholar] [CrossRef]
  30. Soden, M.E.; Chen, L. Fragile X protein FMRP is required for homeostatic plasticity and regulation of synaptic strength by retinoic acid. J. Neurosci. 2010, 30, 16910–16921. [Google Scholar] [CrossRef] [Green Version]
  31. Zhang, Z.; Marro, S.G.; Zhang, Y.; Arendt, K.L.; Patzke, C.; Zhou, B.; Fair, T.; Yang, N.; Sudhof, T.C.; Wernig, M.; et al. The fragile X mutation impairs homeostatic plasticity in human neurons by blocking synaptic retinoic acid signaling. Sci. Transl. Med. 2018, 10, eaar4338. [Google Scholar] [CrossRef] [Green Version]
  32. Davis, G.W.; Muller, M. Homeostatic control of presynaptic neurotransmitter release. Annu. Rev. Physiol. 2015, 77, 251–270. [Google Scholar] [CrossRef]
  33. Turrigiano, G. Homeostatic synaptic plasticity: Local and global mechanisms for stabilizing neuronal function. Cold Spring Harb. Perspect. Biol. 2012, 4, a005736. [Google Scholar] [CrossRef] [Green Version]
  34. Vitureira, N.; Letellier, M.; Goda, Y. Homeostatic synaptic plasticity: From single synapses to neural circuits. Curr. Opin. Neurobiol. 2012, 22, 516–521. [Google Scholar] [CrossRef] [Green Version]
  35. Chen, L.; Lau, A.G.; Sarti, F. Synaptic retinoic acid signaling and homeostatic synaptic plasticity. Neuropharmacology 2014, 78, 3–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Joseph, A.; Turrigiano, G.G. All for One But Not One for All: Excitatory Synaptic Scaling and Intrinsic Excitability Are Coregulated by CaMKIV, Whereas Inhibitory Synaptic Scaling Is Under Independent Control. J. Neurosci. 2017, 37, 6778–6785. [Google Scholar] [CrossRef] [PubMed]
  37. Chambon, P. A decade of molecular biology of retinoic acid receptors. FASEB J. 1996, 10, 940–954. [Google Scholar] [CrossRef]
  38. Poon, M.M.; Chen, L. Retinoic acid-gated sequence-specific translational control by RARalpha. Proc. Natl. Acad. Sci. USA 2008, 105, 20303–20308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Aoto, J.; Nam, C.I.; Poon, M.M.; Ting, P.; Chen, L. Synaptic signaling by all-trans retinoic acid in homeostatic synaptic plasticity. Neuron 2008, 60, 308–320. [Google Scholar] [CrossRef] [Green Version]
  40. Arendt, K.L.; Zhang, Z.; Ganesan, S.; Hintze, M.; Shin, M.M.; Tang, Y.; Cho, A.; Graef, I.A.; Chen, L. Calcineurin mediates homeostatic synaptic plasticity by regulating retinoic acid synthesis. Proc. Natl. Acad. Sci. USA 2015, 112, E5744–E5752. [Google Scholar] [CrossRef] [Green Version]
  41. Wang, H.L.; Zhang, Z.; Hintze, M.; Chen, L. Decrease in calcium concentration triggers neuronal retinoic acid synthesis during homeostatic synaptic plasticity. J. Neurosci. 2011, 31, 17764–17771. [Google Scholar] [CrossRef] [Green Version]
  42. Maghsoodi, B.; Poon, M.M.; Nam, C.I.; Aoto, J.; Ting, P.; Chen, L. Retinoic acid regulates RARalpha-mediated control of translation in dendritic RNA granules during homeostatic synaptic plasticity. Proc. Natl. Acad. Sci. USA 2008, 105, 16015–16020. [Google Scholar] [CrossRef] [Green Version]
  43. Chapellier, B.; Mark, M.; Garnier, J.M.; LeMeur, M.; Chambon, P.; Ghyselinck, N.B. A conditional floxed (loxP-flanked) allele for the retinoic acid receptor alpha (RARalpha) gene. Genesis 2002, 32, 87–90. [Google Scholar] [CrossRef] [PubMed]
  44. Aoto, J.; Martinelli, D.C.; Malenka, R.C.; Tabuchi, K.; Sudhof, T.C. Presynaptic neurexin-3 alternative splicing trans-synaptically controls postsynaptic AMPA receptor trafficking. Cell 2013, 154, 75–88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Arendt, K.L.; Zhang, Y.; Jurado, S.; Malenka, R.C.; Sudhof, T.C.; Chen, L. Retinoic Acid and LTP Recruit Postsynaptic AMPA Receptors Using Distinct SNARE-Dependent Mechanisms. Neuron 2015, 86, 442–456. [Google Scholar] [CrossRef] [Green Version]
  46. Gahwiler, B.H.; Capogna, M.; Debanne, D.; McKinney, R.A.; Thompson, S.M. Organotypic slice cultures: A technique has come of age. Trends Neurosci. 1997, 20, 471–477. [Google Scholar] [CrossRef]
  47. Arendt, K.L.; Sarti, F.; Chen, L. Chronic inactivation of a neural circuit enhances LTP by inducing silent synapse formation. J. Neurosci. 2013, 33, 2087–2096. [Google Scholar] [CrossRef] [Green Version]
  48. Kaeser, P.S.; Deng, L.; Wang, Y.; Dulubova, I.; Liu, X.; Rizo, J.; Sudhof, T.C. RIM proteins tether Ca2+ channels to presynaptic active zones via a direct PDZ-domain interaction. Cell 2011, 144, 282–295. [Google Scholar] [CrossRef] [Green Version]
  49. Sarti, F.; Schroeder, J.; Aoto, J.; Chen, L. Conditional RARalpha knockout mice reveal acute requirement for retinoic acid and RARalpha in homeostatic plasticity. Front. Mol. Neurosci. 2012, 5, 16. [Google Scholar] [CrossRef] [Green Version]
  50. Schenck, A.; Bardoni, B.; Moro, A.; Bagni, C.; Mandel, J.L. A highly conserved protein family interacting with the fragile X mental retardation protein (FMRP) and displaying selective interactions with FMRP-related proteins FXR1P and FXR2P. Proc. Natl. Acad. Sci. USA 2001, 98, 8844–8849. [Google Scholar] [CrossRef] [Green Version]
  51. Napoli, I.; Mercaldo, V.; Boyl, P.P.; Eleuteri, B.; Zalfa, F.; De Rubeis, S.; Di Marino, D.; Mohr, E.; Massimi, M.; Falconi, M.; et al. The fragile X syndrome protein represses activity-dependent translation through CYFIP1, a new 4E-BP. Cell 2008, 134, 1042–1054. [Google Scholar] [CrossRef] [Green Version]
  52. Bardoni, B.; Schenck, A.; Mandel, J.L. A novel RNA-binding nuclear protein that interacts with the fragile X mental retardation (FMR1) protein. Hum. Mol. Genet. 1999, 8, 2557–2566. [Google Scholar] [CrossRef]
  53. Feng, Y.; Absher, D.; Eberhart, D.E.; Brown, V.; Malter, H.E.; Warren, S.T. FMRP associates with polyribosomes as an mRNP, and the I304N mutation of severe fragile X syndrome abolishes this association. Mol. Cell 1997, 1, 109–118. [Google Scholar] [CrossRef]
  54. Zang, J.B.; Nosyreva, E.D.; Spencer, C.M.; Volk, L.J.; Musunuru, K.; Zhong, R.; Stone, E.F.; Yuva-Paylor, L.A.; Huber, K.M.; Paylor, R.; et al. A mouse model of the human Fragile X syndrome I304N mutation. PLoS Genet. 2009, 5, e1000758. [Google Scholar] [CrossRef] [Green Version]
  55. Darnell, J.C.; Jensen, K.B.; Jin, P.; Brown, V.; Warren, S.T.; Darnell, R.B. Fragile X mental retardation protein targets G quartet mRNAs important for neuronal function. Cell 2001, 107, 489–499. [Google Scholar] [CrossRef]
  56. Siomi, H.; Siomi, M.C.; Nussbaum, R.L.; Dreyfuss, G. The protein product of the fragile X gene, FMR1, has characteristics of an RNA-binding protein. Cell 1993, 74, 291–298. [Google Scholar] [CrossRef]
  57. Schaeffer, C.; Bardoni, B.; Mandel, J.L.; Ehresmann, B.; Ehresmann, C.; Moine, H. The fragile X mental retardation protein binds specifically to its mRNA via a purine quartet motif. EMBO J. 2001, 20, 4803–4813. [Google Scholar] [CrossRef] [PubMed]
  58. Huang, H.; Wei, H.; Zhang, X.; Chen, K.; Li, Y.; Qu, P.; Chen, J.; Liu, Y.; Yang, L.; Li, T. Changes in the expression and subcellular localization of RARalpha in the rat hippocampus during postnatal development. Brain Res. 2008, 1227, 26–33. [Google Scholar] [CrossRef] [PubMed]
  59. Eberhart, D.E.; Malter, H.E.; Feng, Y.; Warren, S.T. The fragile X mental retardation protein is a ribonucleoprotein containing both nuclear localization and nuclear export signals. Hum. Mol. Genet. 1996, 5, 1083–1091. [Google Scholar] [CrossRef] [Green Version]
  60. Kim, M.; Bellini, M.; Ceman, S. Fragile X mental retardation protein FMRP binds mRNAs in the nucleus. Mol. Cell. Biol. 2009, 29, 214–228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Darnell, J.C.; Fraser, C.E.; Mostovetsky, O.; Stefani, G.; Jones, T.A.; Eddy, S.R.; Darnell, R.B. Kissing complex RNAs mediate interaction between the Fragile-X mental retardation protein KH2 domain and brain polyribosomes. Genes Dev. 2005, 19, 903–918. [Google Scholar] [CrossRef] [Green Version]
  62. Blackwell, E.; Zhang, X.; Ceman, S. Arginines of the RGG box regulate FMRP association with polyribosomes and mRNA. Hum. Mol. Genet. 2010, 19, 1314–1323. [Google Scholar] [CrossRef] [Green Version]
  63. Mazroui, R.; Huot, M.E.; Tremblay, S.; Boilard, N.; Labelle, Y.; Khandjian, E.W. Fragile X Mental Retardation protein determinants required for its association with polyribosomal mRNPs. Hum. Mol. Genet. 2003, 12, 3087–3096. [Google Scholar] [CrossRef] [Green Version]
  64. Mazroui, R.; Huot, M.E.; Tremblay, S.; Filion, C.; Labelle, Y.; Khandjian, E.W. Trapping of messenger RNA by Fragile X Mental Retardation protein into cytoplasmic granules induces translation repression. Hum. Mol. Genet. 2002, 11, 3007–3017. [Google Scholar] [CrossRef] [PubMed]
  65. Zhong, L.; Chen, X.; Park, E.; Sudhof, T.C.; Chen, L. Retinoic acid receptor RARalpha-dependent synaptic signaling mediates homeostatic synaptic plasticity at the inhibitory synapses of mouse visual cortex. J. Neurosci. 2018, 38, 10454–10466. [Google Scholar] [CrossRef] [PubMed]
  66. Moretti, P.; Zoghbi, H.Y. MeCP2 dysfunction in Rett syndrome and related disorders. Curr. Opin. Genet. Dev. 2006, 16, 276–281. [Google Scholar] [CrossRef]
  67. Blackman, M.P.; Djukic, B.; Nelson, S.B.; Turrigiano, G.G. A critical and cell-autonomous role for MeCP2 in synaptic scaling up. J. Neurosci. 2012, 32, 13529–13536. [Google Scholar] [CrossRef] [PubMed]
  68. Budimirovic, D.B.; Kaufmann, W.E. What can we learn about autism from studying fragile X syndrome? Dev. Neurosci. 2011, 33, 379–394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Gao, R.; Penzes, P. Common mechanisms of excitatory and inhibitory imbalance in schizophrenia and autism spectrum disorders. Curr. Mol. Med. 2015, 15, 146–167. [Google Scholar] [CrossRef] [PubMed]
  70. Nelson, S.B.; Valakh, V. Excitatory/Inhibitory Balance and Circuit Homeostasis in Autism Spectrum Disorders. Neuron 2015, 87, 684–698. [Google Scholar] [CrossRef] [Green Version]
  71. Qiu, Z.; Sylwestrak, E.L.; Lieberman, D.N.; Zhang, Y.; Liu, X.Y.; Ghosh, A. The Rett syndrome protein MeCP2 regulates synaptic scaling. J. Neurosci. 2012, 32, 989–994. [Google Scholar] [CrossRef]
  72. Hsu, Y.T.; Li, J.; Wu, D.; Sudhof, T.C.; Chen, L. Synaptic retinoic acid receptor signaling mediates mTOR-dependent metaplasticity that controls hippocampal learning. Proc. Natl. Acad. Sci. USA 2019, 116, 7113–7122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Martinotti, G.; Di Iorio, G.; Marini, S.; Ricci, V.; De Berardis, D.; Di Giannantonio, M. Nerve growth factor and brain-derived neurotrophic factor concentrations in schizophrenia: A review. J. Biol. Regul. Homeost. Agents 2012, 26, 347–356. [Google Scholar] [PubMed]
  74. Stoll, G.; Pietilainen, O.P.H.; Linder, B.; Suvisaari, J.; Brosi, C.; Hennah, W.; Leppa, V.; Torniainen, M.; Ripatti, S.; Ala-Mello, S.; et al. Deletion of TOP3beta, a component of FMRP-containing mRNPs, contributes to neurodevelopmental disorders. Nat. Neurosci. 2013, 16, 1228–1237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Castren, M.L.; Castren, E. BDNF in fragile X syndrome. Neuropharmacology 2014, 76 Pt C, 729–736. [Google Scholar] [CrossRef]
  76. Duy, P.Q.; Budimirovic, D.B. Fragile X Syndrome: Lessons Learned from the Most Translated Neurodevelopmental Disorder in Clinical Trials. Transl. Neurosci. 2017, 8, 7–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. RARα interacts directly with FMRP. (A) Pull-down of endogenous FMRP from HEK cell lysate by purified GST-RARα coupled to Glutathione sepharose beads (top panel: FMRP immunoblot; bottom panel: Ponceau staining). Purified GST coupled to Glutathione sepharose beads was used as a negative control. (B) Pull-down of endogenous FMRP from P30 mouse hippocampal lysate by purified GST-RARα coupled to Glutathione sepharose beads (top panel: FMRP immunoblot; bottom panel: Commassie staining). Purified GST coupled to Glutathione sepharose beads was used as a negative control. (C) In vitro pull-down of purified T7-FMRP by purified GST-RARα but not GST (top panel: T7 immunoblot; lower panel: Ponceau stain). (D,E) Co-immunoprecipitation of endogenous FMRP with Flag-RARα from DIV14-15 primary hippocampal neurons using anti-FMRP antibodies (D) and anti-Flag antibodies (E). In both experiments, lentivirus expressing Flag-RARα-IRES-GFP-Cre in RARαfl/fl background was used to replace endogenous RARα with Flag-tagged RARα. Hippocampal cultures obtained from RARαffl/fl/Fmr1 KO mice (KO) were used as negative controls for specificity of pull-down.
Figure 1. RARα interacts directly with FMRP. (A) Pull-down of endogenous FMRP from HEK cell lysate by purified GST-RARα coupled to Glutathione sepharose beads (top panel: FMRP immunoblot; bottom panel: Ponceau staining). Purified GST coupled to Glutathione sepharose beads was used as a negative control. (B) Pull-down of endogenous FMRP from P30 mouse hippocampal lysate by purified GST-RARα coupled to Glutathione sepharose beads (top panel: FMRP immunoblot; bottom panel: Commassie staining). Purified GST coupled to Glutathione sepharose beads was used as a negative control. (C) In vitro pull-down of purified T7-FMRP by purified GST-RARα but not GST (top panel: T7 immunoblot; lower panel: Ponceau stain). (D,E) Co-immunoprecipitation of endogenous FMRP with Flag-RARα from DIV14-15 primary hippocampal neurons using anti-FMRP antibodies (D) and anti-Flag antibodies (E). In both experiments, lentivirus expressing Flag-RARα-IRES-GFP-Cre in RARαfl/fl background was used to replace endogenous RARα with Flag-tagged RARα. Hippocampal cultures obtained from RARαffl/fl/Fmr1 KO mice (KO) were used as negative controls for specificity of pull-down.
Ijms 22 06579 g001
Figure 2. Determination of RARα domains important for the interaction with FMRP. (A) Schematics of RARα domain organizations and truncation mutants used for interaction assays. All fusion proteins contain an N-terminal GST tag. (B) GST-tagged purified RARα truncation mutants from (A) were incubated with HEK cell lysates; pull-down of endogenous FMRP in P30 mouse hippocampal lysates was detected by immunoblot (top panel). Coomassie blue staining shows expression levels of truncation mutants (bottom panel). (C) Immunoprecipitation of endogenous FMRP (top panel) in HEK cells transfected with GFP, GFP-RARα, and GFP-RARαΔtC/D (bottom panel) using anti-GFP antibodies.
Figure 2. Determination of RARα domains important for the interaction with FMRP. (A) Schematics of RARα domain organizations and truncation mutants used for interaction assays. All fusion proteins contain an N-terminal GST tag. (B) GST-tagged purified RARα truncation mutants from (A) were incubated with HEK cell lysates; pull-down of endogenous FMRP in P30 mouse hippocampal lysates was detected by immunoblot (top panel). Coomassie blue staining shows expression levels of truncation mutants (bottom panel). (C) Immunoprecipitation of endogenous FMRP (top panel) in HEK cells transfected with GFP, GFP-RARα, and GFP-RARαΔtC/D (bottom panel) using anti-GFP antibodies.
Ijms 22 06579 g002
Figure 3. FMRP domains critical for the interaction with RARα. (A) Schematics of FMRP domain organization and truncation/point mutations used to determine the RARα binding site. All proteins have a C-terminal GFP tag. Numbering represents the position of amino acid in FMRP. (B) Pull-down of FMRP mutants from (A) expressed in Fmr1 KO MEFs with GST or GST-RARα.
Figure 3. FMRP domains critical for the interaction with RARα. (A) Schematics of FMRP domain organization and truncation/point mutations used to determine the RARα binding site. All proteins have a C-terminal GFP tag. Numbering represents the position of amino acid in FMRP. (B) Pull-down of FMRP mutants from (A) expressed in Fmr1 KO MEFs with GST or GST-RARα.
Ijms 22 06579 g003
Figure 4. Retinoic acid enhances the interaction between FMRP and RARα. (A) Immunoprecipitation with anti-GFP antibodies of FMRP with GFP-RARα in RA treated (1 μM for 1 h) HEK cells. (B) Quantification of the amount of FMRP immunoprecipitated by GFP-RARα after treatment with RA (Student’s t-test, n = 3; * p < 0.05). The graph represents average values ± SEM.
Figure 4. Retinoic acid enhances the interaction between FMRP and RARα. (A) Immunoprecipitation with anti-GFP antibodies of FMRP with GFP-RARα in RA treated (1 μM for 1 h) HEK cells. (B) Quantification of the amount of FMRP immunoprecipitated by GFP-RARα after treatment with RA (Student’s t-test, n = 3; * p < 0.05). The graph represents average values ± SEM.
Ijms 22 06579 g004
Figure 5. Blocking interaction between FMRP and RARα abolishes RA-induced synaptic plasticity. (A) A schematic drawing showing the locations of the blocking peptide and the control peptide mapped to RARα. (B) Pull-down of T7-FMRP with GST-RARα in the presence of purified control or blocking peptide. (C) Representative traces of miniature EPSCs recorded from CA1 pyramidal neurons of hippocampal slice cultures infected with lentivirus expressing either the GFP-tagged control or blocking peptide and treated with RA. Scale bars, 10 pA, 1 s. (D) Quantification of miniature EPSC amplitude and frequency after a 4 h treatment with DMSO or 10 μM RA (** p < 0.01; *** p < 0.001; Mann-Whitney). All graphs represent average values ± SEM.
Figure 5. Blocking interaction between FMRP and RARα abolishes RA-induced synaptic plasticity. (A) A schematic drawing showing the locations of the blocking peptide and the control peptide mapped to RARα. (B) Pull-down of T7-FMRP with GST-RARα in the presence of purified control or blocking peptide. (C) Representative traces of miniature EPSCs recorded from CA1 pyramidal neurons of hippocampal slice cultures infected with lentivirus expressing either the GFP-tagged control or blocking peptide and treated with RA. Scale bars, 10 pA, 1 s. (D) Quantification of miniature EPSC amplitude and frequency after a 4 h treatment with DMSO or 10 μM RA (** p < 0.01; *** p < 0.001; Mann-Whitney). All graphs represent average values ± SEM.
Ijms 22 06579 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Park, E.; Lau, A.G.; Arendt, K.L.; Chen, L. FMRP Interacts with RARα in Synaptic Retinoic Acid Signaling and Homeostatic Synaptic Plasticity. Int. J. Mol. Sci. 2021, 22, 6579. https://doi.org/10.3390/ijms22126579

AMA Style

Park E, Lau AG, Arendt KL, Chen L. FMRP Interacts with RARα in Synaptic Retinoic Acid Signaling and Homeostatic Synaptic Plasticity. International Journal of Molecular Sciences. 2021; 22(12):6579. https://doi.org/10.3390/ijms22126579

Chicago/Turabian Style

Park, Esther, Anthony G. Lau, Kristin L. Arendt, and Lu Chen. 2021. "FMRP Interacts with RARα in Synaptic Retinoic Acid Signaling and Homeostatic Synaptic Plasticity" International Journal of Molecular Sciences 22, no. 12: 6579. https://doi.org/10.3390/ijms22126579

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop