Skip to main content

Main menu

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

User menu

Search

  • Advanced search
eNeuro
eNeuro

Advanced Search

 

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

Ubiquitin and Ubiquitin-Like Proteins in the Critical Equilibrium between Synapse Physiology and Intellectual Disability

Alessandra Folci, Filippo Mirabella and Matteo Fossati
eNeuro 27 July 2020, 7 (4) ENEURO.0137-20.2020; https://doi.org/10.1523/ENEURO.0137-20.2020
Alessandra Folci
1Humanitas Clinical and Research Center-IRCCS, via Manzoni 56, 20089, Rozzano (MI), Italy
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Filippo Mirabella
2Department of Biomedical Sciences, Humanitas University, Via Rita Levi Montalcini 4, 20090 Pieve 9 Emanuele – Milan, Italy
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Matteo Fossati
1Humanitas Clinical and Research Center-IRCCS, via Manzoni 56, 20089, Rozzano (MI), Italy
3CNR–Institute of Neuroscience, via Manzoni 56, 20089, Rozzano (MI), Italy
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • ORCID record for Matteo Fossati
  • Article
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF
Loading

Abstract

Posttranslational modifications (PTMs) represent a dynamic regulatory system that precisely modulates the functional organization of synapses. PTMs consist in target modifications by small chemical moieties or conjugation of lipids, sugars or polypeptides. Among them, ubiquitin and a large family of ubiquitin-like proteins (UBLs) share several features such as the structure of the small protein modifiers, the enzymatic cascades mediating the conjugation process, and the targeted aminoacidic residue. In the brain, ubiquitination and two UBLs, namely sumoylation and the recently discovered neddylation orchestrate fundamental processes including synapse formation, maturation and plasticity, and their alteration is thought to contribute to the development of neurological disorders. Remarkably, emerging evidence suggests that these pathways tightly interplay to modulate the function of several proteins that possess pivotal roles for brain homeostasis as well as failure of this crosstalk seems to be implicated in the development of brain pathologies. In this review, we outline the role of ubiquitination, sumoylation, neddylation, and their functional interplay in synapse physiology and discuss their implication in the molecular pathogenesis of intellectual disability (ID), a neurodevelopmental disorder that is frequently comorbid with a wide spectrum of brain pathologies. Finally, we propose a few outlooks that might contribute to better understand the complexity of these regulatory systems in regard to neuronal circuit pathophysiology.

  • intellectual disability
  • ubiquitination
  • sumoylation
  • neddylation
  • synapse development
  • synapse function

Significance Statement

Ubiquitination, sumoylation, and neddylation are related PTMs modulating cellular and molecular pathways that are essential to generate fully functional neuronal circuits. Their impairment is indeed implicated in the pathogenesis of several disorders, including ID. Growing evidence now indicates they also functionally cooperate to govern synapse development and function. The main goals of this review are (1) to provide an overview of the current knowledge on the role of ubiquitination, sumoylation, and neddylation in synapse functions; (2) discuss how altered ubiquitination or sumoylation pathways may contribute to ID development; and (3) highlight evidence of a dynamic cross talk between these PTMs, which represents a novel mechanism that could lead to the identification of new principles underlying synaptic function and dysfunction in ID.

Introduction

Synapses are the basic functional units of the brain ensuring proper information processing and storage. In the forebrain, glutamate and GABA are the main excitatory and inhibitory neurotransmitters (NTs), respectively. Synaptic transmission occurs when an action potential reaches the active zone of the presynaptic terminal and triggers the Ca2+-dependent fusion of synaptic vesicles (SVs) to the presynaptic membrane (Südhof, 2004). SV fusion releases NT molecules into the synaptic cleft and convey a signal to NT receptors localized in the membrane of a specialized domain of the postsynaptic neuron, called the postsynaptic density (PSD). The PSD is composed by a complex network of distinct protein categories. Among them, postsynaptic NT receptors transduce the signal conveyed by NTs into electrical and biochemical cascades. Underneath the postsynaptic membrane, a meshwork of scaffolding proteins generates a structural platform governing synapse organization via multiple bindings to NT receptors, adhesion proteins, signaling molecules, and cytoskeletal elements (Sheng and Kim, 2011).

A prominent feature of synapses is their ability to dynamically modify the strength of synaptic transmission according to the inputs they receive, in a process called synaptic plasticity. Synaptic plasticity is thought to underlie learning and memory and is fundamental to adapt our behavior based on experience. During brain development, synaptic plasticity peaks in specific temporal windows of high sensitivity, called critical periods (Hensch, 2005), which are crucial to orchestrate the formation and refinement of synaptic networks and, ultimately, enables the acquisition of a given skill and/or cognitive function. The drawback of this particularly high sensitivity is a marked susceptibility to genetic and environmental insults that can interfere with molecular and cellular processes critical to synapse development, function, and plasticity. Indeed, perturbation of these pathways leads to neurodevelopmental and psychiatric disorders, such as autism, ID, and schizophrenia. Since these pathologies are characterized by the convergence of distinct pathways onto synaptic impairment, they are collectively defined as synaptopathies (Van Spronsen and Hoogenraad, 2010). To date, no effective therapies are available to treat these diseases. It is therefore crucial to understand the molecular mechanisms of synaptic dysfunction to provide the rationale to develop innovative therapeutic approaches.

Given the plastic nature of the brain, synapses have developed strategies to rapidly modify the strength of synaptic transmission. The dynamic modulation of protein activity via PTMs is a major mechanism to efficiently tune synapse assembly, maturation, and function. PTMs refer to covalent enzymatic modifications, either reversible or irreversible of target proteins, following their translation (Bürkle, 2001). They typically consist in the addition of a functional moiety, which can be either chemical groups or complex molecules, including lipids, sugars, nucleosides, and polypeptides to specific residues of target proteins. PTMs regulate multiple aspects of protein physiology from subcellular localization and activity to conformation and stability/turnover. In neurons, PTMs have been extensively investigated and modulate virtually all pathways that are required to ensure proper synaptic transmission and plasticity, such as presynaptic NT release (Takahashi et al., 2003; Hegde and DiAntonio, 2002; Schorova and Martin, 2016), trafficking and biophysical properties of NT receptors (Luscher et al., 2011; Schorova and Martin, 2016; Diering and Huganir, 2018), PSD organization (Zacchi et al., 2014; Coba, 2019), and synaptic adhesion (Jeong et al., 2017). Beyond the essential roles of PTMs in brain physiology, their impairment is thought to critically contribute to the etiology of several brain disorders including synaptopathies. Among PTMs, ubiquitination and ubiquitin-like proteins (UBLs) such as sumoylation and neddylation share multiple features, tightly interplay, and are vital to synapse assembly, maturation, and function. In this review, we focus on the role of ubiquitination, sumoylation, and neddylation in synapse physiology and their implication in the molecular pathogenesis of ID, a generalized neurodevelopmental disorder that manifests in a wide range of brain pathologies (Verpelli and Sala, 2012; Picker and Walsh, 2013; Vissers et al., 2016).

Ubiquitination and UBL Pathways in Synapse Physiology

Ubiquitination

Ubiquitination occurs in all eukaryotic cells. It consists in the reversible conjugation of the 76 amino acid (aa)-long ubiquitin protein to lysine (K) residues of target proteins. This process is catalyzed by a series of enzymatic reactions. E1 ubiquitin enzymes bind to and activate free ubiquitin through adenylation at ubiquitin C-terminal and thiol transfer. Activated ubiquitin is then transferred to E2 ubiquitin-conjugating enzymes and finally transferred onto a K residue of target proteins by E3 ubiquitin ligases. Protein substrates can be either mono-ubiquitinated or poly-ubiquitinated and ubiquitin chains can vary depending on which of the seven K residues of ubiquitin is used to covalently attach the subsequent ubiquitin (for a comprehensive review, see Komander and Rape, 2012). The human genome encodes two E1, ∼50 E2, and ∼600 E3 enzymes. Thus, substrate specificity mainly relies on E3 ligases and different combinations of E2-E3 proteins. Ubiquitination is counterbalanced by the action of deubiquitinating enzymes (DUBs) that remove ubiquitin from target proteins. The coordinated activity of ubiquitinating enzymes and DUBs is essential to set and maintain ubiquitin homeostasis (Kowalski and Juo, 2012).

Protein degradation through the ubiquitin proteasome system (UPS) is the best characterized function of ubiquitination and K48 poly-ubiquitin chains are the most common signal to target proteins for degradation. Conversely, mono-ubiquitination and K63 poly-ubiquitin chains mediate non-proteasomal functions and typically modulate phosphorylation-dependent protein activation, protein-protein interactions, and membrane protein trafficking (Komander and Rape, 2012). Both proteasomal and non-proteasomal ubiquitin conjugations occur at the synapse, where they regulate molecular processes important for synapse formation, maturation, and plasticity. As a consequence, ubiquitination is critical for long-term memory formation and stability as observed with a variety of behavioral paradigms (for review, see Jarome and Helmstetter, 2013).

Presynaptic ubiquitination

The importance of ubiquitination to the presynaptic function was first described in Drosophila neuromuscular junctions (NMJs). Pharmacological and genetic perturbations of the UPS lead to the accumulation of Dunc-13, a protein important for SV release, and significantly increase presynaptic efficacy (Speese et al., 2003). In line with this, MUN-13, the mouse ortholog of Dunc-13 is also ubiquitinated by the E3 ligase FBXO45 in the hippocampus (Tada et al., 2010; Fig. 1A). Syntaxin 1 (STX1), a major component of the SNAP REceptor (SNARE) complex and the binding partner of the presynaptic Ca2+ sensor synaptotagmin (Rizo and Xu, 2015), is poly-ubiquitinated by the E3 enzyme Staring (Chin et al., 2002; Fig. 1A). Ubiquitination further regulates NT release by targeting the presynaptically expressed Group III metabotropic glutamate (mGlu)7 receptor, which inhibits glutamate and GABA release (Niswender and Conn, 2010). Upon stimulation with L-glutamate, the E3 ligase NEDD4 induces rapid mGlu7 internalization and degradation via both proteasomal and lysosomal pathways (Lee et al., 2019; Fig. 1A). Protein degradation via the UPS is also a major regulator of SV recycling (Willeumier et al., 2006). Interestingly, blocking action potentials with tetrodotoxin (TTX) prevents the effect of proteasome inhibition on the recycling vesicle pool, suggesting that presynaptic UPS may be a negative feedback-regulator of synaptic transmission.

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

Neuronal ubiquitination and ubiquitin-like modifications. A, C, Major components of excitatory and inhibitory synapses targeted by ubiquitin (blue squares) SUMO (purple triangles) and NEDD8 (orange hexagons) pathways. Deubiquitinating enzymes (green clamshell-like shapes) and components of the SUMO (UBC9 and SENPs) and NEDD8 machineries (NAE, UBC12, and UBE2F) are also indicated. Although NEDD8 pathway and targets are also present in the presynaptic compartment, for simplicity they are depicted in the postsynaptic region only. In A, E3 ubiquitin ligases operating at excitatory synapses and their known substrates are listed in the left table. B, Nuclear sumoylation and neddylation critical to synaptic function are indicated.

Beside basal synaptic transmission, presynaptic ubiquitination is also a critical determinant of synaptic plasticity and is required for cognition. Genetic deletion of the gene enconding the anaphase-promoting complex/cyclosome (APC/C) in the forebrain of adult mice impairs hippocampal-dependent memories, such as spatial memory and extinction of fear memory resulting in anxiety-related behaviors (Li et al., 2008; Kuczera et al., 2010). While at Drosophila and Caenorhabditis elegans NMJs it is known that APC/C targets the active zone protein Liprin-α (van Roessel et al., 2004; Kowalski et al., 2014), the presynaptic targets in mammals remain elusive. Recently, the E3 ligase RNF8 was shown to be required to build up cerebellar circuits mediating procedural motor learning by limiting the formation of parallel fiber presynaptic boutons onto Purkinje cells by targeting as yet unidentified substrates (Valnegri et al., 2017). Another presynaptic E3 ligase fundamental to establish neuronal connectivity is SCRAPPER. In mouse hippocampal neurons, SCRAPPER modulates multiple aspects of the presynaptic function by directly ubiquitinating and regulating Rab3-interacting protein 1 (RIM1; Fig. 1A). Consistent with the role of RIM1 as major SV priming factor (Südhof, 2004), axonal boutons of Scrapper-knock-out (KO) neurons show enhanced synaptic transmission, owing a significant increase of NT release probability (Yao et al., 2007; Takagi et al., 2012). Moreover, the absence of Scrapper expression interferes with performances in contextual fear conditioning tests and hippocampi derived from these mice exhibit bidirectional changes of synaptic plasticity regulation (also referred to as metaplasticity; Yao et al., 2011; Takagi et al., 2012). These cellular and behavioral features partially recapitulate defects observed in Rim1-KO animals (Castillo et al., 2002; Powell et al., 2004), although there may be other SCRAPPER synaptic targets. Finally, a spontaneous mutation in the Usp14 gene, encoding the proteasome-associated deubiquitinating enzyme USP14 (Fig. 1A), leads to progressive locomotor defects and ataxia (Wilson et al., 2002). Loss of Usp14 results in impaired short-term facilitation and reduced SV number, indicating the importance of a balanced ubiquitination to preserve presynaptic functions (Walters et al., 2014). However, which proteins are targeted by USP14 and whether USP14 facilitates or inhibits presynaptic UPS are still unclear.

Intriguingly, two large scaffolding proteins of the active zone, Piccolo and Bassoon, were suggested to be important regulators of presynaptic ubiquitination (Waites et al., 2013). Interference with the expression of these genes leads to aberrant degradation of proteins operating in the active zone and degeneration of presynaptic boutons. As this phenotype is partially rescued by either the inhibition of the proteasome or the downregulation of the E3 ligase SIAH1, it is likely that Piccolo and Bassoon limit presynaptic ubiquitination via the suppression of SIAH1 activity.

Altogether, the aforementioned studies clearly indicate that ubiquitination is critical to NT release, the primary function of presynaptic terminals, and at the circuit level participates to memory and learning processes.

Postsynaptic ubiquitination

At the postsynaptic site, ubiquitination is a key mechanism that shapes the functional organization of both excitatory and inhibitory synapses by targeting multiple categories of postsynaptic proteins (Fig. 1A,C). Consistent with this, the proteasome is rapidly recruited and trapped into dendritic spines, the major site for excitatory synapses, after membrane depolarization (Bingol and Schuman, 2006; Bingol et al., 2010).

One of the most well-studied ubiquitinated NT receptors are AMPA receptors (AMPARs; Figs. 1A, 2). They are glutamate-gated ion channels (also referred to as ionotropic glutamate receptors, iGluRs) and mediate the fast excitatory transmission in the brain (Greger et al., 2017). The regulation of their number at the synaptic surface is a major determinant of synaptic strength and plasticity (Choquet and Triller, 2013; Huganir and Nicoll, 2013). The importance of AMPAR ubiquitination was first demonstrated in C. elegans, where it controls synaptic abundance of these receptors and locomotor behavior (Burbea et al., 2002; Juo and Kaplan, 2004; Dreier et al., 2005; Emtage et al., 2009). In mammals, the selective ubiquitination of GluA1 and GluA2 subunits is required for the activity-dependent endocytosis of surface AMPARs directing them toward degradative pathways. Ubiquitination occurs at K residues of their intracellular C-terminal tail via the E3 ligases NEDD4-1 and RNF167 (Schwarz et al., 2010; Lussier et al., 2011, 2012). A more recent study showed that activity-dependent AMPAR ubiquitination involves all four GluA subunits (GluA1-4) and sorts AMPARs from early to late endosomes and subsequent lysosome-dependent degradation (Widagdo et al., 2015). Considering that the modulation of AMPAR number at synaptic sites tunes synaptic strength and plasticity, it is expected that AMPAR ubiquitination has implication in learning and memory paradigms. Consistently, proteasome inhibition increases synaptic surface GluA1-containing AMPARs and enhances fear-conditioned learning (Yeh et al., 2006). UPS-dependent degradation of GluA1 also facilitates the extinction of fear memories, which is mediated by NMDAR-dependent depression of synaptic activity (Mao et al., 2008). AMPAR ubiquitination is finely counteracted by the opposite action of a few DUBs. Overexpression of Usp46 prolongs GluA1 half-life, resulting in enhanced amplitude of excitatory postsynaptic currents (Huo et al., 2015). USP8 antagonizes the E3 ligase NEDD4-1 and is involved in downscaling synaptic strength during homeostatic plasticity, providing the first evidence for an opposed bidirectional control of synaptic strength by ubiquitination (Scudder et al., 2014).

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

Postsynaptic control of glutamate and GABAA receptors by ubiquitination and sumoylation. UPS-dependent degradation (green arrows) of PSD-95 destabilizes surface AMPARs, resulting in enhanced receptor lateral mobility and consequently, endocytosis (box A). Ubiquitination of GluA1 and GluA2 decreases surface AMPARs through clathrin-dependent endocytosis (orange arrows). Ubiquitinated ARC is degraded via the UPS pathway. As ARC is a major regulator of AMPAR internalization, reduced ARC levels suppress AMPAR endocytosis. Conversely, sumoylated ARC triggers AMPAR internalization. Moreover, ubiquitination of intracellular GluA1-4 may also promote AMPAR sorting to the lysosomal degradation pathway (red arrows). Ubiquitination of mGlu1-5 receptors enhanced their clathrin-dependent endocytosis. Surface NMDARs are regulated by the ubiquitination pathway in a subunit-dependent manner. GluN2B undergoes a phosphorylation-dependent ubiquitination (box B), leading to UPS-dependent degradation of NMDARs. GluN2D ubiquitination enhances its degradation, although it is not clear whether it utilizes UPS-dependent or lysosomal-dependent pathways (green and red dotted arrows). Finally, ubiquitination of newly synthetized GluN1 and GluN2A results in NMDAR retrotranslocation from the ER to the cytosol and subsequent degradation through the ERAD pathway (blue arrows). Similar to GluN2D, GluK2 ubiquitination is phosphorylation-dependent and triggers its degradation through an as yet ill-defined pathway (green and red dotted arrows). In contrast, sumoylated GluK2-containing KARs are removed from the synaptic membrane via clathrin-dependent endocytosis. At inhibitory synapses, ubiquitinated γ2-containing GABAAR are sorted to the lysosomal degradation pathway, while β3-containing GABAAR are ubiquitinated in the ER and degraded through the ERAD machinery.

Among the iGluR family, NMDA receptors (NMDARs) play an essential role in synaptic plasticity (Paoletti et al., 2013), and their function, similarly to other iGluR members, is modulated by ubiquitination (Figs. 1A, 2). The GluN2B, GluN2D, GluN1, and GluN2A subunits are targeted by the E3 ubiquitin ligases Mind bomb-2 (MIB2), NEDD4, and FBXO2 (Kato et al., 2005; Jurd et al., 2008; Gautam et al., 2013; Atkin et al., 2015). Upon phosphorylation-dependent ubiquitination of GluN2B by MIB2 in the PSD, NMDAR-mediated currents are significantly reduced. This effect is prevented when cells are treated with the proteasome inhibitor MG132, suggesting that ubiquitinated GluN2B undergoes UPS-dependent degradation (Jurd et al., 2008). Similarly, NEDD4 selectively conjugates poly-ubiquitin chains to GluN2D and reduces NMDAR currents in heterologous cells (Gautam et al., 2013). Whether ubiquitination directly mediates GluN2D proteasome-dependent degradation or modulates receptor endocytosis and sorting to late endosomes and lysosomes is not known. Conversely, FBXO2 ubiquitinates newly synthetized GluN1 and GluN2A subunits in the endoplasmic reticulum (ER) and mediates its degradation via the ER-associated degradation (ERAD) machinery, a mechanism that avoids the formation of supra-numerary synapses in the dendritic shaft and aberrant NMDAR-mediated currents (Kato et al., 2005; Nelson et al., 2006; Gascón et al., 2007; Atkin et al., 2015). Synaptic abundance and clustering of GluN1-containing NMDARs are also modulated by the hominoid-specific DUB USP6 (Zeng et al., 2019). By generating a humanized knock-in (KI) mouse expressing USP6 under the control of Ca2+/calmodulin-dependent kinase II (CamkII) promoter, the authors found that USP6 stabilizes NMDARs at synapses and enhances synaptic function resulting in improved mouse cognitive abilities. Beside the novel synaptic mechanism uncovered here, this study suggests for the first time that ubiquitination may have contributed to the evolution of human-specific synaptic features, which are thought to form, together with other processes the cellular basis of human intelligence (DeFelipe, 2011; Geschwind and Rakic, 2013). In line with this, perturbation of USP6 is associated with human-specific neuropsychiatric disorders, such as ID (Ou et al., 2006) and autism spectrum disorders (ASD; Tentler et al., 2003).

Kainate receptors (KARs) are the third receptor subtypes of iGluRs whose activity and trafficking are modulated by ubiquitination (Figs. 1A, 2). The E3 ligases Cullin 3 (CUL3) and Parkin 2 (PARK2) target the GluK2 subunit of KARs and regulate its surface expression (Salinas et al., 2006; Maraschi et al., 2014). Strikingly, loss-of-function of PARK2, which is causative of the most common form of familial juvenile parkinsonism, leads to abnormal levels of synaptic GluK2, a mechanism that could underlie glutamate excitotoxicity and neurodegeneration in Parkinson’s disease. Yet, it is unclear whether KAR ubiquitination occurs on surface receptors and whether it is degraded through UPS or lysosomal pathway.

mGlu receptors are also targeted by ubiquitination at the postsynaptic compartment (Figs. 1A, 2). Upon dihydroxyphenylglycine (DHPG)-induced activation of Group I mGlu1-5, the E3 ligase SIAH1 attaches K63-linked poly-ubiquitin chains to intracellular K residues of these receptors and induces their internalization (Moriyoshi et al., 2004; Ko et al., 2012; Gulia et al., 2017). Since their activation is critical for the expression of long-term depression (LTD; Niswender and Conn, 2010), mGlu receptor ubiquitination emerged as an essential regulatory mechanism limiting excessive synaptic depression (Gulia et al., 2017).

As aforementioned, postsynaptic scaffolding proteins are the structural core of the PSD. Here, complex protein-protein interactions are tightly regulated by PTMs to control the functional organization of synapses (Sheng and Hoogenraad, 2007). Given the multiple protein-protein interactions that each scaffolding protein engages, the ubiquitination of a few scaffolding proteins allows the precise regulation of a large set of postsynaptic molecules (Fig. 1A). PSD-95, GKAP, AKAP79/150, SHANK, and HOMER1A are abundant scaffolding proteins of excitatory synapses and their levels are bidirectionally modulated by ubiquitination in an activity-dependent manner (Ageta et al., 2001; Colledge et al., 2003; Ehlers, 2003; Rezvani et al., 2007; Na et al., 2012). Importantly, the targeted degradation of scaffolding molecules critically contributes to learning and behavior. Retrieval of fear memory upregulates polyubiquitinated SHANK and GKAP, but not PSD-95. Consistent with this, infusion of the proteasome inhibitor clasto-lactacystin-β-lactone in the CA1 region of the hippocampus immediately after retrieval prevents extinction of fear memory (Lee et al., 2008). To date, only few of the E3 ligases targeting scaffolding proteins were identified. The E3 ligase MDM2 ubiquitinates PSD-95 at multiple residues and induces distinct pathways. The mono-ubiquitination of PSD-95 within the PEST (peptides rich in proline, glutamate, serine, and threonine) motif, a short sequence serving as proteolytic signal, triggers PSD-95 degradation. As a consequence, the number of PSD-95 molecules available to anchor AMPARs is reduced, resulting in less stable surface AMPARs and enhanced endocytosis (Fig. 2), a mechanism that critically contributes to hippocampal LTD (Colledge et al., 2003). Conversely, the CDK5-dependent ubiquitination of PSD-95 at K10 promotes its interaction with and the recruitment of the clathrin adaptor protein complex (AP)2 at the synapse, triggering clathrin-dependent AMPAR endocytosis (Bianchetta et al., 2011). The ubiquitination of GKAP, mediated by the E3 ubiquitin ligase TRIM3, induces structural changes of dendritic spines (Hung et al., 2010). While the enzymes mediating SHANK ubiquitination are not known, it has been recently found that USP8 selectively deubiquitinates SHANK1 and SHANK3 and modulates dendritic spine density and morphology (Campbell and Sheng, 2018).

Among postsynaptic signaling proteins, the ubiquitination of the immediate early gene activity-regulated cytoskeleton-associated protein ARC is best characterized (Figs. 1A, 2). ARC responds to various forms of synaptic plasticity and promotes the endocytosis of AMPARs (Shepherd et al., 2006; Waung et al., 2008). UBE3A is the main causal gene of a severe neurodevelopmental disorder, the Angelman syndrome (AS), and encodes the first E3 ligase proposed to target ARC (Greer et al., 2010). Since UBE3A is also a transcriptional coregulator (Nawaz et al., 1999), it is unclear whether UBE3A-dependent regulation of ARC operates at the protein level through ubiquitination or at the transcriptional level (Kühnle et al., 2013). Other studies indicated that ARC is ubiquitinated by the E3 ligase TRIAD3 (Na et al., 2012; Mabb et al., 2014). Another signaling protein targeted by ubiquitination is the spine-associated Rap GTPase activating protein (SPAR; Fig. 1A). SPAR forms a complex with PSD-95 and NMDARs and is critical for spine structural plasticity (Pak et al., 2001). Upon induction of synaptic down-scaling, phosphorylated SPAR is ubiquitinated by the E3 ligase complex SCFβTRPC resulting in its degradation and spine morphology changes (Pak and Sheng, 2003; Ang et al., 2008). Eventually, proteomic approaches aimed at identifying synaptic ubiquitome in rat brains revealed that the CAMKII, which is crucial for the expression of synaptic plasticity (Coultrap and Bayer, 2012), is ubiquitinated (Na et al., 2012; Fig. 1A). Yet, the identity of the E3 ligase that targets CAMKII, the consequences on CAMKII stability and the functional relevance for plasticity remain to be assessed.

Ubiquitination was also studied at inhibitory synapses (Fig. 1C). The number of surface GABAARs is critically modulated by the ubiquitin-associated chaperon PLIC-1 (also named ubiquilin; Bedford et al., 2001; Zhang et al., 2015). Given the association of PLIC proteins with the proteasome in the ER (Kleijnen et al., 2000), it was proposed that PLIC-1 inhibits GABAAR degradation via ERAD (Bedford et al., 2001; Saliba et al., 2008). Accordingly, synaptic upregulation and downregulation of surface GABAARs are governed by changes in poly-ubiquitination dynamics of newly synthesized β3 subunits in the ER (Saliba et al., 2007; Fig. 2). In contrast, the ubiquitination of the γ2 subunit does not alter the forward-directed biosynthetic route, but instead triggers GABAAR endocytosis and subsequent lysosome-dependent degradation (Arancibia-Cárcamo et al., 2009; Jin et al., 2014; Fig. 2). In the cerebellum, the USP14-dependent deubiquitination of the α1 subunit, which is contained in the majority of cerebellar GABAARs, reduces surface GABAARs and favors receptor sorting toward lysosomal compartments (Lappe-Siefke et al., 2009; Fig. 1C). At inhibitory synapses, unlike their glutamatergic counterparts, no scaffolding proteins were found ubiquitinated, so far. Notwithstanding, gephyrin, the major scaffolding protein of inhibitory synapses (Tyagarajan and Fritschy, 2014), contains two PEST sequences (Tyagarajan et al., 2011). Whether these PEST motifs, as that of PSD-95, are targeted by ubiquitination is unclear.

Collectively, there is strong evidence that ubiquitination targets multiple synaptic components and represents a complex regulatory system controlling synapse formation, maintenance, and plasticity. In line with this, ubiquitination is also required for learning and memory. However, the molecular logic of how the ubiquitin system controls cognition and how targeted ubiquitination determines specific behavioral outputs remain enigmatic.

Sumoylation

Sumoylation consists in the covalent but reversible conjugation of the 100-aa-long small ubiquitin-like modifier (SUMO) protein to specific K residues of substrates (for a more comprehensive review, see Flotho and Melchior, 2013). Like ubiquitination, sumoylation requires a dedicated enzymatic pathway and its homeostasis is finely regulated by conjugating and deconjugating enzymes. SUMO is first synthetized as a non-conjugatable precursor that is cleaved by Sentrin-proteases (SENPs) to generate a mature SUMO molecule. Mature SUMO is subsequently activated by heterodimers of SUMO-activating enzyme (SAE)1 and SAE2 in an ATP-dependent manner and transferred to the catalytic cysteine of the sole SUMO-specific conjugating enzyme UBC9. UBC9 finally catalyzes SUMO conjugation to designated substrates. Targeted K residues typically localize within the consensus motif Ψ-K-x-E, where Ψ is a hydrophobic residue and x is any aa. SUMO deconjugation occurs through the activity of SENPs. Mammalian cells express three different SUMO paralogues (SUMO1-3) and six SENPs (SENP1-3, SENP5-7). As all PTMs, sumoylation modulates the function of target proteins by distinct mechanisms. It can (1) trigger conformational changes that affect protein activity; (2) inhibit or favor protein-protein interactions by either generating or masking a binding site; (3) control substrate stability and turnover through a SUMO-mediated recruitment of members of the SUMO-targeted ubiquitin ligase family (Flotho and Melchior, 2013).

Originally, sumoylation was described as nuclear modification regulating protein translocation across the nuclear membrane and transcription (Matunis et al., 1996; Mahajan et al., 1997). Since then, many other SUMO substrates were identified inside and outside the nuclear compartment, including synapses. It is now well accepted that sumoylation controls several processes critical for neuronal function such as neuronal excitability and synapse development and plasticity (Schorova and Martin, 2016; Henley et al., 2018). During brain development, the SUMO machinery and neuronal sumoylome dynamics are tightly regulated in a spatiotemporal manner. In the brain, the overall expression of SUMO components and SUMO-conjugated proteins peaks during late stages of embryonic development (embryonic day (E)13–E18) and rapidly decreases postnatally (Watanabe et al., 2008; Loriol et al., 2012; Hasegawa et al., 2014; Josa-Prado et al., 2019). In addition, these proteins are redistributed from the nucleus to the synapse after birth (Loriol et al., 2012), where sumoylation still occurs in mature cortical/hippocampal neurons and is rapidly enhanced by neuronal activity (Loriol et al., 2013; Colnaghi et al., 2019). According to the critical role of sumoylation in controlling synaptic function, UBC9 and SENP1 enzymes dynamically diffuse inside and outside dendritic spines in an activity-dependent manner (Loriol et al., 2014; Schorova et al., 2019). The activation of mGlu5 triggers the transient trapping of UBC9 in the head of dendritic spines, leading to a rapid increase in synaptic sumoylation (Loriol et al., 2014). UBC9 recruitment is followed by time-dependent decrease in the exit rate of SENP1 from dendritic spines, resulting in the postsynaptic accumulation of SENP1, which restores synaptic sumoylation to initial levels (Schorova et al., 2019). Moreover, perturbation of SUMO homeostasis affects synapse physiology and consequently impacts cognitive functions. Decreasing sumoylation levels by overexpressing either a dominant negative form of UBC9 or the catalytic domain of SENP1 prevents the increase of surface AMPARs upon long-term potentiation (LTP) induction, indicating that sumoylation is essential for the expression of synaptic plasticity (Jaafari et al., 2013). Accordingly, acute inhibition of sumoylation impairs hippocampal-dependent learning and memory in mice (Lee et al., 2014). Neuron-specific silencing of SUMO1-3 induces anxiety-like responses and impaired episodic memory providing the first evidence that SUMO conjugation is essential for emotionality and cognition (Wang et al., 2014). Conversely, increase of global SUMO1-ylation in SUMO1 transgenic mice or by chronic infusion of exogenous SUMO1 impairs learning and memory performances (Matsuzaki et al., 2015; Yoo et al., 2017). Furthermore, the loss of SENP2 in conditional KO mice results in several behavioral defects including spatial working memory impairment (Huang et al., 2020). RNA sequencing also revealed that the expression of genes critical for learning and memory is finely regulated by SENP2.

Together, these findings pointed out that sumoylation is essential to synapse development and function, and a balanced sumoylation/desumoylation is crucial for learning and memory processes. In line with this, over the last decade, several groups identified novel SUMO targets critical to synaptic function (Schorova and Martin, 2016; Henley et al., 2018).

Extrasynaptic sumoylation critical to the synaptic function

The first identified SUMO substrates bearing synaptic functions were components of the myocyte enhancer factor 2 (MEF2) family of transcription factors (Fig. 1B). In vitro and in vivo studies showed that sumoylation represses the transcriptional activity of MEF2A, MEF2C, and MEF2D (Grégoire et al., 2006; Kang et al., 2006; Shalizi et al., 2006). In particular, MEF2A sumoylation at K403 is critical for dendritic claw differentiation in the cerebellum via transcriptional inhibition of the transcription factor Nur77 (Shalizi et al., 2006). Interestingly, MEF2A sumoylation is tightly modulated by neuronal activity and requires a complex cross talk with other PTMs. The Ca2+-dependent phosphatase calcineurin dephosphorylates MEF2A at serine (S)408 and triggers the switch of K403 from being sumoylated to acetylated, resulting in the activation of MEF2A. Acetylated MEF2A thus enhances Nur77 transcription and leads to the inhibition of dendritic claw formation. Subsequently, it was also demonstrated that MEF2A sumoylation is critical for presynaptic differentiation by regulating the transcription of synaptotagmin 1, thus indicating that MEF2A sumoylation has a broader role in synaptic development (Shalizi et al., 2007).

A second extrasynaptic target of the SUMO machinery is the DNA binding protein and transcriptional repressor Methyl-CpG binding protein 2 (MECP2; Fig. 1B). Sumoylation at K223 of MECP2 is required for the recruitment of histone deacetylase complexes 1/2 (HDAC 1/2) and the consequent repression of gene expression, an essential event for the correct development of hippocampal excitatory synapses in vitro and in vivo (Cheng et al., 2014). Conversely, sumoylation at K412 decreases the physical interaction between MECP2 and the cAMP response element-binding protein (CREB), which ultimately enhances the transcription of the brain-derived neurotrophic factor (Bdnf) transcription (Tai et al., 2016). Similar to MEF2A, MECP2 phosphorylation in nearby residues (S421 and threonine ‘T’308) facilitates MECP2 sumoylation, further highlighting the functional interplay between distinct PTMs. As described below, mutations in MECP2 gene are associated with neurodevelopmental disorders, among which the Rett syndrome (RTT) is the most prevalent (Guy et al., 2011).

Recently, two other transcription factors, FOXP1 and FOXP2, were found sumoylated (Fig. 1B). These factors are required for the assembly of cerebellar circuits underlying vocalization and motor skills (Bacon and Rappold, 2012). FOXP2 sumoylation at K674 regulates its transcriptional activity (Estruch et al., 2016; Meredith et al., 2016) and mediates dendritic outgrowth and arborization of Purkinje cells (Usui et al., 2017). In line with this, individuals from a family affected by speech and language disorders display a significant reduction of sumoylated FOXP2, supporting the pivotal role of FOXP2 sumoylation for the acquisition of these communication skills (Meredith et al., 2016). Similarly to FOXP2, FOXP1 sumoylation at K670 is indispensable for dendritic outgrowth and complexity in cortical neurons via inhibition of FOXP1 transcriptional activity (Rocca et al., 2017). Remarkably, the autism-linked CNTNAP2 gene, which promotes the development of dendritic arbors, is transcriptionally repressed by sumoylated FOXP1, suggesting a potential molecular mechanism underlying FOXP1 function during neuronal development (Rocca et al., 2017).

Fragile X mental retardation protein (FMRP) is the most recently identified extrasynaptic SUMO target with synaptic roles (Khayachi et al., 2018; Fig. 1B). FMRP is an RNA-binding protein and a key component of neuronal mRNA granules. In these granules, FMRP transports translationally-repressed mRNAs along axons and dendrites to the base of active synapses, where their local translation is a key process for synapse maturation and plasticity (Prieto et al., 2020). The rapid activation of mGlu5 results in FMRP sumoylation at K88 and K130, which triggers the dissociation of FMRP from mRNA granules and enables the local translation of mRNAs that control dendritic spine elimination and maturation (Khayachi et al., 2018). It remains to be elucidated whether FMRP sumoylation discharges the whole set of mRNA molecules associated with FMRP or promotes the release of a specific subset of these mRNAs.

Presynaptic sumoylation

At the presynaptic compartment, sumoylation regulates diverse functions. Synaptosomal fractions loaded with purified SUMO1 display a significant reduction of Ca2+ influx and KCl-evoked glutamate release. Accordingly, Ca2+ influx and KCl-evoked glutamate release are enhanced in synapses supplemented with SENP1 (Feligioni et al., 2009). Intriguingly, modulation of synaptic sumoylation produces the reverse effect on glutamate release evoked by kainate stimulation. These results suggest that different stimuli triggers sumoylation of distinct presynaptic proteins to either inhibit or promote NT release (Feligioni et al., 2009). A primary presynaptic function controlled by the SUMO pathway is SV docking/priming through sumoylation of RIM1α and synapsin Ia (SYNIa; Fig. 1A). Sumoylation of SYNIa triggers its association with SVs and facilitates SV anchoring at the presynaptic membrane (Tang et al., 2015). Similarly, RIM1α sumoylation at K502 promotes its interaction with the voltage-gated Ca2+ channels (CaV) 2.1, favoring their clustering at the presynaptic membrane and enhancing the Ca2+ influx required for NT release (Girach et al., 2013). Conversely, non-sumoylated RIM1α prevents CaV2.1 clustering and increases SV docking in the active zone. Thus, the switch between the sumoylated and non-sumoylated forms of RIM1α is a key determinant for the fast and synchronous NT release.

Sumoylation also directly controls SV exocytosis by targeting STX1A (Craig et al., 2015; Fig. 1A). Sumoylated STX1A is more associated with two other components of the SNARE complex, SNAP-25 and VAMP-2, providing the mechanical forces required for SV exocytosis. Indeed, preventing STX1A sumoylation reduces its interaction with SNARE proteins and disrupts the balance of SV endo/exocytosis, skewing toward endocytosis.

In 2015, Matsuzaki and colleagues generated a transgenic mouse line overexpressing SUMO1 in neurons aiming to identify the neuronal sumoylome (Matsuzaki et al., 2015). Among the 95 SUMO1 substrates identified by mass spectrometry (MS), only synaptotagmin-1 was further validated biochemically (Fig. 1A). Nevertheless, further investigations are required to exclude possible off-target sumoylation owing to SUMO1 overexpression.

Regarding mGlu receptors, the majority of Group III were identified as SUMO substrates in primary cultured neurons and heterologous cells (Tang et al., 2005; Seebahn et al., 2008; Dütting et al., 2011; Wilkinson and Henley, 2011; Fig. 1A). mGlu7 is the sole member shown to be sumoylated in brain tissues. Like other SUMO targets, its sumoylation has to be preceded by PKC-dependent phosphorylation at S862 (Choi et al., 2016). Opposed to ubiquitination, mGlu7 sumoylation stabilizes its expression at the cell surface. Therefore, ubiquitination and sumoylation of mGlu7 bidirectionally modulate its surface expression at the presynaptic membrane and ultimately, regulate NT release. The type-1 endocannabinoid (CB1) receptors are another class of metabotropic receptors that regulates presynaptic functions (Castillo et al., 2012). Interestingly, it was shown that conjugated and unconjugated SUMO1 levels transiently increase upon CB1 activation as well as CB1 itself might be targeted by SUMO (Gowran et al., 2009). However, additional experiments are needed to confirm this hypothesis.

Among the voltage-gated ion channels targeted by sumoylation, potassium channels are the most represented (Fig. 1A). Senp2-deficient mice display hypersumoylation of voltage-gated potassium (KV) 1.1 and 7.2/7.3 channels. While sumoylation of KV1.1 does not alter channel activity, enhanced sumoylation of KV7.2 reduces depolarizing M-currents, resulting in neuronal hyperexcitability and epileptic seizures (Qi et al., 2014). Similarly, sumoylation of surface KV2.1 reduces channel activity and, as a consequence, enhances neuronal excitability (Plant et al., 2011). Interestingly, KV4.2 was recently found sumoylated in two distinct sites, producing different effects (Welch et al., 2019). While sumoylation at K437 triggers KV4.2 surface expression, sumoylation at K579 decreases the amplitude of K currents without any change in KV4.2 trafficking. Differently, the sumoylation of voltage-gated sodium (NaV) 1.2 channels enhances the amplitude of Na+ currents (Plant et al., 2016; Fig. 1A).

Sumoylation is also essential to control the axonal trafficking and function of NaV1.7 by targeting the collapsin response mediator protein 2 (CRMP2; Fig. 1A). CRMP2 is highly expressed in the brain and is critical for microtubule remodeling, neuronal polarity, axon outgrowth, and synapse dynamics (Arimura et al., 2004; Zhang et al., 2016). Expression of a CRMP2 SUMO-deficient mutant prevents its physical interaction with NaV1.7 and decreases NaV1.7 surface expression, resulting in reduced amplification of membrane depolarization in presynaptic boutons (Dustrude et al., 2013, 2017; Ju et al., 2013). CRMP2 sumoylation is also phosphorylation dependent (Dustrude et al., 2016). Indeed, interfering with either CRMP2 sumoylation or phosphorylation reduces surface NaV1.7. CRMP2 also interacts with the actin cytoskeleton in dendrites and dendritic spines. It was recently found that desumoylation and dephosphorylation of CRMP2 independently promote the formation and maturation of dendritic spines, broadening the importance of PTMs in regulating the multiple functions of CRMP2 (Zhang et al., 2018).

Beside targeted sumoylation, an intriguing concept that might also apply to synapses (Henley et al., 2018) is that sumoylation may occur simultaneously to components of protein assemblies. This idea of a synchronous “SUMO spray” on multiple targets (named “SUMO velcro”) was first shown to occur in the nucleus, where DNA damage triggers a SUMO wave that activates multiple components of the enzymatic pathway required to repair the DNA (Psakhye and Jentsch, 2012). Although not yet experimentally validated, the model of SUMO spray might properly suit to presynaptic NT release, which requires a coordinated series of protein-protein interactions that are tightly regulated by sumoylation (Henley et al., 2018).

Postsynaptic sumoylation

To date, postsynaptic sumoylation has been less characterized than extrasynaptic and presynaptic sumoylation. Nevertheless, the first identified synaptic target of the SUMO pathway was GluK2 (Martin et al., 2007; Figs. 1A, 2). Kainate stimulation promotes GluK2 sumoylation at K886, triggering receptor endocytosis during LTD expression at mossy fiber–CA3 synapses in the hippocampus (Martin et al., 2007; Konopacki et al., 2011; Chamberlain et al., 2012). As for other SUMO targets, the PKC-dependent phosphorylation of S868-GluK2 is required for its subsequent sumoylation (Konopacki et al., 2011; Chamberlain et al., 2012). As aforementioned, GluK2 is also ubiquitinated. However, whether these two PTMs operate in synergy, reciprocally compete or are independent remains to be investigated.

A second postsynaptic SUMO target is the calcium/calmodulin-dependent serine protein kinase (CASK; Fig. 1A). CASK belongs to the membrane-associated guanylate kinase (MAGUK) family of scaffolding proteins and is involved in dendritic spine formation and maturation through the modulation of the actin cytoskeleton via the interaction with the protein 4.1 (Biederer and Südhof, 2001). The protein 4.1 binds spectrin and bridges the association between F-actin and spectrin, stabilizing actin in dendritic spines. Conjugation of SUMO1 to K679 of CASK interferes with the interaction between CASK and protein 4.1 and, consequently, regulates spine density and morphology in developing neurons (Chao et al., 2008).

Among postsynaptic signaling proteins, ARC, which is also ubiquitinated, is the sole SUMO target identified so far (Craig and Henley, 2012; Craig et al., 2012; Figs. 1A, 2). The expression of a constitutively desumoylated ARC selectively prevents the increase of surface AMPARs during TTX-induced synaptic up-scaling, indicating that sumoylation inhibits ARC function. Interestingly, the same treatment also augments the amount of SUMO1-conjugated proteins and decreases SENP1 levels (Craig et al., 2012). Recently, it was demonstrated that sumoylated ARC is more broadly implicated in synaptic plasticity. During LTP consolidation, sumoylated ARC selectively accumulates in synaptosomal and cytoskeletal fractions, where it forms a complex with the F-actin binding protein drebrin A, stabilizing nascent actin filaments in dendritic spines (Nair et al., 2017). These results further support the role of sumoylation as master regulator of different forms of synaptic plasticity.

At inhibitory synapses, gephyrin is extensively subjected to PTMs (Tyagarajan and Fritschy, 2014). In 2016, it was shown that gephyrin sumoylation at K148 and K724 negatively regulates its postsynaptic clustering and, ultimately, reduces synapse formation and inhibitory transmission (Fig. 1C). Gephyrin sumoylation is intimately linked to other PTMs, which synergistically operate to orchestrate its clustering dynamics (Ghosh et al., 2016). Desumoylation of gephyrin at K148 leads to deacetylation at K666 and dephosphorylation at S268 residues, enhancing its postsynaptic clustering and the assembly of inhibitory synapses.

Recently, the presence of SUMO1-conjugated proteins at synapses was questioned, bringing out an intense scientific debate on this topic (Tirard et al., 2012; Daniel et al., 2017, 2018; Wilkinson et al., 2017). By using KI mice expressing His6-HA-SUMO1, Dr. Brose and Dr. Tirard teams were not able to validate sumoylation of seven previously characterized presynaptic and postsynaptic SUMO targets nor to observe SUMO1 conjugation at synapses (Daniel et al., 2017). On the contrary, they could confirm that extra-synaptic/nuclear sumoylation is present (Tirard et al., 2012), concluding that there is no evidence of synaptic SUMO1-ylation. Notwithstanding, several arguments support the presence of SUMO1-ylation at the synapse questioning whether this animal model is suitable to study sumoylation. (1) In His6-HA-SUMO1 KI mice, there is 20–30% less SUMO1 conjugation than WT mice, and this impairment is associated to a significant increase of SUMO2/3-conjugated proteins (Tirard et al., 2012; Daniel et al., 2017). This reduction in the efficiency of SUMO1 conjugation could make particularly challenging the detection of SUMO targets in the synaptic compartment where sumoylation is already low and extremely transient; (2) a wealth of studies confirmed the functional relevance of synaptic sumoylation (e.g., use of non-sumoylatable mutants), while Daniel and colleagues did not carry out this analysis (for review, see Henley et al., 2018; Schorova and Martin, 2016); (3) consistent data clearly indicate the presence of components of the SUMO machinery at synaptic sites (Watanabe et al., 2008; Loriol et al., 2012, 2014; Hasegawa et al., 2014; Josa-Prado et al., 2019; Schorova et al., 2019); (4) some key technical differences may explain the failure to detect synaptic sumoylation (e.g., weak SUMO1 immunostaining and the use of antibodies that do not recognize sumoylated proteins). Although in our opinion substantial data indicate the presence and the functional relevance of synaptic SUMO1-ylation, the aforementioned studies pointed out the need of defining strict consensus criteria to investigate sumoylation in the brain.

Neddylation

Neddylation consists in the covalent and reversible attachment of a NEDD8 (for neural precursor cell expressed, developmentally downregulated 8) moiety to a K residue of target proteins. Of the UBLs, NEDD8 shows the greatest degree of similarity with ubiquitin (∼80%). Like ubiquitination and other UBLs, neddylation requires a three-step enzymatic cascade to activate and covalently attach NEDD8 to K residues of target proteins (Enchev et al., 2015). While E1 activating and E2 conjugating enzymes have been identified (NEDD8 activating enzyme, NAE as E1, and UBC12 and UBE2F as E2 enzymes), the E3 ligases are as yet ill-defined. Deconjugation of NEDD8 from targets is achieved by NEDD8-specific proteases, namely the metalloprotease CSN5, the cysteine protease NEDP1 and USP21 (Enchev et al., 2015).

Although NEDD8 function and regulation remain largely unexplored, recent papers shed light on the importance of neddylation in synaptic maturation, function and plasticity. Neddylation is observed during embryonic brain development, its expression increases during the first two postnatal weeks and is then maintained throughout adulthood (Kumar et al., 1992; Vogl et al., 2015). As the best characterized function of NEDD8 involves the activation of ubiquitin E3 ligases of the Cullin-RING family (Fig. 1A), neddylation is thought a major regulator of ubiquitination (Enchev et al., 2015). In the hippocampus, NMDAR-dependent neddylation promotes the UPS-dependent degradation of the DNA methyltransferase DNMT3a1 leading to demethylation of the BDNF promoter and enabling BDNF expression during memory consolidation (Fig. 1A; Bayraktar et al., 2019). Among the members of the cullin-RING family, several ligases mediate the targeted degradation of synaptic elements and negatively regulate synapses (e.g., cullin3 and parkin). However, inhibition of the neddylation pathway does not enhance synaptic activity and strength as it would have been expected as a consequence of reduced synaptic ubiquitination. Instead, it destabilizes dendritic spines, affects synaptic transmission and plasticity, and impairs cognitive functions (Scudder and Patrick, 2015; Vogl et al., 2015; Brockmann et al., 2019), thus suggesting that neuronal neddylation targets other substrates relevant to synaptic function. In line with this, Vogl and colleagues elegantly demonstrated that PSD-95 is neddylated at K202 by MDM2 (Vogl et al., 2015), which also mediates PSD-95 ubiquitination (Fig. 1A; Colledge et al., 2003; Bianchetta et al., 2011). Although PSD-95 neddylation does not modify its ubiquitination and degradation rate, neddylation reduces synaptic clustering of PSD-95. Furthermore, the expression of a non-neddylatable mutant recapitulates at least some of the molecular and cellular phenotypes observed with PSD-95 knock-down (Ehrlich et al., 2007), indicating that NEDD8-conjugation is required for PSD-95 proactive functions. PSD-95 is a major molecular hub of excitatory synapses and mediates multiple interactions with several proteins (Sheng and Kim, 2011). How neddylation changes PSD-95 conformation and/or affinities with its partners remains to be uncovered. Very recently, a NEDD8-ubiquitin substrate profiling detected 341 neddylated proteins in HEK293 cells expanding the repertoire of neddylated targets and indicating broader roles of neddylation than the sole activation of cullin-RING E3 ligases (Vogl et al., 2020). Among the identified targets, the authors biochemically and functionally characterized cofilin neddylation in the brain and found that it is required to ensure proper dendrite development in mouse cortical neurons (Fig. 1A; Vogl et al., 2020).

ID and Impairment of Ubiquitination and Sumoylation Pathways

ID is a neurodevelopmental disorder with an estimated prevalence of 1–3%. The formal diagnosis of ID is based on the intelligence quotient (IQ) test, which should be scored <70, and the presence of deficits in at least two adaptive behaviors that affect everyday activities. It is classified as mild, moderate, severe, or profound based on IQ score. ID is defined as non-syndromic if the intellectual deficit is the sole clinical feature, or as syndromic if the mental impairment is comorbid with other neurologic pathologies such as epilepsy, sensory alterations and ASD (Ilyas et al., 2020). ID is a complex multifactorial disease, in which environmental and genetic factors and their reciprocal interaction critically contribute to its etiology. Yet, 25% of ID cases have clear genetic origins and some of these forms are monogenic. The environmental factors that underlie ID comprise stressful events occurring at early stages of neurodevelopment, including drug and alcohol abuse and infections during pregnancy, birth complication, and severe malnutrition. Thanks to recent advances in whole-genome sequencing, the identification of genetic defects is becoming increasingly efficient. To date, ∼700 genes were associated to either syndromic or non-syndromic ID. Notably, more than 50% of these genes encode presynaptic or postsynaptic proteins, or proteins implicated in synapse development, function and plasticity (Ilyas et al., 2020). Among the several genes associated with ID, ∼100 genes are located on chromosome (chr.) X and are responsible for X-linked ID (XLID; Ropers and Hamel, 2005).

Numerous ID-associated genes code for proteins targeted by ubiquitination and sumoylation or components of ubiquitin and SUMO machineries. Furthermore, some extrinsic risk factors of ID seem to be associated with alterations of brain ubiquitome and/or sumoylome. As described above, several studies indicate that ubiquitination and sumoylation are critical to the expression of synaptic plasticity, the cellular correlate of learning and memory processes and whose impairment is a major hallmark of ID (Aicardi, 1998). Collectively, this evidence suggests that ubiquitination and sumoylation failure may be implicated in ID pathogenesis. Despite pharmacological or genetic blockade of neddylation results in cognitive dysfunctions that may manifest in ID patients (e.g., impaired memory and sociability; Vogl et al., 2015), current evidence does not indicate a direct involvement of neddylation in the development of ID. Below, we provide an overview of the ID forms that are linked to defective ubiquitin and SUMO pathways.

Ubiquitination and ID

Beyond the well-established role of ubiquitination in synapse physiology, its dysregulation has been largely linked to synaptic dysfunction and ID pathogenesis (Mabb and Ehlers, 2010; Upadhyay et al., 2017; George et al., 2018). As listed in Table 1, causal genes directly encode either components of the ubiquitin machinery or regulators of ubiquitination (PLAA and MAGEL2 genes). A third category, not discussed here for space constraints comprises genes encoding targets of ubiquitination in which pathogenic mutations may change their ubiquitination profile. The molecular mechanisms underlying the function of the first two categories of genes and the neuronal substrates that they target are poorly characterized. For these reasons, the molecular pathogenesis of these forms of ID is unclear.

View this table:
  • View inline
  • View popup
Table 1

Rare monogenic forms of ID linked to mutations of components or regulators of the ubiquitin pathway

In general, ubiquitin-linked ID forms present comorbidities that are commonly present in ID. Mental impairment is often accompanied by dysmorphic features and multiple neurologic and neuropsychiatric manifestations (e.g., epilepsy, motor dysfunction, and autistic behaviors). Many syndromes are also multisystem disorders underscoring the ubiquitous importance of ubiquitination for tissue and body formation and function. To the best of our knowledge, a clear classification of ID forms, and more generally of brain disorders linked to defective ubiquitination, does not exist. Functional and genetic groups may be however elaborated. For instance, a genetic classification may be based on components of the enzymatic cascade mediating the different steps of ubiquitin conjugation/deconjugation (e.g., E2-E3, and DUB enzymes). Functional classes may include ubiquitinating enzymes involved in distinct cellular functions, such as regulation of transcription, trafficking, cell division or proteasome activity. Although other criteria may certainly be applied to propose alternative classifications including comorbidities, no unique behavioral phenotypes associate with any of the aforementioned groups. Overall, it clearly emerges that ubiquitination and its delicate homeostasis are fundamental for brain physiology. Indeed, the perturbation of this equilibrium caused by the altered expression or activity of a single E3 ubiquitin ligase is hardly compensated by the other ∼600 E3 and is sufficient to lead to severe pathologic states of the brain. Here, we discuss in more details the implications of defective ubiquitination in the molecular pathogenesis of two of the most common forms of syndromic ID, the AS and Down syndrome (DS).

Angelman syndrome

AS (OMIM 105830) is a relatively rare neurodevelopmental disorder (prevalence of 1/15,000 live births) characterized by severe intellectual deficit, motor dysfunction, unusually happy demeanor, seizures and autism-like behavior (Clayton-Smith and Laan, 2003; Margolis et al., 2015; Buiting et al., 2016). AS results from distinct genetic abnormalities that ultimately lead to the loss of the brain-specific imprinted UBE3A (ubiquitin protein ligase E3A, also referred to as E6AP) gene (Mabb et al., 2011). UBE3A encodes three isoforms that are generated by alternative splicing and localize either in the nucleus or cytosol (Miao et al., 2013). In the cytosol, UBE3A is found in axonal terminals and dendritic spines (Dindot et al., 2008; Burette et al., 2018). While UBE3A loss results in AS, elevated expression or activity of UBE3A represent the most identifiable genetic form of ASD, indicating that a tight balance in the dosage of this gene is critical to develop functional neuronal circuits.

Data from AS mouse models suggest that UBE3A plays a key role in modulating synaptic pathways important for cognition and behavior. Neurons from Ube3a-KO mice display abnormal dendritic spines, reduced excitatory neurotransmission and impaired synaptic plasticity (Dindot et al., 2008; Yashiro et al., 2009; Pignatelli et al., 2014; Kim et al., 2016; Avagliano Trezza et al., 2019; Wang et al., 2019a). Given the primary function of UBE3A as E3 ubiquitin ligase, defective ubiquitination is thought to be at the basis of neuronal dysfunctions and clinical manifestations of AS. However, only a few substrates have been identified and shown to be functionally relevant to AS etiology. For these reasons, the molecular underpinnings of UBE3A function remain enigmatic. Different mechanisms, mainly driven by the defective ubiquitination hypothesis, have been put forward on the pathogenic role of UBE3A. (1) In physiological conditions, UBE3A may indirectly modulate the number of AMPARs at the postsynaptic membrane by ubiquitinating the aforementioned signaling protein ARC and promoting its degradation (Greer et al., 2010). Therefore, loss of UBE3A may increase ARC levels and lead to enhanced AMPAR internalization and, ultimately, depression of glutamatergic transmission. (2) A second synaptic substrate of UBE3A is ephexin 5 (Margolis et al., 2010). Ephexin 5 is a guanine nucleotide exchange factor (GEF) activator of the RhoA GTPase that limits the number of excitatory synapses via inhibition of the synaptogenic, trans-synaptic EphB receptor/ephrin ligand complex (Henderson and Dalva, 2018). Ube3a-KO mice display increased amounts of ephexin 5 and abnormal density of excitatory synapses, indicating that UBE3A-dependent degradation of ephexin 5 is critical for the normal development of excitatory synapses. (3) A recent study suggested that UBE3A ubiquitinates the small conductance potassium channel SK2, whose major function is to repolarize neuronal membranes after depolarization (Ngo-Anh et al., 2005) and facilitates its endocytosis (Sun et al., 2015b). In line with defective ubiquitination of SK2, its surface expression is significantly increased in Ube3a-KO mice, resulting in membrane hyperpolarization and impaired synaptic plasticity. (4) A role for CAMKII was proposed based on the observation that hippocampi derived from AS mice display reduced CAMKII activity and excessive inhibitory autophosphorylation (Weeber et al., 2003). In particular, the correction of some AS neurologic deficits obtained preventing CAMKII inhibitory autophosphorylation suggests that CAMKII may be instrumental to UBE3A-dependent pathways underlying AS phenotype (van Woerden et al., 2007). However, it is not known whether UBE3A directly targets CAMKII for degradation or targets unknown kinases or phosphatases that in turn modulate CAMKII phosphorylation. Strikingly, PTPA, which is the activator of a major phosphatase of CAMKII, PP2A, was recently identified as UBE3A target (Wang et al., 2019a). However, the modulation of PTPA levels does not affect the phosphorylation of CAMKII, suggesting other downstream substrates. (5) In cerebellar Purkinje cells of AS mice, mechanistic target of rapamycin (mTOR) signaling is unbalanced, resulting in upregulated mTORC1 and reduced mTORC2 (Sun et al., 2015a). mTORC1 activation relies on its lysosomal recruitment through the interaction with the Rag GTPase-Ragulator complex. UBE3A ubiquitinates the p18 subunit of the Ragulator complex. Therefore, in AS mouse models defective p18 ubiquitination enhances mTORC1 activity by increasing its lysosomal recruitment. (6) A recent proteomic-based study aimed at identifying new UBE3A substrates implicated autophagy in AS pathogenesis (Wang et al., 2019b). UBE3A physically associates with and ubiquitinates the autophagy regulator Huntington-associated protein 1 (HAP1), a key protein for autophagosome trafficking. In AS mouse model, increased HAP1 leads to excessive autophagy and ultimately, dendritic spine defects. Consistently, pharmacological attenuation of autophagy partially alleviates synaptic dysfunction and behavioral deficits in AS mice.

The above studies have focused on mechanisms underlying weakened excitatory synapses. However, the vast majority of AS patients display autistic behavior and seizure susceptibility (Clayton-Smith and Laan, 2003), suggesting that loss of UBE3A may increase the excitation (E)/inhibition (I) ratio in the neocortex. Recent evidence proposed that UBE3A loss might impact GABAergic neurons more severely than the excitatory counterpart, with the potential net outcome favoring hyperexcitability (Wallace et al., 2012; Judson et al., 2016). Future studies aimed at investigating whether UBE3A directly operates at inhibitory synapses and which are its substrates will help clarifying UBE3A-mediated mechanisms modulating neuronal excitability.

Together, these studies clearly indicate that UBE3A regulates several cellular pathways that critically contribute to synapse development and function, and that altered ubiquitination is central to AS etiology. Although the number of identified UBE3A substrates is rapidly increasing, the functional relevance of these interactions is not clear yet and an integrated picture of UBE3A function is lacking. In addition, the weight of individual substrates and pathways in the pathogenesis of AS and its clinical manifestations is still enigmatic.

Down syndrome

DS (OMIM 190685) is the most common cause of ID and accounts for ∼15–20% of all individuals affected by ID. DS is a complex genetic disorder characterized by heterogenous clinical manifestations. Among them, cognitive impairment is present in all DS individuals (Antonarakis et al., 2004). The DS critical region (DSCR) is a relatively small locus on chr.21, and its triplication is necessary and sufficient to generate DS cognitive deficits. One of the genes localized in the DSCR is TTC3, which is consistently upregulated in patients and animal models of DS (Saran et al., 2003). TTC3 encodes an E3 ubiquitin ligase that targets a phosphorylated form of the kinase AKT (Suizu et al., 2009), which is critical for several cellular functions in the brain (Hers et al., 2011). In Ts65Dn mice, the most common animal model of DS, the abnormal levels of TTC3 enhance ubiquitination and degradation of AKT. Among its multiple synaptic functions, AKT phosphorylates the β subunits of GABAARs in hippocampal neurons, thereby increasing the number of surface GABAARs at inhibitory synapses (Wang et al., 2003). Remarkably, DS individuals show a higher incidence of seizures, raising the possibility that enhanced TTC3-mediated degradation of AKT leads to the loss of E/I equilibrium in the brain and hyperexcitability. In developing neurons, TTC3 overexpression also limits neurite outgrowth and modifies the morphology of the Golgi apparatus through the modulation of actin-regulating pathways (Berto et al., 2014). Strikingly, TTC3 also significantly correlates with other DS-unrelated brain pathologies associated with ID, suggesting an essential role for TTC3 in complex cognitive functions (Vilardell et al., 2011).

Dysfunctions of global neuronal ubiquitome in DS are also detected in aged Ts65Dn mice and postmortem brains of DS patients. Such alterations are linked to the accumulation of inclusion bodies in the cerebellum, a cellular hallmark of Alzheimer’s disease (AD), a neurodegenerative disorder that is particularly frequent in DS (Necchi et al., 2011; Tramutola et al., 2017). On the other hand, it was reported that in both DS patients and Ts65Dn mice the USP16 gene, which maps on chr.21 and codes for a deubiquitinase, is critical to DS pathogenesis (Adorno et al., 2013). USP16 triplication excessively removes ubiquitin from histone H2A, a key event for the self-renewal and expansion of different progenitor cells, including neuronal progenitors. In line with this, a reduced expansion of postnatal neuronal progenitors is observed in DS patients.

All of these pieces of evidence suggest that unbalanced ubiquitination critically contributes to neuronal circuit dysregulation and clinical manifestations of DS, including ID. The relative contribution of these pathways to the molecular pathogenesis of DS remains largely unknown.

Sumoylation and ID

As aforementioned, sumoylation is critical to build up proper synaptic connections by modulating the function of several neuronal proteins. The evidence that several SUMO targets are ID-associated proteins raises the possibility that impaired sumoylation may be relevant to ID etiology. An increasing number of molecular studies now supports this hypothesis and suggests that dysregulated neuronal sumoylation may participate to the development of two common syndromic forms of ID, the RTT and DS. Furthermore, defective sumoylation may potentially be implicated in other ID forms, such as Fragile X syndrome (FXS), Phelan–McDermid syndrome (PMS) and the Cc2d1-dependent non-syndromic ID.

Rett syndrome

RTT (OMIM 312750) is a devastating neurodevelopmental disorder and the leading cause of genetic ID in young girls. Almost 95% of RTT individuals carry mutations in the MECP2 gene. As described above, MECP2 is sumoylated at K223 and K412 residues and these modifications regulate the activity of MECP2 as transcriptional repressor. The analysis of sumoylation profiles of seven different mutated MECP2 variants identified in RTT patients reveals that six of these mutations decrease MECP2 sumoylation because of a lower affinity for the SUMO E3 ligase PIAS1 (Tai et al., 2016). Strikingly, while the lentiviral expression of either WT or sumoylated MECP2 in conditional Mecp2-KO mice restores some cellular and behavioral deficits, such as social interaction, fear memory and LTP, the expression of a non-sumoylatable MECP2 mutant fails to rescue these phenotypes. In this study, it was also demonstrated that these pathologic mutations are associated with altered MECP2 phosphorylation at S421, a PTM that is known to promote its sumoylation. In the same study, the authors also demonstrated that mice expressing non-sumoylatable MECP2 display 12-fold decrease in Wnt6 mRNA levels compared to WT animals. Very recently, the same group showed that lentiviral expression of WNT6 rescues defective MECP2 sumoylation in MECP2 T158A mouse model of RTT. Wnt6 transduction and enhanced MECP2 sumoylation also resulted in partial amelioration of locomotor and social behavioral deficits (Hsu et al., 2020). Together, these pioneer studies provided for the first time a link between altered sumoylation and the etiology of ID, representing an important milestone in the field.

ID linked to SUMO machinery-encoding genes

Since the SUMO3 paralogue gene is located in the long arm of chr.21, the excessive dosage of SUMO3 is an intriguing hypothesis in DS pathogenesis. Several transcriptional factors and the glucocorticoid receptor, which are known to control cognitive functions, are targeted by SUMO3 (Schorova and Martin, 2016). In line with this, the levels of free and conjugated SUMO2/3 proteins are significantly higher in hippocampal lysates derived from DS patients compared with healthy individuals (Gardiner, 2006). Yet, a comprehensive view of SUMO3 neuronal targets is lacking and the impact of SUMO3 triplication on DS etiology remains elusive. Recently, Binda and colleagues reported a marked decrease of the SUMO2/3-specific deconjugating enzyme SENP3 in DS patients without any corresponding increase of SUMO2/3-conjugated proteins (Binda et al., 2017). Future studies aiming at systematically analyzing SUMO3 substrates in the brain will be of invaluable help to elucidate the role of enhanced sumoylation in DS pathogenesis.

A woman with profound developmental delay associated with several other clinical manifestations was found to carry a microdeletion in the chr.19 that resulted in the haploinsufficiency of 15 genes, including SAE1, a subunit of the SUMO1 activating enzyme (Leal et al., 2009). Although the authors suggested that the loss of SAE1 gene may correlate with an impaired development of the cleft lip and palate, it remains to be determined whether this alteration might also participate to other pathologic features, comprising defective cognitive functions.

ZMIZ1 gene codes for an androgen receptor coactivator that functions as an E3 SUMO ligase (Sharma et al., 2003) and recently identified mutations were associated with ID and developmental delay (Carapito et al., 2019). Notably, all identified ZMIZ1 mutations occur in the SUMO acceptor site, suggesting that they might compromise the E3 SUMO ligase activity of ZMIZ1, resulting in a failure of sumoylation homeostasis in neurons.

Cc2d1-dependent non-syndromic ID

Mutations in the Coiled-coil and C2 domain containing 1A (CC2D1A) gene are associated with non-syndromic ID (OMIM 608443). This gene codes for CC2D1A, a DNA binding protein that regulates multiple cellular signaling pathways, including serotonin and dopamine receptors, dendritic arborization, synapse maturation, and plasticity (Zhao et al., 2011; Al-Tawashi et al., 2012; Manzini et al., 2014; Oaks et al., 2017; Yang et al., 2019). Consistent with these roles, the loss of CC2D1A in vivo determines the appearance of cognitive and social deficits (Oaks et al., 2017). Very recently, it was shown that conditional deletion of Cc2d1a in excitatory neurons of the forebrain leads to decreased levels of SENP1 and SENP3. As a consequence, the desumoylation of the small GTPase RAC1, one of the major targets of SENP1 and SENP3, is suppressed (Yang et al., 2019). Since RAC1 activity is enhanced by sumoylation, hyperactived RAC1 results in abnormal dendritic morphology and synaptic dysfunction. Remarkably, pharmacological blockade of RAC1 activity partially rescues deficits in synaptic plasticity and memory observed in Cc2d1a-cKO mice (Yang et al., 2019). In conclusion, this work provides further experimental evidence supporting the hypothesis that aberrant sumoylation is a key molecular determinant underlying ID etiology.

mGlu5-related IDs: Phelan–McDermid and Fragile X syndromes

PMS and FXS are two major forms of ID and are associated with dysregulation of the mGlu5-dependent pathway. The loss of SHANK3 gene, which is caused by deletions of the terminal end of chr.22 long arm, is critical to PMS development. SHANK3 is a major scaffolding protein of excitatory synapses that anchors, together with HOMER1, mGlu5 to the postsynaptic membrane. As a consequence, the loss of Shank3 impairs mGlu5 activity by destabilizing the pool of surface receptors (Vicidomini et al., 2017). Conversely, FXS mouse models carrying loss-of-function mutations of Fmr1 gene are characterized by excessive mGlu5-dependent signaling (Prieto et al., 2020).

Considering the functional relevance of mGlu5 to SUMO homeostasis (Loriol et al., 2014; Schorova et al., 2019), the analysis of brain sumoylome in animal models of PMS and FXS is of primary importance and might uncover pivotal roles of sumoylation in the molecular pathogenesis of these neurodevelopmental disorders. Recently, new evidence further supported the possibility that sumoylation might be a key determinant in the pathogenesis of FXS (Khayachi et al., 2018; Prieto et al., 2020). The novel pathologic missense mutation R138Q of FMR1 gene was identified in three unrelated individuals presenting developmental delays, ID, and seizures (Collins et al., 2010; Diaz et al., 2018; Sitzmann et al., 2018). In contrast to other FXS cases, this mutation does not affect the expression levels of FMRP protein. As the R138Q mutation localizes in close proximity to the SUMO site K130 of FMRP, it is possible that the R138Q mutation causes aberrant FMRP sumoylation, and this impairment may be causative of FXS. Similarly, the F126S mutation, which was reported in a FXS male patient presenting ID and autistic features (Quartier et al., 2017), also localizes nearby K130 residue. As this mutation replaces a phenylalanine (F) with an S residue, it likely generates a novel phosphorylation site, which could subsequently promote sumoylation at K130 (Loriol et al., 2014). Further studies are required to better explore this hypothesis.

Functional Interplay Between Ubiquitination and UBLs

Recently, the existence of a dynamic interplay between ubiquitination and UBLs provided an intriguing perspective that could lead to the identification of novel principles underlying synapse development and function. As mentioned above, the best characterized function of NEDD8 is to enhance the activity of Cullin E3 ligases providing evidence of a direct interaction with ubiquitination (Enchev et al., 2015). Moreover, neddylation-dependent ubiquitination and degradation of DNMT3a1 is critical to increase BDNF expression and is required to memory consolidation in the hippocampus (Bayraktar et al., 2019). PSD-95 is targeted by the E3 ligase MDM2, which mediates both ubiquitination and neddylation (Colledge et al., 2003; Bianchetta et al., 2011; Vogl et al., 2015). However, neddylation does not change PSD-95 degradation rate (Vogl et al., 2015), and the possible cross talk between these two PTMs on PSD-95 remains unknown. Altogether, more study is needed to better understand neddylation and its interplay with other PTMs.

Emerging evidence also indicates that ubiquitination and sumoylation reciprocally interact to precisely regulate protein function. Numerous neuronal substrates are targeted by both ubiquitin and SUMO machineries. Proteomic approaches showed that ∼25% of SUMO-acceptor K residues are also ubiquitinated, suggesting that these two pathways likely compete with each other for the same K residues (Tammsalu et al., 2014). However, a few examples suggest that this interplay is not only antagonistic, but ubiquitination and sumoylation can also operate in synergy. The hypoxia-inducible factor 1 α (HIF1α) requires to be sumoylated for its subsequent UPS-dependent degradation (Cheng et al., 2007). Similarly, the sequential sumoylation and ubiquitination of the serine hydroxymethyltransferase 1 (SHMT1) and the regulatory subunit of the kinase IKK are required to control the rate of nuclear import/export of these proteins and, ultimately, to regulate their activity (Huang et al., 2003; Anderson et al., 2012). Moreover, the impairment of ubiquitination/sumoylation cross talk critically contributes to the development of neurodegenerative disorders. For instance, in AD, this interplay is critical for the formation of tau-containing neurofibrillary tangles, a primary hallmark of AD. Indeed, hyperphosphorylation of tau protein enhances its sumoylation in a β-amyloid (Aβ)-dependent manner. Once sumoylated, tau is less ubiquitinated and degraded, favoring tau aggregation and decreasing its solubility (Luo et al., 2014). Another neurodegenerative disorder in which the competition between ubiquitination and sumoylation is relevant for disease pathogenesis is Huntington’s disease (HD). HD is caused by the expansion of huntingtin (HTT) polyglutamine repeats. Proteolytic cleavage of mutated HTT generates a pathogenic fragment (HTTEX1P) that creates toxic aggregates in the nucleus and cytoplasm. Both ubiquitination and sumoylation target K6 and K9 of the HTTEX1P fragment and mediate opposite effects. Sumoylation increases HTTEX1P stability and aggregation, its function as transcriptional repressor and ultimately, its neurotoxicity (Steffan et al., 2004). Conversely, ubiquitination reduces HTTEX1P neurotoxicity probably by promoting its degradation (Kalchman et al., 1996). Consistent with this, in a Drosophila model of HD, sumoylation and ubiquitination of HTTEX1P worsen and abrogate neurodegeneration, respectively (Steffan et al., 2004).

To date, the functional interplay between ubiquitination and sumoylation in ID pathogenesis has been poorly characterized. Growing evidence suggests that a defective cross talk might underlie, at least partially, synaptic dysfunction observed in certain forms of ID. As described above, SUMO3, the E3 ubiquitin ligase TTC3 and the deubiquinating enzyme USP16 are located within the DSCR and are triplicated in DS individuals (Adorno et al., 2013; Gardiner, 2006; Saran et al., 2003). The hyperabundance of SUMO3 could either decrease the ubiquitination of those substrates for which ubiquitination and sumoylation compete or enhance the ubiquitination of those targets for which the two PTMs operate in synergy. Triplication of TTC3 may also lead to an unbalanced ubiquitination/sumoylation that selectively involves TTC3 substrates. For instance, AKT, a major target of TTC3, is also a known SUMO target, raising the possibility that enhanced AKT ubiquitination impairs its sumoylation status in DS patients. Whether other TTC3 substrates are subjected to the same regulatory mechanism is not known as the majority of TTC3 targets are not identified yet. Thus, besides assessing the level of AKT sumoylation in DS, it is of great interest to identify novel TTC3 substrates and investigate whether they are also targeted by sumoylation. A second ID form in which unbalanced ubiquitination/sumoylation could potentially play a central role is FXS. As mentioned above, novel pathologic missense point mutations in the FRM1 gene are closely located to the SUMO site K130 and likely interfere with FMRP sumoylation (Khayachi et al., 2018; Prieto et al., 2020). Furthermore, FMRP is also subjected to activity-dependent ubiquitination by the APC/C complex (Hou et al., 2006; Huang et al., 2015). If ubiquitin and SUMO machineries competed for the same K residues in FMRP sequence, defective sumoylation may promote its ubiquitination. Alternatively, if sumoylation, which is required to release FMRP from dendritic mRNA granules (Khayachi et al., 2018), is instrumental to subsequent ubiquitination, these mutations may affect FMRP degradation, resulting in reduced local translation. Finally, it is also tempting to consider this possibility in regard to the AS. Indeed, ARC, a major synaptic substrate of UBE3A, is also subjected to sumoylation (Nair et al., 2017). Together, the involvement of an unbalanced cross talk between ubiquitination and sumoylation in synaptic dysfunction of ID is an exciting hypothesis and, if confirmed by further experimental data, it would define novel principles underlying the molecular logistics of synapse physiology and the pathogenesis of a spectrum of neurodevelopmental and psychiatric disorders.

Conclusions and Perspectives

As summarized in this review, a wealth of studies clearly demonstrated pivotal roles for ubiquitination and UBLs in the assembly and refinement of neuronal circuits, maintenance of neuronal homeostasis and the emergence of complex cognitive functions. The fine spatiotemporal regulation of these PTMs is of primary importance during brain formation, a process that is characterized by intense synaptic plasticity and, on the other end, by high susceptibility to toxic insults. Therefore, genetic or environmental challenges to ubiquitination and sumoylation pathways critically contribute to the etiology of synaptopathies, including ID. Concerning neddylation, although it might be involved in ID, no evidence is currently available to support this hypothesis. The molecular mechanisms that underpin impaired ubiquitination and sumoylation in ID are unclear as well as a comprehensive view of neuronal ubiquitome and sumoylome is lacking. Here, we provide a few outlooks that are crucial to better understand the role of these two regulatory systems in the brain.

A major limitation to advance our knowledge on the pathophysiological role of ubiquitination and UBLs concerns the technical challenges to detect endogenous substrates. These modifications are transient and restricted to specific subcellular compartments, developmental stages, and states of neuronal activity. This results in scarcely abundant steady-state levels of modified targets. In recent years, new MS-based strategies were developed to efficiently capture ubiquitinated, sumoylated, and neddylated proteins in native tissues. For example, ubiquitomes and sumoylomes may be enriched using a novel monoclonal antibody that recognizes a di-glycine tag on K residues of trypsinized peptides that is present on ubiquitinated and sumoylated proteins solely (Na et al., 2012; Tammsalu et al., 2015). This technique was recently combined with CRISPR-Cas9 genome editing of NEDD8 gene to reveal the neddylome in HEK cells (Vogl et al., 2020). Alternatively, the in vivo proximity-dependent biotin identification (iBioID) approach may efficiently detect substrates of specific ubiquitin and UBL machineries (Coyaud et al., 2015; Pirone et al., 2017).

A fertile area of future research will be the investigation of environmental risk factors for ID that might affect neuronal ubiquitination and UBLs. As previously mentioned, ID-linked environmental factors comprise alcohol and drug abuse and infections during pregnancy, birth complication and severe malnutrition. For instance, prenatal exposure to one of the most commonly used anticonvulsant drugs, the valproic acid, is associated with an increased risk of autism (Christensen et al., 2013) and a few reports suggest that it may regulate the UPS (Kwon et al., 2013). Microarray transcript and proteome profiling of mouse models of fetal alcohol syndrome disorders (FASD) revealed a significant downregulation of the proteasomal function (Green et al., 2007; Mason et al., 2012). In particular, the E2 conjugating enzyme UBE2N, which is known to be critical for neurodevelopment (Muralidhar and Thomas, 1993), is among the identified proteins. Similarly, fetal cortices exposed to alcohol display an increased sumoylation of the Heat shock protein 1 (HSP1), a fundamental cellular sensor of environmental proteotoxic stress, resulting in prolonged activation of HSP1 (El Fatimy et al., 2014). The E3 SUMO ligase PIASy is upregulated in alcohol-treated cultured cells and drives the induction of autophagy (Ran et al., 2018), a well-known mechanism involved in synapse assembly and maturation (Tomoda et al., 2020). Yet, whether alcohol-dependent PIASy upregulation could contribute to synaptic dysfunction in FASD is not known. Perinatal asphyxia (PA) is an obstetric complication derived from impaired gas exchange that inevitably leads to aberrant synaptic networks and severe clinical manifestations, such as epilepsy, schizophrenia, and ID. A number of studies reported enhanced ubiquitination levels in striatal and hippocampal synapses of PA animal models (Capani et al., 2009; Saraceno et al., 2012). Furthermore, hypoxia triggers sumoylation of NaV1.2 channels (Plant et al., 2016), which provides the earliest neuronal response to hypoxia and is a major determinant of hypoxia-induced neuronal death (Weber and Taylor, 1994). Since TTX-induced inhibition of sodium currents prevents neuronal death in acute hypoxia (Taylor et al., 1995), modulation of NaV1.2 sumoylation might represent a novel target to develop neuroprotective therapies (Plant et al., 2016). Finally, nicotine administration at concentrations achieved by smokers inhibits the UPS in the prefrontal cortex of adult mice and influences synaptic plasticity by regulating the levels of multiple PSD proteins, including scaffolding molecules and receptors (Rezvani et al., 2007, 2012). However, the direct effect of smoking during pregnancy on the synaptic ubiquitome of fetal brains remains elusive. Altogether, these examples support the hypothesis that ID-linked environmental risk factors affect ubiquitin and UBLs pathways, which in turn may contribute to ID pathogenesis. As these factors typically alter multiple cellular pathways, it would be of great interest to dissect the weight of defective ubiquitination and UBLs to ID onset.

At present, no effective medical therapies are available to treat ID. Defining the roles of ubiquitination and UBL modifications and the molecular mechanisms underlying these pathways in physiological and pathologic conditions could open new venues to identify novel therapeutic strategies. Given that ubiquitination and UBLs virtually regulate all cellular processes, they are attractive drug targets. For example, compounds that modulate ubiquitination, sumoylation, and neddylation have been tested to treat cancer and other pathologies (Landré et al., 2014; Nawrocki et al., 2012; Bernstock et al., 2018). However, the majority of the molecules that are currently being tested activates or inhibits global cellular ubiquitination or UBL modifications, thus inevitably leading to relevant side effects. Therefore, the development of new compounds that precisely modulate individual components or branches of ubiquitin and UBL systems and their interaction with specific substrates is urgently needed. Finally, a better elucidation of synaptic ubiquitination and UBL pathways will be of invaluable help to move toward precision medicine.

Acknowledgments

Acknowledgements: We thank Michela Matteoli (IN-CNR and Hunimed, Rozzano, Italy), Davide Pozzi (Hunimed, Rozzano, Italy), and Marta Prieto (IPMC, Valbonne, France) for critically reading of this manuscript.

Footnotes

  • The authors declare no competing financial interests.

  • This work was supported by the Italian Ministry of Health (GR-2018-12366478 to M.F.) and the European Union’s Horizon 2020 research and innovation programme (Marie Sklodowska-Curie grant agreement 845466 to A.F.).

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

References

  1. ↵
    Adorno M, Sikandar S, Mitra SS, Kuo A, Di Robilant BN, Haro-Acosta V, Ouadah Y, Quarta M, Rodriguez J, Qian D, Reddy VM, Cheshier S, Garner CC, Clarke MF (2013) Usp16 contributes to somatic stem-cell defects in Down’s syndrome. Nature 501:380–384. doi:10.1038/nature12530 pmid:24025767
    OpenUrlCrossRefPubMed
  2. ↵
    Ageta H, Kato A, Fukazawa Y, Inokuchi K, Sugiyama H (2001) Effects of proteasome inhibitors on the synaptic localization of Vesl-1S/Homer-1a proteins. Brain Res Mol Brain Res 97:186–189. doi:10.1016/s0169-328x(01)00303-5 pmid:11750075
    OpenUrlCrossRefPubMed
  3. Aggarwal S, Das Bhowmik A, Ramprasad VL, Murugan S, Dalal A (2016) A splice site mutation in HERC1 leads to syndromic intellectual disability with macrocephaly and facial dysmorphism: further delineation of the phenotypic spectrum. Am J Med Genet A 170:1868–1873. doi:10.1002/ajmg.a.37654 pmid:27108999
    OpenUrlCrossRefPubMed
  4. ↵
    Aicardi J (1998) The etiology of developmental delay. Semin Pediatr Neurol 5:15–20. doi:10.1016/s1071-9091(98)80013-2 pmid:9548636
    OpenUrlCrossRefPubMed
  5. ↵
    Al-Tawashi A, Jung SY, Liu D, Su B, Qin J (2012) Protein implicated in nonsyndromic mental retardation regulates protein kinase A (PKA) activity. J Biol Chem 287:14644–14658. doi:10.1074/jbc.M111.261875 pmid:22375002
    OpenUrlAbstract/FREE Full Text
  6. ↵
    Anderson DD, Eom JY, Stover PJ (2012) Competition between sumoylation and ubiquitination of serine hydroxymethyltransferase 1 determines its nuclear localization and its accumulation in the nucleus. J Biol Chem 287:4790–4799. doi:10.1074/jbc.M111.302174 pmid:22194612
    OpenUrlAbstract/FREE Full Text
  7. ↵
    Ang XL, Seeburg DP, Sheng M, Harper JW (2008) Regulation of postsynaptic RapGAP SPAR by polo-like kinase 2 and the SCFβ-TRCP ubiquitin ligase in hippocampal neurons. J Biol Chem 283:29424–29432. doi:10.1074/jbc.M802475200 pmid:18723513
    OpenUrlAbstract/FREE Full Text
  8. ↵
    Antonarakis SE, Lyle R, Dermitzakis ET, Reymond A, Deutsch S (2004) Chromosome 21 and Down syndrome: from genomics to pathophysiology. Nat Rev Genet 5:725–738. doi:10.1038/nrg1448 pmid:15510164
    OpenUrlCrossRefPubMed
  9. ↵
    Arancibia-Cárcamo IL, Yuen EY, Muir J, Lumb MJ, Michels G, Saliba RS, Smart TG, Yan Z, Kittler JT, Moss SJ (2009) Ubiquitin-dependent lysosomal targeting of GABAA receptors regulates neuronal inhibition. Proc Natl Acad Sci USA 106:17552–17557. doi:10.1073/pnas.0905502106 pmid:19815531
    OpenUrlAbstract/FREE Full Text
  10. ↵
    Arimura N, Menager C, Fukata Y, Kaibuchi K (2004) Role of CRMP-2 in neuronal polarity. J Neurobiol 58:34–47. doi:10.1002/neu.10269 pmid:14598368
    OpenUrlCrossRefPubMed
  11. ↵
    Atkin G, Moore S, Lu Y, Nelson RF, Tipper N, Rajpal G, Hunt J, Tennant W, Hell JW, Murphy GG, Paulson H (2015) Loss of F-box only protein 2 (Fbxo2) disrupts levels and localization of select NMDA receptor subunits, and promotes aberrant synaptic connectivity. J Neurosci 35:6165–6178. doi:10.1523/JNEUROSCI.3013-14.2015 pmid:25878288
    OpenUrlAbstract/FREE Full Text
  12. Au PYB, Huang L, Broley S, Gallagher L, Creede E, Lahey D, Ordorica S, Mina K, Boycott KM, Baynam G, Dyment DA (2017) Two females with mutations in USP9X highlight the variable expressivity of the intellectual disability syndrome. Eur J Med Genet 60:359–364. doi:10.1016/j.ejmg.2017.03.013 pmid:28377321
    OpenUrlCrossRefPubMed
  13. ↵
    Avagliano Trezza R, Sonzogni M, Bossuyt SNV, Zampeta FI, Punt AM, van den Berg M, Rotaru DC, Koene LMC, Munshi ST, Stedehouder J, Kros JM, Williams M, Heussler H, de Vrij FMS, Mientjes EJ, van Woerden GM, Kushner SA, Distel B, Elgersma Y (2019) Loss of nuclear UBE3A causes electrophysiological and behavioral deficits in mice and is associated with Angelman syndrome. Nat Neurosci 22:1235–1247. doi:10.1038/s41593-019-0425-0 pmid:31235931
    OpenUrlCrossRefPubMed
  14. ↵
    Bacon C, Rappold GA (2012) The distinct and overlapping phenotypic spectra of FOXP1 and FOXP2 in cognitive disorders. Hum Genet 131:1687–1698. doi:10.1007/s00439-012-1193-z pmid:22736078
    OpenUrlCrossRefPubMed
  15. Bainbridge MN, Hu H, Muzny DM, Musante L, Lupski JR, Graham BH, Chen W, Gripp KW, Jenny K, Wienker TF, Yang Y, Sutton VR, Gibbs RA, Ropers HH (2013) De novo truncating mutations in ASXL3 are associated with a novel clinical phenotype with similarities to Bohring-Opitz syndrome. Genome Med 5:11. doi:10.1186/gm415 pmid:23383720
    OpenUrlCrossRefPubMed
  16. Basel-Vanagaite L, Dallapiccola B, Ramirez-Solis R, Segref A, Thiele H, Edwards A, Arends MJ, Miró X, White JK, Désir J, Abramowicz M, Dentici ML, Lepri F, Hofmann K, Har-Zahav A, Ryder E, Karp NA, Estabel J, Gerdin AKB, Podrini C, et al. (2012) Deficiency for the ubiquitin ligase UBE3B in a blepharophimosis-ptosis-intellectual-disability syndrome. Am J Hum Genet 91:998–1010. doi:10.1016/j.ajhg.2012.10.011 pmid:23200864
    OpenUrlCrossRefPubMed
  17. ↵
    Bayraktar G, Yuanxiang P, Gomes GM, Confettura AD, Raza SA, Stork O, Tajima S, Yildirim F, Kreutz MR (2019) Synaptic control of DNA-methylation involves activity-dependent degradation of DNMT3a1 in the nucleus. bioRxiv. doi: https://doi.org/10.1101/602151.
  18. ↵
    Bedford FK, Kittler JT, Muller E, Thomas P, Uren JM, Merlo D, Wisden W, Triller A, Smart TG, Moss SJ (2001) GABAA receptor cell surface number and subunit stability are regulated by the ubiquitin-like protein Plic-1. Nat Neurosci 4:908–916. doi:10.1038/nn0901-908 pmid:11528422
    OpenUrlCrossRefPubMed
  19. ↵
    Bernstock J, Ye D, Lee Y, Gessler F, Friedman G, Zheng W, Hallenbeck J (2018) Drugging SUMOylation for neuroprotection and oncotherapy. Neural Regen Res 13:415–416. doi:10.4103/1673-5374.228718 pmid:29623920
    OpenUrlCrossRefPubMed
  20. ↵
    Berto GE, Iobbi C, Camera P, Scarpa E, Iampietro C, Bianchi F, Gai M, Sgrò F, Cristofani F, Gärtner A, Dotti CG, Di Cunto F (2014) The DCR protein TTC3 affects differentiation and Golgi compactness in neurons through specific actin-regulating pathways. PLoS One 9:e93721. doi:10.1371/journal.pone.0093721 pmid:24695496
    OpenUrlCrossRefPubMed
  21. ↵
    Bianchetta MJ, Lam TKT, Jones SN, Morabito MA (2011) Cyclin-dependent kinase 5 regulates PSD-95 ubiquitination in neurons. J Neurosci 31:12029–12035. doi:10.1523/JNEUROSCI.2388-11.2011 pmid:21849563
    OpenUrlAbstract/FREE Full Text
  22. ↵
    Biederer T, Südhof TC (2001) CASK and protein 4.1 support F-actin nucleation on neurexins. J Biol Chem 276:47869–47876. doi:10.1074/jbc.M105287200 pmid:11604393
    OpenUrlAbstract/FREE Full Text
  23. ↵
    Binda CS, Heimann MJ, Duda JK, Muller M, Henley JM, Wilkinson KA (2017) Analysis of protein sumoylation and SUMO pathway enzyme levels in Alzheimer’s disease and down’s syndrome. Opera Medica Physiol 3:19–24.
    OpenUrl
  24. ↵
    Bingol B, Schuman EM (2006) Activity-dependent dynamics and sequestration of proteasomes in dendritic spines. Nature 441:1144–1148. doi:10.1038/nature04769 pmid:16810255
    OpenUrlCrossRefPubMed
  25. ↵
    Bingol B, Wang C-F, Arnott D, Cheng D, Peng J, Sheng M (2010) Autophosphorylated CaMKIIα acts as a scaffold to recruit proteasomes to dendritic spines. Cell 140:567–578. doi:10.1016/j.cell.2010.01.024 pmid:20178748
    OpenUrlCrossRefPubMed
  26. ↵
    Brockmann MM, Döngi M, Einsfelder U, Körber N, Refojo D, Stein V (2019) Neddylation regulates excitatory synaptic transmission and plasticity. Sci Rep 9:1–10. doi:10.1038/s41598-019-54182-2
    OpenUrlCrossRefPubMed
  27. Budny B, Badura-Stronka M, Materna-Kiryluk A, Tzschach A, Raynaud M, Latos-Bielenska A, Ropers H (2010) Novel missense mutations in the ubiquitination-related gene UBE2A cause a recognizable X-linked mental retardation syndrome. Clin Genet 77:541–551. doi:10.1111/j.1399-0004.2010.01429.x pmid:20412111
    OpenUrlCrossRefPubMed
  28. ↵
    Buiting K, Williams C, Horsthemke B (2016) Angelman syndrome — insights into a rare neurogenetic disorder. Nat Rev Neurol 12:584–593. doi:10.1038/nrneurol.2016.133 pmid:27615419
    OpenUrlCrossRefPubMed
  29. ↵
    Burbea M, Dreier L, Dittman JS, Grunwald ME, Kaplan JM (2002) Ubiquitin and AP180 regulate the abundance of GLR-1 glutamate receptors at postsynaptic elements in C. elegans. Neuron 35:107–120. doi:10.1016/S0896-6273(02)00749-3
    OpenUrlCrossRefPubMed
  30. ↵
    Burette AC, Judson MC, Li AN, Chang EF, Seeley WW, Philpot BD, Weinberg RJ (2018) Subcellular organization of UBE3A in human cerebral cortex. Mol Autism 9:54. doi:10.1186/s13229-018-0238-0 pmid:30364390
    OpenUrlCrossRefPubMed
  31. ↵
    Bürkle A (2001) Posttranslational modification. In: Encyclopedia of Genetics. Amsterdam: Elsevier.
  32. ↵
    Campbell MK, Sheng M (2018) USP8 deubiquitinates SHANK3 to control synapse density and SHANK3 activity-dependent protein levels. J Neurosci 38:5289–5301. doi:10.1523/JNEUROSCI.3305-17.2018 pmid:29735556
    OpenUrlAbstract/FREE Full Text
  33. ↵
    Capani F, Saraceno GE, Botti V, Aon-Bertolino L, Madureira de Oliveira D, Barreto G, Galeano P, Giraldez-Alvarez LD, Coirini H (2009) Protein ubiquitination in postsynaptic densities after hypoxia in rat neostriatum is blocked by hypothermia. Exp Neurol 219:404–413. doi:10.1016/j.expneurol.2009.06.007 pmid:19555686
    OpenUrlCrossRefPubMed
  34. ↵
    Carapito R, Ivanova EL, Morlon A, Meng L, Molitor A, Erdmann E, Kieffer B, Pichot A, Naegely L, Kolmer A, Paul N, Hanauer A, Tran Mau-Them F, Jean-Marçais N, Hiatt SM, Cooper GM, Tvrdik T, Muir AM, Dimartino C, Chopra M, et al. (2019) ZMIZ1 variants cause a syndromic neurodevelopmental disorder. Am J Hum Genet 104:319–330. doi:10.1016/j.ajhg.2018.12.007 pmid:30639322
    OpenUrlCrossRefPubMed
  35. ↵
    Castillo PE, Schoch S, Schmitz F, Südhof TC, Malenka RC (2002) RIM1α is required for presynaptic long-term potentiation. Nature 415:327–330. doi:10.1038/415327a pmid:11797010
    OpenUrlCrossRefPubMed
  36. ↵
    Castillo PE, Younts TJ, Chávez AE, Hashimotodani Y (2012) Endocannabinoid signaling and synaptic function. Neuron 76:70–81. doi:10.1016/j.neuron.2012.09.020 pmid:23040807
    OpenUrlCrossRefPubMed
  37. ↵
    Chamberlain SEL, González-González IM, Wilkinson KA, Konopacki FA, Kantamneni S, Henley JM, Mellor JR (2012) SUMOylation and phosphorylation of GluK2 regulate kainate receptor trafficking and synaptic plasticity. Nat Neurosci 15:845–852. doi:10.1038/nn.3089 pmid:22522402
    OpenUrlCrossRefPubMed
  38. ↵
    Chao HW, Hong CJ, Huang TN, Lin YL, Hsueh YP (2008) SUMOylation of the MAGUK protein CASK regulates dendritic spinogenesis. J Cell Biol 182:141–155. doi:10.1083/jcb.200712094 pmid:18606847
    OpenUrlAbstract/FREE Full Text
  39. ↵
    Cheng J, Kang X, Zhang S, Yeh ETH (2007) SUMO-specific protease 1 is essential for stabilization of HIF1α during hypoxia. Cell 131:584–595. doi:10.1016/j.cell.2007.08.045 pmid:17981124
    OpenUrlCrossRefPubMed
  40. ↵
    Cheng J, Huang M, Zhu Y, Xin YJ, Zhao YK, Huang J, Yu JX, Zhou WH, Qiu Z (2014) SUMOylation of MeCP2 is essential for transcriptional repression and hippocampal synapse development. J Neurochem 128:798–806. doi:10.1111/jnc.12523 pmid:24188180
    OpenUrlCrossRefPubMed
  41. ↵
    Chin LS, Vavalle JP, Li L (2002) Staring, a novel EIII ubiquitin-protein ligase that targets syntaxin 1 for degradation. J Biol Chem 277:35071–35079. doi:10.1074/jbc.M203300200 pmid:12121982
    OpenUrlAbstract/FREE Full Text
  42. ↵
    Choi JH, Park JY, Park SP, Lee H, Han S, Park KH, Suh YH (2016) Regulation of mGluR7 trafficking by SUMOylation in neurons. Neuropharmacology 102:229–235. doi:10.1016/j.neuropharm.2015.11.021 pmid:26631532
    OpenUrlCrossRefPubMed
  43. ↵
    Choquet D, Triller A (2013) The dynamic synapse. Neuron 80:691–703. doi:10.1016/j.neuron.2013.10.013 pmid:24183020
    OpenUrlCrossRefPubMed
  44. ↵
    Christensen J, Grønborg TK, Sørensen MJ, Schendel D, Parner ET, Pedersen LH, Vestergaard M (2013) Prenatal valproate exposure and risk of autism spectrum disorders and childhood autism. JAMA 309:1696–1703. doi:10.1001/jama.2013.2270 pmid:23613074
    OpenUrlCrossRefPubMed
  45. ↵
    Clayton-Smith J, Laan L (2003) Angelman syndrome: a review of the clinical and genetic aspects. J Med Genet 40:87–95. doi:10.1136/jmg.40.2.87 pmid:12566516
    OpenUrlAbstract/FREE Full Text
  46. ↵
    Coba MP (2019) Regulatory mechanisms in postsynaptic phosphorylation networks. Curr Opin Struct Biol 54:86–94. doi:10.1016/j.sbi.2019.01.003 pmid:30807903
    OpenUrlCrossRefPubMed
  47. ↵
    Colledge M, Snyder EM, Crozier RA, Soderling JA, Jin Y, Langeberg LK, Lu H, Bear MF, Scott JD (2003) Ubiquitination regulates PSD-95 degradation and AMPA receptor surface expression. Neuron 40:595–607. doi:10.1016/S0896-6273(03)00687-1
    OpenUrlCrossRefPubMed
  48. ↵
    Collins SC, Bray SM, Suhl JA, Cutler DJ, Coffee B, Zwick ME, Warren ST (2010) Identification of novel FMR1 variants by massively parallel sequencing in developmentally delayed males. Am J Med Genet A 152A:2512–2520. doi:10.1002/ajmg.a.33626 pmid:20799337
    OpenUrlCrossRefPubMed
  49. ↵
    Colnaghi L, Russo L, Natale C, Restelli E, Cagnotto A, Salmona M, Chiesa R, Fioriti L (2019) Super resolution microscopy of SUMO proteins in neurons. Front Cell Neurosci 13:1–12.
    OpenUrlCrossRef
  50. ↵
    Coultrap SJ, Bayer KU (2012) CaMKII regulation in information processing and storage. Trends Neurosci 35:607–618. doi:10.1016/j.tins.2012.05.003 pmid:22717267
    OpenUrlCrossRefPubMed
  51. ↵
    Coyaud E, Mis M, Laurent EMN, Dunham WH, Couzens AL, Robitaille M, Gingras A-C, Angers S, Raught B (2015) BioID-based identification of Skp cullin F-box (SCF) β-TrCP1/2 EIII ligase substrates. Mol Cell Proteomics 14:1781–1795. doi:10.1074/mcp.M114.045658 pmid:25900982
    OpenUrlAbstract/FREE Full Text
  52. ↵
    Craig TJ, Henley JM (2012) SUMOylation, Arc and the regulation homeostatic synaptic scaling. Commun Integr Biol 5:634–636. doi:10.4161/cib.21712 pmid:23739045
    OpenUrlCrossRefPubMed
  53. ↵
    Craig TJ, Jaafari N, Petrovic MM, Jacobs SC, Rubin PP, Mellor JR, Henley JM (2012) Homeostatic synaptic scaling is regulated by protein SUMOylation. J Biol Chem 287:22781–22788. doi:10.1074/jbc.M112.356337 pmid:22582390
    OpenUrlAbstract/FREE Full Text
  54. ↵
    Craig TJ, Anderson D, Evans AJ, Girach F, Henley JM (2015) SUMOylation of Syntaxin1A regulates presynaptic endocytosis. Sci Rep 5:17669. doi:10.1038/srep17669 pmid:26635000
    OpenUrlCrossRefPubMed
  55. ↵
    Daniel JA, Cooper BH, Palvimo JJ, Zhang FP, Brose N, Tirard M (2017) Analysis of SUMO1-conjugation at synapses. Elife 6:e26338. doi:10.7554/eLife.26338
    OpenUrlCrossRef
  56. ↵
    Daniel JA, Cooper BH, Palvimo JJ, Zhang F-P, Brose N, Tirard M (2018) Response: commentary: analysis of SUMO1-conjugation at synapses. Front Cell Neurosci 12:1–6.
    OpenUrlCrossRef
  57. De Rubeis S, He X, Goldberg AP, Poultney CS, Samocha K, Ercument Cicek A, Kou Y, Liu L, Fromer M, Walker S, Singh T, Klei L, Kosmicki J, Fu SC, Aleksic B, Biscaldi M, Bolton PF, Brownfeld JM, Cai J, Campbell NG, et al. (2014) Synaptic, transcriptional and chromatin genes disrupted in autism. Nature 515:209–215. doi:10.1038/nature13772
    OpenUrlCrossRefPubMed
  58. ↵
    DeFelipe J (2011) The evolution of the brain, the human nature of cortical circuits, and intellectual creativity. Front Neuroanat 5:29. doi:10.3389/fnana.2011.00029 pmid:21647212
    OpenUrlCrossRefPubMed
  59. ↵
    Diaz J, Scheiner C, Leon E (2018) Presentation of a recurrent FMR1 missense mutation (R138Q) in an affected female. TRD 3:139–144. doi:10.3233/TRD-180028
    OpenUrlCrossRef
  60. ↵
    Diering GH, Huganir RL (2018) The AMPA receptor code of synaptic plasticity. Neuron 100:314–329. doi:10.1016/j.neuron.2018.10.018 pmid:30359599
    OpenUrlCrossRefPubMed
  61. ↵
    Dindot SV, Antalffy BA, Bhattacharjee MB, Beaudet AL (2008) The Angelman syndrome ubiquitin ligase localizes to the synapse and nucleus, and maternal deficiency results in abnormal dendritic spine morphology. Hum Mol Genet 17:111–118. doi:10.1093/hmg/ddm288 pmid:17940072
    OpenUrlCrossRefPubMed
  62. ↵
    Dreier L, Burbea M, Kaplan JM (2005) LIN-23-mediated degradation of β-catenin regulates the abundance of GLR-1 glutamate receptors in the ventral nerve cord of C. elegans. Neuron 46:51–64. doi:10.1016/j.neuron.2004.12.058 pmid:15820693
    OpenUrlCrossRefPubMed
  63. ↵
    Dustrude ET, Wilson SM, Ju W, Xiao Y, Khanna R (2013) CRMP2 protein SUMOylation modulates NaV1.7 channel trafficking. J Biol Chem 288:24316–24331. doi:10.1074/jbc.M113.474924 pmid:23836888
    OpenUrlAbstract/FREE Full Text
  64. ↵
    Dustrude ET, Moutal A, Yang X, Wang Y, Khanna M, Khanna R (2016) Hierarchical CRMP2 posttranslational modifications control NaV1.7 function. Proc Natl Acad Sci USA 113:E8443–E8452. doi:10.1073/pnas.1610531113 pmid:27940916
    OpenUrlAbstract/FREE Full Text
  65. ↵
    Dustrude ET, Perez-Miller S, François-Moutal L, Moutal A, Khanna M, Khanna R (2017) A single structurally conserved SUMOylation site in CRMP2 controls NaV1.7 function. Channels (Austin) 11:316–328. doi:10.1080/19336950.2017.1299838 pmid:28277940
    OpenUrlCrossRefPubMed
  66. ↵
    Dütting E, Schröder-Kress N, Sticht H, Enz R (2011) SUMO EIII ligases are expressed in the retina and regulate SUMOylation of the metabotropic glutamate receptor 8b. Biochem J 435:365–371. doi:10.1042/BJ20101854 pmid:21288202
    OpenUrlAbstract/FREE Full Text
  67. ↵
    Ehlers MD (2003) Activity level controls postsynaptic composition and signaling via the ubiquitin-proteasome system. Nat Neurosci 6:231–242. doi:10.1038/nn1013 pmid:12577062
    OpenUrlCrossRefPubMed
  68. ↵
    Ehrlich I, Klein M, Rumpel S, Malinow R (2007) PSD-95 is required for activity-driven synapse stabilization. Proc Natl Acad Sci USA 104:4176–4181. doi:10.1073/pnas.0609307104 pmid:17360496
    OpenUrlAbstract/FREE Full Text
  69. ↵
    El Fatimy R, Miozzo F, Mouël A, Abane R, Schwendimann L, Sabéran‐Djoneidi D, Thonel A, Massaoudi I, Paslaru L, Hashimoto ‐Torii K, Christians E, Rakic P, Gressens P, Mezger V (2014) Heat shock factor 2 is a stress‐responsive mediator of neuronal migration defects in models of fetal alcohol syndrome. EMBO Mol Med 6:1043–1061. doi:10.15252/emmm.201303311 pmid:25027850
    OpenUrlAbstract/FREE Full Text
  70. ↵
    Emtage L, Chang H, Tiver R, Rongo C (2009) MAGI-1 modulates AMPA receptor synaptic localization and behavioral plasticity in response to prior experience. PLoS One 4:e4613. doi:10.1371/journal.pone.0004613 pmid:19242552
    OpenUrlCrossRefPubMed
  71. ↵
    Enchev RI, Schulman BA, Peter M (2015) Protein neddylation: beyond cullin-RING ligases. Nat Rev Mol Cell Biol 16:30–44. doi:10.1038/nrm3919 pmid:25531226
    OpenUrlCrossRefPubMed
  72. ↵
    Estruch SB, Graham SA, Deriziotis P, Fisher SE (2016) The language-related transcription factor FOXP2 is post-translationally modified with small ubiquitin-like modifiers. Sci Rep 6:20911. doi:10.1038/srep20911 pmid:26867680
    OpenUrlCrossRefPubMed
  73. Falik Zaccai TC, Savitzki D, Zivony-Elboum Y, Vilboux T, Fitts EC, Shoval Y, Kalfon L, Samra N, Keren Z, Gross B, Chasnyk N, Straussberg R, Mullikin JC, Teer JK, Geiger D, Kornitzer D, Bitterman-Deutsch O, Samson AO, Wakamiya M, Peterson JW, et al. (2017) Phospholipase A 2-activating protein is associated with a novel form of leukoencephalopathy. Brain 140:370–386. doi:10.1093/brain/aww295 pmid:28007986
    OpenUrlCrossRefPubMed
  74. ↵
    Feligioni M, Nishimune A, Henley JM (2009) Protein SUMOylation modulates calcium influx and glutamate release from presynaptic terminals. Eur J Neurosci 29:1348–1356. doi:10.1111/j.1460-9568.2009.06692.x pmid:19344328
    OpenUrlCrossRefPubMed
  75. ↵
    Flotho A, Melchior F (2013) Sumoylation: a regulatory protein modification in health and disease. Annu Rev Biochem 82:357–385. doi:10.1146/annurev-biochem-061909-093311 pmid:23746258
    OpenUrlCrossRefPubMed
  76. Froyen G, Corbett M, Vandewalle J, Jarvela I, Lawrence O, Meldrum C, Bauters M, Govaerts K, Vandeleur L, Van Esch H, Chelly J, Sanlaville D, van Bokhoven H, Ropers HH, Laumonnier F, Ranieri E, Schwartz CE, Abidi F, Tarpey PS, Futreal PA, et al. (2008) Submicroscopic duplications of the hydroxysteroid dehydrogenase HSD17B10 and the EIII ubiquitin ligase HUWE1 are associated with mental retardation. Am J Hum Genet 82:432–443. doi:10.1016/j.ajhg.2007.11.002 pmid:18252223
    OpenUrlCrossRefPubMed
  77. ↵
    Gardiner K (2006) Transcriptional dysregulation in Down syndrome: predictions for altered protein complex stoichiometries and post-translational modifications, and consequences for learning/behavior genes ELK, CREB, and the estrogen and glucocorticoid receptors. Behav Genet 36:439–453. doi:10.1007/s10519-006-9051-1 pmid:16502135
    OpenUrlCrossRefPubMed
  78. ↵
    Gascón S, García-Gallo M, Renart J, Díaz-Guerra M (2007) Endoplasmic reticulum-associated degradation of the NR1 but not the NR2 subunits of the N-methyl-D-aspartate receptor induced by inhibition of the N-glycosylation in cortical neurons. J Neurosci Res 85:1713–1723. doi:10.1002/jnr.21309 pmid:17455306
    OpenUrlCrossRefPubMed
  79. ↵
    Gautam V, Trinidad JC, Rimerman RA, Costa BM, Burlingame AL, Monaghan DT (2013) Nedd4 is a specific EIII ubiquitin ligase for the NMDA receptor subunit GluN2D. Neuropharmacology 74:96–107. doi:10.1016/j.neuropharm.2013.04.035 pmid:23639431
    OpenUrlCrossRefPubMed
  80. Geetha TS, Michealraj KA, Kabra M, Kaur G, Juyal RC, Thelma BK (2014) Targeted deep resequencing identifies MID2 mutation for X-linked intellectual disability with varied disease severity in a large kindred from India. Hum Mutat 35:41–44. doi:10.1002/humu.22453 pmid:24115387
    OpenUrlCrossRefPubMed
  81. ↵
    George AJ, Hoffiz YC, Charles AJ, Zhu Y, Mabb AM (2018) A comprehensive atlas of EIII ubiquitin ligase mutations in neurological disorders. Front Genet 9:1–17.
    OpenUrlCrossRef
  82. ↵
    Geschwind DH, Rakic P (2013) Cortical evolution: judge the brain by its cover. Neuron 80:633–647. doi:10.1016/j.neuron.2013.10.045 pmid:24183016
    OpenUrlCrossRefPubMed
  83. ↵
    Ghosh H, Auguadri L, Battaglia S, Simone Thirouin Z, Zemoura K, Messner S, Acuña MA, Wildner H, Yévenes GE, Dieter A, Kawasaki H, O. Hottiger M, Zeilhofer HU, Fritschy J-M, Tyagarajan SK (2016) Several posttranslational modifications act in concert to regulate gephyrin scaffolding and GABAergic transmission. Nat Commun 7:13365. doi:10.1038/ncomms13365 pmid:27819299
    OpenUrlCrossRefPubMed
  84. ↵
    Girach F, Craig TJ, Rocca DL, Henley JM (2013) RIM1α SUMOylation is required for fast synaptic vesicle exocytosis. Cell Rep 5:1294–1301. doi:10.1016/j.celrep.2013.10.039 pmid:24290762
    OpenUrlCrossRefPubMed
  85. ↵
    Gowran A, Murphy CE, Campbell VA (2009) Delta(9)-tetrahydrocannabinol regulates the p53 post-translational modifiers murine double minute 2 and the small ubiquitin modifier protein in the rat brain. FEBS Lett 583:3412–3418. doi:10.1016/j.febslet.2009.09.056 pmid:19819240
    OpenUrlCrossRefPubMed
  86. ↵
    Green ML, Singh AV, Zhang Y, Nemeth KA, Sulik KK, Knudsen TB (2007) Reprogramming of genetic networks during initiation of the fetal alcohol syndrome. Dev Dyn 236:613–631. doi:10.1002/dvdy.21048 pmid:17200951
    OpenUrlCrossRefPubMed
  87. ↵
    Greer PL, Hanayama R, Bloodgood BL, Mardinly AR, Lipton DM, Flavell SW, Kim TK, Griffith EC, Waldon Z, Maehr R, Ploegh HL, Chowdhury S, Worley PF, Steen J, Greenberg ME (2010) The Angelman syndrome protein Ube3A regulates synapse development by ubiquitinating arc. Cell 140:704–716. doi:10.1016/j.cell.2010.01.026 pmid:20211139
    OpenUrlCrossRefPubMed
  88. ↵
    Greger IH, Watson JF, Cull-Candy SG (2017) Structural and functional architecture of AMPA-type glutamate receptors and their auxiliary proteins. Neuron 94:713–730. doi:10.1016/j.neuron.2017.04.009 pmid:28521126
    OpenUrlCrossRefPubMed
  89. ↵
    Grégoire S, Tremblay AM, Xiao L, Yang Q, Ma K, Nie J, Mao Z, Wu Z, Giguère V, Yang XJ (2006) Control of MEF2 transcriptional activity by coordinated phosphorylation and sumoylation. J Biol Chem 281:4423–4433. doi:10.1074/jbc.M509471200 pmid:16356933
    OpenUrlAbstract/FREE Full Text
  90. ↵
    Gulia R, Sharma R, Bhattacharyya S (2017) A critical role for ubiquitination in the endocytosis of glutamate receptors. J Biol Chem 292:1426–1437. doi:10.1074/jbc.M116.752105 pmid:28011638
    OpenUrlAbstract/FREE Full Text
  91. ↵
    Guy J, Cheval H, Selfridge J, Bird A (2011) The role of MeCP2 in the brain. Annu Rev Cell Dev Biol 27:631–652. doi:10.1146/annurev-cellbio-092910-154121 pmid:21721946
    OpenUrlCrossRefPubMed
  92. Hall EA, Nahorski MS, Murray LM, Shaheen R, Perkins E, Dissanayake KN, Kristaryanto Y, Jones RA, Vogt J, Rivagorda M, Handley MT, Mali GR, Quidwai T, Soares DC, Keighren MA, McKie L, Mort RL, Gammoh N, Garcia-Munoz A, Davey T, et al. (2017) PLAA mutations cause a lethal infantile epileptic encephalopathy by disrupting ubiquitin-mediated endolysosomal degradation of synaptic proteins. Am J Hum Genet 100:706–724. doi:10.1016/j.ajhg.2017.03.008 pmid:28413018
    OpenUrlCrossRefPubMed
  93. Halvardson J, Zhao JJ, Zaghlool A, Wentzel C, Georgii-Hemming P, Månsson E, Sävmarker HE, Brandberg G, Zander CS, Thuresson AC, Feuk L (2016) Mutations in HECW2 are associated with intellectual disability and epilepsy. J Med Genet 53:697–704. doi:10.1136/jmedgenet-2016-103814 pmid:27334371
    OpenUrlAbstract/FREE Full Text
  94. ↵
    Hasegawa Y, Yoshida D, Nakamura Y, Sakakibara SI (2014) Spatiotemporal distribution of SUMOylation components during mouse brain development. J Comp Neurol 522:3020–3036. doi:10.1002/cne.23563 pmid:24639124
    OpenUrlCrossRefPubMed
  95. ↵
    Hegde AN, DiAntonio A (2002) Ubiquitin and the synapse. Nat Rev Neurosci 3:854–861. doi:10.1038/nrn961 pmid:12415293
    OpenUrlCrossRefPubMed
  96. ↵
    Henderson NT, Dalva MB (2018) EphBs and ephrin-Bs: trans-synaptic organizers of synapse development and function. Mol Cell Neurosci 91:108–121. doi:10.1016/j.mcn.2018.07.002 pmid:30031105
    OpenUrlCrossRefPubMed
  97. ↵
    Henley JM, Carmichael RE, Wilkinson KA (2018) Extranuclear SUMOylation in neurons. Trends Neurosci 41:198–210. doi:10.1016/j.tins.2018.02.004 pmid:29530319
    OpenUrlCrossRefPubMed
  98. ↵
    Hensch TK (2005) Critical period plasticity in local cortical circuits. Nat Rev Neurosci 6:877–888. doi:10.1038/nrn1787 pmid:16261181
    OpenUrlCrossRefPubMed
  99. ↵
    Hers I, Vincent EE, Tavaré JM (2011) Akt signalling in health and disease. Cell Signal 23:1515–1527. doi:10.1016/j.cellsig.2011.05.004 pmid:21620960
    OpenUrlCrossRefPubMed
  100. ↵
    Hou L, Antion MD, Hu D, Spencer CM, Paylor R, Klann E (2006) Dynamic translational and proteasomal regulation of fragile X mental retardation protein controls mGluR-dependent long-term depression. Neuron 51:441–454. doi:10.1016/j.neuron.2006.07.005 pmid:16908410
    OpenUrlCrossRefPubMed
  101. ↵
    Hsu WL, Ma YL, Liu YC, Tai DJC, Lee EHY (2020) Restoring Wnt6 signaling ameliorates behavioral deficits in MeCP2 T158A mouse model of Rett syndrome. Sci Rep 10:1074. doi:10.1038/s41598-020-57745-w pmid:31974426
    OpenUrlCrossRefPubMed
  102. ↵
    Huang D, Liu H, Zhu A, Zhou Y, Li Y (2020) Forebrain excitatory neuron-specific SENP2 knockout mouse displays hyperactivity, impaired learning and memory, and anxiolytic-like behavior. Mol Brain 13:59. doi:10.1186/s13041-020-00591-8 pmid:32290845
    OpenUrlCrossRefPubMed
  103. ↵
    Huang J, Ikeuchi Y, Malumbres M, Bonni A (2015) A Cdh1-APC/FMRP ubiquitin signaling link drives mGluR-dependent synaptic plasticity in the mammalian brain. Neuron 86:726–739. doi:10.1016/j.neuron.2015.03.049 pmid:25913861
    OpenUrlCrossRefPubMed
  104. ↵
    Huang TT, Wuerzberger-Davis SM, Wu Z-H, Miyamoto S (2003) Sequential modification of NEMO/IKKgamma by SUMO-1 and ubiquitin mediates NF-kappaB activation by genotoxic stress. Cell 115:565–576. doi:10.1016/s0092-8674(03)00895-x pmid:14651848
    OpenUrlCrossRefPubMed
  105. ↵
    Huganir RL, Nicoll RA (2013) AMPARs and synaptic plasticity: the last 25 years. Neuron 80:704–717. doi:10.1016/j.neuron.2013.10.025 pmid:24183021
    OpenUrlCrossRefPubMed
  106. ↵
    Hung AY, Sung CC, Brito IL, Sheng M (2010) Degradation of postsynaptic scaffold GKAP and regulation of dendritic spine morphology by the TRIM3 ubiquitin ligase in rat hippocampal neurons. PLoS One 5:e9842. doi:10.1371/journal.pone.0009842 pmid:20352094
    OpenUrlCrossRefPubMed
  107. ↵
    Huo Y, Khatri N, Hou Q, Gilbert J, Wang G, Man HY (2015) The deubiquitinating enzyme USP46 regulates AMPA receptor ubiquitination and trafficking. J Neurochem 134:1067–1080. doi:10.1111/jnc.13194 pmid:26077708
    OpenUrlCrossRefPubMed
  108. ↵
    Ilyas M, Mir A, Efthymiou S, Houlden H (2020) The genetics of intellectual disability: advancing technology and gene editing. F1000Res 9:22. doi:10.12688/f1000research.16315.1
    OpenUrlCrossRef
  109. ↵
    Jaafari N, Konopacki FA, Owen TF, Kantamneni S, Rubin P, Craig TJ, Wilkinson KA, Henley JM (2013) SUMOylation is required for glycine-induced increases in AMPA receptor surface expression (ChemLTP) in hippocampal neurons. PLoS One 8:e52345. doi:10.1371/journal.pone.0052345 pmid:23326329
    OpenUrlCrossRefPubMed
  110. ↵
    Jarome TJ, Helmstetter FJ (2013) The ubiquitin–proteasome system as a critical regulator of synaptic plasticity and long-term memory formation. Neurobiol Learn Mem 105:107–116. doi:10.1016/j.nlm.2013.03.009 pmid:23623827
    OpenUrlCrossRefPubMed
  111. ↵
    Jeong J, Paskus JD, Roche KW (2017) Posttranslational modifications of neuroligins regulate neuronal and glial signaling. Curr Opin Neurobiol 45:130–138. doi:10.1016/j.conb.2017.05.017 pmid:28577430
    OpenUrlCrossRefPubMed
  112. ↵
    Jin H, Chiou TT, Serwanski DR, Miralles CP, Pinal N, De Blas AL (2014) Ring finger protein 34 (RNF34) interacts with and promotes γ-aminobutyric acid type-a receptor degradation via ubiquitination of the γ2 subunit. J Biol Chem 289:29420–29436. doi:10.1074/jbc.M114.603068 pmid:25193658
    OpenUrlAbstract/FREE Full Text
  113. ↵
    Josa-Prado F, Luo J, Rubin P, Henley JM, Wilkinson KA (2019) Developmental profiles of SUMOylation pathway proteins in rat cerebrum and cerebellum. PLoS One 14:e0212857. doi:10.1371/journal.pone.0212857 pmid:30794696
    OpenUrlCrossRefPubMed
  114. ↵
    Ju W, Li Q, Wilson SM, Brittain JM, Meroueh L, Khanna R (2013) SUMOylation alters CRMP2 regulation of calcium influx in sensory neurons. Channels (Austin) 7:153–159. doi:10.4161/chan.24224 pmid:23510938
    OpenUrlCrossRefPubMed
  115. ↵
    Judson MC, Wallace ML, Sidorov MS, Burette AC, Gu B, van Woerden GM, King IF, Han JE, Zylka MJ, Elgersma Y, Weinberg RJ, Philpot BD (2016) GABAergic neuron-specific loss of Ube3a causes Angelman syndrome-like EEG abnormalities and enhances seizure susceptibility. Neuron 90:56–69. doi:10.1016/j.neuron.2016.02.040 pmid:27021170
    OpenUrlCrossRefPubMed
  116. ↵
    Juo P, Kaplan JM (2004) The anaphase-promoting complex regulates the abundance of GLR-1 glutamate receptors in the ventral nerve cord of C. elegans. Curr Biol 14:2057–2062. doi:10.1016/j.cub.2004.11.010 pmid:15556870
    OpenUrlCrossRefPubMed
  117. ↵
    Jurd R, Thornton C, Wang J, Luong K, Phamluong K, Kharazia V, Gibb SL, Ron D (2008) Mind Bomb-2 is an EIII ligase that ubiquitinates the N-methyl-d-aspartate receptor NR2B subunit in a phosphorylation-dependent manner. J Biol Chem 283:301–310. doi:10.1074/jbc.M705580200 pmid:17962190
    OpenUrlAbstract/FREE Full Text
  118. ↵
    Kalchman MA, Graham RK, Xia G, Koide HB, Hodgson JG, Graham KC, Goldberg YP, Gietz RD, Pickart CM, Hayden MR (1996) Huntingtin is ubiquitinated and interacts with a specific ubiquitin-conjugating enzyme. J Biol Chem 271:19385–19394. doi:10.1074/jbc.271.32.19385 pmid:8702625
    OpenUrlAbstract/FREE Full Text
  119. ↵
    Kang J, Gocke CB, Yu H (2006) Phosphorylation-facilitated sumoylation of MEF2C negatively regulates its transcriptional activity. BMC Biochem 7:5. doi:10.1186/1471-2091-7-5
    OpenUrlCrossRefPubMed
  120. ↵
    Kato A, Rouach N, Nicoll RA, Bredt DS (2005) Activity-dependent NMDA receptor degradation mediated by retrotranslocation and ubiquitination. Proc Natl Acad Sci USA 102:5600–5605. doi:10.1073/pnas.0501769102 pmid:15809437
    OpenUrlAbstract/FREE Full Text
  121. ↵
    Khayachi A, Gwizdek C, Poupon G, Alcor D, Chafai M, Cassé F, Maurin T, Prieto M, Folci A, De Graeve F, Castagnola S, Gautier R, Schorova L, Loriol C, Pronot M, Besse F, Brau F, Deval E, Bardoni B, Martin S (2018) Sumoylation regulates FMRP-mediated dendritic spine elimination and maturation. Nat Commun 9:757. doi:10.1038/s41467-018-03222-y pmid:29472612
    OpenUrlCrossRefPubMed
  122. ↵
    Kim H, Kunz PA, Mooney R, Philpot BD, Smith SL (2016) Maternal loss of Ube3a impairs experience-driven dendritic spine maintenance in the developing visual cortex. J Neurosci 36:4888–4894. doi:10.1523/JNEUROSCI.4204-15.2016 pmid:27122043
    OpenUrlAbstract/FREE Full Text
  123. ↵
    Kleijnen MF, Shih AH, Zhou P, Kumar S, Soccio RE, Kedersha NL, Gill G, Howley PM (2000) The hPLIC proteins may provide a link between the ubiquitination machinery and the proteasome. Mol Cell 6:409–419. doi:10.1016/s1097-2765(00)00040-x pmid:10983987
    OpenUrlCrossRefPubMed
  124. ↵
    Ko SJ, Isozaki K, Kim I, Lee JH, Cho HJ, Sohn SY, Oh SR, Park S, Kim DG, Kim CH, Roche KW (2012) PKC phosphorylation regulates mGluR5 trafficking by enhancing binding of Siah-1A. J Neurosci 32:16391–16401. doi:10.1523/JNEUROSCI.1964-12.2012 pmid:23152621
    OpenUrlAbstract/FREE Full Text
  125. ↵
    Komander D, Rape M (2012) The ubiquitin code. Annu Rev Biochem 81:203–229. doi:10.1146/annurev-biochem-060310-170328 pmid:22524316
    OpenUrlCrossRefPubMed
  126. ↵
    Konopacki FA, Jaafari N, Rocca DL, Wilkinson KA, Chamberlain S, Rubin P, Kantamneni S, Mellor JR, Henley JM (2011) Agonist-induced PKC phosphorylation regulates GluK2 SUMOylation and kainate receptor endocytosis. Proc Natl Acad Sci USA 108:19772–19777. doi:10.1073/pnas.1111575108 pmid:22089239
    OpenUrlAbstract/FREE Full Text
  127. ↵
    Kowalski JR, Juo P (2012) The role of deubiquitinating enzymes in synaptic function and nervous system diseases. Neural Plast 2012:892749–892713. doi:10.1155/2012/892749 pmid:23316392
    OpenUrlCrossRefPubMed
  128. ↵
    Kowalski JR, Dube H, Touroutine D, Rush KM, Goodwin PR, Carozza M, Didier Z, Francis MM, Juo P (2014) The anaphase-promoting complex (APC) ubiquitin ligase regulates GABA transmission at the C. elegans neuromuscular junction. Mol Cell Neurosci 58:62–75. doi:10.1016/j.mcn.2013.12.001 pmid:24321454
    OpenUrlCrossRefPubMed
  129. ↵
    Kuczera T, Stilling RM, Hsia H-E, Bahari-Javan S, Irniger S, Nasmyth K, Sananbenesi F, Fischer A (2010) The anaphase promoting complex is required for memory function in mice. Learn Mem 18:49–57. doi:10.1101/lm.1998411 pmid:21191042
    OpenUrlAbstract/FREE Full Text
  130. ↵
    Kühnle S, Mothes B, Matentzoglu K, Scheffner M (2013) Role of the ubiquitin ligase E6AP/UBE3A in controlling levels of the synaptic protein Arc. Proc Natl Acad Sci USA 110:8888–8893. doi:10.1073/pnas.1302792110 pmid:23671107
    OpenUrlAbstract/FREE Full Text
  131. ↵
    Kumar S, Tomooka Y, Noda M (1992) Identification of a set of genes with developmentally down-regulated expression in the mouse brain. Biochem Biophys Res Commun 185:1155–1161. doi:10.1016/0006-291x(92)91747-e pmid:1378265
    OpenUrlCrossRefPubMed
  132. ↵
    Kwon KJ, Kim JN, Kim MK, Kim SY, Cho KS, Jeon SJ, Kim HY, Ryu JH, Han S-Y, Cheong JH, Ignarro LJ, Han S-H, Shin CY (2013) Neuroprotective effects of valproic acid against hemin toxicity: possible involvement of the down-regulation of heme oxygenase-1 by regulating ubiquitin–proteasomal pathway. Neurochem Int 62:240–250. doi:10.1016/j.neuint.2012.12.019 pmid:23298644
    OpenUrlCrossRefPubMed
  133. ↵
    Landré V, Rotblat B, Melino S, Bernassola F, Melino G (2014) Screening for EIII-ubiquitin ligase inhibitors: challenges and opportunities. Oncotarget 5:7988–8013. doi:10.18632/oncotarget.2431 pmid:25237759
    OpenUrlCrossRefPubMed
  134. ↵
    Lappe-Siefke C, Loebrich S, Hevers W, Waidmann OB, Schweizer M, Fehr S, Fritschy J-M, Dikic I, Eilers J, Wilson SM, Kneussel M (2009) The ataxia (axJ) mutation causes abnormal GABAA receptor turnover in mice. PLoS Genet 5:e1000631. doi:10.1371/journal.pgen.1000631 pmid:19759851
    OpenUrlCrossRefPubMed
  135. ↵
    Leal T, Andrieux J, Duban-Bedu B, Bouquillon S, Brevière GM, Delobel B (2009) Array-CGH detection of a de novo 0.8 Mb deletion in 19q13.32 associated with mental retardation, cardiac malformation, cleft lip and palate, hearing loss and multiple dysmorphic features. Eur J Med Genet 52:62–66. doi:10.1016/j.ejmg.2008.09.007 pmid:19022414
    OpenUrlCrossRefPubMed
  136. ↵
    Lee L, Dale E, Staniszewski A, Zhang H, Saeed F, Sakurai M, Fa' M, Orozco I, Michelassi F, Akpan N, Lehrer H, Arancio O (2014) Regulation of synaptic plasticity and cognition by SUMO in normal physiology and Alzheimer’s disease. Sci Rep 4:7190. doi:10.1038/srep07190 pmid:25448527
    OpenUrlCrossRefPubMed
  137. ↵
    Lee SH, Choi JH, Lee N, Lee HR, Kim JI, Yu NK, Choi SL, Lee SH, Kim H, Kaang BK (2008) Synaptic protein degradation underlies destabilization of retrieved fear memory. Science 319:1253–1256. doi:10.1126/science.1150541 pmid:18258863
    OpenUrlAbstract/FREE Full Text
  138. ↵
    Lee S, Park S, Lee H, Han S, Song J, Han D, Suh YH (2019) Nedd4 EIII ligase and beta-arrestins regulate ubiquitination, trafficking, and stability of the mGlu7 receptor. Elife 8:e44502. doi:10.7554/eLife.44502
    OpenUrlCrossRef
  139. ↵
    Li M, Shin Y-H, Hou L, Huang X, Wei Z, Klann E, Zhang P (2008) The adaptor protein of the anaphase promoting complex Cdh1 is essential in maintaining replicative lifespan and in learning and memory. Nat Cell Biol 10:1083–1089. doi:10.1038/ncb1768 pmid:19160489
    OpenUrlCrossRefPubMed
  140. ↵
    Loriol C, Parisot J, Poupon G, Gwizdek C, Martin S (2012) Developmental regulation and spatiotemporal redistribution of the sumoylation machinery in the rat central nervous system. PLoS One 7:e33757. doi:10.1371/journal.pone.0033757 pmid:22438991
    OpenUrlCrossRefPubMed
  141. ↵
    Loriol C, Khayachi A, Poupon G, Gwizdek C, Martin S (2013) Activity-dependent regulation of the sumoylation machinery in rat hippocampal neurons. Biol Cell 105:30–45. doi:10.1111/boc.201200016 pmid:23066795
    OpenUrlCrossRefPubMed
  142. ↵
    Loriol C, Cassé F, Khayachi A, Poupon G, Chafai M, Deval E, Gwizdek C, Martin S (2014) mGlu5 receptors regulate synaptic sumoylation via a transient PKC-dependent diffusional trapping of Ubc9 into spines. Nat Commun 5:5113. doi:10.1038/ncomms6113 pmid:25311713
    OpenUrlCrossRefPubMed
  143. ↵
    Luo HB, Xia YY, Shu XJ, Liu ZC, Feng Y, Liu XH, Yu G, Yin G, Xiong YS, Zeng K, Jiang J, Ye K, Wang XC, Wang JZ (2014) SUMOylation at K340 inhibits tau degradation through deregulating its phosphorylation and ubiquitination. Proc Natl Acad Sci USA 111:16586–16591. doi:10.1073/pnas.1417548111 pmid:25378699
    OpenUrlAbstract/FREE Full Text
  144. ↵
    Luscher B, Fuchs T, Kilpatrick CL (2011) GABAA receptor trafficking-mediated plasticity of inhibitory synapses. Neuron 70:385–409. doi:10.1016/j.neuron.2011.03.024 pmid:21555068
    OpenUrlCrossRefPubMed
  145. ↵
    Lussier MP, Nasu-Nishimura Y, Roche KW (2011) Activity-dependent ubiquitination of the AMPA receptor subunit GluA2. J Neurosci 31:3077–3081. doi:10.1523/JNEUROSCI.5944-10.2011 pmid:21414928
    OpenUrlAbstract/FREE Full Text
  146. ↵
    Lussier MP, Herring BE, Nasu-Nishimura Y, Neutzner A, Karbowski M, Youle RJ, Nicoll RA, Roche KW (2012) Ubiquitin ligase RNF167 regulates AMPA receptor-mediated synaptic transmission. Proc Natl Acad Sci USA 109:19426–19431. doi:10.1073/pnas.1217477109 pmid:23129617
    OpenUrlAbstract/FREE Full Text
  147. ↵
    Mabb AM, Ehlers MD (2010) Ubiquitination in postsynaptic function and plasticity. Annu Rev Cell Dev Biol 26:179–210. doi:10.1146/annurev-cellbio-100109-104129 pmid:20604708
    OpenUrlCrossRefPubMed
  148. ↵
    Mabb AM, Judson MC, Zylka MJ, Philpot BD (2011) Angelman syndrome: insights into genomic imprinting and neurodevelopmental phenotypes. Trends Neurosci 34:293–303. doi:10.1016/j.tins.2011.04.001 pmid:21592595
    OpenUrlCrossRefPubMed
  149. ↵
    Mabb AM, Je HS, Wall MJ, Robinson CG, Larsen RS, Qiang Y, Corrêa SAL, Ehlers MD (2014) Triad3A regulates synaptic strength by ubiquitination of arc. Neuron 82:1299–1316. doi:10.1016/j.neuron.2014.05.016 pmid:24945773
    OpenUrlCrossRefPubMed
  150. ↵
    Mahajan R, Delphin C, Guan T, Gerace L, Melchior F (1997) A small ubiquitin-related polypeptide involved in targeting RanGAP1 to nuclear pore complex protein RanBP2. Cell 88:97–107. doi:10.1016/s0092-8674(00)81862-0 pmid:9019411
    OpenUrlCrossRefPubMed
  151. ↵
    Manzini MC, Xiong L, Shaheen R, Tambunan DE, Di Costanzo S, Mitisalis V, Tischfield DJ, Cinquino A, Ghaziuddin M, Christian M, Jiang Q, Laurent S, Nanjiani ZA, Rasheed S, Hill RS, Lizarraga SB, Gleason D, Sabbagh D, Salih MA, Alkuraya FS, et al. (2014) CC2D1A regulates human intellectual and social function as well as NF-κB signaling homeostasis. Cell Rep 8:647–655. doi:10.1016/j.celrep.2014.06.039 pmid:25066123
    OpenUrlCrossRefPubMed
  152. ↵
    Mao SC, Lin HC, Gean PW (2008) Augmentation of fear extinction by D-cycloserine is blocked by proteasome inhibitors. Neuropsychopharmacology 33:3085–3095. doi:10.1038/npp.2008.30 pmid:18368037
    OpenUrlCrossRefPubMed
  153. ↵
    Maraschi A, Ciammola A, Folci A, Sassone F, Ronzitti G, Cappelletti G, Silani V, Sato S, Hattori N, Mazzanti M, Chieregatti E, Mulle C, Passafaro M, Sassone J (2014) Parkin regulates kainate receptors by interacting with the GluK2 subunit. Nat Commun 5:5182. doi:10.1038/ncomms6182 pmid:25316086
    OpenUrlCrossRefPubMed
  154. ↵
    Margolis SS, Salogiannis J, Lipton DM, Mandel-Brehm C, Wills ZP, Mardinly AR, Hu L, Greer PL, Bikoff JB, Ho H-YH, Soskis MJ, Sahin M, Greenberg ME (2010) EphB-mediated degradation of the RhoA GEF Ephexin5 relieves a developmental brake on excitatory synapse formation. Cell 143:442–455. doi:10.1016/j.cell.2010.09.038 pmid:21029865
    OpenUrlCrossRefPubMed
  155. ↵
    Margolis SS, Sell GL, Zbinden MA, Bird LM (2015) Angelman syndrome. Neurotherapeutics 12:641–650. doi:10.1007/s13311-015-0361-y pmid:26040994
    OpenUrlCrossRefPubMed
  156. ↵
    Martin S, Nishimune A, Mellor JR, Henley JM (2007) SUMOylation regulates kainate-receptor-mediated synaptic transmission. Nature 447:321–325. doi:10.1038/nature05736 pmid:17486098
    OpenUrlCrossRefPubMed
  157. ↵
    Mason S, Anthony B, Lai X, Ringham HN, Wang M, Witzmann FA, You J-S, Zhou FC (2012) Ethanol exposure alters protein expression in a mouse model of fetal alcohol spectrum disorders. Int J Proteomics 2012:1–10. doi:10.1155/2012/867141
    OpenUrlCrossRef
  158. ↵
    Matsuzaki S, Lee L, Knock E, Srikumar T, Sakurai M, Hazrati L-N, Katayama T, Staniszewski A, Raught B, Arancio O, Fraser PE (2015) SUMO1 affects synaptic function, spine density and memory. Sci Rep 5:10730. doi:10.1038/srep10730 pmid:26022678
    OpenUrlCrossRefPubMed
  159. ↵
    Matunis MJ, Coutavas E, Blobel G (1996) A novel ubiquitin-like modification modulates the partitioning of the Ran-GTPase-activating protein RanGAP1 between the cytosol and the nuclear pore complex. J Cell Biol 135:1457–1470. doi:10.1083/jcb.135.6.1457 pmid:8978815
    OpenUrlAbstract/FREE Full Text
  160. ↵
    Meredith LJ, Wang C-M, Nascimento L, Liu R, Wang L, Yang W-H (2016) The key regulator for language and speech development, FOXP2, is a novel substrate for SUMOylation. J Cell Biochem 117:426–438. doi:10.1002/jcb.25288 pmid:26212494
    OpenUrlCrossRefPubMed
  161. ↵
    Miao S, Chen R, Ye J, Tan GH, Li S, Zhang J, Jiang Yh, Xiong ZQ (2013) The Angelman syndrome protein Ube3a is required for polarized dendrite morphogenesis in pyramidal neurons. J Neurosci 33:327–333. doi:10.1523/JNEUROSCI.2509-12.2013 pmid:23283345
    OpenUrlAbstract/FREE Full Text
  162. Micale L, Fusco C, Augello B, Napolitano LMR, Dermitzakis ET, Meroni G, Merla G, Reymond A (2008) Williams-Beuren syndrome TRIM50 encodes an EIII ubiquitin ligase. Eur J Hum Genet 16:1038–1049. doi:10.1038/ejhg.2008.68 pmid:18398435
    OpenUrlCrossRefPubMed
  163. Moortgat S, Berland S, Aukrust I, Maystadt I, Baker L, Benoit V, Caro-Llopis A, Cooper NS, Debray FG, Faivre L, Gardeitchik T, Haukanes BI, Houge G, Kivuva E, Martinez F, Mehta SG, Nassogne MC, Powell-Hamilton N, Pfundt R, Rosello M, et al. (2018) HUWE1 variants cause dominant X-linked intellectual disability: a clinical study of 21 patients. Eur J Hum Genet 26:64–74. doi:10.1038/s41431-017-0038-6 pmid:29180823
    OpenUrlCrossRefPubMed
  164. ↵
    Moriyoshi K, Iijima K, Fujii H, Ito H, Cho Y, Nakanishi S (2004) Seven in absentia homolog 1A mediates ubiquitination and degradation of group 1 metabotropic glutamate receptors. Proc Natl Acad Sci USA 101:8614–8619. doi:10.1073/pnas.0403042101 pmid:15163799
    OpenUrlAbstract/FREE Full Text
  165. ↵
    Muralidhar MG, Thomas JB (1993) The Drosophila bendless gene encodes a neural protein related to ubiquitin-conjugating enzymes. Neuron 11:253–266. doi:10.1016/0896-6273(93)90182-q pmid:8394720
    OpenUrlCrossRefPubMed
  166. ↵
    Na CH, Jones DR, Yang Y, Wang X, Xu Y, Peng J (2012) Synaptic protein ubiquitination in rat brain revealed by antibody-based ubiquitome analysis. J Proteome Res 11:4722–4732. doi:10.1021/pr300536k pmid:22871113
    OpenUrlCrossRefPubMed
  167. ↵
    Nair RR, Patil S, Tiron A, Kanhema T, Panja D, Schiro L, Parobczak K, Wilczynski G, Bramham CR (2017) Dynamic Arc SUMOylation and selective interaction with F-actin-binding protein Drebrin A in LTP consolidation in vivo. Front Synaptic Neurosci 9:1–14.
    OpenUrl
  168. Nascimento RMP, Otto PA, de Brouwer APM, Vianna-Morgante AM (2006) UBE2A, which encodes a ubiquitin-conjugating enzyme, is mutated in a novel X-linked mental retardation syndrome. Am J Hum Genet 79:549–555. doi:10.1086/507047 pmid:16909393
    OpenUrlCrossRefPubMed
  169. ↵
    Nawaz Z, Lonard DM, Smith CL, Lev-Lehman E, Tsai SY, Tsai MJ, O'Malley BW (1999) The Angelman syndrome-associated protein, EVI-AP, is a coactivator for the nuclear hormone receptor superfamily. Mol Cell Biol 19:1182–1189. doi:10.1128/mcb.19.2.1182 pmid:9891052
    OpenUrlAbstract/FREE Full Text
  170. ↵
    Nawrocki ST, Griffin P, Kelly KR, Carew JS (2012) MLN4924: a novel first-in-class inhibitor of NEDD8-activating enzyme for cancer therapy. Expert Opin Investig Drugs 21:1563–1573. doi:10.1517/13543784.2012.707192 pmid:22799561
    OpenUrlCrossRefPubMed
  171. ↵
    Necchi D, Lomoio S, Scherini E (2011) Dysfunction of the ubiquitin-proteasome system in the cerebellum of aging Ts65Dn mice. Exp Neurol 232:114–118. doi:10.1016/j.expneurol.2011.08.009 pmid:21867700
    OpenUrlCrossRefPubMed
  172. ↵
    Nelson RF, Glenn KA, Miller VM, Wen H, Paulson HL (2006) A novel route for F-box protein-mediated ubiquitination links CHIP to glycoprotein quality control. J Biol Chem 281:20242–20251. doi:10.1074/jbc.M602423200 pmid:16682404
    OpenUrlAbstract/FREE Full Text
  173. ↵
    Ngo-Anh TJ, Bloodgood BL, Lin M, Sabatini BL, Maylie J, Adelman JP (2005) SK channels and NMDA receptors form a Ca2+-mediated feedback loop in dendritic spines. Nat Neurosci 8:642–649. doi:10.1038/nn1449 pmid:15852011
    OpenUrlCrossRefPubMed
  174. Nguyen LS, Schneider T, Rio M, Moutton S, Siquier-Pernet K, Verny F, Boddaert N, Desguerre I, Munich A, Rosa JL, Cormier-Daire V, Colleaux L (2016) A nonsense variant in HERC1 is associated with intellectual disability, megalencephaly, thick corpus callosum and cerebellar atrophy. Eur J Hum Genet 24:455–458. doi:10.1038/ejhg.2015.140 pmid:26153217
    OpenUrlCrossRefPubMed
  175. ↵
    Niswender CM, Conn PJ (2010) Metabotropic glutamate receptors: physiology, pharmacology, and disease. Annu Rev Pharmacol Toxicol 50:295–322. doi:10.1146/annurev.pharmtox.011008.145533 pmid:20055706
    OpenUrlCrossRefPubMed
  176. ↵
    Oaks AW, Zamarbide M, Tambunan DE, Santini E, Di Costanzo S, Pond HL, Johnson MW, Lin J, Gonzalez DM, Boehler JF, Wu GK, Klann E, Walsh CA, Manzini MC (2017) Cc2d1a loss of function disrupts functional and morphological development in forebrain neurons leading to cognitive and social deficits. Cereb Cortex 27:1670–1685. doi:10.1093/cercor/bhw009 pmid:26826102
    OpenUrlCrossRefPubMed
  177. Ortega-Recalde O, Beltrán OI, Gálvez JM, Palma-Montero A, Restrepo CM, Mateus HE, Laissue P (2015) Biallelic HERC1 mutations in a syndromic form of overgrowth and intellectual disability. Clin Genet 88:e1–e3. doi:10.1111/cge.12634 pmid:26138117
    OpenUrlCrossRefPubMed
  178. ↵
    Ou Z, Jarmuz M, Sparagana SP, Michaud J, Décarie JC, Yatsenko SA, Nowakowska B, Furman P, Shaw CA, Shaffer LG, Lupski JR, Chinault AC, Cheung SW, Stankiewicz P (2006) Evidence for involvement of TRE-2 (USP6) oncogene, low-copy repeat and acrocentric heterochromatin in two families with chromosomal translocations. Hum Genet 120:227–237. doi:10.1007/s00439-006-0200-7 pmid:16791615
    OpenUrlCrossRefPubMed
  179. ↵
    Pak DTS, Sheng M (2003) Targeted protein degradation and synapse remodeling by an inducible protein kinase. Science 302:1368–1373. doi:10.1126/science.1082475 pmid:14576440
    OpenUrlAbstract/FREE Full Text
  180. ↵
    Pak DTS, Yang S, Rudolph-Correia S, Kim E, Sheng M (2001) Regulation of dendritic spine morphology by SPAR, a PSD-95-associated RapGAP. Neuron 31:289–303. doi:10.1016/S0896-6273(01)00355-5 pmid:11502259
    OpenUrlCrossRefPubMed
  181. ↵
    Paoletti P, Bellone C, Zhou Q (2013) NMDA receptor subunit diversity: impact on receptor properties, synaptic plasticity and disease. Nat Rev Neurosci 14:383–400. doi:10.1038/nrn3504 pmid:23686171
    OpenUrlCrossRefPubMed
  182. ↵
    Picker JD, Walsh CA (2013) New innovations: therapeutic opportunities for intellectual disabilities. Ann Neurol 74:382–390. doi:10.1002/ana.24002 pmid:24038210
    OpenUrlCrossRefPubMed
  183. ↵
    Pignatelli M, Piccinin S, Molinaro G, Di Menna L, Riozzi B, Cannella M, Motolese M, Vetere G, Catania MV, Battaglia G, Nicoletti F, Nisticò R, Bruno V (2014) Changes in mGlu5 receptor-dependent synaptic plasticity and coupling to homer proteins in the hippocampus of Ube3A hemizygous mice modeling Angelman syndrome. J Neurosci 34:4558–4566. doi:10.1523/JNEUROSCI.1846-13.2014 pmid:24672001
    OpenUrlAbstract/FREE Full Text
  184. ↵
    Pirone L, Xolalpa W, Sigurðsson JO, Ramirez J, Pérez C, González M, de Sabando AR, Elortza F, Rodriguez MS, Mayor U, Olsen JV, Barrio R, Sutherland JD (2017) A comprehensive platform for the analysis of ubiquitin-like protein modifications using in vivo biotinylation. Sci Rep 7:40756. doi:10.1038/srep40756 pmid:28098257
    OpenUrlCrossRefPubMed
  185. ↵
    Plant LD, Dowdell EJ, Dementieva IS, Marks JD, Goldstein SAN (2011) SUMO modification of cell surface Kv2.1 potassium channels regulates the activity of rat hippocampal neurons. J Gen Physiol 137:441–454. doi:10.1085/jgp.201110604 pmid:21518833
    OpenUrlAbstract/FREE Full Text
  186. ↵
    Plant LD, Marks JD, Goldstein SA (2016) SUMOylation of NaV1.2 channels mediates the early response to acute hypoxia in central neurons. Elife 5:e20054. doi:10.7554/eLife.20054
    OpenUrlCrossRefPubMed
  187. ↵
    Powell CM, Schoch S, Monteggia L, Barrot M, Matos MF, Feldmann N, Südhof TC, Nestler EJ (2004) The presynaptic active zone protein RIM1alpha is critical for normal learning and memory. Neuron 42:143–153. doi:10.1016/s0896-6273(04)00146-1 pmid:15066271
    OpenUrlCrossRefPubMed
  188. ↵
    Prieto M, Folci A, Martin S (2020) Post-translational modifications of the Fragile X Mental Retardation Protein in neuronal function and dysfunction. Mol Psychiatry 25:1688–1703.
    OpenUrl
  189. ↵
    Psakhye I, Jentsch S (2012) Protein group modification and synergy in the SUMO pathway as exemplified in DNA repair. Cell 151:807–820. doi:10.1016/j.cell.2012.10.021 pmid:23122649
    OpenUrlCrossRefPubMed
  190. Puffenberger EG, Jinks RN, Wang H, Xin B, Fiorentini C, Sherman EA, Degrazio D, Shaw C, Sougnez C, Cibulskis K, Gabriel S, Kelley RI, Morton DH, Strauss KA (2012) A homozygous missense mutation in HERC2 associated with global developmental delay and autism spectrum disorder. Hum Mutat 33:1639–1646. doi:10.1002/humu.22237 pmid:23065719
    OpenUrlCrossRefPubMed
  191. ↵
    Qi Y, Wang J, Bomben VC, Li D-P, Chen S-R, Sun H, Xi Y, Reed JG, Cheng J, Pan H-L, Noebels JL, Yeh ETH (2014) Hyper-SUMOylation of the Kv7 potassium channel diminishes the M-current leading to seizures and sudden death. Neuron 83:1159–1171. doi:10.1016/j.neuron.2014.07.042 pmid:25189211
    OpenUrlCrossRefPubMed
  192. ↵
    Quartier A, Poquet H, Gilbert-Dussardier B, Rossi M, Casteleyn AS, Portes Vd, Feger C, Nourisson E, Kuentz P, Redin C, Thevenon J, Mosca-Boidron AL, Callier P, Muller J, Lesca G, Huet F, Geoffroy V, El Chehadeh S, Jung M, Trojak B, et al. (2017) Intragenic FMR1 disease-causing variants: a significant mutational mechanism leading to Fragile-X syndrome. Eur J Hum Genet 25:423–431. doi:10.1038/ejhg.2016.204 pmid:28176767
    OpenUrlCrossRefPubMed
  193. ↵
    Ran M, Chen H, Liang B, Liao W, Jiang J, Huang J, Ning C, Zang N, Zhou B, Liao Y, Liu H, Qin F, Yang Q, Li J, Ho W, Liang H, Ye L (2018) Alcohol-induced autophagy via upregulation of PIASy promotes HCV replication in human hepatoma cells. Cell Death Dis 9:898. doi:10.1038/s41419-018-0845-x pmid:30185779
    OpenUrlCrossRefPubMed
  194. ↵
    Rezvani K, Teng Y, Shim D, De Biasi M (2007) Nicotine regulates multiple synaptic proteins by inhibiting proteasomal activity. J Neurosci 27:10508–10519. doi:10.1523/JNEUROSCI.3353-07.2007 pmid:17898222
    OpenUrlAbstract/FREE Full Text
  195. ↵
    Rezvani K, Baalman K, Teng Y, Mee MP, Dawson SP, Wang H, De Biasi M, Mayer RJ (2012) Proteasomal degradation of the metabotropic glutamate receptor 1α is mediated by Homer-3 via the proteasomal S8 ATPase: signal transduction and synaptic transmission. J Neurochem 122:24–37. doi:10.1111/j.1471-4159.2012.07752.x pmid:22486777
    OpenUrlCrossRefPubMed
  196. ↵
    Rizo J, Xu J (2015) The synaptic vesicle release machinery. Annu Rev Biophys 44:339–367. doi:10.1146/annurev-biophys-060414-034057 pmid:26098518
    OpenUrlCrossRefPubMed
  197. ↵
    Rocca DL, Wilkinson KA, Henley JM (2017) SUMOylation of FOXP1 regulates transcriptional repression via CtBP1 to drive dendritic morphogenesis. Sci Rep 7:877. doi:10.1038/s41598-017-00707-6 pmid:28408745
    OpenUrlCrossRefPubMed
  198. ↵
    Ropers HH, Hamel BCJ (2005) X-linked mental retardation. Nat Rev Genet 6:46–57. doi:10.1038/nrg1501 pmid:15630421
    OpenUrlCrossRefPubMed
  199. ↵
    Saliba RS, Michels G, Jacob TC, Pangalos MN, Moss SJ (2007) Activity-dependent ubiquitination of GABAA receptors regulates their accumulation at synaptic sites. J Neurosci 27:13341–13351. doi:10.1523/JNEUROSCI.3277-07.2007 pmid:18045928
    OpenUrlAbstract/FREE Full Text
  200. ↵
    Saliba RS, Pangalos M, Moss SJ (2008) The ubiquitin-like protein Plic-1 enhances the membrane insertion of GABA A receptors by increasing their stability within the endoplasmic reticulum. J Biol Chem 283:18538–18544. doi:10.1074/jbc.M802077200 pmid:18467327
    OpenUrlAbstract/FREE Full Text
  201. ↵
    Salinas GD, Blair LAC, Needleman LA, Gonzales JD, Chen Y, Li M, Singer JD, Marshall J (2006) Actinfilin is a Cul3 substrate adaptor, linking GluR6 kainate receptor subunits to the ubiquitin-proteasome pathway. J Biol Chem 281:40164–40173. doi:10.1074/jbc.M608194200 pmid:17062563
    OpenUrlAbstract/FREE Full Text
  202. Santiago-Sim T, Burrage LC, Ebstein F, Tokita MJ, Miller M, Bi W, Braxton AA, Rosenfeld JA, Shahrour M, Lehmann A, Cogné B, Küry S, Besnard T, Isidor B, Bézieau S, Hazart I, Nagakura H, Immken LL, Littlejohn RO, Roeder E, Kara B, et al. (2017) Biallelic variants in OTUD6B cause an intellectual disability syndrome associated with seizures and dysmorphic features. Am J Hum Genet 100:676–688. doi:10.1016/j.ajhg.2017.03.001 pmid:28343629
    OpenUrlCrossRefPubMed
  203. ↵
    Saraceno GE, Castilla R, Barreto GE, Gonzalez J, Kölliker-Frers RA, Capani F (2012) Hippocampal Dendritic Spines Modifications Induced by Perinatal Asphyxia. Neural Plast 2012:1–10.
    OpenUrlCrossRefPubMed
  204. ↵
    Saran NG, Pletcher MT, Natale JE, Cheng Y, Reeves RH (2003) Global disruption of the cerebellar transcriptome in a Down syndrome mouse model. Hum Mol Genet 12:2013–2019. doi:10.1093/hmg/ddg217 pmid:12913072
    OpenUrlCrossRefPubMed
  205. ↵
    Schorova L, Martin S (2016) Sumoylation in synaptic function and dysfunction. Front Synaptic Neurosci 8:1–24.
    OpenUrlCrossRefPubMed
  206. ↵
    Schorova L, Pronot M, Poupon G, Prieto M, Folci A, Khayachi A, Brau F, Cassé F, Gwizdek C, Martin S (2019) The synaptic balance between sumoylation and desumoylation is maintained by the activation of metabotropic mGlu5 receptors. Cell Mol Life Sci 76:3019–3031. doi:10.1007/s00018-019-03075-8 pmid:30904951
    OpenUrlCrossRefPubMed
  207. ↵
    Schwarz LA, Hall BJ, Patrick GN (2010) Activity-dependent ubiquitination of GluA1 mediates a distinct AMPA receptor endocytosis and sorting pathway. J Neurosci 30:16718–16729. doi:10.1523/JNEUROSCI.3686-10.2010 pmid:21148011
    OpenUrlAbstract/FREE Full Text
  208. ↵
    Scudder SL, Goo MS, Cartier AE, Molteni A, Schwarz LA, Wright R, Patrick GN (2014) Synaptic strength is bidirectionally controlled by opposing activity-dependent regulation of nedd4-1 and USP8. J Neurosci 34:16637–16649. doi:10.1523/JNEUROSCI.2452-14.2014 pmid:25505317
    OpenUrlAbstract/FREE Full Text
  209. ↵
    Scudder SL, Patrick GN (2015) Synaptic structure and function are altered by the neddylation inhibitor MLN4924. Mol Cell Neurosci 65:52–57. doi:10.1016/j.mcn.2015.02.010 pmid:25701678
    OpenUrlCrossRefPubMed
  210. ↵
    Seebahn A, Rose M, Enz R (2008) RanBPM is expressed in synaptic layers of the mammalian retina and binds to metabotropic glutamate receptors. FEBS Lett 582:2453–2457. doi:10.1016/j.febslet.2008.06.010 pmid:18555800
    OpenUrlCrossRefPubMed
  211. ↵
    Shalizi A, Gaudillière B, Yuan Z, Stegmüller J, Shirogane T, Ge Q, Tan Y, Schulman B, Harper JW, Bonni A (2006) A calcium-regulated MEF2 sumoylation switch controls postsynaptic differentiation. Science 311:1012–1017. doi:10.1126/science.1122513 pmid:16484498
    OpenUrlAbstract/FREE Full Text
  212. ↵
    Shalizi A, Bilimoria PM, Stegmüller J, Gaudillière B, Yang Y, Shuai K, Bonni A (2007) PIASx is a MEF2 SUMO EIII ligase that promotes postsynaptic dendritic morphogenesis. J Neurosci 27:10037–10046. doi:10.1523/JNEUROSCI.0361-07.2007 pmid:17855618
    OpenUrlAbstract/FREE Full Text
  213. ↵
    Sharma M, Li X, Wang Y, Zarnegar M, Huang C-Y, Palvimo JJ, Lim B, Sun Z (2003) hZimp10 is an androgen receptor co-activator and forms a complex with SUMO-1 at replication foci. EMBO J 22:6101–6114. doi:10.1093/emboj/cdg585 pmid:14609956
    OpenUrlAbstract/FREE Full Text
  214. ↵
    Sheng M, Hoogenraad CC (2007) The postsynaptic architecture of excitatory synapses: a more quantitative view. Annu Rev Biochem 76:823–847. doi:10.1146/annurev.biochem.76.060805.160029 pmid:17243894
    OpenUrlCrossRefPubMed
  215. ↵
    Sheng M, Kim E (2011) The postsynaptic organization of synapses. Cold Spring Harb Perspect Biol 3:a005678. doi:10.1101/cshperspect.a005678
    OpenUrlAbstract/FREE Full Text
  216. ↵
    Shepherd JD, Rumbaugh G, Wu J, Chowdhury S, Plath N, Kuhl D, Huganir RL, Worley PF (2006) Arc/Arg3.1 mediates homeostatic synaptic scaling of AMPA receptors. Neuron 52:475–484. doi:10.1016/j.neuron.2006.08.034 pmid:17088213
    OpenUrlCrossRefPubMed
  217. ↵
    Sitzmann AF, Hagelstrom RT, Tassone F, Hagerman RJ, Butler MG (2018) Rare FMR1 gene mutations causing fragile X syndrome: a review. Am J Med Genet A 176:11–18. doi:10.1002/ajmg.a.38504 pmid:29178241
    OpenUrlCrossRefPubMed
  218. ↵
    Speese SD, Trotta N, Rodesch CK, Aravamudan B, Broadie K (2003) The ubiquitin proteasome system acutely regulates presynaptic protein turnover and synaptic efficacy. Curr Biol 13:899–910. doi:10.1016/S0960-9822(03)00338-5
    OpenUrlCrossRefPubMed
  219. ↵
    Steffan JS, Agrawal N, Pallos J, Rockabrand E, Trotman LC, Slepko N, Illes K, Lukacsovich T, Zhu Y-Z, Cattaneo E, Pandolfi PP, Thompson LM, Marsh JL (2004) SUMO modification of Huntingtin and Huntington’s disease pathology. Science 304:100–104. doi:10.1126/science.1092194 pmid:15064418
    OpenUrlAbstract/FREE Full Text
  220. ↵
    Südhof TC (2004) The synaptic vesicle cycle. Annu Rev Neurosci 27:509–547. doi:10.1146/annurev.neuro.26.041002.131412
    OpenUrlCrossRefPubMed
  221. ↵
    Suizu F, Hiramuki Y, Okumura F, Matsuda M, Okumura AJ, Hirata N, Narita M, Kohno T, Yokota J, Bohgaki M, Obuse C, Hatakeyama S, Obata T, Noguchi M (2009) The EIII ligase TTC3 facilitates ubiquitination and degradation of phosphorylated Akt. Dev Cell 17:800–810. doi:10.1016/j.devcel.2009.09.007 pmid:20059950
    OpenUrlCrossRefPubMed
  222. ↵
    Sun J, Liu Y, Moreno S, Baudry M, Bi X (2015a) Imbalanced mechanistic target of rapamycin C1 and C2 activity in the cerebellum of Angelman syndrome mice impairs motor function. J Neurosci 35:4706–4718. doi:10.1523/JNEUROSCI.4276-14.2015 pmid:25788687
    OpenUrlAbstract/FREE Full Text
  223. ↵
    Sun J, Zhu G, Liu Y, Standley S, Ji A, Tunuguntla R, Wang Y, Claus C, Luo Y, Baudry M, Bi X (2015b) UBE3A regulates synaptic plasticity and learning and memory by controlling SK2 channel endocytosis. Cell Rep 12:449–461. doi:10.1016/j.celrep.2015.06.023 pmid:26166566
    OpenUrlCrossRefPubMed
  224. Tacer KF, Potts PR (2017) Cellular and disease functions of the Prader–Willi syndrome gene MAGEL2. Biochem J 474:2177–2190. doi:10.1042/BCJ20160616 pmid:28626083
    OpenUrlAbstract/FREE Full Text
  225. ↵
    Tada H, Okano HJ, Takagi H, Shibata S, Yao I, Matsumoto M, Saiga T, Nakayama KI, Kashima H, Takahashi T, Setou M, Okano H (2010) Fbxo45, a novel ubiquitin ligase, regulates synaptic activity. J Biol Chem 285:3840–3849. doi:10.1074/jbc.M109.046284 pmid:19996097
    OpenUrlAbstract/FREE Full Text
  226. ↵
    Tai DJC, Liu YC, Hsu WL, Ma YL, Cheng SJ, Liu SY, Lee EHY (2016) MeCP2 SUMOylation rescues Mecp2-mutant-induced behavioural deficits in a mouse model of Rett syndrome. Nat Commun 7:10552. doi:10.1038/ncomms10552 pmid:26842955
    OpenUrlCrossRefPubMed
  227. ↵
    Takagi H, Setou M, Ito S, Yao I (2012) SCRAPPER regulates the thresholds of long-term potentiation/depression, the bidirectional synaptic plasticity in hippocampal CA3-CA1 synapses. Neural Plast 2012:352829. doi:10.1155/2012/352829 pmid:23316391
    OpenUrlCrossRefPubMed
  228. ↵
    Takahashi M, Itakura M, Kataoka M (2003) New aspects of neurotransmitter release and exocytosis: regulation of neurotransmitter release by phosphorylation. J Pharmacol Sci 93:41–45. doi:10.1254/jphs.93.41 pmid:14501150
    OpenUrlCrossRefPubMed
  229. ↵
    Tammsalu T, Matic I, Jaffray EG, Ibrahim AFM, Tatham MH, Hay RT (2014) Proteome-wide identification of SUMO2 modification sites. Sci Signal 7:rs2. doi:10.1126/scisignal.2005146 pmid:24782567
    OpenUrlAbstract/FREE Full Text
  230. ↵
    Tammsalu T, Matic I, Jaffray EG, Ibrahim AFM, Tatham MH, Hay RT (2015) Proteome-wide identification of SUMO modification sites by mass spectrometry. Nat Protoc 10:1374–1388. doi:10.1038/nprot.2015.095 pmid:26292070
    OpenUrlCrossRefPubMed
  231. ↵
    Tang LTH, Craig TJ, Henley JM (2015) SUMOylation of synapsin Ia maintains synaptic vesicle availability and is reduced in an autism mutation. Nat Commun 6:7728. doi:10.1038/ncomms8728 pmid:26173895
    OpenUrlCrossRefPubMed
  232. ↵
    Tang Z, Far O, El Betz H, Scheschonka A (2005) Pias1 interaction and sumoylation of metabotropic glutamate receptor 8. J Biol Chem 280:38153–38159. doi:10.1074/jbc.M508168200 pmid:16144832
    OpenUrlAbstract/FREE Full Text
  233. ↵
    Taylor CP, Burke SP, Weber ML (1995) Hippocampal slices: glutamate overflow and cellular damage from ischemia are reduced by sodium-channel blockade. J Neurosci Methods 59:121–128. doi:10.1016/0165-0270(94)00202-r pmid:7475242
    OpenUrlCrossRefPubMed
  234. ↵
    Tentler D, Johannesson T, Johansson M, Råstam M, Gillberg C, Orsmark C, Carlsson B, Wahlström J, Dahl N (2003) A candidate region for Asperger syndrome defined by two 17p breakpoints. Eur J Hum Genet 11:189–195. doi:10.1038/sj.ejhg.5200939 pmid:12634868
    OpenUrlCrossRefPubMed
  235. ↵
    Tirard M, Hsiao H-H, Nikolov M, Urlaub H, Melchior F, Brose N (2012) In vivo localization and identification of SUMOylated proteins in the brain of His6-HA-SUMO1 knock-in mice. Proc Natl Acad Sci USA 109:21122–21127. doi:10.1073/pnas.1215366110 pmid:23213215
    OpenUrlAbstract/FREE Full Text
  236. ↵
    Tomoda T, Yang K, Sawa A (2020) Neuronal autophagy in synaptic functions and psychiatric disorders. Biol Psychiatry 87:787–796. doi:10.1016/j.biopsych.2019.07.018 pmid:31542152
    OpenUrlCrossRefPubMed
  237. Tønne E, Holdhus R, Stansberg C, Stray-Pedersen A, Petersen K, Brunner HG, Gilissen C, Hoischen A, Prescott T, Steen VM, Fiskerstrand T (2015) Syndromic X-linked intellectual disability segregating with a missense variant in RLIM. Eur J Hum Genet 23:1652–1656. doi:10.1038/ejhg.2015.30 pmid:25735484
    OpenUrlCrossRefPubMed
  238. ↵
    Tramutola A, Di Domenico F, Barone E, Arena A, Giorgi A, di Francesco L, Schininà ME, Coccia R, Head E, Butterfield DA, Perluigi M (2017) Polyubiquitinylation profile in Down syndrome brain before and after the development of Alzheimer neuropathology. Antioxid Redox Signal 26:280–298. doi:10.1089/ars.2016.6686 pmid:27627691
    OpenUrlCrossRefPubMed
  239. ↵
    Tyagarajan SK, Ghosh H, Yévenes GE, Nikonenko I, Ebeling C, Schwerdel C, Sidler C, Zeilhofer HU, Gerrits B, Muller D, Fritschy JM (2011) Regulation of GABAergic synapse formation and plasticity by GSK3beta-dependent phosphorylation of gephyrin. Proc Natl Acad Sci USA 108:379–384. doi:10.1073/pnas.1011824108 pmid:21173228
    OpenUrlAbstract/FREE Full Text
  240. ↵
    Tyagarajan SK, Fritschy J-M (2014) Gephyrin: a master regulator of neuronal function? Nat Rev Neurosci 15:141–156. doi:10.1038/nrn3670 pmid:24552784
    OpenUrlCrossRefPubMed
  241. ↵
    Upadhyay A, Joshi V, Amanullah A, Mishra R, Arora N, Prasad A, Mishra A (2017) EIII ubiquitin ligases neurobiological mechanisms: development to degeneration. Front Mol Neurosci 10:1–21.
    OpenUrlCrossRefPubMed
  242. ↵
    Usui N, Co M, Harper M, Rieger MA, Dougherty JD, Konopka G (2017) Sumoylation of FOXP2 regulates motor function and vocal communication through Purkinje cell development. Biol Psychiatry 81:220–230. doi:10.1016/j.biopsych.2016.02.008 pmid:27009683
    OpenUrlCrossRefPubMed
  243. ↵
    Valnegri P, Huang J, Yamada T, Yang Y, Mejia LA, Cho HY, Oldenborg A, Bonni A (2017) RNF8/UBC13 ubiquitin signaling suppresses synapse formation in the mammalian brain. Nat Commun 8:1271. doi:10.1038/s41467-017-01333-6 pmid:29097665
    OpenUrlCrossRefPubMed
  244. ↵
    van Roessel P, Elliott DA, Robinson IM, Prokop A, Brand AH (2004) Independent regulation of synaptic size and activity by the anaphase-promoting complex. Cell 119:707–718. doi:10.1016/j.cell.2004.11.028 pmid:15550251
    OpenUrlCrossRefPubMed
  245. ↵
    Van Spronsen M, Hoogenraad CC (2010) Synapse pathology in psychiatric and neurologic disease. Curr Neurol Neurosci Rep 10:207–214. doi:10.1007/s11910-010-0104-8 pmid:20425036
    OpenUrlCrossRefPubMed
  246. ↵
    van Woerden GM, Harris KD, Hojjati MR, Gustin RM, Qiu S, de Avila Freire R, Jiang Y, Elgersma Y, Weeber EJ (2007) Rescue of neurological deficits in a mouse model for Angelman syndrome by reduction of alphaCaMKII inhibitory phosphorylation. Nat Neurosci 10:280–282. doi:10.1038/nn1845 pmid:17259980
    OpenUrlCrossRefPubMed
  247. ↵
    Verpelli C, Sala C (2012) Molecular and synaptic defects in intellectual disability syndromes. Curr Opin Neurobiol 22:530–536. doi:10.1016/j.conb.2011.09.007 pmid:22000839
    OpenUrlCrossRefPubMed
  248. ↵
    Vicidomini C, Ponzoni L, Lim D, Schmeisser MJ, Reim D, Morello N, Orellana D, Tozzi A, Durante V, Scalmani P, Mantegazza M, Genazzani AA, Giustetto M, Sala M, Calabresi P, Boeckers TM, Sala C, Verpelli C (2017) Pharmacological enhancement of mGlu5 receptors rescues behavioral deficits in SHANK3 knock-out mice. Mol Psychiatry 22:689–702. doi:10.1038/mp.2016.30 pmid:27021819
    OpenUrlCrossRefPubMed
  249. ↵
    Vilardell M, Rasche A, Thormann A, Maschke-Dutz E, Pérez-Jurado LA, Lehrach H, Herwig R (2011) Meta-analysis of heterogeneous Down Syndrome data reveals consistent genome-wide dosage effects related to neurological processes. BMC Genomics 12:229. doi:10.1186/1471-2164-12-229 pmid:21569303
    OpenUrlCrossRefPubMed
  250. ↵
    Vissers LELM, Gilissen C, Veltman JA (2016) Genetic studies in intellectual disability and related disorders. Nat Rev Genet 17:9–18. doi:10.1038/nrg3999 pmid:26503795
    OpenUrlCrossRefPubMed
  251. ↵
    Vogl AM, Brockmann MM, Giusti SA, MacCarrone G, Vercelli CA, Bauder CA, Richter JS, Roselli F, Hafner AS, Dedic N, Wotjak CT, Vogt-Weisenhorn DM, Choquet D, Turck CW, Stein V, Deussing JM, Refojo D (2015) Neddylation inhibition impairs spine development, destabilizes synapses and deteriorates cognition. Nat Neurosci 18:239–251. doi:10.1038/nn.3912 pmid:25581363
    OpenUrlCrossRefPubMed
  252. ↵
    Vogl AM, Phu L, Becerra R, Giusti SA, Verschueren E, Hinkle TB, Bordenave MD, Adrian M, Heidersbach A, Yankilevich P, Stefani FD, Wurst W, Hoogenraad CC, Kirkpatrick DS, Refojo D, Sheng M (2020) Global site-specific neddylation profiling reveals that NEDDylated cofilin regulates actin dynamics. Nat Struct Mol Biol 27:210–220. doi:10.1038/s41594-019-0370-3 pmid:32015554
    OpenUrlCrossRefPubMed
  253. ↵
    Waites CL, Leal-Ortiz SA, Okerlund N, Dalke H, Fejtova A, Altrock WD, Gundelfinger ED, Garner CC (2013) Bassoon and Piccolo maintain synapse integrity by regulating protein ubiquitination and degradation. EMBO J 32:954–969. doi:10.1038/emboj.2013.27 pmid:23403927
    OpenUrlCrossRefPubMed
  254. ↵
    Wallace ML, Burette AC, Weinberg RJ, Philpot BD (2012) Maternal loss of Ube3a produces an excitatory/inhibitory imbalance through neuron type-specific synaptic defects. Neuron 74:793–800. doi:10.1016/j.neuron.2012.03.036 pmid:22681684
    OpenUrlCrossRefPubMed
  255. ↵
    Walters BJ, Hallengren JJ, Theile CS, Ploegh HL, Wilson SM, Dobrunz LE (2014) A catalytic independent function of the deubiquitinating enzyme USP14 regulates hippocampal synaptic short-term plasticity and vesicle number. J Physiol 592:571–586. doi:10.1113/jphysiol.2013.266015 pmid:24218545
    OpenUrlCrossRefPubMed
  256. Wang HL, Chang NC, Weng YH, Yeh TH (2013) XLID CUL4B mutants are defective in promoting TSC2 degradation and positively regulating mTOR signaling in neocortical neurons. Biochim Biophys Acta 1832:585–593. doi:10.1016/j.bbadis.2013.01.010 pmid:23348097
    OpenUrlCrossRefPubMed
  257. ↵
    Wang L, Rodriguiz R, Wetsel W, Sheng H, Zhao S, Liu X, Paschen W, Yang W (2014) Neuron-specific Sumo1–3 knockdown in mice impairs episodic and fear memories. J Psychiatry Neurosci 39:259–266. doi:10.1503/jpn.130148 pmid:24690371
    OpenUrlAbstract/FREE Full Text
  258. ↵
    Wang Q, Liu L, Pei L, Ju W, Ahmadian G, Lu J, Wang Y, Liu F, Wang YT (2003) Control of synaptic strength, a novel function of Akt. Neuron 38:915–928. doi:10.1016/s0896-6273(03)00356-8 pmid:12818177
    OpenUrlCrossRefPubMed
  259. ↵
    Wang J, Lou SS, Wang T, Wu RJ, Li G, Zhao M, Lu B, Li YY, Zhang J, Cheng X, Shen Y, Wang X, Zhu ZC, Li MJ, Takumi T, Yang H, Yu X, Liao L, Xiong ZQ (2019a) UBE3A-mediated PTPA ubiquitination and degradation regulate PP2A activity and dendritic spine morphology. Proc Natl Acad Sci USA 116:12500–12505. doi:10.1073/pnas.1820131116 pmid:31160454
    OpenUrlAbstract/FREE Full Text
  260. ↵
    Wang T, Wang J, Wang J, Mao L, Tang B, Vanderklish PW, Liao X, Xiong ZQ, Liao L (2019b) HAP1 is an in vivo UBE3A target that augments autophagy in a mouse model of Angelman syndrome. Neurobiol Dis 132:104585. doi:10.1016/j.nbd.2019.104585 pmid:31445164
    OpenUrlCrossRefPubMed
  261. ↵
    Watanabe M, Takahashi K, Tomizawa K, Mizusawa H, Takahashi H (2008) Developmental regulation of Ubc9 in the rat nervous system. Acta Biochim Pol 55:681–686. pmid:19039338
    OpenUrlPubMed
  262. ↵
    Waung MW, Pfeiffer BE, Nosyreva ED, Ronesi JA, Huber KM (2008) Rapid translation of Arc/Arg3.1 selectively mediates mGluR-dependent LTD through persistent increases in AMPAR endocytosis rate. Neuron 59:84–97. doi:10.1016/j.neuron.2008.05.014 pmid:18614031
    OpenUrlCrossRefPubMed
  263. ↵
    Weber ML, Taylor CP (1994) Damage from oxygen and glucose deprivation in hippocampal slices is prevented by tetrodotoxin, lidocaine and phenytoin without blockade of action potentials. Brain Res 664:167–177. doi:10.1016/0006-8993(94)91967-4 pmid:7895026
    OpenUrlCrossRefPubMed
  264. ↵
    Weeber EJ, Jiang Y-H, Elgersma Y, Varga AW, Carrasquillo Y, Brown SE, Christian JM, Mirnikjoo B, Silva A, Beaudet AL, Sweatt JD (2003) Derangements of hippocampal calcium/calmodulin-dependent protein kinase II in a mouse model for Angelman mental retardation syndrome. J Neurosci 23:2634–2644. doi:10.1523/JNEUROSCI.23-07-02634.2003
    OpenUrlAbstract/FREE Full Text
  265. ↵
    Welch MA, Forster LA, Atlas SI, Baro DJ (2019) SUMOylating two distinct sites on the A-type potassium channel, Kv4.2, increases surface expression and decreases current amplitude. Front Mol Neurosci 12:144. doi:10.3389/fnmol.2019.00144 pmid:31213982
    OpenUrlCrossRefPubMed
  266. ↵
    Widagdo J, Chai YJ, Ridder MC, Chau YQ, Johnson RC, Sah P, Huganir RL, Anggono V (2015) Activity-dependent ubiquitination of GluA1 and GluA2 regulates AMPA receptor intracellular sorting and degradation. Cell Rep 10:783–795. doi:10.1016/j.celrep.2015.01.015 pmid:25660027
    OpenUrlCrossRefPubMed
  267. ↵
    Wilkinson KA, Henley JM (2011) Analysis of metabotropic glutamate receptor 7 as a potential substrate for SUMOylation. Neurosci Lett 491:181–186. doi:10.1016/j.neulet.2011.01.032 pmid:21255632
    OpenUrlCrossRefPubMed
  268. ↵
    Wilkinson KA, Martin S, Tyagarajan SK, Arancio O, Craig TJ, Guo C, Fraser PE, Goldstein SAN, Henley JM (2017) Commentary: analysis of SUMO1-conjugation at synapses. Front Cell Neurosci 11:2016–2018.
    OpenUrl
  269. ↵
    Willeumier K, Pulst SM, Schweizer FE (2006) Proteasome inhibition triggers activity-dependent increase in the size of the recycling vesicle pool in cultured hippocampal neurons. J Neurosci 26:11333–11341. doi:10.1523/JNEUROSCI.1684-06.2006 pmid:17079661
    OpenUrlAbstract/FREE Full Text
  270. ↵
    Wilson SM, Bhattacharyya B, Rachel RA, Coppola V, Tessarollo L, Householder DB, Fletcher CF, Miller RJ, Copeland NG, Jenkins NA (2002) Synaptic defects in ataxia mice result from a mutation in Usp14, encoding a ubiquitin-specific protease. Nat Genet 32:420–425. doi:10.1038/ng1006 pmid:12368914
    OpenUrlCrossRefPubMed
  271. ↵
    Yang CY, Yu TH, Wen WL, Ling P, Hsu KS (2019) Conditional deletion of CC2D1A reduces hippocampal synaptic plasticity and impairs cognitive function through Rac1 hyperactivation. J Neurosci 39:4959–4975. doi:10.1523/JNEUROSCI.2395-18.2019 pmid:30992372
    OpenUrlAbstract/FREE Full Text
  272. ↵
    Yao I, Takagi H, Ageta H, Kahyo T, Sato S, Hatanaka K, Fukuda Y, Chiba T, Morone N, Yuasa S, Inokuchi K, Ohtsuka T, MacGregor GR, Tanaka K, Setou M (2007) SCRAPPER-dependent ubiquitination of active zone protein RIM1 regulates synaptic vesicle release. Cell 130:943–957. doi:10.1016/j.cell.2007.06.052 pmid:17803915
    OpenUrlCrossRefPubMed
  273. ↵
    Yao I, Takao K, Miyakawa T, Ito S, Setou M (2011) Synaptic EIII ligase SCRAPPER in contextual fear conditioning: extensive behavioral phenotyping of scrapper heterozygote and overexpressing mutant mice. PLoS One 6:e17317. doi:10.1371/journal.pone.0017317 pmid:21390313
    OpenUrlCrossRefPubMed
  274. ↵
    Yashiro K, Riday TT, Condon KH, Roberts AC, Bernardo DR, Prakash R, Weinberg RJ, Ehlers MD, Philpot BD (2009) Ube3a is required for experience-dependent maturation of the neocortex. Nat Neurosci 12:777–783. doi:10.1038/nn.2327 pmid:19430469
    OpenUrlCrossRefPubMed
  275. ↵
    Yeh SH, Mao SC, Lin HC, Gean PW (2006) Synaptic expression of glutamate receptor after encoding of fear memory in the rat amygdala. Mol Pharmacol 69:299–308. doi:10.1124/mol.105.017194 pmid:16219906
    OpenUrlAbstract/FREE Full Text
  276. ↵
    Yoo DY, Won Kim D, Kwon HJ, Jung HY, Nam SM, Kim JW, Chung JY, Won MH, Yoon YS, Choi SY, Hwang IK (2017) Chronic administration of SUMO-1 has negative effects on novel object recognition memory as well as cell proliferation and neuroblast differentiation in the mouse dentate gyrus. Mol Med Rep 16:3427–3432. doi:10.3892/mmr.2017.6946 pmid:28713906
    OpenUrlCrossRefPubMed
  277. ↵
    Zacchi P, Antonelli R, Cherubini E (2014) Gephyrin phosphorylation in the functional organization and plasticity of GABAergic synapses. Front Cell Neurosci 8:103. doi:10.3389/fncel.2014.00103 pmid:24782709
    OpenUrlCrossRefPubMed
  278. ↵
    Zeng F, Ma X, Zhu L, Xu Q, Zeng Y, Gao Y, Li G, Guo T, Zhang H, Tang X, Wang Z, Ye Z, Zheng L, Zhang H, Zheng Q, Li K, Lu J, Qi X, Luo H, Zhang X, et al. (2019) The deubiquitinase USP6 affects memory and synaptic plasticity through modulating NMDA receptor stability. PLOS Biol 17:e3000525. doi:10.1371/journal.pbio.3000525 pmid:31841517
    OpenUrlCrossRefPubMed
  279. ↵
    Zhang H, Kang E, Wang Y, Yang C, Yu H, Wang Q, Chen Z, Zhang C, Christian KM, Song H, Ming G, Xu Z (2016) Brain-specific Crmp2 deletion leads to neuronal development deficits and behavioural impairments in mice. Nat Commun 7:11773. doi:10.1038/ncomms11773
    OpenUrlCrossRefPubMed
  280. Zhang J, Gambin T, Yuan B, Szafranski P, Rosenfeld JA, Balwi MA, Alswaid A, Al-Gazali L, Shamsi AMA, Komara M, Ali BR, Roeder E, McAuley L, Roy DS, Manchester DK, Magoulas P, King LE, Hannig V, Bonneau D, Denommé-Pichon AS, et al. (2017) Haploinsufficiency of the EIII ubiquitin-protein ligase gene TRIP12 causes intellectual disability with or without autism spectrum disorders, speech delay, and dysmorphic features. Hum Genet 136:377–386. doi:10.1007/s00439-017-1763-1 pmid:28251352
    OpenUrlCrossRefPubMed
  281. ↵
    Zhang J, Zhao B, Zhu X, Li J, Wu F, Li S, Gong X, Cha C, Guo G (2018) Phosphorylation and SUMOylation of CRMP2 regulate the formation and maturation of dendritic spines. Brain Res Bull 139:21–30. doi:10.1016/j.brainresbull.2018.02.004 pmid:29425794
    OpenUrlCrossRefPubMed
  282. ↵
    Zhang Y, Li Z, Gu J, Zhang Y, Wang W, Shen H, Chen G, Wang X (2015) Plic-1, a new target in repressing epileptic seizure by regulation of GABAAR function in patients and a rat model of epilepsy. Clin Sci 129:1207–1223. doi:10.1042/CS20150202 pmid:26415648
    OpenUrlAbstract/FREE Full Text
  283. ↵
    Zhao M, Raingo J, Chen ZJ, Kavalali ET (2011) Cc2d1a, a C2 domain containing protein linked to nonsyndromic mental retardation, controls functional maturation of central synapses. J Neurophysiol 105:1506–1515. doi:10.1152/jn.00950.2010 pmid:21273312
    OpenUrlCrossRefPubMed
  284. Zou Y, Liu Q, Chen B, Zhang X, Guo C, Zhou H, Li J, Gao G, Guo Y, Yan C, Wei J, Shao C, Gong Y (2007) Mutation in CUL4B, which encodes a member of cullin-RING ubiquitin ligase complex, causes X-linked mental retardation. Am J Hum Genet 80:561–566. doi:10.1086/512489 pmid:17273978
    OpenUrlCrossRefPubMed

Synthesis

Reviewing Editor: Juan Burrone, King’s College London

Decisions are customarily a result of the Reviewing Editor and the peer reviewers coming together and discussing their recommendations until a consensus is reached. When revisions are invited, a fact-based synthesis statement explaining their decision and outlining what is needed to prepare a revision will be listed below. The following reviewer(s) agreed to reveal their identity: NONE.

The review provides a comprehensive description of both Ubiquitination and Sumoylation in the process of synapse function and how alterations in these pathways may be linked to intellectual disabilities. Although many reviews exist that discuss Ubiquitination and Sumoylation independently, the novelty here is that both are discussed in the same review and evidence for crosstalk between the two pathways is presented. The link to disease is timely. Overall, the review is well written and will provide an important source of information on the role of post-translational modifications on synaptic physiology and pathophysiology. However, there are a few aspects of the review that could be improved before publication. Below are comments that the authors should consider as suggestions to help improve the manuscript.

MAIN CONCEPTUAL COMMENTS AND SUGGESTIONS

1. In the view of this reviewer the broadness and robustness of the manuscript can be further improved by adding in the first “physiological” section a small sub-chapter describing the role of the Ubiquitin-like protein (UBL) Nedd8 (neural precursor cell-expressed developmentally downregulated gene 8) in neuronal and synaptic function. Nedd8 is the evolutionary closest UBL to ubiquitin and several relevant papers have recently shown the importance of Neddylation in synaptic function. Nedd8 binds to and allows the activity of the Ubiquitin Cullin-RING E3 ligase complexes. The different 7 Cullins control hundreds of targets, many of them of synaptic localization and in fact Cullin themselves have also been mechanistically involved in different neurodevelopmental disorders including intellectual disabilities. But beyond Culllins, and similarly to Ubiquitin and SUMO1-3, Nedd8 also directly targets hundreds of specific proteins (see Vogl et al. 2020, Nat Struct Mol Biol). In recent years, several papers have been published and deposited in electronic archives describing the involvement of Neddylation in the maturation, maintenance and plasticity of the synaptic compartment (Vogl et al. 2015 Nat Neurosci, Scudder et al. 2015 Mol Cell Neurosci., Bayraktar et al 2019 (BioRx), Brockmann et al 2019, Sci Reports, Vogl et al., Nat Struct Mol Biol 2020).

This topic has never been reviewed before and will therefore add novelty, broadness and impact to the article, though the title might probably need to slightly change The final chapter discussing target commonalities and Ub-Sumo molecular cross-talks is interesting too and Neddylation has also direct interactions with the ubiquitin system for instance via the Cullins system.

2. On the other hand, the first (functional) part of the manuscript is some times too focused on describing enzymes and targets, becoming a bit repetitive and becoming closer to a catalog than a conceptual piece. A comprehensive list of mediators and targets is indeed valuable but I think that section of the article would be benefited with additional information and discussion either of phenotypic (e.g. deep discussion of behavioral phenotypes accompanying the synaptic findings) or technical (e.g. recent controversies on the actual existence of SUMO1-conjugation) nature. Regarding this last point, I am particularly surprise with the absence of the papers of Nils Brose Lab which has been challenging with very accurate and extremely well-controlled studies the notion of SUMO1-conjugation to synaptic proteins using even the only tagged-SUMO knock-in mouse model available so far. Particularly the paper of Tirard et al PNAS 2011 (coauthored by Frauke Melchiord) and a subsequent eLife 2017 paper by Daniel et al, as well as another open discussion by Daniel et al in Front. Cellular Neurosci. are excellent materials to bring very relevant discussions (about unspecific antibodies, affinity purification troubles, synaptic localizations, etc.) that have not been properly addressed so far in the field of Ub and UBLs. I do not want to suggest with these comments that SUMOylation does not exist in synapses; there is substantial evidence that strongly suggests it, as the presence of the enzymatic SUMOylation apparatus in the synaptic compartment. But perhaps not all versions of SUMO have the same presence at distal neural sites or this only occurs under certain stimulation conditions, this kind of arguments could also be examined.

Our scientific community is acknowledging that we are facing a reproducibility problem and consequently about the importance of publishing papers with negative results, mainly if they are very very-controlled. This is great opportunity to follow and promote that path. In case the Editor considers that the final length might be an issue, I would recommend shortening the long description of targets to prioritize more conceptual discussions.

3. Finally, the section about ID is much more fluent and eclectic even though the pathophysiological mechanisms of these disorder are rather unclear yet. This section is an excellent counter side to balance physiology and disease. The two main syndromes chosen are thoroughly described with all the mechanistic insight available in the literature. In my opinion, the authors’ decision to leave out of the description, other syndromes (e.g. the ones listed in Table I or Prader-Willy or Kaufman Oculocerebral Syndromes) and enzymes responsible for ID disorders, although it may be justified for reasons of space, also makes this section lose some deepness and perspective. Many of these cases are actually listed in Table I, which is not even discussed. Taking this into account, I would suggest adding a small section after the Down Syndrome description to discuss at least in a general and conceptual rather than descriptive manner, for example, an analysis could be carried out according to groups of enzymes, classes of E3 ligases, types of potential targets, validations with rodent models, clinical typology that distinguishes between disorders with essentially functional pathology vs. anatomical comorbidities, etc.

MINOR COMMENTS

4. In lane 3-5 the author/s claim “Among PTMs, ubiquitination and sumoylation share several features, such as the structure of the small protein modifiers (ubiquitin and Small Ubiquitin-like Modifier, SUMO), the enzymatic cascades mediating the conjugation process and the targeted aminoacidic residue.”

The UBLs are grouped in a large family of proteins that include SUMO but also Nedd8, Fub1, Isg15, Fat10, Hub1, Ufm1, Urm1 and the autophagy-related proteins Atg8 and Atg12. In fact the family does not include ubiquitin itself. This sentence should be corrected accordingly. In addition, in the present form the authors could involuntarily and in a tacit way, favor the interpretation of three erroneous concepts. 1. That ubiquitin is part of the UBLs family. 2. That Sumo is the only UBL. 3. That there are no UBLs with relevance in the central nervous system and synaptic physiology.

To avoid this potential misunderstanding, the sentence should be modified. Furthermore it would be desirable that the authors provide the complete list of UBLs.

5. In line 49: “...that can interfere with neuronal pathways...”

Do the authors mean neuronal intracellular, signaling pathways? It´s a bit misleading.

6. In line 58: “PTMs refer to covalent but reversible enzymatic modifications of target proteins, following their translation (Bürkle, 2001). They typically consist in the addition of a functional moiety, which can be either a chemical group or a small peptide, to specific residues of target proteins.”

This statement is right in principle though it also sounds incomplete in different aspects. The PTMs might also be irreversible like in the case of proteolysis (there are several families of highly relevant proteases, particularly relevant in this review describing several DUBs) and amino acid modifications (e.g. deamination). To be accurate, these two other variants of PTMs might be mentioned. In addition, the not only chemical groups or peptides can be bound to proteins but also complex molecules (e.g. glycosylation, AMPylatoin, ADP ribosylation, etc). It might be desirable that the authors modify the paragraph accordingly.

7. In line 62: “In neurons, PTMs have been extensively investigated and modulate virtually all pathways that are required to ensure proper synaptic transmission, such as presynaptic NT release (Hegde and DiAntonio, 2002; Schorova and Martin, 2016; Takahashi et al., 2003) and trafficking of NT receptors (Diering and Huganir, 2018; Luscher et al., 2011; 65 Schorova and Martin, 2016; Widagdo et al., 2017).

This paragraph attempts to be an exemplified introduction of the multiple actions that PTMs can exert in synapses. I agree with the authors that it is an important sentence. As such it might be desirable to include the description of a broader list of examples of the actions executed by all PTMs on the many different aspects of the synaptic physiology.

8. In line 69: “In this review, we focus on the role of ubiquitination and sumoylation in synapse physiology and their implication in the molecular pathogenesis of intellectual disability (ID), a generalized neurodevelopmental disorder that manifests in a wide range of brain pathologies (Aicardi, 1998).”

Considering that ID would be a central topic along the review I would suggest the authors to provide additional and more recent references.

9. In line 77: “E1 ubiquitin enzymes bind to and activate free ubiquitin.”

I would suggest to very briefly explain what “Ub activation” means.

10. In line 82: “The human genome encodes two E1, ∼50 E2 and ∼600 E3 82 enzymes. Thus, substrate specificity mainly relies on different combinations of E2-E3 proteins.”

This is partially correct because in some cases the target specificity is mainly given by a large set of substrate receptor and adaptors. In fact this is the case of CRLs E3 ligases, which targets the largest amount of proteins in the Ub system. This should be clarified.

11. In line 137 “As this phenotype is partially rescued by the inhibition of either the proteasome or the E3 ligase Siah1, “

-It might be more accurate to rephrase it as “ As this phenotype is partially rescued by either the inhibition the proteasome or the downregulation E3 ligase Siah1,...”

12. In line 234 “ Eventually, proteomic approaches aimed at identifying synaptic ubiquitome in rat brains revealed that the Ca2+/Calmodulin-dependent kinase II (CaMKII), which is crucial for the expression of synaptic plasticity (Coultrap and Bayer, 2012), is ubiquitinated (Fig. 1A).”

-What is the citation describing this?

13. In line 239: “Ubiquitination was also studied at inhibitory synapses (Fig. 1C). The number of surface GABAARs is critically modulated by the ubiquitin-like protein Plic-1 (also named ubiquilin)...”

- Ubiquilin, is actually an ubiquitin-associated chaperon and the term ubiquitin-like protein (originally employed in the rather old referred paper) nowadays is used to name UBLs and therefore it would be misleading. To avoid misunderstanding of non-expert readers, the term UBL should be reserved to describe the UBL family members.

14. In line 278: “For example, neuronal overexpression of SUMO1 in vivo affects the density of dendritic spines and worsens behavioral performances in learning and memory tests (Matsuzaki et al., 2015).”

- It is somehow surprising that the authors choice to start the description of the role of SUMO in synapses with this paper. The overexpression of ubiquitin and UBLs is an ambiguous, misinforming and thus outdated approach that should be used and reserved for very special experimental cases. The lost of endogenous stoichiometry upon SUMO (or other Ub/UBLs) overexpression is known to potentially give raise to artefactual targeting and downstream effects. The paper will be mentioned again later on in the review. If the authors are still interested in mentioning this paper should at least express concerns about this well-known technical drawbacks.

15. In line 356: “A primary presynaptic function controlled by the SUMO pathway is SV docking/priming through sumoylation of RIM1α and synapsin Ia (SynIa) (Fig. 1A). Sumoylation of SynIa triggers its association with SVs and facilitates SV anchoring at the presynaptic membrane (Tang et al., 2015).”

- I already mentioned the papers of Daniel et al above (eLife and Frontiers Cell Neurosci). In those papers, the authors specifically test in different ways the SUMO1 conjugation to several of the best-described synaptic proteins of the literature including RIM1 and Synla. They could not see the described conjugation. In line 410 something similar happen for the case of GluK2. These cases should be discussed in the light of the Daniel´s findings too.

16. In line 376: “However, mGlu7R is the sole member shown to be sumoylated in vivo”.

The in vivo concept often has different interpretation in the different fields. Meanwhile cell cultures models can be considered “in vivo” by many biochemists accustom to work with recombinant proteins and assays, the same approach is rated as “in vitro” for many neuroscientists. To avoid misunderstandings, it would be ideal to clarify the experimental model employed.

17. In line 767: “For example, compounds that modulate sumoylation have been tested to treat cancer and ischemia (Bernstock et al., 2018).”

I would add the word ubiquitination “For example, compounds that modulate ubiquitination and sumoylation have been tested to treat cancer and ischemia (Bernstock et al., 2018).”, adding of course an adequate citation.

18. A recent remarkable paper in Plos Biol. by Feng et al. (2019) “The deubiquitinase USP6 affects memory and synaptic plasticity through modulating NMDA receptor stability” describes the effects of Usp6, a hominoid deubiquitinating enzyme previously implicated in intellectual disability and Asperger Syndrome. This is a interesting paper involving the characterization of the humanized mouse models expressing Usp6 in forebrain pyramidal neurons, a kind of approach that is substantially different to the studies described in the paper which are sometimes repetitive in their experimental nature. A comment on this paper could be included either associated to the role of ubiquitination and NMDA function and/or associated to intellectual disabilities.

Author Response

Back to top

In this issue

eneuro: 7 (4)
eNeuro
Vol. 7, Issue 4
July/August 2020
  • Table of Contents
  • Index by author
  • Ed Board (PDF)
Email

Thank you for sharing this eNeuro article.

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

Enter multiple addresses on separate lines or separate them with commas.
Ubiquitin and Ubiquitin-Like Proteins in the Critical Equilibrium between Synapse Physiology and Intellectual Disability
(Your Name) has forwarded a page to you from eNeuro
(Your Name) thought you would be interested in this article in eNeuro.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Print
View Full Page PDF
Citation Tools
Ubiquitin and Ubiquitin-Like Proteins in the Critical Equilibrium between Synapse Physiology and Intellectual Disability
Alessandra Folci, Filippo Mirabella, Matteo Fossati
eNeuro 27 July 2020, 7 (4) ENEURO.0137-20.2020; DOI: 10.1523/ENEURO.0137-20.2020

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Respond to this article
Share
Ubiquitin and Ubiquitin-Like Proteins in the Critical Equilibrium between Synapse Physiology and Intellectual Disability
Alessandra Folci, Filippo Mirabella, Matteo Fossati
eNeuro 27 July 2020, 7 (4) ENEURO.0137-20.2020; DOI: 10.1523/ENEURO.0137-20.2020
Twitter logo Facebook logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Google Plus One

Jump to section

  • Article
    • Abstract
    • Significance Statement
    • Introduction
    • Ubiquitination and UBL Pathways in Synapse Physiology
    • ID and Impairment of Ubiquitination and Sumoylation Pathways
    • Conclusions and Perspectives
    • Acknowledgments
    • Footnotes
    • References
    • Synthesis
    • Author Response
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF

Keywords

  • intellectual disability
  • ubiquitination
  • SUMOylation
  • neddylation
  • synapse development
  • synapse function

Responses to this article

Respond to this article

Jump to comment:

No eLetters have been published for this article.

Related Articles

Cited By...

More in this TOC Section

Review

  • My 50 Year Odyssey to Develop Behavioral Methods to Let Me See Quickly How Well Kittens See
  • A Systematic Review and Meta-Analysis Assessing the Accuracy of Blood Biomarkers for the Diagnosis of Ischemic Stroke in Adult and Elderly Populations
  • Neuroscientist’s Behavioral Toolbox for Studying Episodic-Like Memory
Show more Review

Disorders of the Nervous System

  • The PDGFBB-PDGFRβ Pathway and Laminins in Pericytes Are Involved in the Temporal Change of AQP4 Polarity during Temporal Lobe Epilepsy Pathogenesis
  • The Ventral Pallidum Innervates a Distinct Subset of Midbrain Dopamine Neurons
  • Rethinking Alzheimer's: Harnessing Cannabidiol to Modulate IDO and cGAS Pathways for Neuroinflammation Control
Show more Disorders of the Nervous System

Subjects

  • Disorders of the Nervous System
  • Reviews
  • Home
  • Alerts
  • Follow SFN on BlueSky
  • Visit Society for Neuroscience on Facebook
  • Follow Society for Neuroscience on Twitter
  • Follow Society for Neuroscience on LinkedIn
  • Visit Society for Neuroscience on Youtube
  • Follow our RSS feeds

Content

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

Information

  • For Authors
  • For the Media

About

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

Copyright © 2025 by the Society for Neuroscience.
eNeuro eISSN: 2373-2822

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