Cell membrane repair is a critical process used to maintain cell integrity and survival from potentially lethal chemical, and mechanical membrane injury. Rapid increases in local calcium levels due to a membrane rupture have been widely accepted as a trigger for multiple membrane-resealing models that utilize exocytosis, endocytosis, patching, and shedding mechanisms. Calcium-sensor proteins, such as synaptotagmins (Syt), dysferlin, S100 proteins, and annexins, have all been identified to regulate, or participate in, multiple modes of membrane repair. Dysfunction of membrane repair from inefficiencies or genetic alterations in these proteins contributes to diseases such as muscular dystrophy (MD) and heart disease. The present review covers the role of some of the key calcium-sensor proteins and their involvement in membrane repair.

The plasma membrane is a protective barrier for cells, made up of a phospholipid bilayer with embedded proteins that carry out specific functions. An intact plasma membrane is vital for cell integrity and maintaining homeostasis providing a stable, yet semipermeable, platform for proteins and lipids to conduct nutrient intake, vesical transport, enzyme secretion and signaling within and outside of the cell. Unlike cell walls in plant and bacterial cells that provide greater structural support and strength, the plasma membranes of animals are more fragile and susceptible to injuries from the surrounding environment that can damage the phospholipid bilayer, eventually leading to cell death. Multiple mechanisms are possible to tear or rupture the plasma membrane including injury from mechanical, chemical, or bacterial/viral sources [1,2]. For example, in cardiac and skeletal muscle, the high mechanical stress placed on muscle fibers due to repeated contractions causes the plasma membrane in these cells to frequently rupture. Over time, the constant breaching of the plasma membrane in cardiac muscle cells can lead to heart failure. Reactive oxygen species through mitochondrial dysfunction can cause oxidation of proteins and lipids leading to pores in cell membranes, especially in cells with large numbers of mitochondria and increased metabolic activity such as liver and brain [3–5]. Chemical and mechanical cell damage occurs in oral cavities due to brushing, chewing, and tobacco usage [6,7]. Damage to skin cells is induced by ultraviolet radiation and physical abrasion [8]. Defects in neuronal membrane integrity are common in neurological conditions, including traumatic brain and spinal cord injuries, Parkinson’s disease [9], and Alzheimer’s disease [10]. Myocardial damage including ischemia-reperfusion injury can lead to heart failure [11–13]. Infection by pathogenic bacteria, viruses, and pore-forming toxins (PFTs) can result in transmembrane lesions at the cell membrane [14,15]. Bacteria associated with Shigella or Salmonella that cause intestinal infections utilize a ‘needle-like’ protein apparatus to breach a membrane and secrete proteins into host cells [16]. The spike protein of SARS-CoV2 latches on to cells and then uses the host’s enzymatic system to fuse with the plasma membrane to create a pore that facillitates infection [17]. In these cases, it is unclear what damage to the plasma membrane has been left behind. Several excellent reviews on the mechanisms and outcomes of plasma membrane rupture are available [11,12,18,19].

Nonetheless, cells have a remarkable ability to survive membrane disruption, frequently involving complex networks of proteins and lipids that patch a membrane lesion and return the cell to normal function. Multiple studies have demonstrated that calcium is a necessary requirement for membrane repair following disruption. For example, membrane-fusion that seals a torn plasma membrane is observed in sea urchin eggs, triggered by the rapid influx of extracellular calcium ions (Ca2+) [20]. In contrast, membrane damage is not repaired when Ca2+ is absent from the external environment, resulting in continuous spillage of cytoplasmic contents and cell death. Furthermore, upon microelectrode puncture wounding of frog skeletal muscle fibers seal effectively in the presence of Ca2+ but is suppressed at low Ca2+ concentrations [19]. In addition, frog oocytes wounded by microneedle fail to reorganize microtubules and heal the lesion in a Ca2+-free environment [21]. The key step in these membrane repair processes is the sensing of abnormal cytosolic Ca2+ concentrations by specific membrane repair proteins including dysferlin, synaptotagmins (Syt), caveolins, TRIM72, annexins, and others. Several excellent reviews have been published that describe different aspects of membrane repair [22–27]. In the present review, we focus on the roles of calcium-sensor proteins in membrane repair and the implications of dysfunctional membrane repair proteins on human disease.

Pathological conditions and chemical exposure that compromise the membrane repair process or membrane integrity pose immediate threats to cell survival. Generally, the location of the defective repair process dictates the type of illness or disease. Mutations in several genes that encode membrane repair proteins contribute to diseases such as muscular dystrophy (MD), gastrointestinal diseases, and heart disease [28,29]. The development and maintenance of skeletal and cardiac muscle is particularly affected (Figure 1) [30]. For instance, according to the Human Gene Mutation Database (www.hgmd.cf.ac.uk), nearly 400 missense mutations in the DYSF gene, responsible for the membrane repair protein dysferlin, are documented to cause weakness and deterioration in muscles of the arms and legs and manifested in multiple forms of limb-grindle muscular dystrophy (LGMD), Miyoshi myopathy, or general dysferlinopathies. Patients with Miyoshi myopathy also display early myocardial dysfunction prior to developing more serious cardiovascular problems. Similarly, over 350 mutations in the CAPN3 gene (calpain-3) have been identified where >90% are linked to LGMD. Mutations in the transmembrane protein caveolin-3, also involved in membrane repair, are causative for LGMD and ‘rippling-muscle’ disease but also implicated in cardiomyopathies and sudden-infant death syndrome [31]. Some rare disorders such as chronic inflammation of muscle cells are also characterized by changes in proteins involved in membrane repair [32]. An uncontrolled autoimmune response in skeletal muscle is believed to begin with a deficiency of Syt VII (Syt7) that results in autoantibodies-targeting TRIM72. The continued inflammatory response eventually leads to muscle degeneration [30,33,34]. While not involved in membrane repair dystrophin co-ordinates a large protein complex near the sarcolemma of muscle cells to maintain membrane integrity and prevent injury, especially during contraction. Deletions, truncations, or mutations of the gene (DMD) for dystrophin are responsible for Duchenne or Becker forms of MD [30,35,36].

Pathological conditions and symptoms associated with membrane repair pathways

Figure 1
Pathological conditions and symptoms associated with membrane repair pathways

Duchenne and Becker MD result from loss of muscle mass. Duchenne MD is also associated with increased risk of heart and respiratory failure, while Becker MD patients develop symptoms slower than Duchenne MD. LGMD (LGMD1 and LGMD2) manifests in the proximal muscles around the hips and shoulders. Miyoshi myopathy manifests in distal muscles, mainly in legs. Inflammatory myopathy is associated with weakness in the large muscles around the neck, shoulders, and hips. Two types of inflammatory myopathy are polymyositis and dermatomyositis that cause lung inflammation. Acute brain injury can lead to chronic and progressive neurodegeneration such as Parkinson’s disease and Alzheimer’s disease. Defects in membrane repair contribute to different cardiac and lung diseases.

Figure 1
Pathological conditions and symptoms associated with membrane repair pathways

Duchenne and Becker MD result from loss of muscle mass. Duchenne MD is also associated with increased risk of heart and respiratory failure, while Becker MD patients develop symptoms slower than Duchenne MD. LGMD (LGMD1 and LGMD2) manifests in the proximal muscles around the hips and shoulders. Miyoshi myopathy manifests in distal muscles, mainly in legs. Inflammatory myopathy is associated with weakness in the large muscles around the neck, shoulders, and hips. Two types of inflammatory myopathy are polymyositis and dermatomyositis that cause lung inflammation. Acute brain injury can lead to chronic and progressive neurodegeneration such as Parkinson’s disease and Alzheimer’s disease. Defects in membrane repair contribute to different cardiac and lung diseases.

Close modal

Multiple studies show that defects in membrane repair are manifested in different cardiac and lung diseases. The oxidative stress of ischemic injury and subsequent reperfusion is thought to trigger membrane repair proteins. Dysferlin, annexins, and TRIM72 are up-regulated in the latter stages of heart failure [11,37], perhaps in a last-ditch effort to rescue cardiac cells. Annexin A1 (ANX1) deficiency results in cardiac necrosis and fibrosis following myocardial infarction and other annexin proteins are up-regulated in a failing heart [38,39]. Recombinant TRIM72 has been shown to have a protective role and stabilize myofibers [40] and improve membrane repair in the heart [41,42]. As a result, administration of TRIM72 has been suggested as a possible therapeutic route following myocardial infarction [43]. Truncation and missense mutations have been observed in the repair protein AHNAK in congenital heart disease patients [44] and a missense mutation in apoptosis-linked gene 2 (ALG-2, also called programmed cell death 6 protein [PDCD6]) is observed in patients with left ventricular noncompaction [45]. Impaired left ventricle function has also been identified in patients carrying missense mutations in the dysferlin gene. In the lungs, both caveolin-1 and TRIM72 are important in the repair of epithelial cells in lung tissue [46,47]. The absence of either of these proteins causes abnormalities and diminishes the repair process.

Currently permanent solutions for any form of MD or inflammatory myopathies are not available for patients [11,30]. However, stimulating membrane sealing with gene therapy approaches appears to be a promising therapeutic strategy. Gene replacement strategies using overlapping viral vectors have been used in mice [48]. Exon skipping has been used to replace protein-coding and noncoding regions in the DYSF gene to rescue membrane repair in patient fibroblast cells [49]. In addition, CRISPR methods in patient-derived stem cells can be used to replace missense mutations in the DYSF gene [50].

The Ca2+ ion is accepted as an evolutionarily conserved cellular signaling molecule originally shown to control heart contraction and neurotransmitter release [54,55]. Through binding to a wide variety of proteins, Ca2+ has the ability to initiate conformational changes that signal muscle contraction, synaptic transmissions, and enzymic function [56]. Furthermore, Ca2+ can neutralize the negative potential of proteins and membrane surfaces fine-tuning electrostatic gradients in the cell. The observation that nearly all aspects of cellular life are affected by Ca2+ requires intracellular Ca2+ concentrations to be tightly regulated. Cytosolic Ca2+ concentrations are maintained near ∼100 nM by calcium transporters that actively pump Ca2+ ions from the cytosol to extracellular space (Na+/Ca2+ exchangers, voltage-dependent Ca2+ channels), from the cytosol to the endoplasmic reticulum (sarco/endoplasmic reticulum calcium ATPase, SERCA) or Ca2+-channels (ryanodine receptor, IP3 receptor) that release calcium from the endoplasmic reticulum to the cytosol [56,57]. In addition, calcium-sensor proteins with high affinities (Kd = 10−9 M) or high calcium capacity act as sponges to limit Ca2+ in the cytosol [58]. This results in steep concentration gradients between cytosolic Ca2+ levels and extracellular (2–3 mM) or intracellular storage (0.5–1 mM) environments. In many signaling processes, intracellular Ca2+ concentrations increase to 1–10 µM, controlled by Na+/Ca2+ exchangers or voltage-gated Ca2+ channels, in response to ligand binding or changes in membrane potential. Here, EF-hand Ca2+-binding proteins such as troponin-C and calmodulin have calcium affinities that are finely tuned to bind Ca2+ ions, undergo conformational changes and modulate interactions with other proteins to stimulate muscle contraction or enzymatic action [59–62].

Unlike this controlled import of Ca2+ in response to stimuli, membrane rupture causes a rapid, unrestrained influx of extracellular Ca2+ into a cell around the rupture site, resulting in an overwhelming surge in cytosolic Ca2+ [63]. In the absence of appropriate mechanisms to reseal the membrane fault, toxic levels of intracellular Ca2+ promote apoptosis and cell death. For small pores (<1 nm), membrane repair likely occurs using thermal fluctuations of phospholipids to reseal a hole, first observed in liposomes and red blood cells [64,65]. The mechanism relies on the unfavorable lipid disorder energetics of the hydrophobic phospholipid backbone exposed to the hydrophilic surroundings created at the rupture site [64,66]. As a result, an increase in surface energy drives lipids at the rupture site to form curved edges to minimize pore edge tension. Most injuries induced by mechanical or chemical stress produce larger membrane holes (>10 µm2) that are repaired rapidly (<10–30 s) [67,68]. In the presence of additional oxidative stress, this time frame can be extended (100–300 s) [69]. In all cases, membrane lesions require specialized repair mechanisms to detect and determine the nature of the injury, then treat it with Ca2+-sensitive proteomic machinery.

In general, there are about four calcium-sensitive mechanisms used by cells to repair larger holes in the plasma membrane: exocytosis, endocytosis (internalization), patching, and shedding (externalization) (Figure 2) [20,26,70–72]. In exocytosis, damage to the plasma membrane triggers Ca2+ influx and the localization of intracellular vesicles to the damaged site within seconds. Subsequent fusion of vesicles to the inner surface of the plasma membrane has been observed in sea urchin eggs, endothelial cells, and fibroblasts (Figure 2A) [20,73]. This mechanism that artificially adds more lipids via vesicle fusion to a broken plasma membrane reduces membrane tension efficiently. Membrane patching is similar to the exocytosis-mediated repair mechanism and also utilizes intracellular compartments. This Ca2+-dependent vesicle–vesicle fusion event is observed in starfish oocytes and sea urchin eggs [67]. The patching model relies on the rapid Ca2+-dependent fusion of vesicles to form a large multicompartment barrier that encompasses and fuses with the plasma membrane at the rupture site. This model allows sealing of much larger holes in the membrane (>10 nm in diameter) within seconds. The underlying proteomic machinery for exocytosis and patching requires Ca2+-sensor proteins that detect the injury sites and promote vesicle–vesicle and vesicle–plasma fusion events. Candidate Ca2+ sensor proteins for these mechanisms include members of the Syt and annexin protein families and dysferlin [12,74–77].

Active membrane repair pathways

Figure 2
Active membrane repair pathways

Larger lesions in a plasma membrane are repaired by (A) Exocytosis, (B) Endocytosis, (C) Patching and (D) Shedding. (A) Calcium influx activates vesicular exocytosis events. (B) Inward budding reseals small lesions (<100 nm) by creating lesion-containing vesicles that fuse with lysosomes targeted for endo-lysosomal degradation. (C) Exocytotic-induced vesicles fuse at the rupture site, forming a large patch that isolates the rupture site. (D) Shedding repairs holes by outward budding and pinching off of vesicles containing the damaged portion of the plasma membrane.

Figure 2
Active membrane repair pathways

Larger lesions in a plasma membrane are repaired by (A) Exocytosis, (B) Endocytosis, (C) Patching and (D) Shedding. (A) Calcium influx activates vesicular exocytosis events. (B) Inward budding reseals small lesions (<100 nm) by creating lesion-containing vesicles that fuse with lysosomes targeted for endo-lysosomal degradation. (C) Exocytotic-induced vesicles fuse at the rupture site, forming a large patch that isolates the rupture site. (D) Shedding repairs holes by outward budding and pinching off of vesicles containing the damaged portion of the plasma membrane.

Close modal

The endocytosis mechanism reseals membrane holes (<100 nm) through inward budding where the damaged plasma membrane is pinched off. The resulting lesion-containing vesicle then fuses with lysosomes and is targeted for endo-lysosomal degradation [78]. Membrane lesions induced by the bacterial toxin streptolysin O (SLO) show rapid Ca2+-dependent resealing by this mechanism [79]. The shedding mechanism repairs holes by outward budding of vesicles containing the damaged portion of the plasma membrane [80]. This mechanism utilizes proteins of the Endosomal Sorting Complex Required for Transport (ESCRT) machinery to repair smaller holes (<100 nm) [71]. As with endocytosis, the shedding mechanism appears to be used to repair SLO-infected cells [72,79]. In addition, Ca2+ activation promotes the accumulation of multiple proteins near the site of the membrane lesion, assembling a barrier that plugs the lesion to limit the loss of cytosolic media. The annexin protein family are major Ca2+ sensors for this protein accumulation process [81], and annexins appear to work with other proteins including the S100 proteins, dysferlin, and TRIM72 (MG53) during membrane repair [82–84].

Syt1 and Syt7 and soluble N-ethylmaleimide-sensitive factor attachment protein receptor proteins in exocytosis

The Syt family of Ca2+-binding proteins regulates Ca2+-dependent exocytosis in several cell types. The process is especially well characterized in neurons for neurotransmitter release from synaptic vesicles [85]. Syt proteins consist of 17 isoforms in mammals that each have an N-terminal transmembrane region, a variable length linker region and two cytosolic C2 domains (C2A, C2B) [86–88], first identified in protein kinase C [89,90]. Multiple structures of C2 domains from different Syt proteins show these share a conserved β-sandwich fold of eight antiparallel β-sheets connected by surface loops [91–96]. Most isoforms show tissue-specific distribution, whereas only nine isoforms (Syt1–Syt7, Syt9, Syt10) have C2 domains that bind calcium [97]. Typically, C2 domains bind two to three Ca2+ ions although up to four sites are identified when comparing multiple family members [98–100]. Calcium binding is controlled by a series of loops and water molecules at one end of the β-sandwich structure. The C2A domain of Syt1 binds three Ca2+ ions with Kd of 54, 530 μM, and a very low affinity of >20 mM. Its C2B domain binds two Ca2+ ions with Kd of 300–400 and 500–600 μM [100–102]. However, the intrinsic Ca2+ affinities of the Syt proteins are distinct in different isoforms. For instance, Syt7 has the highest Ca2+ sensitivity, having a half-maximum binding of Ca2+ at 1.5 μM for the C2A domain and 2.5 μM for the C2B domain [103]. Ca2+-binding does not cause a significant conformational change in the C2 domains, rather it modifies the electrostatic surface potential that allows the C2 domains to interact with phospholipids and intensifies Ca2+ binding [91,99,104]. As a result, substitutions to ligating residues in Syt severely dampens both Ca2+ binding and phospholipid interactions.

The most comprehensively studied Syt isoform is the neuronal Syt1, shown to be the principal Ca2+ sensor that triggers neurotransmitter release through exocytosis [105,106]. Syt1 functions with the Soluble N-ethylmaleimide-sensitive factor Attachment protein Receptor (SNARE) proteins synaptobrevin 2, SNAP-25 (25 kDa synaptosomal-associated protein) and syntaxin 1 to trigger membrane fusion between a synaptic vesicle and the plasma membrane [107]. SNARE proteins are anchored to the vesicle (synaptobrevin 2) or plasma (SNAP-25, syntaxin 1) membrane via either a transmembrane domain or post-translation modification (Figure 3A). Although multiple models have been proposed [108–111], one recent model based on X-ray structures indicates assembly of the SNARE proteins into a partial four-helix bundle structure may be important to understand the exocytosis model for the repair progress [112,113]. This remains to be validated.

Model of SNARE assembly and membrane fusion

Figure 3
Model of SNARE assembly and membrane fusion

(A) Locations of the SNARE and Syt1 near the plasma membrane vesicle membrane interface in the absence of calcium influx. The SNARE proteins syntaxin1 (cyan) and SNAP-25 (green) localize on the plasma membrane while synaptobrevin 2 (red) and Syt1 (gray) localize to the vesicle membrane. The C2A and C2B domains of Syt1 are not bound to Ca2+. (B) In the prefusion state, Syt1 interacts with the partially assembled SNAREs to form an inhibitory complex in a Ca2+-independent manner. (C) Ca2+ channel opening causes Ca2+ influx, resulting in Ca2+ binding to the Syt1 C2 domains. (D) The Ca2+-bound C2 domains intercalate into the neighboring plasma or vesicle phospholipid membrane releasing the SNARE complex.

Figure 3
Model of SNARE assembly and membrane fusion

(A) Locations of the SNARE and Syt1 near the plasma membrane vesicle membrane interface in the absence of calcium influx. The SNARE proteins syntaxin1 (cyan) and SNAP-25 (green) localize on the plasma membrane while synaptobrevin 2 (red) and Syt1 (gray) localize to the vesicle membrane. The C2A and C2B domains of Syt1 are not bound to Ca2+. (B) In the prefusion state, Syt1 interacts with the partially assembled SNAREs to form an inhibitory complex in a Ca2+-independent manner. (C) Ca2+ channel opening causes Ca2+ influx, resulting in Ca2+ binding to the Syt1 C2 domains. (D) The Ca2+-bound C2 domains intercalate into the neighboring plasma or vesicle phospholipid membrane releasing the SNARE complex.

Close modal

Evidence suggests that in the absence of Ca2+, a prefusion complex is formed between the three SNARE proteins, the cytosolic protein complexin and Syt1 that brings the vesicle and plasma membrane surfaces into close proximity (<4 nm, Figure 3B) [112,114,115]. This allows the SNARE proteins to form a partial four-helix bundle in complex with two molecules of Syt1 that reside on a synaptic vesicle. One molecule of Syt1 interacts with the syntaxin/synaptobrevin/complexin components while a second molecule interacts on the opposite side of the SNARE complex with SNAP-25. In both cases, the C2B domain appears to be the major interacting region with the SNARE complex. Upon calcium entry, and Ca2+ binding to Syt1 (Figure 3C), the Syt1 interaction with synaptobrevin/complexin is released and the Ca2+-filled C2 domains intercalate with either the plasma or vesicle membranes that allows the SNARE proteins to ‘fully-zipper’ and facilitate fusion of the two membranes (Figure 3D) [111]. The sensing role of Syt1 is critical for full SNARE assembly and release of inhibition since deletion of Syt1 or substitutions in the C2B domain impair neurotransmitter release [116]. The energy released from the full assembly of the SNAREs, and neutralization of the acidic Ca2+-binding loops in Syt1 appear to be the driving force to overcome fusion of the similarly negative-charged vesicle and plasma membranes [115,117,118]. Loss of Syt1 causes a reduction in membrane integrity that is magnified under stress conditions such as salt and freezing in plants [119,120].

Syt7, widely expressed in neuronal and non-neuronal tissues, is proposed to have a calcium-sensing role in membrane repair. Syt7 is located on the membrane surfaces of lysosomes and nonsynaptic secretory granules. Early work showed that Syt7 regulates membrane repair in fibroblasts via calcium-sensitive lysosome exocytosis [121]. Knockdown of Syt7 results in a reduction in exocytosis of pancreatic β-cells’ secretory granules [122] and defects in lysosomal exocytosis and membrane repair are noted in Syt7-deficient mice [34]. Syt7 shares some features with Syt1 and mechanistic features for possible membrane repair by Syt7 can be gleaned from the role of Syt1 in synaptic vesicle fusion. For example, in vitro assays show Ca2+ binding promotes SNARE-mediated membrane fusion. Syt7 exhibits strong Ca2+-sensitive binding to phospholipids and the plasma membrane SNARE proteins syntaxin 1A and SNAP-25 [123]. Furthermore, lysosomal exocytosis studies show interactions between Syt7 and SNAP-23 are facilitated by syntaxin 4, but not by syntaxins 2, 3, or 6 [124]. Pulldown assays detect interactions between TI-VAMP/VAMP7 and SNAP-23 with the C2A domain of Syt7 under 1 mM Ca2+ but not with the Syt1 C2A domain [124]. Botulinum neurotoxin E cleavage of SNAP-23 severely inhibits lysosomal exocytosis, indicating that interactions of Syt7 with some isoforms of SNARE proteins may play a key role in membrane resealing [124].

Membrane patching by dysferlin

The membrane repair protein dysferlin is a large, tail-anchored transmembrane protein in the Ferlin family of vesicle fusion proteins. It is expressed in many tissues [76] but is particularly abundant in skeletal and cardiac muscle where dysferlin is localized to cytoplasmic vesicles and the sarcolemma. Dysferlin-deficient mice display defective Ca2+-dependent sarcolemma repair and develop forms of slow, progressive MD [125–127]. Consistent with its role in membrane repair, mutations in the dysferlin gene are linked to two diseases characterized by muscle weakness, Miyoshi myopathy and LGMD type 2B (LGMD2B) [51,52]. Furthermore, immunolabeling experiments of dystrophic skeletal muscle display extreme reductions of dysferlin at the sarcolemma [128]. Dysferlin-null myofibers initially survive contraction-induced injury but eventually succumb to inflammation and necrosis in the absence of membrane repair. On the other hand, wild-type muscle cells can survive large-strain injury by sarcolemma membrane repair [126]. Reintroduction of the dysferlin gene into dysferlin-null mice promotes rapid recovery of skeletal muscles from contraction-induced injury [129]. Dysferlin-deficient cells are more prone to injury following mechanical membrane stress [130,131]. These findings suggest that dysferlin is vital in mediating the recovery of membrane rupture and maintaining membrane integrity in muscle fiber cells.

Dysferlin comprises seven C2 Ca2+-binding domains (C2A–C2G) and two accessory domains (C2-FerA, DysF) anchored to vesicle surfaces and the sarcolemma in muscle cells by a C-terminal transmembrane domain. In vitro studies show the C2 domains bind one to two Ca2+ ions and display a wide range of Ca2+-binding affinities [132]. For instance, dissociation constants (Kd) between 1 and 3 μM have been measured for individual sites in the C2A and C2C domains, whereas sites in the C2D and C2E domains have affinities well above 500 μM [132]. Protein–lipid overlay assays show that the dysferlin C2 domains have diverse affinities and Ca2+ dependencies for different lipid types. All C2 domains of dysferlin bind phosphatidylserine (PS) lipids in a Ca2+-independent fashion that is enhanced in four of the C2 domains (C2A, C2E, C2F, and C2G) upon Ca2+ binding [132]. The C2A domain exhibits the highest Ca2+-induced association with PS and phosphoinositide [133]. High-resolution structures are available for the C2A domain [134,135] and the DysF domain [136,137] from dysferlin. Like the Syt proteins, structures show that the C2A domain forms a β-sandwich structure with the Ca2+ loops at one end of the fold [134,135]. Missense mutations in each of the C2 domains of the dysferlin gene cause rare forms of MD, implying that altering a single C2 domain is sufficient to disrupt its membrane repair function [138]. This could arise from defects in domain folding, association with binding partners, improper lipid interactions or alterations in the overall tertiary structure [138,139]. Ca2+ binding to the C2A domain does not induce large conformational changes but rigidifies the Ca2+-binding loops important for lipid binding [135]. Structures of the remaining C2 domains are not available but have been predicted by RoseTTAFold and AlphaFold2 programs. Early predictions suggested that dysferlin adopts a flexible ‘beads on a string’ tertiary structure with the C2A domain furthest from the membrane surface [140,141]. More recently, modeling algorithms predict a more compact arrangement where all the C2 domains are packed together, along with the C2-FerA and DysF domains [142].

Dysferlin is proposed to direct calcium-sensitive membrane patching of the sarcolemma in skeletal and cardiac muscle cells (Figure 2C) [143]. This process is triggered by a rapid Ca2+influx into cells causing a high Ca2+-concentration zone around the rupture site. Vesicles recruited to the injury site fuse with one another forming a patch that anneals with the plasma membrane to reseal the rupture sites (Figure 4). Live-cell imaging using mechanically wounded rat myocytes shows dysferlin-containing vesicles transport along microtubules toward the injury site [144]. It is unclear what vesicles are involved, yet these undergo rapid fusion to form larger vesicles in the cytoplasm rather than directly accumulating at the rupture sites. The proposed role of Ca2+ is to trigger formation of large dysferlin-containing vesicles that act as a ‘patch’ to repair large lesions [144]. Since dysferlin contains a transmembrane domain anchored to phospholipids in both vesicle and plasma membranes, it is thought to co-ordinate vesicle docking and fusion with help from cytosolic protein-binding partners that do not contain membrane-spanning regions such as S100A10, Annexin A2 (ANX2), AHNAK, and TRIM72 (MG53) [82,145,146]. Supporting this, multiple large-scale proteomics experiments have identified interactions between dysferlin and over 100 other proteins, including many implicated in membrane repair [145,147,148]. Fluorescence-imaging experiments using mouse myoblasts show possible associations between dysferlin, TRIM72, and caveolin-3 (Cav3) that facilitates intracellular vesicle trafficking during acute damage membrane repair [82]. Furthermore, scrape injury experiments show that TRIM72 colocalizes and accumulates with dysferlin at the injury site, forming a lattice-like network near the lesion edges [68,82]. Substitutions in Cav3 influence dysferlin trafficking with TRIM72, in which significant amounts of TRIM72 and dysferlin are redistributed to the Golgi apparatus suggesting functional interactions between TRIM72 and dysferlin that is mediated by Cav3 [82,149]. Substitutions in Cav3 or TRIM72 knockout experiments display abnormal vesicular trafficking in mouse myoblasts, whereas dysferlin null experiments show normal vesicle accumulation near the sarcolemma, but defective membrane repair [76,82]. This suggests that dysferlin participates in the resealing process that requires protein partners to facilitate vesicle recruitment to injury sites [82]. Supporting this idea, pulldown assays show association between the dysferlin C2A domain and TRIM72 in the absence and presence of Ca2+-free environment, that is dependent on the TRIM72 oligomerization state in an oxidation-dependent manner [150]. Substitutions in the dysferlin domain (W52R and V67D) that are causative for muscle disease also alter its Ca2+ association with TRIM72 [150].

Model of dysferlin-mediated vesicle fusion membrane repair

Figure 4
Model of dysferlin-mediated vesicle fusion membrane repair

(A) Dysferlin associates with plasma and vesicle membranes. Other proteins including TRIM72, AHNAK, and S100A10:ANX2 are mostly cytosolic. (B) Membrane rupture causes a local Ca2+ influx, binding to and activating Ca2+-sensitive proteins such as dysferlin and ANX2. This enables multiple interactions between the proteins and causes recruitment of TRIM72 and AHNAK by dysferlin and S100A10:ANX2 to vesicle surfaces. (C) Ca2+-dependent protein–protein interactions recruit vesicles such as enlargosomes to the injured plasma membrane facilitated by the complete dysferlin complex that spans the vesicle-plasma membrane space. (D) Multiple vesicles containing dysferlin and interacting proteins accumulate at the injured site forming a patch across the membrane lesion.

Figure 4
Model of dysferlin-mediated vesicle fusion membrane repair

(A) Dysferlin associates with plasma and vesicle membranes. Other proteins including TRIM72, AHNAK, and S100A10:ANX2 are mostly cytosolic. (B) Membrane rupture causes a local Ca2+ influx, binding to and activating Ca2+-sensitive proteins such as dysferlin and ANX2. This enables multiple interactions between the proteins and causes recruitment of TRIM72 and AHNAK by dysferlin and S100A10:ANX2 to vesicle surfaces. (C) Ca2+-dependent protein–protein interactions recruit vesicles such as enlargosomes to the injured plasma membrane facilitated by the complete dysferlin complex that spans the vesicle-plasma membrane space. (D) Multiple vesicles containing dysferlin and interacting proteins accumulate at the injured site forming a patch across the membrane lesion.

Close modal

Amongst the many dysferlin partners is the giant muscle-cell protein AHNAK, shown to migrate to a membrane injury site under cellular stress. AHHAK is a cytosolic protein that associates with vesicles (enlargosomes) involved in exocytosis [151] in a Ca2+-dependent manner. The absence of a transmembrane domain in AHNAK indicates its localization to a membrane surface is controlled by interactions with other partner proteins. Immunostaining of skeletal muscle shows colocalization of dysferlin and AHNAK at the sarcolemma of skeletal muscle [146]. Furthermore, coimmunoprecipitation coupled to mass spectrometry experiments shows interactions between dysferlin and AHNAK [146]. GST pulldown experiments indicate that the dysferlin C2A domain is most important for the recruitment of AHNAK [146]. This interaction appears to be calcium insensitive and dependent on the C-terminus of AHNAK. Proteomic experiments also identify the EF-hand protein S100A10 and phospholipid-binding protein ANX2 as members of the dysferlin complex [145,147]. Dysferlin and ANX2 co-accumulate at injury sites of the plasma membrane in a Ca2+-sensitive fashion [152].

ANX2 forms a tight complex with S100A10 and the two proteins are frequently coisolated as a S100A10:ANX2 heterotetramer [153,154]. Yeast three hybrid experiments show S100A10:ANX2 interacts with a C-terminal region from AHNAK different from that used by dysferlin [155]. Furthermore, structural experiments show that AHNAK asymmetrically spans the complete heterotetramer utilizing two both S100A10 and ANX2 protomers [156–158]. These interactions paint a picture where AHNAK acts as a scaffold to recruit both S100A10:ANX2 and dysferlin [146,156]. However, the precise arrangement of these proteins, along with other candidate repair proteins such as TRIM72 and Cav-3, is unknown as are the mechanisms used for the repair process.

Cysteine protease calpain enzymes sense membrane ruptures

Calpains are a large family of Ca2+-sensitive cysteine proteases ubiquitously expressed as inactive proenzymes in most cell types [159]. Rather than being a degradative protease, calpains modify their substrate by cleavage and often confer a Ca2+-signaling function. Two of the best studied enzymes, calpain-1 and calpain-2 are composed of a large 80 kDa catalytic subunit (CAPN1, CAPN2) with two protease core domains (PC1, PC2) that bind two Ca2+ ions and a small 30 kDa noncatalytic subunit (CAPNS1) [160–163]. Each subunit also contains a penta EF-hand domain that binds four additional Ca2+ ions. The cleavage activities of calpains are closely correlated to extracellular Ca2+ levels and typically require high Ca2+ levels such as those that occur during plasma membrane rupture. For example, calpain-1 and calpain-2 require between 5–50 μM or 200–800 μM Ca2+ concentrations, respectively, to reach half-maximum activity [164]. Calpain-3 deficiency results in LGMD-recessive type 1 (LGMDR1) but has little effect on the survival of laser-wounded skeletal myotubes [165]. Decreased expression of calpain-1 and calpain-2 in mouse embryonic fibroblasts shows loss of sarcolemma-resealing activity, suggesting calpain-1 and -2 are required for calcium-dependent membrane repair [165]. Moreover, knockout of CAPNS1 inhibits Ca2+-dependent resealing of membrane ruptures. However, knockout of either calpain-1 or calpain-2 has less effect on membrane repair, suggesting redundancy for protein cleavage events during membrane repair [165]. Consistent with a role in the repair of larger lesions, loss of both calpains decreases cell survival following mechanical membrane injury but is not impacted by small ruptures formed by SLO and saponin [166]. High-resolution microscopy shows that calpain inhibition prevents resealing of crayfish medial giant axons but does not prevent vesicle accumulation [167]. These findings indicate that calpains are required for Ca2+-dependent survival after membrane injury [164] likely due to enhancing membrane fusion rather than vesical formation and protein recruitment at the injury site [167].

Calpains-1 and -2 proteolyze dysferlin between the C2E and C2F domains yielding a membrane bound C-terminal fragment [168]. Proteolysis requires extracellular Ca2+ provided by membrane rupture [68]. Interestingly, a genetic deletion in the dysferlin gene has also been identified that yields a similar C-terminal protein fragment (‘mini-dysferlin’; includes C2F, C2G, and transmembrane domains) [169]. This dysferlin fragment resembles inverted Syt1, having its transmembrane anchor at the C-terminus rather than the N-terminus. When truncated dysferlin is expressed in dysferlin-null mice, isolated mice muscle cells show restored membrane repair function [169]. Furthermore, immunolabeling experiments show that the C-terminal fragment of dysferlin is detected at sites of membrane injury in human myoblasts, rather than the N-terminus [68,168]. Other experiments reveal that TRIM72 and mini-dysferlin-containing vesicles cluster at injury sites within 10 s following injury [68,170]. Mini-dysferlin and TRIM72 aggregate with exposed phospholipids to encircle the injury site, thus stabilizing the plasma membrane [68]. This mechanism appears to be most prominent in lung, liver, pancreas, and placenta (40–60%) rather than cardiac or skeletal muscle cells (10–15%) where low abundance the mini-dysferlin protein is detected [168]. Calpain-3 is also proposed to work as a regulatory enzyme for the dysferlin complex [171]. Immunostaining in cell cultures suggests cleavage of AHNAK and the muscle specific-protein titin is regulated by calpain-3 perhaps, suggesting a role following membrane repair linked to cytoskeletal maintenance [166,172].

Annexin complexes seal a membrane

The annexins are a family of Ca2+-regulated phospholipid-binding proteins, with Annexins A1, A2, A4, A5, and A6 most known for their roles in membrane repair. Annexins include four to eight core domains that bind Ca2+ arranged in a right-handed superhelix and a unique, variable length N-terminal domain [173]. Under resting Ca2+ concentrations, most annexins diffusely distribute in their Ca2+-free, inactivated state in the cytosol. In response to a Ca2+ influx from membrane injury, Ca2+ binding to the outer surface of the annexin protein facilitates interaction with the acidic phospholipid membrane. Calcium-binding sites in the annexins are located in the outer loop regions of the protein and display different calcium co-ordination patterns from EF-hand calcium-binding motifs [174,175]. Although the core domains are highly conserved in the annexin family, their relative orientations and numbers of bound Ca2+ ions differ amongst isoforms [175]. Crystal structures of Annexins A1 and A5 show six and four Ca2+-binding sites, respectively [176,178]. It is thought that the range of calcium sensitivities by the annexins may be correlated with the size of the membrane hole. Furthermore, Ca2+ sensitivities for individual annexin proteins vary in the presence of phospholipids. For example, ANX1 and ANX2 have calcium Kd values of >1000 and 100–500 μM, respectively, without phospholipid [179–181,228]. In contrast, their affinities for Ca2+ increase in 10- to 100-fold with phospholipid present [177,179,228,229]. The higher calcium sensitivity of ANX2 may be due to its association with S100A10 that is unable to bind calcium (see below) [182]. The variable N-terminal domains of annexins are thought to provide specificity for different binding partners and in most cases these interactions are controlled by Ca2+ binding to the annexin protein. For instance, the N-terminal domain of ANX1 is buried within the core domains in the absence of Ca2+ and is released upon Ca2+ binding due to rearrangement of the core domains [183].

The relatively weak affinities and preference for phospholipid membranes make members of the annexin family perfect candidates for membrane repair. Furthermore, the variable N-termini provides a platform for binding partners found in different membrane-resealing pathways. Frequent binding partners for the annexin proteins are the S100 proteins, a family of dimeric, EF-hand proteins. For instance, the ANX1 and ANX2 co-ordinate with S100A11 and S100A10, respectively [157,158,184,185]. Most interactions of annexins with S100 proteins are also controlled by Ca2+ binding to the S100 protein, which typically bind two Ca2+ ions per protomer and have Ca2+ dissociation constants ranging from 10 to 500 μM for the two Ca2+-binding sites [186–188]. Calcium binding to most S100 proteins causes a significant conformational change that provides a hydrophobic surface used to recruit the N-terminal region from an annexin protein [184]. Numerous three-dimensional structures show the annexin N-terminus forms an amphipathic helix that bridges the S100 dimer [184,185,188]. Although annexins are found as monomers in the cytosol, binding to a dimeric S100 protein results in the formation of a noncovalent heterotetramer. One outlier in this mechanism is S100A10 that recruits ANX2 in the absence of Ca2+. S100A10 is a unique member of the S100 family that has lost its ability to bind Ca2+ yet retains the Ca2+-activated structure exhibited by other Ca2+-bound S100 proteins [154,190]. The S100A10:ANX2 complex was originally isolated as a Ca2+-insensitive p11p36 heterotetramer from intestinal epithelium cells [154]. Although other S100 proteins such as S100A4 and S100A11 can recruit ANX2 [191–195], S100A10 has been shown to bind to this annexin protein nearly three-orders of magnitude stronger. The S100A10:ANX2 complex is likely the predominant form used in membrane repair. In the absence of ANX2, S100A10 is rapidly degraded in cells [196]. In addition, it has been observed that oxidative stress results in glutathionylation of ANX2 in the S100A10:ANX2 heterotetramer, resulting in a loss of its ability to interact with phospholipids and F-actin, suggesting ANX2 is sensitive to oxidation that alters its conformation and function [197].

ANX1 and ANX2 aggregate with intracellular vesicles by interacting with dysferlin in a Ca2+-dependent manner. Both ANX1 and ANX2 colocalize with dysferlin at the sarcolemma [53]. This association is disrupted in muscle cells lacking dysferlin [53], suggesting an interaction between dysferlin and annexins may be essential for vesical aggregation and fusion during membrane repair. ANX2 is known to stabilize and organize lipid microdomains and is proposed to regulate exocytosis events in chromaffin cells [198]. The S100A10:ANX2 complex also localizes at membrane injury sites [156]. Both ANX1 and ANX2 translocate to the membrane in damaged human and zebrafish muscle cells, although at different rates that may depend on localization prior to a rupture [199,200]. Annexin knockout experiments display poor myofiber repair or muscle regeneration abilities [201–203]. A Ca2+-insensitive version of ANX1 also inhibits membrane sealing [204]. Association between dysferlin and the S100A10:ANX2 complex is Ca2+ and injury dependent. Furthermore, a proposed model depicts that membrane-patching repair requires the interaction between dysferlin and ANX1 and ANX2 [205]. Abnormal localization of ANX1 and ANX2 has been found in dysferlin-deficient mouse muscle cells [53]. Interestingly, fluorescence lifetime imaging microscopy (FLIM) experiments demonstrate significant interactions between dysferlin and ANX1 during resting physiological Ca2+ levels, and this interaction is lost in injured cells [53]. On the other hand, interactions between dysferlin and ANX2 are observed in resting and ruptured mice muscle cells, suggesting that each annexin protein may have a specific role in repair processes [53,206]. The recruitment of Ca2+-activated ANX1 to a damaged region seems independent of S100A11. Immunolocalization assays show ANX1 concentrates around the rupture site in human breast cancer cells lacking S100A11 expression [205]. Interestingly, ANX1-deficient mice show little effect in Ca2+-dependent myofiber repair compared with wild-type myofibers. Instead, increasing numbers of unfused differentiating myoblast suggest that the cell–cell fusion stage of myofiber formation was affected by the absence of ANX1, implying that ANX1 potentially participates in regulating membrane fusion, and other possible proteins may substitute its role in membrane repair [201].

Other annexin proteins are proposed to be important in membrane repair. Annexin A5 (ANX5) is rapidly recruited to a membrane injury site and binds to the curved edges of a damaged membrane. ANX5 self-assembles into two-dimensional (2D) protein arrays on PS-containing membranes, possibly stabilizing the membrane, and limiting wound expansion during membrane repair [207–209]. ANX5-null cells exhibit defective membrane repair that is reversed upon addition of recombinant ANX5 [210]. In contrast, an ANX5 variant that is unable to assemble into arrays, fails to rescue membrane repair ability.

Annexin A4 (ANX4) and A6 (ANX6) appear to work together to induce membrane curvature and wound closure. The smallest annexin member, ANX4, self-assembles into trimers on a membrane surface to induce curvature at rupture edges [211,212], allowing ANX6 to constrict and close of the wound [212]. Truncated ANX6 competes with and prevents translocation of wild-type ANX6 to alter assembly of the annexin repair complex [213]. Conversely, overexpression of ANX6 promotes membrane repair by increasing the formation of external vesicles at the site of membrane injury [213]. This repair strategy seems to depend on cell type and wound size, in which smaller lesions in metastatic cancer cells require actin remodeling and membrane restoration by the S100A10:ANX2 complex. On the other hand, larger wounds in metastatic cancer cells involve S100A10:ANX2 coupled to ANX4 and ANX6 for repair, suggesting diverse membrane repair mechanisms may be used to maintain membrane integrity [212].

ALG-2 and ESCRT-mediated repair

ALG-2 (also PDCD6) is a highly conserved Ca2+-binding protein in the penta-EF-hand (PEF) family. Although originally thought to be a proapoptotic signaling protein, ALG-2-deficient mice display little phenotypic changes to support this function [214]. Instead, ALG-2 displays an array of interacting protein partners that support a role in cell proliferation, vesicle transport, and membrane repair in many cell types [215–217]. ALG-2 includes a PEF domain characterized by utilizing the fifth EF hand for Ca2+-independent dimer formation [218,219]. The first (EF-1) and third (EF-3) EF hands in ALG-2 possess the two highest-affinity Ca2+-binding sites with Kd of 1.2 and 300 μM, while EF-5 displays very weak Ca2+ binding. Ca2+ binding to EF-1 and EF-3 induces conformational changes that open a hydrophobic pocket near the N-terminus of ALG-2 for target recognition and binding, suggesting ALG-2 is a calcium sensor [218,219]. As a result, substitutions to the Ca2+ co-ordinating residues in EF-1 and EF-3 impairs Ca2+-induced conformational changes that may directly alter ALG-2 function [220]. Consistent with this, immunofluorescence assays show ALG-2 localizes to the cytoplasm and nucleus of cells [221] but moves to the membrane fraction in the presence of Ca2+ [222].

A variety of ALG-2-interacting proteins have been identified that associate with vesicle and plasma membranes, including Annexins A7 (ANX7) and A11, ALG-2-interacting protein X (ALIX), and components of the ESCRT complex [218,227]. Membrane-wounding assays show Ca2+-bound ANX7 nucleates at ruptured membrane edges along with ALG-2 following scrape injury, suggesting that ANX7 regulates the proper localization of ALG-2 to the injured membrane [227]. ESCRT proteins are organized into five functionally distinct complexes, shown to facilitate the repair of narrow membrane wounds (<100 nm) observed in PFT-injured cells [71]. Laser-induced membrane damage experiments show that ESCRTs facilitate cleavage of damaged regions of the cell membrane using ALG-2 and ALIX [223]. Particularly, ESCRT-III induces membrane deformation and scission to shed extracellular vesicles to remove damaged membranes [71,223,224]. The essential regulator protein ALIX has an important role in the recruitment of ESCRT-III components including CHMP4 [216,223,225,226]. On the other hand, knockdown of ALG-2 inhibits the recruitment of ALIX at the injury site but not vice versa [223]. Thus, Scheffer et al. proposed a model whereby ESCRT-mediated membrane repair requires the use of Ca2+-activated ALG-2 to initiate the sequential recruitment of ALIX and ESCRT proteins to the lesion for membrane repair (Figure 5) [223,228].

Externalization repair via ESCRT-mediated shedding to remove bacterial PFTs

Figure 5
Externalization repair via ESCRT-mediated shedding to remove bacterial PFTs

(A) PFTs induce stable transmembrane protein-based lesions, which cannot simply repair by spontaneous membrane reseal. Ca2+ activates ANX7 and PEF protein ALG-2. ANX7 associate with ALG-2, and Ca2+-activated ALG-2 associate with ALIX. (B) ANX7, ALG-2, and ALIX form a complex that migrates to the plasma membrane. (C) ALG-2 complex initiates the assembly of ESCRT-III and Vps4 at the membrane injury. (D) Sequential membrane remodeling and membrane scission by ESCRT pathway. Other ESCRT distinct complexes involve in the recruitment and membrane repair process are not shown here.

Figure 5
Externalization repair via ESCRT-mediated shedding to remove bacterial PFTs

(A) PFTs induce stable transmembrane protein-based lesions, which cannot simply repair by spontaneous membrane reseal. Ca2+ activates ANX7 and PEF protein ALG-2. ANX7 associate with ALG-2, and Ca2+-activated ALG-2 associate with ALIX. (B) ANX7, ALG-2, and ALIX form a complex that migrates to the plasma membrane. (C) ALG-2 complex initiates the assembly of ESCRT-III and Vps4 at the membrane injury. (D) Sequential membrane remodeling and membrane scission by ESCRT pathway. Other ESCRT distinct complexes involve in the recruitment and membrane repair process are not shown here.

Close modal

Recent studies used cell viability following electroporation membrane lesions as a measure for membrane repair. An ALG-2 variant lacking the ability to bind Ca2+ due to substitutions in the EF-1 and EF-3 domains shows lower cell survival compared with controls [217]. On the other hand, overexpression of wild-type ALG-2 increases cell viability of lesion-damaged cells [217]. Furthermore, cell viability decreases when the ALIX-binding site on ALG-2 is blocked, supporting the importance of an ALG-2/ALIX interaction for the repair model proposed by Scheffer [217,223].

Plasma membrane repair is an emergency response system, essential for maintaining cellular homeostasis and ensuring cell survival. The rapid influx of extracellular Ca2+ upon membrane rupture is one of the primary signals to initiate the repair process. Recruitment of multiple Ca2+-dependent sensor proteins has been identified that display multiple membrane repair mechanisms. It is likely that these protein sensors co-ordinate a diverse set of repair responses tailored to the location, size, and nature of the membrane injury. Defective membrane repair, frequently due to mutations in genes that code for repair proteins, is now recognized as a contributor to different forms of MD muscle disease and heart disease. Considering all known protein components work together and have distinct roles in membrane repair it is likely that no single mechanism, nor therapeutic solution, applies in all injury scenarios. Although some three-dimensional structure information is available for some individual proteins the atomic level details of the interactions between nearly all repair proteins are missing. This information is urgently needed as the foundation for possible pharmaceutical therapies. In addition, it is likely that other possible mediators and physiological conditions exist that control membrane repair mechanisms that have yet to be resolved. Many of these aspects should be identified to gain a full understanding of the membrane repair machinery and provide new avenues for future treatments of inherited diseases.

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

This research was supported by a research [grant number G-20-0028739] from the Heart and Stroke Foundation of Canada (GSS).

ZiWei Li: Conceptualization, Data curation, Writing—original draft, Writing—review & editing. Gary S. Shaw: Conceptualization, Supervision, Funding acquisition, Writing—original draft, Project administration, Writing—review & editing.

2D

two-dimensional

ALG-2

apoptosis-linked gene 2

ALIX

ALG-2-interacting protein X

ANX1

annexin A1

ANX2

annexin A2

ANX4

annexin A4

ANX5

annexin A5

ANX6

annexin A6

Ca2+

calcium

CRISPR

clustered regularly interspaced short palindromic repeats

ESCRT

endosomal sorting complex required for transport

FLIM

fluorescence lifetime imaging microscopy

Kd

dissociation constant

LGMD2B

limb-girdle muscular dystrophy type 2B

LGMDR1

limb-girdle muscular dystrophy recessive type 1

MD

muscular dystrophy

PEF

penta-EF-hand

PFT

pore-forming toxin

PS

phosphatidylserine

SNAP-25

25 kDa synaptosomal-associated protein

SNARE

Soluble N-ethylmaleimide sensitive factor Attachment protein Receptor

Syt

synaptotagmin

Syt7

synaptotagmin VII

1.
McNeil
P.L.
and
Khakee
R.
(
1992
)
Disruptions of muscle fiber plasma membranes. Role in exercise-induced damage
.
Am. J. Pathol.
140
,
1097
1109
[PubMed]
2.
Howard
A.C.
,
McNeil
A.K.
and
McNeil
P.L.
(
2011
)
Promotion of plasma membrane repair by vitamin E
.
Nat. Commun.
2
,
597
[PubMed]
3.
Forrester
S.J.
,
Kikuchi
D.S.
,
Hernandes
M.S.
,
Xu
Q.
and
Griendling
K.K.
(
2018
)
Reactive oxygen species in metabolic and inflammatory signaling
.
Circ. Res.
122
,
877
902
[PubMed]
4.
Su
L.J.
,
Zhang
J.H.
,
Gomez
H.
,
Murugan
R.
,
Hong
X.
,
Xu
D.
et al.
(
2019
)
Reactive oxygen species-induced lipid peroxidation in apoptosis, autophagy, and ferroptosis
.
Oxidative Med. Cell. Longevity
2019
,
5080843
5.
Zorov
D.B.
,
Juhaszova
M.
and
Sollott
S.J.
(
2006
)
Mitochondrial ROS-induced ROS release: an update and review
.
Biochim. Biophys. Acta
1757
,
509
517
[PubMed]
6.
Anura
A.
(
2014
)
Traumatic oral mucosal lesions: a mini review and clinical update
.
Oral Health Dent. Manag.
13
,
254
259
7.
Khowal
S.
and
Wajid
S.
(
2019
)
Role of smoking-mediated molecular events in the genesis of oral cancers
.
Toxicol. Mech. Methods
29
,
665
685
[PubMed]
8.
Cobb
J.P.
,
Hotchkiss
R.S.
,
Karl
I.E.
and
Buchman
T.G.
(
1996
)
Mechanisms of cell injury and death
.
Br. J. Anaesth.
77
,
3
10
[PubMed]
9.
Maiti
P.
,
Manna
J.
and
Dunbar
G.L.
(
2017
)
Current understanding of the molecular mechanisms in Parkinson's disease: Targets for potential treatments
.
Transl. Neurodegener.
6
,
28
10.
Agnihotri
A.
and
Aruoma
O.I.
(
2020
)
Alzheimer's Disease and Parkinson's Disease: a nutritional toxicology perspective of the impact of oxidative stress, mitochondrial dysfunction, nutrigenomics and environmental chemicals
.
J. Am. Coll. Nutr.
39
,
16
27
[PubMed]
11.
Dias
C.
and
Nylandsted
J.
(
2021
)
Plasma membrane integrity in health and disease: significance and therapeutic potential
.
Cell Discov.
7
,
4
[PubMed]
12.
Cooper
S.T.
and
McNeil
P.L.
(
2015
)
Membrane repair: mechanisms and pathophysiology
.
Physiol. Rev.
95
,
1205
1240
[PubMed]
13.
Kitmitto
A.
,
Baudoin
F.
and
Cartwright
E.J.
(
2019
)
Cardiomyocyte damage control in heart failure and the role of the sarcolemma
.
J. Muscle Res. Cell Motil.
40
,
319
333
[PubMed]
14.
Stow
J.L.
and
Condon
N.D.
(
2016
)
The cell surface environment for pathogen recognition and entry
.
Clin. Transl. Immunol.
5
,
e71
15.
Dal Peraro
M.
and
van der Goot
F.G.
(
2016
)
Pore-forming toxins: ancient, but never really out of fashion
.
Nat. Rev. Microbiol.
14
,
77
92
[PubMed]
16.
Coburn
B.
,
Sekirov
I.
and
Finlay
B.B.
(
2007
)
Type III secretion systems and disease
.
Clin. Microbiol. Rev.
20
,
535
549
[PubMed]
17.
Tang
T.
,
Bidon
M.
,
Jaimes
J.A.
,
Whittaker
G.R.
and
Daniel
S.
(
2020
)
Coronavirus membrane fusion mechanism offers a potential target for antiviral development
.
Antiviral Res.
178
,
104792
[PubMed]
18.
McNeil
P.L.
and
Steinhardt
R.A.
(
1997
)
Loss, restoration, and maintenance of plasma membrane integrity
.
J. Cell Biol.
137
,
1
4
[PubMed]
19.
Ammendolia
D.A.
,
Bement
W.M.
and
Brumell
J.H.
(
2021
)
Plasma membrane integrity: implications for health and disease
.
BMC Biol.
19
,
71
[PubMed]
20.
Bi
G.Q.
,
Alderton
J.M.
and
Steinhardt
R.A.
(
1995
)
Calcium-regulated exocytosis is required for cell membrane resealing
.
J. Cell Biol.
131
,
1747
1758
[PubMed]
21.
Bement
W.M.
,
Mandato
C.A.
and
Kirsch
M.N.
(
1999
)
Wound-induced assembly and closure of an actomyosin purse string in Xenopus oocytes
.
Curr. Biol.
9
,
579
587
[PubMed]
22.
Zhen
Y.
,
Radulovic
M.
,
Vietri
M.
and
Stenmark
H.
(
2021
)
Sealing holes in cellular membranes
.
EMBO J.
40
,
e106922
[PubMed]
23.
Andrews
N.W.
,
Almeida
P.E.
and
Corrotte
M.
(
2014
)
Damage control: cellular mechanisms of plasma membrane repair
.
Trends Cell Biol.
24
,
734
742
[PubMed]
24.
Andrews
N.W.
and
Corrotte
M.
(
2018
)
Plasma membrane repair
.
Curr. Biol.
28
,
R392
R397
[PubMed]
25.
Draeger
A.
,
Schoenauer
R.
,
Atanassoff
A.P.
,
Wolfmeier
H.
and
Babiychuk
E.B.
(
2014
)
Dealing with damage: plasma membrane repair mechanisms
.
Biochimie
107
,
66
72
[PubMed]
26.
Blazek
A.D.
,
Paleo
B.J.
and
Weisleder
N.
(
2015
)
Plasma membrane repair: a central process for maintaining cellular homeostasis
.
Physiology
30
,
438
448
[PubMed]
27.
Jimenez
A.J.
and
Perez
F.
(
2017
)
Plasma membrane repair: the adaptable cell life-insurance
.
Curr. Opin. Cell Biol.
47
,
99
107
[PubMed]
28.
Lovering
R.M.
,
Porter
N.C.
and
Bloch
R.J.
(
2005
)
The muscular dystrophies: from genes to therapies
.
Phys. Ther.
85
,
1372
1388
[PubMed]
29.
Howell
G.J.
,
Holloway
Z.G.
,
Cobbold
C.
,
Monaco
A.P.
and
Ponnambalam
S.
(
2006
)
Cell biology of membrane trafficking in human disease
.
Int. Rev. Cytol.
252
,
1
69
[PubMed]
30.
Davies
K.E.
and
Nowak
K.J.
(
2006
)
Molecular mechanisms of muscular dystrophies: old and new players
.
Nat. Rev. Mol. Cell Biol.
7
,
762
773
[PubMed]
31.
Gazzerro
E.
,
Sotgia
F.
,
Bruno
C.
,
Lisanti
M.P.
and
Minetti
C.
(
2010
)
Caveolinopathies: from the biology of caveolin-3 to human diseases
.
Eur. J. Human Genet.
18
,
137
145
32.
Han
R.
(
2011
)
Muscle membrane repair and inflammatory attack in dysferlinopathy
.
Skelet. Muscle
1
,
10
[PubMed]
33.
McElhanon
K.E.
,
Young
N.
,
Hampton
J.
,
Paleo
B.J.
,
Kwiatkowski
T.A.
,
Beck
E.X.
et al.
(
2020
)
Autoantibodies targeting TRIM72 compromise membrane repair and contribute to inflammatory myopathy
.
J. Clin. Invest.
130
,
4440
4455
[PubMed]
34.
Chakrabarti
S.
,
Kobayashi
K.S.
,
Flavell
R.A.
,
Marks
C.B.
,
Miyake
K.
,
Liston
D.R.
et al.
(
2003
)
Impaired membrane resealing and autoimmune myositis in synaptotagmin VII-deficient mice
.
J. Cell Biol.
162
,
543
549
[PubMed]
35.
Koenig
M.
,
Monaco
A.P.
and
Kunkel
L.M.
(
1988
)
The complete sequence of dystrophin predicts a rod-shaped cytoskeletal protein
.
Cell
53
,
219
228
[PubMed]
36.
Gao
Q.Q.
and
McNally
E.M.
(
2015
)
The dystrophin complex: structure, function, and implications for therapy
.
Compr. Physiol.
5
,
1223
1239
37.
Kitmitto
A.
,
Baudoin
F.
and
Cartwright
E.J.
(
2019
)
Cardiomyocyte damage control in heart failure and the role of the sarcolemma
.
J. Muscle Res. Cell Motil.
40
,
319
333
[PubMed]
38.
Benevolensky
D.
,
Belikova
Y.
,
Mohammadzadeh
R.
,
Trouvé
P.
,
Marotte
F.
,
Russo-Marie
F.
et al.
(
2000
)
Expression and localization of the annexins II, V, and VI in myocardium from patients with end-stage heart failure
.
Lab. Invest.
80
,
123
133
[PubMed]
39.
Matteo
R.G.
and
Moravec
C.S.
(
2000
)
Immunolocalization of annexins IV, V and VI in the failing and non-failing human heart
.
Cardiovasc. Res.
45
,
961
970
[PubMed]
40.
Weisleder
N.
,
Takizawa
N.
,
Lin
P.
,
Wang
X.
,
Cao
C.
,
Zhang
Y.
et al.
(
2012
)
Recombinant MG53 protein modulates therapeutic cell membrane repair in treatment of muscular dystrophy
.
Sci. Transl. Med.
4
,
139ra85
[PubMed]
41.
He
B.
,
Tang
R.H.
,
Weisleder
N.
,
Xiao
B.
,
Yuan
Z.
,
Cai
C.
et al.
(
2012
)
Enhancing muscle membrane repair by gene delivery of MG53 ameliorates muscular dystrophy and heart failure in δ-Sarcoglycan-deficient hamsters
.
Mol. Ther.
20
,
727
735
42.
Wang
X.
,
Xie
W.
,
Zhang
Y.
,
Lin
P.
,
Han
L.
,
Han
P.
et al.
(
2010
)
Cardioprotection of ischemia/reperfusion injury by cholesterol-dependent MG53-mediated membrane repair
.
Circ. Res.
107
,
76
83
[PubMed]
43.
Liu
J.
,
Zhu
H.
,
Zheng
Y.
,
Xu
Z.
,
Li
L.
,
Tan
T.
et al.
(
2015
)
Cardioprotection of recombinant human MG53 protein in a porcine model of ischemia and reperfusion injury
.
J. Mol. Cell Cardiol.
80
,
10
19
[PubMed]
44.
Homsy
J.
,
Zaidi
S.
,
Shen
Y.
,
Ware
J.S.
,
Samocha
K.E.
,
Karczewski
K.J.
et al.
(
2015
)
De novo mutations in congenital heart disease with neurodevelopmental and other congenital anomalies
.
Science
350
,
1262
1266
[PubMed]
45.
Zhou
Y.
,
Qian
Z.
,
Yang
J.
,
Zhu
M.
,
Hou
X.
,
Wang
Y.
et al.
(
2018
)
Whole exome sequencing identifies novel candidate mutations in a Chinese family with left ventricular noncompaction
.
Mol. Med. Rep.
17
,
7325
7330
46.
Kim
S.C.
,
Kellett
T.
,
Wang
S.
,
Nishi
M.
,
Nagre
N.
,
Zhou
B.
et al.
(
2014
)
TRIM72 is required for effective repair of alveolar epithelial cell wounding
.
Am. J. Physiol. Lung Cell. Mol. Physiol.
307
,
L449
L459
47.
Jin
Y.
,
Lee
S.J.
,
Minshall
R.D.
and
Choi
A.M.
(
2011
)
Caveolin-1: a critical regulator of lung injury
.
Am. J. Physiol. Lung Cell. Mol. Physiol.
300
,
L151
L160
48.
Potter
R.A.
,
Griffin
D.A.
,
Sondergaard
P.C.
,
Johnson
R.W.
,
Pozsgai
E.R.
,
Heller
K.N.
et al.
(
2018
)
Systemic delivery of dysferlin overlap vectors provides long-term gene expression and functional improvement for dysferlinopathy
.
Hum. Gene Ther.
29
,
749
762
[PubMed]
49.
Lee
J.J.A.
,
Maruyama
R.
,
Duddy
W.
,
Sakurai
H.
and
Yokota
T.
(
2018
)
Identification of novel antisense-mediated exon skipping targets in DYSF for therapeutic treatment of dysferlinopathy
.
Mol. Ther. Nucleic Acids
13
,
596
604
[PubMed]
50.
Turan
S.
,
Farruggio
A.P.
,
Srifa
W.
,
Day
J.W.
and
Calos
M.P.
(
2016
)
Precise correction of disease mutations in induced pluripotent stem cells derived from patients with limb girdle muscular dystrophy
.
Mol. Ther.
24
,
685
696
51.
Liu
J.
,
Aoki
M.
,
Illa
I.
,
Wu
C.
,
Fardeau
M.
,
Angelini
C.
et al.
(
1998
)
Dysferlin, a novel skeletal muscle gene, is mutated in Miyoshi myopathy and limb girdle muscular dystrophy
.
Nat. Genet.
20
,
31
36
[PubMed]
52.
Nguyen
K.
,
Bassez
G.
,
Bernard
R.
,
Krahn
M.
,
Labelle
V.
,
Figarella-Branger
D.
et al.
(
2005
)
Dysferlin mutations in LGMD2B, Miyoshi myopathy, and atypical dysferlinopathies
.
Hum. Mutat.
26
,
165
[PubMed]
53.
Lennon
N.J.
,
Kho
A.
,
Bacskai
B.J.
,
Perlmutter
S.L.
,
Hyman
B.T.
and
Brown
R.H.
Jr
(
2003
)
Dysferlin interacts with annexins A1 and A2 and mediates sarcolemmal wound-healing
.
J. Biol. Chem.
278
,
50466
50473
[PubMed]
54.
Berridge
M.J.
,
Lipp
P.
and
Bootman
M.D.
(
2000
)
The versatility and universality of calcium signalling
.
Nat. Rev. Mol. Cell Biol.
1
,
11
21
[PubMed]
55.
Berridge
M.J.
,
Bootman
M.D.
and
Roderick
H.L.
(
2003
)
Calcium signalling: dynamics, homeostasis and remodelling
.
Nat. Rev. Mol. Cell Biol.
4
,
517
529
[PubMed]
56.
Clapham
D.E.
(
2007
)
Calcium signaling
.
Cell
131
,
1047
1058
[PubMed]
57.
Inesi
G.
and
Tadini-Buoninsegni
F.
(
2014
)
Ca(2+)/H (+) exchange, lumenal Ca(2+) release and Ca (2+)/ATP coupling ratios in the sarcoplasmic reticulum ATPase
.
J. Cell Commun. Signal.
8
,
5
11
58.
Schwaller
B.
(
2010
)
Cytosolic Ca2+ buffers
.
Cold Spring Harbor Perspect. Biol.
2
,
a004051
59.
Grabarek
Z.
(
2006
)
Structural basis for diversity of the EF-hand calcium-binding proteins
.
J. Mol. Biol.
359
,
509
525
[PubMed]
60.
Gifford
J.L.
,
Walsh
M.P.
and
Vogel
H.J.
(
2007
)
Structures and metal-ion-binding properties of the Ca2+-binding helix-loop-helix EF-hand motifs
.
Biochem. J.
405
,
199
221
[PubMed]
61.
Chazin
W.J.
(
2011
)
Relating form and function of EF-hand calcium binding proteins
.
Acc. Chem. Res.
44
,
171
179
[PubMed]
62.
Heizmann
C.W.
(
2019
)
Ca2+-binding proteins of the EF-hand superfamily: diagnostic and prognostic biomarkers and novel therapeutic targets
.
Methods Mol. Biol.
1929
,
157
186
[PubMed]
63.
Cheng
X.
,
Zhang
X.
,
Yu
L.
and
Xu
H.
(
2015
)
Calcium signaling in membrane repair
.
Semin. Cell Dev. Biol.
45
,
24
31
64.
Karatekin
E.
,
Sandre
O.
,
Guitouni
H.
,
Borghi
N.
,
Puech
P.H.
and
Brochard-Wyart
F.
(
2003
)
Cascades of transient pores in giant vesicles: line tension and transport
.
Biophys. J.
84
,
1734
1749
[PubMed]
65.
McNeil
P.L.
,
Miyake
K.
and
Vogel
S.S.
(
2003
)
The endomembrane requirement for cell surface repair
.
PNAS
100
,
4592
4597
[PubMed]
66.
Gozen
I.
and
Dommersnes
P.
(
2014
)
Pore dynamics in lipid membranes
.
Eur. Phys. J. Spec. Top.
223
,
1813
1829
67.
Terasaki
M.
,
Miyake
K.
and
McNeil
P.L.
(
1997
)
Large plasma membrane disruptions are rapidly resealed by Ca2+-dependent vesicle-vesicle fusion events
.
J. Cell Biol.
139
,
63
74
[PubMed]
68.
Lek
A.
,
Evesson
F.J.
,
Lemckert
F.A.
,
Redpath
G.M.
,
Lueders
A.K.
,
Turnbull
L.
et al.
(
2013
)
Calpains, cleaved mini-dysferlinC72, and L-type channels underpin calcium-dependent muscle membrane repair
.
J. Neurosci.
33
,
5085
5094
[PubMed]
69.
Duan
X.
,
Chan
K.T.
,
Lee
K.K.
and
Mak
A.F.
(
2015
)
Oxidative stress and plasma membrane repair in single myoblasts after femtosecond laser photoporation
.
Ann. Biomed. Eng.
43
,
2735
2744
[PubMed]
70.
Reddy
A.
,
Caler
E.V.
and
Andrews
N.W.
(
2001
)
Plasma membrane repair is mediated by Ca(2+)-regulated exocytosis of lysosomes
.
Cell
106
,
157
169
[PubMed]
71.
Jimenez
A.J.
,
Maiuri
P.
,
Lafaurie-Janvore
J.
,
Divoux
S.
,
Piel
M.
and
Perez
F.
(
2014
)
ESCRT machinery is required for plasma membrane repair
.
Science
343
,
1247136
[PubMed]
72.
Keyel
P.A.
,
Loultcheva
L.
,
Roth
R.
,
Salter
R.D.
,
Watkins
S.C.
,
Yokoyama
W.M.
et al.
(
2011
)
Streptolysin O clearance through sequestration into blebs that bud passively from the plasma membrane
.
J. Cell Sci.
124
,
2414
2423
[PubMed]
73.
McNeil
P.L.
and
Kirchhausen
T.
(
2005
)
An emergency response team for membrane repair
.
Nat. Rev. Mol. Cell Biol.
6
,
499
505
[PubMed]
74.
Andrews
N.W.
(
2005
)
Membrane resealing: synaptotagmin VII keeps running the show
.
Sci. STKE
2005
,
pe19
[PubMed]
75.
Boye
T.L.
and
Nylandsted
J.
(
2016
)
Annexins in plasma membrane repair
.
Biol. Chem.
397
,
961
969
[PubMed]
76.
Bansal
D.
and
Campbell
K.P.
(
2004
)
Dysferlin and the plasma membrane repair in muscular dystrophy
.
Trends Cell Biol.
14
,
206
213
[PubMed]
77.
Horn
A.
and
Jaiswal
J.K.
(
2018
)
Cellular mechanisms and signals that coordinate plasma membrane repair
.
Cell. Mol. Life Sci.
75
,
3751
3770
[PubMed]
78.
Idone
V.
,
Tam
C.
,
Goss
J.W.
,
Toomre
D.
,
Pypaert
M.
and
Andrews
N.W.
(
2008
)
Repair of injured plasma membrane by rapid Ca2+-dependent endocytosis
.
J. Cell Biol.
180
,
905
914
[PubMed]
79.
Corrotte
M.
,
Fernandes
M.C.
,
Tam
C.
and
Andrews
N.W.
(
2012
)
Toxin pores endocytosed during plasma membrane repair traffic into the lumen of MVBs for degradation
.
Traffic
13
,
483
494
[PubMed]
80.
Babiychuk
E.B.
and
Draeger
A.
(
2015
)
Defying death: cellular survival strategies following plasmalemmal injury by bacterial toxins
.
Semin. Cell Dev. Biol.
45
,
39
47
81.
Gerke
V.
,
Creutz
C.E.
and
Moss
S.E.
(
2005
)
Annexins: linking Ca2+ signalling to membrane dynamics
.
Nat. Rev. Mol. Cell Biol.
6
,
449
461
[PubMed]
82.
Cai
C.
,
Masumiya
H.
,
Weisleder
N.
,
Matsuda
N.
,
Nishi
M.
,
Hwang
M.
et al.
(
2009
)
MG53 nucleates assembly of cell membrane repair machinery
.
Nat. Cell Biol.
11
,
56
64
[PubMed]
83.
Draeger
A.
,
Monastyrskaya
K.
and
Babiychuk
E.B.
(
2011
)
Plasma membrane repair and cellular damage control: the annexin survival kit
.
Biochem. Pharmacol.
81
,
703
712
[PubMed]
84.
Koerdt
S.N.
,
Ashraf
A.
and
Gerke
V.
(
2019
)
Annexins and plasma membrane repair
.
Curr. Top. Membr.
84
,
43
65
[PubMed]
85.
Lee
J.
and
Littleton
J.T.
(
2015
)
Transmembrane tethering of synaptotagmin to synaptic vesicles controls multiple modes of neurotransmitter release
.
PNAS
112
,
3793
3798
[PubMed]
86.
Craxton
M.
(
2004
)
Synaptotagmin gene content of the sequenced genomes
.
BMC Genomics
5
,
43
[PubMed]
87.
Südhof
T.C.
(
2002
)
Synaptotagmins: why so many?
J. Biol. Chem.
277
,
7629
7632
[PubMed]
88.
Öhrfelt
A.
,
Brinkmalm
A.
,
Dumurgier
J.
,
Brinkmalm
G.
,
Hansson
O.
,
Zetterberg
H.
et al.
(
2016
)
The pre-synaptic vesicle protein synaptotagmin is a novel biomarker for Alzheimer's disease
.
Alzheimers Res. Ther.
8
,
41
89.
Coussens
L.
,
Parker
P.J.
,
Rhee
L.
,
Yang-Feng
T.L.
,
Chen
E.
,
Waterfield
M.D.
et al.
(
1986
)
Multiple, distinct forms of bovine and human protein kinase C suggest diversity in cellular signaling pathways
.
Science
233
,
859
866
[PubMed]
90.
Kohout
S.C.
,
Corbalán-García
S.
,
Torrecillas
A.
,
Goméz-Fernandéz
J.C.
and
Falke
J.J.
(
2002
)
C2 domains of protein kinase C isoforms alpha, beta, and gamma: activation parameters and calcium stoichiometries of the membrane-bound state
.
Biochemistry
41
,
11411
11424
[PubMed]
91.
Sutton
R.B.
,
Davletov
B.A.
,
Berghuis
A.M.
,
Südhof
T.C.
and
Sprang
S.R.
(
1995
)
Structure of the first C2 domain of synaptotagmin I: a novel Ca2+/phospholipid-binding fold
.
Cell
80
,
929
938
[PubMed]
92.
Sutton
R.B.
,
Ernst
J.A.
and
Brunger
A.T.
(
1999
)
Crystal structure of the cytosolic C2A-C2B domains of synaptotagmin III. Implications for Ca(+2)-independent snare complex interaction
.
J. Cell Biol.
147
,
589
598
[PubMed]
93.
Dai
H.
,
Shin
O.H.
,
Machius
M.
,
Tomchick
D.R.
,
Südhof
T.C.
and
Rizo
J.
(
2004
)
Structural basis for the evolutionary inactivation of Ca2+ binding to synaptotagmin 4
.
Nat. Struct. Mol. Biol.
11
,
844
849
94.
Qiu
X.
,
Ge
J.
,
Gao
Y.
,
Teng
M.
and
Niu
L.
(
2017
)
Structural analysis of Ca2+-binding pocket of synaptotagmin 5 C2A domain
.
Int. J. Biol. Macromol.
95
,
946
953
[PubMed]
95.
Xue
M.
,
Craig
T.K.
,
Shin
O.H.
,
Li
L.
,
Brautigam
C.A.
,
Tomchick
D.R.
et al.
(
2010
)
Structural and mutational analysis of functional differentiation between synaptotagmins-1 and -7
.
PloS ONE
5
,
e12544
[PubMed]
96.
Corbalan-Garcia
S.
and
Gómez-Fernández
J.C.
(
2014
)
Signaling through C2 domains: more than one lipid target
.
Biochim. Biophys. Acta
1838
,
1536
1547
[PubMed]
97.
Südhof
T.C.
(
2012
)
Calcium control of neurotransmitter release
.
Cold Spring Harbor Perspect. Biol.
4
,
a011353
98.
Guillén
J.
,
Ferrer-Orta
C.
,
Buxaderas
M.
,
Pérez-Sánchez
D.
,
Guerrero-Valero
M.
,
Luengo-Gil
G.
et al.
(
2013
)
Structural insights into the Ca2+ and PI(4,5)P2 binding modes of the C2 domains of rabphilin 3A and synaptotagmin 1
.
PNAS
110
,
20503
20508
[PubMed]
99.
Ubach
J.
,
Zhang
X.
,
Shao
X.
,
Südhof
T.C.
and
Rizo
J.
(
1998
)
Ca2+ binding to synaptotagmin: how many Ca2+ ions bind to the tip of a C2-domain?
EMBO J.
17
,
3921
3930
[PubMed]
100.
Nalefski
E.A.
and
Falke
J.J.
(
1996
)
The C2 domain calcium-binding motif: structural and functional diversity
.
Protein Sci.
5
,
2375
2390
[PubMed]
101.
Radhakrishnan
A.
,
Stein
A.
,
Jahn
R.
and
Fasshauer
D.
(
2009
)
The Ca2+ affinity of synaptotagmin 1 is markedly increased by a specific interaction of its C2B domain with phosphatidylinositol 4,5-bisphosphate
.
J. Biol. Chem.
284
,
25749
25760
[PubMed]
102.
Guerrero-Valero
M.
,
Ferrer-Orta
C.
,
Querol-Audí
J.
,
Marin-Vicente
C.
,
Fita
I.
,
Gómez-Fernández
J.C.
et al.
(
2009
)
Structural and mechanistic insights into the association of PKCalpha-C2 domain to PtdIns(4,5)P2
.
PNAS
106
,
6603
6607
[PubMed]
103.
Sugita
S.
,
Shin
O.H.
,
Han
W.
,
Lao
Y.
and
Südhof
T.C.
(
2002
)
Synaptotagmins form a hierarchy of exocytotic Ca(2+) sensors with distinct Ca(2+) affinities
.
EMBO J.
21
,
270
280
[PubMed]
104.
Shao
X.
,
Fernandez
I.
,
Südhof
T.C.
and
Rizo
J.
(
1998
)
Solution structures of the Ca2+-free and Ca2+-bound C2A domain of synaptotagmin I: does Ca2+ induce a conformational change?
Biochemistry
37
,
16106
16115
[PubMed]
105.
Geppert
M.
,
Goda
Y.
,
Hammer
R.E.
,
Li
C.
,
Rosahl
T.W.
,
Stevens
C.F.
et al.
(
1994
)
Synaptotagmin I: a major Ca2+ sensor for transmitter release at a central synapse
.
Cell
79
,
717
727
[PubMed]
106.
Ramakrishnan
S.
,
Bera
M.
,
Coleman
J.
,
Rothman
J.E.
and
Krishnakumar
S.S.
(
2020
)
Synergistic roles of synaptotagmin-1 and complexin in calcium-regulated neuronal exocytosis
.
eLife
9
,
e54506
[PubMed]
107.
de Wit
H.
,
Walter
A.M.
,
Milosevic
I.
,
Gulyás-Kovács
A.
,
Riedel
D.
,
Sørensen
J.B.
et al.
(
2009
)
Synaptotagmin-1 docks secretory vesicles to syntaxin-1/SNAP-25 acceptor complexes
.
Cell
138
,
935
946
[PubMed]
108.
Carr
C.M.
and
Munson
M.
(
2007
)
Tag team action at the synapse
.
EMBO Rep.
8
,
834
838
[PubMed]
109.
Kim
J.Y.
,
Choi
B.K.
,
Choi
M.G.
,
Kim
S.A.
,
Lai
Y.
,
Shin
Y.K.
et al.
(
2012
)
Solution single-vesicle assay reveals PIP2-mediated sequential actions of synaptotagmin-1 on SNAREs
.
EMBO J.
31
,
2144
2155
[PubMed]
110.
Yoon
T.Y.
and
Munson
M.
(
2018
)
SNARE complex assembly and disassembly
.
Curr. Biol.
28
,
R397
R401
[PubMed]
111.
Voleti
R.
,
Jaczynska
K.
and
Rizo
J.
(
2020
)
Ca2+-dependent release of synaptotagmin-1 from the SNARE complex on phosphatidylinositol 4,5-bisphosphate-containing membranes
.
eLife
9
,
e57154
[PubMed]
112.
Zhou
Q.
,
Zhou
P.
,
Wang
A.L.
,
Wu
D.
,
Zhao
M.
,
Südhof
T.C.
et al.
(
2017
)
The primed SNARE-complexin-synaptotagmin complex for neuronal exocytosis
.
Nature
548
,
420
425
[PubMed]
113.
Brunger
A.T.
,
Choi
U.B.
,
Lai
Y.
,
Leitz
J.
,
White
K.I.
and
Zhou
Q.
(
2019
)
The pre-synaptic fusion machinery
.
Curr. Opin. Struct. Biol.
54
,
179
188
[PubMed]
114.
Zhou
Q.
,
Lai
Y.
,
Bacaj
T.
,
Zhao
M.
,
Lyubimov
A.Y.
,
Uervirojnangkoorn
M.
et al.
(
2015
)
Architecture of the synaptotagmin-SNARE machinery for neuronal exocytosis
.
Nature
525
,
62
67
[PubMed]
115.
Li
F.
,
Pincet
F.
,
Perez
E.
,
Eng
W.S.
,
Melia
T.J.
,
Rothman
J.E.
et al.
(
2007
)
Energetics and dynamics of SNAREpin folding across lipid bilayers
.
Nat. Struct. Mol. Biol.
14
,
890
896
116.
Courtney
N.A.
,
Bao
H.
,
Briguglio
J.S.
and
Chapman
E.R.
(
2019
)
Synaptotagmin 1 clamps synaptic vesicle fusion in mammalian neurons independent of complexin
.
Nat. Commun.
10
,
4076
[PubMed]
117.
Grote
E.
,
Baba
M.
,
Ohsumi
Y.
and
Novick
P.J.
(
2000
)
Geranylgeranylated SNAREs are dominant inhibitors of membrane fusion
.
J. Cell Biol.
151
,
453
466
[PubMed]
118.
Han
J.
,
Pluhackova
K.
and
Böckmann
R.A.
(
2017
)
The multifaceted role of SNARE proteins in membrane fusion
.
Front. Physiol.
8
,
5
[PubMed]
119.
Schapire
A.L.
,
Voigt
B.
,
Jasik
J.
,
Rosado
A.
,
Lopez-Cobollo
R.
,
Menzel
D.
et al.
(
2008
)
Arabidopsis synaptotagmin 1 is required for the maintenance of plasma membrane integrity and cell viability
.
Plant Cell
20
,
3374
3388
[PubMed]
120.
Yamazaki
T.
,
Kawamura
Y.
,
Minami
A.
and
Uemura
M.
(
2008
)
Calcium-dependent freezing tolerance in Arabidopsis involves membrane resealing via synaptotagmin SYT1
.
Plant Cell
20
,
3389
3404
[PubMed]
121.
Martinez
I.
,
Chakrabarti
S.
,
Hellevik
T.
,
Morehead
J.
,
Fowler
K.
and
Andrews
N.W.
(
2000
)
Synaptotagmin VII regulates Ca(2+)-dependent exocytosis of lysosomes in fibroblasts
.
J. Cell Biol.
148
,
1141
1149
[PubMed]
122.
Dolai
S.
,
Xie
L.
,
Zhu
D.
,
Liang
T.
,
Qin
T.
,
Xie
H.
et al.
(
2016
)
Synaptotagmin-7 functions to replenish insulin granules for exocytosis in human islet β-cells
.
Diabetes
65
,
1962
1976
[PubMed]
123.
Bhalla
A.
,
Tucker
W.C.
and
Chapman
E.R.
(
2005
)
Synaptotagmin isoforms couple distinct ranges of Ca2+, Ba2+, and Sr2+ concentration to SNARE-mediated membrane fusion
.
Mol. Biol. Cell
16
,
4755
4764
[PubMed]
124.
Rao
S.K.
,
Huynh
C.
,
Proux-Gillardeaux
V.
,
Galli
T.
and
Andrews
N.W.
(
2004
)
Identification of SNAREs involved in synaptotagmin VII-regulated lysosomal exocytosis
.
J. Biol. Chem.
279
,
20471
20479
[PubMed]
125.
Bansal
D.
,
Miyake
K.
,
Vogel
S.S.
,
Groh
S.
,
Chen
C.C.
,
Williamson
R.
et al.
(
2003
)
Defective membrane repair in dysferlin-deficient muscular dystrophy
.
Nature
423
,
168
172
[PubMed]
126.
Roche
J.A.
,
Lovering
R.M.
,
Roche
R.
,
Ru
L.W.
,
Reed
P.W.
and
Bloch
R.J.
(
2010
)
Extensive mononuclear infiltration and myogenesis characterize recovery of dysferlin-null skeletal muscle from contraction-induced injuries
.
Am. J. Physiol. Cell Physiol.
298
,
C298
C312
127.
Díaz-Manera
J.
,
Touvier
T.
,
Dellavalle
A.
,
Tonlorenzi
R.
,
Tedesco
F.S.
,
Messina
G.
et al.
(
2010
)
Partial dysferlin reconstitution by adult murine mesoangioblasts is sufficient for full functional recovery in a murine model of dysferlinopathy
.
Cell Death Dis.
1
,
e61
128.
Piccolo
F.
,
Moore
S.A.
,
Ford
G.C.
and
Campbell
K.P.
(
2000
)
Intracellular accumulation and reduced sarcolemmal expression of dysferlin in limb–girdle muscular dystrophies
.
Ann. Neurol.
48
,
902
912
[PubMed]
129.
Millay
D.P.
,
Maillet
M.
,
Roche
J.A.
,
Sargent
M.A.
,
McNally
E.M.
,
Bloch
R.J.
et al.
(
2009
)
Genetic manipulation of dysferlin expression in skeletal muscle: novel insights into muscular dystrophy
.
Am. J. Pathol.
175
,
1817
1823
[PubMed]
130.
Kobayashi
K.
,
Izawa
T.
,
Kuwamura
M.
and
Yamate
J.
(
2011
)
Comparative gene expression analysis in the skeletal muscles of dysferlin-deficient SJL/J and A/J mice
.
J. Toxicologic Pathol.
24
,
49
62
131.
Kerr
J.P.
,
Ziman
A.P.
,
Mueller
A.L.
,
Muriel
J.M.
,
Kleinhans-Welte
E.
,
Gumerson
J.D.
et al.
(
2013
)
Dysferlin stabilizes stress-induced Ca2+ signaling in the transverse tubule membrane
.
PNAS
110
,
20831
20836
[PubMed]
132.
Abdullah
N.
,
Padmanarayana
M.
,
Marty
N.J.
and
Johnson
C.P.
(
2014
)
Quantitation of the calcium and membrane binding properties of the C2 domains of dysferlin
.
Biophys. J.
106
,
382
389
[PubMed]
133.
Therrien
C.
,
Di Fulvio
S.
,
Pickles
S.
and
Sinnreich
M.
(
2009
)
Characterization of lipid binding specificities of dysferlin C2 domains reveals novel interactions with phosphoinositides
.
Biochemistry
48
,
2377
2384
[PubMed]
134.
Fuson
K.
,
Rice
A.
,
Mahling
R.
,
Snow
A.
,
Nayak
K.
,
Shanbhogue
P.
et al.
(
2014
)
Alternate splicing of dysferlin C2A confers Ca²2+-dependent and Ca²2+-independent binding for membrane repair
.
Structure
22
,
104
115
[PubMed]
135.
Wang
Y.
,
Tadayon
R.
,
Santamaria
L.
,
Mercier
P.
,
Forristal
C.J.
and
Shaw
G.S.
(
2021
)
Calcium binds and rigidifies the dysferlin C2A domain in a tightly coupled manner
.
Biochem. J.
478
,
197
215
[PubMed]
136.
Sula
A.
,
Cole
A.R.
,
Yeats
C.
,
Orengo
C.
and
Keep
N.H.
(
2014
)
Crystal structures of the human Dysferlin inner DysF domain
.
BMC Struct. Biol.
14
,
3
[PubMed]
137.
Patel
P.
,
Harris
R.
,
Geddes
S.M.
,
Strehle
E.M.
,
Watson
J.D.
,
Bashir
R.
et al.
(
2008
)
Solution structure of the inner DysF domain of myoferlin and implications for limb girdle muscular dystrophy type 2b
.
J. Mol. Biol.
379
,
981
990
[PubMed]
138.
Therrien
C.
,
Dodig
D.
,
Karpati
G.
and
Sinnreich
M.
(
2006
)
Mutation impact on dysferlin inferred from database analysis and computer-based structural predictions
.
J. Neurol. Sci.
250
,
71
78
[PubMed]
139.
Fujita
E.
,
Kouroku
Y.
,
Isoai
A.
,
Kumagai
H.
,
Misutani
A.
,
Matsuda
C.
et al.
(
2007
)
Two endoplasmic reticulum-associated degradation (ERAD) systems for the novel variant of the mutant dysferlin: ubiquitin/proteasome ERAD(I) and autophagy/lysosome ERAD(II)
.
Hum. Mol. Genet.
16
,
618
629
[PubMed]
140.
Xu
L.
,
Pallikkuth
S.
,
Hou
Z.
,
Mignery
G.A.
,
Robia
S.L.
and
Han
R.
(
2011
)
Dysferlin forms a dimer mediated by the C2 domains and the transmembrane domain in vitro and in living cells
.
PloS ONE
6
,
e27884
[PubMed]
141.
Woolger
N.
,
Bournazos
A.
,
Sophocleous
R.A.
,
Evesson
F.J.
,
Lek
A.
,
Driemer
B.
et al.
(
2017
)
Limited proteolysis as a tool to probe the tertiary conformation of dysferlin and structural consequences of patient missense variant L344P
.
J. Biol. Chem.
292
,
18577
18591
[PubMed]
142.
Dominguez
M.J.
,
McCord
J.J.
and
Sutton
R.B.
(
2022
)
Redefining the architecture of ferlin proteins: insights into multi-domain protein structure and function
.
PloS ONE
17
,
e0270188
[PubMed]
143.
Han
R.
and
Campbell
K.P.
(
2007
)
Dysferlin and muscle membrane repair
.
Curr. Opin. Cell Biol.
19
,
409
416
[PubMed]
144.
McDade
J.R.
and
Michele
D.E.
(
2014
)
Membrane damage-induced vesicle-vesicle fusion of dysferlin-containing vesicles in muscle cells requires microtubules and kinesin
.
Hum. Mol. Genet.
23
,
1677
1686
[PubMed]
145.
Leung
C.
,
Utokaparch
S.
,
Sharma
A.
,
Yu
C.
,
Abraham
T.
,
Borchers
C.
et al.
(
2011
)
Proteomic identification of dysferlin-interacting protein complexes in human vascular endothelium
.
Biochem. Biophys. Res. Commun.
415
,
263
269
[PubMed]
146.
Huang
Y.
,
Laval
S.H.
,
van Remoortere
A.
,
Baudier
J.
,
Benaud
C.
,
Anderson
L.V.
et al.
(
2007
)
AHNAK, a novel component of the dysferlin protein complex, redistributes to the cytoplasm with dysferlin during skeletal muscle regeneration
.
FASEB J.
21
,
732
742
[PubMed]
147.
de Morrée
A.
,
Hensbergen
P.J.
,
van Haagen
H.H.
,
Dragan
I.
,
Deelder
A.M.
,
‘t Hoen
P.A.
et al.
(
2010
)
Proteomic analysis of the dysferlin protein complex unveils its importance for sarcolemmal maintenance and integrity
.
PloS ONE
5
,
e13854
[PubMed]
148.
Assadi
M.
,
Schindler
T.
,
Muller
B.
,
Porter
J.D.
,
Ruegg
M.A.
and
Langen
H.
(
2008
)
Identification of proteins interacting with dysferlin using the tandem affinity purification method
.
Open Cell Dev. Biol. J.
1
,
17
23
149.
Hernández-Deviez
D.J.
,
Howes
M.T.
,
Laval
S.H.
,
Bushby
K.
,
Hancock
J.F.
and
Parton
R.G.
(
2008
)
Caveolin regulates endocytosis of the muscle repair protein, dysferlin
.
J. Biol. Chem.
283
,
6476
6488
[PubMed]
150.
Matsuda
C.
,
Miyake
K.
,
Kameyama
K.
,
Keduka
E.
,
Takeshima
H.
,
Imamura
T.
et al.
(
2012
)
The C2A domain in dysferlin is important for association with MG53 (TRIM72)
.
PLoS Curr.
4
,
e5035add8caff4
[PubMed]
151.
Borgonovo
B.
,
Cocucci
E.
,
Racchetti
G.
,
Podini
P.
,
Bachi
A.
and
Meldolesi
J.
(
2002
)
Regulated exocytosis: a novel, widely expressed system
.
Nat. Cell Biol.
4
,
955
962
[PubMed]
152.
Bittel
D.C.
,
Chandra
G.
,
Tirunagri
L.
,
Deora
A.B.
,
Medikayala
S.
,
Scheffer
L.
et al.
(
2020
)
Annexin A2 mediates dysferlin accumulation and muscle cell membrane repair
.
Cells
9
,
1919
[PubMed]
153.
Erikson
E.
,
Tomasiewicz
H.G.
and
Erikson
R.L.
(
1984
)
Biochemical characterization of a 34-kilodalton normal cellular substrate of pp60v-src and an associated 6-kilodalton protein
.
Mol. Cell. Biol.
4
,
77
85
[PubMed]
154.
Gerke
V.
and
Weber
K.
(
1985
)
The regulatory chain in the p36-kd substrate complex of viral tyrosine-specific protein kinases is related in sequence to the S-100 protein of glial cells
.
EMBO J.
4
,
2917
2920
[PubMed]
155.
De Seranno
S.
,
Benaud
C.
,
Assard
N.
,
Khediri
S.
,
Gerke
V.
,
Baudier
J.
et al.
(
2006
)
Identification of an AHNAK binding motif specific for the Annexin2/S100A10 tetramer
.
J. Biol. Chem.
281
,
35030
35038
[PubMed]
156.
Rezvanpour
A.
,
Santamaria-Kisiel
L.
and
Shaw
G.S.
(
2011
)
The S100A10-annexin A2 complex provides a novel asymmetric platform for membrane repair
.
J. Biol. Chem.
286
,
40174
40183
[PubMed]
157.
Dempsey
B.R.
,
Rezvanpour
A.
,
Lee
T.W.
,
Barber
K.R.
,
Junop
M.S.
and
Shaw
G.S.
(
2012
)
Structure of an asymmetric ternary protein complex provides insight for membrane interaction
.
Structure
20
,
1737
1745
[PubMed]
158.
Ozorowski
G.
,
Milton
S.
and
Luecke
H.
(
2013
)
Structure of a C-terminal AHNAK peptide in a 1:2:2 complex with S100A10 and an acetylated N-terminal peptide of annexin A2
.
Acta Crystallogr. D.
69
,
92
104
159.
Johnson
G.V.
and
Guttmann
R.P.
(
1997
)
Calpains: intact and active?
Bioessays
19
,
1011
1018
160.
Campbell
R.L.
and
Davies
P.L.
(
2012
)
Structure-function relationships in calpains
.
Biochem. J.
447
,
335
351
[PubMed]
161.
Hanna
R.A.
,
Campbell
R.L.
and
Davies
P.L.
(
2008
)
Calcium-bound structure of calpain and its mechanism of inhibition by calpastatin
.
Nature
456
,
409
412
[PubMed]
162.
Moldoveanu
T.
,
Hosfield
C.M.
,
Lim
D.
,
Elce
J.S.
,
Jia
Z.
and
Davies
P.L.
(
2002
)
A Ca(2+) switch aligns the active site of calpain
.
Cell
108
,
649
660
[PubMed]
163.
Hosfield
C.M.
,
Elce
J.S.
,
Davies
P.L.
and
Jia
Z.
(
1999
)
Crystal structure of calpain reveals the structural basis for Ca(2+)-dependent protease activity and a novel mode of enzyme activation
.
EMBO J.
18
,
6880
6889
[PubMed]
164.
Mellgren
R.L.
,
Zhang
W.
,
Miyake
K.
and
McNeil
P.L.
(
2007
)
Calpain is required for the rapid, calcium-dependent repair of wounded plasma membrane
.
J. Biol. Chem.
282
,
2567
2575
[PubMed]
165.
Mellgren
R.L.
,
Miyake
K.
,
Kramerova
I.
,
Spencer
M.J.
,
Bourg
N.
,
Bartoli
M.
et al.
(
2009
)
Calcium-dependent plasma membrane repair requires m- or mu-calpain, but not calpain-3, the proteasome, or caspases
.
Biochim. Biophys. Acta
1793
,
1886
1893
[PubMed]
166.
Piper
A.K.
,
Sophocleous
R.A.
,
Ross
S.E.
,
Evesson
F.J.
,
Saleh
O.
,
Bournazos
A.
et al.
(
2020
)
Loss of calpains-1 and -2 prevents repair of plasma membrane scrape injuries, but not small pores, and induces a severe muscular dystrophy
.
Am. J. Physiol. Cell Physiol.
318
,
C1226
C1237
[PubMed]
167.
Godell
C.M.
,
Smyers
M.E.
,
Eddleman
C.S.
,
Ballinger
M.L.
,
Fishman
H.M.
and
Bittner
G.D.
(
1997
)
Calpain activity promotes the sealing of severed giant axons
.
PNAS
94
,
4751
4756
[PubMed]
168.
Redpath
G.M.
,
Woolger
N.
,
Piper
A.K.
,
Lemckert
F.A.
,
Lek
A.
,
Greer
P.A.
et al.
(
2014
)
Calpain cleavage within dysferlin exon 40a releases a synaptotagmin-like module for membrane repair
.
Mol. Biol. Cell
25
,
3037
3048
[PubMed]
169.
Krahn
M.
,
Wein
N.
,
Bartoli
M.
,
Lostal
W.
,
Courrier
S.
,
Bourg-Alibert
N.
et al.
(
2010
)
A naturally occurring human minidysferlin protein repairs sarcolemmal lesions in a mouse model of dysferlinopathy
.
Sci. Transl. Med.
2
,
50ra69
[PubMed]
170.
Ballouhey
O.
,
Courrier
S.
,
Kergourlay
V.
,
Gorokhova
S.
,
Cerino
M.
,
Krahn
M.
et al.
(
2021
)
The dysferlin transcript containing the alternative exon 40a is essential for myocyte functions
.
Front. Cell Dev. Biol.
9
,
754555
171.
Huang
Y.
,
de Morrée
A.
,
van Remoortere
A.
,
Bushby
K.
,
Frants
R.R.
,
den Dunnen
J.T.
et al.
(
2008
)
Calpain 3 is a modulator of the dysferlin protein complex in skeletal muscle
.
Hum. Mol. Genet.
17
,
1855
1866
[PubMed]
172.
Taveau
M.
,
Bourg
N.
,
Sillon
G.
,
Roudaut
C.
,
Bartoli
M.
and
Richard
I.
(
2003
)
Calpain 3 is activated through autolysis within the active site and lyses sarcomeric and sarcolemmal components
.
Mol. Cell. Biol.
23
,
9127
9135
[PubMed]
173.
Kourie
J.I.
and
Wood
H.B.
(
2000
)
Biophysical and molecular properties of annexin-formed channels
.
Prog. Biophys. Mol. Biol.
73
,
91
134
[PubMed]
174.
Swairjo
M.A.
and
Seaton
B.A.
(
1994
)
Annexin structure and membrane interactions: a molecular perspective
.
Annu. Rev. Biophys. Biomol. Struct.
23
,
193
213
[PubMed]
175.
Jin
M.
,
Smith
C.
,
Hsieh
H.Y.
,
Gibson
D.F.
and
Tait
J.F.
(
2004
)
Essential role of B-helix calcium binding sites in annexin V-membrane binding
.
J. Biol. Chem.
279
,
40351
40357
[PubMed]
176.
Sopkova
J.
,
Renouard
M.
and
Lewit-Bentley
A.
(
1993
)
The crystal structure of a new high-calcium form of annexin V
.
J. Mol. Biol.
234
,
816
825
[PubMed]
177.
Ando
Y.
,
Imamura
S.
,
Hong
Y.M.
,
Owada
M.K.
,
Kakunaga
T.
and
Kannagi
R.
(
1989
)
Enhancement of calcium sensitivity of lipocortin I in phospholipid binding induced by limited proteolysis and phosphorylation at the amino terminus as analyzed by phospholipid affinity column chromatography
.
J. Biol. Chem.
264
,
6948
6955
[PubMed]
178.
Weng
X.
,
Luecke
H.
,
Song
I.S.
,
Kang
D.S.
,
Kim
S.H.
and
Huber
R.
(
1993
)
Crystal structure of human annexin I at 2.5 A resolution
.
Protein Sci.
2
,
448
458
[PubMed]
179.
Raynal
P.
and
Pollard
H.B.
(
1994
)
Annexins: the problem of assessing the biological role for a gene family of multifunctional calcium- and phospholipid-binding proteins
.
Biochim. Biophys. Acta
1197
,
63
93
[PubMed]
180.
Pigault
C.
,
Follénius-Wund
A.
,
Lux
B.
and
Gérard
D.
(
1990
)
A fluorescence spectroscopy study of the calpactin I complex and its subunits p11 and p36: calcium-dependent conformation changes
.
Biochim. Biophys. Acta
1037
,
106
114
[PubMed]
181.
Shadle
P.J.
,
Gerke
V.
and
Weber
K.
(
1985
)
Three Ca2+-binding proteins from porcine liver and intestine differ immunologically and physicochemically and are distinct in Ca2+ affinities
.
J. Biol. Chem.
260
,
16354
16360
[PubMed]
182.
Monastyrskaya
K.
,
Babiychuk
E.B.
,
Hostettler
A.
,
Rescher
U.
and
Draeger
A.
(
2007
)
Annexins as intracellular calcium sensors
.
Cell Calcium
41
,
207
219
[PubMed]
183.
Rosengarth
A.
and
Luecke
H.
(
2003
)
A calcium-driven conformational switch of the N-terminal and core domains of annexin A1
.
J. Mol. Biol.
326
,
1317
1325
[PubMed]
184.
Réty
S.
,
Osterloh
D.
,
Arié
J.P.
,
Tabaries
S.
,
Seeman
J.
,
Russo-Marie
F.
et al.
(
2000
)
Structural basis of the Ca(2+)-dependent association between S100C (S100A11) and its target, the N-terminal part of annexin I
.
Structure
8
,
175
184
[PubMed]
185.
Réty
S.
,
Sopkova
J.
,
Renouard
M.
,
Osterloh
D.
,
Gerke
V.
,
Tabaries
S.
et al.
(
1999
)
The crystal structure of a complex of p11 with the annexin II N-terminal peptide
.
Nat. Struct. Biol.
6
,
89
95
[PubMed]
186.
Santamaria-Kisiel
L.
,
Rintala-Dempsey
A.C.
and
Shaw
G.S.
(
2006
)
Calcium-dependent and -independent interactions of the S100 protein family
.
Biochem. J.
396
,
201
214
[PubMed]
187.
Baudier
J.
,
Deloulme
J.C.
and
Shaw
G.S.
(
2020
)
The Zn2+ and Ca2+ -binding S100B and S100A1 proteins: beyond the myths
.
Biol. Rev. Camb. Philos. Soc.
95
,
738
758
[PubMed]
188.
Weisz
J.
and
Uversky
V.N.
(
2020
)
Zooming into the dark side of human annexin-S100 complexes: dynamic alliance of flexible partners
.
Int. J. Mol. Sci.
21
,
5879
[PubMed]
189.
Sopkova-De Oliveira Santos
J.
,
Fischer
S.
,
Guilbert
C.
,
Lewit-Bentley
A.
and
Smith
J.C.
(
2000
)
Pathway for large-scale conformational change in annexin V
.
Biochemistry
39
,
14065
14074
[PubMed]
190.
Bharadwaj
A.
,
Kempster
E.
and
Waisman
D.M.
(
2021
)
The annexin A2/S100A10 complex: the mutualistic symbiosis of two distinct proteins
.
Biomolecules
11
,
1849
[PubMed]
191.
Rintala-Dempsey
A.C.
,
Santamaria-Kisiel
L.
,
Liao
Y.
,
Lajoie
G.
and
Shaw
G.S.
(
2006
)
Insights into S100 target specificity examined by a new interaction between S100A11 and annexin A2
.
Biochemistry
45
,
14695
14705
[PubMed]
192.
Ecsédi
P.
,
Kiss
B.
,
Gógl
G.
,
Radnai
L.
,
Buday
L.
,
Koprivanacz
K.
et al.
(
2017
)
Regulation of the equilibrium between closed and open conformations of annexin A2 by N-terminal phosphorylation and S100A4-binding
.
Structure
25
,
1195.e5
1207.e5
[PubMed]
193.
Ecsédi
P.
,
Gógl
G.
,
Hóf
H.
,
Kiss
B.
,
Harmat
V.
and
Nyitray
L.
(
2020
)
Structure determination of the transactivation domain of p53 in complex with S100A4 using annexin A2 as a crystallization chaperone
.
Structure
28
,
943.e4
953.e4
194.
Nedjadi
T.
,
Kitteringham
N.
,
Campbell
F.
,
Jenkins
R.E.
,
Park
B.K.
,
Navarro
P.
et al.
(
2009
)
S100A6 binds to annexin 2 in pancreatic cancer cells and promotes pancreatic cancer cell motility
.
Br. J. Cancer
101
,
1145
1154
195.
Semov
A.
,
Moreno
M.J.
,
Onichtchenko
A.
,
Abulrob
A.
,
Ball
M.
,
Ekiel
I.
et al.
(
2005
)
Metastasis-associated protein S100A4 induces angiogenesis through interaction with annexin II and accelerated plasmin formation
.
J. Biol. Chem.
280
,
20833
20841
196.
Puisieux
A.
,
Ji
J.
and
Ozturk
M.
(
1996
)
Annexin II up-regulates cellular levels of p11 protein by a post-translational mechanisms
.
Biochem. J.
313
,
51
55
197.
Caplan
J.F.
,
Filipenko
N.R.
,
Fitzpatrick
S.L.
and
Waisman
D.M.
(
2004
)
Regulation of annexin A2 by reversible glutathionylation
.
J. Biol. Chem.
279
,
7740
7750
198.
Chasserot-Golaz
S.
,
Vitale
N.
,
Umbrecht-Jenck
E.
,
Knight
D.
,
Gerke
V.
and
Bader
M.F.
(
2005
)
Annexin 2 promotes the formation of lipid microdomains required for calcium-regulated exocytosis of dense-core vesicles
.
Mol. Biol. Cell
16
,
1108
1119
199.
Croissant
C.
,
Gounou
C.
,
Bouvet
F.
,
Tan
S.
and
Bouter
A.
(
2022
)
Trafficking of annexins during membrane repair in human skeletal muscle cells
.
Membranes
12
,
153
200.
Roostalu
U.
and
Strähle
U.
(
2012
)
In vivo imaging of molecular interactions at damaged sarcolemma
.
Dev. Cell
22
,
515
529
201.
Leikina
E.
,
Defour
A.
,
Melikov
K.
,
Van der Meulen
J.H.
,
Nagaraju
K.
,
Bhuvanendran
S.
et al.
(
2015
)
Annexin A1 deficiency does not affect myofiber repair but delays regeneration of injured muscles
.
Sci. Rep.
5
,
18246
[PubMed]
202.
Defour
A.
,
Medikayala
S.
,
Van der Meulen
J.H.
,
Hogarth
M.W.
,
Holdreith
N.
,
Malatras
A.
et al.
(
2017
)
Annexin A2 links poor myofiber repair with inflammation and adipogenic replacement of the injured muscle
.
Hum. Mol. Genet.
26
,
1979
1991
[PubMed]
203.
Grewal
T.
,
Wason
S.J.
,
Enrich
C.
and
Rentero
C.
(
2016
)
Annexins - insights from knockout mice
.
Biol. Chem.
397
,
1031
1053
[PubMed]
204.
McNeil
A.K.
,
Rescher
U.
,
Gerke
V.
and
McNeil
P.L.
(
2006
)
Requirement for annexin A1 in plasma membrane repair
.
J. Biol. Chem.
281
,
35202
35207
[PubMed]
205.
Jaiswal
J.K.
,
Lauritzen
S.P.
,
Scheffer
L.
,
Sakaguchi
M.
,
Bunkenborg
J.
,
Simon
S.M.
et al.
(
2014
)
S100A11 is required for efficient plasma membrane repair and survival of invasive cancer cells
.
Nat. Commun.
5
,
3795
[PubMed]
206.
Glover
L.
and
Brown
R.H.
Jr
(
2007
)
Dysferlin in membrane trafficking and patch repair
.
Traffic
8
,
785
794
[PubMed]
207.
Reviakine
I.
,
Bergsma-Schutter
W.
and
Brisson
A.
(
1998
)
Growth of protein 2-D crystals on supported planar lipid bilayers imaged in situ by AFM
.
J. Struct. Biol.
121
,
356
361
[PubMed]
208.
Carmeille
R.
,
Bouvet
F.
,
Tan
S.
,
Croissant
C.
,
Gounou
C.
,
Mamchaoui
K.
et al.
(
2016
)
Membrane repair of human skeletal muscle cells requires Annexin-A5
.
Biochim. Biophys. Acta
1863
,
2267
2279
[PubMed]
209.
Lin
Y.C.
,
Chipot
C.
and
Scheuring
S.
(
2020
)
Annexin-V stabilizes membrane defects by inducing lipid phase transition
.
Nat. Commun.
11
,
230
[PubMed]
210.
Bouter
A.
,
Gounou
C.
,
Bérat
R.
,
Tan
S.
,
Gallois
B.
,
Granier
T.
,
d'Estaintot
B.L.
et al.
(
2011
)
Annexin-A5 assembled into two-dimensional arrays promotes cell membrane repair
.
Nat. Commun.
2
,
270
[PubMed]
211.
Kaetzel
M.A.
,
Mo
Y.D.
,
Mealy
T.R.
,
Campos
B.
,
Bergsma-Schutter
W.
,
Brisson
A.
et al.
(
2001
)
Phosphorylation mutants elucidate the mechanism of annexin IV-mediated membrane aggregation
.
Biochemistry
40
,
4192
4199
[PubMed]
212.
Boye
T.L.
,
Maeda
K.
,
Pezeshkian
W.
,
Sønder
S.L.
,
Haeger
S.C.
,
Gerke
V.
et al.
(
2017
)
Annexin A4 and A6 induce membrane curvature and constriction during cell membrane repair
.
Nat. Commun.
8
,
1623
[PubMed]
213.
Demonbreun
A.R.
,
Fallon
K.S.
,
Oosterbaan
C.C.
,
Bogdanovic
E.
,
Warner
J.L.
,
Sell
J.J.
et al.
(
2019
)
Recombinant annexin A6 promotes membrane repair and protects against muscle injury
.
J. Clin. Invest.
129
,
4657
4670
[PubMed]
214.
Jang
I.K.
,
Hu
R.
,
Lacaná
E.
,
D'Adamio
L.
and
Gu
H.
(
2002
)
Apoptosis-linked gene 2-deficient mice exhibit normal T-cell development and function
.
Mol. Cell. Biol.
22
,
4094
4100
[PubMed]
215.
Missotten
M.
,
Nichols
A.
,
Rieger
K.
and
Sadoul
R.
(
1999
)
Alix, a novel mouse protein undergoing calcium-dependent interaction with the apoptosis-linked-gene 2 (ALG-2) protein
.
Cell Death Differ.
6
,
124
129
[PubMed]
216.
Katoh
K.
,
Shibata
H.
,
Suzuki
H.
,
Nara
A.
,
Ishidoh
K.
,
Kominami
E.
et al.
(
2003
)
The ALG-2-interacting protein Alix associates with CHMP4b, a human homologue of yeast Snf7 that is involved in multivesicular body sorting
.
J. Biol. Chem.
278
,
39104
39113
[PubMed]
217.
la Cour
J.M.
,
Winding Gojkovic
P.
,
Ambjørner
S.
,
Bagge
J.
,
Jensen
S.M.
,
Panina
S.
et al.
(
2018
)
ALG-2 participates in recovery of cells after plasma membrane damage by electroporation and digitonin treatment
.
PloS ONE
13
,
e0204520
[PubMed]
218.
Subramanian
L.
,
Crabb
J.W.
,
Cox
J.
,
Durussel
I.
,
Walker
T.M.
,
van Ginkel
P.R.
et al.
(
2004
)
Ca2+ binding to EF hands 1 and 3 is essential for the interaction of apoptosis-linked gene-2 with Alix/AIP1 in ocular melanoma
.
Biochemistry
43
,
11175
11186
[PubMed]
219.
Maki
M.
,
Suzuki
H.
and
Shibata
H.
(
2011
)
Structure and function of ALG-2, a penta-EF-hand calcium-dependent adaptor protein
.
Sci. China Life Sci.
54
,
770
779
220.
Lo
K.W.
,
Zhang
Q.
,
Li
M.
and
Zhang
M.
(
1999
)
Apoptosis-linked gene product ALG-2 is a new member of the calpain small subunit subfamily of Ca2+-binding proteins
.
Biochemistry
38
,
7498
7508
[PubMed]
221.
Kitaura
Y.
,
Matsumoto
S.
,
Satoh
H.
,
Hitomi
K.
and
Maki
M.
(
2001
)
Peflin and ALG-2, members of the penta-EF-hand protein family, form a heterodimer that dissociates in a Ca2+-dependent manner
.
J. Biol. Chem.
276
,
14053
14058
[PubMed]
222.
Maki
M.
,
Yamaguchi
K.
,
Kitaura
Y.
,
Satoh
H.
and
Hitomi
K.
(
1998
)
Calcium-induced exposure of a hydrophobic surface of mouse ALG-2, which is a member of the penta-EF-hand protein family
.
J. Biochem.
124
,
1170
1177
223.
Scheffer
L.L.
,
Sreetama
S.C.
,
Sharma
N.
,
Medikayala
S.
,
Brown
K.J.
,
Defour
A.
et al.
(
2014
)
Mechanism of Ca²⁺-triggered ESCRT assembly and regulation of cell membrane repair
.
Nat. Commun.
5
,
5646
[PubMed]
224.
Hurley
J.H.
(
2015
)
ESCRTs are everywhere
.
EMBO J.
34
,
2398
2407
[PubMed]
225.
McCullough
J.
,
Fisher
R.D.
,
Whitby
F.G.
,
Sundquist
W.I.
and
Hill
C.P.
(
2008
)
ALIX-CHMP4 interactions in the human ESCRT pathway
.
PNAS
105
,
7687
7691
[PubMed]
226.
Larios
J.
,
Mercier
V.
,
Roux
A.
and
Gruenberg
J.
(
2020
)
ALIX- and ESCRT-III-dependent sorting of tetraspanins to exosomes
.
J. Cell Biol.
219
,
e201904113
[PubMed]
227.
Sønder
S.L.
,
Boye
T.L.
,
Tölle
R.
,
Dengjel
J.
,
Maeda
K.
,
Jäättelä
M.
et al.
(
2019
)
Annexin A7 is required for ESCRT III-mediated plasma membrane repair
.
Sci. Rep.
9
,
6726
[PubMed]
228.
Schlaepfer
D.D.
and
Haigler
H.T.
(
1987
)
Characterization of Ca2+-dependent phospholipid binding and phosphorylation of lipocortin I
.
J. Biol. Chem.
262
,
6931
6937
[PubMed]
229.
Glenney
J.
(
1986
)
Phospholipid-dependent Ca2+ binding by the 36-kDa tyrosine kinase substrate (calpactin) and its 33-kDa core
.
J. Biol. Chem.
261
,
7247
7252
[PubMed]
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).