Small ubiquitin-like modifier (SUMO) conjugation (or SUMOylation) is a post-translational protein modification implicated in alterations to protein expression, localization and function. Despite a number of nuclear roles for SUMO being well characterized, this process has only started to be explored in relation to membrane proteins, such as ion channels. Calcium ion (Ca2+) signalling is crucial for the normal functioning of cells and is also involved in the pathophysiological mechanisms underlying relevant neurological and cardiovascular diseases. Intracellular Ca2+ levels are tightly regulated; at rest, most Ca2+ is retained in organelles, such as the sarcoplasmic reticulum, or in the extracellular space, whereas depolarization triggers a series of events leading to Ca2+ entry, followed by extrusion and reuptake. The mechanisms that maintain Ca2+ homoeostasis are candidates for modulation at the post-translational level. Here, we review the effects of protein SUMOylation, including Ca2+ channels, their proteome and other proteins associated with Ca2+ signalling, on vital cellular functions, such as neurotransmission within the central nervous system (CNS) and in additional systems, most prominently here, in the cardiac system.

The small ubiquitin-like modifier (SUMO) was first described as targeting nuclear proteins that regulate transcription factors, gene expression and DNA integrity [1]. Experiments with knockout mice for the sole SUMO conjugating enzyme, ubiquitin-like conjugating enzyme 9 (Ubc9), demonstrated nuclear dysfunction and embryonic lethality, confirming that SUMOylation is physiologically indispensable [2]. Reports that are more recent have shown that SUMO can also target cytosolic and membrane proteins, including ion channels, to regulate crucial cellular functions, such as plasma membrane depolarization and neurotransmission [3,4]. So far, the majority of studies have focused on SUMOylation of potassium (K+) channels, which are involved in setting the duration and firing pattern of action potentials [5]. For example, SUMOylation can modulate both two-pore domain K+ (K2P) channels [3,69], responsible for the regulation of background leak currents, and voltage-dependent K+ (KV) channels [1013] that repolarize cell membrane during action potential input. However, there is also recent evidence that voltage-gated Ca2+ channels (VGCCs) [14] and transient receptor potential (TRP) channels [15], both of which can mediate Ca2+ influx, are SUMO targets. Considering the utmost relevance of Ca2+ in physiological and pathophysiological processes, and the growing evidence that SUMO can modify ion channels, our review focused on the potential roles of SUMOylation of Ca2+ channels and proteins related with Ca2+ signalling with a focus on the central nervous system (CNS) and, also, the cardiac system.

Post-translational modifications of proteins can affect their function, localization and degradation depending on the stimulus applied, to control cellular response [16,17]. SUMOylation is a reversible lysine-targeted post-translational modification, whereby covalently conjugated SUMO regulates proteins in numerous pathways [18,19]. Currently, there are five proposed SUMO isoforms, with SUMO-1, 2 and 3 being the best-characterized paralogs. SUMO-1 shares approximately 50% of its amino acid sequence with both SUMO-2 and SUMO-3, which are typically known as SUMO-2/3 since they differ by only three N-terminal amino acids and antibodies are usually unable to distinguish between them [20,21]. Despite the similarities, there are functional differences between SUMO-1 and SUMO-2/3. For instance, under basal conditions, unconjugated SUMO-1 is scarce, but free SUMO-2/3 is widely expressed in mammalian cells [22]. Although the exact role for SUMO-4 remains uncertain, it has been associated with the pathophysiological mechanisms underlying diabetes [23,24]. Finally, the existence of a fifth SUMO isoform, SUMO-5, that regulates promyelocytic leukaemia nuclear bodies, has recently been suggested [25]. The same enzymes conjugate all SUMO isoforms [19].

The first step in the SUMOylation process requires the maturation of SUMO by SUMO-specific isopeptidases/proteases; next, SUMO is activated in an ATP-dependent step by E1 complex, which in humans consists of an heterodimer formed by SUMO-activating enzyme subunits 1 and 2 (SAE1 and SAE2 respectively). Subsequently, SUMO is transferred from the E1 activating enzyme to the E2 conjugating enzyme, also known as Ubc9, which is able to conjugate SUMO to target proteins both in E3 ligase-dependent and -independent manners. Most target proteins carry the same consensus motif that is directly recognized by Ubc9: the –K–x–D/E sequence, with representing a large hydrophobic residue (commonly isoleucine, leucine or valine), K is the modified lysine, x is any residue and D/E are acidic residues [22,26]. Nevertheless, non-covalent interactions between SUMO and target proteins can occur through SUMO interacting motifs (SIMs) [17,27]. These SIMs consist of a short stretch of branched hydrophobic residues, typically comprising isoleucine (I) or valine (V) residues organized as (V/I)–x–(V/I)–(V/I) or (V/I)–(V/I)–x–(V/I), flanked NH2– or COOH– terminally by serine residues and/or acidic residues [28]. Alternatively, SUMO E3 ligases can directly bind to target proteins [17]. The SUMOylation process is highly reversible by the same enzymes responsible for SUMO maturation and also SUMO deconjugation from substrate proteins [29].

Recently, three distinct families of SUMO-specific isopeptidases and proteases have been identified in mammals: the ubiquitin-like protease/sentrin-specific protease (Ulp/SENP), the deSUMOylating isopeptidase (Desi) and ubiquitin-specific peptidase-like protein 1 (USPL1) [30,31]. The SENPs are the best characterized and, so far, six SENP isoforms have been identified in humans: SENP1, 2, 3, 5, 6 and 7 [17]. SENP1 is highly expressed in the nucleus, in the nuclear pore and as discrete nuclear ‘dots’ [32], but can also be found in all neuronal processes and at synapses at lower levels [3335]. During the maturation phase, SENP1 cleaves pro-SUMO preferentially to generate SUMO-1 and SUMO-2/3 [36,37], while it deconjugates both SUMO isoforms [37,38]. SENP2 is similar to SENP1 with respect to its localization and characteristics regarding the maturation step, but differs from SENP1 regarding its highly selectivity for SUMO-2/3 deconjugation [3740]. SENP3 is found in the nucleus, but also in the mitochondria and participates in neuronal signalling [41]. The role of SENP3 in cleaving pro-SUMO has not been elucidated as yet, but it is suggested that SENP3 is somehow selective for removing SUMO-2/3 from target proteins [37,38]. As for SENP3, SENP5 has a nuclear localization [37,42] and is important for SUMO-2/3 maturation and deconjugation [37,38,43]. Finally, SENP6 and SENP7 are located throughout the nucleoplasm [17,44] and, although neither participates in the maturation step, they are both important for removal of SUMO-2/3 [17,44,45]. Regarding the Desi family, two isoforms have been identified so far: Desi-1 and Desi-2. Whereas Desi-1 is found both in the cytoplasm and the nucleus, where it promotes deconjugation of all SUMO isoforms, Desi-2 is exclusively cytoplasmatic and its properties remain undefined [30,31]. Lastly, USPL1 preferably promotes SUMO-2/3 deconjugation and is located in Cajal bodies [30,31].

Disruption of basal SUMOylation has been implicated in multiple neurological disorders, including neurodegenerative diseases, such as Alzheimer and Parkinson’s diseases (AD and PD respectively), spinocerebellar ataxias (SCAs), cerebral ischaemia and epilepsy [46]. More specifically, amyloid precursor protein (APP) and tau, which are key proteins in AD, have been identified as SUMO targets in HeLa and HEK293 cells [4749]. APP undergoes proteolytic cleavage by α- or β-secretases, and both are followed by further γ-secretase processing [50]. While α-secretases cleave APP to peptides that are proposed to participate in neuroprotection and neuroplasticity, characterizing the non-amyloidogenic pathway [51], cleavage by β-secretases leads to the amyloidogenic pathway, generating toxic amyloid β (Aβ) that accumulates and forms amyloid plaques [52]. A reduction in Aβ aggregates was found in HeLa cells when APP was SUMOylated by either SUMO-1 or SUMO-2 at lysines 587 and 595, which are located adjacently to the β-secretase site [48]. Moreover, poly-SUMOylation of APP by SUMO-3 has been reported to regulate APP cleavage and decrease Aβ production in HEK293 cells [53]. Conversely, SUMO-3, as well as SUMO-1, was found to increase γ-secretase levels [54], thus increasing Aβ production in a transgenic mice model for AD [55]. It is important to note that SUMO-3 effects on Aβ deposition might not be dependent on the ability of SUMO-3 to conjugate to target proteins [54]. Another AD hallmark is the hyperphosphorylation of tau [56] that decreases its affinity for microtubules, resulting in tau accumulation and formation of neurofibrillary tangles [57]. Tau can undergo SUMOylation at lysine 340 in HEK293 cells, which triggered its phosphorylation and inhibited its degradation by the ubiquitin–proteasome pathway, thus increasing tau aggregation [47].

As for mouse models of AD [55], increased levels of SUMO-1 were found in the plasma of patients with dementia [58]. Conversely, SUMO-1 conjugates were not altered in the post-mortem hippocampus of AD patients, whereas SUMO-2/3 high molecular weight conjugates were decreased [59]. These observations are in agreement with previous reports that found increased SUMO-1 and decreased SUMO-2 conjugation levels in the cortex and hippocampus respectively, of Tg2576 mice [60,61]. However, a recent study demonstrated absence of gross changes in global SUMOylation levels in the post-mortem cortex of AD patients [62].

α-Synuclein, parkin and DJ-1 are examples of SUMO targets relevant to PD [17,63,64]. Cytosolic inclusions known as Lewy bodies, comprised mostly by aggregated α-synuclein, contribute to the synaptic dysfunction and consequent dopaminergic neuronal death predominantly in the substantia nigra, a well-described characteristic of PD [6568]. Promisingly, SUMO-1 conjugation to α-synuclein reduced its aggregation and toxicity in a transgenic mice model for PD [69]. Interestingly, in an early communication, lysosomal SUMO-1 labelling was identified in human olfactory mucosa-neurospheres obtained from biopsies of patients with idiopathic PD [70]. A similar finding was observed in post-mortem tissue from patients with multiple system atrophy and progressive supranuclear palsy, diseases in which α-synuclein and tau seem to be involved [70,71]. In both familial and sporadic PD, parkin, which is an ubiquitin ligase, can be found together with α-synuclein in Lewy bodies, where SUMO-1 was shown to non-covalently and selectively interact with parkin, increasing its auto-ubiquitination and transportion to the nucleus [72]. Moreover, SUMOylation of DJ-1, a transcriptional regulator mutated in 1–2% of early-onset PD cases, maintained its cytoprotective function in response to oxidative stress [73,74], whereas incomplete SUMOylation of DJ-1 led to its proteasomal degradation [75]. In a similar way to SUMOylated α-synuclein, increased SUMO conjugation to ataxin-7 decreased its aggregation and cytotoxicity in SCAs [76].

Despite several reports from our group and others showing that SUMOylation can protect cells from metabolic stress caused by low levels of oxygen and glucose in different models of cerebral ischaemia and hypoxic conditions [7781], disease-modified SUMO targets remain largely unknown. However, one such target is the mitochondrial GTPase dynamin-related protein 1 (Drp1), which regulates mitochondrial fission [41,82]. Under stress conditions, Drp1-mediated mitochondrial fission can release cytochrome c and induce caspase cleavage followed by cell apoptosis [83]. In an in vitro model of ischaemia, oxygen and glucose deprivation led to SENP3 degradation and consequent increase in SUMO-2/3 conjugation to Drp1, thus preventing mitochondrial fission and cytochrome c release, as well as promoting cell survival [41]. Another ischaemia-modified SUMO target is the isoform 3 of the sodium (Na+)/Ca2+ exchanger (NCX), which controls ionic homoeostasis during cerebral ischaemia [84]. NCX3 f-loop lysine 590 is required for SUMOylation, and the absence of this residue increased NCX3 degradation, exacerbating ischaemic damage induced by permanent and transient middle cerebral artery occlusion (MCAO) [85]. Following preconditioning and transient MCAO, SUMO-1 basal expression led to increased NCX3 levels, whereas SUMO silencing decreased NCX3 levels, suggesting that NCX3 SUMOylation participates in the protective role that SUMO-1 plays during ischaemic preconditioning [85].

Evidence shows that SUMOylation may be involved in mechanisms implicated in the development and maintenance of epilepsy, since it was demonstrated that neuronal K+ channels could be SUMOylated, thus modulating neuronal excitability [3,610]. Moreover, SUMOylation of excitatory receptor subunits can modulate receptor trafficking and interfere with synaptic transmission [8690]. For example, SUMOylation of the GluK2 subunit of kainate receptors led to receptor internalization, which could be neuroprotective against excitotoxicity [33]. More recently, the major cause of premature death in epilepsy, known as sudden unexplained death in epilepsy, has been linked with the hyper-SUMOylation of the KV7 K+ channel, which functionally reduces the depolarizing M-current conducted by this channel [13].

Unique amongst other ions, Ca2+ can modulate both membrane potential and function as an important signalling entity. Several cellular processes, ranging from neurotransmitter/hormone release [91] and muscle contraction [92] to gene transcription [93,94], require an increase in the intracellular Ca2+ levels, which under basal conditions are maintained approximately 100 nM [95]. This temporary increase occurs by either release from intracellular Ca2+ stores or influx into the cell by agonist-operated channels, G-protein coupled receptors, store-operated channels and, predominantly, through VGCCs located at the plasma membrane [96].

VGCCs were initially classified based on their voltage-dependent activation (high or low voltage-activated channels) [97,98] and subsequently subdivided by pharmacological and biophysical function (high voltage-activated and low voltage-activated) [99] and then by CaVα1 subunits [100]. CaVα1 structure allows selectivity for Ca2+ over monovalent ions and contains a sensor motif that detects membrane depolarization leading to channel opening [96]. Based on their CaVα1 subunits, three families of VGCCs have been defined: CaV1 – present mainly in skeletal muscle, heart, neurons and endocrine cells, CaV2 – found mainly at presynaptic terminals in the CNS, but also in peripheral synapses, and CaV3 – localized mainly in the sinoatrial node, adrenal glomerulosa cells, neurons and sperm acrosome [100,101]. CaV1 subunits form L-type Ca2+ current; CaV2.1 forms P/Q-type, CaV2.2 N-type and CaV2.3 form R-type current, whereas CaV3 subunits form T-type current. In addition to the three CaVα1 family subunits (CaV1, CaV2 and CaV3), there are auxiliary β, α2, δ and also γ subunits that comprise the channel complex and have various functions including transporting channels from the endoplasmic reticulum to the plasma membrane, maintaining channel stability and contributing to physiological and pharmacological properties [100].

Pathological changes in Ca2+ homoeostasis and deregulation of Ca2+ channels are implicated in a range of neurological disorders, including epilepsy, cerebral ischaemia, pain, neurodegenerative, and psychiatric diseases [102104]. Ca2+ levels control neuronal hyperexcitability and mutations in VGCCs have been identified in familial CNS diseases (so-called ‘channelopathies’). For example, CaV2.1 and CaV3.2 channelopathies have been widely associated with forms of absence epilepsy and episodic ataxia [105]. Furthermore, acquired epilepsy and cerebral ischaemia can occur due to insults resulting from increased Ca2+ influx [105,106]. Moreover, exocytosis of synaptic vesicles mediated by VGCCs, whereby membrane depolarization triggered by action potentials causes transmitter release, may be targeted in pain pathways, in particular at central terminals of sensory nociceptive afferents. For example, both CaV2.2 and CaV3.2 channels are crucial for control of neurotransmitter release at the dorsal horn [107,108]. CaV2.2 is targeted therapeutically by ziconotide [109,110], a drug used to treat cancer-derived pain, and other drugs targeting CaV2.2 are in development [96]. CaV3.2 also acts to regulate afferent fibre excitability [111] and there is good evidence that these channels are up-regulated under chronic pain conditions [112–115].

Neurodegenerative diseases and psychiatric disorders have been related to Ca2+ handling often with respect to mitochondrial function, since rises in Ca2+ levels lead to mitochondrial stress and generation of reactive oxygen species [96]. In AD, deregulation of Ca2+ homoeostasis contributes to Aβ production and accumulated Aβ interferes with Ca2+ influx. Under physiological conditions, Ca2+ entry is reported to contribute to APP cleavage by α-secretase, while improper intracellular Ca2+ mobilization can affect APP processing and lead to increased Aβ levels, neuroinflammation and metabolic stress [115,116]. Aβ is proposed to modulate Ca2+ influx in various ways including: by direct effects of oligomeric Aβ on the CaVα1 subunit [117,118], inducing membrane-associated oxidative stress or contributing to excitotoxicity [116,119]. Moreover, mutations in CaV1.2 and CaVβ2 have been linked to both bipolar disorder and schizophrenia, while mutations in CaV1.3 have also been linked to bipolar disorder [96]. In addition, CaV1.3 contributed to neuronal loss in PD as a consequence of inherent voltage-dependent activation of the subunit, rather than their selectivity for Ca2+ [120]. Moreover, α-synuclein aggregation can modulate the influx of Ca2+, and, in turn, increases in Ca2+ concentration can promote α-synuclein aggregation [121,122].

SUMOylation of proteins involved in Ca2+ signalling affects the maintenance of neurotransmission from synapse formation (Figure 1A) to neurotransmitter release (Figure 1B) and synaptic plasticity. Mutations in the CACNA1A gene, which encodes the CaV2.1 subunit, are found in SCA type 6 (SCA6) and lead to impaired VGCC function [123]. In an early communication, SUMO-1 overexpression was reported to decrease wild-type CaV2.1 current density in HEK293 cells, whereas it had no effects on SCA6 CaV2.1 mutants [124]. Interestingly, either SUMO-1 overexpression or SENP1 silencing enhanced cAMP-dependent exocytosis and glucagon secretion from both mouse and human pancreatic α-cells via effects on CaV1 channels [14].

Potential roles played by SUMO on Ca2+ signalling in neurotransmission

Figure 1
Potential roles played by SUMO on Ca2+ signalling in neurotransmission

(A) Decreased calcium signalling leads to phosphorylation and SUMOylation of MEF2A, thus promoting synapse formation. As a result of VGCC activation, MEF2 is dephosphorylated and switches SUMOylation to acetylation inhibiting synaptic processes. (B) SUMOylated RIM1α facilitates the clustering of CaV2.1 Ca2+ channels and enhances Ca2+ influx necessary for vesicular release. When SUMO is conjugated to CRMP2, it inhibits Ca2+ entry through CaV2.2 channels, and increases surface expression of NaV1.7 channels. SUMOylation of syntaxin-1A, synaptotagmin-1 and synapsin la can regulate neurotransmission by participating in docking/priming of synaptic vesicles; CRMP2, collapsin response mediator protein 2; MEF2, myocyte enhancer factor 2.

Figure 1
Potential roles played by SUMO on Ca2+ signalling in neurotransmission

(A) Decreased calcium signalling leads to phosphorylation and SUMOylation of MEF2A, thus promoting synapse formation. As a result of VGCC activation, MEF2 is dephosphorylated and switches SUMOylation to acetylation inhibiting synaptic processes. (B) SUMOylated RIM1α facilitates the clustering of CaV2.1 Ca2+ channels and enhances Ca2+ influx necessary for vesicular release. When SUMO is conjugated to CRMP2, it inhibits Ca2+ entry through CaV2.2 channels, and increases surface expression of NaV1.7 channels. SUMOylation of syntaxin-1A, synaptotagmin-1 and synapsin la can regulate neurotransmission by participating in docking/priming of synaptic vesicles; CRMP2, collapsin response mediator protein 2; MEF2, myocyte enhancer factor 2.

Close modal

Increased SUMO-1 conjugation to presynaptic target proteins was shown to regulate Ca2+ influx and neurotransmitter release in synaptosomes [125]. Depending on the applied stimulus, SUMOylation of presynaptic proteins could either increase or decrease neurotransmitter release. For example, loading synaptosomes with SUMO-1 and SENP1 peptides decreased and increased Ca2+ influx and KCl-evoked glutamate release respectively. Conversely, kainate-induced Ca2+ influx and neurotransmitter release were increased in synaptosomes loaded with SUMO-1 and decreased in synaptosomes loaded with SENP1 [125]. These results suggest that SUMO may be conjugated to distinct presynaptic proteins and act in an activity-dependent and stimulus-specific manner to modulate presynaptic release.

Crucial proteins in neurotransmitter release, CRMP2 and Rab3a-interacting molecule (RIM) have been identified as members of the CaV2 proteome [126]. SUMOylation of VGCC interacting proteins has been reported to play an important role in neurotransmission within pain pathways. CRMP2 interacts with CaV2.2 subunits in sensory neurons or nociceptors to modulate neurotransmitter release [127]. SUMO-1–3 modified CRMP2 at lysine 374 in cultured cathecholamine A differentiated cells [128]. Overexpression of SUMO, Ubc9 and CRMP2 in adult dorsal root ganglion neurons decreased, whereas overexpression of non-SUMOylatable CRMP2 increased, KCl depolarization-induced Ca2+ entry. In addition, CRMP2 SUMOylation increased surface expression of NaV1.7 channels [129]. Mutations in NaV1.7 channels, which are highly expressed in peripheral sensory neurons, where they are responsible for regulating neuronal excitability, are directly related with pain disorders [130].

RIM1α interacts either directly or indirectly with most presynaptic active zone proteins and participates in the docking and priming of synaptic vesicles [131] by modulating Ca2+ influx through regulation of VGCCs clustering [132,133]. SUMO-1 conjugation to RIM1α at lysine 502 was shown to be crucial for normal presynaptic exocytosis in neurons [133]. Knockdown of endogenous RIM1α, and its replacement with a non-SUMOylatable mutant, led to impairment of Ca2+-induced depolarization and consequent removal of the fast component of vesicle exocytosis. SUMOylated RIM1α facilitated the clustering of CaV2.1 channels and enhanced Ca2+ influx necessary for vesicular release, whereas de-SUMOylated RIM1α participated in the docking/priming of synaptic vesicles and structural maintenance of the active zone [133].

Presynaptic soluble N-ethylmaleimide sensitive factor attachment protein receptors (SNARE) proteins, such as syntaxin 1, are fundamental for neurotransmitter release [134] and might also participate in vesicle endocytosis [135,136]. Syntaxin 1A can be modified by SUMO-1 at any of three lysine residues (K252, K253 or K256) near the C-terminal transmembrane domain [137]. Preventing syntaxin 1A SUMOylation reduced its interaction with other SNARE proteins and disrupted the balance of synaptic vesicle endo/exocytosis, resulting in increased endocytosis. Another key protein that is SUMOylated is synapsin Ia: preventing SUMO-1 conjugation to synapsin Ia at lysine 687 caused impaired exocytosis due to a reduction in the number of releasable synaptic vesicles [138]. Proteomic analysis from a neuron-specific SUMO-1 overexpressing transgenic mouse model led to the identification of a number of previously unrecognized SUMO-1 targets in vivo, including the Ca2+ sensor synaptotagmin-1 [139]. Increased SUMO-1 conjugation to synaptotagmin-1 resulted in impaired performed paired pulse facilitation (PPF), which involves the facilitation of neurotransmitter release caused by residual Ca2+ from a previous stimulus.

Homologs of the SUMOylation machinery were identified in Drosophila, and an interaction with Ca2+/calmodulin-dependent protein kinase II (CaMKII) that modulates synaptic plasticity by regulating glutamatergic synapses [140] was demonstrated by yeast two-hybrid screening [141]. Drosophila SUMO-1 (DmSUMO-1) modification has potential to change the subcellular localization of CaMKII, but the functional consequences for this interaction remain to be confirmed.

Dendritic claws in cerebellar granular neurons, in which mossy fibre terminals and Golgi neurons form synapses [142], are regulated by the myocyte enhancer factor 2A (MEF2A). MEF2A transcription factor activity is regulated by several post-translational protein modifications, including phosphorylation [143145], ubiquitination [146] and SUMOylation [147]. Lack of Ca2+ signalling led to phosphorylation of MEF2A at serine 408, which in turn led to SUMO-1 conjugation at lysine 403 and inactivation of MEF2A, promoting dendritic claw differentiation, synapse formation and maturation. Activity-dependent Ca2+ signalling via CaV1 VGCCs induced calcineurin-mediated dephosphorylation of MEF2A at serine 408, promoting a switch from SUMOylation to acetylation at lysine 403, which in turn activated MEF2A and inhibited dendritic claw differentiation and synapse formation [147].

As previously described, deregulation of Ca2+ homoeostasis contributes to aggregation of proteins such as Aβ and α-synuclein, known as aggregation-prone proteins, which can interfere with neurotransmission. Also, production and accumulation of these proteins interfere with Ca2+ influx [148]. Two lysines of APP can be modified by SUMO in vivo leading to decreased levels of Aβ aggregates [48]. SUMOylation of α-synuclein seems to inhibit α-synuclein aggregation and toxicity both in vitro and in vivo [149]. This inhibition depends on the SUMO isoform (SUMO-1 conjugation is better than SUMO-3) and on the SUMOylated lysine (K102 is better than K96) [150]. Interestingly, raised concentrations of monomeric α-synuclein in the extracellular medium promoted dopamine release in the striatum via CaV2.2 channels in vivo and in vitro, modifying plasma membrane structure and altering raft partitioning of this channel, suggesting the early reorganization of synaptic terminals as the mechanism to sensitizing dopaminergic neurons [151]. Paradoxically, SUMOylation of α-synuclein promoted its aggregation in COS-7 cells and had an intriguing protective effect [152].

Other than the brain, SUMOylation is well characterized in the heart. Both Ubc9 inhibition and SUMO-2 knockout caused early embryonic lethality in mice [2,153], whereas SUMO-1 knockout led to specific cardiac septal defects [154]. Activating the SUMOylation pathway can also evoke cardiac abnormalities, such as cardiac specific SUMO-2 overexpression that induced premature death and severe cardiomyopathy [155]. Conversely, SUMO-1 overexpression improved heart failure [154156], suggesting that tightly regulated SUMOylation levels are essential for normal cardiac development [154,157].

SUMOylation also influences cardiac metabolism, controlling crucial proteins for the maintenance of cardiac energy homoeostasis and mitochondrial biogenesis, such as peroxisome proliferator-activated receptor (PPAR) and its associated co-regulators [158]. Similarly, under metabolic stress conditions, increased cellular SUMOylation (mainly by SUMO-2/3) can protect the brain during ischaemia or hibernation torpor [158160]. Both in animal models and human patients, a fine balance between SUMO conjugation/deconjugation is critical for cardiac stress adaptation [155,156,161,162].

SUMOylation is not only essential for cardiac development, predominantly by regulating transcription factors, but also implicated in the onset of cardiac diseases [163165]. Several K+ channels found in the heart can be modulated by SUMO, such as KV2.1 [11,12], a channel that helps set the cell resting potential [166]; KV1.5 [10], which controls excitability of atrial cells [167]; and K2P1 [3,69], which helps set resting membrane potential. SUMOylation also regulates the cardiac non-selective cationic channel TRPM4, which is localized predominantly in human atrial myocardium, and can act as a Ca2+ regulator [15,168]. Progressive familial heart block type I, an autosomal dominant disease, has been linked to a mutation in the TRPM4 amino-terminal region that leads to increased TRPM4 SUMOylation and prevention of its ubiquitination and consequent proteasomal degradation [15]. Other proteins crucial for the maintenance of cardiomyocyte physiology, such as lamin A that plays a structural and functional role in the nucleus, are also reported to be SUMOylated [169,170]. Familial cardiomyopathy has been linked with mutations in the human laminin A gene, which were in turn associated with decreases in laminin A SUMOylation and accelerated cell death [169].

Disrupting Ca2+ dynamics by interfering with other proteins or transcriptional factors that maintain Ca2+ homoeostasis, such as some of TRP protein Ca2+ entry channels or N-terminal serine residues of the nuclear factor of activated T cells (NFAT), can contribute to the onset of cardiac dysfunctions [171]. Increased intracellular Ca2+ levels activate calcineurin, a Ca2+-calmodulin dependent serine–threonine protein phosphatase that dephosphorylates NFATs, leading to nuclear translocation of NFATs and activation of pro-hypertrophic genes [172]. SUMO-2 can activate calcineurin-NFAT signalling in cardiomyocytes leading to a hyperthophic phenotype, both in vitro and in vivo [173]. Unexpectedly, a conjugation-deficient SUMO-2 mutant (SUMO-2ΔGG) was equally capable to activate the pathway and promote hypertrophic effects, suggesting a SUMOylation-independent mechanism.

Proteins such as sarcoendoplasmic reticulum calcium ATPase (SERCA) in the sarcoplasmic reticulum and NCX in the cardiomyocyte membrane help to restore Ca2+ concentrations at baseline following contraction [174]. The reduced expression or activity of SERCA2a is a hallmark of heart failure [175]. A proteomic screen has identified SERCA2a as a target for SUMO-1 (but not SUMO-2/3) at lysines 480 and 585 [156]. SUMO-1 and SERCA2a protein levels were decreased in animal models of heart failure, as well as in human cardiomyocytes isolated from failing ventricles. SUMO-1 overexpression restored SERCA2a levels, whereas either SUMO-1 or SERCA2a overexpression improved Ca2+ handling, improving cardiac function. However, increased global SUMOylation in SERCA2a knockdown cardiomyocytes did not prevent contractile dysfunction, further confirming that SUMOylated SERCA2a is essential for cardiac function [156]. The small molecule N106 (N-(4-methoxybenzo[d]thiazol-2-yl)-5-(4-methoxyphenyl)-1,3,4-oxadiazol-2-amine) was identified using an α-screen assay that detects SUMO-1 conjugation to nuclear RanGAP1 (the first and one of the most stable SUMO targets identified so far [176]). N106 promoted SERCA2a SUMOylation, resulting in enhanced contractility both in cultured cardiomyocytes and in vivo, significantly improving ventricular function in mice with heart failure [177]. N106 was proposed to directly activate the SUMO-activating enzyme [177].

Both alterations in Ca2+ homoeostasis and protein SUMOylation may lead to severe neurological, and also, cardiac pathologies. For example, SUMOylation of proteins involved in Ca2+ signalling can modulate synapse formation and alter neurotransmitter release. Furthermore, SUMOylation of proteins can modulate Ca2+ reuptake in cardiomyocytes and thus affect contractility. As described above and summarized in Table 1, it is clear that a wide range of proteins involved in these key physiological processes are subject to, potentially temporal, post-translational modification by different SUMO isoforms. Thus, at the presynapse, proteins involved in Ca2+ homoeostasis, including VGCCs and their proteome, are emerging as SUMO targets; equally, synaptic proteins involved in exocytosis and endocytosis are known to be SUMOylated. Postsynaptic receptor SUMOylation can also impact synaptic function. There is clear potential to exploit this knowledge to improve synaptic function in neurodegenerative and hyperexcitability disorders and to improve cardiac function. Thus, understanding how SUMOylation affects Ca2+ signalling in physiological and pathophysiological conditions is key to novel therapeutic strategies to prevent and/or cure important human diseases.

Table 1
Potential functional consequences of SUMOylation in Ca2+ signalling
Target (direct or indirect)SUMO isoformModified lysineMechanism or Ca2+ channel typeProposed SUMOylation effectReference
CaV2.1 subunit (indirect) SUMO-1 Unknown Inhibition of P/Q-type Ca2+ channels Role in SCA6 pathogenesis [124
CAMKII (indirect) SUMO-1 Unknown – Differentiation of Drosophila’s nervous system [141
CRMP2 (direct) SUMO-1 SUMO-2/3 K374 Inhibition of N-type Ca2+ channels Reduces Ca2+ influx in sensory neurons [128
MEF2 (direct) SUMO-1 K403 – Promotes dendritic claw differentiation [145,147
NCX3 (direct) SUMO-1 K590 – Inhibits NCX3 degradation [85
NFAT (indirect) SUMO-2 Unknown – Activates pro-hypertrophic genes [173
RIM1α (direct) SUMO-1 K502 Increase in P/Q-type Ca2+ channel activity Promotes synaptic vesicles release [133
SERCA2a (direct) SUMO-1 K480 and K585 – Increases Ca2+reuptake to sarcoendoplasmic reticulum [156,177
Synapsin Ia (direct) SUMO-1 K687 – Sets up releasable synaptic vesicles [138
Synaptotagmin-1 (indirect) SUMO-1 Unknown – Impairs neurotransmitter release [139
Syntaxin 1A (direct) SUMO-1 K252, K253 or K256 – Increases vesicular endocytosis [137
Target (direct or indirect)SUMO isoformModified lysineMechanism or Ca2+ channel typeProposed SUMOylation effectReference
CaV2.1 subunit (indirect) SUMO-1 Unknown Inhibition of P/Q-type Ca2+ channels Role in SCA6 pathogenesis [124
CAMKII (indirect) SUMO-1 Unknown – Differentiation of Drosophila’s nervous system [141
CRMP2 (direct) SUMO-1 SUMO-2/3 K374 Inhibition of N-type Ca2+ channels Reduces Ca2+ influx in sensory neurons [128
MEF2 (direct) SUMO-1 K403 – Promotes dendritic claw differentiation [145,147
NCX3 (direct) SUMO-1 K590 – Inhibits NCX3 degradation [85
NFAT (indirect) SUMO-2 Unknown – Activates pro-hypertrophic genes [173
RIM1α (direct) SUMO-1 K502 Increase in P/Q-type Ca2+ channel activity Promotes synaptic vesicles release [133
SERCA2a (direct) SUMO-1 K480 and K585 – Increases Ca2+reuptake to sarcoendoplasmic reticulum [156,177
Synapsin Ia (direct) SUMO-1 K687 – Sets up releasable synaptic vesicles [138
Synaptotagmin-1 (indirect) SUMO-1 Unknown – Impairs neurotransmitter release [139
Syntaxin 1A (direct) SUMO-1 K252, K253 or K256 – Increases vesicular endocytosis [137

Abbreviations: CAMKII, Ca2+/calmodulin-dependent protein kinase II; CRMP2, collapsin response mediator protein 2; MEF2, myocyte enhancer factor 2; NCX3, isoform 3 of the Na+/Ca2+ exchanger; NFAT, N-terminal serine residues of the nuclear factor of activated T-cells; RIM1α, Rab3a-interacting molecule 1α; SERCA2a, isoform 2a of sarcoendoplasmic reticulum Ca2+ ATPase.

Royal Society Newton Advanced and CNPq Fellowships together with IBRO and ISN/CAEN Return Home Awards to H.C. supported this work. L.C.S. is recipient of a CAPES MSc Studentship.

The authors declare that there are no competing interests associated with the manuscript.

amyloid β

AD

Alzheimer's disease

APP

amyloid precursor protein

CNS

central nervous system

CRMP2

collapsin response mediator protein 2

Desi

deSUMOylating isopeptidase

DJ-1

PD (autosomal recessive, early onset) 7

DmSUMO-1

Drosophila SUMO-1

Drp1

dynamin-related protein 1

MEF2A

myocyte enhancer factor 2A

NCX

sodium/calcium exchanger

NFAT

N-terminal serine residues of the nuclear factor of activated T cells

PD

Parkinson’s disease

PPAR

peroxisome proliferator-activated receptor

PPF

paired pulse facilitation

RIM

Rab3a-interacting molecule

SCA

spinocerebellar ataxia

SERCA

sarcoendoplasmic reticulum calcium ATPase

SIM

SUMO interacting motif

SNARE

soluble N-ethylmaleimide sensitive factor attachment protein receptors

SUMO

small ubiquitin-like modifier

TRP

transient receptor potential

TRPM4

transient receptor potential cation channel subfamily M member 4

Ubc9

ubiquitin-like conjugating enzyme 9

USPL1

ubiquitin-specific peptidase-like protein 1

VGCC

voltage-gated calcium channel

1
Hay
R.T.
(
2005
)
SUMO: a history of modification
.
Mol. Cell.
18
,
1
12
2
Nacerddine
K.
,
Lehembre
F.
,
Bhaumik
M.
,
Artus
J.
,
Cohen-Tannoudji
M.
,
Babinet
C.
et al
(
2005
)
The SUMO pathway is essential for nuclear integrity and chromosome segregation in mice
.
Dev. Cell
9
,
769
779
3
Plant
L.D.
,
Dementieva
I.S.
,
Kollewe
A.
,
Olikara
S.
,
Marks
J.D.
and
Goldstein
S.A.
(
2010
)
One SUMO is sufficient to silence the dimeric potassium channel K2P1
.
Proc. Natl Acad. Sci. U.S.A.
107
,
10743
10748
4
Silveirinha
V.
,
Stephens
G.J.
and
Cimarosti
H.
(
2013
)
Molecular targets underlying SUMO-mediated neuroprotection in brain ischemia
.
J. Neurochem.
127
,
580
591
5
Doyle
D.A.
,
Morais-Cabral
J.
,
Pfuetzner
R.A.
,
Kuo
A.
,
Gulbis
J.M.
,
Cohen
S.L.
et al
(
1998
)
The structure of the potassium channel: molecular basis of k1 conduction and selectivity
.
Science
280
,
69
77
6
Rajan
S.
,
Plant
L.D.
,
Rabin
M.L.
,
Butler
M. H.
and
Goldstein
S.A.
(
2005
)
Sumoylation silences the plasma membrane leak K+ channel K2P1
.
Cell
121
,
37
47
7
Feliciangeli
S.
,
Bendahhou
S.
,
Sandoz
G.
,
Gounon
P.
,
Reichold
M.
,
Warth
R.
et al
(
2007
)
Does sumoylation control K2P1/TWIK1 background K+ channels?
Cell
130
,
563
569
8
Es-Salah-Lamoureux
Z.
,
Steele
D.F.
and
Fedida
D.
(
2010
)
Research into the therapeutic roles of two-pore-domain potassium channels
.
Trends Pharmacol. Sci.
31
,
587
595
9
Plant
L.D.
,
Zuniga
L.
,
Araki
D.
,
Marks
J.D.
and
Goldstein
S.A.
(
2012
)
SUMOylation silences heterodimeric TASK potassium channels containing K2P1 subunits in cerebellar granule neurons
.
Sci. Signal.
5
,
ra84
10
Benson
M.D.
,
Li
Q.J.
,
Kieckhafer
K.
,
Dudek
D.
,
Whorton
M.R.
,
Sunahara
R.K.
et al
(
2007
)
SUMO modification regulates inactivation of the voltage-gated potassium channel KV1.5
.
Proc. Natl Acad. Sci. U.S.A.
104
,
1805
1810
11
Dai
X.Q.
,
Kolic
J.
,
Marchi
P.
,
Sipione
S.
and
Macdonald
P.E.
(
2009
)
SUMOylation regulates KV2.1 and modulates pancreatic beta-cell excitability
.
J. Cell Sci.
122
,
775
779
12
Plant
L.D.
,
Dowdell
E.J.
,
Dementieva
I.S.
,
Marks
J.D.
and
Goldstein
S.A.
(
2011
)
SUMO modification of cell surface KV2.1 potassium channels regulates the activity of rat hippocampal neurons
.
J. Gen. Physiol.
137
,
441
454
13
Qi
Y.
,
Wang
J.
,
Bomben
V.C.
,
Li
D.P.
,
Chen
S.-R.
,
Sun
H.
et al
(
2014
)
Hyper-SUMOylation of the KV7 potassium channel diminishes the M-current leading to seizures and sudden death
.
Neuron
83
,
1159
1171
14
Dai
X.Q.
,
Spigelman
A.F.
,
Khan
S.
,
Braun
M.
,
Manning-Fox
J.E.
and
Macdonald
P.E.
(
2014
)
SUMO1 enhances cAMP-dependent exocytosis and glucagon secretion from pancreatic α-cells
.
J. Physiol.
592
,
3715
3726
15
Kruse
M.
,
Schulze-Bahr
E.
,
Corfield
V.
,
Beckmann
A.
,
Stallmeyer
B.
,
Kurtbay
G.
et al
(
2009
)
Impaired endocytosis of the ion channel TRPM4 is associated with human progressive familial heart block type I
.
J. Clin. Invest.
119
,
2737
274
16
Walsh
C.T.
,
Garneau-Tsodikova
S.
and
Gatto
G.J.
Jr
(
2005
)
Protein posttranslational modifications: the chemistry of proteome diversifications
.
Angew. Chem. Int. Ed. Engl.
45
,
7342
7372
17
Henley
J.M.
,
Craig
T.J.
and
Wilkinson
K.A.
(
2014
)
Neuronal SUMOylation: mechanisms, physiology, and roles in neuronal dysfunction
.
Physiol. Rev.
94
,
1249
1258
18
Hendriks
I.A.
,
D’Souza
R.C.J.
,
Yang
B.
,
Verlaan-de Vries
M.
,
Mann
M.
and
Vertegaal
A.C.O.
(
2014
)
Uncovering Global SUMOylation Signaling Networks in a Site-Specific Manner
.
Nat. Struct. Mol. Biol.
10
,
927
936
19
Hendriks
I.A.
and
Vertegaal
A.C.O.
(
2016
)
A comprehensive compilation of SUMO proteomics
.
Nat. Rev. Mol. Cell Biol.
17
,
581
595
20
Johnson
E.S.
(
2004
)
Protein modification by SUMO
.
Annu. Rev. Biochem.
73
,
355
382
21
Wang
Y.
and
Dasso
M.
(
2009
)
SUMOylation and deSUMOylation at a glance
.
J. Cell Sci.
122
,
4249
4252
22
Sampson
D.A.
,
Wang
M.
and
Matunis
M.J.
(
2001
)
The small ubiquitin-like modifier-1 (SUMO-1) consensus sequence mediates Ubc9 binding and is essential for SUMO-1 modification
.
J. Biol. Chem.
276
,
21664
21669
23
Sozen
S.
,
Horozoglu
C.
,
Bireller
E.S.
,
Karaali
Z.
and
Cakmakoglu
B.
(
2014
)
Association of SUMO4 M55V and -94ins/del gene variants with type-2 diabetes
.
In Vivo
28
,
919
923
24
Sinha
N.
,
Yadav
A.K.
,
Kumar
V.
,
Dutta
P.
,
Bhansali
A.
and
Jha
V.
(
2016
)
SUMO4 163 G>A variation is associated with kidney disease in Indian subjects with type 2 diabetes
.
Mol. Biol. Rep.
43
,
345
348
25
Liang
Y.C.
,
Lee
C.C.
,
Yao
Y.L.
,
Lai
C.C.
,
Schmitz
M.L.
and
Yang
W.M.
(
2016
)
SUMO5, a novel poly-SUMO isoform, regulates PML nuclear bodies
.
Sci. Rep.
6
,
26509
26
Rodriguez
M.S.
,
Dargemont
C.
and
Hay
R.T.
(
2001
)
SUMO-1 conjugation in vivo requires both a consensus modification motif and nuclear targeting
.
J. Biol. Chem.
276
,
12654
12659
27
Jardin
C.
,
Anselm
H.C.
and
Sticht
H.
(
2015
)
Binding properties of SUMO-interacting motifs (SIMs) in yeast
.
J. Mol. Model.
21
,
50
28
Flotho
A.
and
Melchior
F.
(
2013
)
Sumoylation: a regulatory protein modification in health and disease
.
Annu. Rev. Biochem.
82
,
357
385
29
Gareau
J.R.
and
Lima
C.D.
(
2010
)
The SUMO pathway: emerging mechanisms that shape specificity, conjugation and recognition
.
Nat. Rev. Mol. Cell Biol.
11
,
861
871
30
Nayak
A.
and
Müller
S.
(
2014
)
SUMO-specific proteases/isopeptidases: SENPs and beyond
.
Genome Biol.
15
,
422
31
Hickey
C.M.
,
Wilson
N.R.
and
Hochstrasser
M.
(
2012
)
Function and regulation of SUMO proteases
.
Nat. Rev. Mol. Cell Biol.
13
,
755
766
32
Gong
L.
,
Millas
S.
,
Maul
G.G.
and
Yeh
E.T.
(
2000
)
Differential regulation of sentrinized proteins by a novel sentrin-specific protease
.
J. Biol. Chem.
275
,
3355
3359
33
Martin
S.
,
Nishimune
A.
,
Mellor
J.R.
and
Henley
J.M.
(
2007
)
SUMOylation regulates kainate-receptor-mediated synaptic transmission
.
Nature
447
,
321
325
34
Loriol
C.
,
Parisot
J.
,
Poupon
G.
,
Gwizdek
C.
and
Martin
S.
(
2012
)
Developmental regulation andspatiotemporal redistribution of the sumoylation machinery in the rat central nervous system
.
PLoS One
7
,
e33757
35
Loriol
C.
,
Khayachi
A.
,
Poupon
G.
,
Gwizdek
C.
and
Martin
S.
(
2013
)
Activity-dependent regulation of the sumoylation machinery in rat hippocampal neurons
.
Biol. Cell
105
,
30
45
36
Xu
Z.
and
Au
S.W.
(
2005
)
Mapping residues of SUMO precursors essential in differential maturation by SUMO-specific protease, SENP1
.
Biochem. J.
386
,
325
330
37
Gong
L.
and
Yeh
E.T.
(
2006
)
Characterization of a family of nucleolar SUMO-specific proteases with preference for SUMO-2 or SUMO-3
.
J. Biol. Chem.
281
,
15869
15877
38
Kolli
N.
,
Mikolajczyk
J.
,
Drag
M.
,
Mukhopadhyay
D.
,
Moffatt
N.
,
Dasso
M.
et al
(
2010
)
Distribution and paralogue specificity of mammalian deSUMOylating enzymes
.
Biochem. J.
430
,
335
344
39
Hang
J.
and
Dasso
M.
(
2002
)
Association of the human SUMO-1 protease SENP2 with the nuclear pore
.
J. Biol. Chem.
277
,
19961
19966
40
Reverter
D.
and
Lima
C.D.
(
2004
)
A basis for SUMO protease specificity provided by analysis of human Senp2 and a Senp2-SUMO complex
.
Structure
12
,
1519
1531
41
Guo
C.
,
Hildick
K.L.
,
Luo
J.
,
Dearden
L.
,
Wilkinson
K.A.
and
Henley
J.M.
(
2013
)
SENP3-mediated deSUMOylation of dynamin-related protein 1 promotes cell death following ischaemia
.
EMBO J.
32
,
1514
1528
42
Zunino
R.
,
Schauss
A.
,
Rippstein
P.
,
Andrade-Navarro
M.
and
McBride
H.M.
(
2007
)
The SUMO protease SENP5 is required to maintain mitochondrial morphology and function
.
J. Cell Sci.
120
,
1178
1188
43
Di-Bacco
A.
,
Ouyang
J.
,
Lee
H.Y.
,
Catic
A.
,
Ploegh
H.
and
Gill
G.
(
2006
)
The SUMO-specific protease SENP5 is required for cell division
.
Mol. Cell Biol.
26
,
4489
4498
44
Shen
L.N.
,
Geoffroy
M.C.
,
Jaffray
E.G.
and
Hay
R.T.
(
2009
)
Characterization of SENP7, a SUMO-2/3-specific isopeptidase
.
Biochem. J.
421
,
223
230
45
Lima
C.D.
and
Reverter
D.
(
2008
)
Structure of the human SENP7 catalytic domain and poly- SUMO deconjugation activities for SENP6 and SENP7
.
J. Biol. Chem.
283
,
32045
32055
46
Anderson
D.B.
,
Zanella
C.A.
,
Henley
J.M.
and
Cimarosti
H.
(
2017
)
Sumoylation: implications for neurodegenerative diseases
.
Adv. Exp. Med. Biol.
963
,
261
281
47
Dorval
V.
and
Fraser
P.E.
(
2006
)
Small ubiquitin-like modifier (SUMO) modification of natively unfolded proteins tau and alpha-synuclein
.
J. Biol. Chem.
281
,
9919
9924
48
Zhang
Y.Q.
and
Sarge
K.D.
(
2008
)
Sumoylation of amyloid precursor protein negatively regulates Abeta aggregate levels
.
Biochem. Biophys. Res. Commun.
374
,
673
678
49
Geoffroy
M.C.
and
Hay
R.T.
(
2009
)
An additional role for SUMO in ubiquitin-mediated proteolysis
.
Nat. Rev. Mol. Cell Biol.
10
,
564
568
50
Haass
C.
and
Selkoe
D.J.
(
2007
)
Soluble protein oligomers in neurodegeneration: lessons from the Alzheimer’s amyloid β-peptide
.
Nat. Rev. Mol. Cell Biol.
8
,
101
112
51
Thornton
E.
,
Vink
R.
,
Blumbergs
P.C.
and
Heuvel
C.V.D.
(
2006
)
Soluble amyloid precursor protein a reduces neuronal injury and improves functional outcome following diffuse traumatic brain injury in rats
.
Brain Res.
1094
,
38
46
52
Harris
M.E.
,
Wang
Y.
,
Pedigo
N.W.
Jr
,
Hensley
K.
,
Butterfield
D.A.
and
Carney
J.M.
(
1996
)
Amyloid β peptide (25–35) inhibits Na+-dependent glutamate uptake in rat hippocampal astrocyte cultures
.
J. Neurochem.
67
,
277
286
53
Li
Y.
,
Wang
H.
,
Wang
S.
,
Quon
D.
,
Liu
Y.W.
and
Cordell
B.
(
2003
)
Positive and negative regulation of APP amyloidogenesis by sumoylation
.
Proc. Natl. Acad. Sci. U.S.A.
100
,
259
264
54
Dorval
V.
,
Mazzella
M.J.
,
Mathews
P.M.
,
Hay
R.T.
and
Fraser
P.E.
(
2007
)
Modulation of Abeta generation by small ubiquitin-like modifiers does not require conjugation to target proteins
.
Biochem. J.
404
,
309
316
55
Yun
S.M.
,
Cho
S.J.
,
Song
J.C.
,
Song
S.Y.
,
Jo
S.A.
,
Jo
C.
et al
(
2013
)
SUMO1 modulates Abeta generation via BACE1 accumulation
.
Neurobiol. Aging
34
,
650
662
56
Weingarten
M.D.
,
Lockwood
A.H.
,
Hwo
S.Y.
and
Kirschner
M.W.
(
1975
)
A protein factor essential for microtubule assembly
.
Proc. Natl Acad. Sci. U.S.A.
72
,
1858
1862
57
Selkoe
D.
(
2001
)
Alzheimer’s disease: genes, proteins, and therapy
.
Physiol. Rev.
81
,
741
766
58
Cho
S.J.
,
Yun
S.M.
,
Lee
D.H.
,
Jo
C.
,
Ho-Park
M.
,
Han
C.
et al
(
2015
)
Plasma SUMO-1 protein is elevated in Alzheimer's disease
.
J. Alzheimers Dis.
47
,
639
643
59
Lee
L.
,
Dale
E.
,
Staniszewski
A.
,
Zhang
H.
,
Saeed
F.
,
Sakurai
M.
et al
(
2014
)
Regulation of synaptic plasticity and cognition by SUMO in normal physiology and Alzheimer's disease
.
Sci. Rep.
4
,
7190
60
McMillan
L.E.
,
Brown
J.T.
,
Henley
J.M.
and
Cimarosti
H.
(
2011
)
Profiles of SUMO and ubiquitin conjugation in an Alzheimer's disease model
.
Neurosci. Lett.
502
,
201
208
61
Nistico
R.
,
Ferraina
C.
,
Marconi
V.
,
Blandini
F.
,
Negri
L.
,
Egebjerg
J.
et al
(
2014
)
Age-related changes of protein SUMOylation balance in the AbetaPP Tg2576 mouse model of Alzheimer's disease
.
Front. Pharmacol.
5
,
63
62
Binda
C.S.
,
Heimann
M.J.
,
Duda
J.K.
,
Muller
M.
,
Henley
J.M.
and
Wilkinson
K.A.
(
2017
)
Analysis of protein SUMOylation and SUMO pathway enzyme levels in Alzheimer’s disease and Down's syndrome
.
Opera Med. Phys.
3
,
19
24
63
Eckermann
K.
(
2013
)
SUMO and Parkinson’s disease
.
Neuromolecular Med.
15
,
737
759
64
Guerra de Souza
A.C.
,
Prediger
R.D.
and
Cimarosti
H.
(
2016
)
SUMO-regulated mitochondrial function in Parkinson’s disease
.
J. Neurochem.
137
,
673
686
65
Maroteaux
L.
,
Campanelli
J.T.
and
Scheller
R.H.
(
1988
)
Synuclein: a neuron-specific protein localized to the nucleus and presynaptic nerve terminal
.
J. Neurosci.
8
,
2804
2815
66
Golbe
L.I.
,
Iorio
G.
,
Di-Bonavita
V.
,
Miller
D.C.
and
Duvoisin
R.C.
(
1990
)
A large kindred with autosomal dominant Parkinson’s disease
.
Ann. Neurol.
27
,
276
282
67
Chandra
S.
,
Fornai
F.
,
Kwon
H.B.
,
Yazdani
U.
,
Atasoy
D.
,
Liu
X.
et al
(
2004
)
Double-knockout mice for alpha- and beta-synucleins: effect on synaptic functions
.
Proc. Natl. Acad. Sci. U.S.A.
101
,
14966
14971
68
Chandra
S.
,
Gallardo
G.
,
Fernandez-Chacon
R.
,
Schluter
O.M.
and
Sudhof
T.C.
(
2005
)
Alpha-synuclein cooperates with CSPalpha in preventing neurodegeneration
.
Cell
123
,
383
396
69
Oh
Y.
,
Kim
Y.M.
,
Mouradian
M.M.
and
Chung
K.C.
(
2011
)
Human polycomb protein 2 promotes alpha-synuclein aggregate formation through covalent SUMOylation
.
Brain Res.
1381
,
78
89
70
Wong
M.J.L.
,
Cook
A.L.
,
Mackay-Sim
A.
and
Poutney
D.L.
(
2012
)
Differential SUMO-1 distribution in Parkinson’s disease patient neurosphere-derived cells in response to proteolytic stress (Abstract)
.
Proteostasis and Disease Symposium
71
Wong
M.B.
,
Goodwin
J.
,
Norazit
A.
,
Meedeniya
A.C.
,
Richter-Landsberg
C.
,
Gai
W.P.
et al
(
2013
)
SUMO-1 is associated with a subset of lysosomes in glial protein aggregate diseases
.
Neurotox. Res.
23
,
1
21
72
Um
J.W.
and
Chung
K.C.
(
2006
)
Functional modulation of parkin through physical interaction with SUMO-1
.
J. Neurosc. Res.
84
,
1543
1554
73
Bonifati
V.
,
Rizzu
P.
,
van-Baren
M.J.
,
Schaap
O.
,
Breedveld
G.J.
,
Krieger
E.
et al
(
2003
)
Mutations in the DJ-1 gene associated with autosomal recessive early-onset parkinsonism
.
Science
299
,
256
259
74
Taira
T.
,
Saito
Y.
,
Niki
T.
,
Iguchi-Ariga
S.M.
,
Takahashi
K.
and
Ariga
H.
(
2004
)
DJ-1 has a role in antioxidative stress to prevent cell death
.
EMBO Rep.
5
,
213
218
75
Shinbo
Y.
,
Niki
T.
,
Taira
T.
,
Ooe
H.
,
Takahashi-Niki
K.
,
Maita
C.
et al
(
2006
)
Proper SUMO-1 conjugation is essential to DJ-1 to exert its full activities
.
Cell Death Differ.
13
,
96
108
76
Janer
A.
,
Werner
A.
,
Takahashi-Fujigasaki
J.
,
Daret
A.
,
Fujigasaki
H.
,
Takada
K.
et al
(
2010
)
SUMOylation attenuates the aggregation propensity and cellular toxicity of the polyglutamine expanded ataxin-7
.
Hum. Mol. Genet.
19
,
181
195
77
Lee
Y.J.
,
Miyake
S.
,
Wakita
H.
,
McMullen
D.C.
,
Azuma
Y.
,
Auh
S.
et al
(
2007
)
Protein SUMOylation is massively increased in hibernation torpor and is critical for the cytoprotection provided by ischemic preconditioning and hypothermia in SHSY5Y cells
.
J. Cereb. Blood Flow Metab.
27
,
950
962
78
Cimarosti
H.
,
Lindberg
C.
,
Bomholt
S.F.
,
Ronn
L.C.
and
Henley
J.M.
(
2008
)
Increased protein SUMOylation following focal cerebral ischemia
.
Neuropharmacology
54
,
280
289
79
Yang
W.
,
Sheng
H.
,
Homi
H.M.
,
Warner
D.S.
and
Paschen
W.
(
2008
)
Cerebral ischemia/stroke and small ubiquitin-like modifier (SUMO) conjugation–a new target for therapeutic intervention?
J. Neurochem.
106
,
989
999
80
Sarge
K.D.
and
Park-Sarge
O.K.
(
2011
)
SUMO and its role in human diseases
.
Int. Rev. Cell Mol. Biol.
288
,
167
183
81
Cimarosti
H.
,
Ashikaga
E.
,
Jaafari
N.
,
Dearden
L.
,
Rubin
P.
,
Wilkinson
K. A.
et al
(
2012
)
Enhanced SUMOylation and SENP-1 protein levels following oxygen and glucose deprivation in neurones
.
J. Cereb. Blood Flow Metab.
32
,
17
22
82
Wilson
T.J.
,
Slupe
A.M.
and
Strack
S.
(
2013
)
Cell signaling and mitochondrial dynamics: Implications for neuronal function and neurodegenerative disease
.
Neurobiol. Dis.
51
,
13
26
83
Chang
C.R.
and
Blackstone
C.
(
2010
)
Dynamic regulation of mitochondrial fission through modification of the dynamin-related protein Drp1
.
Ann. NY Acad. Sci.
1201
,
34
39
84
Molinaro
P.
,
Cuomo
O.
,
Pignataro
G.
,
Boscia
F.
,
Sirabella
R.
,
Pannaccione
A.
et al
(
2008
)
Targeted disruption of Na+/Ca2+ exchanger 3 (NCX3) gene leads to a worsening of ischemic brain damage
.
J. Neurosci.
28
,
1179
1184
85
Cuomo
O.
,
Pignataro
G.
,
Sirabella
R.
,
Molinaro
P.
,
Anzilotti
S.
,
Scorziello
A.
et al
(
2016
)
SUMOylation of LYS590 of NCX3 f-Loop by SUMO-1 participates in brain neuroprotection induced by ischemic preconditioning
.
Stroke
47
,
1085
1093
86
Dutting
E.
,
Schroder-Kress
N.
,
Sticht
H.
and
Enz
R.
(
2011
)
SUMO E3 ligases are expressed in the retina and regulate SUMOylation of the metabotropic glutamate receptor 8b
.
Biochem. J.
435
,
365
371
87
Konopacki
F.A.
,
Jaafari
N.
,
Rocca
D.L.
,
Wilkinson
K.A.
,
Chamberlain
S.
,
Rubin
P.
et al
(
2011
)
Agonist-induced PKC phosphorylation regulates GluK2 SUMOylation and kainate receptor endocytosis
.
Proc. Natl. Acad. Sci. U.S.A.
108
,
19772
19777
88
Caraci
F.
,
Battaglia
G.
,
Sortino
M. A.
,
Spampinato
S.
,
Molinaro
G.
,
Copani
A.
et al
(
2012
)
Metabotropic glutamate receptors in neurodegeneration/neuroprotection: Still a hot topic?
Neurochem. Int.
61
,
559
565
89
Zhu
Q.J.
,
Xu
Y.
,
Du
C. P.
and
Hou
X.Y.
(
2012
)
SUMOylation of the kainate receptor subunit GluK2 contributes to the activation of the MLK3-JNK3 pathway following kainate stimulation
.
FEBS Lett.
586
,
1259
1264
90
Schorova
L.
and
Martin
S.
(
2016
)
SUMOylation in synaptic function and dysfunction
.
Front. Synaptic Neurosci.
8
,
1
24
91
Wheeler
D.B.
,
Randall
A.
and
Tsien
R.W.
(
1994
)
Roles of N-type and Q-type Ca2+ channels in supporting hippocampal synaptic transmission
.
Science
264
,
107
111
92
Tanabe
T.
,
Beam
K.G.
,
Adams
B.A.
,
Niidome
T.
and
Numa
S.
(
1990
)
Regions of the skeletal muscle dihydropyridine receptor critical for excitation-contraction coupling
.
Nature
346
,
567
569
93
Dolmetsch
R.E.
,
Pajvani
U.
,
Fife
K.
,
Spotts
J.M.
and
Greenberg
M.E.
(
2001
)
Signaling to the nucleus by an L-type calcium channel-calmodulin complex through the MAP kinase pathway
.
Science
294
,
333
339
94
Wheeler
D.G.
,
Groth
R.D.
,
Ma
H.
,
Barret
C.F.
,
Owen
S.F.
,
Safa
P.
et al
(
2012
)
CaV.l1 and CaV2 channels engage distinct modes of Ca2+ signaling to control CREB dependent gene expression
.
Cell
149
,
1112
1124
95
Clapham
D.E.
(
2007
)
Calcium signaling
.
Cell
131
,
1047
1058
96
Zamponi
G.W.
(
2016
)
Targeting voltage-gated calcium channels in neurological and psychiatric diseases
.
Nat. Rev.
15
,
19
34
97
Hagiwara
S.
,
Ozawa
S.
and
Sand
O.
(
1975
)
Voltage clamp analysis of two inward current mechanisms in the egg cell membrane of a starfish
.
J. Gen. Physiol.
65
,
617
644
98
Bean
B.P.
(
1989
)
Classes of calcium channels in vertebrate cells
.
Annu. Rev. Physiol.
51
,
367
384
99
Nowycky
M.C.
,
Fox
A.P.
and
Tsien
R.W.
(
1985
)
Three types of neuronal calcium channel with different calcium agonist sensitivity
.
Nature
316
,
440
443
100
Zamponi
G.W.
,
Striessnig
J.
,
Koschak
A.
and
Dolphin
A.C
(
2015
)
The physiology, pathology, and pharmacology of voltage-gated calcium channels and their future therapeutic potential
.
Pharmacol. Rev.
67
,
821
870
101
Berridge
M.J.
(
2014
)
Ion channels
.
Cell Signal. Biol.
1
74
102
Felix
R.
(
2006
)
Calcium channelopathies
.
Neuromolecular Med.
8
,
307
318
103
Lory
P.
and
Mezghrani
A.
(
2010
)
Calcium channelopathies in inherited neurological disorders: relevance to drug screening for acquired channel disorders
.
IDrugs
13
,
467
471
104
Oliveira
A.M.
,
Bading
H.
and
Mauceri
D.
(
2014
)
Dysfunction of neuronal calcium signaling in aging and disease
.
Cell Tissue Res.
2
,
381
383
105
Steinlein
O.K.
(
2014
)
Calcium signaling and epilepsy
.
Cell Tissue Res.
2
,
385
393
106
Schäfer
M.K.E.
,
Pfeiffer
A.
,
Jaeckel
M.
,
Pouya
A.
,
Dolga
A.M.
and
Methner
A.
(
2014
)
Regulators of mitochondrial Ca2+ homeostasis in cerebral ischemia
.
Cell Tissue Res.
357
,
395
405
107
Bourinet
E.
,
Altier
C.
,
Hildebrand
M.E.
,
Trang
T.
,
Salter
M.W.
and
Zamponi
G.W.
(
2014
)
Calcium-permeable ion channels in pain signaling
.
Physiol. Rev.
94
,
81
140
108
Waxman
S.G.
and
Zamponi
G.W.
(
2014
)
Regulating excitability of peripheral afferents: emerging ion channel targets
.
Nat. Neurosci.
17
,
153
163
109
Miljanich
G.P.
(
2004
)
Ziconotide: neuronal calcium channel blocker for treating severe chronic pain
.
Curr. Med. Chem.
11
,
3029
3040
110
Rauck
R.L.
,
Wallace
M.S.
,
Burton
A.W.
,
Kapural
L.
and
North
J.M.
(
2009
)
Intrathecal ziconotide for neuropathic pain: a review
.
Pain Pract.
9
,
327
337
111
Smith
H.S.
and
Deer
T.R.
(
2009
)
Safety and efficacy of intrathecal ziconotide in the management of severe chronic pain
.
Ther. Clin. Risk Manag.
5
,
521
534
112
Cizkova
D.
,
Marsala
J.
,
Lukacova
N.
,
Marsala
M.
,
Jergova
S.
,
Orendacova
J.
et al
(
2002
)
Localization of N-type Ca2+ channels in the rat spinal cord following chronic constrictive nerve injury
.
Exp. Brain Res.
147
,
456
463
113
Jagodic
M.M.
,
Pathirathna
S.
,
Joksovic
P.M.
,
Lee
W.
,
Nelson
M.T.
,
Naik
A.K.
et al
(
2008
)
Upregulation of the T-type calcium current in small rat sensory neurons after chronic constrictive injury of the sciatic nerve
.
J. Neurophysiol.
99
,
3151
3156
114
Marger
F.
,
Gelot
A.
,
Alloui
A.
,
Matricon
J.
,
Ferrer
J.F.
,
Barrère
C.
et al
(
2011
)
T-type calcium channels contribute to colonic hypersensitivity in a rat model of irritable bowel syndrome
.
Proc. Natl. Acad. Sci. U.S.A.
108
,
11268
11273
115
Bezprozvanny
I.
and
Mattson
M.P.
(
2008
)
Neuronal calcium mishandling and the pathogenesis of Alzheimer’s disease
.
Trends Neurosci.
31
,
454
463
116
Brawek
B.
and
Garaschuk
O.
(
2014
)
Network-wide dysregulation of calcium homeostasis in Alzheimer’s disease
.
Cell Tissue Res.
357
,
427
438
117
Mezler
M.
,
Barghorn
S.
,
Schoemaker
H.
,
Gross
G.
and
Nimmrich
V.
(
2012
)
A β-amyloid oligomer directly modulates P/Q-type calcium currents in Xenopus oocytes
.
Br. J. Pharmacol.
165
,
1572
1583
118
Hermann
D.
,
Mezler
M.
,
Müller
M.K.
,
Wicke
K.
,
Gross
G.
,
Draguhn
A.
et al
(
2013
)
Synthetic Aβ oligomers (Aβ(1-42) globulomer) modulate presynaptic calcium currents: prevention of Aβ-induced synaptic deficits by calcium channel blockers
.
Eur. J. Pharmacol.
702
,
44
55
119
Rush
T.
and
Buisson
A.
(
2014
)
Reciprocal disruption of neuronal signaling and Aβ production mediated by extrasynaptic NMDA receptors: a downward spiral
.
Cell Tissue Res.
356
,
279
286
120
Putzier
I.
,
Kullmann
P.H.
,
Horn
J.P.
and
Levitan
E.S.
(
2009
)
CaV1.3 channel voltage dependence, not Ca2+ selectivity, drives pacemaker activity and amplifies bursts in nigral dopamine neurons
.
J. Neurosci.
29
,
15414
15419
121
Surmeier
D.J.
and
Schumacker
P.T.
(
2013
)
Calcium, bioenergetics, and neuronal vulnerability in Parkinson’s disease
.
J. Biol. Chem.
288
,
10736
10741
122
Rcom-H’cheo-Gauthier
A.
,
Goodwin
J.
and
Pountney
D.L.
(
2014
)
Interactions between calcium and alpha-synuclein in neurodegeneration Biomolecules
4
,
795
811
123
Zhuchenko
O.
,
Bailey
J.
,
Bonnen
P.
,
Ashizawa
T.
,
Stockton
D.W.
,
Amos
C.
et al
(
1997
)
Autosomal dominant cerebellar ataxia (SCA6) associated with small polyglutamine expansions in the alpha 1A-voltage-dependent calcium channel
.
Nat. Genet.
15
,
62
69
124
Davila
M.A.
,
Chan
H.
and
Piedras-Renteria
E.S.
(
2010
)
SUMOylation of voltage-gated alphaA1a calcium channels
.
Biophys. J.
98
,
692a
693a
125
Feligioni
M.
,
Nishimune
A.
and
Henley
J.M.
(
2009
)
Protein SUMOylation modulates calcium influx and glutamate release from presynaptic terminals
.
Eur. J. Neurosci.
29
,
1348
1356
126
Müller
C.S.
,
Haupt
A.
,
Bildl
W.
,
Schindler
J.
,
Knaus
H.G.
,
Meissner
M.
et al
(
2010
)
Quantitative proteomics of the CaV2 channel nano-environments in the mammalian brain
.
Proc. Natl. Acad. Sci. U.S.A.
107
,
14950
14957
127
Brittain
J.M.
,
Piekarz
A.D.
,
Wang
Y.
,
Kondo
T.
,
Cummins
T.R.
and
Khanna
R.
(
2009
)
An atypical role for collapsin response mediator protein 2 (CRMP-2) in neurotransmitter release via interaction with presynaptic voltage-gated calcium channels
.
J. Biol. Chem.
284
,
31375
31390
128
Ju
W.
,
Li
Q.
,
Wilson
S.M.
,
Brittain
J.M.
,
Meroueh
L.
and
Khanna
R.
(
2013
)
SUMOylation alters CRMP2 regulation of calcium influx in sensory neurons
.
Channels (Austin)
7
,
153
159
129
Dustrude
E.T.
,
Wilson
S.M.
,
Ju
W.
,
Xiao
Y.
and
Khanna
R.
(
2013
)
CRMP2 protein SUMOylation modulates NaV1.7 channel trafficking
.
J. Biol. Chem.
288
,
24316
24331
130
Dib-Hajj
S. D.
,
Yang
Y.
,
Black
J.A.
and
Waxman
S.G.
(
2013
)
The NaV1.7 sodium channel: from molecule to man
.
Nat. Rev. Neurosci.
14
,
49
62
131
Deng
L.
,
Kaeser
P.S.
,
Xu
W.
and
Sudhof
T.C.
(
2011
)
RIM proteins activate vesicle priming by reversing autoinhibitory homodimerization of Munc13
.
Neuron
69
,
317
331
132
Kaeser
P.S.
,
Deng
L.
,
Wang
Y.
,
Dulubova
I.
,
Liu
X.
,
Rizo
J.
et al
(
2011
)
RIM proteins tether Ca2+ channels to presynaptic active zones via a direct PDZ-domain interaction
.
Cell
144
,
282
295
133
Girach
F.
,
Craig
T.J.
,
Rocca
D.L.
and
Henley
J.M.
(
2013
)
RIM1a SUMOylation is required for fast synaptic vesicle exocytosis
.
Cell Rep.
5
,
1294
1301
134
Südhof
T.C.
(
2013
)
Neurotransmitter release: the last millisecond in the life of a synaptic vesicle
.
Neuron
80
,
675
690
135
Xu
J.
,
Luo
F.
,
Zhang
Z.
,
Xue
L.
,
Wu
X.S.
,
Chiang
H.C.
et al
(
2013
)
SNARE proteins synaptobrevin, SNAP-25, and syntaxin are involved in rapid and slow endocytosis at synapses
.
Cell Rep.
3
,
1414
1421
136
Zhang
Z.
,
Wang
D.
,
Sun
T.
,
Xu
J.
,
Chiang
H.C.
,
Shin
W.
et al
(
2013
)
The SNARE proteins SNAP25 and synaptobrevin are involved in endocytosis at hippocampal synapses
.
J. Neurosci.
33
,
9169
9175
137
Craig
T.J.
,
Anderson
D.
,
Evans
A.J.
,
Girach
F.
and
Henley
J.M.
(
2015
)
SUMOylation of syntaxin1A regulates presynaptic endocytosis
.
Sci. Rep.
5
,
17669
138
Tang
L.T.
,
Craig
T.J.
and
Henley
J.M.
(
2015
)
SUMOylation of synapsin Ia maintains synaptic vesicle availability and is reduced in an autism mutation
.
Nat. Commun.
6
,
7728
139
Matsuzaki
S.
,
Lee
L.
,
Knock
E.
,
Srikumar
T.
,
Sakurai
M.
,
Hazrati
L.N.
et al
(
2015
)
SUMO-1 affects synaptic function, spine density and memory
.
Sci. Rep.
5
,
10730
140
Lisman
J.
,
Schulman
H.
and
Cline
H.
(
2002
)
The molecular basis of CaMKII function in synaptic and behavioural memory
.
Nat. Rev. Neurosci.
3
,
175
190
141
Long
X.
and
Griffith
L.C.
(
2000
)
Identification and characterization of a SUMO-1 conjugation system that modifies neuronal calcium/calmodulin-dependent protein kinase II in Drosophila melanogaster
.
J. Biol. Chem.
275
,
40765
40776
142
Flavell
S.W.
,
Cowan
C.W.
,
Kim
T.K.
,
Greer
P.L.
,
Lin
Y.
,
Paradis
S.
et al
(
2006
)
Activity-dependent regulation of MEF2 transcription factors suppresses excitatory synapse number
.
Science
311
,
1008
1012
143
Hietakangas
V.
,
Anckar
J.
,
Blomster
H.A.
,
Fujimoto
M.
,
Palvimo
J.J.
,
Nakai
A.
et al
(
2006
)
PDSM, a motif for phosphorylation-dependent SUMO modification
.
Proc. Natl. Acad. Sci. U.S.A.
103
,
45
50
144
Kang
J.
,
Gocke
C.B.
and
Yu
H.
(
2006
)
Phosphorylation-facilitated SUMOylation of MEF2C negatively regulates its transcriptional activity
.
BMC Biochem.
7
,
5
145
Riquelme
C.
,
Barthel
K.K.
and
Liu
X.
(
2006
)
SUMO-1 modification of MEF2A regulates its transcriptional activity
.
J. Cell. Mol. Med.
10
,
132
144
146
Zhao
X.
,
Sternsdorf
T.
,
Bolger
T.A.
,
Evans
R.M.
and
Yao
T.P.
(
2005
)
Regulation of MEF2 by histone deacetylase 4- and SIRT1 deacetylase-mediated lysine modifications
.
Mol. Cell. Biol.
25
,
8456
8464
147
Shalizi
A.
,
Gaudillière
B.
,
Yuan
Z.
,
Stegmüller
J.
,
Shirogane
T.
,
Ge
Q.
et al
(
2006
)
A calcium-regulated MEF2 SUMOylation switch controls postsynaptic differentiation
.
Science
311
,
1012
1017
148
Zündorf
G.
and
Reiser
G.
(
2011
)
Calcium dysregulation and homeostasis of neural calcium in the molecular mechanisms of neurodegenerative diseases provide multiple targets for neuroprotection
.
Antioxid. Redox Signal.
14
,
1275
1288
149
Krumova
P.
,
Meulmeester
E.
,
Garrido
M.
,
Tirard
M.
,
Hsiao
H. H.
,
Bossis
G
et al
(
2011
)
SUMOylation inhibits α-synuclein aggregation and toxicity
.
J. Cell Biol.
194
,
49
60
150
Abeywardana
T.
and
Pratt
M.R.
(
2015
)
Extent of inhibition of α synuclein aggregation in vitro by SUMOylation is conjugation site- and SUMO isoform-selective
.
Biochem
54
,
959
961
151
Ronzitti
G.
,
Bucci
G.
,
Emanuele
M.
,
Leo
D.
,
Sotnikova
T.D.
,
Mus
L.V.
et al
(
2014
)
Exogenous α-synuclein decreases raft partitioning of Cav2.2 channels inducing dopamine release
.
J. Neurosci.
34
,
10603
10615
152
Oh
Y.
,
Kim
Y.M.
,
Mouradian
M.M.
and
Chung
K.C.
(
2011
)
Human polycomb protein 2 promotes α-synuclein aggregate formation through covalent SUMOylation
.
Brain Res.
1381
,
78
89
153
Wang
L.
,
Wansleeben
C.
,
Zhao
S.
,
Miao
P.
,
Paschen
W.
and
Yang
W.
(
2014
)
SUMO-2 is essential while SUMO3 is dispensable for mouse embryonic development
.
EMBO Rep.
15
,
878
885
154
Wang
J.
,
Chen
L.
,
Wen
S.
,
Zhu
H.
,
Yu
W.
,
Moskowitz
I.P.
et al
(
2011
)
Defective SUMOylation pathway directs congenital heart disease
.
Birth Defects Res. A Clin. Mol. Teratol.
91
,
468
476
155
Kim
E.Y.
,
Zhang
Y.
,
Ye
B.
,
Segura
A.M.
,
Beketaev
I.
,
Xi
Y.
et al
(
2015
)
Involvement of activated SUMO-2 conjugation in cardiomyopathy
.
Biochim. Biophys. Acta
1852
,
1388
1399
156
Kho
C.
,
Lee
A.
,
Jeong
D.
,
Oh
J.G.
,
Chaanine
A.H.
,
Kizana
E.
et al
(
2011
)
SUMO-1-dependent modulation of SERCA2a in heart failure
.
Nature
477
,
601
605
157
Kim
E.Y.
,
Chen
L.
,
Ma
Y.
,
Yu
W.
,
Chang
J.
,
Moskowitz
I.P.
et al
(
2012
)
Enhanced deSUMOylation in murine hearts by overexpressed SENP2 leads to congenital heart defects and cardiac dysfunction
.
J. Mol. Cell. Cardiol.
52
,
638
649
158
Mendler
L.
,
Braun
T.
and
Müller
S.
(
2016
)
The ubiquitin-like SUMO system and heart function from development to disease
.
Circ. Res.
118
,
132
144
159
Lee
Y.J.
and
Hallenbeck
J.M.
(
2013
)
SUMO and ischemic tolerance
.
Neuromol. Med.
15
,
771
781
160
Guo
C.
and
Henley
J.M.
(
2014
)
Wrestling with stress: roles of protein SUMOylation and deSUMOylation in cell stress response
.
IUBMB Life
66
,
71
77
161
Gupta
M.K.
,
Gulick
J.
,
Liu
R.
,
Wang
X.
,
Molkentin
J.D.
and
Robbins
J.
(
2014
)
SUMO E2 enzyme UBC9 is required for efficient protein quality control in cardiomyocytes
.
Circ. Res.
115
,
721
729
162
Maejima
Y.
and
Sadoshima
J.
(
2014
)
SUMOylation: a novel protein quality control modifier in the heart
.
Circ. Res.
115
,
686
689
163
Matsuzaki
K.
,
Minami
T.
,
Tojo
M.
,
Honda
Y.
,
Uchimura
Y.
,
Saitoh
H.
et al
(
2003
)
Serum response factor is modulated by the SUMO-1 conjugation system
.
Biochem. Biophys. Res. Commun.
306
,
32
38
164
Wang
J.
,
Feng
X.H.
and
Schwartz
R.J.
(
2004
)
SUMO-1 modification activated GATA4-dependent cardiogenic gene activity
.
J. Biol. Chem.
279
,
49091
49098
165
Wang
J.
,
Li
A.
,
Wang
Z.
,
Feng
X.
,
Olson
E.N.
and
Schwartz
R.J.
(
2007
)
Myocardin SUMOylation transactivates cardiogenic genes in pluripotent 10T1/2 fibroblasts
.
Mol. Cell Biol.
27
,
622
632
166
Pal
S.
,
Hartnett
K.A.
,
Nerbonne
J.M.
,
Levitan
E.S.
and
Aizenman
E.
(
2003
)
Mediation of neuronal apoptosis by KV2.1-encoded potassium channels
.
J. Neurosci.
23
,
4798
4802
167
Stapels
M.
,
Piper
C.
,
Yang
T.
,
Li
M.
,
Stowell
C.
,
Xiong
Z.G.
et al
(
2010
)
Polycomb group proteins as epigenetic mediators of neuroprotection in ischemic tolerance
.
Sci. Signal.
3
,
ra15
168
Rougier
J.S.
,
Albesa
M.
and
Abrie
I.H.
(
2010
)
Ubiquitylation and SUMOylation of cardiac ion channels
.
J. Cardiovasc. Pharmacol.
56
,
22
28
169
Zhang
Y.Q.
and
Sarge
K.D.
(
2008
)
SUMOylation regulates lamin A function and is lost in lamin A mutants associated with familial cardiomyopathies
.
J. Cell Biol.
182
,
35
39
170
Broers
J.L.
,
Ramaekers
F.C.
,
Bonne
G.
,
Yaou
R.B.
and
Hutchison
C.J.
(
2006
)
Nuclear lamins: laminopathies and their role in premature ageing
.
Physiol. Rev.
86
,
967
1008
171
Cartwright
E.J.
,
Mohamed
T.
,
Oceandy
D.
and
Neyses
L.
(
2011
)
Calcium signaling dysfunction in heart disease
.
Biofactors
37
,
175
181
172
Wilkins
B.J.
and
Molkentin
J.D.
(
2004
)
Calcium-calcineurin signaling in the regulation of cardiac hypertrophy
.
Biochem. Biophys. Res. Commun.
322
,
1178
1191
173
Bernt
A.
,
Rangrez
A.Y.
,
Eden
M.
,
Jungmann
A.
,
Katz
S.
,
Rohr
C.
et al
(
2016
)
SUMOylation-independent activation of Calcineurin-NFAT signaling via SUMO-2 mediates cardiomyocyte hypertrophy
.
Sci. Rep.
6
,
35758
174
Woodcock
E.A.
and
Matkovich
S.J.
(
2005
)
Cardiomyocytes structure, function and associated pathologies
.
Int. J. Biochem. Cell Biol.
37
,
1746
1751
175
Meyer
M.
,
Schillinger
W.
,
Pieske
B.
,
Holubarsch
C.
,
Heilmann
C.
,
Posival
H.
et al
(
1995
)
Alterations of sarcoplasmic reticulum proteins in failing human dilated cardiomyopathy
.
Circulation
92
,
778
784
176
Melchior
F.
,
Paschal
B.
,
Evans
J.
and
Gerace
L.
(
1993
)
Inhibition of nuclear protein import by nonhydrolyzable analogues of GTP and identification of the small GTPase Ran/TC4 as an essential transport factor
.
J. Cell Biol.
123
,
1649
1659
177
Kho
C.
,
Lee
A.
,
Jeong
D.
,
Oh
J.G.
,
Gorski
P.A.
,
Fish
K.
et al
(
2015
)
Small-molecule activation of SERCA2a SUMOylation for the treatment of heart failure
.
Nat. Commun.
7229
,
1
11
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).