Next Article in Journal
Reprogramming Metabolism of Macrophages as a Target for Kidney Dysfunction Treatment in Autoimmune Diseases
Next Article in Special Issue
Agomelatine, Ketamine and Vortioxetine Attenuate Energy Cell Metabolism—In Vitro Study
Previous Article in Journal
Electrically Conductive and Antimicrobial Agro-Food Waste Biochar Functionalized with Zinc Oxide Particles
Previous Article in Special Issue
Association of DNA Methylation of the NLRP3 Gene with Changes in Cortical Thickness in Major Depressive Disorder
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

SUMOylation and Major Depressive Disorder

1
BK21 FOUR KNU Creative BioResearch Group, School of Life Sciences, College of National Sciences, Kyungpook National University, Daegu 41566, Korea
2
Brain Science and Engineering Institute, Kyungpook National University, Daegu 41566, Korea
3
Department of Biochemistry, Inje University College of Medicine, Busan 50834, Korea
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(14), 8023; https://doi.org/10.3390/ijms23148023
Submission received: 27 June 2022 / Revised: 19 July 2022 / Accepted: 19 July 2022 / Published: 21 July 2022
(This article belongs to the Special Issue Pathogenesis and Treatment Mechanisms of Depression)

Abstract

:
Since the discovery of the small ubiquitin-like modifier (SUMO) protein in 1995, SUMOylation has been considered a crucial post-translational modification in diverse cellular functions. In neurons, SUMOylation has various roles ranging from managing synaptic transmitter release to maintaining mitochondrial integrity and determining neuronal health. It has been discovered that neuronal dysfunction is a key factor in the development of major depressive disorder (MDD). PubMed and Google Scholar databases were searched with keywords such as ‘SUMO’, ‘neuronal plasticity’, and ‘depression’ to obtain relevant scientific literature. Here, we provide an overview of recent studies demonstrating the role of SUMOylation in maintaining neuronal function in participants suffering from MDD.

1. Introduction

Synaptic plasticity is a process involved in the change of synaptic strength through specific patterns of synaptic activity. Changes in synaptic plasticity (also known as Hebbian plasticity) are known to modify thoughts, feelings, behavior, memory, and brain volume [1]. Long-term potentiation (LTP) and depression (LTD) are the best-studied forms of synaptic plasticity. LTP is defined as a long-term increase in synaptic strength and functionality caused by certain patterns of synaptic activity, such as increased postsynaptic α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptors (AMPARs), whereas LTD is defined as a long-term decrease in synaptic strength and function [2]. As reviewed by Richer-Levin and Xu, stress has been shown to prompt LTP and TLD alterations in neurons in the CA1 region of the hippocampus [3]. Other studies have reported negative alterations in rodent CA3 neuron dendrites after exposure to chronic stress, further strengthening the theoretical link between stress, synaptic plasticity, and major depressive disorder (MDD) [4]. Synaptic plasticity is largely controlled by the activation level, location, and state of AMPARs and N-methyl D-aspartate (NMDARs). AMPARs and NMDARs are non-selective ionotropic glutamate receptors responsible for relaying excitatory synaptic transmission through the nervous system [5]. During early LTP, the synapse needs to adapt to rapid changes and thus relies on the movement and control of pre-existing proteins, activation of NMDARs, and trafficking of AMPARs [6,7,8,9]. During late-LTP, the control of gene expression and RNA translation is much more significant [10,11]. There are many reports of synaptic functions being regulated by postsynaptic glutamate receptor modifications such as phosphorylation and ubiquitination. CaMKII is known to cause the phosphorylation of Ser831 of the GluA1 subunit of AMPAR, leading to the increase in channel conductivity, which causes expression of LTP in the hippocampus [12]. The ubiquitin-proteasome system, which is responsible for the targeting and degradation of proteins, has been shown to relocate into dendritic spines during synaptic activation to alter the levels of synaptic proteins [13]. Larger-scale alterations caused by post-translational modifications (PTMs) have also been reported. Applying stress in rodent models via various methods, such as predator and social defeat stress, resulted in a sharp increase in H3 phosphorylation and acetylation in the hippocampus. Predator stress is induced by introducing natural predator odors to rodents, and social stress is initiated by placing a male rodent in the cage of a more dominant male [14,15]. It has been shown that H3 PTMs in the nucleus accumbens can alter susceptibility and resilience to stress. Mice with increased levels of H3 acetylation were shown to be more resilient to stress and subsequent depression-like phenotype, whereas increased levels of H3K9me2 caused the mice to be more susceptible to stress [16,17]. The observation of LTP-related genes shows rapid methylation and demethylation following LTP induction [18]. PTMs play significant roles in the maintenance of stress and depression.
MDD, as defined by the American Psychiatric Association, is a common and serious mental disorder accompanied by the loss of interest and pleasure in enjoyable activities (anhedonia), changes in appetite, loss of energy, insomnia, psychomotor retardation, and suicidal thoughts [19]. MDD can be potentially fatal, with an estimated one in six people having experienced depression once during their lifetime [20,21,22]. The exact causes of MDD remain elusive, with possibilities ranging from neural biochemistry to personality [23,24]. The lack of biomarkers, inconclusive genetic studies, and personal effects such as stress all contribute to the poor understanding of MDD, which also plays a role in antidepressant studies that have led to disheartening results [25,26,27]. Serotonin selective reuptake inhibitors (SSRIs), the most widely prescribed antidepressant drugs, are ineffective in up to 45% of participants even after weeks to months of drug treatment [22,28,29,30]. Although the exact mechanisms of MDD pathology remain unclear, global efforts have been made to clarify the cellular mechanisms leading to MDD. Post-mortem studies of depressed participants have shown a reduced number of prefrontal cortex (PFC) synapses, which have been logically connected to reduced synaptic functions [31]. Furthermore, the volumes of the PFC and hippocampus have been observed to shrink in individuals with depression, with the extent of shrinkage positively correlated to the severity of depression the person is experiencing [32,33].
Small ubiquitin-like modifier (SUMO)ylation is a relatively new form of PTM that is currently under careful investigation by scientists worldwide. The first mention of the direct control of the synapse by SUMOylation was in 2009, by Feligioni et al. By increasing and decreasing presynaptic SUMOylation levels in synaptosomes, they discovered fluctuations in Ca2+ influx. Increasing levels of SUMOylation resulted in reduced Ca2+ influx and decreased glutamate release, whereas the opposite was observed by reducing presynaptic SUMOylation levels [34]. In 2014, SUMOylation was demonstrated to be required for LTP activation. Using inactive forms of SUMO-conjugating enzymes (Ubc9), LTP was shown to be significantly reduced, and increasing sentrin/SUMO-specific protease 1 (SNEP1) levels showed the same effect in CA1 hippocampal slices [35]. Compared to other forms of PTMs, such as acetylation and methylation, the mechanisms and effects of SUMOylation are still elusive, but they seem to have a noteworthy effect on synaptic plasticity and overall neuronal health. Here, we review the regulation of neuronal function by SUMOylation of proteins associated with various methods of synaptic plasticity control connected to the pathogenesis of MDD.

2. SUMOylation

SUMO proteins are a family of proteins that are conjugated to lysine residues on target proteins. SUMO is ligated to a ΨKX(D/E) consensus motif of the target proteins, where Ψ represents a hydrophobic residue, K is a target lysine, X is any residue, and D/E represents acidic residues [36]. There are five isoforms of SUMO found in humans, SUMO1–5. While SUMO4 and 5 still require further investigation, the functions of SUMO 1~3 have been extensively investigated. SUMO1 is involved in monoSUMOylation (attachment of SUMO to a single lysine residue) and regulates protein localization and ubiquitination. SUMO2 and 3 are nearly indistinguishable and are thus referred to as SUMO2/3. SUMO2/3 is involved in polySUMOylation, forming SUMO chains on target proteins, and seems to respond to cellular stresses such as heat shock, DNA damage, and nutrient deficiency [37,38]. SUMOylation occurs in a cycle (Figure 1) and relies on the activation of four types of enzymes known as the Ulp/SENP family, E1 SUMO-activating, E2 SUMO-conjugating, and E3 SUMO-ligase. The SUMO peptide is first matured for conjugation by cleaving the C-terminus to reveal a glycine-glycine motif. The E1 enzyme, an SAE1/2 heterodimer, binds to the exposed motif, transferring SUMO to the E2 enzyme Ubc9. The E3 enzyme then transfers SUMO from Ubc9 to the target protein, completing SUMOylation [39]. There are three main E3 ligase families: SP-RING ligases, which include the Siz and PIAS family; RAN binding protein 2 (RanBP2), which is a part of the nuclear pore complex; and human polycomb protein 2 (hPc2) [40]. SUMOylation is highly dynamic; thus, SUMO peptides can be removed from the substrate by Ulp/SENP proteases, allowing SUMO peptides to be recycled [39]. There are six known members of the SENP protease family, SENP1–7 (excluding SENP4), each with specific functions and localizations. SENP1 functions in removing SUMO1 preferentially over SUMO2/3, whereas other members specialize in removing SUMO2/3 [41,42]. Three other SUMO proteases have been recently discovered, but their functions have not been extensively investigated [43]. SUMOylation can affect target proteins by blocking protein-protein interaction sites, providing binding sites for other regulatory factors, controlling protein stability, or causing conformational changes in target proteins. These changes by SUMOylation have numerous effects, including regulating transcription, cell cycle progression, DNA repair, protein trafficking, and mitochondrial dynamics, among many other functions [44,45,46,47,48,49,50,51,52]. Because of these effects, SUMOylation has been of interest to neurobiologists worldwide. SUMO1 and SUMO2/3 have been detected in all parts of the neuron, from the nucleus to the synapse. In conjunction, SUMO isoforms, Ubc9 SENPs, and PIAS proteins have been detected in synaptosomes, implying that SUMOylation may have a role in regulating neurotransmitter release and general synaptic function [53,54].

3. SUMOylation in the Synapse

Synaptic plasticity is associated with many cellular mechanisms, such as protein trafficking, translation, and transcription [55]. LTP and LTD are mediated through NMDAR and subsequent AMPAR translocation and activity, mainly through the Ca2+ influx caused by NMDARs. Increased levels of cellular Ca2+ lead to the activation of multiple signal transduction pathways, including the neuronal nitric oxide synthase/nitric oxide (nNOS/NO) signaling and Ras-ERK pathways. In addition to NMDARs and AMPARs, G-protein coupled receptors and serotonin receptors have also been shown to be instrumental in plasticity regulation. In this section, we summarize current studies regarding the effects of SUMOylation on specific proteins and pathways.

3.1. AMPARs

AMPARs are glutamate-gated ion channels responsible for rapid excitatory synaptic transmission in the central nervous system. AMPARs are comprised of four subunits that form ion channels with different functions [56]. To induce NMDAR-mediated LTP, AMPARs must be concentrated in the postsynaptic membrane. It is widely believed that this trafficking is possible due to a large number of PTM-induced modifications on AMPAR subunits, including glycosylation, palmitoylation, and phosphorylation [57]. However, a bacterial SUMOylation assay of neuronal proteins to test for SUMOylation core motifs has shown that AMPAR subunits are not SUMOylated [58]. When inducing LTP in cultured hippocampal slices, an increase in SUMO2/3 was observed, stipulating a connection between LTP induction and SUMOylation [35]. When LTP was induced in cultured hippocampal slices, an increase in SUMO2/3 was observed, indicating a connection between LTP induction and SUMOylation [25]. Furthermore, increasing surface level AMPARs by chemically inducing LTP caused an increase in SUMO1, Ubc9, and SUMO1 RNAs; however, overexpression of SENP1 prevented these increases [59]. Additionally, by silencing SUMO1-3 expression in mice, malfunctions in episodic and fear memory conditioning have been observed [60]. Synaptic plasticity plays an imperative role in fear memory conditioning; thus, SUMOylation is required for the proper trafficking of AMPARs. An AMPAR trafficking-related factor that has been identified as SUMOylated is an activity-regulated cytoskeleton-associated protein (Arc) (Figure 2). Arc levels increase rapidly in response to neuronal activity and have been shown to increase during periods of stress in rodent models [61]. Arc has been shown to participate in AMPAR trafficking and ensuing LTP and LTD induction [62,63]. Arc has been reported to be SUMOylated at lysine residues 110 and 268, and SUMOylation of these residues regulates protein-protein interactions and synaptic scaling [64]. In vivo studies have shown that Arc SUMOylation is involved in LTP induction and vice versa; LTP induction gives rise to Arc SUMOylation, which is repressed when LTP is inhibited [65].

3.2. nNOS/NO Signaling

Nitric oxide is a secondary messenger found in mammalian cells and is produced by NOS enzymes. nNOS is an NOS isoform found in neurons and is important for proper neuronal function. Post-mortem studies of MDD participants have found increased levels of nNOS expression in the hippocampus, whereas inhibition of nNOS prevented depression-like behavior in rodent models exposed to chronic mild stress [66,67]. A study by Du et al. has shed light on nNOS SUMOylation, showing that nNOS is SUMOylated at lysine residues 725 and 739. SUMOylation of these residues promotes nNOS phosphorylation, which increases NO production (Figure 2). By disrupting nNOS pathways, activity-induced nNOS SUMOylation also decreases LTP-related Arc expression and blocks ERK 1/2 signaling [68].

3.3. Ras/Raf/MEK/ERK Pathway

The Ras/Raf/mitogen-activated protein kinase/ERK kinase (MEK)/extracellular signal-regulated kinase (ERK) signaling cascade (shortened to the ERK pathway) relays signals from cell surface receptors to transcription factors to regulate gene expression [69]. The ERK pathway has been the most studied signal transduction pathway, with many reports of decreased ERK signaling in the PFC and hippocampus of suicidal MDD participants [70,71,72]. Decreased MEK-ERK signaling was detected in the hippocampus of suicidal participants, while MAPK phosphatase (MKP), a MAPK inhibitor, was increased [73]. Duric et al. (2010) discovered increased levels of MKP in mice undergoing chronic unpredictable stress, with ERK1/2 signaling decreased in the PFC and hippocampus, while deletion of the Mkp-1 gene resulted in resistance to stress [72]. Ras activation has also been implicated in the trafficking of AMPARs during LTP [74]. All isoforms of Ras are SUMOylated at lysine 42, and SUMOylation of K42 is crucial for Ras-promoted signaling, as the K42R mutation causes reductions in downstream activation of the ERK signaling cascade (Figure 2) [75,76].
Downstream of the ERK signaling cascade, cAMP-responsive element-binding protein (CREB) has also been shown to decrease in the hippocampus of MDD participants [72]. As a phosphorylation target of the ERK signaling cascade, CREB is a transcription factor required for the transcription of genes that regulate neuronal plasticity. Chronic stress models have also been used to test these results; indeed, chronic stress leads to lower CREB activity in rodent models, indicating that stress can lower CREB levels in the brain, which could lead to MDD development [77,78]. By increasing levels of PIAS1 or NMDA in rodent models, a research team has shown enhanced SUMOylation of CREB in the hippocampus. They demonstrated two predominant lysine residues that are SUMOylated (lysine 271 and 290). Interestingly, preventing SUMOylation increases the phosphorylation of CREB, while preventing CREB phosphorylation inhibits CREB SUMOylation. These data imply cross-talk between the two types of PTMs and require further research [79]. In addition to the above-mentioned effects, CREB SUMOylation also increases the transcription levels of brain-derived neurotrophic factor (BDNF), which we will discuss in greater detail in the next section of this review.

3.4. BDNF

BDNF is a growth factor involved in many central functions of the brain, including synaptic plasticity [80]. BDNF functions by binding to tropomyosin receptor kinase B (TrkB) in the cellular membrane and triggering a signaling cascade that activates various pathways, including the ERK cascade and CREB [81]. Thus, BDNF has been studied with great interest regarding mood disorders, and some antidepressants act upon BDNF-TrkB [82]. Environmental stress that triggers depression can lower BDNF mRNA levels [83]. While BDNF has been thoroughly studied in other fields, SUMOylation of BDNF has not yet been elucidated. A recent study showed that BDNF-TrkB signaling can alter the subcellular localization of SUMOylation enzymes. PIAS3 is functionally regulated and translocated from the nucleus to the dendrites via the ERK1/2 pathway, which is downstream of BDNF-TrkB [84].
A downstream event of BDNF-TrkB signaling is the activation of the P13K-Akt pathway. In MDD participants, reduced levels of P13K-Akt were observed in the PFC, while antidepressant treatment was shown to increase Akt signaling [85,86]. Akt1 is SUMOylated at lysine 64 and 276 by PIAS3 following NMDAR activity. SUMOylation of these residues increases Akt enzymatic activity, and inactivation of PIAS3 impairs LTP expression [87].

3.5. SNARE Complex

In yeast, mammalian, and plant cells, vesicle fusion with the cellular membrane is mediated by a complex of proteins in the soluble N-ethylmaleimide-sensitive factor attachment protein receptor (SNARE) protein family, commonly known as the SNARE complex [88]. SNAREs have been thoroughly studied in the brain, as they mediate neurotransmitter release through synaptic vesicles. SNARE proteins are categorized into synaptosomal-associated proteins (SNAP), syntaxins (stx), and target SNAREs (t-SNAREs), depending on their location within the synapse [89]. Synapsin 1a (Syn1a) is a protein that regulates vesicle availability at the pre-synapse. There is evidence showing that Syn1a is SUMOylated at lysine 687 and mutations to K687R result in a decrease in available releasable vesicles and impaired exocytosis, indicating SUMOylation as a Syn1a function regulator [90]. Stx1a is another member of the SNARE family, which is involved in the release of neurotransmitters and the recycling of synaptic vesicle membranes and proteins from the plasma membrane [91]. Upon NMDAR activation, Stx1a is SUMOylated at three lysine residues (K252, K253, and K256), and mutations of the three lysine residues to arginine cause an imbalance in synaptic vesicle endocytosis and exocytosis, indicating that Stx1a SUMOylation is a regulator of neurotransmitter release [92]. Rab3-Interacting molecule 1α (RIM1α) regulates vesicle release from synapses. Further research has shown that RIM1α plays a crucial role in the docking of synaptic vesicles [93]. Studies have shown that RIM1α is SUMOylated at lysine 502, which aids in the clustering of Cav2.1 calcium channels. K502R mutations in RIM1α decrease synaptic exocytosis [94].

3.6. Ion Channels

Neuronal excitability is maintained by the properties and functionality of ion channels found in the plasma membranes of neurons. Voltage-dependent potassium channels (Kv) are gated neuronal transmembrane channels and are the most abundant type of ion channel. Kvs are essential for action potential generation and are phosphorylated during LTP to increase neuronal excitability [95]. There have been several reports of Kv SUMOylation. Kv1.1 was modified by SUMO1/2 and interacted with SENP2 in hippocampal neurons. Through SENP2 knock-out, hyperSUMOylation was observed on Kv1.1, Kv7.2, and Kv7.3 in rodent model hippocampal neurons. Kv1.1 hyperSUMOylation did not appear to have any effect on channel activity. Kv7.2 and 7.3 SUMOylation resulted in a diminished M-current, which is responsible for neuronal hyperexcitability. Neuronal hyperexcitability is known to cause epilepsy, neuropathic pain, and degeneration [96,97]. Kv1.5 SUMOylation regulates channel inactivation, while K4.2 is SUMOylated at two lysine residues, K437 and K579, which increase Kv4.2 surface expression while decreasing current conductivity [98,99]. Similar effects have been observed by SUMOylation of the Kv2.1, Kv7.1, and Kv11.1 channels [100,101,102]. However, studies on sodium channels have been scarce, with only one study implicating the results of direct voltage-gated sodium channel (NaV) SUMOylation. SUMOylation of NaV1.2 resulted in increased current conductivity. Collapsin response mediator protein 2 (CRMP2) has been previously shown to regulate NaV1.7 channels and has been reported as a SUMOylation target. A K374A mutation in CRMP2 causes a reduction in NaV1.7 current conductivity by interfering with CRMP2 and NaV1.7 interactions [103,104].

3.7. Receptors

3.7.1. Serotonin Receptors

Changes in serotonin receptors are a critical step in MDD development. There are seven classes of serotonin receptors, each with a different function. Of these classes, 5-HT1A primarily regulates reward processing, mobility, appetite, and anxiety [105]. Recently, 5-HT1A was reported to be SUMOylated by SUMO1 in rodent brains. Overexpression of PIAS in cells has been shown to increase SUMOylation, while SENP2 decreases SUMOylation of 5-HT1A, although the implications of 5-HT1A SUMOylation are yet to be discovered [106].

3.7.2. Dopamine Transporters

Dopamine is a neuromodulator involved in stimulus-reward learning processes and behavioral control. Dopamine functions by altering the properties of neurons, such as membrane excitability, neurotransmitter release, and protein trafficking [107]. Dopamine is regulated by dopamine transporters (DAT), which reuptake dopamine back into the pre-synapse. Overexpression of Ubc9 and SUMO1 enhanced DAT surface expression and stability, and increased DAT functionality, whereas knockdown of Ubc9 resulted in lower SUMO1-DAT levels and increased DAT degradation [108].

3.7.3. G-Protein Coupled Receptors (GPCRs)

Only a handful of GPCRs are SUMOylated in eukaryotes. mGluR7 and mGluR8 are located in the presynaptic region and have different functions. mGluR7 plays a vital role in controlling excitatory synapse function, whereas mGluR8 controls Ca2+-dependent neurotransmitter release [106,109]. mGluR7 is SUMOylated at lysine 889, whereas mGluR8 is SUMOylated at lysine 882 and 903. SUMOylation was found to regulate the endocytosis of these receptors [110,111]. Interestingly, mGluR7 shows different SUMOylation patterns depending on the cell type. In HEK293T cells, mGluR7 is SUMOylated by both SUMO1 and SUMO2/3, whereas only SUMO1 is found in hippocampal cells [112]. Additionally, mGluR5 activation reduces the diffusion of Ubc9 out of the dendritic membrane, increasing Ubc9 levels inside neurons [113]. Similarly, SENP1 levels also increase in dendrites upon mGluR5 activation [114]. The M1 muscarinic acetylcholine receptors (M1Rs) are another type of GPCR that is SUMOylated. They mediate signals via acetylcholine and play vital roles in learning and memory. M1R is SUMOylated at lysine 237 and receptor activation removes SUMOylation at this residue. SUMOylation at this site increases the binding capacity, sequentially raising signal transduction levels. K237R mutations decrease SUMOylation and signal transduction through M1R [115]. Another GPCR that has been researched is cannabinoid receptor 1 (CB1). They function as modulators of the voltage-gated calcium and potassium channels. Activation of CB1 has been shown to increase SUMO1 conjugation and free SUMO1 levels in cortical neurons of rodent models, although the exact functions of SUMOylation are yet to be discovered [116].

3.7.4. Kainite Receptors (KARs)

KARs are relatively under-studied members of the glutamate receptor family. AMPARs and NMDARs play important roles in neuronal function and the regulation of plasticity, such as regulating neurotransmitter release and membrane excitability [117]. A subunit of KAR, GluK2, is SUMOylated singly at the lysine 886 residue in rodent hippocampal neurons. KAR activation by glutamate binding caused K886 to be SUMOylated, and the KAR was internalized, which was confirmed by infusing SUMO1 into the post-synapse and detecting lower levels of KAR signals [53].

3.8. Cytoplasmic Polyadenylation Element-Binding Protein 3 (CPEB3)

CPEB3 exists in a soluble inactive form or as an insoluble active aggregate and regulates the synthesis of synaptic proteins. Appropriate localization and activation are crucial for the function of CPEB3 in LTP [118]. SUMOylation is necessary for the accurate localization of CPEB3 to processing bodies, implicating functions related to translation. Under basal conditions, CPEB3 is SUMOylated by SUMO2/3, which is then rapidly deSUMOylated during LTP transduction, and subsequently translocated into polysomes to aid translation [118,119]. Other reports have shown that SUMO2 conjugation may play a role in silencing the local translation after a pulse of neuronal activity. SUMOylation by SUMO2 shifts the binding affinity of mRNA from deSUMOylated CPEB3 aggregates to inactive SUMOylated CPEB3 monomers [107].

4. Mitochondrial Dysfunction

Commonly known as ‘the powerhouse of the cell’, mitochondria have major functions in neuronal processes, such as Ca2+ regulation, plasma membrane potential maintenance, the release of neurotransmitters, and maintenance of high energy levels [120]. It is not surprising that MDD is linked to mitochondrial function. Chronic mild stress induced in mice resulted in the inhibition of oxidative phosphorylation and dissipated mitochondrial membrane potential in various brain regions, including the hippocampus [121]. Glucose utilization by the PFC and other brain regions is reduced in depressed human subjects, implying that mitochondrial health is a requirement for regulating mood [122]. Mitochondrial health is balanced by the tight regulation of mitochondrial fission, fusion, and mitophagy. Fission is responsible for mitochondrial renewal, redistribution, and partitioning of damaged parts of the mitochondria, while fusion is the process of neighboring mitochondria mixing their contents to restore function after being damaged [120,123]. Mitophagy is a ‘cleaning’ mechanism that removes dysfunctional mitochondria from the cell through autophagy [124]. Although no studies have directly linked SUMO-mediated mitochondrial dysfunction to MDD, depression is a common symptom in participants with Parkinson’s disease [125].

4.1. Dynamin-Related Protein 1 (Drp1)

A SUMOylation target that has been extensively studied in relation to neuronal dysfunction in Parkinson’s disease is the cytosolic GTPase Drp1. Drp1 is recruited to the outer mitochondrial membrane to regulate mitochondrial fission [126,127]. Drp1 is regulated by several PTMs, including SUMOylation. Drp1 has been found to be SUMOylated at various lysine residues throughout the protein. Drp1 interacts with Ubc9 and is SUMOylated by all SUMO isoforms [128]. SUMO-1 overexpression prevents Drp1 degradation, leading to enhanced mitochondrial fission and subsequent apoptosis (Figure 3) [129]. Intriguingly, a study in 2013 by Guo et al. showed that SENP3 degradation caused increased conjugation of SUMO2/3 to Drp1, which inhibited mitochondrial fission [130]. Drp1 activity is regulated by SENPs. SENP5 overexpression causes a reduction in SUMO-1-induced mitochondrial fission, while the opposite is true with the knockdown of SENP2 and SENP5 [131,132].

4.2. DJ-1

There have been previous reports of increased reactive oxygen species (ROS) production in participants with depression [133]. Under oxidative stress conditions, the protein deglycase DJ-1 is translocated into the mitochondria, where it functions as a mitochondrial complex I regulator and contributes to protecting the cell from oxidative stress [134,135]. DJ-1 must be suitably SUMOylated at K130 to function. Mutations that interfere with DJ-1 SUMOylation were found to perturb the antioxidant functions of DJ-1. It was found that improper SUMOylation decreases DJ-1 solubility, leaving the protein susceptible to proteasomal degradation [136]. SENP3 regulates mitophagy by interacting with Fis1, a protein required for mitophagy. Fis1 was discovered to be SUMOylated at lysine 149, and it was found that K149 SUMOylation prevents Fis1 from causing mitophagy (Figure 3) [137]. Interestingly, DJ-1 was found to promote proteasomal degradation of Fis1 in response to oxidative stress [138]. Additionally, Fis1 recruits Drp1 from the cytosol to the mitochondrial outer membrane [139]. These results suggest an interesting response to oxidative stress, involving lower levels of mitophagy and increased fission mediated by SUMOylation.

4.3. Peroxisomal Proliferator-Activated Receptor-γ Coactivator 1 α (PGC-1α)

PGC-1α is a transcriptional coactivator of energy metabolism genes and a key regulator of mitochondrial biogenesis and function [140]. SUMO-1 conjugation at K138 inhibits PGC-1α transcriptional activity [141]. SENP1-mediated removal of SUMO-1 can recover the negative effects caused by SUMOylation and PGC-1α transcriptional activity [142]. PGC-1α is under the indirect control of DJ-1, mediated by pyrimidine tract-binding protein-associated splicing factor (PSF, Figure 3). DJ-1 inhibits SUMOylation of PSF, which in turn binds to PGC-1α and suppresses its transcriptional activity [143]. This demonstrates a link between oxidative stress and SUMO-mediated abnormal mitochondrial gene expression.

4.4. Parkin

Parkin is a ubiquitin E3 ligase, which is a crucial regulator of mitochondrial fission and fusion. Parkin can bind and ubiquitinate Drp1 to promote Drp1 degradation and prevent mitochondrial fission. This effect can be nullified by SUMOylation of parkin by SUMO-1. SUMO-1 conjugation results in parkin being translocated into the nucleus, reducing the amount available for activity in the mitochondria [144].

4.5. Parkin-Interacting Substrate (PARIS)

PARIS is a transcriptional repressor expression of PGC-1α. PARIS levels are regulated by the ubiquitin E3 ligase activity of parkin (Figure 3) [145]. PARIS is SUMOylated with SUMO1 and SUMO2/3 at K189 and K286, respectively, by PIASy, while the effects of SUMOylation are unclear and dependent on cell type. SUMOylation of HepG2 cells did not affect PARIS activity, while SH-SY5Y cells with 2KR PARIS mutations showed restoration of repression effects. Conversely, in HEK293 cells, 2KR mutants show increased transcriptional repression [146]. In subsequent studies, SUMOylation of PARIS with SUMO2/3 induced ubiquitination and proteasomal degradation, thereby relieving PARIS-mediated transcriptional repression [147]. Although there have been few direct connections between MDD and mitochondria-related SUMOylation, the above data show that SUMOylation plays a significant role in mitochondria-mediated neuronal health, and future studies into MDD-related mitochondrial dysfunction could reveal prospective methods of treatment.

5. Conclusions

Since its discovery in 1995, SUMOylation has been identified as a key regulator of many functions in eukaryotic cells. Although yet unclear, the above results do show that SUMOylation plays an integral role in the proper regulation of neural function. Recently, SUMOylation has been researched using non-neuronal models, showing promising results. An example would be the control of Fis1 localization in HeLa cells by SENP3-mediated SUMOylation or the localization of mitochondria in HEK-293 embryonic kidney cells by mitofusin1/2 mediated by SUMO2 modifications [137,148]. Mitophagy in neurons has recently been reviewed by Doxaki and Palikaras, indicating the importance of mitophagy in neuronal survival and cellular homeostasis [149]. As there are currently no studies on the effects of mitophagy-related SUMOylation on neuronal function, this seems to be a promising field for further study. In addition, chronic pain has been linked to depression in many clinical studies, with some reports stating that up to 85% of participants experiencing chronic pain are also affected by severe depression [150]. Recent studies have reported on the involvement of SUMOylation in pain regulation. One study showed that preventing SUMOylation of CRMP2, a NaV1.7 channel interacting protein, was antinociceptive in chronic and acute pain models using rodents [151]. A different study on neuropathic pain management reported that enhancing SUMOylation of DGCR8 elevated sensitivity to pain in rat models of spinal nerve ligation [152]. These results imply that targeting SUMOylation is a promising new approach for pain relief and antidepressant medication. Experiments with existing antidepressants have shown the integral role of SUMOylation. Tricyclic antidepressants inhibited FKBP51 (a glucocorticoid receptor inhibitor) SUMOylation by blocking interactions with PIAS4 and subsequently restoring glucocorticoid receptor activity [153]. Although inconclusive, recent advances have shown promising results, suggesting that targeting of SUMOylation-related enzymes or SUMOylated proteins may serve as a potential therapeutic target in the future.

Author Contributions

Conceptualization and writing, S.-W.J. and H.-Y.R.; review and editing, H.-S.P., Z.Y.R., D.-H.C. and H.-S.L.; funding acquisition, H.-Y.R. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by a National Research Foundation of Korea (NRF) grant funded by the South Korean government (MSIT) (nos. 2020R1C1C1009367 and 2020R1A4A1018280) and Korean Environment Industry & Technology Institute (KEITI) through Core Technology Development Project for Environmental Diseases Prevention and Management funded by Korean Ministry of Environment (MOE) (no. 2022003310001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Citri, A.; Malenka, R.C. Synaptic Plasticity: Multiple Forms, Functions, and Mechanisms. Neuropsychopharmacology 2008, 33, 18–41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Bliss, T.V.P.; Cooke, S.F. Long-term potentiation and long-term depression: A clinical perspective. Clinics 2011, 66 (Suppl. 1), 3–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Richter-Levin, G.; Xu, L. How could stress lead to major depressive disorder? IBRO Rep. 2018, 4, 38–43. [Google Scholar] [CrossRef]
  4. Magariños, A.M.; McEwen, B.S.; Flügge, G.; Fuchs, E. Chronic psychosocial stress causes apical dendritic atrophy of hippocampal CA3 pyramidal neurons in subordinate tree shrews. J. Neurosci. 1996, 16, 3534–3540. [Google Scholar] [CrossRef] [Green Version]
  5. Traynelis, S.F.; Wollmuth, L.P.; McBain, C.J.; Menniti, F.S.; Vance, K.M.; Ogden, K.K.; Hansen, K.B.; Yuan, H.; Myers, S.J.; Dingledine, R. Glutamate receptor ion channels: Structure, regulation, and function. Pharmacol. Rev. 2010, 62, 405–496. [Google Scholar] [CrossRef] [Green Version]
  6. Malenka, R.C.; Kauer, J.A.; Perkel, D.J.; Mauk, M.D.; Kelly, P.T.; Nicoll, R.A.; Waxham, M.N. An essential role for postsynaptic calmodulin and protein kinase activity in long-term potentiation. Nature 1989, 340, 554–557. [Google Scholar] [CrossRef]
  7. Silva Alcino, J.; Stevens Charles, F.; Tonegawa, S.; Wang, Y. Deficient Hippocampal Long-Term Potentiation in α-Calcium-Calmodulin Kinase II Mutant Mice. Science 1992, 257, 201–206. [Google Scholar] [CrossRef]
  8. Lisman, J. The CaM kinase II hypothesis for the storage of synaptic memory. Trends Neurosci. 1994, 17, 406–412. [Google Scholar] [CrossRef]
  9. Malenka Robert, C.; Nicoll, R.A. Long-Term Potentiation—A Decade of Progress? Science 1999, 285, 1870–1874. [Google Scholar] [CrossRef] [Green Version]
  10. Slipczuk, L.; Bekinschtein, P.; Katche, C.; Cammarota, M.; Izquierdo, I.; Medina, J.H. BDNF Activates mTOR to Regulate GluR1 Expression Required for Memory Formation. PLoS ONE 2009, 4, e6007. [Google Scholar] [CrossRef]
  11. Wu, H.; Zhou, Y.; Xiong, Z.-Q. Transducer of regulated CREB and late phase long-term synaptic potentiation. FEBS J. 2007, 274, 3218–3223. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Kristensen, A.S.; Jenkins, M.A.; Banke, T.G.; Schousboe, A.; Makino, Y.; Johnson, R.C.; Huganir, R.; Traynelis, S.F. Mechanism of Ca2+/calmodulin-dependent kinase II regulation of AMPA receptor gating. Nat. Neurosci. 2011, 14, 727–735. [Google Scholar] [CrossRef] [PubMed]
  13. Bingol, B.; Schuman, E.M. Activity-dependent dynamics and sequestration of proteasomes in dendritic spines. Nature 2006, 441, 1144–1148. [Google Scholar] [CrossRef] [Green Version]
  14. Burgado, J.; Harrell, C.S.; Eacret, D.; Reddy, R.; Barnum, C.J.; Tansey, M.G.; Miller, A.H.; Wang, H.; Neigh, G.N. Two weeks of predatory stress induces anxiety-like behavior with co-morbid depressive-like behavior in adult male mice. Behav. Brain Res. 2014, 275, 120–125. [Google Scholar] [CrossRef] [Green Version]
  15. Hollis, F.; Kabbaj, M. Social defeat as an animal model for depression. ILAR J. 2014, 55, 221–232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Covington, H.E., 3rd; Maze, I.; LaPlant, Q.C.; Vialou, V.F.; Ohnishi, Y.N.; Berton, O.; Fass, D.M.; Renthal, W.; Rush, A.J., 3rd; Wu, E.Y.; et al. Antidepressant actions of histone deacetylase inhibitors. J. Neurosci. Off. J. Soc. Neurosci. 2009, 29, 11451–11460. [Google Scholar] [CrossRef]
  17. Covington, H.E., 3rd; Maze, I.; Sun, H.; Bomze, H.M.; DeMaio, K.D.; Wu, E.Y.; Dietz, D.M.; Lobo, M.K.; Ghose, S.; Mouzon, E.; et al. A role for repressive histone methylation in cocaine-induced vulnerability to stress. Neuron 2011, 71, 656–670. [Google Scholar] [CrossRef] [Green Version]
  18. Maag, J.L.V.; Kaczorowski, D.C.; Panja, D.; Peters, T.J.; Bramham, C.R.; Wibrand, K.; Dinger, M.E. Widespread promoter methylation of synaptic plasticity genes in long-term potentiation in the adult brain in vivo. BMC Genom. 2017, 18, 250. [Google Scholar] [CrossRef] [Green Version]
  19. Edition, F. Diagnostic and statistical manual of mental disorders. Am. Psychiatr. Assoc. 2013, 21, 591–643. [Google Scholar]
  20. Kessler, R.C.; Berglund, P.; Demler, O.; Jin, R.; Koretz, D.; Merikangas, K.R.; Rush, A.J.; Walters, E.E.; Wang, P.S. The epidemiology of major depressive disorder: Results from the National Comorbidity Survey Replication (NCS-R). JAMA 2003, 289, 3095–3105. [Google Scholar] [CrossRef]
  21. Walker, E.R.; McGee, R.E.; Druss, B.G. Mortality in mental disorders and global disease burden implications: A systematic review and meta-analysis. JAMA Psychiatry 2015, 72, 334–341. [Google Scholar] [CrossRef] [PubMed]
  22. Trivedi, M.H.; Rush, A.J.; Wisniewski, S.R.; Nierenberg, A.A.; Warden, D.; Ritz, L.; Norquist, G.; Howland, R.H.; Lebowitz, B.; McGrath, P.J.; et al. Evaluation of outcomes with citalopram for depression using measurement-based care in STAR*D: Implications for clinical practice. Am. J. Psychiatry 2006, 163, 28–40. [Google Scholar] [CrossRef]
  23. Farabaugh, A.M.Y.; Mischoulon, D.; Fava, M.; Guyker, W.; Alpert, J. The Overlap Between Personality Disorders and Major Depressive Disorder (MDD). Ann. Clin. Psychiatry 2004, 16, 217–224. [Google Scholar] [CrossRef] [PubMed]
  24. Boku, S.; Nakagawa, S.; Toda, H.; Hishimoto, A. Neural basis of major depressive disorder: Beyond monoamine hypothesis. Psychiatry Clin. Neurosci. 2018, 72, 3–12. [Google Scholar] [CrossRef] [Green Version]
  25. Al-Harbi, K.S. Treatment-resistant depression: Therapeutic trends, challenges, and future directions. Patient Prefer. Adherence 2012, 6, 369–388. [Google Scholar] [CrossRef] [Green Version]
  26. Duncan, L.E.; Keller, M.C. A Critical Review of the First 10 Years of Candidate Gene-by-Environment Interaction Research in Psychiatry. Am. J. Psychiatry 2011, 168, 1041–1049. [Google Scholar] [CrossRef] [PubMed]
  27. Hacimusalar, Y.; Eşel, E. Suggested Biomarkers for Major Depressive Disorder. Noro Psikiyatr. Ars. 2018, 55, 280–290. [Google Scholar] [CrossRef]
  28. Mayberg, H.S. Defining the Neural Circuitry of Depression: Toward a New Nosology with Therapeutic Implications. Biol. Psychiatry 2007, 61, 729–730. [Google Scholar] [CrossRef]
  29. Kessler, R.C.; Chiu, W.T.; Demler, O.; Merikangas, K.R.; Walters, E.E. Prevalence, severity, and comorbidity of 12-month DSM-IV disorders in the National Comorbidity Survey Replication. Arch. Gen. Psychiatry 2005, 62, 617–627. [Google Scholar] [CrossRef] [Green Version]
  30. Mrazek, D.A.; Hornberger, J.C.; Altar, C.A.; Degtiar, I. A review of the clinical, economic, and societal burden of treatment-resistant depression: 1996–2013. Psychiatr. Serv. 2014, 65, 977–987. [Google Scholar] [CrossRef]
  31. Kang, H.J.; Voleti, B.; Hajszan, T.; Rajkowska, G.; Stockmeier, C.A.; Licznerski, P.; Lepack, A.; Majik, M.S.; Jeong, L.S.; Banasr, M.; et al. Decreased expression of synapse-related genes and loss of synapses in major depressive disorder. Nat. Med. 2012, 18, 1413–1417. [Google Scholar] [CrossRef] [PubMed]
  32. MacQueen, G.M.; Yucel, K.; Taylor, V.H.; Macdonald, K.; Joffe, R. Posterior hippocampal volumes are associated with remission rates in patients with major depressive disorder. Biol. Psychiatry 2008, 64, 880–883. [Google Scholar] [CrossRef]
  33. Chan, S.W.; Harmer, C.J.; Norbury, R.; O’Sullivan, U.; Goodwin, G.M.; Portella, M.J. Hippocampal volume in vulnerability and resilience to depression. J. Affect. Disord. 2016, 189, 199–202. [Google Scholar] [CrossRef] [PubMed]
  34. Feligioni, M.; Nishimune, A.; Henley, J.M. Protein SUMOylation modulates calcium influx and glutamate release from presynaptic terminals. Eur. J. Neurosci. 2009, 29, 1348–1356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Lee, L.; Dale, E.; Staniszewski, A.; Zhang, H.; Saeed, F.; Sakurai, M.; Fa, M.; Orozco, I.; Michelassi, F.; Akpan, N.; et al. Regulation of synaptic plasticity and cognition by SUMO in normal physiology and Alzheimer’s disease. Sci. Rep. 2014, 4, 7190. [Google Scholar] [CrossRef] [Green Version]
  36. Zhao, X. SUMO-Mediated Regulation of Nuclear Functions and Signaling Processes. Mol. Cell 2018, 71, 409–418. [Google Scholar] [CrossRef] [Green Version]
  37. Zhou, W.; Ryan, J.J.; Zhou, H. Global analyses of sumoylated proteins in Saccharomyces cerevisiae. Induction of protein sumoylation by cellular stresses. J. Biol. Chem. 2004, 279, 32262–32268. [Google Scholar] [CrossRef] [Green Version]
  38. Enserink, J.M. Sumo and the cellular stress response. Cell Div. 2015, 10, 4. [Google Scholar] [CrossRef] [Green Version]
  39. Johnson, P.R.; Hochstrasser, M. SUMO-1: Ubiquitin gains weight. Trends Cell Biol. 1997, 7, 408–413. [Google Scholar] [CrossRef]
  40. Krumova, P.; Weishaupt, J.H. Sumoylation in neurodegenerative diseases. Cell Mol. Life Sci. 2013, 70, 2123–2138. [Google Scholar] [CrossRef]
  41. Hickey, C.M.; Wilson, N.R.; Hochstrasser, M. Function and regulation of SUMO proteases. Nat. Rev. Mol. Cell Biol. 2012, 13, 755–766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Kunz, K.; Piller, T.; Müller, S. SUMO-specific proteases and isopeptidases of the SENP family at a glance. J. Cell Sci. 2018, 131, jcs211904. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Henley, J.M.; Craig, T.J.; Wilkinson, K.A. Neuronal SUMOylation: Mechanisms, physiology, and roles in neuronal dysfunction. Physiol. Rev. 2014, 94, 1249–1285. [Google Scholar] [CrossRef] [PubMed]
  44. Berndt, A.; Wilkinson, K.A.; Henley, J.M. Regulation of Neuronal Protein Trafficking and Translocation by SUMOylation. Biomolecules 2012, 2, 256–268. [Google Scholar] [CrossRef] [Green Version]
  45. Flotho, A.; Melchior, F. Sumoylation: A Regulatory Protein Modification in Health and Disease. Annu. Rev. Biochem. 2013, 82, 357–385. [Google Scholar] [CrossRef]
  46. Ryu, H.-Y.; Hochstrasser, M. Histone sumoylation and chromatin dynamics. Nucleic Acids Res. 2021, 49, 6043–6052. [Google Scholar] [CrossRef]
  47. Zhao, J. Sumoylation regulates diverse biological processes. Cell. Mol. Life Sci. 2007, 64, 3017–3033. [Google Scholar] [CrossRef]
  48. Ryu, H.Y.; Wilson, N.R.; Mehta, S.; Hwang, S.S.; Hochstrasser, M. Loss of the SUMO protease Ulp2 triggers a specific multichromosome aneuploidy. Genes Dev. 2016, 30, 1881–1894. [Google Scholar] [CrossRef] [Green Version]
  49. Ryu, H.-Y.; López-Giráldez, F.; Knight, J.; Hwang, S.S.; Renner, C.; Kreft, S.G.; Hochstrasser, M. Distinct adaptive mechanisms drive recovery from aneuploidy caused by loss of the Ulp2 SUMO protease. Nat. Commun. 2018, 9, 5417. [Google Scholar] [CrossRef] [Green Version]
  50. Ryu, H.-Y.; Su, D.; Wilson-Eisele, N.R.; Zhao, D.; López-Giráldez, F.; Hochstrasser, M. The Ulp2 SUMO protease promotes transcription elongation through regulation of histone sumoylation. EMBO J. 2019, 38, e102003. [Google Scholar] [CrossRef]
  51. Ryu, H.Y.; Zhao, D.; Li, J.; Su, D.; Hochstrasser, M. Histone sumoylation promotes Set3 histone-deacetylase complex-mediated transcriptional regulation. Nucleic Acids Res. 2020, 48, 12151–12168. [Google Scholar] [CrossRef] [PubMed]
  52. Ryu, H.-Y.; Ahn, S.H.; Hochstrasser, M. SUMO and cellular adaptive mechanisms. Exp. Mol. Med. 2020, 52, 931–939. [Google Scholar] [CrossRef] [PubMed]
  53. Martin, S.; Nishimune, A.; Mellor, J.R.; Henley, J.M. SUMOylation regulates kainate-receptor-mediated synaptic transmission. Nature 2007, 447, 321–325. [Google Scholar] [CrossRef] [PubMed]
  54. Loriol, C.; Parisot, J.; Poupon, G.; Gwizdek, C.; Martin, S. Developmental regulation and spatiotemporal redistribution of the sumoylation machinery in the rat central nervous system. PLoS ONE 2012, 7, e33757. [Google Scholar] [CrossRef]
  55. Bae, J.R.; Kim, S.H. Synapses in neurodegenerative diseases. BMB Rep. 2017, 50, 237–246. [Google Scholar] [CrossRef] [Green Version]
  56. Shepherd, J.D.; Huganir, R.L. The cell biology of synaptic plasticity: AMPA receptor trafficking. Annu. Rev. Cell Dev. Biol. 2007, 23, 613–643. [Google Scholar] [CrossRef] [Green Version]
  57. Diering, G.H.; Huganir, R.L. The AMPA Receptor Code of Synaptic Plasticity. Neuron 2018, 100, 314–329. [Google Scholar] [CrossRef] [Green Version]
  58. Wilkinson, K.A.; Nishimune, A.; Henley, J.M. Analysis of SUMO-1 modification of neuronal proteins containing consensus SUMOylation motifs. Neurosci. Lett. 2008, 436, 239–244. [Google Scholar] [CrossRef] [Green Version]
  59. Jaafari, N.; Konopacki, F.A.; Owen, T.F.; Kantamneni, S.; Rubin, P.; Craig, T.J.; Wilkinson, K.A.; Henley, J.M. SUMOylation Is Required for Glycine-Induced Increases in AMPA Receptor Surface Expression (ChemLTP) in Hippocampal Neurons. PLoS ONE 2013, 8, e52345. [Google Scholar] [CrossRef] [Green Version]
  60. Wang, L.; Rodriguiz, R.M.; Wetsel, W.C.; Sheng, H.; Zhao, S.; Liu, X.; Paschen, W.; Yang, W. Neuron-specific Sumo1-3 knockdown in mice impairs episodic and fear memories. J. Psychiatry Neurosci. 2014, 39, 259–266. [Google Scholar] [CrossRef] [Green Version]
  61. Ons, S.; Martí, O.; Armario, A. Stress-induced activation of the immediate early gene Arc (activity-regulated cytoskeleton-associated protein) is restricted to telencephalic areas in the rat brain: Relationship to c-fos mRNA. J. Neurochem. 2004, 89, 1111–1118. [Google Scholar] [CrossRef] [PubMed]
  62. Henley, J.M.; Carmichael, R.E.; Wilkinson, K.A. Extranuclear SUMOylation in Neurons. Trends Neurosci. 2018, 41, 198–210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Chowdhury, S.; Shepherd, J.D.; Okuno, H.; Lyford, G.; Petralia, R.S.; Plath, N.; Kuhl, D.; Huganir, R.L.; Worley, P.F. Arc/Arg3.1 interacts with the endocytic machinery to regulate AMPA receptor trafficking. Neuron 2006, 52, 445–459. [Google Scholar] [CrossRef] [Green Version]
  64. Craig, T.J.; Jaafari, N.; Petrovic, M.M.; Rubin, P.P.; Mellor, J.R.; Henley, J.M. Homeostatic Synaptic Scaling Is Regulated by Protein SUMOylation. J. Biol. Chem. 2012, 287, 22781–22788. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Nair, R.R.; Patil, S.; Tiron, A.; Kanhema, T.; Panja, D.; Schiro, L.; Parobczak, K.; Wilczynski, G.; Bramham, C.R. Dynamic Arc SUMOylation and Selective Interaction with F-Actin-Binding Protein Drebrin A in LTP Consolidation In Vivo. Front. Synaptic Neurosci. 2017, 9, 8. [Google Scholar]
  66. Oh, D.H.; Park, Y.C.; Kim, S.H. Increased glycogen synthase kinase-3β mRNA level in the hippocampus of patients with major depression: A study using the stanley neuropathology consortium integrative database. Psychiatry Investig. 2010, 7, 202–207. [Google Scholar] [CrossRef] [Green Version]
  67. Yazir, Y.; Utkan, T.; Aricioglu, F. Inhibition of Neuronal Nitric Oxide Synthase and Soluble Guanylate Cyclase Prevents Depression-Like Behaviour in Rats Exposed to Chronic Unpredictable Mild Stress. Basic Clin. Pharmacol. Toxicol. 2012, 111, 154–160. [Google Scholar] [CrossRef]
  68. Du, C.-P.; Wang, M.; Geng, C.; Hu, B.; Meng, L.; Xu, Y.; Cheng, B.; Wang, N.; Zhu, Q.-J.; Hou, X.-Y. Activity-Induced SUMOylation of Neuronal Nitric Oxide Synthase Is Associated with Plasticity of Synaptic Transmission and Extracellular Signal-Regulated Kinase 1/2 Signaling. Antioxid. Redox Signal. 2019, 32, 18–34. [Google Scholar] [CrossRef]
  69. Chang, F.; Steelman, L.S.; Lee, J.T.; Shelton, J.G.; Navolanic, P.M.; Blalock, W.L.; Franklin, R.A.; McCubrey, J.A. Signal transduction mediated by the Ras/Raf/MEK/ERK pathway from cytokine receptors to transcription factors: Potential targeting for therapeutic intervention. Leukemia 2003, 17, 1263–1293. [Google Scholar] [CrossRef]
  70. Dwivedi, Y.; Rizavi, H.S.; Zhang, H.; Roberts, R.C.; Conley, R.R.; Pandey, G.N. Aberrant extracellular signal-regulated kinase (ERK)1/2 signalling in suicide brain: Role of ERK kinase 1 (MEK1). Int. J. Neuropsychopharmacol. 2009, 12, 1337–1354. [Google Scholar] [CrossRef] [Green Version]
  71. Dwivedi, Y.; Zhang, H. Altered ERK1/2 Signaling in the Brain of Learned Helpless Rats: Relevance in Vulnerability to Developing Stress-Induced Depression. Neural Plast. 2016, 2016, 7383724. [Google Scholar] [CrossRef] [Green Version]
  72. Duric, V.; Banasr, M.; Licznerski, P.; Schmidt, H.D.; Stockmeier, C.A.; Simen, A.A.; Newton, S.S.; Duman, R.S. A negative regulator of MAP kinase causes depressive behavior. Nat. Med. 2010, 16, 1328–1332. [Google Scholar] [CrossRef]
  73. Dwivedi, Y.; Rizavi, H.S.; Roberts, R.C.; Conley, R.C.; Tamminga, C.A.; Pandey, G.N. Reduced activation and expression of ERK1/2 MAP kinase in the post-mortem brain of depressed suicide subjects. J. Neurochem. 2001, 77, 916–928. [Google Scholar] [CrossRef]
  74. Zhu, J.J.; Qin, Y.; Zhao, M.; Van Aelst, L.; Malinow, R. Ras and Rap Control AMPA Receptor Trafficking during Synaptic Plasticity. Cell 2002, 110, 443–455. [Google Scholar] [CrossRef] [Green Version]
  75. Choi, B.H.; Philips, M.R.; Chen, Y.; Lu, L.; Dai, W. K-Ras Lys-42 is crucial for its signaling, cell migration, and invasion. J. Biol. Chem. 2018, 293, 17574–17581. [Google Scholar] [CrossRef] [Green Version]
  76. Choi, B.H.; Chen, C.; Philips, M.; Dai, W. RAS GTPases are modified by SUMOylation. Oncotarget 2017, 9, 4440–4450. [Google Scholar] [CrossRef] [Green Version]
  77. Qi, X.; Lin, W.; Li, J.; Li, H.; Wang, W.; Wang, D.; Sun, M. Fluoxetine increases the activity of the ERK-CREB signal system and alleviates the depressive-like behavior in rats exposed to chronic forced swim stress. Neurobiol. Dis. 2008, 31, 278–285. [Google Scholar] [CrossRef]
  78. Grønli, J.; Bramham, C.; Murison, R.; Kanhema, T.; Fiske, E.; Bjorvatn, B.; Ursin, R.; Portas, C.M. Chronic mild stress inhibits BDNF protein expression and CREB activation in the dentate gyrus but not in the hippocampus proper. Pharmacol. Biochem. Behav. 2006, 85, 842–849. [Google Scholar] [CrossRef]
  79. Chen, Y.C.; Hsu, W.L.; Ma, Y.L.; Tai, D.J.; Lee, E.H. CREB SUMOylation by the E3 ligase PIAS1 enhances spatial memory. J. Neurosci. 2014, 34, 9574–9589. [Google Scholar] [CrossRef]
  80. Mann, J.J. Role of the serotonergic system in the pathogenesis of major depression and suicidal behavior. Neuropsychopharmacology 1999, 21, 99s–105s. [Google Scholar] [CrossRef] [Green Version]
  81. Kowiański, P.; Lietzau, G.; Czuba, E.; Waśkow, M.; Steliga, A.; Moryś, J. BDNF: A Key Factor with Multipotent Impact on Brain Signaling and Synaptic Plasticity. Cell. Mol. Neurobiol. 2018, 38, 579–593. [Google Scholar] [CrossRef]
  82. Nibuya, M.; Morinobu, S.; Duman, R.S. Regulation of BDNF and trkB mRNA in rat brain by chronic electroconvulsive seizure and antidepressant drug treatments. J. Neurosci. 1995, 15, 7539–7547. [Google Scholar] [CrossRef]
  83. Smith, M.A.; Makino, S.; Kvetnansky, R.; Post, R.M. Stress and glucocorticoids affect the expression of brain-derived neurotrophic factor and neurotrophin-3 mRNAs in the hippocampus. J. Neurosci. 1995, 15, 1768–1777. [Google Scholar] [CrossRef] [Green Version]
  84. Thirouin, Z.S.; Figueiredo, M.; Hleihil, M.; Gill, R.; Bosshard, G.; McKinney, R.A.; Tyagarajan, S.K. Trophic factor BDNF inhibits GABAergic signaling by facilitating dendritic enrichment of SUMO E3 ligase PIAS3 and altering gephyrin scaffold. J. Biol. Chem. 2022, 298, 101840. [Google Scholar] [CrossRef]
  85. Karege, F.; Perroud, N.; Burkhardt, S.; Fernandez, R.; Ballmann, E.; La Harpe, R.; Malafosse, A. Alterations in Phosphatidylinositol 3-Kinase Activity and PTEN Phosphatase in the Prefrontal Cortex of Depressed Suicide Victims. Neuropsychobiology 2011, 63, 224–231. [Google Scholar] [CrossRef] [Green Version]
  86. Briones, A.; Gagno, S.; Martisova, E.; Dobarro, M.; Aisa, B.; Solas, M.; Tordera, R.; Ramírez, M.J. Stress-induced anhedonia is associated with an increase in Alzheimer’s disease-related markers. Br. J. Pharmacol. 2012, 165, 897–907. [Google Scholar] [CrossRef] [Green Version]
  87. Meng, L.; Du, C.-P.; Lu, C.-Y.; Zhang, K.; Li, L.; Yan, J.-Z.; Hou, X.-Y. Neuronal activity-induced SUMOylation of Akt1 by PIAS3 is required for long-term potentiation of synaptic transmission. FASEB J. 2021, 35, e21769. [Google Scholar] [CrossRef]
  88. Yoon, T.-Y.; Munson, M. SNARE complex assembly and disassembly. Curr. Biol. 2018, 28, R397–R401. [Google Scholar] [CrossRef] [Green Version]
  89. Madrigal, M.P.; Portalés, A.; SanJuan, M.P.; Jurado, S. Postsynaptic SNARE Proteins: Role in Synaptic Transmission and Plasticity. Neuroscience 2019, 420, 12–21. [Google Scholar] [CrossRef]
  90. Tang, L.T.H.; Craig, T.J.; Henley, J.M. SUMOylation of synapsin Ia maintains synaptic vesicle availability and is reduced in an autism mutation. Nat. Commun. 2015, 6, 7728. [Google Scholar] [CrossRef] [Green Version]
  91. Watanabe, S. Slow or fast? A tale of synaptic vesicle recycling. Science 2015, 350, 46–47. [Google Scholar] [CrossRef] [Green Version]
  92. Craig, T.J.; Anderson, D.; Evans, A.J.; Girach, F.; Henley, J.M. SUMOylation of Syntaxin1A regulates presynaptic endocytosis. Sci. Rep. 2015, 5, 17669. [Google Scholar] [CrossRef] [Green Version]
  93. Dulubova, I.; Lou, X.; Lu, J.; Huryeva, I.; Alam, A.; Schneggenburger, R.; Südhof, T.C.; Rizo, J. A Munc13/RIM/Rab3 tripartite complex: From priming to plasticity? EMBO J. 2005, 24, 2839–2850. [Google Scholar] [CrossRef]
  94. Girach, F.; Craig, T.J.; Rocca, D.L.; Henley, J.M. RIM1α SUMOylation Is Required for Fast Synaptic Vesicle Exocytosis. Cell Rep. 2013, 5, 1294–1301. [Google Scholar] [CrossRef] [Green Version]
  95. Voglis, G.; Tavernarakis, N. The role of synaptic ion channels in synaptic plasticity. EMBO Rep. 2006, 7, 1104–1110. [Google Scholar] [CrossRef] [Green Version]
  96. Chai, Z.; Ma, C.; Jin, X. Homeostatic activity regulation as a mechanism underlying the effect of brain stimulation. Bioelectron. Med. 2019, 5, 16. [Google Scholar] [CrossRef] [Green Version]
  97. Qi, Y.; Wang, J.; Bomben, V.C.; Li, D.-P.; Chen, S.-R.; Sun, H.; Xi, Y.; Reed, J.G.; Cheng, J.; Pan, H.-L.; et al. Hyper-SUMOylation of the Kv7 potassium channel diminishes the M-current leading to seizures and sudden death. Neuron 2014, 83, 1159–1171. [Google Scholar] [CrossRef] [Green Version]
  98. Mark, D.B.; Li, Q.-J.; Kieckhafer, K.; Dudek, D.; Matthew, R.W.; Sunahara Roger, K.; Iñiguez-Lluhí Jorge, A.; Martens Jeffrey, R. SUMO modification regulates inactivation of the voltage-gated potassium channel Kv1.5. Proc. Natl. Acad. Sci. USA 2007, 104, 1805–1810. [Google Scholar] [CrossRef] [Green Version]
  99. Welch, M.A.; Forster, L.A.; Atlas, S.I.; Baro, D.J. SUMOylating Two Distinct Sites on the A-type Potassium Channel, Kv4.2, Increases Surface Expression and Decreases Current Amplitude. Front. Mol. Neurosci. 2019, 12, 144. [Google Scholar]
  100. Plant, L.D.; Dowdell, E.J.; Dementieva, I.S.; Marks, J.D.; Goldstein, S.A.N. SUMO modification of cell surface Kv2.1 potassium channels regulates the activity of rat hippocampal neurons. J. Gen. Physiol. 2011, 137, 441–454. [Google Scholar] [CrossRef] [Green Version]
  101. Xiong, D.; Li, T.; Dai, H.; Arena Anthony, F.; Plant Leigh, D.; Goldstein Steve, A.N. SUMOylation determines the voltage required to activate cardiac IKs channels. Proc. Natl. Acad. Sci. USA 2017, 114, E6686–E6694. [Google Scholar] [CrossRef] [Green Version]
  102. Steffensen, A.B.; Andersen, M.N.; Mutsaers, N.; Mujezinovic, A.; Schmitt, N. SUMO co-expression modifies KV11.1 channel activity. Acta Physiol. 2018, 222, e12974. [Google Scholar] [CrossRef]
  103. Ju, W.; Li, Q.; Wilson, S.M.; Brittain, J.M.; Meroueh, L.; Khanna, R. SUMOylation alters CRMP2 regulation of calcium influx in sensory neurons. Channels 2013, 7, 153–159. [Google Scholar] [CrossRef] [Green Version]
  104. Gomez, K.; Ran, D.; Madura, C.L.; Moutal, A.; Khanna, R. Non-SUMOylated CRMP2 decreases NaV1.7 currents via the endocytic proteins Numb, Nedd4-2 and Eps15. Mol. Brain 2021, 14, 20. [Google Scholar] [CrossRef]
  105. Andrews, P.W.; Bharwani, A.; Lee, K.R.; Fox, M.; Thomson, J.A. Is serotonin an upper or a downer? The evolution of the serotonergic system and its role in depression and the antidepressant response. Neurosci. Biobehav. Rev. 2015, 51, 164–188. [Google Scholar] [CrossRef]
  106. Palazzo, E.; Marabese, I.; de Novellis, V.; Rossi, F.; Maione, S. Metabotropic Glutamate Receptor 7: From Synaptic Function to Therapeutic Implications. Curr. Neuropharmacol. 2016, 14, 504–513. [Google Scholar] [CrossRef] [Green Version]
  107. Gu, X.; Schafer, N.P.; Bueno, C.; Lu, W.; Wolynes, P.G. A structural dynamics model for how CPEB3 binding to SUMO2 can regulate translational control in dendritic spines. bioRxiv 2021. [Google Scholar] [CrossRef]
  108. Cartier, E.; Garcia-Olivares, J.; Janezic, E.; Viana, J.; Moore, M.; Lin, M.L.; Caplan, J.L.; Torres, G.; Kim, Y.-H. The SUMO-Conjugase Ubc9 Prevents the Degradation of the Dopamine Transporter, Enhancing Its Cell Surface Level and Dopamine Uptake. Front. Cell. Neurosci. 2019, 13, 35. [Google Scholar]
  109. Koulen, P.; Liu, J.; Nixon, E.; Madry, C. Interaction between mGluR8 and Calcium Channels in Photoreceptors Is Sensitive to Pertussis Toxin and Occurs Via G Protein βγ Subunit Signaling. Investig. Ophthalmol. Vis. Sci. 2005, 46, 287–291. [Google Scholar] [CrossRef]
  110. Wilkinson, K.A.; Henley, J.M. Analysis of metabotropic glutamate receptor 7 as a potential substrate for SUMOylation. Neurosci. Lett. 2011, 491, 181–186. [Google Scholar] [CrossRef] [Green Version]
  111. Dütting, E.; Schröder-Kress, N.; Sticht, H.; Enz, R. SUMO E3 ligases are expressed in the retina and regulate SUMOylation of the metabotropic glutamate receptor 8b. Biochem. J. 2011, 435, 365–371. [Google Scholar] [CrossRef] [Green Version]
  112. Choi, J.-H.; Park, J.-Y.; Park, S.P.; Lee, H.; Han, S.; Park, K.H.; Suh, Y.H. Regulation of mGluR7 trafficking by SUMOylation in neurons. Neuropharmacology 2016, 102, 229–235. [Google Scholar] [CrossRef] [PubMed]
  113. Loriol, C.; Cassé, F.; Khayachi, A.; Poupon, G.; Chafai, M.; Deval, E.; Gwizdek, C.; Martin, S. mGlu5 receptors regulate synaptic sumoylation via a transient PKC-dependent diffusional trapping of Ubc9 into spines. Nat. Commun. 2014, 5, 5113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Schorova, L.; Pronot, M.; Poupon, G.; Prieto, M.; Folci, A.; Khayachi, A.; Brau, F.; Cassé, F.; Gwizdek, C.; Martin, S. The synaptic balance between sumoylation and desumoylation is maintained by the activation of metabotropic mGlu5 receptors. Cell Mol. Life Sci. 2019, 76, 3019–3031. [Google Scholar] [CrossRef] [PubMed]
  115. Xu, J.; Tan, P.; Li, H.; Cui, Y.; Qiu, Y.; Wang, H.; Zhang, X.; Li, J.; Zhu, L.; Zhou, W.; et al. Direct SUMOylation of M1 muscarinic acetylcholine receptor increases its ligand-binding affinity and signal transduction. FASEB J. 2019, 33, 3237–3251. [Google Scholar] [CrossRef]
  116. Gowran, A.; Murphy, C.E.; Campbell, V.A. Δ9-Tetrahydrocannabinol regulates the p53 post-translational modifiers Murine double minute 2 and the Small Ubiquitin MOdifier protein in the rat brain. FEBS Lett. 2009, 583, 3412–3418. [Google Scholar] [CrossRef] [Green Version]
  117. Sihra, T.S.; Flores, G.; Rodríguez-Moreno, A. Kainate Receptors: Multiple Roles in Neuronal Plasticity. Neuroscientist 2013, 20, 29–43. [Google Scholar] [CrossRef]
  118. Fioriti, L.; Myers, C.; Huang, Y.-Y.; Li, X.; Stephan, J.S.; Trifilieff, P.; Colnaghi, L.; Kosmidis, S.; Drisaldi, B.; Pavlopoulos, E.; et al. The Persistence of Hippocampal-Based Memory Requires Protein Synthesis Mediated by the Prion-like Protein CPEB3. Neuron 2015, 86, 1433–1448. [Google Scholar] [CrossRef] [Green Version]
  119. Ford, L.; Asok, A.; Tripp, A.D.; Parro, C.; Fitzpatrick, M.; de Solis, C.A.; Shafiian, N.; Fioriti, L.; Soni, R.K.; Kandel, E.R. The low complexity motif of cytoplasmic polyadenylation element binding protein 3 (CPEB3) is critical for the trafficking of its targets in neurons. bioRxiv 2020. [Google Scholar] [CrossRef]
  120. Knott, A.B.; Perkins, G.; Schwarzenbacher, R.; Bossy-Wetzel, E. Mitochondrial fragmentation in neurodegeneration. Nat. Rev. Neurosci. 2008, 9, 505–518. [Google Scholar] [CrossRef] [Green Version]
  121. Gong, Y.; Chai, Y.; Ding, J.H.; Sun, X.L.; Hu, G. Chronic mild stress damages mitochondrial ultrastructure and function in mouse brain. Neurosci. Lett. 2011, 488, 76–80. [Google Scholar] [CrossRef] [PubMed]
  122. Baxter, L.R., Jr.; Schwartz, J.M.; Phelps, M.E.; Mazziotta, J.C.; Guze, B.H.; Selin, C.E.; Gerner, R.H.; Sumida, R.M. Reduction of prefrontal cortex glucose metabolism common to three types of depression. Arch. Gen. Psychiatry 1989, 46, 243–250. [Google Scholar] [CrossRef] [PubMed]
  123. Youle, R.J.; van der Bliek, A.M. Mitochondrial fission, fusion, and stress. Science 2012, 337, 1062–1065. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Tatsuta, T.; Langer, T. Quality control of mitochondria: Protection against neurodegeneration and ageing. EMBO J. 2008, 27, 306–314. [Google Scholar] [CrossRef] [Green Version]
  125. Lieberman, A. Depression in Parkinson’s disease–A review. Acta Neurol. Scand. 2006, 113, 1–8. [Google Scholar] [CrossRef]
  126. Detmer, S.A.; Chan, D.C. Functions and dysfunctions of mitochondrial dynamics. Nat. Rev. Mol. Cell Biol. 2007, 8, 870–879. [Google Scholar] [CrossRef]
  127. Cho, H.M.; Sun, W. The coordinated regulation of mitochondrial structure and function by Drp1 for mitochondrial quality surveillance. BMB Rep. 2019, 52, 109–110. [Google Scholar] [CrossRef] [Green Version]
  128. Figueroa-Romero, C.; Iñiguez-Lluhí, J.A.; Stadler, J.; Chang, C.-R.; Arnoult, D.; Keller, P.J.; Hong, Y.; Blackstone, C.; Feldman, E.L. SUMOylation of the mitochondrial fission protein Drp1 occurs at multiple nonconsensus sites within the B domain and is linked to its activity cycle. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2009, 23, 3917–3927. [Google Scholar] [CrossRef] [Green Version]
  129. Harder, Z.; Zunino, R.; McBride, H. Sumo1 conjugates mitochondrial substrates and participates in mitochondrial fission. Curr. Biol. 2004, 14, 340–345. [Google Scholar] [CrossRef]
  130. Guo, C.; Hildick, K.L.; Luo, J.; Dearden, L.; Wilkinson, K.A.; Henley, J.M. SENP3-mediated deSUMOylation of dynamin-related protein 1 promotes cell death following ischaemia. EMBO J. 2013, 32, 1514–1528. [Google Scholar] [CrossRef]
  131. Braschi, E.; Zunino, R.; McBride, H.M. MAPL is a new mitochondrial SUMO E3 ligase that regulates mitochondrial fission. EMBO Rep. 2009, 10, 748–754. [Google Scholar] [CrossRef] [PubMed]
  132. Fu, J.; Yu, H.M.I.; Chiu, S.-Y.; Mirando, A.J.; Maruyama, E.O.; Cheng, J.-G.; Hsu, W. Disruption of SUMO-Specific Protease 2 Induces Mitochondria Mediated Neurodegeneration. PLoS Genet. 2014, 10, e1004579. [Google Scholar] [CrossRef] [Green Version]
  133. Bansal, Y.; Kuhad, A. Mitochondrial Dysfunction in Depression. Curr. Neuropharmacol. 2016, 14, 610–618. [Google Scholar] [PubMed] [Green Version]
  134. Hayashi, T.; Ishimori, C.; Takahashi-Niki, K.; Taira, T.; Kim, Y.C.; Maita, H.; Maita, C.; Ariga, H.; Iguchi-Ariga, S.M. DJ-1 binds to mitochondrial complex I and maintains its activity. Biochem. Biophys. Res. Commun. 2009, 390, 667–672. [Google Scholar] [CrossRef] [Green Version]
  135. Junn, E.; Jang, W.H.; Zhao, X.; Jeong, B.S.; Mouradian, M.M. Mitochondrial localization of DJ-1 leads to enhanced neuroprotection. J. Neurosci. Res. 2009, 87, 123–129. [Google Scholar] [CrossRef] [Green Version]
  136. Shinbo, Y.; Niki, T.; Taira, T.; Ooe, H.; Takahashi-Niki, K.; Maita, C.; Seino, C.; Iguchi-Ariga, S.M.; Ariga, H. Proper SUMO-1 conjugation is essential to DJ-1 to exert its full activities. Cell Death Differ. 2006, 13, 96–108. [Google Scholar] [CrossRef]
  137. Waters, E.; Wilkinson, K.A.; Harding, A.L.; Carmichael, R.E.; Robinson, D.; Colley, H.E.; Guo, C. The SUMO protease SENP3 regulates mitochondrial autophagy mediated by Fis1. EMBO Rep. 2022, 23, e48754. [Google Scholar] [CrossRef]
  138. Zhang, Q.; Wu, J.; Wu, R.; Ma, J.; Du, G.; Jiao, R.; Tian, Y.; Zheng, Z.; Yuan, Z. DJ-1 promotes the proteasomal degradation of Fis1: Implications of DJ-1 in neuronal protection. Biochem. J. 2012, 447, 261–269. [Google Scholar] [CrossRef]
  139. Yoon, Y.; Krueger, E.W.; Oswald, B.J.; McNiven, M.A. The mitochondrial protein hFis1 regulates mitochondrial fission in mammalian cells through an interaction with the dynamin-like protein DLP1. Mol. Cell. Biol. 2003, 23, 5409–5420. [Google Scholar] [CrossRef] [Green Version]
  140. Villena, J.A. New insights into PGC-1 coactivators: Redefining their role in the regulation of mitochondrial function and beyond. FEBS J. 2015, 282, 647–672. [Google Scholar] [CrossRef]
  141. Rytinki, M.M.; Palvimo, J.J. SUMOylation attenuates the function of PGC-1alpha. J. Biol. Chem. 2009, 284, 26184–26193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Cai, R.; Yu, T.; Huang, C.; Xia, X.; Liu, X.; Gu, J.; Xue, S.; Yeh, E.T.H.; Cheng, J. SUMO-specific protease 1 regulates mitochondrial biogenesis through PGC-1α. J. Biol. Chem. 2012, 287, 44464–44470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Zhong, N.; Xu, J. Synergistic activation of the human MnSOD promoter by DJ-1 and PGC-1alpha: Regulation by SUMOylation and oxidation. Hum. Mol. Genet. 2008, 17, 3357–3367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Um, J.W.; Chung, K.C. Functional modulation of parkin through physical interaction with SUMO-1. J. Neurosci. Res. 2006, 84, 1543–1554. [Google Scholar] [CrossRef] [PubMed]
  145. Shin, J.H.; Ko, H.S.; Kang, H.; Lee, Y.; Lee, Y.I.; Pletinkova, O.; Troconso, J.C.; Dawson, V.L.; Dawson, T.M. PARIS (ZNF746) repression of PGC-1α contributes to neurodegeneration in Parkinson’s disease. Cell 2011, 144, 689–702. [Google Scholar] [CrossRef] [Green Version]
  146. Nishida, T.; Yamada, Y. SUMOylation of the KRAB zinc-finger transcription factor PARIS/ZNF746 regulates its transcriptional activity. Biochem. Biophys. Res. Commun. 2016, 473, 1261–1267. [Google Scholar] [CrossRef]
  147. Nishida, T.; Yamada, Y. RNF4-mediated SUMO-targeted ubiquitination relieves PARIS/ZNF746-mediated transcriptional repression. Biochem. Biophys. Res. Commun. 2020, 526, 110–116. [Google Scholar] [CrossRef]
  148. Kim, C.; Juncker, M.; Reed, R.; Haas, A.; Guidry, J.; Matunis, M.; Yang, W.C.; Schwartzenburg, J.; Desai, S. SUMOylation of mitofusins: A potential mechanism for perinuclear mitochondrial congression in cells treated with mitochondrial stressors. Biochim. Biophys. Acta Mol. Basis Dis. 2021, 1867, 166104. [Google Scholar] [CrossRef]
  149. Doxaki, C.; Palikaras, K. Neuronal Mitophagy: Friend or Foe? Front. Cell Dev. Biol. 2021, 8, 611938. [Google Scholar]
  150. Williams, L.S.; Jones, W.J.; Shen, J.; Robinson, R.L.; Weinberger, M.; Kroenke, K. Prevalence and impact of depression and pain in neurology outpatients. J. Neurol. Neurosurg. Psychiatry 2003, 74, 1587. [Google Scholar] [CrossRef] [Green Version]
  151. Cai, S.; Moutal, A.; Yu, J.; Chew Lindsey, A.; Isensee, J.; Chawla, R.; Gomez, K.; Luo, S.; Zhou, Y.; Chefdeville, A.; et al. Selective targeting of NaV1.7 via inhibition of the CRMP2-Ubc9 interaction reduces pain in rodents. Sci. Transl. Med. 2021, 13, eabh1314. [Google Scholar] [CrossRef] [PubMed]
  152. Shi, C.; Jin, J.; Xu, H.; Ma, J.; Li, T.; Xie, Y.; Li, Z. CCR1 enhances SUMOylation of DGCR8 by up-regulating ERK phosphorylation to promote spinal nerve ligation-induced neuropathic pain. Gene Therapy 2021, 29, 379–389. [Google Scholar] [CrossRef] [PubMed]
  153. Budziñski, M.L.; Sokn, C.; Gobbini, R.; Ugo, B.; Antunica-Noguerol, M.; Senin, S.; Bajaj, T.; Gassen, N.C.; Rein, T.; Schmidt, M.V.; et al. Tricyclic antidepressants target FKBP51 SUMOylation to restore glucocorticoid receptor activity. Mol. Psychiatry 2022, 27, 2533–2545. [Google Scholar] [CrossRef] [PubMed]
Figure 1. A diagram of the SUMOylation pathway. Small Ubiquitin-like MOdifier (SUMO) is first synthesized as an inactive precursor, which is subsequently cleaved by sentrin/SUMO-specific protease (SENPs) to reveal the double glycine motif and create mature SUMO proteins. SUMO is then activated by E1 proteins (SAE1 and SAE2 heterodimer). The activated SUMO protein is transferred to the Ubc9 E2 enzyme, which works in unison with E3 enzymes to target and ligate the SUMO protein onto substrates. The SUMO protein can be recycled by SENP-mediated cleavage of SUMO. SUMOylation either enhances or blocks protein-protein interactions, modifies substrate conformation, blocks sites for other post-translational modifications (i.e., ubiquitin), and aids in protein localization.
Figure 1. A diagram of the SUMOylation pathway. Small Ubiquitin-like MOdifier (SUMO) is first synthesized as an inactive precursor, which is subsequently cleaved by sentrin/SUMO-specific protease (SENPs) to reveal the double glycine motif and create mature SUMO proteins. SUMO is then activated by E1 proteins (SAE1 and SAE2 heterodimer). The activated SUMO protein is transferred to the Ubc9 E2 enzyme, which works in unison with E3 enzymes to target and ligate the SUMO protein onto substrates. The SUMO protein can be recycled by SENP-mediated cleavage of SUMO. SUMOylation either enhances or blocks protein-protein interactions, modifies substrate conformation, blocks sites for other post-translational modifications (i.e., ubiquitin), and aids in protein localization.
Ijms 23 08023 g001
Figure 2. SUMOylation in the neuron. SUMOylation plays an integral function in signal transduction. α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptors (AMPARs) are required for proper synaptic transmission. Although AMPARs are not directly SUMOylated, SUMOylation is required for proper AMPAR trafficking by Arc proteins. N-methyl-D-aspartate receptor (NMDAR) stimulation causes SUMOylation of neuronal nitric oxides synthase (nNOS), which allows signal relay from the nNOS-NO pathway to the extracellular signal-regulated kinase (ERK) pathway. Prevention of nNOS SUMOylation blocks ERK1/2 signaling. Ras SUMOylation is required for signal transduction through the Ras/Raf/MEK/ERK cascade. cAMP response element-binding protein (CREB) SUMOylation increases brain-derived neurotrophic factor (BDNF) levels. BDNF-TrkB signaling activates CREB binding through the ERK pathway. BDNF-TrkB regulates PIAS3 translocation from the nucleus to the dendrites. Downstream of BDNF-TrkB signaling is the P13K-Akt pathway. Akt is SUMOylated at two lysine residues, leading to an increase in Akt activity.
Figure 2. SUMOylation in the neuron. SUMOylation plays an integral function in signal transduction. α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptors (AMPARs) are required for proper synaptic transmission. Although AMPARs are not directly SUMOylated, SUMOylation is required for proper AMPAR trafficking by Arc proteins. N-methyl-D-aspartate receptor (NMDAR) stimulation causes SUMOylation of neuronal nitric oxides synthase (nNOS), which allows signal relay from the nNOS-NO pathway to the extracellular signal-regulated kinase (ERK) pathway. Prevention of nNOS SUMOylation blocks ERK1/2 signaling. Ras SUMOylation is required for signal transduction through the Ras/Raf/MEK/ERK cascade. cAMP response element-binding protein (CREB) SUMOylation increases brain-derived neurotrophic factor (BDNF) levels. BDNF-TrkB signaling activates CREB binding through the ERK pathway. BDNF-TrkB regulates PIAS3 translocation from the nucleus to the dendrites. Downstream of BDNF-TrkB signaling is the P13K-Akt pathway. Akt is SUMOylated at two lysine residues, leading to an increase in Akt activity.
Ijms 23 08023 g002
Figure 3. SUMOylation is involved in mitochondrial function. Drp1 is a fission regulatory protein that is found in the cytosol and recruited to the mitochondrial outer membrane by Fis1. Conjugation of SUMO1 to Drp1 increases fission, while SUMO2/3 conjugation represses fission activity. Cellular level of Drp1 is regulated by Parkin E3 ligase, whose activity is regulated by SUMO1 conjugation. Mitophagy is regulated by Fis1, but SUMOylation of Fis1 can repress mitophagy activity. Fis1 levels are regulated by DJ-1, which has antioxidant effects and is recruited into the mitochondria during oxidative stress. SUMOylation is necessary for DJ-1 antioxidant activity. DJ-1 has also been shown to regulate mitochondrial gene expression through indirect control of peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC-1α) through polypyrimidine tract-binding protein-associated splicing factor (PSF). DJ-1 inhibits PSF SUMOylation, which increases PSF/PGC-1α binding affinity and suppresses PGC-1α transcriptional activity. PGC-1α levels are regulated by parkin through parkin-interacting substrate (PARIS). PARIS, acting as a suppressor of PGC-1α expression, is regulated by ubiquitin-mediated proteasome via parkin E3 activity. The conjugation of SUMO2/3 to PARIS has been demonstrated to relieve PARIS-mediated transcriptional repression.
Figure 3. SUMOylation is involved in mitochondrial function. Drp1 is a fission regulatory protein that is found in the cytosol and recruited to the mitochondrial outer membrane by Fis1. Conjugation of SUMO1 to Drp1 increases fission, while SUMO2/3 conjugation represses fission activity. Cellular level of Drp1 is regulated by Parkin E3 ligase, whose activity is regulated by SUMO1 conjugation. Mitophagy is regulated by Fis1, but SUMOylation of Fis1 can repress mitophagy activity. Fis1 levels are regulated by DJ-1, which has antioxidant effects and is recruited into the mitochondria during oxidative stress. SUMOylation is necessary for DJ-1 antioxidant activity. DJ-1 has also been shown to regulate mitochondrial gene expression through indirect control of peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC-1α) through polypyrimidine tract-binding protein-associated splicing factor (PSF). DJ-1 inhibits PSF SUMOylation, which increases PSF/PGC-1α binding affinity and suppresses PGC-1α transcriptional activity. PGC-1α levels are regulated by parkin through parkin-interacting substrate (PARIS). PARIS, acting as a suppressor of PGC-1α expression, is regulated by ubiquitin-mediated proteasome via parkin E3 activity. The conjugation of SUMO2/3 to PARIS has been demonstrated to relieve PARIS-mediated transcriptional repression.
Ijms 23 08023 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jeoung, S.-W.; Park, H.-S.; Ryoo, Z.Y.; Cho, D.-H.; Lee, H.-S.; Ryu, H.-Y. SUMOylation and Major Depressive Disorder. Int. J. Mol. Sci. 2022, 23, 8023. https://doi.org/10.3390/ijms23148023

AMA Style

Jeoung S-W, Park H-S, Ryoo ZY, Cho D-H, Lee H-S, Ryu H-Y. SUMOylation and Major Depressive Disorder. International Journal of Molecular Sciences. 2022; 23(14):8023. https://doi.org/10.3390/ijms23148023

Chicago/Turabian Style

Jeoung, Seok-Won, Hyun-Sun Park, Zae Young Ryoo, Dong-Hyung Cho, Hyun-Shik Lee, and Hong-Yeoul Ryu. 2022. "SUMOylation and Major Depressive Disorder" International Journal of Molecular Sciences 23, no. 14: 8023. https://doi.org/10.3390/ijms23148023

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

Article Metrics

Back to TopTop