Next Article in Journal
Hybrid Imaging of Aspergillus fumigatus Pulmonary Infection with Fluorescent, 68Ga-Labelled Siderophores
Next Article in Special Issue
Brain Cytoplasmic RNAs in Neurons: From Biosynthesis to Function
Previous Article in Journal
The Impact of Phytases on the Release of Bioactive Inositols, the Profile of Inositol Phosphates, and the Release of Selected Minerals in the Technology of Buckwheat Beer Production
Previous Article in Special Issue
Translation from the Ribosome to the Clinic: Implication in Neurological Disorders and New Perspectives from Recent Advances
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cataloguing and Selection of mRNAs Localized to Dendrites in Neurons and Regulated by RNA-Binding Proteins in RNA Granules

1
Life Science Research Center, University of Toyama, Toyama 930-0194, Japan
2
Laboratory of Neuronal Cell Biology, National Institute for Basic Biology, Okazaki, Aichi 444-8585, Japan
3
Department of Basic Biology, SOKENDAI, Okazaki, Aichi 444-8585, Japan
4
Exploratory Research Center on Life and Living Systems, Okazaki, Aichi 444-8585, Japan
*
Authors to whom correspondence should be addressed.
Submission received: 16 December 2019 / Revised: 18 January 2020 / Accepted: 20 January 2020 / Published: 22 January 2020
(This article belongs to the Special Issue RNA Trafficking and Local Translation in Neuronal Health and Disease)

Abstract

:
Spatiotemporal translational regulation plays a key role in determining cell fate and function. Specifically, in neurons, local translation in dendrites is essential for synaptic plasticity and long-term memory formation. To achieve local translation, RNA-binding proteins in RNA granules regulate target mRNA stability, localization, and translation. To date, mRNAs localized to dendrites have been identified by comprehensive analyses. In addition, mRNAs associated with and regulated by RNA-binding proteins have been identified using various methods in many studies. However, the results obtained from these numerous studies have not been compiled together. In this review, we have catalogued mRNAs that are localized to dendrites and are associated with and regulated by the RNA-binding proteins fragile X mental retardation protein (FMRP), RNA granule protein 105 (RNG105, also known as Caprin1), Ras-GAP SH3 domain binding protein (G3BP), cytoplasmic polyadenylation element binding protein 1 (CPEB1), and staufen double-stranded RNA binding proteins 1 and 2 (Stau1 and Stau2) in RNA granules. This review provides comprehensive information on dendritic mRNAs, the neuronal functions of mRNA-encoded proteins, the association of dendritic mRNAs with RNA-binding proteins in RNA granules, and the effects of RNA-binding proteins on mRNA regulation. These findings provide insights into the mechanistic basis of protein-synthesis-dependent synaptic plasticity and memory formation and contribute to future efforts to understand the physiological implications of local regulation of dendritic mRNAs in neurons.

1. Introduction

Spatiotemporal translational regulation is key to increasing the concentration of specific proteins to exert their functions at specific timings and locations in cells. In neurons, mRNAs are translated not only in the cell soma but also in axons and dendrites [1,2]. Local translation in axons is mainly required for axon outgrowth and maintenance, and local translation in dendrites is necessary for synaptic plasticity and long-term memory formation [1,2,3,4].
In 1996, Kang and Schuman revealed that long-term potentiation (LTP) occurs in the stratum radiatum (SR), which is a dendrite-enriched region in hippocampal CA1, even after the SR was isolated from the cell-soma-enriched stratum pyramidale (SP). LTP in the isolated SR was lost after inhibition of protein synthesis [5]. Their findings suggest the existence of “local translation” in neurites, and that proteins translated locally in neurites are sufficient for LTP induction. Since these findings, our understanding of local translation in dendrites has deepened substantially: specific mRNAs are recruited to “RNA granules”, which are membrane-less RNA–protein complexes containing mRNAs, RNA-binding proteins, ribosomes, and translational regulators, and are transported to dendrites using microtubules as rails [6,7,8]. RNA-binding proteins in RNA granules are key regulators of mRNA localization and protein synthesis. Dysfunction of these proteins causes abnormalities in dendritic mRNA localization and translation, resulting in impairment of higher-order brain functions, such as neurodevelopmental disorders, intellectual and mental disorders, and loss of long-term memory formation [9,10].
This review catalogues and selects mRNAs that are localized to dendrites and regulated by RNA-binding proteins in RNA granules. Here, we focus on dendritic mRNAs that have been identified in common in several studies that identified dendritic mRNAs based on various criteria using comprehensive RNA sequencing (RNA-seq) of the SR from rodent hippocampal CA1. These mRNAs are further classified into groups according to the functions of the mRNA-encoded proteins. In addition, we focus on the major RNA-binding proteins in RNA granules and list their target mRNAs and effects on mRNAs. Furthermore, we compared our selected set of dendritic mRNAs with the RNA-binding protein target mRNAs and discuss the contribution of RNA-binding proteins to mRNA regulation, such as mRNA expression, stability, localization, and translation, for the regulation of local translation in dendrites.

2. mRNAs That Are Localized in the Dendrite-Enriched Layer of the Hippocampus

In the 1990s, several mRNAs were identified as dendritically localized mRNAs, among which Camk2a, Map2, Arc, and Insp3r1 mRNAs were intensively studied [11,12]. Later, with the help of advances in deep sequencing technology, thousands of mRNAs that are localized to dendrites have been identified from the dendrite-enriched layer (SR) in the rodent hippocampus [13,14,15,16,17]. We briefly review studies that comprehensively identified the candidates of dendritic mRNAs from the SR of hippocampal CA1 using RNA-seq (Figure 1) and focus on selective mRNAs identified in common among these studies.

2.1. mRNAs That Are Identified in SR Isolated from Rodent Hippocampal CA1

Cajigas et al. identified mRNAs that are abundant in the synaptic neuropil in the rat hippocampus (Figure 1) [15]. They microdissected the SR and stratum lacunosum moleculare (SLM) as the neuropil layer from the hippocampal CA1 region of adult rat brains and subjected them to RNA-seq. From the raw dataset of RNA-seq, they subtracted mRNAs enriched in various types of cells other than pyramidal neurons, such as glial cells, interneurons, blood vessels, nuclei, and mitochondria, and identified 2550 transcripts as dendritic and/or axonal mRNAs. Subsequent mRNA labeling with high-resolution in situ hybridization revealed that all 71 mRNAs for which the authors developed probes were detected in dendrites, suggesting that most of the mRNAs identified from the SR and SLM layers using this method are dendritic mRNAs.
Nakayama et al. identified mRNAs more enriched in the SR layer compared with the SP layer in the hippocampus of adult mice (Figure 1) [16]. They microdissected and isolated the SP and SR layers from mouse hippocampal CA1 and subjected them to RNA-seq. From the RNA-seq raw data, as in the study of Cajigas et al., mRNAs enriched in various cell types other than pyramidal neurons were subtracted. Relative read counts of transcripts in the SR compared with the SP were calculated, which identified 2106 SP-enriched mRNAs and 1122 SR-enriched mRNAs as somatically and dendritically enriched mRNAs, respectively. The authors also revealed that the enrichment of SR-enriched mRNAs in the SR layer was reduced by conditional knockout (cKO) of RNA granule protein 105 (RNG105, also known as Caprin1), deficiency of which severely impaired long-term memory formation.
Ainsley et al. identified ribosome-bound mRNAs in the hippocampus of mice that had received a novel experience causing activity-induced memory formation (Figure 1) [17]. They collected tissues from the SP and SR layers of the hippocampal CA1 of adult mice expressing EGFP-tagged ribosomal protein L10a (Rpl10a) specifically in pyramidal neurons but not in glia or interneurons. SP- and SR-enriched mRNAs bound to EGFP-tagged ribosomes were obtained by immunoprecipitation and then sequenced using RNA-seq. The RNA-seq data from resting mice and mice exposed to a novel experience consisting of a contextual-fear-conditioning trial were analyzed by machine learning classification, which listed 1860 ribosome-bound dendritically localized mRNAs after fear conditioning. Because mRNA binding to ribosomes increases during translation, these ribosome-bound mRNAs were thought to be translated locally in dendrites in an activity-dependent manner.
These three studies used different strategies to identify dendritic mRNAs in term of abundance, relative enrichment, and ribosome binding. In particular, the strategy used by Ainsley et al. differed significantly from the other two studies in that they used transgenic mice overexpressing Rpl10a and analyzed selective mRNAs that bound to ribosomes. Due to this difference, it may be supposed that the overlap of dendritic mRNAs between Ainsley et al. and the other two studies may be smaller than that between the other two. However, this was not the case (Figure 2a), suggesting that dendritic translation may not occur in a highly mRNA-selective manner. We found that the number of overlapped dendritic mRNAs among the three studies was 78, representing only 3%–7% of the total number of dendritic mRNAs identified in each study (Figure 2a, Table 1, and Supplementary Table S1). Therefore, these 78 mRNAs are considered fairly reliable candidates for mRNAs localized to dendrites and locally translated in an activity-dependent manner. Recently, Farris et al. also identified mRNAs localized in the hippocampal subregions (CA1, CA2, CA3, and the dentate gyrus) of the adult mouse hippocampus [18]. They found that 68 of the 78 mRNAs overlapped as mRNAs enriched in the SR layer of the CA1 region among their study and the three studies above [18], supporting the reliable dendritic localization of the commonly identified mRNAs.
The general functions and neuronal/brain functions of these 78 mRNA-encoded proteins are summarized in Supplementary Table S1 [19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194]. Of note, most of the 78 mRNAs encoded proteins involved in the regulation of synaptic functions, such as α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptor trafficking and transmission, N-methyl-d-aspartate (NMDA) receptor activity, ion channel activity, spine growth and maintenance, synaptic plasticity, LTP, and learning and memory. In addition, their dysfunctions were associated with a variety of neuropsychiatric disorders such as fragile X syndrome, Alzheimer’s disease, Down syndrome, schizophrenia, Parkinson’s disease, and autism spectrum disorder (ASD) (Supplementary Table S1). In order to understand the major biological and functional categories in which the proteins encoded by these mRNAs are involved, they were classified by gene ontology (GO) enrichment analysis using DAVID 6.8. As a result, these mRNAs were enriched mainly in the categories of “ribosome”, “synapse”, “elongation factor”, “dendritic spine”, and “ionotropic glutamate receptor complex” (Figure 2b,c).

2.2. Unexpected GO Categories of the Catalogued Dendritic mRNAs: “Ribosome” and “Elongation Factor”

The fact that many mRNAs encoding ribosomal proteins were found in the dendritic mRNAs was surprising. It is well known that ribosome subunits are assembled in the nucleus, where ribosome assembly requires many steps and with the help of various regulatory factors [195]. The resulting ribosomes are recruited to RNA granules and transported to dendrites [1,7,196]. Thus, although the possibility of ribosome biogenesis in dendrites cannot be ruled out, it seems almost impossible. Another possible involvement of local translation of ribosomal proteins is ribosome heterogeneity in neurons. It has been suggested that the structure of ribosomes is often not uniform in different tissues, in cells at different developmental stages, and even in distinct subcellular locations within the same cell. Such ribosomes recognize specific mRNAs through differences in components and chemical modifications of ribosomal subunit proteins, which is known as the “ribosome filter hypothesis” [197]. For example, the components of ribosomal subunits are different between testis and liver [198], ribosomes in mouse embryonic stem cells are heterogeneous and translate distinct subpools of mRNAs [199], and changes in the composition of ribosomal proteins regulate the translation of specific mRNAs in neocortical development [200,201]. In yeast, translation of ASH mRNA, localized at the tip of daughter cells, requires ribosomes composed of specific ribosomal protein paralogs [202]. Given that mRNA translation differs between soma and dendrites, local translation in dendrites of mRNAs encoding ribosomal proteins could cause subcellular ribosome heterogeneity in neurons. This mechanism may require the exchange of ribosomal subunit proteins in dendrites. Ribosomal protein exchange has been reported in Escherichia coli in several studies [203,204,205], and notably, damaged ribosomes are repaired by exchanging ribosomal subunit proteins [206]. Thus, the exchange of ribosomal subunit proteins could occur in neurons, which may contribute not only to ribosome heterogeneity but also to the maintenance of functional ribosomes in dendrites. Another possibility is that ribosomal subunit proteins have their own functions by themselves. The physiological relevance of the localization of ribosomal-protein-encoding mRNAs to dendrites is unknown, but it is very interesting whether the dendritic localization is involved in the local regulation of synaptic plasticity and higher-order brain functions.
The GO category “elongation factor” included Eef1a1, Eef1b2, and Eef2 mRNAs. Many translational regulators are known to be recruited to RNA granules as proteins and transported to dendrites [6,196]. In contrast, the above elongation factors are recruited as mRNAs and may be translated locally in dendrites. It is interesting that, among several steps of translation such as initiation, elongation, and termination, the factors involved in the elongation step were selectively enriched in dendritic mRNAs. Translation initiation steps are mainly regulated by phosphorylation and dephosphorylation of the initiation factors. For example, phosphorylation of eIF4E promotes the initiation step, whereas phosphorylation of eIF2α in stress conditions inhibits translation initiation. In contrast, in the elongation step, the exchange between GTP and GDP forms of eEF1a and eEF2 is a key regulatory mechanism [207]. In addition, control of the amount of elongation factors may be important for the regulation of elongation. This may be related to the fact that decreased expression levels of Eef1a1 and Eef1b2 are associated with Alzheimer’s disease and intellectual disability, respectively [67,69]. Increasing concentrations of elongation factors at specific timings and locations mediated by local translation could be important for local synaptic regulation.

2.3. Expected GO Categories of the Catalogued Dendritic mRNAs: “Synapse”, “Dendritic Spine”, and “Ionotropic Glutamate Receptor Complex”

The other GO categories “synapse”, “dendritic spine”, and “ionotropic glutamate receptor complex” included mRNAs encoding postsynaptic proteins such as Arc, Kcnab2, Crtc1, Ncs1, Camk2a, Cdk16, Shank1, Homer2, Dlg2, Dlg4, Dlgap3, Sipa1l1, and Psd3 (Figure 2b,c). Most of these are known to play key roles in synaptic plasticity and higher-order brain functions (Supplementary Table S1). Arc accumulates at inactive synapses to prevent their enhancement and is involved in memory formation and neuronal diseases such as fragile X syndrome and Alzheimer’s disease [21,22,23,24,25,26]. Kcnab2, a potassium voltage-gated channel subunit, interacts with Shank3, and mice deficient in Kcnab2 exhibit amygdala hyperexcitability and impaired associative learning and memory [92,93]. Crtc1 is a CREB-regulated transcription factor that is transported from the synapse to the nucleus in response to late-phase LTP (L-LTP), where it enhances spatial memory and memory consolidation/reconsolidation, and its deficiency in mice showed depression-like behaviors and impaired memory formation [44,45,46,47,48,49]. Ncs1 is a calcium-binding protein involved in short- and long-term synaptic plasticity, dopaminergic signaling, learning, and memory, and its dysfunction leads to various neuronal diseases such as schizophrenia, bipolar disorder, Parkinson’s disease, and fragile X syndrome [97,98,99]. Camk2a is a calcium/calmodulin-dependent protein kinase subunit that plays key roles in synaptic plasticity, AMPA receptor transmission, LTP, and long-term memory formation, and its dysfunction underlies neuropsychiatric disorders such as schizophrenia, epilepsy, and fragile X syndrome [35,36,37,38,39,40,41,42]. Cdk16, a cyclin-dependent kinase, is responsible for dendrite development [43]. Shank1, Homer2, Dlg2, and Dlg4 are scaffold proteins of excitatory postsynaptic density. Shank1 regulates spine morphology and neurotransmission and is associated with ASD and neuropsychiatric disorders [156,157,158,159,160]. Homer2 interacts with dendritic spine actin regulators such as Cdc42 and drebrin, enhances cell surface expression of NMDA receptors, and is involved in spine maturation. Homer2 also interacts with amyloid precursor protein to inhibit Aβ production [86,87,88,89]. Dlg2 is associated with neurodevelopmental disorders, its deficiency in mice causes LTP impairment and hypersocial behavior, and its overexpression improves Aβ-mediated cognitive dysfunction [56,57,58,59,60]. Dlg4, also known as PSD-95, plays key roles in synaptic plasticity, spine growth, and AMPA and NMDA receptor regulation, and its dysfunction underlies neuropsychiatric disorders such as schizophrenia and autism [61,62]. Dlgap3, also known as SAPAP3, binds to Dlg4, regulates mGluR5-driven AMPA receptor trafficking, and its deficiency in mice exhibits obsessive–compulsive-disorder-like behaviors and altered ultrasonic vocalizations [63,64,65,66]. Sipa1l1 is a Rap GTPase-activating protein (GAP) involved in spine morphogenesis, homeostatic synaptic plasticity, learning, and memory, and its expression is altered by Aβ treatment [161,162,163,164,165,166,167,168,169,170,171,172,173]. Finally, Psd3 is a guanine nucleotide exchange factor of ADP-ribosylation factor 6 (Arf6), but its neuronal function is unknown. Thus, these dendritic mRNA-encoded proteins are implicated in synaptic plasticity and learning and memory, well-known roles of local translation in dendrites.
Of note, Homer2 and Psd3 were also classified as proteins that contain membrane-bound pleckstrin-homology (PH) domain in the GO enrichment analysis, although the Benjamini value (0.0535) did not reach significance. The GO categories of “PH domain” and/or “PH-like domain” included Homer2, Psd3, Psd, Plekhm2, Sptbn2, and Arhgap25. Compared with soluble proteins that diffuse easily in the cytosol, membrane-associated proteins are difficult to diffuse freely due to physical obstacles and electrostatic interactions within the membrane [208]. PH domains bind to phosphoinositides, which provide such electrostatic interactions [208,209]. Therefore, local translation of membrane-associated proteins appears to be an effective way to deliver low-mobility proteins at the required time and place.

3. mRNAs Bound to and Regulated by RNA-Binding Proteins of RNA Granules

RNA-binding proteins in RNA granules play a variety of roles in regulating mRNA stability, localization, and translation in neurons. Among them, we focus on five RNA-binding proteins: fragile X mental retardation protein (FMRP), RNG105/Caprin1, Ras-GAP SH3 domain binding protein (G3BP), cytoplasmic polyadenylation element binding protein 1 (CPEB1), and staufen double-stranded RNA binding proteins 1 and 2 (Stau1 and Stau2), of which associated mRNAs have been relatively extensively identified, and as summarized below, their knockout (KO)/knockdown (KD) in rodents are known to affect neural functions and higher-order brain functions, including memory formation.

3.1. Effects of RNA-Binding Proteins of RNA Granules on Higher-Order Brain Functions

FMRP is encoded by the Fmr1 gene, in which mutations cause fragile X syndrome, a mental retardation disorder. FMRP regulates protein-synthesis-dependent synaptic plasticity, because Fmr1 KO mice exhibited enhanced metabotropic glutamate receptor (mGluR)-dependent long-term depression (LTD) that requires protein synthesis [210,211,212]. Although controversial, several studies reported the impact of FMRP on memory formation. Contextual fear memory in Fmr1 KO mice was both impaired [213,214,215] and unaffected [216]. Spatial memory in the Morris water maze (MWM) in Fmr1 KO mice was also both impaired [217,218] and unaffected [213,219,220].
RNG105/Caprin1 is responsible for the formation and/or maintenance of dendrites and synaptic connections [221] and is essential for long-term memory formation. Even a slight deficiency of RNG105/Caprin1 in heterozygous mice impaired reversal learning in spatial learning tasks [222]. A more severe deficiency of RNG105/Caprin1 in forebrain-specific cKO mice significantly impaired long-term memory in both spatial MWM and contextual-fear-conditioning tasks [16].
G3BP plays a central role in the assembly of stress granules (SGs) in response to various types of stress [223,224]. G3BP1 KO mice showed impaired spatial working memory in the Y-maze test, but they had normal acquisition of nonspatial long-term memory in the passive avoidance test [225].
CPEB1 binds to cytoplasmic polyadenylation elements (CPEs) within the 3’UTR of mRNAs and mediates cytoplasmic polyadenylation of mRNAs to promote translation [226]. CPEB1 regulates synaptic plasticity as indicated by a deficit in LTP in CPEB1 KO mice [227]. CPEB1 KO mice showed normal acquisition and retention of MWM spatial memory and fear-conditioned contextual memory but reduced memory extinction in these tests [228].
Staufen is well known for its role in mRNA localization during Drosophila embryo development, whereas its orthologs are known to regulate synaptic plasticity in rodents: Stau1 and Stau2 are required for L-LTP and LTD, respectively [229,230]. Although Stau1 KO mice showed no abnormalities in memory formation [231], Stau2 deficiency affected several types of memory. Stau2 KD rats exhibited impairment in spatial working memory in a delayed matching to place task in the MWM and delayed nonmatching to place task in an eight-arm radial maze [232]. In addition, Stau2 KD rats showed deficits in temporal association memory in a trace fear conditioning task and spatial association memory in an inhibitory avoidance task [232]. Downregulation of Stau2 in mice displayed deficits in discriminating different spatial contexts in the Barnes maze task [233].

3.2. Methods for Identifying RNA–Protein Interactions

To better understand the mechanistic basis of the RNA-binding-protein-mediated regulation of biological functions, including brain functions, it is important to identify target mRNAs for RNA-binding proteins. Various methods have been developed for this purpose [234,235].
In one method, RNA–protein direct binding in vitro has been analyzed using systematic evolution of ligands by exponential enrichment (SELEX). In SELEX, binding cycles between RNA-binding proteins and RNA pools are repeated to select high-affinity RNAs, and these RNAs are amplified by reverse transcription PCR, which detects RNAs that bind to proteins with high probability [236,237,238].
Second, RNA–protein association in cells and tissues has been identified with RNA-immunoprecipitation (IP), UV-crosslinking IP (CLIP), and electrophoretic mobility shift assay (EMSA). In CLIP, UV irradiation of cells or tissues generates covalent bonds between RNA and RNA-binding proteins when they are in close proximity to each other, followed by co-IP of the RNA with the RNA-binding proteins [239,240,241]. Other modified CLIP methods have been developed, such as high-throughput sequencing CLIP (HITS-CLIP) and photoactivatable ribonucleoside-enhanced CLIP (PAR-CLIP). HITS-CLIP is a combination of CLIP and high-throughput sequencing that enables comprehensive identification of associated RNAs [240,241]. In PAR-CLIP, RNA is labeled with photoreactive nucleoside analogs, and cells are irradiated with 365 nm UV. As a result, crosslinking efficiency is improved compared with conventional CLIP [240]. In EMSA, labeled RNA pools are incubated with cell and tissue extracts, and the binding of RNA to proteins can be detected by reduced mobility of the RNA in gel electrophoresis [242,243].
Third, RNA-protein association has also been identified using gene overexpression, KD, and KO of RNA-binding proteins in cells and animals. These altered expression levels affect the stability, localization, and translation efficiency of the target mRNAs depending on the effect of the RNA-binding proteins on the mRNAs. Differential analysis of these changes compared with control cells and animals, using microarrays, RNA-seq, ribosome footprints, and their combination with the isolation of specific regions of cells, can identify mRNAs regulated by RNA-binding proteins. There are also various other methods to analyze RNA-protein associations, such as sequence- and structure-based methods using computational approaches [235].
Using these methods, target mRNAs of RNA-binding proteins in RNA granules have been identified. We comprehensively list the target mRNAs of FMRP, RNG105/Caprin1, G3BP, CPEB1, Stau1, and Stau2 and summarize the effects of the RNA-binding proteins on the mRNAs (Table 1 and Supplementary Table S2a–e). Furthermore, we compared the list of the dendritic mRNAs (Supplementary Table S1) with the target mRNAs of the RNA-binding proteins (Supplementary Table S2a–e) to find dendritic mRNAs that are associated with and/or regulated by FMRP, RNG105/Caprin1, G3BP, CPEB1, Stau1, and Stau2 (Supplementary Table S1). In the following subsection, we discuss how the RNA-binding proteins regulate their target mRNAs during mRNA localization and local translation in dendrites.

3.3. Effects of the RNA-Binding Proteins on mRNA Regulation

FMRP target mRNAs have been identified in many studies and are listed in Supplementary Table S2a [244,245,246,247,248,249,250,251,252,253,254,255,256,257,258,259,260,261,262,263,264,265,266,267,268,269,270,271,272,273,274,275,276,277,278,279,280,281,282,283,284,285,286,287,288,289,290,291,292,293,294,295,296,297,298,299,300,301,302,303,304]. FMRP has various effects on mRNA regulation such as mRNA expression, stability, localization, and translation, among which the major role of FMRP is translational regulation, as shown in a number of studies (Supplementary Table S2a). FMRP represses translation through multiple mechanisms in the basal state, but the repression is relieved by dephosphorylation of FMRP upon mGluR stimulation (references in Supplementary Table S2a). CPEB1 also regulates translation of its target mRNAs. The CPEB1 target mRNAs are listed in Supplementary Table S2d [305,306,307,308,309,310,311,312,313,314,315,316,317,318,319,320,321,322,323,324,325,326,327,328,329,330,331,332,333,334]. CPEB1 represses translation in the basal state but is phosphorylated upon synaptic stimulation, thereby polyadenylating its target mRNAs and promoting translation (references in Supplementary Table S2d). In neurons, both FMRP and CPEB1 are localized in dendrites and in/near spines [335,336,337,338]. Given that many of the 78 dendritic mRNAs have been identified to be associated with FMRP and CPEB1 (i.e., 28 with FMRP and 11 with CPEB1 (Supplementary Table S1)), FMRP and CPEB1 may regulate activity-dependent translation of dendritic mRNAs in dendrites near spines.
In contrast, RNG105/Caprin1, for which target mRNAs are listed in Supplementary Table S2b [16,221,339,340,341,342], is involved in mRNA localization. RNG105/Caprin1-containing RNA granules are localized throughout the proximal to distal regions of dendrites [221,339]. RNG105 cKO reduced the dendritic localization of many (46 of 78) of the dendritic mRNAs (Supplementary Table S1) [16]. This suggests that RNG105 plays an important role in mRNA localization to proximal and distal dendrites to achieve local translation, which may be necessary for the formation of long-term memory. In addition to mRNA localization, RNG105/Caprin1 may play a role in spatially biasing translation to RNA granules: overexpression of RNG105/Caprin1 in cells locally increased translation in/near RNA granules but suppressed translation in the cytoplasm [343]. Staufen also regulates mRNA localization in addition to participating in mRNA decay, known as Staufen-mediated mRNA decay. The target mRNAs of Staufen are listed in Supplementary Table S2e [230,312,344,345,346,347,348,349,350,351,352,353,354,355,356,357,358,359,360]. Of the 78 dendritic mRNAs, 44 were associated with Stau1 and/or Stau2 (Supplementary Table S1). Although both Stau1 and Stau2 show dendritic localization in neurons, Stau2 is localized to more distal dendrites than Stau1 [361]. It was reported that some dendritic mRNAs are restricted to proximal dendrites, but others are localized not only to proximal but also distal dendrites [15]. Given the different distributions along the proximal–distal dendritic axis of RNG105/Caprin1, Stau1, and Stau2, these RNA-binding proteins may regulate mRNA localization to the appropriate region of dendrites.
The association of G3BP with mRNA has been analyzed in several studies, including comprehensive identifications of G3BP-associated mRNAs (Supplementary Table S2c) [362,363,364,365,366,367,368,369,370,371,372]. However, no G3BP-associated mRNAs have been found in the 78 dendritic mRNAs (Supplementary Table S1). The dendritic mRNAs were identified in the hippocampus [15,16,17], but fibroblasts were used to comprehensively identify G3BP-associated mRNAs [370,371]. This difference in mRNA sources may be a reason for the absence of G3BP-associated mRNAs in dendritic mRNAs. Alternative reasons may come from the G3BP distribution pattern in neurons: although G3BP is localized to dendrites in addition to the soma in primary cultured neurons [221], G3BP is almost restricted to the soma in hippocampal CA1 pyramidal neurons in vivo [225]. Thus, G3BP may preferentially target somatic mRNAs rather than dendritic mRNAs in hippocampal pyramidal neurons. It should also be noted that G3BP is a key protein that facilitates the assembly of SGs, of which many components are shared with neuronal RNA granules. SGs are formed only under stress, but neuronal RNA granules are always present, even when neurons are unstressed. Arsenite treatment of neurons that already have RNA granules has been reported to induce the assembly of another population of granules containing RNA-binding proteins such as G3BP, TIAR, and FMRP [129,373,374]. Thus, stress-induced granules may differ from granules present in unstressed neurons. In this case, G3BP may primarily target mRNAs in SGs rather than regulating dendritic mRNAs for neuronal synaptic plasticity.
Each RNA-binding protein has its own functions that regulate the expression, stability, localization, and translation of its target mRNAs. In addition to executing these intrinsic functions, RNA-binding proteins in RNA granules may also play roles in regulating mRNAs through controlling the condensation and dissolution of RNA granules. Recently, it has been revealed that RNA granules are formed by liquid–liquid phase separation and their liquid-like and solid-like states can be dynamically controlled [375,376,377,378,379,380,381]. Each RNA-binding protein exhibits different dynamics in RNA granules, and their combinations in RNA granules affect the physical properties of RNA granules and the translation efficiency in/near RNA granules [343]. Understanding the effects of physical properties and condensation/dissolution behavior of RNA-binding proteins on mRNA regulation may be also important to better understand the mechanistic basis of mRNA localization and local translation in neurons.

4. Future Perspectives

The functions of RNA-binding proteins have been revealed in many studies. In addition, their target and dendritic mRNAs have been identified, and the functions of the proteins encoded by these mRNAs have been elucidated. However, there is limited understanding of the physiological relevance of locally increasing the concentration of mRNA-encoded proteins through mRNA localization and local translation. Exceptionally, Camk2a mRNA localization to dendrites has been reported to be required for L-LTP and long-term memory formation. This was revealed by the deletion of the 3’UTR, which caused a loss of dendritic localization of Camk2a mRNA but did not affect its translation [382]. However, the relevance of such local regulation of most other dendritic mRNAs to synaptic plasticity and brain functions remains elusive. To determine this, high-throughput methods cannot be used, and each mRNA has to be analyzed individually. Nevertheless, recent advances in comprehensive identification technology have identified a large number of candidates for dendritic mRNAs and RNA-binding protein target mRNAs, and selective ones need to be prioritized. In this review, we have catalogued and selected mRNAs that are localized to dendrites and regulated by RNA-binding proteins in RNA granules. A future challenge in this research field is to elucidate the biological and physiological relevance of local regulation of gene expression, which can be solved by artificially manipulating local regulation, such as altering mRNA localization without affecting translation efficiency and affecting RNA granule assembly without losing essential functions of RNA-binding proteins. We hope this review will help researchers in this field find gemstones in the wilderness.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2218-273X/10/2/167/s1, Table S1: Dendritic mRNAs and their associated RNA-binding proteins. Table S2a: FMRP target mRNAs. Table S2b: RNG105/Caprin1 target mRNAs. Table S2c: G3BP target mRNAs. Table S2d: CPEB1 target mRNAs. Table S2e: Staufen, Stau1 and Stau2 target mRNAs.

Author Contributions

Conceptualization, R.O. and N.S.; Analysis, R.O. and N.S.; Writing-original draft, R.O.; Writing-review and editing, N.S.; Supervision, N.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Grant-in-Aid for Scientific Research from the Japan Society for the Promotion of Science (JSPS) (19H03161 (N.S.)) and Grant-in-Aid for JSPS fellows (19J01826 (R.O.)).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Martin, K.C.; Ephrussi, A. mRNA localization: Gene expression in the spatial dimension. Cell 2009, 136, 719–730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Costa-Mattioli, M.; Sossin, W.S.; Klann, E.; Sonenberg, N. Translational control of long-lasting synaptic plasticity and memory. Neuron 2009, 61, 10–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Sutton, M.A.; Schuman, E.M. Dendritic protein synthesis, synaptic plasticity, and memory. Cell 2006, 127, 49–58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Cioni, J.-M.; Koppers, M.; Holt, C.E. Molecular control of local translation in axon development and maintenance. Curr. Opin. Neurobiol. 2018, 51, 86–94. [Google Scholar] [CrossRef] [PubMed]
  5. Kang, H.; Schuman, E.M. A requirement for local protein synthesis in neurotrophin-induced hippocampal synaptic plasticity. Science 1996, 273, 1402–1406. [Google Scholar] [CrossRef]
  6. Krichevsky, A.M.; Kosik, K.S. Neuronal RNA granules: A link between RNA localization and stimulation-dependent translation. Neuron 2001, 32, 683–696. [Google Scholar] [CrossRef] [Green Version]
  7. Kiebler, M.A.; Bassell, G.J. Neuronal RNA granules: Movers and makers. Neuron 2006, 51, 685–690. [Google Scholar] [CrossRef] [Green Version]
  8. Bramham, C.R.; Wells, D.G. Dendritic mRNA: Transport, translation and function. Nat. Rev. Neurosci. 2007, 8, 776–789. [Google Scholar] [CrossRef]
  9. Wang, W.; van Niekerk, E.; Willis, D.E.; Twiss, J.L. RNA transport and localized protein synthesis in neurological disorders and neural repair. Dev. Neurobiol. 2007, 67, 1166–1182. [Google Scholar] [CrossRef]
  10. Sudhakaran, I.P.; Ramaswami, M. Long-term memory consolidation: The role of RNA-binding proteins with prion-like domains. RNA Biol. 2017, 14, 568–586. [Google Scholar] [CrossRef]
  11. Kuhl, D.; Skehel, P. Dendritic localization of mRNAs. Curr. Opin. Neurobiol. 1998, 8, 600–606. [Google Scholar] [CrossRef]
  12. Steward, O.; Schuman, E.M. Protein synthesis at synaptic sites on dendrites. Annu. Rev. Neurosci. 2001, 24, 299–325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Poon, M.M.; Choi, S.-H.; Jamieson, C.A.M.; Geschwind, D.H.; Martin, K.C. Identification of process-localized mRNAs from cultured rodent hippocampal neurons. J. Neurosci. 2006, 26, 13390–13399. [Google Scholar] [CrossRef] [Green Version]
  14. Zhong, J.; Zhang, T.; Bloch, L.M. Dendritic mRNAs encode diversified functionalities in hippocampal pyramidal neurons. BMC Neurosci. 2006, 7, 17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Cajigas, I.J.; Tushev, G.; Will, T.J.; tom Dieck, S.; Fuerst, N.; Schuman, E.M. The local transcriptome in the synaptic neuropil revealed by deep sequencing and high-resolution imaging. Neuron 2012, 74, 453–466. [Google Scholar] [CrossRef] [Green Version]
  16. Nakayama, K.; Ohashi, R.; Shinoda, Y.; Yamazaki, M.; Abe, M.; Fujikawa, A.; Shigenobu, S.; Futatsugi, A.; Noda, M.; Mikoshiba, K.; et al. RNG105/caprin1, an RNA granule protein for dendritic mRNA localization, is essential for long-term memory formation. Elife 2017, 6. [Google Scholar] [CrossRef] [Green Version]
  17. Ainsley, J.A.; Drane, L.; Jacobs, J.; Kittelberger, K.A.; Reijmers, L.G. Functionally diverse dendritic mRNAs rapidly associate with ribosomes following a novel experience. Nat. Commun. 2014, 5, 4510. [Google Scholar] [CrossRef] [Green Version]
  18. Farris, S.; Ward, J.M.; Carstens, K.E.; Samadi, M.; Wang, Y.; Dudek, S.M. Hippocampal Subregions Express Distinct Dendritic Transcriptomes that Reveal Differences in Mitochondrial Function in CA2. Cell Rep. 2019, 29, 522–539.e6. [Google Scholar] [CrossRef] [Green Version]
  19. Wang, H.; Ferguson, G.D.; Pineda, V.V.; Cundiff, P.E.; Storm, D.R. Overexpression of type-1 adenylyl cyclase in mouse forebrain enhances recognition memory and LTP. Nat. Neurosci. 2004, 7, 635–642. [Google Scholar] [CrossRef]
  20. Garelick, M.G.; Chan, G.C.K.; DiRocco, D.P.; Storm, D.R. Overexpression of type I adenylyl cyclase in the forebrain impairs spatial memory in aged but not young mice. J. Neurosci. 2009, 29, 10835–10842. [Google Scholar] [CrossRef]
  21. Korb, E.; Finkbeiner, S. Arc in synaptic plasticity: From gene to behavior. Trends Neurosci. 2011, 34, 591–598. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Nikolaienko, O.; Patil, S.; Eriksen, M.S.; Bramham, C.R. Arc protein: A flexible hub for synaptic plasticity and cognition. Semin. Cell Dev. Biol. 2018, 77, 33–42. [Google Scholar] [CrossRef] [PubMed]
  23. Newpher, T.M.; Harris, S.; Pringle, J.; Hamilton, C.; Soderling, S. Regulation of spine structural plasticity by Arc/Arg3.1. Semin. Cell Dev. Biol. 2018, 77, 25–32. [Google Scholar] [CrossRef] [PubMed]
  24. Tzingounis, A.V.; Nicoll, R.A. Arc/Arg3.1: Linking gene expression to synaptic plasticity and memory. Neuron 2006, 52, 403–407. [Google Scholar] [CrossRef] [Green Version]
  25. Shepherd, J.D.; Bear, M.F. New views of Arc, a master regulator of synaptic plasticity. Nat. Neurosci. 2011, 14, 279–284. [Google Scholar] [CrossRef]
  26. Okuno, H.; Minatohara, K.; Bito, H. Inverse synaptic tagging: An inactive synapse-specific mechanism to capture activity-induced Arc/arg3.1 and to locally regulate spatial distribution of synaptic weights. Semin. Cell Dev. Biol. 2018, 77, 43–50. [Google Scholar] [CrossRef]
  27. Yoshihara, T.; Sugihara, K.; Kizuka, Y.; Oka, S.; Asano, M. Learning/memory impairment and reduced expression of the HNK-1 carbohydrate in beta4-galactosyltransferase-II-deficient mice. J. Biol. Chem. 2009, 284, 12550–12561. [Google Scholar] [CrossRef] [Green Version]
  28. Ohta, T.; Ohba, T.; Suzuki, T.; Watanabe, H.; Sasano, H.; Murakami, M. Decreased calcium channel currents and facilitated epinephrine release in the Ca2+ channel beta3 subunit-null mice. Biochem. Biophys. Res. Commun. 2010, 394, 464–469. [Google Scholar] [CrossRef]
  29. Belkacemi, A.; Hui, X.; Wardas, B.; Laschke, M.W.; Wissenbach, U.; Menger, M.D.; Lipp, P.; Beck, A.; Flockerzi, V. IP3 Receptor-Dependent Cytoplasmic Ca2+ Signals Are Tightly Controlled by Cavβ3. Cell Rep. 2018, 22, 1339–1349. [Google Scholar] [CrossRef] [Green Version]
  30. Namkung, Y.; Smith, S.M.; Lee, S.B.; Skrypnyk, N.V.; Kim, H.L.; Chin, H.; Scheller, R.H.; Tsien, R.W.; Shin, H.S. Targeted disruption of the Ca2+ channel beta3 subunit reduces N- and L-type Ca2+ channel activity and alters the voltage-dependent activation of P/Q-type Ca2+ channels in neurons. Proc. Natl. Acad. Sci. USA 1998, 95, 12010–12015. [Google Scholar] [CrossRef] [Green Version]
  31. Jeon, D.; Song, I.; Guido, W.; Kim, K.; Kim, E.; Oh, U.; Shin, H.-S. Ablation of Ca2+ channel beta3 subunit leads to enhanced N-methyl-D-aspartate receptor-dependent long term potentiation and improved long term memory. J. Biol. Chem. 2008, 283, 12093–12101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Murakami, M.; Nakagawasai, O.; Yanai, K.; Nunoki, K.; Tan-No, K.; Tadano, T.; Iijima, T. Modified behavioral characteristics following ablation of the voltage-dependent calcium channel beta3 subunit. Brain Res. 2007, 1160, 102–112. [Google Scholar] [CrossRef] [PubMed]
  33. Griessmeier, K.; Cuny, H.; Rötzer, K.; Griesbeck, O.; Harz, H.; Biel, M.; Wahl-Schott, C. Calmodulin is a functional regulator of Cav1.4 L-type Ca2+ channels. J. Biol. Chem. 2009, 284, 29809–29816. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Hisatsune, C.; Umemori, H.; Inoue, T.; Michikawa, T.; Kohda, K.; Mikoshiba, K.; Yamamoto, T. Phosphorylation-dependent regulation of N-methyl-D-aspartate receptors by calmodulin. J. Biol. Chem. 1997, 272, 20805–20810. [Google Scholar] [CrossRef] [Green Version]
  35. Miyamoto, E.; Fukunaga, K. A role of Ca2+/calmodulin-dependent protein kinase II in the induction of long-term potentiation in hippocampal CA1 area. Neurosci. Res. 1996, 24, 117–122. [Google Scholar] [CrossRef]
  36. Zalcman, G.; Federman, N.; Romano, A. CaMKII Isoforms in Learning and Memory: Localization and Function. Front. Mol. Neurosci. 2018, 11, 445. [Google Scholar] [CrossRef]
  37. Colbran, R.J.; Brown, A.M. Calcium/calmodulin-dependent protein kinase II and synaptic plasticity. Curr. Opin. Neurobiol. 2004, 14, 318–327. [Google Scholar] [CrossRef]
  38. Fink, C.C.; Meyer, T. Molecular mechanisms of CaMKII activation in neuronal plasticity. Curr. Opin. Neurobiol. 2002, 12, 293–299. [Google Scholar] [CrossRef]
  39. Ataei, N.; Sabzghabaee, A.M.; Movahedian, A. Calcium/Calmodulin-dependent Protein Kinase II is a Ubiquitous Molecule in Human Long-term Memory Synaptic Plasticity: A Systematic Review. Int. J. Prev. Med. 2015, 6, 88. [Google Scholar]
  40. Lisman, J.; Schulman, H.; Cline, H. The molecular basis of CaMKII function in synaptic and behavioural memory. Nat. Rev. Neurosci. 2002, 3, 175–190. [Google Scholar] [CrossRef]
  41. Hell, J.W. CaMKII: Claiming center stage in postsynaptic function and organization. Neuron 2014, 81, 249–265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Robison, A.J. Emerging role of CaMKII in neuropsychiatric disease. Trends Neurosci. 2014, 37, 653–662. [Google Scholar] [CrossRef] [PubMed]
  43. Fu, W.-Y.; Cheng, K.; Fu, A.K.Y.; Ip, N.Y. Cyclin-dependent kinase 5-dependent phosphorylation of Pctaire1 regulates dendrite development. Neuroscience 2011, 180, 353–359. [Google Scholar] [CrossRef] [PubMed]
  44. Sekeres, M.J.; Mercaldo, V.; Richards, B.; Sargin, D.; Mahadevan, V.; Woodin, M.A.; Frankland, P.W.; Josselyn, S.A. Increasing CRTC1 function in the dentate gyrus during memory formation or reactivation increases memory strength without compromising memory quality. J. Neurosci. 2012, 32, 17857–17868. [Google Scholar] [CrossRef] [Green Version]
  45. Ch’ng, T.H.; Uzgil, B.; Lin, P.; Avliyakulov, N.K.; O’Dell, T.J.; Martin, K.C. Activity-dependent transport of the transcriptional coactivator CRTC1 from synapse to nucleus. Cell 2012, 150, 207–221. [Google Scholar] [CrossRef] [Green Version]
  46. Parra-Damas, A.; Valero, J.; Chen, M.; España, J.; Martín, E.; Ferrer, I.; Rodríguez-Alvarez, J.; Saura, C.A. Crtc1 activates a transcriptional program deregulated at early Alzheimer’s disease-related stages. J. Neurosci. 2014, 34, 5776–5787. [Google Scholar] [CrossRef] [Green Version]
  47. Breuillaud, L.; Rossetti, C.; Meylan, E.M.; Mérinat, C.; Halfon, O.; Magistretti, P.J.; Cardinaux, J.-R. Deletion of CREB-regulated transcription coactivator 1 induces pathological aggression, depression-related behaviors, and neuroplasticity genes dysregulation in mice. Biol. Psychiatry 2012, 72, 528–536. [Google Scholar] [CrossRef]
  48. Nonaka, M.; Kim, R.; Fukushima, H.; Sasaki, K.; Suzuki, K.; Okamura, M.; Ishii, Y.; Kawashima, T.; Kamijo, S.; Takemoto-Kimura, S.; et al. Region-specific activation of CRTC1-CREB signaling mediates long-term fear memory. Neuron 2014, 84, 92–106. [Google Scholar] [CrossRef] [Green Version]
  49. Parra-Damas, A.; Chen, M.; Enriquez-Barreto, L.; Ortega, L.; Acosta, S.; Perna, J.C.; Fullana, M.N.; Aguilera, J.; Rodríguez-Alvarez, J.; Saura, C.A. CRTC1 Function During Memory Encoding Is Disrupted in Neurodegeneration. Biol. Psychiatry 2017, 81, 111–123. [Google Scholar] [CrossRef] [Green Version]
  50. Li, D.-P.; Zhou, J.-J.; Pan, H.-L. Endogenous casein kinase-1 modulates NMDA receptor activity of hypothalamic presympathetic neurons and sympathetic outflow in hypertension. J. Physiol. (Lond.) 2015, 593, 4439–4452. [Google Scholar] [CrossRef] [Green Version]
  51. Chergui, K.; Svenningsson, P.; Greengard, P. Physiological role for casein kinase 1 in glutamatergic synaptic transmission. J. Neurosci. 2005, 25, 6601–6609. [Google Scholar] [CrossRef] [PubMed]
  52. Gross, S.D.; Hoffman, D.P.; Fisette, P.L.; Baas, P.; Anderson, R.A. A phosphatidylinositol 4,5-bisphosphate-sensitive casein kinase I alpha associates with synaptic vesicles and phosphorylates a subset of vesicle proteins. J. Cell Biol. 1995, 130, 711–724. [Google Scholar] [CrossRef] [PubMed]
  53. Yinan, M.; Yu, Q.; Zhiyue, C.; Jianjun, L.; Lie, H.; Liping, Z.; Jianhui, Z.; Fang, S.; Dingfang, B.; Qing, L.; et al. Polymorphisms of casein kinase I gamma 2 gene associated with simple febrile seizures in Chinese Han population. Neurosci. Lett. 2004, 368, 2–6. [Google Scholar] [CrossRef] [PubMed]
  54. Ji, Z.; Li, H.; Yang, Z.; Huang, X.; Ke, X.; Ma, S.; Lin, Z.; Lu, Y.; Zhang, M. Kibra Modulates Learning and Memory via Binding to Dendrin. Cell Rep. 2019, 26, 2064–2077.e7. [Google Scholar] [CrossRef] [Green Version]
  55. Kremerskothen, J.; Kindler, S.; Finger, I.; Veltel, S.; Barnekow, A. Postsynaptic recruitment of Dendrin depends on both dendritic mRNA transport and synaptic anchoring. J. Neurochem. 2006, 96, 1659–1666. [Google Scholar] [CrossRef]
  56. Reggiani, C.; Coppens, S.; Sekhara, T.; Dimov, I.; Pichon, B.; Lufin, N.; Addor, M.-C.; Belligni, E.F.; Digilio, M.C.; Faletra, F.; et al. Novel promoters and coding first exons in DLG2 linked to developmental disorders and intellectual disability. Genome Med. 2017, 9, 67. [Google Scholar] [CrossRef]
  57. Winkler, D.; Daher, F.; Wüstefeld, L.; Hammerschmidt, K.; Poggi, G.; Seelbach, A.; Krueger-Burg, D.; Vafadari, B.; Ronnenberg, A.; Liu, Y.; et al. Hypersocial behavior and biological redundancy in mice with reduced expression of PSD95 or PSD93. Behav. Brain Res. 2018, 352, 35–45. [Google Scholar] [CrossRef]
  58. Parker, M.J.; Zhao, S.; Bredt, D.S.; Sanes, J.R.; Feng, G. PSD93 regulates synaptic stability at neuronal cholinergic synapses. J. Neurosci. 2004, 24, 378–388. [Google Scholar] [CrossRef] [Green Version]
  59. Yu, L.; Liu, Y.; Yang, H.; Zhu, X.; Cao, X.; Gao, J.; Zhao, H.; Xu, Y. PSD-93 Attenuates Amyloid-β-Mediated Cognitive Dysfunction by Promoting the Catabolism of Amyloid-β. J. Alzheimers Dis. 2017, 59, 913–927. [Google Scholar] [CrossRef] [Green Version]
  60. Carlisle, H.J.; Fink, A.E.; Grant, S.G.N.; O’Dell, T.J. Opposing effects of PSD-93 and PSD-95 on long-term potentiation and spike timing-dependent plasticity. J. Physiol. (Lond.) 2008, 586, 5885–5900. [Google Scholar] [CrossRef]
  61. Iasevoli, F.; Tomasetti, C.; de Bartolomeis, A. Scaffolding proteins of the post-synaptic density contribute to synaptic plasticity by regulating receptor localization and distribution: Relevance for neuropsychiatric diseases. Neurochem. Res. 2013, 38, 1–22. [Google Scholar] [CrossRef] [PubMed]
  62. Coley, A.A.; Gao, W.-J. PSD95: A synaptic protein implicated in schizophrenia or autism? Prog. Neuropsychopharmacol. Biol. Psychiatry 2018, 82, 187–194. [Google Scholar] [CrossRef] [PubMed]
  63. Wan, Y.; Feng, G.; Calakos, N. Sapap3 deletion causes mGluR5-dependent silencing of AMPAR synapses. J. Neurosci. 2011, 31, 16685–16691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Chen, M.; Wan, Y.; Ade, K.; Ting, J.; Feng, G.; Calakos, N. Sapap3 deletion anomalously activates short-term endocannabinoid-mediated synaptic plasticity. J. Neurosci. 2011, 31, 9563–9573. [Google Scholar] [CrossRef] [PubMed]
  65. Welch, J.M.; Lu, J.; Rodriguiz, R.M.; Trotta, N.C.; Peca, J.; Ding, J.-D.; Feliciano, C.; Chen, M.; Adams, J.P.; Luo, J.; et al. Cortico-striatal synaptic defects and OCD-like behaviours in Sapap3-mutant mice. Nature 2007, 448, 894–900. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Tesdahl, N.S.; King, D.K.; McDaniel, L.N.; Pieper, A.A. Altered ultrasonic vocalization in neonatal SAPAP3-deficient mice. Neuroreport 2017, 28, 1115–1118. [Google Scholar] [CrossRef]
  67. Beckelman, B.C.; Zhou, X.; Keene, C.D.; Ma, T. Impaired Eukaryotic Elongation Factor 1A Expression in Alzheimer’s Disease. Neurodegener. Dis. 2016, 16, 39–43. [Google Scholar] [CrossRef]
  68. Tsokas, P.; Grace, E.A.; Chan, P.; Ma, T.; Sealfon, S.C.; Iyengar, R.; Landau, E.M.; Blitzer, R.D. Local protein synthesis mediates a rapid increase in dendritic elongation factor 1A after induction of late long-term potentiation. J. Neurosci. 2005, 25, 5833–5843. [Google Scholar] [CrossRef]
  69. Najmabadi, H.; Hu, H.; Garshasbi, M.; Zemojtel, T.; Abedini, S.S.; Chen, W.; Hosseini, M.; Behjati, F.; Haas, S.; Jamali, P.; et al. Deep sequencing reveals 50 novel genes for recessive cognitive disorders. Nature 2011, 478, 57–63. [Google Scholar] [CrossRef]
  70. Heise, C.; Gardoni, F.; Culotta, L.; di Luca, M.; Verpelli, C.; Sala, C. Elongation factor-2 phosphorylation in dendrites and the regulation of dendritic mRNA translation in neurons. Front. Cell. Neurosci. 2014, 8, 35. [Google Scholar] [CrossRef] [Green Version]
  71. Martinetz, S.; Meinung, C.-P.; Jurek, B.; von Schack, D.; van den Burg, E.H.; Slattery, D.A.; Neumann, I.D. De Novo Protein Synthesis Mediated by the Eukaryotic Elongation Factor 2 Is Required for the Anxiolytic Effect of Oxytocin. Biol. Psychiatry 2019, 85, 802–811. [Google Scholar] [CrossRef] [PubMed]
  72. Li, X.; Alafuzoff, I.; Soininen, H.; Winblad, B.; Pei, J.-J. Levels of mTOR and its downstream targets 4E-BP1, eEF2, and eEF2 kinase in relationships with tau in Alzheimer’s disease brain. FEBS J. 2005, 272, 4211–4220. [Google Scholar] [CrossRef] [PubMed]
  73. Iizuka, A.; Sengoku, K.; Iketani, M.; Nakamura, F.; Sato, Y.; Matsushita, M.; Nairn, A.C.; Takamatsu, K.; Goshima, Y.; Takei, K. Calcium-induced synergistic inhibition of a translational factor eEF2 in nerve growth cones. Biochem. Biophys. Res. Commun. 2007, 353, 244–250. [Google Scholar] [CrossRef] [PubMed]
  74. Sutton, M.A.; Taylor, A.M.; Ito, H.T.; Pham, A.; Schuman, E.M. Postsynaptic decoding of neural activity: eEF2 as a biochemical sensor coupling miniature synaptic transmission to local protein synthesis. Neuron 2007, 55, 648–661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Park, S.; Park, J.M.; Kim, S.; Kim, J.-A.; Shepherd, J.D.; Smith-Hicks, C.L.; Chowdhury, S.; Kaufmann, W.; Kuhl, D.; Ryazanov, A.G.; et al. Elongation factor 2 and fragile X mental retardation protein control the dynamic translation of Arc/Arg3.1 essential for mGluR-LTD. Neuron 2008, 59, 70–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Takei, N.; Kawamura, M.; Ishizuka, Y.; Kakiya, N.; Inamura, N.; Namba, H.; Nawa, H. Brain-derived neurotrophic factor enhances the basal rate of protein synthesis by increasing active eukaryotic elongation factor 2 levels and promoting translation elongation in cortical neurons. J. Biol. Chem. 2009, 284, 26340–26348. [Google Scholar] [CrossRef] [Green Version]
  77. Verpelli, C.; Piccoli, G.; Zibetti, C.; Zanchi, A.; Gardoni, F.; Huang, K.; Brambilla, D.; Di Luca, M.; Battaglioli, E.; Sala, C. Synaptic activity controls dendritic spine morphology by modulating eEF2-dependent BDNF synthesis. J. Neurosci. 2010, 30, 5830–5842. [Google Scholar] [CrossRef]
  78. McCamphill, P.K.; Farah, C.A.; Anadolu, M.N.; Hoque, S.; Sossin, W.S. Bidirectional regulation of eEF2 phosphorylation controls synaptic plasticity by decoding neuronal activity patterns. J. Neurosci. 2015, 35, 4403–4417. [Google Scholar] [CrossRef] [Green Version]
  79. Martin, H.C.; Jones, W.D.; McIntyre, R.; Sanchez-Andrade, G.; Sanderson, M.; Stephenson, J.D.; Jones, C.P.; Handsaker, J.; Gallone, G.; Bruntraeger, M.; et al. Quantifying the contribution of recessive coding variation to developmental disorders. Science 2018, 362, 1161–1164. [Google Scholar] [CrossRef] [Green Version]
  80. Del Prete, D.; Rice, R.C.; Rajadhyaksha, A.M.; D’Adamio, L. Amyloid Precursor Protein (APP) May Act as a Substrate and a Recognition Unit for CRL4CRBN and Stub1 E3 Ligases Facilitating Ubiquitination of Proteins Involved in Presynaptic Functions and Neurodegeneration. J. Biol. Chem. 2016, 291, 17209–17227. [Google Scholar] [CrossRef] [Green Version]
  81. Smith, A.R.; Smith, R.G.; Pishva, E.; Hannon, E.; Roubroeks, J.A.Y.; Burrage, J.; Troakes, C.; Al-Sarraj, S.; Sloan, C.; Mill, J.; et al. Parallel profiling of DNA methylation and hydroxymethylation highlights neuropathology-associated epigenetic variation in Alzheimer’s disease. Clin. Epigenetics 2019, 11, 52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Gokoffski, K.K.; Wu, H.-H.; Beites, C.L.; Kim, J.; Kim, E.J.; Matzuk, M.M.; Johnson, J.E.; Lander, A.D.; Calof, A.L. Activin and GDF11 collaborate in feedback control of neuroepithelial stem cell proliferation and fate. Development 2011, 138, 4131–4142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Augustin, H.; McGourty, K.; Steinert, J.R.; Cochemé, H.M.; Adcott, J.; Cabecinha, M.; Vincent, A.; Halff, E.F.; Kittler, J.T.; Boucrot, E.; et al. Myostatin-like proteins regulate synaptic function and neuronal morphology. Development 2017, 144, 2445–2455. [Google Scholar] [CrossRef] [Green Version]
  84. Shi, Y.; Liu, J.-P. Gdf11 facilitates temporal progression of neurogenesis in the developing spinal cord. J. Neurosci. 2011, 31, 883–893. [Google Scholar] [CrossRef] [PubMed]
  85. Phiel, C.J.; Wilson, C.A.; Lee, V.M.-Y.; Klein, P.S. GSK-3alpha regulates production of Alzheimer’s disease amyloid-beta peptides. Nature 2003, 423, 435–439. [Google Scholar] [CrossRef] [PubMed]
  86. Parisiadou, L.; Bethani, I.; Michaki, V.; Krousti, K.; Rapti, G.; Efthimiopoulos, S. Homer2 and Homer3 interact with amyloid precursor protein and inhibit Abeta production. Neurobiol. Dis. 2008, 30, 353–364. [Google Scholar] [CrossRef]
  87. Smothers, C.T.; Szumlinski, K.K.; Worley, P.F.; Woodward, J.J. Altered NMDA receptor function in primary cultures of hippocampal neurons from mice lacking the Homer2 gene. Synapse 2016, 70, 33–39. [Google Scholar] [CrossRef] [Green Version]
  88. Shiraishi-Yamaguchi, Y.; Sato, Y.; Sakai, R.; Mizutani, A.; Knöpfel, T.; Mori, N.; Mikoshiba, K.; Furuichi, T. Interaction of Cupidin/Homer2 with two actin cytoskeletal regulators, Cdc42 small GTPase and Drebrin, in dendritic spines. BMC Neurosci. 2009, 10, 25. [Google Scholar] [CrossRef] [Green Version]
  89. McGuier, N.S.; Padula, A.E.; Mulholland, P.J.; Chandler, L.J. Homer2 deletion alters dendritic spine morphology but not alcohol-associated adaptations in GluN2B-containing N-methyl-D-aspartate receptors in the nucleus accumbens. Front. Pharmacol. 2015, 6, 28. [Google Scholar] [CrossRef] [Green Version]
  90. Lamprecht, R.; Dracheva, S.; Assoun, S.; LeDoux, J.E. Fear conditioning induces distinct patterns of gene expression in lateral amygdala. Genes Brain Behav. 2009, 8, 735–743. [Google Scholar] [CrossRef] [Green Version]
  91. Few, A.P.; Lautermilch, N.J.; Westenbroek, R.E.; Scheuer, T.; Catterall, W.A. Differential regulation of CaV2.1 channels by calcium-binding protein 1 and visinin-like protein-2 requires N-terminal myristoylation. J. Neurosci. 2005, 25, 7071–7080. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Proepper, C.; Putz, S.; Russell, R.; Boeckers, T.M.; Liebau, S. The Kvβ2 subunit of voltage-gated potassium channels is interacting with ProSAP2/Shank3 in the PSD. Neuroscience 2014, 261, 133–143. [Google Scholar] [CrossRef] [PubMed]
  93. Perkowski, J.J.; Murphy, G.G. Deletion of the mouse homolog of KCNAB2, a gene linked to monosomy 1p36, results in associative memory impairments and amygdala hyperexcitability. J. Neurosci. 2011, 31, 46–54. [Google Scholar] [CrossRef] [PubMed]
  94. Hanlon, D.W.; Yang, Z.; Goldstein, L.S. Characterization of KIFC2, a neuronal kinesin superfamily member in mouse. Neuron 1997, 18, 439–451. [Google Scholar] [CrossRef] [Green Version]
  95. Saito, N.; Okada, Y.; Noda, Y.; Kinoshita, Y.; Kondo, S.; Hirokawa, N. KIFC2 is a novel neuron-specific C-terminal type kinesin superfamily motor for dendritic transport of multivesicular body-like organelles. Neuron 1997, 18, 425–438. [Google Scholar] [CrossRef] [Green Version]
  96. Kelley, C.M.; Ginsberg, S.D.; Alldred, M.J.; Strupp, B.J.; Mufson, E.J. Maternal Choline Supplementation Alters Basal Forebrain Cholinergic Neuron Gene Expression in the Ts65Dn Mouse Model of Down Syndrome. Dev. Neurobiol. 2019, 79, 664–683. [Google Scholar] [CrossRef]
  97. Bandura, J.; Feng, Z.-P. Current Understanding of the Role of Neuronal Calcium Sensor 1 in Neurological Disorders. Mol. Neurobiol. 2019, 56, 6080–6094. [Google Scholar] [CrossRef]
  98. Nakamura, T.Y.; Nakao, S.; Wakabayashi, S. Emerging Roles of Neuronal Ca2+ Sensor-1 in Cardiac and Neuronal Tissues: A Mini Review. Front. Mol. Neurosci. 2019, 12, 56. [Google Scholar] [CrossRef] [Green Version]
  99. Dason, J.S.; Romero-Pozuelo, J.; Atwood, H.L.; Ferrús, A. Multiple roles for frequenin/NCS-1 in synaptic function and development. Mol. Neurobiol. 2012, 45, 388–402. [Google Scholar] [CrossRef] [Green Version]
  100. Heyne, H.O.; Lautenschläger, S.; Nelson, R.; Besnier, F.; Rotival, M.; Cagan, A.; Kozhemyakina, R.; Plyusnina, I.Z.; Trut, L.; Carlborg, Ö.; et al. Genetic influences on brain gene expression in rats selected for tameness and aggression. Genetics 2014, 198, 1277–1290. [Google Scholar] [CrossRef] [Green Version]
  101. Yamada, M.; Toba, S.; Takitoh, T.; Yoshida, Y.; Mori, D.; Nakamura, T.; Iwane, A.H.; Yanagida, T.; Imai, H.; Yu-Lee, L.-Y.; et al. mNUDC is required for plus-end-directed transport of cytoplasmic dynein and dynactins by kinesin-1. EMBO J. 2010, 29, 517–531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Cappello, S.; Monzo, P.; Vallee, R.B. NudC is required for interkinetic nuclear migration and neuronal migration during neocortical development. Dev. Biol. 2011, 357, 326–335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Aumais, J.P.; Tunstead, J.R.; McNeil, R.S.; Schaar, B.T.; McConnell, S.K.; Lin, S.H.; Clark, G.D.; Yu-Lee, L.Y. NudC associates with Lis1 and the dynein motor at the leading pole of neurons. J. Neurosci. 2001, 21, RC187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Cruchaga, C.; Karch, C.M.; Jin, S.C.; Benitez, B.A.; Cai, Y.; Guerreiro, R.; Harari, O.; Norton, J.; Budde, J.; Bertelsen, S.; et al. Rare coding variants in the phospholipase D3 gene confer risk for Alzheimer’s disease. Nature 2014, 505, 550–554. [Google Scholar] [CrossRef] [Green Version]
  105. Wang, J.; Yu, J.-T.; Tan, L. PLD3 in Alzheimer’s disease. Mol. Neurobiol. 2015, 51, 480–486. [Google Scholar] [CrossRef]
  106. Satoh, J.-I.; Kino, Y.; Yamamoto, Y.; Kawana, N.; Ishida, T.; Saito, Y.; Arima, K. PLD3 is accumulated on neuritic plaques in Alzheimer’s disease brains. Alzheimers Res. Ther. 2014, 6, 70. [Google Scholar] [CrossRef] [Green Version]
  107. Lambert, J.-C.; Grenier-Boley, B.; Bellenguez, C.; Pasquier, F.; Campion, D.; Dartigues, J.-F.; Berr, C.; Tzourio, C.; Amouyel, P. PLD3 and sporadic Alzheimer’s disease risk. Nature 2015, 520, E1. [Google Scholar] [CrossRef]
  108. van der Lee, S.J.; Holstege, H.; Wong, T.H.; Jakobsdottir, J.; Bis, J.C.; Chouraki, V.; van Rooij, J.G.J.; Grove, M.L.; Smith, A.V.; Amin, N.; et al. PLD3 variants in population studies. Nature 2015, 520, E2–E3. [Google Scholar] [CrossRef] [Green Version]
  109. Heilmann, S.; Drichel, D.; Clarimon, J.; Fernández, V.; Lacour, A.; Wagner, H.; Thelen, M.; Hernández, I.; Fortea, J.; Alegret, M.; et al. PLD3 in non-familial Alzheimer’s disease. Nature 2015, 520, E3–E5. [Google Scholar] [CrossRef]
  110. Fazzari, P.; Horre, K.; Arranz, A.M.; Frigerio, C.S.; Saito, T.; Saido, T.C.; De Strooper, B. PLD3 gene and processing of APP. Nature 2017, 541, E1–E2. [Google Scholar] [CrossRef]
  111. Zhang, D.-F.; Fan, Y.; Wang, D.; Bi, R.; Zhang, C.; Fang, Y.; Yao, Y.-G. PLD3 in Alzheimer’s Disease: A Modest Effect as Revealed by Updated Association and Expression Analyses. Mol. Neurobiol. 2016, 53, 4034–4045. [Google Scholar] [CrossRef]
  112. Roginski, R.S.; Lau, C.W.; Santoiemma, P.P.; Weaver, S.J.; Du, P.; Soteropoulos, P.; Yang, J. The human GCOM1 complex gene interacts with the NMDA receptor and internexin-alpha. Gene 2018, 648, 42–53. [Google Scholar] [CrossRef] [Green Version]
  113. Roginski, R.S.; Goubaeva, F.; Mikami, M.; Fried-Cassorla, E.; Nair, M.R.; Yang, J. GRINL1A colocalizes with N-methyl D-aspartate receptor NR1 subunit and reduces N-methyl D-aspartate toxicity. Neuroreport 2008, 19, 1721–1726. [Google Scholar] [CrossRef]
  114. Wong, M.Y.; Liu, C.; Wang, S.S.H.; Roquas, A.C.F.; Fowler, S.C.; Kaeser, P.S. Liprin-α3 controls vesicle docking and exocytosis at the active zone of hippocampal synapses. Proc. Natl. Acad. Sci. USA 2018, 115, 2234–2239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Bruschetta, G.; Jin, S.; Kim, J.D.; Diano, S. Prolyl carboxypeptidase in Agouti-related Peptide neurons modulates food intake and body weight. Mol. Metab. 2018, 10, 28–38. [Google Scholar] [CrossRef]
  116. Sakagami, H.; Kamata, A.; Fukunaga, K.; Kondo, H. Functional assay of EFA6A, a guanine nucleotide exchange factor for ADP-ribosylation factor 6 (ARF6), in dendritic formation of hippocampal neurons. Meth. Enzymol. 2005, 404, 232–242. [Google Scholar]
  117. Choi, S.; Ko, J.; Lee, J.-R.; Lee, H.W.; Kim, K.; Chung, H.S.; Kim, H.; Kim, E. ARF6 and EFA6A regulate the development and maintenance of dendritic spines. J. Neurosci. 2006, 26, 4811–4819. [Google Scholar] [CrossRef] [Green Version]
  118. Fukaya, M.; Fukushima, D.; Hara, Y.; Sakagami, H. EFA6A, a guanine nucleotide exchange factor for Arf6, interacts with sorting nexin-1 and regulates neurite outgrowth. J. Neurochem. 2014, 129, 21–36. [Google Scholar] [CrossRef]
  119. Sironi, C.; Teesalu, T.; Muggia, A.; Fontana, G.; Marino, F.; Savaresi, S.; Talarico, D. EFA6A encodes two isoforms with distinct biological activities in neuronal cells. J. Cell. Sci. 2009, 122, 2108–2118. [Google Scholar] [CrossRef] [Green Version]
  120. Sakagami, H. The EFA6 family: Guanine nucleotide exchange factors for ADP ribosylation factor 6 at neuronal synapses. Tohoku J. Exp. Med. 2008, 214, 191–198. [Google Scholar] [CrossRef] [Green Version]
  121. Dong, H.; Zhu, M.; Meng, L.; Ding, Y.; Yang, D.; Zhang, S.; Qiang, W.; Fisher, D.W.; Xu, E.Y. Pumilio2 regulates synaptic plasticity via translational repression of synaptic receptors in mice. Oncotarget 2018, 9, 32134–32148. [Google Scholar] [CrossRef] [Green Version]
  122. Siemen, H.; Colas, D.; Heller, H.C.; Brüstle, O.; Pera, R.A.R. Pumilio-2 function in the mouse nervous system. PLoS ONE 2011, 6, e25932. [Google Scholar] [CrossRef] [Green Version]
  123. Vessey, J.P.; Schoderboeck, L.; Gingl, E.; Luzi, E.; Riefler, J.; Di Leva, F.; Karra, D.; Thomas, S.; Kiebler, M.A.; Macchi, P. Mammalian Pumilio 2 regulates dendrite morphogenesis and synaptic function. Proc. Natl. Acad. Sci. USA 2010, 107, 3222–3227. [Google Scholar] [CrossRef] [Green Version]
  124. Marrero, E.; Rossi, S.G.; Darr, A.; Tsoulfas, P.; Rotundo, R.L. Translational regulation of acetylcholinesterase by the RNA-binding protein Pumilio-2 at the neuromuscular synapse. J. Biol. Chem. 2011, 286, 36492–36499. [Google Scholar] [CrossRef] [Green Version]
  125. Follwaczny, P.; Schieweck, R.; Riedemann, T.; Demleitner, A.; Straub, T.; Klemm, A.H.; Bilban, M.; Sutor, B.; Popper, B.; Kiebler, M.A. Pumilio2-deficient mice show a predisposition for epilepsy. Dis. Model Mech. 2017, 10, 1333–1342. [Google Scholar] [CrossRef] [Green Version]
  126. Zhang, M.; Chen, D.; Xia, J.; Han, W.; Cui, X.; Neuenkirchen, N.; Hermes, G.; Sestan, N.; Lin, H. Post-transcriptional regulation of mouse neurogenesis by Pumilio proteins. Genes Dev. 2017, 31, 1354–1369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Wu, X.-L.; Huang, H.; Huang, Y.-Y.; Yuan, J.-X.; Zhou, X.; Chen, Y.-M. Reduced Pumilio-2 expression in patients with temporal lobe epilepsy and in the lithium-pilocarpine induced epilepsy rat model. Epilepsy Behav. 2015, 50, 31–39. [Google Scholar] [CrossRef]
  128. Driscoll, H.E.; Muraro, N.I.; He, M.; Baines, R.A. Pumilio-2 regulates translation of Nav1.6 to mediate homeostasis of membrane excitability. J. Neurosci. 2013, 33, 9644–9654. [Google Scholar] [CrossRef] [Green Version]
  129. Vessey, J.P.; Vaccani, A.; Xie, Y.; Dahm, R.; Karra, D.; Kiebler, M.A.; Macchi, P. Dendritic localization of the translational repressor Pumilio 2 and its contribution to dendritic stress granules. J. Neurosci. 2006, 26, 6496–6508. [Google Scholar] [CrossRef]
  130. Muto, A.; Arai, K.-I.; Watanabe, S. Rab11-FIP4 is predominantly expressed in neural tissues and involved in proliferation as well as in differentiation during zebrafish retinal development. Dev. Biol. 2006, 292, 90–102. [Google Scholar] [CrossRef] [Green Version]
  131. Nishimura, N.; Van Huyen Pham, T.; Hartomo, T.B.; Lee, M.J.; Hasegawa, D.; Takeda, H.; Kawasaki, K.; Kosaka, Y.; Yamamoto, T.; Morikawa, S.; et al. Rab15 expression correlates with retinoic acid-induced differentiation of neuroblastoma cells. Oncol. Rep. 2011, 26, 145–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Hao, H.; Veleri, S.; Sun, B.; Kim, D.S.; Keeley, P.W.; Kim, J.-W.; Yang, H.-J.; Yadav, S.P.; Manjunath, S.H.; Sood, R.; et al. Regulation of a novel isoform of Receptor Expression Enhancing Protein REEP6 in rod photoreceptors by bZIP transcription factor NRL. Hum. Mol. Genet. 2014, 23, 4260–4271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Agrawal, S.A.; Burgoyne, T.; Eblimit, A.; Bellingham, J.; Parfitt, D.A.; Lane, A.; Nichols, R.; Asomugha, C.; Hayes, M.J.; Munro, P.M.; et al. REEP6 deficiency leads to retinal degeneration through disruption of ER homeostasis and protein trafficking. Hum. Mol. Genet. 2017, 26, 2667–2677. [Google Scholar] [CrossRef] [PubMed]
  134. Veleri, S.; Nellissery, J.; Mishra, B.; Manjunath, S.H.; Brooks, M.J.; Dong, L.; Nagashima, K.; Qian, H.; Gao, C.; Sergeev, Y.V.; et al. REEP6 mediates trafficking of a subset of Clathrin-coated vesicles and is critical for rod photoreceptor function and survival. Hum. Mol. Genet. 2017, 26, 2218–2230. [Google Scholar] [CrossRef] [Green Version]
  135. Arno, G.; Agrawal, S.A.; Eblimit, A.; Bellingham, J.; Xu, M.; Wang, F.; Chakarova, C.; Parfitt, D.A.; Lane, A.; Burgoyne, T.; et al. Mutations in REEP6 Cause Autosomal-Recessive Retinitis Pigmentosa. Am. J. Hum. Genet. 2016, 99, 1305–1315. [Google Scholar] [CrossRef] [PubMed]
  136. Dinamarca, M.C.; Guzzetti, F.; Karpova, A.; Lim, D.; Mitro, N.; Musardo, S.; Mellone, M.; Marcello, E.; Stanic, J.; Samaddar, T.; et al. Ring finger protein 10 is a novel synaptonuclear messenger encoding activation of NMDA receptors in hippocampus. Elife 2016, 5, e12430. [Google Scholar] [CrossRef] [Green Version]
  137. Carrano, N.; Samaddar, T.; Brunialti, E.; Franchini, L.; Marcello, E.; Ciana, P.; Mauceri, D.; Di Luca, M.; Gardoni, F. The Synaptonuclear Messenger RNF10 Acts as an Architect of Neuronal Morphology. Mol. Neurobiol. 2019, 56, 7583–7593. [Google Scholar] [CrossRef]
  138. Mateu-Huertas, E.; Rodriguez-Revenga, L.; Alvarez-Mora, M.I.; Madrigal, I.; Willemsen, R.; Milà, M.; Martí, E.; Estivill, X. Blood expression profiles of fragile X premutation carriers identify candidate genes involved in neurodegenerative and infertility phenotypes. Neurobiol. Dis. 2014, 65, 43–54. [Google Scholar] [CrossRef]
  139. Marcello, E.; Di Luca, M.; Gardoni, F. Synapse-to-nucleus communication: From developmental disorders to Alzheimer’s disease. Curr. Opin. Neurobiol. 2018, 48, 160–166. [Google Scholar] [CrossRef]
  140. Malik, Y.S.; Sheikh, M.A.; Lai, M.; Cao, R.; Zhu, X. RING finger protein 10 regulates retinoic acid-induced neuronal differentiation and the cell cycle exit of P19 embryonic carcinoma cells. J. Cell. Biochem. 2013, 114, 2007–2015. [Google Scholar] [CrossRef]
  141. Hoshikawa, S.; Ogata, T.; Fujiwara, S.; Nakamura, K.; Tanaka, S. A novel function of RING finger protein 10 in transcriptional regulation of the myelin-associated glycoprotein gene and myelin formation in Schwann cells. PLoS ONE 2008, 3, e3464. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Klauck, S.M.; Felder, B.; Kolb-Kokocinski, A.; Schuster, C.; Chiocchetti, A.; Schupp, I.; Wellenreuther, R.; Schmötzer, G.; Poustka, F.; Breitenbach-Koller, L.; et al. Mutations in the ribosomal protein gene RPL10 suggest a novel modulating disease mechanism for autism. Mol. Psychiatry 2006, 11, 1073–1084. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Gong, X.; Delorme, R.; Fauchereau, F.; Durand, C.M.; Chaste, P.; Betancur, C.; Goubran-Botros, H.; Nygren, G.; Anckarsäter, H.; Rastam, M.; et al. An investigation of ribosomal protein L10 gene in autism spectrum disorders. BMC Med. Genet. 2009, 10, 7. [Google Scholar] [CrossRef]
  144. Chiocchetti, A.G.; Haslinger, D.; Boesch, M.; Karl, T.; Wiemann, S.; Freitag, C.M.; Poustka, F.; Scheibe, B.; Bauer, J.W.; Hintner, H.; et al. Protein signatures of oxidative stress response in a patient specific cell line model for autism. Mol. Autism 2014, 5, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Zanni, G.; Kalscheuer, V.M.; Friedrich, A.; Barresi, S.; Alfieri, P.; Di Capua, M.; Haas, S.A.; Piccini, G.; Karl, T.; Klauck, S.M.; et al. A Novel Mutation in RPL10 (Ribosomal Protein L10) Causes X-Linked Intellectual Disability, Cerebellar Hypoplasia, and Spondylo-Epiphyseal Dysplasia. Hum. Mutat. 2015, 36, 1155–1158. [Google Scholar] [CrossRef] [PubMed]
  146. Vandewalle, J.; Van Esch, H.; Govaerts, K.; Verbeeck, J.; Zweier, C.; Madrigal, I.; Mila, M.; Pijkels, E.; Fernandez, I.; Kohlhase, J.; et al. Dosage-dependent severity of the phenotype in patients with mental retardation due to a recurrent copy-number gain at Xq28 mediated by an unusual recombination. Am. J. Hum. Genet. 2009, 85, 809–822. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Brooks, S.S.; Wall, A.L.; Golzio, C.; Reid, D.W.; Kondyles, A.; Willer, J.R.; Botti, C.; Nicchitta, C.V.; Katsanis, N.; Davis, E.E. A novel ribosomopathy caused by dysfunction of RPL10 disrupts neurodevelopment and causes X-linked microcephaly in humans. Genetics 2014, 198, 723–733. [Google Scholar] [CrossRef]
  148. Kalendar, R.; Belyayev, A.; Zachepilo, T.; Vaido, A.; Maidanyuk, D.; Schulman, A.H.; Dyuzhikova, N. Copy-number variation of housekeeping gene rpl13a in rat strains selected for nervous system excitability. Mol. Cell. Probes 2017, 33, 11–15. [Google Scholar] [CrossRef]
  149. Perucho, L.; Artero-Castro, A.; Guerrero, S.; Ramón y Cajal, S.; LLeonart, M.E.; Wang, Z.-Q. RPLP1, a crucial ribosomal protein for embryonic development of the nervous system. PLoS ONE 2014, 9, e99956. [Google Scholar] [CrossRef] [Green Version]
  150. Evans, H.T.; Benetatos, J.; van Roijen, M.; Bodea, L.-G.; Götz, J. Decreased synthesis of ribosomal proteins in tauopathy revealed by non-canonical amino acid labelling. EMBO J. 2019, 38, e101174. [Google Scholar] [CrossRef]
  151. Miyata, S.; Mori, Y.; Tohyama, M. PRMT3 is essential for dendritic spine maturation in rat hippocampal neurons. Brain Res. 2010, 1352, 11–20. [Google Scholar] [CrossRef] [PubMed]
  152. Grupe, A.; Li, Y.; Rowland, C.; Nowotny, P.; Hinrichs, A.L.; Smemo, S.; Kauwe, J.S.K.; Maxwell, T.J.; Cherny, S.; Doil, L.; et al. A scan of chromosome 10 identifies a novel locus showing strong association with late-onset Alzheimer disease. Am. J. Hum. Genet. 2006, 78, 78–88. [Google Scholar] [CrossRef] [Green Version]
  153. Kalathur, R.K.R.; Giner-Lamia, J.; Machado, S.; Barata, T.; Ayasolla, K.R.S.; Futschik, M.E. The unfolded protein response and its potential role in Huntington’s disease elucidated by a systems biology approach. F1000Res 2015, 4, 103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. De Graeve, S.; Marinelli, S.; Stolz, F.; Hendrix, J.; Vandamme, J.; Engelborghs, Y.; Van Dijck, P.; Thevelein, J.M. Mammalian ribosomal and chaperone protein RPS3A counteracts α-synuclein aggregation and toxicity in a yeast model system. Biochem. J. 2013, 455, 295–306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Watkins-Chow, D.E.; Cooke, J.; Pidsley, R.; Edwards, A.; Slotkin, R.; Leeds, K.E.; Mullen, R.; Baxter, L.L.; Campbell, T.G.; Salzer, M.C.; et al. Mutation of the diamond-blackfan anemia gene Rps7 in mouse results in morphological and neuroanatomical phenotypes. PLoS Genet. 2013, 9, e1003094. [Google Scholar] [CrossRef] [Green Version]
  156. Gong, X.; Wang, H. SHANK1 and autism spectrum disorders. Sci. China Life Sci. 2015, 58, 985–990. [Google Scholar] [CrossRef] [Green Version]
  157. Monteiro, P.; Feng, G. SHANK proteins: Roles at the synapse and in autism spectrum disorder. Nat. Rev. Neurosci. 2017, 18, 147–157. [Google Scholar] [CrossRef]
  158. Guilmatre, A.; Huguet, G.; Delorme, R.; Bourgeron, T. The emerging role of SHANK genes in neuropsychiatric disorders. Dev. Neurobiol. 2014, 74, 113–122. [Google Scholar] [CrossRef]
  159. Sungur, A.Ö.; Schwarting, R.K.W.; Wöhr, M. Behavioral phenotypes and neurobiological mechanisms in the Shank1 mouse model for autism spectrum disorder: A translational perspective. Behav. Brain Res. 2018, 352, 46–61. [Google Scholar] [CrossRef]
  160. Jiang, Y.-H.; Ehlers, M.D. Modeling autism by SHANK gene mutations in mice. Neuron 2013, 78, 8–27. [Google Scholar] [CrossRef] [Green Version]
  161. Wu, L.-X.; Sun, C.-K.; Zhang, Y.-M.; Fan, M.; Xu, J.; Ma, H.; Zhang, J. Involvement of the Snk-SPAR pathway in glutamate-induced excitotoxicity in cultured hippocampal neurons. Brain Res. 2007, 1168, 38–45. [Google Scholar] [CrossRef] [PubMed]
  162. Pak, D.T.S.; Sheng, M. Targeted protein degradation and synapse remodeling by an inducible protein kinase. Science 2003, 302, 1368–1373. [Google Scholar] [CrossRef] [PubMed]
  163. Meyer, G.; Brose, N. Neuroscience. SPARring with spines. Science 2003, 302, 1341–1344. [Google Scholar] [CrossRef] [PubMed]
  164. Maruoka, H.; Konno, D.; Hori, K.; Sobue, K. Collaboration of PSD-Zip70 with its binding partner, SPAR, in dendritic spine maturity. J. Neurosci. 2005, 25, 1421–1430. [Google Scholar] [CrossRef]
  165. Lu, X.-J.; Chen, X.-Q.; Weng, J.; Zhang, H.-Y.; Pak, D.T.; Luo, J.-H.; Du, J.-Z. Hippocampal spine-associated Rap-specific GTPase-activating protein induces enhancement of learning and memory in postnatally hypoxia-exposed mice. Neuroscience 2009, 162, 404–414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Seeburg, D.P.; Feliu-Mojer, M.; Gaiottino, J.; Pak, D.T.S.; Sheng, M. Critical role of CDK5 and Polo-like kinase 2 in homeostatic synaptic plasticity during elevated activity. Neuron 2008, 58, 571–583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Herrick, S.; Evers, D.M.; Lee, J.-Y.; Udagawa, N.; Pak, D.T.S. Postsynaptic PDLIM5/Enigma Homolog binds SPAR and causes dendritic spine shrinkage. Mol. Cell. Neurosci. 2010, 43, 188–200. [Google Scholar] [CrossRef] [Green Version]
  168. Sui, H.; Lu, X.-G.; Zhan, L.-B.; Jiang, W.-Z.; Qi, X.; Gong, X.-Y.; Niu, X.-P. Decreased expression of spine-associated Rap guanosine triphosphatase-activating protein (SPAR) in glutamate-treated primary hippocampal neurons. J. Clin. Neurosci. 2010, 17, 1042–1046. [Google Scholar] [CrossRef]
  169. Chen, Y.; Yuanxiang, P.; Knöpfel, T.; Thomas, U.; Behnisch, T. Hippocampal LTP triggers proteasome-mediated SPAR degradation in CA1 neurons. Synapse 2012, 66, 142–150. [Google Scholar] [CrossRef]
  170. Mihalas, A.B.; Araki, Y.; Huganir, R.L.; Meffert, M.K. Opposing action of nuclear factor κB and Polo-like kinases determines a homeostatic end point for excitatory synaptic adaptation. J. Neurosci. 2013, 33, 16490–16501. [Google Scholar] [CrossRef] [Green Version]
  171. Ang, X.L.; Seeburg, D.P.; Sheng, M.; Harper, J.W. Regulation of postsynaptic RapGAP SPAR by Polo-like kinase 2 and the SCFbeta-TRCP ubiquitin ligase in hippocampal neurons. J. Biol. Chem. 2008, 283, 29424–29432. [Google Scholar] [CrossRef] [Green Version]
  172. Gong, X.; Lu, X.; Zhan, L.; Sui, H.; Qi, X.; Ji, Z.; Niu, X.; Liu, L. Role of the SNK-SPAR pathway in the development of Alzheimer’s disease. IUBMB Life 2010, 62, 214–221. [Google Scholar] [CrossRef]
  173. Sui, H.; Zhan, L.; Niu, X.; Liang, L.; Li, X. The SNK and SPAR signaling pathway changes in hippocampal neurons treated with amyloid-beta peptide in vitro. Neuropeptides 2017, 63, 43–48. [Google Scholar] [CrossRef]
  174. Al-Muhaizea, M.A.; AlMutairi, F.; Almass, R.; AlHarthi, S.; Aldosary, M.S.; Alsagob, M.; AlOdaib, A.; Colak, D.; Kaya, N. A Novel Homozygous Mutation in SPTBN2 Leads to Spinocerebellar Ataxia in a Consanguineous Family: Report of a New Infantile-Onset Case and Brief Review of the Literature. Cerebellum 2018, 17, 276–285. [Google Scholar] [CrossRef]
  175. Mizuno, T.; Kashimada, A.; Nomura, T.; Moriyama, K.; Yokoyama, H.; Hasegawa, S.; Takagi, M.; Mizutani, S. Infantile-onset spinocerebellar ataxia type 5 associated with a novel SPTBN2 mutation: A case report. Brain Dev. 2019, 41, 630–633. [Google Scholar] [CrossRef]
  176. Durr, A. Autosomal dominant cerebellar ataxias: Polyglutamine expansions and beyond. Lancet Neurol. 2010, 9, 885–894. [Google Scholar] [CrossRef]
  177. Parolin Schnekenberg, R.; Perkins, E.M.; Miller, J.W.; Davies, W.I.L.; D’Adamo, M.C.; Pessia, M.; Fawcett, K.A.; Sims, D.; Gillard, E.; Hudspith, K.; et al. De novo point mutations in patients diagnosed with ataxic cerebral palsy. Brain 2015, 138, 1817–1832. [Google Scholar] [CrossRef] [Green Version]
  178. Jacob, F.-D.; Ho, E.S.; Martinez-Ojeda, M.; Darras, B.T.; Khwaja, O.S. Case of infantile onset spinocerebellar ataxia type 5. J. Child Neurol. 2013, 28, 1292–1295. [Google Scholar] [CrossRef]
  179. Ikeda, Y.; Dick, K.A.; Weatherspoon, M.R.; Gincel, D.; Armbrust, K.R.; Dalton, J.C.; Stevanin, G.; Dürr, A.; Zühlke, C.; Bürk, K.; et al. Spectrin mutations cause spinocerebellar ataxia type 5. Nat. Genet. 2006, 38, 184–190. [Google Scholar] [CrossRef]
  180. Németh, A.H.; Kwasniewska, A.C.; Lise, S.; Parolin Schnekenberg, R.; Becker, E.B.E.; Bera, K.D.; Shanks, M.E.; Gregory, L.; Buck, D.; Zameel Cader, M.; et al. Next generation sequencing for molecular diagnosis of neurological disorders using ataxias as a model. Brain 2013, 136, 3106–3118. [Google Scholar] [CrossRef]
  181. Armbrust, K.R.; Wang, X.; Hathorn, T.J.; Cramer, S.W.; Chen, G.; Zu, T.; Kangas, T.; Zink, A.N.; Öz, G.; Ebner, T.J.; et al. Mutant β-III spectrin causes mGluR1α mislocalization and functional deficits in a mouse model of spinocerebellar ataxia type 5. J. Neurosci. 2014, 34, 9891–9904. [Google Scholar] [CrossRef] [Green Version]
  182. Perkins, E.M.; Clarkson, Y.L.; Sabatier, N.; Longhurst, D.M.; Millward, C.P.; Jack, J.; Toraiwa, J.; Watanabe, M.; Rothstein, J.D.; Lyndon, A.R.; et al. Loss of beta-III spectrin leads to Purkinje cell dysfunction recapitulating the behavior and neuropathology of spinocerebellar ataxia type 5 in humans. J. Neurosci. 2010, 30, 4857–4867. [Google Scholar] [CrossRef] [Green Version]
  183. Lise, S.; Clarkson, Y.; Perkins, E.; Kwasniewska, A.; Sadighi Akha, E.; Schnekenberg, R.P.; Suminaite, D.; Hope, J.; Baker, I.; Gregory, L.; et al. Recessive mutations in SPTBN2 implicate β-III spectrin in both cognitive and motor development. PLoS Genet. 2012, 8, e1003074. [Google Scholar] [CrossRef] [Green Version]
  184. Baltussen, L.L.; Rosianu, F.; Ultanir, S.K. Kinases in synaptic development and neurological diseases. Prog. Neuropsychopharmacol. Biol. Psychiatry 2018, 84, 343–352. [Google Scholar] [CrossRef]
  185. Yadav, S.; Oses-Prieto, J.A.; Peters, C.J.; Zhou, J.; Pleasure, S.J.; Burlingame, A.L.; Jan, L.Y.; Jan, Y.-N. TAOK2 Kinase Mediates PSD95 Stability and Dendritic Spine Maturation through Septin7 Phosphorylation. Neuron 2017, 93, 379–393. [Google Scholar] [CrossRef] [Green Version]
  186. Richter, M.; Murtaza, N.; Scharrenberg, R.; White, S.H.; Johanns, O.; Walker, S.; Yuen, R.K.C.; Schwanke, B.; Bedürftig, B.; Henis, M.; et al. Altered TAOK2 activity causes autism-related neurodevelopmental and cognitive abnormalities through RhoA signaling. Mol. Psychiatry 2019, 24, 1329–1350. [Google Scholar] [CrossRef] [Green Version]
  187. Chen, J.; Yu, S.; Fu, Y.; Li, X. Synaptic proteins and receptors defects in autism spectrum disorders. Front. Cell. Neurosci. 2014, 8, 276. [Google Scholar] [CrossRef] [Green Version]
  188. de Anda, F.C.; Rosario, A.L.; Durak, O.; Tran, T.; Gräff, J.; Meletis, K.; Rei, D.; Soda, T.; Madabhushi, R.; Ginty, D.D.; et al. Autism spectrum disorder susceptibility gene TAOK2 affects basal dendrite formation in the neocortex. Nat. Neurosci. 2012, 15, 1022–1031. [Google Scholar] [CrossRef]
  189. Roque, C.G.; Wong, H.H.-W.; Lin, J.Q.; Holt, C.E. Tumor protein Tctp regulates axon development in the embryonic visual system. Development 2016, 143, 1134–1148. [Google Scholar] [CrossRef] [Green Version]
  190. Filatova, E.V.; Shadrina, M.I.; Alieva, A.K.; Kolacheva, A.A.; Slominsky, P.A.; Ugrumov, M.V. Expression analysis of genes of ubiquitin-proteasome protein degradation system in MPTP-induced mice models of early stages of Parkinson’s disease. Dokl. Biochem. Biophys. 2014, 456, 116–118. [Google Scholar] [CrossRef]
  191. Vogl, A.M.; Brockmann, M.M.; Giusti, S.A.; Maccarrone, G.; Vercelli, C.A.; Bauder, C.A.; Richter, J.S.; Roselli, F.; Hafner, A.-S.; Dedic, N.; et al. Neddylation inhibition impairs spine development, destabilizes synapses and deteriorates cognition. Nat. Neurosci. 2015, 18, 239–251. [Google Scholar] [CrossRef]
  192. Deng, G.; Jin, L. The effects of vasoactive intestinal peptide in neurodegenerative disorders. Neurol. Res. 2017, 39, 65–72. [Google Scholar] [CrossRef]
  193. Fahrenkrug, J. Transmitter role of vasoactive intestinal peptide. Pharmacol. Toxicol. 1993, 72, 354–363. [Google Scholar] [CrossRef]
  194. Hill, J.M. Vasoactive intestinal peptide in neurodevelopmental disorders: Therapeutic potential. Curr. Pharm. Des. 2007, 13, 1079–1089. [Google Scholar] [CrossRef] [Green Version]
  195. Baßler, J.; Hurt, E. Eukaryotic Ribosome Assembly. Annu. Rev. Biochem. 2019, 88, 281–306. [Google Scholar] [CrossRef]
  196. Knowles, R.B.; Sabry, J.H.; Martone, M.E.; Deerinck, T.J.; Ellisman, M.H.; Bassell, G.J.; Kosik, K.S. Translocation of RNA granules in living neurons. J. Neurosci. 1996, 16, 7812–7820. [Google Scholar] [CrossRef] [Green Version]
  197. Mauro, V.P.; Edelman, G.M. The ribosome filter hypothesis. Proc. Natl. Acad. Sci. USA 2002, 99, 12031–12036. [Google Scholar] [CrossRef] [Green Version]
  198. Sugihara, Y.; Honda, H.; Iida, T.; Morinaga, T.; Hino, S.; Okajima, T.; Matsuda, T.; Nadano, D. Proteomic analysis of rodent ribosomes revealed heterogeneity including ribosomal proteins L10-like, L22-like 1, and L39-like. J. Proteome Res. 2010, 9, 1351–1366. [Google Scholar] [CrossRef]
  199. Shi, Z.; Fujii, K.; Kovary, K.M.; Genuth, N.R.; Röst, H.L.; Teruel, M.N.; Barna, M. Heterogeneous Ribosomes Preferentially Translate Distinct Subpools of mRNAs Genome-wide. Mol. Cell 2017, 67, 71–83.e7. [Google Scholar] [CrossRef] [Green Version]
  200. Kraushar, M.L.; Viljetic, B.; Wijeratne, H.R.S.; Thompson, K.; Jiao, X.; Pike, J.W.; Medvedeva, V.; Groszer, M.; Kiledjian, M.; Hart, R.P.; et al. Thalamic WNT3 Secretion Spatiotemporally Regulates the Neocortical Ribosome Signature and mRNA Translation to Specify Neocortical Cell Subtypes. J. Neurosci. 2015, 35, 10911–10926. [Google Scholar] [CrossRef] [Green Version]
  201. Kraushar, M.L.; Popovitchenko, T.; Volk, N.L.; Rasin, M.-R. The frontier of RNA metamorphosis and ribosome signature in neocortical development. Int. J. Dev. Neurosci. 2016, 55, 131–139. [Google Scholar] [CrossRef] [Green Version]
  202. Komili, S.; Farny, N.G.; Roth, F.P.; Silver, P.A. Functional specificity among ribosomal proteins regulates gene expression. Cell 2007, 131, 557–571. [Google Scholar] [CrossRef] [Green Version]
  203. Ulbrich, B.; Nierhaus, K.H. Pools of ribosomal proteins in Escherichia coli. Studies on the exchange of proteins between pools and ribosomes. Eur. J. Biochem. 1975, 57, 49–54. [Google Scholar] [CrossRef]
  204. Robertson, W.R.; Dowsett, S.J.; Hardy, S.J. Exchange of ribosomal proteins among the ribosomes of Escherichia coli. Mol. Gen. Genet. 1977, 157, 205–214. [Google Scholar] [CrossRef]
  205. Subramanian, A.R.; van Duin, J. Exchange of individual ribosomal proteins between ribosomes as studied by heavy isotope-transfer experiments. Mol. Gen. Genet. 1977, 158, 1–9. [Google Scholar] [CrossRef]
  206. Pulk, A.; Liiv, A.; Peil, L.; Maiväli, U.; Nierhaus, K.; Remme, J. Ribosome reactivation by replacement of damaged proteins. Mol. Microbiol. 2010, 75, 801–814. [Google Scholar] [CrossRef]
  207. Dever, T.E.; Green, R. The elongation, termination, and recycling phases of translation in eukaryotes. Cold Spring Harb. Perspect. Biol. 2012, 4, a013706. [Google Scholar] [CrossRef] [Green Version]
  208. Trimble, W.S.; Grinstein, S. Barriers to the free diffusion of proteins and lipids in the plasma membrane. J. Cell Biol. 2015, 208, 259–271. [Google Scholar] [CrossRef] [Green Version]
  209. Lemmon, M.A. Membrane recognition by phospholipid-binding domains. Nat. Rev. Mol. Cell Biol. 2008, 9, 99–111. [Google Scholar] [CrossRef]
  210. 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]
  211. Hou, L.; Antion, M.D.; Hu, D.; Spencer, C.M.; Paylor, R.; Klann, E. Dynamic translational and proteasomal regulation of fragile X mental retardation protein controls mGluR-dependent long-term depression. Neuron 2006, 51, 441–454. [Google Scholar] [CrossRef] [Green Version]
  212. Michalon, A.; Sidorov, M.; Ballard, T.M.; Ozmen, L.; Spooren, W.; Wettstein, J.G.; Jaeschke, G.; Bear, M.F.; Lindemann, L. Chronic pharmacological mGlu5 inhibition corrects fragile X in adult mice. Neuron 2012, 74, 49–56. [Google Scholar] [CrossRef] [Green Version]
  213. Paradee, W.; Melikian, H.E.; Rasmussen, D.L.; Kenneson, A.; Conn, P.J.; Warren, S.T. Fragile X mouse: Strain effects of knockout phenotype and evidence suggesting deficient amygdala function. Neuroscience 1999, 94, 185–192. [Google Scholar] [CrossRef]
  214. Ding, Q.; Sethna, F.; Wang, H. Behavioral analysis of male and female Fmr1 knockout mice on C57BL/6 background. Behav. Brain Res. 2014, 271, 72–78. [Google Scholar] [CrossRef] [Green Version]
  215. Qin, M.; Kang, J.; Smith, C.B. Increased rates of cerebral glucose metabolism in a mouse model of fragile X mental retardation. Proc. Natl. Acad. Sci. USA 2002, 99, 15758–15763. [Google Scholar] [CrossRef] [Green Version]
  216. Uutela, M.; Lindholm, J.; Louhivuori, V.; Wei, H.; Louhivuori, L.M.; Pertovaara, A.; Akerman, K.; Castrén, E.; Castrén, M.L. Reduction of BDNF expression in Fmr1 knockout mice worsens cognitive deficits but improves hyperactivity and sensorimotor deficits. Genes Brain Behav. 2012, 11, 513–523. [Google Scholar] [CrossRef]
  217. Baker, K.B.; Wray, S.P.; Ritter, R.; Mason, S.; Lanthorn, T.H.; Savelieva, K.V. Male and female Fmr1 knockout mice on C57 albino background exhibit spatial learning and memory impairments. Genes Brain Behav. 2010, 9, 562–574. [Google Scholar] [CrossRef]
  218. Kooy, R.F.; D’Hooge, R.; Reyniers, E.; Bakker, C.E.; Nagels, G.; De Boulle, K.; Storm, K.; Clincke, G.; De Deyn, P.P.; Oostra, B.A.; et al. Transgenic mouse model for the fragile X syndrome. Am. J. Med. Genet. 1996, 64, 241–245. [Google Scholar] [CrossRef]
  219. D’Hooge, R.; Nagels, G.; Franck, F.; Bakker, C.E.; Reyniers, E.; Storm, K.; Kooy, R.F.; Oostra, B.A.; Willems, P.J.; De Deyn, P.P. Mildly impaired water maze performance in male Fmr1 knockout mice. Neuroscience 1997, 76, 367–376. [Google Scholar] [CrossRef]
  220. The Dutch-Belgian Fragile X Consortium. Fmr1 knockout mice: A model to study fragile X mental retardation. Cell 1994, 78, 23–33. [Google Scholar]
  221. Shiina, N.; Yamaguchi, K.; Tokunaga, M. RNG105 deficiency impairs the dendritic localization of mRNAs for Na+/K+ ATPase subunit isoforms and leads to the degeneration of neuronal networks. J. Neurosci. 2010, 30, 12816–12830. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Ohashi, R.; Takao, K.; Miyakawa, T.; Shiina, N. Comprehensive behavioral analysis of RNG105 (Caprin1) heterozygous mice: Reduced social interaction and attenuated response to novelty. Sci. Rep. 2016, 6, 20775. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Kedersha, N.; Panas, M.D.; Achorn, C.A.; Lyons, S.; Tisdale, S.; Hickman, T.; Thomas, M.; Lieberman, J.; McInerney, G.M.; Ivanov, P.; et al. G3BP-Caprin1-USP10 complexes mediate stress granule condensation and associate with 40S subunits. J. Cell Biol. 2016, 212, 845–860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Tourrière, H.; Chebli, K.; Zekri, L.; Courselaud, B.; Blanchard, J.M.; Bertrand, E.; Tazi, J. The RasGAP-associated endoribonuclease G3BP assembles stress granules. J. Cell Biol. 2003, 160, 823–831. [Google Scholar] [CrossRef]
  225. Martin, S.; Zekri, L.; Metz, A.; Maurice, T.; Chebli, K.; Vignes, M.; Tazi, J. Deficiency of G3BP1, the stress granules assembly factor, results in abnormal synaptic plasticity and calcium homeostasis in neurons. J. Neurochem. 2013, 125, 175–184. [Google Scholar] [CrossRef]
  226. Richter, J.D. CPEB: A life in translation. Trends Biochem. Sci. 2007, 32, 279–285. [Google Scholar] [CrossRef]
  227. Alarcon, J.M.; Hodgman, R.; Theis, M.; Huang, Y.-S.; Kandel, E.R.; Richter, J.D. Selective modulation of some forms of schaffer collateral-CA1 synaptic plasticity in mice with a disruption of the CPEB-1 gene. Learn. Mem. 2004, 11, 318–327. [Google Scholar] [CrossRef] [Green Version]
  228. Berger-Sweeney, J.; Zearfoss, N.R.; Richter, J.D. Reduced extinction of hippocampal-dependent memories in CPEB knockout mice. Learn. Mem. 2006, 13, 4–7. [Google Scholar] [CrossRef] [Green Version]
  229. Heraud-Farlow, J.E.; Kiebler, M.A. The multifunctional Staufen proteins: Conserved roles from neurogenesis to synaptic plasticity. Trends Neurosci. 2014, 37, 470–479. [Google Scholar] [CrossRef] [Green Version]
  230. Lebeau, G.; Miller, L.C.; Tartas, M.; McAdam, R.; Laplante, I.; Badeaux, F.; DesGroseillers, L.; Sossin, W.S.; Lacaille, J.-C. Staufen 2 regulates mGluR long-term depression and Map1b mRNA distribution in hippocampal neurons. Learn. Mem. 2011, 18, 314–326. [Google Scholar] [CrossRef] [Green Version]
  231. Vessey, J.P.; Macchi, P.; Stein, J.M.; Mikl, M.; Hawker, K.N.; Vogelsang, P.; Wieczorek, K.; Vendra, G.; Riefler, J.; Tübing, F.; et al. A loss of function allele for murine Staufen1 leads to impairment of dendritic Staufen1-RNP delivery and dendritic spine morphogenesis. Proc. Natl. Acad. Sci. USA 2008, 105, 16374–16379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Berger, S.M.; Fernández-Lamo, I.; Schönig, K.; Fernández Moya, S.M.; Ehses, J.; Schieweck, R.; Clementi, S.; Enkel, T.; Grothe, S.; von Bohlen Und Halbach, O.; et al. Forebrain-specific, conditional silencing of Staufen2 alters synaptic plasticity, learning, and memory in rats. Genome Biol. 2017, 18, 222. [Google Scholar] [CrossRef] [PubMed]
  233. Popper, B.; Demleitner, A.; Bolivar, V.J.; Kusek, G.; Snyder-Keller, A.; Schieweck, R.; Temple, S.; Kiebler, M.A. Staufen2 deficiency leads to impaired response to novelty in mice. Neurobiol. Learn. Mem. 2018, 150, 107–115. [Google Scholar] [CrossRef] [PubMed]
  234. Popova, V.V.; Kurshakova, M.M.; Kopytova, D.V. Methods to study the RNA-protein interactions. Mol. Biol. (Mosk.) 2015, 49, 472–481. [Google Scholar] [CrossRef]
  235. Dasti, A.; Cid-Samper, F.; Bechara, E.; Tartaglia, G.G. RNA-centric approaches to study RNA-protein interactions in vitro and in silico. Methods 2019. [Google Scholar] [CrossRef] [PubMed]
  236. Djordjevic, M. SELEX experiments: New prospects, applications and data analysis in inferring regulatory pathways. Biomol. Eng. 2007, 24, 179–189. [Google Scholar] [CrossRef]
  237. Zimmermann, B.; Bilusic, I.; Lorenz, C.; Schroeder, R. Genomic SELEX: A discovery tool for genomic aptamers. Methods 2010, 52, 125–132. [Google Scholar] [CrossRef]
  238. Darmostuk, M.; Rimpelova, S.; Gbelcova, H.; Ruml, T. Current approaches in SELEX: An update to aptamer selection technology. Biotechnol. Adv. 2015, 33, 1141–1161. [Google Scholar] [CrossRef]
  239. Ule, J.; Jensen, K.; Mele, A.; Darnell, R.B. CLIP: A method for identifying protein-RNA interaction sites in living cells. Methods 2005, 37, 376–386. [Google Scholar] [CrossRef]
  240. Milek, M.; Wyler, E.; Landthaler, M. Transcriptome-wide analysis of protein-RNA interactions using high-throughput sequencing. Semin. Cell Dev. Biol. 2012, 23, 206–212. [Google Scholar] [CrossRef]
  241. Darnell, J.C.; Mele, A.; Hung, K.Y.S.; Darnell, R.B. Mapping of In Vivo RNA-Binding Sites by Ultraviolet (UV)-Cross-Linking Immunoprecipitation (CLIP). Cold Spring Harb. Protoc. 2018, 2018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Rio, D.C. Electrophoretic mobility shift assays for RNA-protein complexes. Cold Spring Harb. Protoc. 2014, 2014, 435–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Daras, G.; Alatzas, A.; Tsitsekian, D.; Templalexis, D.; Rigas, S.; Hatzopoulos, P. Detection of RNA-protein interactions using a highly sensitive non-radioactive electrophoretic mobility shift assay. Electrophoresis 2019, 40, 1365–1371. [Google Scholar] [CrossRef] [PubMed]
  244. Lee, E.K.; Kim, H.H.; Kuwano, Y.; Abdelmohsen, K.; Srikantan, S.; Subaran, S.S.; Gleichmann, M.; Mughal, M.R.; Martindale, J.L.; Yang, X.; et al. hnRNP C promotes APP translation by competing with FMRP for APP mRNA recruitment to P bodies. Nat. Struct. Mol. Biol. 2010, 17, 732–739. [Google Scholar] [CrossRef] [Green Version]
  245. Zalfa, F.; Giorgi, M.; Primerano, B.; Moro, A.; Di Penta, A.; Reis, S.; Oostra, B.; Bagni, C. The fragile X syndrome protein FMRP associates with BC1 RNA and regulates the translation of specific mRNAs at synapses. Cell 2003, 112, 317–327. [Google Scholar] [CrossRef] [Green Version]
  246. Zalfa, F.; Adinolfi, S.; Napoli, I.; Kühn-Hölsken, E.; Urlaub, H.; Achsel, T.; Pastore, A.; Bagni, C. Fragile X mental retardation protein (FMRP) binds specifically to the brain cytoplasmic RNAs BC1/BC200 via a novel RNA-binding motif. J. Biol. Chem. 2005, 280, 33403–33410. [Google Scholar] [CrossRef] [Green Version]
  247. Iacoangeli, A.; Rozhdestvensky, T.S.; Dolzhanskaya, N.; Tournier, B.; Schütt, J.; Brosius, J.; Denman, R.B.; Khandjian, E.W.; Kindler, S.; Tiedge, H. On BC1 RNA and the fragile X mental retardation protein. Proc. Natl. Acad. Sci. USA 2008, 105, 734–739. [Google Scholar] [CrossRef] [Green Version]
  248. Louhivuori, V.; Vicario, A.; Uutela, M.; Rantamäki, T.; Louhivuori, L.M.; Castrén, E.; Tongiorgi, E.; Akerman, K.E.; Castrén, M.L. BDNF and TrkB in neuronal differentiation of Fmr1-knockout mouse. Neurobiol. Dis. 2011, 41, 469–480. [Google Scholar] [CrossRef]
  249. Gray, E.E.; Murphy, J.G.; Liu, Y.; Trang, I.; Tabor, G.T.; Lin, L.; Hoffman, D.A. Disruption of GpI mGluR-Dependent Cav2.3 Translation in a Mouse Model of Fragile X Syndrome. J. Neurosci. 2019, 39, 7453–7464. [Google Scholar] [CrossRef]
  250. Sudhakaran, I.P.; Hillebrand, J.; Dervan, A.; Das, S.; Holohan, E.E.; Hülsmeier, J.; Sarov, M.; Parker, R.; VijayRaghavan, K.; Ramaswami, M. FMRP and Ataxin-2 function together in long-term olfactory habituation and neuronal translational control. Proc. Natl. Acad. Sci. USA 2014, 111, E99–E108. [Google Scholar] [CrossRef] [Green Version]
  251. Kao, D.-I.; Aldridge, G.M.; Weiler, I.J.; Greenough, W.T. Altered mRNA transport, docking, and protein translation in neurons lacking fragile X mental retardation protein. Proc. Natl. Acad. Sci. USA 2010, 107, 15601–15606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  252. Daroles, L.; Gribaudo, S.; Doulazmi, M.; Scotto-Lomassese, S.; Dubacq, C.; Mandairon, N.; Greer, C.A.; Didier, A.; Trembleau, A.; Caillé, I. Fragile X Mental Retardation Protein and Dendritic Local Translation of the Alpha Subunit of the Calcium/Calmodulin-Dependent Kinase II Messenger RNA Are Required for the Structural Plasticity Underlying Olfactory Learning. Biol. Psychiatry 2016, 80, 149–159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. 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]
  254. Muddashetty, R.S.; Kelić, S.; Gross, C.; Xu, M.; Bassell, G.J. Dysregulated metabotropic glutamate receptor-dependent translation of AMPA receptor and postsynaptic density-95 mRNAs at synapses in a mouse model of fragile X syndrome. J. Neurosci. 2007, 27, 5338–5348. [Google Scholar] [CrossRef] [PubMed]
  255. Feuge, J.; Scharkowski, F.; Michaelsen-Preusse, K.; Korte, M. FMRP Modulates Activity-Dependent Spine Plasticity by Binding Cofilin1 mRNA and Regulating Localization and Local Translation. Cereb. Cortex 2019. [Google Scholar] [CrossRef]
  256. Estes, P.S.; O’Shea, M.; Clasen, S.; Zarnescu, D.C. Fragile X protein controls the efficacy of mRNA transport in Drosophila neurons. Mol. Cell. Neurosci. 2008, 39, 170–179. [Google Scholar] [CrossRef]
  257. Akins, M.R.; Berk-Rauch, H.E.; Kwan, K.Y.; Mitchell, M.E.; Shepard, K.A.; Korsak, L.I.T.; Stackpole, E.E.; Warner-Schmidt, J.L.; Sestan, N.; Cameron, H.A.; et al. Axonal ribosomes and mRNAs associate with fragile X granules in adult rodent and human brains. Hum. Mol. Genet. 2017, 26, 192–209. [Google Scholar] [CrossRef] [Green Version]
  258. 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]
  259. Brown, V.; Small, K.; Lakkis, L.; Feng, Y.; Gunter, C.; Wilkinson, K.D.; Warren, S.T. Purified recombinant Fmrp exhibits selective RNA binding as an intrinsic property of the fragile X mental retardation protein. J. Biol. Chem. 1998, 273, 15521–15527. [Google Scholar] [CrossRef] [Green Version]
  260. Antar, L.N.; Afroz, R.; Dictenberg, J.B.; Carroll, R.C.; Bassell, G.J. Metabotropic glutamate receptor activation regulates fragile x mental retardation protein and FMR1 mRNA localization differentially in dendrites and at synapses. J. Neurosci. 2004, 24, 2648–2655. [Google Scholar] [CrossRef]
  261. Sabanov, V.; Braat, S.; D’Andrea, L.; Willemsen, R.; Zeidler, S.; Rooms, L.; Bagni, C.; Kooy, R.F.; Balschun, D. Impaired GABAergic inhibition in the hippocampus of Fmr1 knockout mice. Neuropharmacology 2017, 116, 71–81. [Google Scholar] [CrossRef]
  262. Maurin, T.; Melko, M.; Abekhoukh, S.; Khalfallah, O.; Davidovic, L.; Jarjat, M.; D’Antoni, S.; Catania, M.V.; Moine, H.; Bechara, E.; et al. The FMRP/GRK4 mRNA interaction uncovers a new mode of binding of the Fragile X mental retardation protein in cerebellum. Nucleic Acids Res. 2015, 43, 8540–8550. [Google Scholar] [CrossRef] [Green Version]
  263. Strumbos, J.G.; Brown, M.R.; Kronengold, J.; Polley, D.B.; Kaczmarek, L.K. Fragile X mental retardation protein is required for rapid experience-dependent regulation of the potassium channel Kv3.1b. J. Neurosci. 2010, 30, 10263–10271. [Google Scholar] [CrossRef] [PubMed]
  264. Gross, C.; Yao, X.; Pong, D.L.; Jeromin, A.; Bassell, G.J. Fragile X mental retardation protein regulates protein expression and mRNA translation of the potassium channel Kv4.2. J. Neurosci. 2011, 31, 5693–5698. [Google Scholar] [CrossRef] [PubMed]
  265. Ceolin, L.; Bouquier, N.; Vitre-Boubaker, J.; Rialle, S.; Severac, D.; Valjent, E.; Perroy, J.; Puighermanal, E. Cell Type-Specific mRNA Dysregulation in Hippocampal CA1 Pyramidal Neurons of the Fragile X Syndrome Mouse Model. Front. Mol. Neurosci. 2017, 10, 340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. Lu, R.; Wang, H.; Liang, Z.; Ku, L.; O’donnell, W.T.; Li, W.; Warren, S.T.; Feng, Y. The fragile X protein controls microtubule-associated protein 1B translation and microtubule stability in brain neuron development. Proc. Natl. Acad. Sci. USA 2004, 101, 15201–15206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Wei, Z.-X.; Yi, Y.-H.; Sun, W.-W.; Wang, R.; Su, T.; Bai, Y.-J.; Liao, W.-P. Expression changes of microtubule associated protein 1B in the brain of Fmr1 knockout mice. Neurosci. Bull. 2007, 23, 203–208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. Janusz, A.; Milek, J.; Perycz, M.; Pacini, L.; Bagni, C.; Kaczmarek, L.; Dziembowska, M. The Fragile X mental retardation protein regulates matrix metalloproteinase 9 mRNA at synapses. J. Neurosci. 2013, 33, 18234–18241. [Google Scholar] [CrossRef] [Green Version]
  269. Jeon, S.J.; Kim, J.-W.; Kim, K.C.; Han, S.M.; Go, H.S.; Seo, J.E.; Choi, C.S.; Ryu, J.H.; Shin, C.Y.; Song, M.-R. Translational regulation of NeuroD1 expression by FMRP: Involvement in glutamatergic neuronal differentiation of cultured rat primary neural progenitor cells. Cell. Mol. Neurobiol. 2014, 34, 297–305. [Google Scholar] [CrossRef]
  270. Chmielewska, J.J.; Kuzniewska, B.; Milek, J.; Urbanska, K.; Dziembowska, M. Neuroligin 1, 2, and 3 Regulation at the Synapse: FMRP-Dependent Translation and Activity-Induced Proteolytic Cleavage. Mol. Neurobiol. 2019, 56, 2741–2759. [Google Scholar] [CrossRef] [Green Version]
  271. Kwan, K.Y.; Lam, M.M.S.; Johnson, M.B.; Dube, U.; Shim, S.; Rašin, M.-R.; Sousa, A.M.M.; Fertuzinhos, S.; Chen, J.-G.; Arellano, J.I.; et al. Species-dependent posttranscriptional regulation of NOS1 by FMRP in the developing cerebral cortex. Cell 2012, 149, 899–911. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  272. Zhang, M.; Wang, Q.; Huang, Y. Fragile X mental retardation protein FMRP and the RNA export factor NXF2 associate with and destabilize Nxf1 mRNA in neuronal cells. Proc. Natl. Acad. Sci. USA 2007, 104, 10057–10062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Maurin, T.; Lebrigand, K.; Castagnola, S.; Paquet, A.; Jarjat, M.; Popa, A.; Grossi, M.; Rage, F.; Bardoni, B. HITS-CLIP in various brain areas reveals new targets and new modalities of RNA binding by fragile X mental retardation protein. Nucleic Acids Res. 2018, 46, 6344–6355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. Majumder, P.; Chu, J.-F.; Chatterjee, B.; Swamy, K.B.S.; Shen, C.-K.J. Co-regulation of mRNA translation by TDP-43 and Fragile X Syndrome protein FMRP. Acta Neuropathol. 2016, 132, 721–738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Tervonen, T.; Akerman, K.; Oostra, B.A.; Castrén, M. Rgs4 mRNA expression is decreased in the brain of Fmr1 knockout mouse. Brain Res. Mol. Brain Res. 2005, 133, 162–165. [Google Scholar] [CrossRef] [PubMed]
  276. Bechara, E.G.; Didiot, M.C.; Melko, M.; Davidovic, L.; Bensaid, M.; Martin, P.; Castets, M.; Pognonec, P.; Khandjian, E.W.; Moine, H.; et al. A novel function for fragile X mental retardation protein in translational activation. PLoS Biol. 2009, 7, e16. [Google Scholar] [CrossRef] [Green Version]
  277. Paul, A.; Nawalpuri, B.; Shah, D.; Sateesh, S.; Muddashetty, R.S.; Clement, J.P. Differential Regulation of Syngap1 Translation by FMRP Modulates eEF2 Mediated Response on NMDAR Activity. Front. Mol. Neurosci. 2019, 12, 97. [Google Scholar] [CrossRef] [Green Version]
  278. Higashimori, H.; Morel, L.; Huth, J.; Lindemann, L.; Dulla, C.; Taylor, A.; Freeman, M.; Yang, Y. Astroglial FMRP-dependent translational down-regulation of mGluR5 underlies glutamate transporter GLT1 dysregulation in the fragile X mouse. Hum. Mol. Genet. 2013, 22, 2041–2054. [Google Scholar] [CrossRef] [Green Version]
  279. Wang, H.; Ku, L.; Osterhout, D.J.; Li, W.; Ahmadian, A.; Liang, Z.; Feng, Y. Developmentally-programmed FMRP expression in oligodendrocytes: A potential role of FMRP in regulating translation in oligodendroglia progenitors. Hum. Mol. Genet. 2004, 13, 79–89. [Google Scholar] [CrossRef] [Green Version]
  280. Rackham, O.; Brown, C.M. Visualization of RNA-protein interactions in living cells: FMRP and IMP1 interact on mRNAs. EMBO J. 2004, 23, 3346–3355. [Google Scholar] [CrossRef]
  281. Fähling, M.; Mrowka, R.; Steege, A.; Kirschner, K.M.; Benko, E.; Förstera, B.; Persson, P.B.; Thiele, B.J.; Meier, J.C.; Scholz, H. Translational regulation of the human achaete-scute homologue-1 by fragile X mental retardation protein. J. Biol. Chem. 2009, 284, 4255–4266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  282. Hoeffer, C.A.; Sanchez, E.; Hagerman, R.J.; Mu, Y.; Nguyen, D.V.; Wong, H.; Whelan, A.M.; Zukin, R.S.; Klann, E.; Tassone, F. Altered mTOR signaling and enhanced CYFIP2 expression levels in subjects with fragile X syndrome. Genes Brain Behav. 2012, 11, 332–341. [Google Scholar] [CrossRef] [PubMed]
  283. Sung, Y.J.; Dolzhanskaya, N.; Nolin, S.L.; Brown, T.; Currie, J.R.; Denman, R.B. The fragile X mental retardation protein FMRP binds elongation factor 1A mRNA and negatively regulates its translation in vivo. J. Biol. Chem. 2003, 278, 15669–15678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Li, Z.; Zhang, Y.; Ku, L.; Wilkinson, K.D.; Warren, S.T.; Feng, Y. The fragile X mental retardation protein inhibits translation via interacting with mRNA. Nucleic Acids Res. 2001, 29, 2276–2283. [Google Scholar] [CrossRef] [Green Version]
  285. Castets, M.; Schaeffer, C.; Bechara, E.; Schenck, A.; Khandjian, E.W.; Luche, S.; Moine, H.; Rabilloud, T.; Mandel, J.-L.; Bardoni, B. FMRP interferes with the Rac1 pathway and controls actin cytoskeleton dynamics in murine fibroblasts. Hum. Mol. Genet. 2005, 14, 835–844. [Google Scholar] [CrossRef] [Green Version]
  286. 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]
  287. Ascano, M.; Mukherjee, N.; Bandaru, P.; Miller, J.B.; Nusbaum, J.D.; Corcoran, D.L.; Langlois, C.; Munschauer, M.; Dewell, S.; Hafner, M.; et al. FMRP targets distinct mRNA sequence elements to regulate protein expression. Nature 2012, 492, 382–386. [Google Scholar] [CrossRef]
  288. Darnell, J.C.; Van Driesche, S.J.; Zhang, C.; Hung, K.Y.S.; 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]
  289. Brown, V.; Jin, P.; Ceman, S.; Darnell, J.C.; O’Donnell, W.T.; Tenenbaum, S.A.; Jin, X.; Feng, Y.; Wilkinson, K.D.; Keene, J.D.; et al. Microarray identification of FMRP-associated brain mRNAs and altered mRNA translational profiles in fragile X syndrome. Cell 2001, 107, 477–487. [Google Scholar] [CrossRef] [Green Version]
  290. Das Sharma, S.; Metz, J.B.; Li, H.; Hobson, B.D.; Hornstein, N.; Sulzer, D.; Tang, G.; Sims, P.A. Widespread Alterations in Translation Elongation in the Brain of Juvenile Fmr1 Knockout Mice. Cell Rep. 2019, 26, 3313–3322.e5. [Google Scholar] [CrossRef] [Green Version]
  291. Liu, B.; Li, Y.; Stackpole, E.E.; Novak, A.; Gao, Y.; Zhao, Y.; Zhao, X.; Richter, J.D. Regulatory discrimination of mRNAs by FMRP controls mouse adult neural stem cell differentiation. Proc. Natl. Acad. Sci. USA 2018, 115, E11397–E11405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  292. Xu, S.; Poidevin, M.; Han, E.; Bi, J.; Jin, P. Circadian rhythm-dependent alterations of gene expression in Drosophila brain lacking fragile X mental retardation protein. PLoS ONE 2012, 7, e37937. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Bittel, D.C.; Kibiryeva, N.; Butler, M.G. Whole genome microarray analysis of gene expression in subjects with fragile X syndrome. Genet. Med. 2007, 9, 464–472. [Google Scholar] [CrossRef] [Green Version]
  294. Stefanovic, S.; DeMarco, B.A.; Underwood, A.; Williams, K.R.; Bassell, G.J.; Mihailescu, M.R. Fragile X mental retardation protein interactions with a G quadruplex structure in the 3’-untranslated region of NR2B mRNA. Mol. Biosyst. 2015, 11, 3222–3230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  295. Menon, L.; Mader, S.A.; Mihailescu, M.-R. Fragile X mental retardation protein interactions with the microtubule associated protein 1B RNA. RNA 2008, 14, 1644–1655. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Menon, L.; Mihailescu, M.-R. Interactions of the G quartet forming semaphorin 3F RNA with the RGG box domain of the fragile X protein family. Nucleic Acids Res. 2007, 35, 5379–5392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  297. Zhang, Y.; Gaetano, C.M.; Williams, K.R.; Bassell, G.J.; Mihailescu, M.R. FMRP interacts with G-quadruplex structures in the 3’-UTR of its dendritic target Shank1 mRNA. RNA Biol. 2014, 11, 1364–1374. [Google Scholar] [CrossRef] [Green Version]
  298. McAninch, D.S.; Heinaman, A.M.; Lang, C.N.; Moss, K.R.; Bassell, G.J.; Rita Mihailescu, M.; Evans, T.L. Fragile X mental retardation protein recognizes a G quadruplex structure within the survival motor neuron domain containing 1 mRNA 5’-UTR. Mol. Biosyst. 2017, 13, 1448–1457. [Google Scholar] [CrossRef]
  299. 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]
  300. Ramos, A.; Hollingworth, D.; Pastore, A. G-quartet-dependent recognition between the FMRP RGG box and RNA. RNA 2003, 9, 1198–1207. [Google Scholar] [CrossRef] [Green Version]
  301. Suhl, J.A.; Chopra, P.; Anderson, B.R.; Bassell, G.J.; Warren, S.T. Analysis of FMRP mRNA target datasets reveals highly associated mRNAs mediated by G-quadruplex structures formed via clustered WGGA sequences. Hum. Mol. Genet. 2014, 23, 5479–5491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  302. Vasilyev, N.; Polonskaia, A.; Darnell, J.C.; Darnell, R.B.; Patel, D.J.; Serganov, A. Crystal structure reveals specific recognition of a G-quadruplex RNA by a β-turn in the RGG motif of FMRP. Proc. Natl. Acad. Sci. USA 2015, 112, E5391–E5400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. 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]
  304. Zhang, F.; Kang, Y.; Wang, M.; Li, Y.; Xu, T.; Yang, W.; Song, H.; Wu, H.; Shu, Q.; Jin, P. Fragile X mental retardation protein modulates the stability of its m6A-marked messenger RNA targets. Hum. Mol. Genet. 2018, 27, 3936–3950. [Google Scholar] [CrossRef]
  305. McEvoy, M.; Cao, G.; Montero Llopis, P.; Kundel, M.; Jones, K.; Hofler, C.; Shin, C.; Wells, D.G. Cytoplasmic polyadenylation element binding protein 1-mediated mRNA translation in Purkinje neurons is required for cerebellar long-term depression and motor coordination. J. Neurosci. 2007, 27, 6400–6411. [Google Scholar] [CrossRef] [Green Version]
  306. Oe, S.; Yoneda, Y. Cytoplasmic polyadenylation element-like sequences are involved in dendritic targeting of BDNF mRNA in hippocampal neurons. FEBS Lett. 2010, 584, 3424–3430. [Google Scholar] [CrossRef] [Green Version]
  307. Kim, K.C.; Oh, W.J.; Ko, K.H.; Shin, C.Y.; Wells, D.G. Cyclin B1 expression regulated by cytoplasmic polyadenylation element binding protein in astrocytes. J. Neurosci. 2011, 31, 12118–12128. [Google Scholar] [CrossRef]
  308. Kundel, M.; Jones, K.J.; Shin, C.Y.; Wells, D.G. Cytoplasmic polyadenylation element-binding protein regulates neurotrophin-3-dependent beta-catenin mRNA translation in developing hippocampal neurons. J. Neurosci. 2009, 29, 13630–13639. [Google Scholar] [CrossRef] [Green Version]
  309. Alves-Sampaio, A.; Troca-Marín, J.A.; Montesinos, M.L. NMDA-mediated regulation of DSCAM dendritic local translation is lost in a mouse model of Down’s syndrome. J. Neurosci. 2010, 30, 13537–13548. [Google Scholar] [CrossRef]
  310. Gerstner, J.R.; Vanderheyden, W.M.; LaVaute, T.; Westmark, C.J.; Rouhana, L.; Pack, A.I.; Wickens, M.; Landry, C.F. Time of day regulates subcellular trafficking, tripartite synaptic localization, and polyadenylation of the astrocytic Fabp7 mRNA. J. Neurosci. 2012, 32, 1383–1394. [Google Scholar] [CrossRef] [Green Version]
  311. Zearfoss, N.R.; Alarcon, J.M.; Trifilieff, P.; Kandel, E.; Richter, J.D. A molecular circuit composed of CPEB-1 and c-Jun controls growth hormone-mediated synaptic plasticity in the mouse hippocampus. J. Neurosci. 2008, 28, 8502–8509. [Google Scholar] [CrossRef] [PubMed]
  312. Liu, J.; Hu, J.-Y.; Wu, F.; Schwartz, J.H.; Schacher, S. Two mRNA-binding proteins regulate the distribution of syntaxin mRNA in Aplysia sensory neurons. J. Neurosci. 2006, 26, 5204–5214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  313. Kochanek, D.M.; Wells, D.G. CPEB1 regulates the expression of MTDH/AEG-1 and glioblastoma cell migration. Mol. Cancer Res. 2013, 11, 149–160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  314. Bava, F.-A.; Eliscovich, C.; Ferreira, P.G.; Miñana, B.; Ben-Dov, C.; Guigó, R.; Valcárcel, J.; Méndez, R. CPEB1 coordinates alternative 3’-UTR formation with translational regulation. Nature 2013, 495, 121–125. [Google Scholar] [CrossRef]
  315. Nishimura, Y.; Kano, K.; Naito, K. Porcine CPEB1 is involved in Cyclin B translation and meiotic resumption in porcine oocytes. Anim. Sci. J. 2010, 81, 444–452. [Google Scholar] [CrossRef]
  316. Han, S.J.; Martins, J.P.S.; Yang, Y.; Kang, M.K.; Daldello, E.M.; Conti, M. The Translation of Cyclin B1 and B2 is Differentially Regulated during Mouse Oocyte Reentry into the Meiotic Cell Cycle. Sci. Rep. 2017, 7, 14077. [Google Scholar] [CrossRef] [Green Version]
  317. Sousa Martins, J.P.; Liu, X.; Oke, A.; Arora, R.; Franciosi, F.; Viville, S.; Laird, D.J.; Fung, J.C.; Conti, M. DAZL and CPEB1 regulate mRNA translation synergistically during oocyte maturation. J. Cell. Sci. 2016, 129, 1271–1282. [Google Scholar] [CrossRef] [Green Version]
  318. Nagaoka, K.; Fujii, K.; Zhang, H.; Usuda, K.; Watanabe, G.; Ivshina, M.; Richter, J.D. CPEB1 mediates epithelial-to-mesenchyme transition and breast cancer metastasis. Oncogene 2016, 35, 2893–2901. [Google Scholar] [CrossRef] [Green Version]
  319. Galardi, S.; Petretich, M.; Pinna, G.; D’Amico, S.; Loreni, F.; Michienzi, A.; Groisman, I.; Ciafrè, S.A. CPEB1 restrains proliferation of Glioblastoma cells through the regulation of p27(Kip1) mRNA translation. Sci. Rep. 2016, 6, 25219. [Google Scholar] [CrossRef] [Green Version]
  320. Jones, K.J.; Korb, E.; Kundel, M.A.; Kochanek, A.R.; Kabraji, S.; McEvoy, M.; Shin, C.Y.; Wells, D.G. CPEB1 regulates beta-catenin mRNA translation and cell migration in astrocytes. Glia 2008, 56, 1401–1413. [Google Scholar] [CrossRef] [Green Version]
  321. Rutledge, C.E.; Lau, H.-T.; Mangan, H.; Hardy, L.L.; Sunnotel, O.; Guo, F.; MacNicol, A.M.; Walsh, C.P.; Lees-Murdock, D.J. Efficient translation of Dnmt1 requires cytoplasmic polyadenylation and Musashi binding elements. PLoS ONE 2014, 9, e88385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  322. Hägele, S.; Kühn, U.; Böning, M.; Katschinski, D.M. Cytoplasmic polyadenylation-element-binding protein (CPEB)1 and 2 bind to the HIF-1alpha mRNA 3’-UTR and modulate HIF-1alpha protein expression. Biochem. J. 2009, 417, 235–246. [Google Scholar] [CrossRef] [PubMed]
  323. Caldeira, J.; Simões-Correia, J.; Paredes, J.; Pinto, M.T.; Sousa, S.; Corso, G.; Marrelli, D.; Roviello, F.; Pereira, P.S.; Weil, D.; et al. CPEB1, a novel gene silenced in gastric cancer: A Drosophila approach. Gut 2012, 61, 1115–1123. [Google Scholar] [CrossRef] [PubMed]
  324. Takahashi, K.; Yi, H.; Gu, J.; Ikegami, D.; Liu, S.; Iida, T.; Kashiwagi, Y.; Dong, C.; Kunisawa, T.; Hao, S. The mitochondrial calcium uniporter contributes to morphine tolerance through pCREB and CPEB1 in rat spinal cord dorsal horn. Br. J. Anaesth. 2019, 123, e226–e238. [Google Scholar] [CrossRef]
  325. Xu, M.; Fang, S.; Song, J.; Chen, M.; Zhang, Q.; Weng, Q.; Fan, X.; Chen, W.; Wu, X.; Wu, F.; et al. CPEB1 mediates hepatocellular carcinoma cancer stemness and chemoresistance. Cell Death Dis. 2018, 9, 957. [Google Scholar] [CrossRef] [Green Version]
  326. Shin, J.; Paek, K.Y.; Ivshina, M.; Stackpole, E.E.; Richter, J.D. Essential role for non-canonical poly(A) polymerase GLD4 in cytoplasmic polyadenylation and carbohydrate metabolism. Nucleic Acids Res. 2017, 45, 6793–6804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  327. Tay, J.; Richter, J.D. Germ cell differentiation and synaptonemal complex formation are disrupted in CPEB knockout mice. Dev. Cell 2001, 1, 201–213. [Google Scholar] [CrossRef] [Green Version]
  328. Nairismägi, M.-L.; Vislovukh, A.; Meng, Q.; Kratassiouk, G.; Beldiman, C.; Petretich, M.; Groisman, R.; Füchtbauer, E.-M.; Harel-Bellan, A.; Groisman, I. Translational control of TWIST1 expression in MCF-10A cell lines recapitulating breast cancer progression. Oncogene 2012, 31, 4960–4966. [Google Scholar] [CrossRef] [Green Version]
  329. Yin, J.; Park, G.; Lee, J.E.; Park, J.Y.; Kim, T.-H.; Kim, Y.-J.; Lee, S.-H.; Yoo, H.; Kim, J.H.; Park, J.B. CPEB1 modulates differentiation of glioma stem cells via downregulation of HES1 and SIRT1 expression. Oncotarget 2014, 5, 6756–6769. [Google Scholar] [CrossRef] [Green Version]
  330. Alexandrov, I.M.; Ivshina, M.; Jung, D.Y.; Friedline, R.; Ko, H.J.; Xu, M.; O’Sullivan-Murphy, B.; Bortell, R.; Huang, Y.-T.; Urano, F.; et al. Cytoplasmic polyadenylation element binding protein deficiency stimulates PTEN and Stat3 mRNA translation and induces hepatic insulin resistance. PLoS Genet. 2012, 8, e1002457. [Google Scholar] [CrossRef] [Green Version]
  331. Zhang, J.-H.; Panicker, L.M.; Seigneur, E.M.; Lin, L.; House, C.D.; Morgan, W.; Chen, W.C.; Mehta, H.; Haj-Ali, M.; Yu, Z.-X.; et al. Cytoplasmic polyadenylation element binding protein is a conserved target of tumor suppressor HRPT2/CDC73. Cell Death Differ. 2010, 17, 1551–1565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  332. Udagawa, T.; Farny, N.G.; Jakovcevski, M.; Kaphzan, H.; Alarcon, J.M.; Anilkumar, S.; Ivshina, M.; Hurt, J.A.; Nagaoka, K.; Nalavadi, V.C.; et al. Genetic and acute CPEB1 depletion ameliorate fragile X pathophysiology. Nat. Med. 2013, 19, 1473–1477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  333. Charlesworth, A.; Meijer, H.A.; de Moor, C.H. Specificity factors in cytoplasmic polyadenylation. Wiley Interdiscip. Rev. RNA 2013, 4, 437–461. [Google Scholar] [CrossRef] [PubMed]
  334. Stepien, B.K.; Oppitz, C.; Gerlach, D.; Dag, U.; Novatchkova, M.; Krüttner, S.; Stark, A.; Keleman, K. RNA-binding profiles of Drosophila CPEB proteins Orb and Orb2. Proc. Natl. Acad. Sci. USA 2016, 113, E7030–E7038. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  335. Antar, L.N.; Dictenberg, J.B.; Plociniak, M.; Afroz, R.; Bassell, G.J. Localization of FMRP-associated mRNA granules and requirement of microtubules for activity-dependent trafficking in hippocampal neurons. Genes Brain Behav. 2005, 4, 350–359. [Google Scholar] [CrossRef] [PubMed]
  336. El Fatimy, R.; Davidovic, L.; Tremblay, S.; Jaglin, X.; Dury, A.; Robert, C.; De Koninck, P.; Khandjian, E.W. Tracking the Fragile X Mental Retardation Protein in a Highly Ordered Neuronal RiboNucleoParticles Population: A Link between Stalled Polyribosomes and RNA Granules. PLoS Genet. 2016, 12, e1006192. [Google Scholar] [CrossRef] [PubMed]
  337. Wu, L.; Wells, D.; Tay, J.; Mendis, D.; Abbott, M.A.; Barnitt, A.; Quinlan, E.; Heynen, A.; Fallon, J.R.; Richter, J.D. CPEB-mediated cytoplasmic polyadenylation and the regulation of experience-dependent translation of alpha-CaMKII mRNA at synapses. Neuron 1998, 21, 1129–1139. [Google Scholar] [CrossRef] [Green Version]
  338. Huang, Y.-S.; Jung, M.-Y.; Sarkissian, M.; Richter, J.D. N-methyl-D-aspartate receptor signaling results in Aurora kinase-catalyzed CPEB phosphorylation and alpha CaMKII mRNA polyadenylation at synapses. EMBO J. 2002, 21, 2139–2148. [Google Scholar] [CrossRef] [Green Version]
  339. Shiina, N.; Shinkura, K.; Tokunaga, M. A novel RNA-binding protein in neuronal RNA granules: Regulatory machinery for local translation. J. Neurosci. 2005, 25, 4420–4434. [Google Scholar] [CrossRef] [Green Version]
  340. Solomon, S.; Xu, Y.; Wang, B.; David, M.D.; Schubert, P.; Kennedy, D.; Schrader, J.W. Distinct structural features of caprin-1 mediate its interaction with G3BP-1 and its induction of phosphorylation of eukaryotic translation initiation factor 2alpha, entry to cytoplasmic stress granules, and selective interaction with a subset of mRNAs. Mol. Cell. Biol. 2007, 27, 2324–2342. [Google Scholar] [CrossRef] [Green Version]
  341. Arguello, A.E.; DeLiberto, A.N.; Kleiner, R.E. RNA Chemical Proteomics Reveals the N6-Methyladenosine (m6A)-Regulated Protein-RNA Interactome. J. Am. Chem. Soc. 2017, 139, 17249–17252. [Google Scholar] [CrossRef] [PubMed]
  342. Friedersdorf, M.B.; Keene, J.D. Advancing the functional utility of PAR-CLIP by quantifying background binding to mRNAs and lncRNAs. Genome Biol. 2014, 15, R2. [Google Scholar] [CrossRef] [PubMed]
  343. Shiina, N. Liquid- and solid-like RNA granules form through specific scaffold proteins and combine into biphasic granules. J. Biol. Chem. 2019, 294, 3532–3548. [Google Scholar] [CrossRef] [Green Version]
  344. Kim, Y.K.; Furic, L.; Desgroseillers, L.; Maquat, L.E. Mammalian Staufen1 recruits Upf1 to specific mRNA 3’UTRs so as to elicit mRNA decay. Cell 2005, 120, 195–208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  345. Ferrandon, D.; Elphick, L.; Nüsslein-Volhard, C.; St Johnston, D. Staufen protein associates with the 3’UTR of bicoid mRNA to form particles that move in a microtubule-dependent manner. Cell 1994, 79, 1221–1232. [Google Scholar] [CrossRef]
  346. Gardiol, A.; St Johnston, D. Staufen targets coracle mRNA to Drosophila neuromuscular junctions and regulates GluRIIA synaptic accumulation and bouton number. Dev. Biol. 2014, 392, 153–167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  347. Roussis, I.M.; Guille, M.; Myers, F.A.; Scarlett, G.P. RNA Whole-Mount In situ Hybridisation Proximity Ligation Assay (rISH-PLA), an Assay for Detecting RNA-Protein Complexes in Intact Cells. PLoS ONE 2016, 11, e0147967. [Google Scholar] [CrossRef] [Green Version]
  348. Gong, C.; Kim, Y.K.; Woeller, C.F.; Tang, Y.; Maquat, L.E. SMD and NMD are competitive pathways that contribute to myogenesis: Effects on PAX3 and myogenin mRNAs. Genes Dev. 2009, 23, 54–66. [Google Scholar] [CrossRef] [Green Version]
  349. Broadus, J.; Fuerstenberg, S.; Doe, C.Q. Staufen-dependent localization of prospero mRNA contributes to neuroblast daughter-cell fate. Nature 1998, 391, 792–795. [Google Scholar] [CrossRef]
  350. Li, P.; Yang, X.; Wasser, M.; Cai, Y.; Chia, W. Inscuteable and Staufen mediate asymmetric localization and segregation of prospero RNA during Drosophila neuroblast cell divisions. Cell 1997, 90, 437–447. [Google Scholar] [CrossRef] [Green Version]
  351. Heber, S.; Gáspár, I.; Tants, J.-N.; Günther, J.; Moya, S.M.F.; Janowski, R.; Ephrussi, A.; Sattler, M.; Niessing, D. Staufen2-mediated RNA recognition and localization requires combinatorial action of multiple domains. Nat. Commun. 2019, 10, 1659. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  352. Lucas, B.A.; Lavi, E.; Shiue, L.; Cho, H.; Katzman, S.; Miyoshi, K.; Siomi, M.C.; Carmel, L.; Ares, M.; Maquat, L.E. Evidence for convergent evolution of SINE-directed Staufen-mediated mRNA decay. Proc. Natl. Acad. Sci. USA 2018, 115, 968–973. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  353. Sugimoto, Y.; Vigilante, A.; Darbo, E.; Zirra, A.; Militti, C.; D’Ambrogio, A.; Luscombe, N.M.; Ule, J. hiCLIP reveals the in vivo atlas of mRNA secondary structures recognized by Staufen 1. Nature 2015, 519, 491–494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  354. Kim, Y.K.; Furic, L.; Parisien, M.; Major, F.; DesGroseillers, L.; Maquat, L.E. Staufen1 regulates diverse classes of mammalian transcripts. EMBO J. 2007, 26, 2670–2681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  355. Heraud-Farlow, J.E.; Sharangdhar, T.; Li, X.; Pfeifer, P.; Tauber, S.; Orozco, D.; Hörmann, A.; Thomas, S.; Bakosova, A.; Farlow, A.R.; et al. Staufen2 regulates neuronal target RNAs. Cell Rep. 2013, 5, 1511–1518. [Google Scholar] [CrossRef] [Green Version]
  356. Laver, J.D.; Li, X.; Ancevicius, K.; Westwood, J.T.; Smibert, C.A.; Morris, Q.D.; Lipshitz, H.D. Genome-wide analysis of Staufen-associated mRNAs identifies secondary structures that confer target specificity. Nucleic Acids Res. 2013, 41, 9438–9460. [Google Scholar] [CrossRef] [Green Version]
  357. LeGendre, J.B.; Campbell, Z.T.; Kroll-Conner, P.; Anderson, P.; Kimble, J.; Wickens, M. RNA targets and specificity of Staufen, a double-stranded RNA-binding protein in Caenorhabditis elegans. J. Biol. Chem. 2013, 288, 2532–2545. [Google Scholar] [CrossRef] [Green Version]
  358. de Lucas, S.; Oliveros, J.C.; Chagoyen, M.; Ortín, J. Functional signature for the recognition of specific target mRNAs by human Staufen1 protein. Nucleic Acids Res. 2014, 42, 4516–4526. [Google Scholar] [CrossRef]
  359. Furic, L.; Maher-Laporte, M.; DesGroseillers, L. A genome-wide approach identifies distinct but overlapping subsets of cellular mRNAs associated with Staufen1- and Staufen2-containing ribonucleoprotein complexes. RNA 2008, 14, 324–335. [Google Scholar] [CrossRef] [Green Version]
  360. Sharangdhar, T.; Sugimoto, Y.; Heraud-Farlow, J.; Fernández-Moya, S.M.; Ehses, J.; Ruiz de Los Mozos, I.; Ule, J.; Kiebler, M.A. A retained intron in the 3’-UTR of Calm3 mRNA mediates its Staufen2- and activity-dependent localization to neuronal dendrites. EMBO Rep. 2017, 18, 1762–1774. [Google Scholar] [CrossRef]
  361. Duchaîne, T.F.; Hemraj, I.; Furic, L.; Deitinghoff, A.; Kiebler, M.A.; DesGroseillers, L. Staufen2 isoforms localize to the somatodendritic domain of neurons and interact with different organelles. J. Cell. Sci. 2002, 115, 3285–3295. [Google Scholar] [PubMed]
  362. Sahoo, P.K.; Lee, S.J.; Jaiswal, P.B.; Alber, S.; Kar, A.N.; Miller-Randolph, S.; Taylor, E.E.; Smith, T.; Singh, B.; Ho, T.S.-Y.; et al. Axonal G3BP1 stress granule protein limits axonal mRNA translation and nerve regeneration. Nat. Commun. 2018, 9, 3358. [Google Scholar] [CrossRef] [PubMed]
  363. Atlas, R.; Behar, L.; Elliott, E.; Ginzburg, I. The insulin-like growth factor mRNA binding-protein IMP-1 and the Ras-regulatory protein G3BP associate with tau mRNA and HuD protein in differentiated P19 neuronal cells. J. Neurochem. 2004, 89, 613–626. [Google Scholar] [CrossRef] [PubMed]
  364. Ortega, A.D.; Willers, I.M.; Sala, S.; Cuezva, J.M. Human G3BP1 interacts with beta-F1-ATPase mRNA and inhibits its translation. J. Cell. Sci. 2010, 123, 2685–2696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  365. Taniuchi, K.; Nishimori, I.; Hollingsworth, M.A. Intracellular CD24 inhibits cell invasion by posttranscriptional regulation of BART through interaction with G3BP. Cancer Res. 2011, 71, 895–905. [Google Scholar] [CrossRef] [Green Version]
  366. Bikkavilli, R.K.; Malbon, C.C. Arginine methylation of G3BP1 in response to Wnt3a regulates β-catenin mRNA. J. Cell. Sci. 2011, 124, 2310–2320. [Google Scholar] [CrossRef] [Green Version]
  367. Lypowy, J.; Chen, I.-Y.; Abdellatif, M. An alliance between Ras GTPase-activating protein, filamin C, and Ras GTPase-activating protein SH3 domain-binding protein regulates myocyte growth. J. Biol. Chem. 2005, 280, 25717–25728. [Google Scholar] [CrossRef] [Green Version]
  368. Gallouzi, I.E.; Parker, F.; Chebli, K.; Maurier, F.; Labourier, E.; Barlat, I.; Capony, J.P.; Tocque, B.; Tazi, J. A novel phosphorylation-dependent RNase activity of GAP-SH3 binding protein: A potential link between signal transduction and RNA stability. Mol. Cell. Biol. 1998, 18, 3956–3965. [Google Scholar] [CrossRef] [Green Version]
  369. Winslow, S.; Leandersson, K.; Larsson, C. Regulation of PMP22 mRNA by G3BP1 affects cell proliferation in breast cancer cells. Mol. Cancer 2013, 12, 156. [Google Scholar] [CrossRef] [Green Version]
  370. Zekri, L.; Chebli, K.; Tourrière, H.; Nielsen, F.C.; Hansen, T.V.O.; Rami, A.; Tazi, J. Control of fetal growth and neonatal survival by the RasGAP-associated endoribonuclease G3BP. Mol. Cell. Biol. 2005, 25, 8703–8716. [Google Scholar] [CrossRef] [Green Version]
  371. Taniuchi, K.; Nishimori, I.; Hollingsworth, M.A. The N-terminal domain of G3BP enhances cell motility and invasion by posttranscriptional regulation of BART. Mol. Cancer Res. 2011, 9, 856–866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  372. Tourrière, H.; Gallouzi, I.E.; Chebli, K.; Capony, J.P.; Mouaikel, J.; van der Geer, P.; Tazi, J. RasGAP-associated endoribonuclease G3Bp: Selective RNA degradation and phosphorylation-dependent localization. Mol. Cell. Biol. 2001, 21, 7747–7760. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  373. Martin, S.; Tazi, J. Visualization of G3BP stress granules dynamics in live primary cells. J. Vis. Exp. 2014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  374. Chen, Y.; Cohen, T.J. Aggregation of the nucleic acid-binding protein TDP-43 occurs via distinct routes that are coordinated with stress granule formation. J. Biol. Chem. 2019, 294, 3696–3706. [Google Scholar] [CrossRef] [Green Version]
  375. Weber, S.C.; Brangwynne, C.P. Getting RNA and protein in phase. Cell 2012, 149, 1188–1191. [Google Scholar] [CrossRef] [Green Version]
  376. Lin, Y.; Protter, D.S.W.; Rosen, M.K.; Parker, R. Formation and Maturation of Phase-Separated Liquid Droplets by RNA-Binding Proteins. Mol. Cell 2015, 60, 208–219. [Google Scholar] [CrossRef] [Green Version]
  377. Shin, Y.; Brangwynne, C.P. Liquid phase condensation in cell physiology and disease. Science 2017, 357. [Google Scholar] [CrossRef] [Green Version]
  378. Polymenidou, M. The RNA face of phase separation. Science 2018, 360, 859–860. [Google Scholar] [CrossRef]
  379. Jain, S.; Wheeler, J.R.; Walters, R.W.; Agrawal, A.; Barsic, A.; Parker, R. ATPase-Modulated Stress Granules Contain a Diverse Proteome and Substructure. Cell 2016, 164, 487–498. [Google Scholar] [CrossRef] [Green Version]
  380. Banani, S.F.; Lee, H.O.; Hyman, A.A.; Rosen, M.K. Biomolecular condensates: Organizers of cellular biochemistry. Nat. Rev. Mol. Cell Biol. 2017, 18, 285–298. [Google Scholar] [CrossRef]
  381. Hofweber, M.; Dormann, D. Friend or foe-Post-translational modifications as regulators of phase separation and RNP granule dynamics. J. Biol. Chem. 2019, 294, 7137–7150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  382. Miller, S.; Yasuda, M.; Coats, J.K.; Jones, Y.; Martone, M.E.; Mayford, M. Disruption of dendritic translation of CaMKIIalpha impairs stabilization of synaptic plasticity and memory consolidation. Neuron 2002, 36, 507–519. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Identification of mRNAs localized to the dendrite-enriched stratum radiatum (SR) layer in hippocampal CA1. Three different groups comprehensively identified mRNAs localized in the hippocampal SR layer using next-generation RNA sequencing. In the hippocampus, somas align in the stratum pyramidale (SP) and dendrites elongate into the SR. Cajigas et al., Nakayama et al., and Ainsley et al. identified mRNAs localized to the hippocampal SP and SR layers after isolating the layers from rodents. Cajigas et al. identified mRNAs abundant in the SR of the rat hippocampal CA1 region. Nakayama et al. identified mRNAs that are more enriched in the SR layer compared with the SP layer in the mouse hippocampus. They also identified mRNAs that were localized to the SR but were reduced in the SR of RNA granule protein 105 (RNG105, also known as Caprin1) conditional knockout (cKO) mice that showed long-term memory impairment. Ainsley et al. identified ribosome-bound mRNAs in the hippocampal SR of fear-conditioned mice. In this review, we compared the SR-enriched mRNA lists from these studies (colored in brown) and focused on dendritic mRNAs identified in common in all of these studies (Supplementary Table S1).
Figure 1. Identification of mRNAs localized to the dendrite-enriched stratum radiatum (SR) layer in hippocampal CA1. Three different groups comprehensively identified mRNAs localized in the hippocampal SR layer using next-generation RNA sequencing. In the hippocampus, somas align in the stratum pyramidale (SP) and dendrites elongate into the SR. Cajigas et al., Nakayama et al., and Ainsley et al. identified mRNAs localized to the hippocampal SP and SR layers after isolating the layers from rodents. Cajigas et al. identified mRNAs abundant in the SR of the rat hippocampal CA1 region. Nakayama et al. identified mRNAs that are more enriched in the SR layer compared with the SP layer in the mouse hippocampus. They also identified mRNAs that were localized to the SR but were reduced in the SR of RNA granule protein 105 (RNG105, also known as Caprin1) conditional knockout (cKO) mice that showed long-term memory impairment. Ainsley et al. identified ribosome-bound mRNAs in the hippocampal SR of fear-conditioned mice. In this review, we compared the SR-enriched mRNA lists from these studies (colored in brown) and focused on dendritic mRNAs identified in common in all of these studies (Supplementary Table S1).
Biomolecules 10 00167 g001
Figure 2. Gene ontology (GO) categories in which the identified dendritic mRNAs were enriched. The three studies (Cajigas et al., 2012; Ainsley et al., 2014; Nakayama et al., 2017) were compared, and 78 mRNAs were found to be commonly identified in the studies (a). The 78 mRNAs were classified by GO enrichment analysis using DAVID 6.8. GO categories in which the mRNAs were significantly enriched (b), and the list of mRNAs included in the GO categories (c) are shown.
Figure 2. Gene ontology (GO) categories in which the identified dendritic mRNAs were enriched. The three studies (Cajigas et al., 2012; Ainsley et al., 2014; Nakayama et al., 2017) were compared, and 78 mRNAs were found to be commonly identified in the studies (a). The 78 mRNAs were classified by GO enrichment analysis using DAVID 6.8. GO categories in which the mRNAs were significantly enriched (b), and the list of mRNAs included in the GO categories (c) are shown.
Biomolecules 10 00167 g002
Table 1. Summary of Supplementary Tables S1 and S2.
Table 1. Summary of Supplementary Tables S1 and S2.
TableTitleContent
Supplementary Table S1Dendritic mRNAs and their associated RNA-binding proteins
  • Candidate dendritic mRNAs commonly identified in studies that used RNA sequencing (RNA-seq) (Cajigas et al., 2012; Ainsley et al., 2014; Nakayama et al., 2017) are listed, and the general and neuronal functions of the encoded proteins are summarized.
  • Association of these mRNAs with RNA-binding proteins of RNA granules—FMRP, RNG105/Caprin1, G3BP, CPEB1, Stau1, and Stau2—are indicated.
Supplementary Table S2a–e1,2Target mRNAs of RNA-binding proteins of RNA granules
  • mRNAs associated with and regulated by the RNA-binding proteins of RNA granules are listed.
  • Effects of the RNA-binding proteins on the mRNAs are summarized.
RNA-binding proteins
2aFMRP
2bRNG105/Caprin1
2cG3BP
2dCPEB1
2eStaufen, Stau1, Stau2
1 “Target mRNA (Gene Name)” indicates the gene name shown in the references. “Gene Symbol” indicates the official gene symbol in Mus musculus. 2 Because the comprehensive analyses identified a large number of target mRNAs, the identified mRNAs are not concretely listed but indicated by asterisks (*). However, some comprehensive studies analyzed particular mRNAs more specifically. In such cases, those particular mRNAs are listed separately. Abbreviations: FMRP—fragile X mental retardation protein; RNG105—RNA granule protein 105; G3BP—Ras-GAP SH3 domain binding protein; CPEB1—cytoplasmic polyadenylation element binding protein 1; Stau1 and Stau2—staufen double-stranded RNA binding proteins 1 and 2.

Share and Cite

MDPI and ACS Style

Ohashi, R.; Shiina, N. Cataloguing and Selection of mRNAs Localized to Dendrites in Neurons and Regulated by RNA-Binding Proteins in RNA Granules. Biomolecules 2020, 10, 167. https://0-doi-org.brum.beds.ac.uk/10.3390/biom10020167

AMA Style

Ohashi R, Shiina N. Cataloguing and Selection of mRNAs Localized to Dendrites in Neurons and Regulated by RNA-Binding Proteins in RNA Granules. Biomolecules. 2020; 10(2):167. https://0-doi-org.brum.beds.ac.uk/10.3390/biom10020167

Chicago/Turabian Style

Ohashi, Rie, and Nobuyuki Shiina. 2020. "Cataloguing and Selection of mRNAs Localized to Dendrites in Neurons and Regulated by RNA-Binding Proteins in RNA Granules" Biomolecules 10, no. 2: 167. https://0-doi-org.brum.beds.ac.uk/10.3390/biom10020167

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