Next Article in Journal
Plasma microRNA Signature as Companion Diagnostic for Abiraterone Acetate Treatment in Metastatic Castration-Resistant Prostate Cancer: A Pilot Study
Next Article in Special Issue
The Role of MicroRNA in the Pathogenesis of Duchenne Muscular Dystrophy
Previous Article in Journal
Phytonanotherapy for the Treatment of Metabolic Dysfunction-Associated Steatotic Liver Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Limb Girdle Muscular Dystrophy Type 2B (LGMD2B): Diagnosis and Therapeutic Possibilities

by
Bal Hari Poudel
1,2,3,
Sue Fletcher
1,
Steve D. Wilton
1,2 and
May Aung-Htut
1,2,*
1
Centre for Molecular Medicine and Innovative Therapeutics, Health Futures Institute, Murdoch University, Perth, WA 6150, Australia
2
Perron Institute for Neurological and Translational Science, The University of Western Australia, Perth, WA 6009, Australia
3
Central Department of Biotechnology, Tribhuvan University, Kirtipur, Kathmandu 44618, Nepal
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(11), 5572; https://doi.org/10.3390/ijms25115572
Submission received: 18 April 2024 / Revised: 11 May 2024 / Accepted: 16 May 2024 / Published: 21 May 2024

Abstract

:
Dysferlin is a large transmembrane protein involved in critical cellular processes including membrane repair and vesicle fusion. Mutations in the dysferlin gene (DYSF) can result in rare forms of muscular dystrophy; Miyoshi myopathy; limb girdle muscular dystrophy type 2B (LGMD2B); and distal myopathy. These conditions are collectively known as dysferlinopathies and are caused by more than 600 mutations that have been identified across the DYSF gene to date. In this review, we discuss the key molecular and clinical features of LGMD2B, the causative gene DYSF, and the associated dysferlin protein structure. We also provide an update on current approaches to LGMD2B diagnosis and advances in drug development, including splice switching antisense oligonucleotides. We give a brief update on clinical trials involving adeno-associated viral gene therapy and the current progress on CRISPR/Cas9 mediated therapy for LGMD2B, and then conclude by discussing the prospects of antisense oligomer-based intervention to treat selected mutations causing dysferlinopathies.

1. Introduction

Dysferlinopathies are a family of autosomal recessive muscular disorders caused by mutations in the dysferlin gene (DYSF) and include the following diseases: Limb-girdle muscular dystrophy type 2B (LGMD2B; MIM #253601), Miyoshi myopathy (MM; MIM #254130), and distal myopathy anterior tibial onset (DMAT; OMIM #606768). Dysferlin (DYSF) is primarily expressed in skeletal and cardiac muscles, where it encodes a ~237 kDa plasma membrane protein that plays a crucial role in the repair of damaged muscle cell membranes [1,2]. The protein encoded by dysferlin belongs to the ferlin protein family that is characterized by the presence of a type II transmembrane domain, with the majority of the domain adjacent to the cytoplasm, as well as multiple calcium binding (C2) domains that are involved in calcium-dependent membrane fusion events [3,4]. As with most muscular dystrophies, there is currently no curative treatment available for dysferlinopathies. However, preliminary in vitro and in vivo studies using viral gene therapy, CRISPR/CaS9, and antisense oligonucleotide (AO) technologies have shown promise in treating limb girdle muscular dystrophy 2B (LGMD2B). Among these, AO-mediated splice switching presents a promising strategy to treat some mutations causing LGMD2B. Antisense oligonucleotide-mediated intervention during pre-mRNA processing has been clinically approved for selected mutations causing Duchenne muscular dystrophy, reactivating inappropriately processed gene transcripts as a treatment for spinal muscular atrophy and a single patient with Batten’s disease [5,6].

1.1. Dysferlin

The dysferlin gene (DYSF) is located on chromosome 2p12-14 and composed of 55 exons, spread over 150 Kb of genomic DNA, and encodes a protein of ~237 kDa. As mentioned previously, defects in DYSF can manifest as different types of muscular dystrophy based on clinical presentation [7,8]. The weakening and degeneration of muscles in the pelvic and shoulder girdles, often accompanied by a deficiency or absence of the dysferlin protein, typically exhibit in one’s third or fourth decade of life. Nonetheless, in severe instances, this reduction or absence of dysferlin may become evident earlier [9].
The contraction of voluntary skeletal muscles generates the energy required for both motility and strength. As such, skeletal muscle cells are subjected to significant contractile forces and stresses, frequently resulting in disruption of the muscle surface membrane (sarcolemma). Dysferlin assists in repairing the disrupted membrane in association with Ca+2 binding proteins annexins (A1, A2 and A6), mitsugumin, and S100A [10]. As a consequence of insufficient amounts of, or non-functional, dysferlin, the accumulation of membrane damage eventually leads to muscular dystrophy.
The DYSF promoter region contains two CpG islands, a TATA box, and two clusters of binding sites for various transcription factors, including those involved in muscle-specific dysferlin expression. Alternative splicing of exons 5, 17, and 40 has been previously reported [11]. Using 5′ rapid amplification of the cDNA ends on the adult skeletal muscle total RNA, Pramono et al. (2006) identified a splice variant of DYSF, DYSF-V1 (Ref seq NM_003494.4), that arises from alternate exon inclusion in the 5′ UTR [12]. Pramono and colleagues had previously reported other novel transcripts due to the inclusion of exons derived from either intron 5 or 40 [12]. An alternative transcript generated through omission of exon 17 was also reported [12]. High levels of DYSF mRNA expression have been reported in blood, skeletal muscle, heart, and the placenta, with lower expression observed in the brain, kidney, lung, and even less in the liver and pancreas. According to Aoki et al. (2001), the largest reported dysferlin cDNA sequence spans 6.9 Kb and is found in skeletal muscle [13], whereas in the brain, the major transcript reported has a length of only 3.8 Kb, with the highest expression being reported in the putamen and essentially no expression in the spinal cord and fetal brain [14,15].
The cytoplasmic region of dysferlin contains seven calcium-binding domains (C2A to C2G, from N-terminus to C-terminus) that are thought to mediate vesicle fusion with the plasma membrane. In addition, dysferlin has multiple domains including ferlin “fer” and dysferlin “dysF” domains. Various calcium-dependent lipid-binding C2 (C2), “fer”, and “dysF” domains are shown in Figure 1 [4]. Ferlin domain A (FerA) of dysferlin binds to phospholipids and plays an integral role in membrane fusion activity in a calcium-dependent manner [16]. The C2 domains are independently folded blocks of approximately 130 residues that assemble into a beta-sandwich motif containing eight anti-parallel beta-sheets [4]. These C2 domains also participate in lipid and protein binding [17] and are generally involved in membrane interactions or fusion events, or in the generation of secondary messenger lipids involved in signal transduction pathways [18,19].

1.2. The Role of Dysferlin in Membrane Repair and the Intracellular Vesicular System

In muscle cells, dystrophin–glycoprotein complex (DGC) proteins maintain muscle membrane integrity and structure, while the muscle membrane repair complex rapidly repairs sarcolemmal tears and ruptures. Dysferlin is one of the most important proteins in the sarcolemmal repair process, although deficiencies of any protein in this repair complex would lead to muscular dystrophy, a heterogeneous group of muscle-wasting diseases [4].
According to the membrane repair model for dysferlin, damage to the membrane influences the diffusion of calcium within the muscle fibers, resulting in a zone of high calcium around the disrupted site [21]. In the presence of localized high levels of calcium, dysferlin-carrying repair vesicles are directed to the site of damage, where they accumulate and fuse with one another and the plasma membrane. Dysferlin then facilitates vesicle docking and fusion with the plasma membrane by interacting with annexin A2, other dysferlin molecules, and additional unknown protein-binding molecules [2,21,22]. Fusion of the repair vesicles with the plasma membrane acts as a “patch” across the disruption and thereby reseals the damaged plasma membrane [10,21]. Research into other pathways associated with dysferlin-linked membrane repair is still ongoing.
Dysferlin also interacts with other proteins to repair membrane degeneration. Sharma et al. (2010) indicated a novel dysferlin interacting partner—platelet endothelial cellular adhesion molecule-1 (PECAM-1). PECAM-1 is an essential adhesion molecule for angiogenesis regulation, which adds to our understanding of the functional roles of ferlins in angiogenesis and selective protein trafficking in a vascular setting [23]. Lek et al. (2013) described the interaction of dysferlin with a previously uncharacterized membrane repair protein, mitsugumin-53 (MG53), an E3 ubiquitin ligase that is rapidly recruited to injury sites [24]. They showed that an injury-specific calpain cleaves dysferlin into a ~72 kDa C-terminal dysferlin isoform, a mini-dysferlinC72 that is recruited to the sites of membrane injury rather than the full-length dysferlin. They reported that mini-dysferlinC72-rich vesicles are rapidly recruited to injury sites and fuse with plasma membrane compartments decorated by MG53 in a process coordinated by L-type calcium channels [24]. The role of dysferlin in the repair of a damaged sarcolemma membrane is shown in Figure 2.
Dysferlin also plays an important role in regulating calcium (Ca2+) signaling by maintaining the balance of Ca2+ levels in T-tubule membranes, particularly around the triad junction [26,27]. Muriel et al. conducted a study investigating the function of each dysferlin C2 domain [28]. They found that specific domains are involved in Ca2+ signaling, while others contribute independently to the process of membrane repair. Additionally, they observed that the absence of dysferlin leads to a reduction in the strength of the voltage-induced Ca2+ signals. They suggested that dysferlin may aid in optimizing the interaction between the L-type calcium channel (LTCC) and the ryanodine receptor (RyR1) at the triad junction [28]. Furthermore, Wang et al. demonstrated that the dysferlin C2A domain binds with two calcium ions, resulting in a more stable structure that enhances calcium binding [29].
Mitochondria, cellular energy generators that produce adenosine triphosphate (ATP) by oxidative phosphorylation, are also involved in Ca2+ buffering. Vincent et al. (2016) hypothesized that mitochondrial defects may be evident in the skeletal muscle biopsies from dysferlinopathy patients. Their analyses revealed that the percentage of mitochondrial complex I- and complex IV (cytochrome c oxidase)-deficient muscle fibers was higher in patients with DYSF mutations than in the healthy controls [30]. However, it was noted that these patients did not exhibit any rearrangements in their mitochondrial DNA.
Furthermore, the membrane repair function of dysferlin has been linked to the soluble N-ethylmaleimide-sensitive factor (SNARE). The direct interaction between dysferlin and the SNARE proteins, syntaxin 4 and SNAP-23, was reported by Codding et al. (2016). In addition, dysferlin accelerated the syntaxin4/SNAP23 complex formation and SNARE mediated lipid mixing in a calcium-dependent manner [31].
Through interactions with other proteins, dysferlin also contributes to tubule formation and many studies have provided additional evidence for this function [32,33]. Demonbreun et al. (2014) showed that transverse tubule formation and glycerol sensitivity were regulated through dysferlin and myoferlin [34]. The lack of functional ferlins and reduced vesicle trafficking causes the accumulation of lipids. In the absence of functional ferlins, T-tubules and plasma membranes are unable to repair damage. A leaky sarcolemma results in the release of lipids, including glycerol, which promotes myopathy by functionally separating the T-tubules from the sarcolemma and sarcoplasmic reticulum [35]. Ultimately, adipogenesis and the accumulation of adipocytes are promoted within the muscle, which leads to an increased production of glycerol, a by-product of lipolysis [36]. Furthermore, dysferlin assists with the arrangement of liposomes, and produces a T-tubule-like membrane system in non-muscle cells that facilitates the biogenesis of the T-tubules system [32]. Finally, dysferlin has also been reported to interact with the proteins associated with human neuroblast differentiation, such as AHNAK, affixin, S100A10 [37], calpain, and dihydropyridine receptor [38]. In summary, dysferlin is a protein crucial for the maintenance of membrane repair; hence, reduced dysferlin expression or the production of non-functional dysferlin protein isoforms due to null or missense mutations can cause muscle-wasting diseases.

1.3. Mutations in Dysferlin and LGMD2B

Liu et al. (1998) identified DYSF mutations in patients with dysferlinopathies and observed that the same DYSF mutations can present more than one disease phenotype in different individuals [1]. Sinnreich et al. (2006) identified a single base substitution in a highly conserved branch point sequence of intron 31 that led to the exclusion of exon 32 from the mature mRNA [39]. Western blot analysis revealed a reduction in dysferlin levels to about 10% of those observed in healthy controls, contributing to a mild phenotype. This observation indicated that a proportion of the transcript may have escaped the exon 32 exclusion, producing some dysferlin protein. Mutation studies by Mafalda et al. (2011) confirmed the primary involvement of DYSF in the LGMD 2B/MM phenotypes [40], providing the first direct and conclusive evidence that dysferlin protein levels lower than 10% of those in healthy individuals are considered pathogenic [40].
Due to the constantly increasing spectrum of DYSF mutations, the Universal Mutation Database for Dysferlin (UMD-DYSF) was established to manage the expanding mutational dataset [41]. UMD-DYSF is a locus-specific database that contains extensive data related to DYSF mutations [8,41] and provides a comprehensive account of the DYSF disease-causing mutations reported in the literature [8]. According to the UMD-DYSF, different types of mutations, which include nonsense, missense, and indels, are dispersed relatively evenly throughout DYSF, as shown in Figure 3. Missense mutations are the most abundant followed by deletion and nonsense mutations (Figure 3) and insertion is the least reported. In addition to the recurrent mutations, the UMD-DYSF database also lists seven founder mutations [41]. The Portuguese population shows two founder mutations c.1180_1180+7delAGTGCGTG (r.1054_1284del, p.Glu353_Leu429del) and c.5492G>A. Founder mutations in the Canadian (c.2372C>G (p.Pro791Arg)), Italian (c.2875C>T (p.Arg959Trp), Caucasian Jewish c.2779delG (p.Ala927LeufsX21), and Lebanese Jewish (c.4872_4876delinsCCCC (p.Glu1624AspfsX9)) populations are reported. Lastly, 2% of Spanish dysferlinopathy patients from the region of Sueca have a c.5713C>T (p.Arg1905X) founder mutation. Izumi et al. (2020) reported a c.29997G>T; p.Trp999Cys mutation as the most frequent (22.9%) in a Japanese cohort of 209 cases [42]. Similarly, they also found that the frequency of missense mutations was higher (70.6%) in the inner dysferlin domain [42]. However, in a French dysferlinopathy cohort, the incidence of nonsense mutations was found to be higher [43]. The frequency of missense, nonsense, and indel mutations appears to vary across diverse populations as evidenced by several studies on different cohorts [41,44,45].
A DYSF mutation study conducted on 245 dysferlinopathy patients of Chinese ethnicity by Zhong et al. (2021) identified 40 novel mutations and observed c.1375dupin 6.5% of the patients [44]. Genomic deletions of exons 3, 33, 34, 35, 40, and 41, and 42 were reported in the same cohort. Yi-Ying Hu and colleagues (2018) reported cases where two different mutations were found (compound heterozygous mutations) in a patient, including a de novo mutation c.613C>T in exon 6 and a novel missense mutation c.968T>C in exon 11 [46].
To further study the consequences of different mutations, disease phenotype, muscle pathology and the effect of drugs, researchers have developed several dysferlin-deficient mouse models, as well as studied naturally occurring dystrophic mouse strains.

1.4. Mouse Models to Study Dysferlinopathies

In recent years, several mouse dysferlinopathy models have been identified or developed, including SJL/J, A/J, BLA/J, and Dysf-knock out mice [47,48]. SJL/J mice possess a 171 bp genomic deletion affecting the 3′ splice site of exon 45 [49] in the DYSF gene, leading to decreased dysferlin protein levels compared to wild-type mice. Consequently, this alteration results in the spontaneous development of myopathy. The A/J mouse strain is another naturally occurring dysferlin-deficient animal arising from a retrotransposon insertion in DYSF intron 4, while BLA/J mice were derived from the SJL/L strain [50]. Despite a compromised dysferlin, the SJL/J and A/J murine models do not exhibit significant muscle weakness. This is not unusual, as some other mouse muscular dystrophy models such as the mdx mouse that carries a nonsense mutation in the dystrophin gene do not show obvious symptoms until the animals are of advanced age [51]. Therefore, the screening and analysis of any dysferlin therapies are normally dependent on histopathological examinations, such as observing the fiber shape and size, central nucleation, fibrosis, and inflammation.
To study the consequences of one particular missense mutation, Malcher et al. (2018) developed a new mouse model called MMex38, which carries a missense mutation in DYSF exon 38. Similar to the human DYSF variant p.Leu1341Pro, this particular variant induces all the characteristics of a dysferlinopathy arising from a missense mutation, including progressive muscular dystrophy, amyloid formation, and defects in membrane repair [52]. The MMex38 mouse model was employed for a study where U7 small nuclear RNA (snRNA)-mediated splice switching was used to induce dual exon 37 and 38 skipping from the DYSF transcript. Dysferlin exon skipping was assessed in vitro in C2C12 murine cells with a reported efficiency of 12.9%. The subsequent in vivo study carried out in MMex38 mice showed less efficient exon skipping as determined by RT-PCR, but an increase in dysferlin protein functionality was reported [52].

Stem Cells as a Model

Patient-derived induced pluripotent stem cells (iPSCs) have been used as an in vitro model to test potential drug compounds. Kokubu et al. (2019) screened small molecules using iPSCs generated from a patient with compound heterozygous mutations; a missense mutation c.2997G>T(p.W999C) and a nonsense mutation c.1958delG, and observed that nocodazole was found to increase the level of misfolded dysferlin expression in cells with increased membrane resealing following injury by irradiation when compared to healthy controls [53]. Further research is required to assess the effectiveness of nocodazole against various pathogenic missense mutations present in dysferlin, as illustrated in Figure 3. Additionally, it is necessary to establish optimized growth and differentiation protocols to enable reliable safety assessments using iPSCs as a cellular model and for cell replacement therapy.

2. LGMD2B Disease Symptoms and Diagnosis

Dysferlinopathies are characterized by the atrophy and weakness of the gastrocnemius muscle and/or the anterior tibial muscles, along with elevated creatine kinase (CK). In muscle biopsies, mononuclear cell infiltration may be observed [54,55] and can often be misdiagnosed as an inflammatory myopathy, such as polymyositis. Misdiagnosis can lead to inappropriate treatment with anti-inflammatory drugs such as corticosteroids, that are ineffective in treating the consequences of dysferlin mutations and typically result in many adverse side effects, including loss of muscle strength, reduced bone density, hypertension, cataracts, and diabetes [56].
Mutations affecting the structure and functions of the dysferlin protein manifest with chronic muscle fiber loss, fat replacement, and fibrosis [57]. Clinical symptoms of dysferlinopathies include progressive muscle weakness, increased serum kinase (CK), inflammation, and abnormal muscle morphology. Approximately 30 percent of LGMD2B patients become wheelchair dependent within 15 years from disease onset, but severity and progression can vary significantly between populations [58,59]. Although many studies have been conducted, the exact prevalence of LGMD2B is still not known and can only be estimated [60], which is between 1 in 14,286 and 1 in 200,000 [41]. Others have estimated a smaller prevalence range, from 1 in 14,500 to 1 in 123,000 people, with LGMD2B thought to account for between 3% and 19% of all LGMDs [41].
Immunohistochemistry on muscle biopsy can be inconsistent when using the Novocastra antibodies NCL-Hamlet and NCL-Hamlet-2 to confirm diagnosis [61]. Western blotting is the most proficient method for determining diagnosis based on quantity and molecular size of the 237 kDa protein. In affected subjects, dysferlin levels can vary from partial deficiency to total absence. Matching the phenotype with dysferlin levels has been challenging. Anderson et al. (2000) demonstrated low levels in homozygous frameshifting, none in nonsense/frameshift, mid-range in frameshift, and deletion mutations and high levels in missense mutations [62]. In another study, all reported pathogenic dysferlin mutations affected the protein expression level in skeletal muscle [63].
Genomic analysis of the DYSF gene is the most effective approach to detect pathogenic variants [58]. Through whole genome sequencing, changes such as synonymous mutations that influence splicing, splice site mutations, deletions and duplications in coding and noncoding regions can be discovered. Similarly, whole transcriptome sequencing allows for the characterization of all types of transcripts (both coding and noncoding). After confirming the presence of a pathogenic mutation, genetic counseling becomes accessible, marking the initial phase in the implementation of gene/cell replacement or mutation-specific therapies.

3. Therapeutic Strategies for Dysferlinopathies

Although there are currently no effective drugs available to treat dysferlinopathies, clinical trials are ongoing with an adeno-associated virus-mediated gene replacement strategy (NCT02710500) [64], as well as various preliminary studies using gene and AO therapies in mouse models and patient-derived cells. In the following sections, the current studies and outcomes of adeno-associated viral (AAV) vector-based gene therapy, CRISPR (clustered regularly interspaced short palindromic repeats)/Cas9 gene correction, small molecule and nonsense mutation readthrough approaches, as well as AO interventions are discussed and shown in Figure 4.

3.1. Adeno-Associated Virus-Mediated Gene Therapy

Adeno-associated virus vectors infect dividing and non-dividing cells and are extensively used for gene therapy applications; favored for their relatively low immunogenicity, lack of pathogenicity, and ability to establish long-term transgene expression [65]. AAV-mediated gene therapy for dysferlinopathy involves introducing functional copies of the dysferlin gene into affected cells to compensate for the defective or absent dysferlin protein. This can potentially restore normal muscle function and alleviate symptoms associated with dysferlinopathies, such as muscle weakness and degeneration.
One of the major limitations of recombinant AAVs is the limited cargo capacity of 4.7 Kb. The DYSF-protein-coding sequence spans 6.9 Kb, thus severely restricting its suitability for AAV-mediated gene transfer therapy, although it has been suggested that larger genomes can be packaged into some AAV vectors generated through rare recombination events [66].
An alternative gene transfer strategy is packaging the first (exons 1–29) and second half (exons 29–55) of the DYSF cDNA into two separate AAV vectors. These two dysferlin transcripts contain donor and acceptor splice site sequences to facilitate trans-splicing and generate a full-length dysferlin coding transcript. A preclinical study conducted by Lostal et al. (2010) demonstrated that intravenous injection of viral constructs into dysferlin-deficient mice resulted in the expression of full-length dysferlin and led to an improvement in function compared to untreated or placebo-treated controls [67]. Dysferlin function was assessed histologically by a reduction in the number of necrotic fibers, restoration of membrane repair capacity in the muscle, and a global improvement in locomotor activity [67]. A similar preclinical study was performed by Sondergaard and colleagues who used the dual-vector system to package the DYSF cDNA into AAV serotype rh.74 using two discrete vectors with a 1 Kb region of homology [68]. They showed an increase in dysferlin expression in mice and non-human primates after intramuscular and vascular delivery. Similarly, in a separate in vivo study, Potter et al. (2018) delivered full-length 6.9 Kb dysferlin cDNA using dual AAV vectors. Dysferlin-deficient mice were injected with the AAV viral constructs systemically and an improvement in membrane repair, comparable to that of wild-type levels, was reported [69]. An AAV construct (rAAVrh74.MHCK7.DYSF.DV) developed by Sarepta Therapeutics is another dual vector system under clinical evaluation in a phase I safety and tolerability study. This AAV construct rAAVrh74.MHCK7.DYSF.DV is a dual recombinant AAV carrying each half of the dysferlin transgene under control of the muscle- and heart-specific MHCK7 enhancer [70,71] (NCT02710500).
A separate clinical trial has assessed the safety, effectiveness, and tolerability of SRP-6004, a dual-vector AAV gene therapy, when delivered intravenously to ambulatory patients (NCT05906251) [72]. Adaptive and innate immune responses are major limitations of viral vector-mediated gene therapies [73]. A number of clinical trials using AAVs are on hold due to serious adverse events, including deaths from 2020 to 2021 [74,75,76]. Astellas reported the deaths of four participants enrolled in their AT132 gene therapy trial for X-linked myotubular myopathy, which was halted twice due to safety concerns [75]. The deaths of three participants were linked to the high dosages; however, the cause of the fourth, not related to high dosage, is unknown. Similarly, Biomarin’s BMN 307 gene therapy drug for phenylketonuria has been placed on hold after liver tumors were observed in the mice administered with BMN 307 [75].

3.2. CRISPR/Cas9-Mediated Precise Correction for Pathogenic DYSF Mutations

CRISPR/Cas9 is an antiviral mechanism present in some bacteria that has become an exciting and popular tool for gene editing [77,78]. The CRISPR/Cas9 system relies on the Cas9 nuclease to be directed to specifically bind and cleave a nucleic acid sequence through the annealing of the guide RNA sequence (gRNA sequence). After cleavage, the DNA can be edited, replaced, or re-joined by nonhomologous end joining (NHEJ) or through homology-directed repair (HDR) mechanisms [79]. There are two classes of CRISPR/Cas systems (class I and class II) present in different bacteria, classified according to the structure and functions of the Cas protein. Class I includes type I, III, and IV, and class II includes type II, V, and VI, based on the target. DNA is recognized and cleaved by Type I, II, and V systems, whereas type VI edits RNAs and type III can edit both RNA and DNA. Turan et al. (2016) used CRISPR/Cas9 for the precise in vitro correction of disease-causing mutations in iPSCs derived from patients with LGMD2B [80,81]. They were able to correct a dysferlin nonsense mutation, c. 5713C>T; p.R1905X and the most common alpha-sarcoglycan mutation, a missense mutation, c.229C>T; p. R77C, using the CRISPR/Cas9 gene editing system [81]. Although the authors claimed this technique was promising, the efficiency of allele-specific correction was only 0.7–1.5%, and off-target effects were also observed, another major limitation that needs to be addressed and overcome.

3.3. Readthrough of Nonsense Mutations to Treat Dysferlinopathies

Bypassing nonsense mutations with some compounds that influence recognition of premature termination codons could generate functional dysferlin protein. However, this readthrough approach can be applied only to nonsense mutations, which account for ~25% of recurrent mutations in dysferlinopathy, as shown in Figure 3. Some antibiotics, in particular aminoglycosides such as gentamicin can induce readthrough of nonsense mutations. Gentamicin was tested in Duchenne muscular dystrophy (DMD) patients and dystrophin protein was induced; however, there were severe side effects such as nephrotoxicity and ototoxicity which precluded gentamicin for long-term usage [82]. Ataluren (PTC124) is another small molecule reported to bypass premature stop codons in DMD [83] and other disorders, like Leber congenital amaurosis type 4 (LCA4) [84], cystic fibrosis transmembrane conductance regulator (CFTR) [85], and Usher Syndrome 2A [86]. It was developed by PTC Therapeutics Inc (South Plainfield, NJ) and is currently undergoing phase 3 clinical trials for DMD (NCT03179631) [83,87].
Seo et al., 2021 demonstrated that a premature stop codon in a humanized DYSF knock-in mouse model (dqx) could respond to ataluren treatment, whereby readthrough of the p.Q832* mutation was reported to induce some functional recovery. The dqx mouse lacks dysferlin in skeletal muscle from birth but after two weeks of oral ataluren treatment at 0.9 mg/mL, some restoration of dysferlin expression and reduced skeletal muscle pathology was evident. As a negative control, the same treatments in the A/J mouse did not show any improvement, since this animal carries a unique ETn retrotransposon inserted in intron 4 that cannot respond to ataluren treatment [88]. These outcomes support potential treatment of DYSF nonsense mutations, providing hope for dysferlinopathy patients carrying a nonsense mutation [88].

3.4. Small Molecule Restoration of Membrane Repair Function

4-phenylbutyric acid (4-PBA) can partially restore the membrane repair ability of some mutant dysferlins. Tominaga et al. (2022) demonstrated that 4-PBA can partially restore the membrane localization of 25 different dysferlin missense variants in an HEK cell assay [89]. Subsequently, they showed the rescue of membrane repair ability of the variants in patient-derived cells after treatment with 4-PBA. A two-day oral administration of 4-PBA solution (2 mg/mL in sterile drinking water) in MMex38 mice resulted in expression and localization of dysferlin similar to the untreated, age-matched wild-type animals, whereas the untreated MMex38 mice were membrane-repair deficient [89].

3.5. Antisense Oligonucleotide-Mediated Therapies

Antisense oligonucleotides are short, synthetic single-stranded oligonucleotides that can be designed to bind to specific regions of mRNA or pre-mRNA to modify gene expression through several distinct mechanisms, as determined by the AO chemistry [5,90] and target sequence. DNA analogue antisense sequences can activate RNaseH degradation of the target RNA and hence downregulate specific gene expression [91], whereas RNA-like analogues do not induce RNaseH activity and can be used as steric blockers to influence translation or modulate pre-mRNA splicing in order to correct aberrant splicing [92], remove specific exons [93], or retain selected exons as required. Due to rapid degradation by endogenous nucleases, natural DNA and RNA oligonucleotides have limited potential as therapeutic agents. To overcome this limitation, oligonucleotides with modified bases and backbones have been developed to improve stability, enhance binding affinity, and minimize the immune-stimulatory response [94,95].
The splice-switching drug Exondys 51 was approved by the US Food and Drug Administration in September 2016 for the treatment of the most common subset of DMD-causing dystrophin deletions [96,97,98]. Exondys 51 is a phosphorodiamidate morpholino oligomer and reframes the dystrophin transcripts disrupted by some deletions flanking DMD exon 51. Restoration of the reading frame generates internally truncated but partially functional protein isoforms, as found in some mildly affected Becker muscular dystrophy (BMD) patients [96]. The more recent approval of additional splice-modulating AO drugs [99], Golodirsen, Viltolarsen [100] and Casimersen [101], to treat other DMD deletions and the approval of Nusinersen to treat spinal muscular atrophy has expanded the repertoire of antisense drugs to treat rare diseases. Nusinersen is another example of an AO drug that specifically targets a region of the SMN2 pre-mRNA downstream of exon 7 [102,103]. By binding to a strong silencer element in this pre-mRNA, Nusinersen alters the splicing process by promoting the inclusion of exon 7 in the mature SMN2 mRNA and the production of functional SMN protein [104].
While the exon skipping or exon retention drugs were developed to treat these common rare diseases, perhaps a more remarkable outcome was seen with the approval of the personalized splice-switching AO, Milasen, in 2018. The drug was developed to treat a six-year-old child, Mila, who was suffering from Batten’s disease arising from mutations in her neuronal ceroid lipofuscinosis 7 (CLN7) gene. Mutational analysis showed a genomic insertion of an SVA(SINE-VNTR-Alu) retrotransposon-a DNA sequence that can replicate and insert itself into different parts of the genome. SVAs are composite elements composed of three main parts, namely the SINE (Short Interspersed Nuclear Element), VNTR (Variable Number Tandem Repeat), and Alu (another type of retrotransposon) [105]. They can influence gene expression and may cause genetic mutations or alterations when inserted into genes or regulatory regions of the genome. The SVA insertion creates a cryptic splice-acceptor site in intron 6 in one of her CLN7 alleles [92], resulting in the retention of a partial CLN7 intron 6 sequence in the mature mRNA. Milasen, a 22 nucleotide (2′-O-methoxyethyl modified nucleotides on a phosphorothioate backbone, 2′ OMe-PS) AO targeting the CLN7 intron 6 cryptic splice site reduced this aberrant splice form and forced the splicing machinery to default to the normal splicing pattern. Milasen was the first FDA approved drug customized for the use of just one person and progressed from identification of the disease-causing mutation to patient treatment in less than a year. Similar approaches to treat the same type of splicing defect (cryptic splice site activation and pseudo-exon retention) should be feasible for some rare pathogenic mutations in DYSF. Furthermore, some nonsense, missense or frame-shifting indels in redundant DYSF exons could be by-passed if the loss of that exon in the mature mRNA does not completely compromise dysferlin function.

3.5.1. Antisense Oligonucleotide Mediated Strategies to Address Dysferlinopathy

Sinnreich et al. described two sisters with severe dystrophic symptoms who both possessed homozygous dysferlin null mutations, {(4872delG) and (G4876C)} [39]. Both parents were identified as heterozygous carriers of these mutations. Their mother carried an additional novel substitution of A to G at position –33 in intron 31 (A344333>G) of DYSF. This A to G variation compromised the branch point, resulting in skipping of the in-frame exon 32 from the mature mRNA. The dysferlin protein produced from this latter allele appeared to provide partial compensation for the null mutation, consistent with the mother’s less severe phenotype [39]. This report suggests that DYSF exon 32 can be removed from the mature mRNA without seriously compromising dysferlin function. Barthelemy et al. (2015) undertook another in vitro study and induced DYSF exon 32 skipping using AOs [106]. They also performed an in vitro functional analysis of dysferlin and reported an increase in functional dysferlin expression [106].
Dominov et al. (2014) identified a novel deep intronic point mutation within intron 44, (c.4886+1249 G>T). This mutation activated a cryptic donor splice site that altered the normal splicing of the DYSF mRNA by creating an aberrant transcript that contained an extra 177 nucleotides from intron 44. The additional 59 amino acids encoded within the conserved C2F domain of the dysferlin protein completely disrupted function [107]. To correct this aberrant splicing, three 2′ OMe-PS AOs targeting the acceptor and donor sites of the pseudoexon were designed and used to block the abnormal splicing event. Two AOs targeting the donor site of pseudoexon, after transfection into the patient’s cells, restored synthesis of the full-length DYSF mRNA with increased dysferlin expression [107]. Subsequently, Dominov et al. (2019) identified a deep intronic point mutation (c.5668-824 C>T) in intron 50 of DYSF in another dysferlinopathy patient. The mutation within intron 50 triggers cryptic splicing, leading to the insertion of an additional 180 bases, which results in the premature termination of protein translation within the DYSF C2G domain. As with the other examples, these cryptic splicing scenarios could be suppressed by the splice-switching AOs, and efficacy assessed by transcript analysis and Western blot analysis [108].
Pathogenic mutations in dysferlin are dispersed throughout the gene [8,45], and like many other genes, not all mutations can be addressed with splice-switching AOs and hence precise identification and characterization of mutations is crucial. The development of AOs to induce dystrophin exon skipping was informed by the clear genotype–phenotype correlation, with in-frame genomic deletions typically (but not always) associated with the milder form of Becker muscular dystrophy (BMD) [109]. Hence, restoration of the dystrophin reading frame for a frame-shifting deletion or by-passing a premature termination of translation mutation within an in-frame exon should result in a protein that retains some function. Similarly, it will be imperative to thoroughly investigate which DYSF exons other than exon 32 are dispensable, and therefore amenable to AO-mediated exon skipping. Non-essential exons carrying nonsense, missense, or frame-shifting indel mutations may be removed by steric-blocking AOs to yield isoforms that retain some dysferlin function. Similarly, cryptic exons and some splicing mutations could be addressed as was achieved with Milasen to restore normal splicing.

3.5.2. Overcoming Limitations and Advancing Delivery Strategies of AO

Antisense therapy presents a promising and potentially adaptable approach to treating a variety of disorders. However, this therapeutic class has some limitations that necessitate careful consideration to maximize benefits while minimizing risks and challenges. These limitations include restricted tissue penetration, off-target effects, limited duration of action, potential immune activation, high cost, and delivery challenges [110].
Chemical modifications are essential for improving the delivery, efficiency, stability, and target specificity of AOs. Altering the phosphate backbone of AOs can enhance their stability and resistance to nuclease degradation. Phosphorothioate (PS) backbone modification, where a sulfur atom replaces one of the non-bridging oxygen atoms in the phosphate backbone, is one of the most commonly used modifications for this purpose. Modifying the sugar moiety of nucleotides by incorporating 2′-O-methyl (2′-OMe), 2′-O-methoxyethyl (2′-MOE) or 2′-fluoro (2′-F) modifications, can further improve the stability of AOs while maintaining their RNA binding affinity [111,112]. Substituting specific nucleotide bases with modified analogs, such as locked nucleic acid (LNA) or peptide nucleic acid (PNA) can also enhance the binding affinity and specificity of AOs to their target RNA sequences [94,113,114]. Phosphorodiamidate morpholino oligomers (PMO) replace the deoxyribose/ribose moiety with a morpholine ring and substitute the charged phosphodiester inter-subunit linkage with an uncharged phosphorodiamidate linkage. This alteration renders PMOs nuclease-resistant and charge-neutral [115]. Another variation of the morpholino oligomers are thiomorpholino oligonucleotides (TMOs), a recently developed novel nucleic acid analog consisting of a morpholino nucleoside connected by thiophosphoramidate internucleotide linkages [116]. Several studies are currently underway to evaluate TMO chemistry in vitro and in vivo [116,117,118].
Conjugating AOs with lipophilic moieties, such as cholesterol or fatty acids, can promote their cellular uptake and intracellular delivery by facilitating their interaction with cell membranes and endosomal escape [119]. In addition, conjugating AOs with cell-penetrating peptides (CPPs) or other cell-targeting ligands can enhance their cellular uptake and tissue-specific delivery by promoting receptor-mediated endocytosis or translocation across cell membranes [120,121,122]. The safety profile of the neutrally charged PMOs could be due in part to the relatively poor cellular uptake, an issue that is being addressed by conjugation to cell-penetrating peptides [123,124]. An ongoing study showed that SRP-5051 (a PMO targeting DMD exon 51 conjugated to a proprietary cell-penetrating peptide) could offer greater efficacy with less frequent dosing than Exondys 51, the FDA-approved PMO targeting exon 51 for the most common subset of Duchenne MD patients. Sarepta’s predictive model shows that SRP-5051 at 30 mg/Kg after 12 weeks dosed monthly resulted in 18 times higher exon skipping and eight times higher dystrophin production compared to Exondys 51 that is administered weekly at 30 mg/Kg [125]. Treatment with SRP-5051 greatly reduces the impost on patients; the frequency of clinic visits is reduced, and a lower dosage is required while generating more of the induced dystrophin isoform. Encapsulating AOs within nanoparticles, such as lipid nanoparticles (LNPs) [126,127,128] or polymeric nanoparticles [129,130], can protect them from degradation, hence enhancing their circulation time, and facilitate their targeted delivery to specific tissues or cells.

4. Conclusions

As is the case with many inherited diseases, LGMD2B is currently considered to be an untreatable neuromuscular condition with poor prognosis [131]. Several researchers are studying potential treatments, and few have reached the clinical trial stage. Multiple therapeutic approaches including gene replacement, CRISPR/Cas9 gene editing, readthrough strategies, small molecule modulators of pathology, and AO-mediated exon skipping are being explored as potential treatments for LGMD2B. As with all ongoing research, there is an expectation that forthcoming treatments will greatly enhance quality of life for individuals with dysferlinopathy. However, more studies are required to better assess the novel therapeutic approaches to verify both efficacy and safety. The trajectory of dysferlinopathy treatment hinges on integrated research, novel therapeutic approaches, and cooperative endeavors that prioritize patient welfare and enhances the quality of life.

Author Contributions

B.H.P., M.A.-H. and S.D.W. performed the conceptualization; B.H.P. drafted the review, S.D.W., S.F. and M.A.-H. edited. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by funding from the National Health and Medical Research Council (Grant No. AP1144791). B.H.P. received an International Tuition Fee Scholarship from Murdoch University and scholarships from Perron institute to conduct his PhD studies.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Liu, J.; Aoki, M.; Illa, I.; Wu, C.; Fardeau, M.; Angelini, C.; Serrano, C.; Urtizberea, J.A.; Hentati, F.; Hamida, M.B.; et al. Dysferlin, a novel skeletal muscle gene, is mutated in Miyoshi myopathy and limb girdle muscular dystrophy. Nat. Genet. 1998, 20, 31–36. [Google Scholar] [CrossRef]
  2. Lennon, N.J.; Kho, A.; Bacskai, B.J.; Perlmutter, S.L.; Hyman, B.T.; Brown, R.H. Dysferlin interacts with Annexins A1 and A2 and mediates sarcolemmal wound-healing. J. Biol. Chem. 2003, 278, 50466–50473. [Google Scholar] [CrossRef] [PubMed]
  3. Britton, S.; Freeman, T.; Vafiadaki, E.; Keers, S.; Harrison, R.; Bushby, K.; Bashir, R. The third human FER-1-like protein is highly similar to dysferlin. Genomics 2000, 68, 313–321. [Google Scholar] [CrossRef]
  4. Bulankina, A.V.; Thoms, S. Functions of Vertebrate Ferlins. Cells 2020, 9, 534. [Google Scholar] [CrossRef]
  5. karishma Dhuri, C.B.; Quijano, E.; Pham, H.; Gupta, A.; Bikram, A.; Bahal, R. Antisense Oligonucleotides: An Emerging Area in Drug Discovery and Development. Clin. Med. 2020, 9, 2004. [Google Scholar]
  6. Aung-Htut, M.T.; Ham, K.A.; Tchan, M.; Johnsen, R.; Schnell, F.J.; Fletcher, S.; Wilton, S.D. Splice modulating antisense oligonucleotides restore some acid-alpha-glucosidase activity in cells derived from patients with late-onset Pompe disease. Sci. Rep. 2020, 10, 6702. [Google Scholar] [CrossRef]
  7. Pegoraro, E.; Hoffman, E.P. Limb-Girdle Muscular Dystrophy Overview. In GeneReviews®; Adam, M.P., Ardinger, H.H., Pagon, R.A., Wallace, S.E., Bean, L.J.H., Stephens, K., Amemiya, A., Eds.; University of Washington, Seattle: Seattle, WA, USA, 1993. [Google Scholar]
  8. Blandin, G.; Beroud, C.; Labelle, V.; Nguyen, K.; Wein, N.; Hamroun, D.; Williams, B.; Monnier, N.; Rufibach, L.E.; Urtizberea, J.A.; et al. UMD-DYSF, a novel locus specific database for the compilation and interactive analysis of mutations in the dysferlin gene. Hum. Mutat. 2012, 33, E2317–E2331. [Google Scholar] [CrossRef] [PubMed]
  9. Aoki, M.; Takahashi, T. Dysferlinopathy. In GeneReviews®; Adam, M.P., Ardinger, H.H., Pagon, R.A., Wallace, S.E., Bean, L.J.H., Gripp, K.W., Mirzaa, G.M., Amemiya, A., Eds.; University of Washington, Seattle: Seattle, WA, USA, 1993. [Google Scholar]
  10. Han, R.; Campbell, K.P. Dysferlin and muscle membrane repair. Curr. Opin. Cell Biol. 2007, 19, 409–416. [Google Scholar] [CrossRef] [PubMed]
  11. Salani, S.; Lucchiari, S.; Fortunato, F.; Crimi, M.; Corti, S.; Locatelli, F.; Bossolasco, P.; Bresolin, N.; Comi, G.P. Developmental and tissue-specific regulation of a novel dysferlin isoform. Muscle Nerve 2004, 30, 366–374. [Google Scholar] [CrossRef]
  12. Pramono, Z.A.; Lai, P.S.; Tan, C.L.; Takeda, S.; Yee, W.C. Identification and characterization of a novel human dysferlin transcript: Dysferlin_v1. Hum. Genet. 2006, 120, 410–419. [Google Scholar] [CrossRef]
  13. Aoki, M.; Liu, J.; Richard, I.; Bashir, R.; Britton, S.; Keers, S.M.; Oeltjen, J.; Brown, H.E.; Marchand, S.; Bourg, N.; et al. Genomic organization of the dysferlin gene and novel mutations in Miyoshi myopathy. Neurology 2001, 57, 271–278. [Google Scholar] [CrossRef] [PubMed]
  14. Anderson, L.V.B.; Davison, K.; Moss, J.A.; Young, C.; Cullen, M.J.; Walsh, J.; Johnson, M.A.; Bashir, R.; Britton, S.; Keers, S.; et al. Dysferlin is a plasma membrane protein and is expressed early in human development. Hum. Mol. Genet. 1999, 8, 855–861. [Google Scholar] [CrossRef]
  15. Gallardo, E.; de Luna, N.; Diaz-Manera, J.; Rojas-Garcia, R.; Gonzalez-Quereda, L.; Flix, B.; de Morree, A.; van der Maarel, S.; Illa, I. Comparison of Dysferlin Expression in Human Skeletal Muscle with That in Monocytes for the Diagnosis of Dysferlin Myopathy. PLoS ONE 2011, 6, e0029061. [Google Scholar] [CrossRef] [PubMed]
  16. Harsini, F.M.; Chebrolu, S.; Fuson, K.L.; White, M.A.; Rice, A.M.; Sutton, R.B. FerA is a Membrane-Associating Four-Helix Bundle Domain in the Ferlin Family of Membrane-Fusion Proteins. Sci. Rep. 2018, 8, 10949. [Google Scholar] [CrossRef]
  17. Abdullah, N.; Padmanarayana, M.; Marty, N.J.; Johnson, C.P. Quantitation of the calcium and membrane binding properties of the C2 domains of dysferlin. Biophys. J. 2014, 106, 382–389. [Google Scholar] [CrossRef] [PubMed]
  18. Shao, X.; Davletov, B.A.; Sutton, R.B.; Sudhof, T.C.; Rizo, J. Bipartite Ca2+-binding motif in C2 domains of synaptotagmin and protein kinase C. Science 1996, 273, 248–251. [Google Scholar] [CrossRef]
  19. Marty, N.J.; Holman, C.L.; Abdullah, N.; Johnson, C.P. The C2 Domains of Otoferlin, Dysferlin, and Myoferlin Alter the Packing of Lipid Bilayers. Biochemistry 2013, 52, 5585–5592. [Google Scholar] [CrossRef]
  20. Sula, A.; Cole, A.R.; Yeats, C.; Orengo, C.; Keep, N.H. Crystal structures of the human Dysferlin inner DysF domain. BMC Struct. Biol. 2014, 14, 3. [Google Scholar] [CrossRef] [PubMed]
  21. Bansal, D.; Campbell, K.P. Dysferlin and the plasma membrane repair in muscular dystrophy. Trends Cell Biol. 2004, 14, 206–213. [Google Scholar] [CrossRef]
  22. Glover, L.; Brown, R.H. Dysferlin in membrane trafficking and patch repair. Traffic 2007, 8, 785–794. [Google Scholar] [CrossRef]
  23. Sharma, A.; Yu, C.; Leung, C.; Trane, A.; Lau, M.; Utokaparch, S.; Shaheen, F.; Sheibani, N.; Bernatchez, P. A new role for the muscle repair protein dysferlin in endothelial cell adhesion and angiogenesis. Arterioscler. Thromb. Vasc. Biol. 2010, 30, 2196–2204. [Google Scholar] [CrossRef] [PubMed]
  24. Lek, A.; Evesson, F.J.; Lemckert, F.A.; Redpath, G.M.; Lueders, A.K.; Turnbull, L.; Whitchurch, C.B.; North, K.N.; Cooper, S.T. Calpains, cleaved mini-dysferlinC72, and L-type channels underpin calcium-dependent muscle membrane repair. J. Neurosci. 2013, 33, 5085–5094. [Google Scholar] [CrossRef] [PubMed]
  25. Han, R. Muscle membrane repair and inflammatory attack in dysferlinopathy. Skelet. Muscle 2011, 1, 10. [Google Scholar] [CrossRef] [PubMed]
  26. Kerr, J.P.; Ziman, A.P.; Mueller, A.L.; Muriel, J.M.; Kleinhans-Welte, E.; Gumerson, J.D.; Vogel, S.S.; Ward, C.W.; Roche, J.A.; Bloch, R.J. Dysferlin stabilizes stress-induced Ca2+ signaling in the transverse tubule membrane. Proc. Natl. Acad. Sci. USA 2013, 110, 20831–20836. [Google Scholar] [CrossRef] [PubMed]
  27. Kerr, J.P.; Ward, C.W.; Bloch, R.J. Dysferlin at transverse tubules regulates Ca2+ homeostasis in skeletal muscle. Front. Physiol. 2014, 5, 77998. [Google Scholar] [CrossRef] [PubMed]
  28. Muriel, J.; Lukyanenko, V.; Kwiatkowski, T.; Bhattacharya, S.; Garman, D.; Weisleder, N.; Bloch, R.J. The C2 domains of dysferlin: Roles in membrane localization, Ca(2+) signalling and sarcolemmal repair. J. Physiol. 2022, 600, 1953–1968. [Google Scholar] [CrossRef]
  29. Wang, Y.; Tadayon, R.; Santamaria, L.; Mercier, P.; Forristal, C.J.; Shaw, G.S. Calcium binds and rigidifies the dysferlin C2A domain in a tightly coupled manner. Biochem. J. 2021, 478, 197–215. [Google Scholar] [CrossRef]
  30. Vincent, A.E.; Rosa, H.S.; Alston, C.L.; Grady, J.P.; Rygiel, K.A.; Rocha, M.C.; Barresi, R.; Taylor, R.W.; Turnbull, D.M. Dysferlin mutations and mitochondrial dysfunction. Neuromuscular Disord. 2016, 26, 782–788. [Google Scholar] [CrossRef]
  31. Codding, S.J.; Marty, N.; Abdullah, N.; Johnson, C.P. Dysferlin Binds SNAREs (Soluble N-Ethylmaleimide-sensitive Factor (NSF) Attachment Protein Receptors) and Stimulates Membrane Fusion in a Calcium-sensitive Manner. J. Biol. Chem. 2016, 291, 14575–14584. [Google Scholar] [CrossRef]
  32. Hofhuis, J.; Bersch, K.; Bussenschutt, R.; Drzymalski, M.; Liebetanz, D.; Nikolaev, V.; Wagner, S.; Maier, L.S.; Gartner, J.; Klinge, L.; et al. Dysferlin mediates membrane tubulation and links T-tubule biogenesis to muscular dystrophy. J. Cell Sci. 2017, 130, 841–852. [Google Scholar] [CrossRef]
  33. Hofhuis, J.; Bersch, K.; Wagner, S.; Molina, C.; Fakuade, F.E.; Iyer, L.M.; Streckfuss-Bomeke, K.; Toischer, K.; Zelarayan, L.C.; Voigt, N.; et al. Dysferlin links excitation-contraction coupling to structure and maintenance of the cardiac transverse-axial tubule system. Europace 2020, 22, 1119–1131. [Google Scholar] [CrossRef] [PubMed]
  34. Demonbreun, A.R.; Rossi, A.E.; Alvarez, M.G.; Swanson, K.E.; Deveaux, H.K.; Earley, J.U.; Hadhazy, M.; Vohra, R.; Walter, G.A.; Pytel, P.; et al. Dysferlin and myoferlin regulate transverse tubule formation and glycerol sensitivity. Am. J. Pathol. 2014, 184, 248–259. [Google Scholar] [CrossRef]
  35. Demonbreun, A.R.; Allen, M.V.; Warner, J.L.; Barefield, D.Y.; Krishnan, S.; Swanson, K.E.; Earley, J.U.; McNally, E.M. Enhanced Muscular Dystrophy from Loss of Dysferlin Is Accompanied by Impaired Annexin A6 Translocation after Sarcolemmal Disruption. Am. J. Pathol. 2016, 186, 1610–1622. [Google Scholar] [CrossRef]
  36. Demonbreun, A.R.; Quattrocelli, M.; Barefield, D.Y.; Allen, M.V.; Swanson, K.E.; McNally, E.M. An actin-dependent annexin complex mediates plasma membrane repair in muscle. J. Cell Biol. 2016, 213, 705–718. [Google Scholar] [CrossRef]
  37. Rezvanpour, A.; Shaw, G.S. Unique S100 target protein interactions. Gen. Physiol. Biophys. 2009, 28, F39–F46. [Google Scholar]
  38. Ampong, B.N.; Imamura, M.; Matsumiya, T.; Yoshida, M.; Takeda, S. Intracellular localization of dysferlin and its association with the dihydropyridine receptor. Acta Myol. 2005, 24, 134–144. [Google Scholar]
  39. Sinnreich, M.; Therrien, C.; Karpati, G. Lariat branch point mutation in the dysferlin gene with mild limb-girdle muscular dystrophy. Neurology 2006, 66, 1114–1116. [Google Scholar] [CrossRef] [PubMed]
  40. Cacciottolo, M.; Numitone, G.; Aurino, S.; Caserta, I.R.; Fanin, M.; Politano, L.; Minetti, C.; Ricci, E.; Piluso, G.; Angelini, C.; et al. Muscular dystrophy with marked Dysferlin deficiency is consistently caused by primary dysferlin gene mutations. Eur. J. Hum. Genet. 2011, 19, 974–980. [Google Scholar] [CrossRef] [PubMed]
  41. Krahn, M. The UMD-DYSF Locus-Specific Database. [Web Page] 2011 06/26/2015. Available online: http://www.umd.be/DYSF/ (accessed on 15 September 2018).
  42. Izumi, R.; Takahashi, T.; Suzuki, N.; Niihori, T.; Ono, H.; Nakamura, N.; Katada, S.; Kato, M.; Warita, H.; Tateyama, M.; et al. The genetic profile of dysferlinopathy in a cohort of 209 cases: Genotype-phenotype relationship and a hotspot on the inner DysF domain. Hum. Mutat. 2020, 41, 1540–1554. [Google Scholar] [CrossRef]
  43. Krahn, M.; Beroud, C.; Labelle, V.; Nguyen, K.; Bernard, R.; Bassez, G.; Figarella-Branger, D.; Fernandez, C.; Bouvenot, J.; Richard, I.; et al. Analysis of the DYSF mutational spectrum in a large cohort of patients. Hum. Mutat. 2009, 30, E345–E375. [Google Scholar] [CrossRef]
  44. Zhong, H.; Yu, M.; Lin, P.; Zhao, Z.; Zheng, X.; Xi, J.; Zhu, W.; Zheng, Y.; Zhang, W.; Lv, H.; et al. Molecular landscape of DYSF mutations in dysferlinopathy: From a Chinese multicenter analysis to a worldwide perspective. Hum. Mutat. 2021, 42, 1615–1623. [Google Scholar] [CrossRef] [PubMed]
  45. Jin, S.Q.; Yu, M.; Zhang, W.; Lyu, H.; Yuan, Y.; Wang, Z.X. Dysferlin Gene Mutation Spectrum in a Large Cohort of Chinese Patients with Dysferlinopathy. Chinese Med. J.-Peking. 2016, 129, 2287. [Google Scholar] [CrossRef] [PubMed]
  46. Hu, Y.Y.; Lian, Y.J.; Xu, H.L.; Zheng, Y.K.; Li, C.F.; Zhang, J.W.; Yan, S.P. Novel, de novo dysferlin gene mutations in a patient with Miyoshi myopathy. Neurosci. Lett. 2018, 664, 107–109. [Google Scholar] [CrossRef] [PubMed]
  47. Kobayashi, K.; Izawa, T.; Kuwamura, M.; Yamate, J. Dysferlin and animal models for dysferlinopathy. J. Toxicol. Pathol. 2012, 25, 135–147. [Google Scholar] [CrossRef] [PubMed]
  48. Rayavarapu, S.; Van der Meulen, J.H.; Gordish-Dressman, H.; Hoffman, E.P.; Nagaraju, K.; Knoblach, S.M. Characterization of dysferlin deficient SJL/J mice to assess preclinical drug efficacy: Fasudil exacerbates muscle disease phenotype. PLoS ONE 2010, 5, e12981. [Google Scholar] [CrossRef]
  49. Vafiadaki, E.; Reis, A.; Keers, S.; Harrison, R.; Anderson, L.V.; Raffelsberger, T.; Ivanova, S.; Hoger, H.; Bittner, R.E.; Bushby, K.; et al. Cloning of the mouse dysferlin gene and genomic characterization of the SJL-Dysf mutation. Neuroreport 2001, 12, 625–629. [Google Scholar] [CrossRef]
  50. The Jackson Laboratory. Mouse Strain Datasheet. Available online: https://www.jax.org(accessed on 11 August 2018).
  51. Bulfield, G.; Siller, W.G.; Wight, P.A.; Moore, K.J. X chromosome-linked muscular dystrophy (mdx) in the mouse. Proc. Natl. Acad. Sci. USA 1984, 81, 1189–1192. [Google Scholar] [CrossRef]
  52. Malcher, J.; Heidt, L.; Goyenvalle, A.; Escobar, H.; Marg, A.; Beley, C.; Benchaouir, R.; Bader, M.; Spuler, S.; Garcia, L.; et al. Exon Skipping in a Dysf-Missense Mutant Mouse Model. Mol. Ther. Nucleic Acids 2018, 13, 198–207. [Google Scholar] [CrossRef]
  53. Kokubu, Y.; Nagino, T.; Sasa, K.; Oikawa, T.; Miyake, K.; Kume, A.; Fukuda, M.; Fuse, H.; Tozawa, R.; Sakurai, H. Phenotypic Drug Screening for Dysferlinopathy Using Patient-Derived Induced Pluripotent Stem Cells. Stem Cells Transl. Med. 2019, 8, 1017–1029. [Google Scholar] [CrossRef]
  54. Kesari, A.; Fukuda, M.; Knoblach, S.; Bashir, R.; Nader, G.A.; Rao, D.; Nagaraju, K.; Hoffman, E.P. Dysferlin deficiency shows compensatory induction of Rab27A/Slp2a that may contribute to inflammatory onset. Am. J. Pathol. 2008, 173, 1476–1487. [Google Scholar] [CrossRef]
  55. Nagaraju, K.; Rawat, R.; Veszelovszky, E.; Thapliyal, R.; Kesari, A.; Sparks, S.; Raben, N.; Plotz, P.; Hoffman, E.P. Dysferlin deficiency enhances monocyte phagocytosis: A model for the inflammatory onset of limb-girdle muscular dystrophy 2B. Am. J. Pathol. 2008, 172, 774–785. [Google Scholar] [CrossRef] [PubMed]
  56. Nguyen, K.; Bassez, G.; Bernard, R.; Krahn, M.; Labelle, V.; Figarella-Branger, D.; Pouget, J.; Hammouda el, H.; Beroud, C.; Urtizberea, A.; et al. Dysferlin mutations in LGMD2B, Miyoshi myopathy, and atypical dysferlinopathies. Hum. Mutat. 2005, 26, 165. [Google Scholar] [CrossRef] [PubMed]
  57. Fanin, M.; Angelini, C. Muscle pathology in dysferlin deficiency. Neuropath Appl. Neuro 2002, 28, 461–470. [Google Scholar] [CrossRef] [PubMed]
  58. Harris, E.; Bladen, C.L.; Mayhew, A.; James, M.; Bettinson, K.; Moore, U.; Smith, F.E.; Rufibach, L.; Cnaan, A.; Bharucha-Goebel, D.X.; et al. The Clinical Outcome Study for dysferlinopathy: An international multicenter study. Neurol. Genet. 2016, 2, e89. [Google Scholar] [CrossRef] [PubMed]
  59. Dastur, R.S.; Gaitonde, P.S.; Kachwala, M.; Nallamilli, B.R.R.; Ankala, A.; Khadilkar, S.V.; Atchayaram, N.; Gayathri, N.; Meena, A.K.; Rufibach, L.; et al. Detection of Dysferlin Gene Pathogenic Variants in the Indian Population in Patients Predicted to have a Dysferlinopathy Using a Blood-based Monocyte Assay and Clinical Algorithm: A Model for Accurate and Cost-effective Diagnosis. Ann. Indian. Acad. Neur 2017, 20, 302–308. [Google Scholar] [CrossRef] [PubMed]
  60. Aoki , M.; Takahashi, T. Mutational and clinical features of Japanese patients with dysferlinopathy (Miyoshi myopathy and limb girdle muscular dystrophy type 2B). Rinsho Shinkeigaku 2005, 45, 938–942. [Google Scholar] [PubMed]
  61. Fanin, M.; Angelini, C. Progress and challenges in diagnosis of dysferlinopathy. Muscle Nerve 2016, 54, 821–835. [Google Scholar] [CrossRef]
  62. Anderson, L.V.; Harrison, R.M.; Pogue, R.; Vafiadaki, E.; Pollitt, C.; Davison, K.; Moss, J.A.; Keers, S.; Pyle, A.; Shaw, P.J.; et al. Secondary reduction in calpain 3 expression in patients with limb girdle muscular dystrophy type 2B and Miyoshi myopathy (primary dysferlinopathies). Neuromuscul. Disord. 2000, 10, 553–559. [Google Scholar] [CrossRef] [PubMed]
  63. Therrien, C.; Dodig, D.; Karpati, G.; Sinnreich, M. Mutation impact on dysferlin inferred from database analysis and computer-based structural predictions. J. Neurol. Sci. 2006, 250, 71–78. [Google Scholar] [CrossRef]
  64. Clinicaltrials.gov. rAAVrh74.MHCK7.DYSF.DV for Treatment of Dysferlinopathies. The Proposed Clinical Trial is a Double-Blind, Randomized Controlled Study with Direct Intramuscular Injection of rAAVrh.74.MHCK7.DYSF.DV Gene Vector to the Extensor Digitorum Brevis Muscle (EDB). Two Cohorts of Subjects with Dysferlin Deficiency, Each with Proven Mutations will Undergo Gene Transfer. A Minimum of Three Subjects will be Enrolled into Each Cohort.]. 2021. Available online: https://classic.clinicaltrials.gov/ct2/show/NCT02710500 (accessed on 7 March 2024).
  65. Mendell, J.R.; Al-Zaidy, S.A.; Rodino-Klapac, L.R.; Goodspeed, K.; Gray, S.J.; Kay, C.N.; Boye, S.L.; Boye, S.E.; George, L.A.; Salabarria, S.; et al. Current Clinical Applications of In Vivo Gene Therapy with AAVs. Mol. Ther. 2021, 29, 464–488. [Google Scholar] [CrossRef]
  66. Allocca, M.; Doria, M.; Petrillo, M.; Colella, P.; Garcia-Hoyos, M.; Gibbs, D.; Kim, S.R.; Maguire, A.; Rex, T.S.; Di Vicino, U.; et al. Serotype-dependent packaging of large genes in adeno-associated viral vectors results in effective gene delivery in mice. J. Clin. Investig. 2008, 118, 1955–1964. [Google Scholar] [CrossRef] [PubMed]
  67. Lostal, W.; Bartoli, M.; Bourg, N.; Roudaut, C.; Bentaïb, A.; Miyake, K.; Guerchet, N.; Fougerousse, F.; McNeil, P.; Richard, I. Efficient recovery of dysferlin deficiency by dual adeno-associated vector-mediated gene transfer. Hum. Mol. Genet. 2010, 19, 1897–1907. [Google Scholar] [CrossRef] [PubMed]
  68. Sondergaard, P.C.; Griffin, D.A.; Pozsgai, E.R.; Johnson, R.W.; Grose, W.E.; Heller, K.N.; Shontz, K.M.; Montgomery, C.L.; Liu, J.; Clark, K.R.; et al. AAV.Dysferlin Overlap Vectors Restore Function in Dysferlinopathy Animal Models. Ann. Clin. Transl. Neurol. 2015, 2, 256–270. [Google Scholar] [CrossRef] [PubMed]
  69. Potter, R.A.; Griffin, D.A.; Sondergaard, P.C.; Johnson, R.W.; Pozsgai, E.R.; Heller, K.N.; Peterson, E.L.; Lehtimaki, K.K.; Windish, H.P.; Mittal, P.J.; et al. Systemic Delivery of Dysferlin Overlap Vectors Provides Long-Term Gene Expression and Functional Improvement for Dysferlinopathy. Hum. Gene Ther. 2018, 29, 749–762. [Google Scholar] [CrossRef] [PubMed]
  70. Grose, W.E.; Clark, K.R.; Griffin, D.; Malik, V.; Shontz, K.M.; Montgomery, C.L.; Lewis, S.; Brown, R.H.; Janssen, P.M.L.; Mendell, J.R.; et al. Homologous Recombination Mediates Functional Recovery of Dysferlin Deficiency following AAV5 Gene Transfer. PLoS ONE 2012, 7, e0039233. [Google Scholar] [CrossRef]
  71. Pryadkina, M.; Lostal, W.; Bourg, N.; Charton, K.; Roudaut, C.; Hirsch, M.L.; Richard, I. A comparison of AAV strategies distinguishes overlapping vectors for efficient systemic delivery of the 6.2 kb Dysferlin coding sequence. Mol. Ther.-Meth Clin. D 2015, 2, 15009. [Google Scholar] [CrossRef] [PubMed]
  72. Clinicaltrials.gov. A Gene Transfer Study to Evaluate the Safety, Tolerability and Efficacy of SRP-6004 in Ambulatory Participants with Limb Girdle Muscular Dystrophy, Type 2B/R2 (LGMD2B/R2, Dysferlin [DYSF] Related). 2023. Available online: https://clinicaltrials.gov/study/NCT05906251#publications (accessed on 7 March 2024).
  73. Shirley, J.L.; de Jong, Y.P.; Terhorst, C.; Herzog, R.W. Immune Responses to Viral Gene Therapy Vectors. Mol. Ther. 2020, 28, 709–722. [Google Scholar] [CrossRef] [PubMed]
  74. Arnold, C. Record number of gene-therapy trials, despite setbacks. Nat. Med. 2021, 27, 1312–1315. [Google Scholar] [CrossRef] [PubMed]
  75. Venditti, C.P. Safety questions for AAV gene therapy. Nat. Biotechnol. 2021, 39, 24–26. [Google Scholar] [CrossRef]
  76. Wills, C.A.; Drago, D.; Pietrusko, R.G. Clinical holds for cell and gene therapy trials: Risks, impact, and lessons learned. Mol. Ther. Methods Clin. Dev. 2023, 31, 101125. [Google Scholar] [CrossRef]
  77. Shen, B.; Zhang, W.; Zhang, J.; Zhou, J.; Wang, J.; Chen, L.; Wang, L.; Hodgkins, A.; Iyer, V.; Huang, X.; et al. Efficient genome modification by CRISPR-Cas9 nickase with minimal off-target effects. Nat. Methods 2014, 11, 399–402. [Google Scholar] [CrossRef] [PubMed]
  78. Horodecka, K.; Duchler, M. CRISPR/Cas9: Principle, Applications, and Delivery through Extracellular Vesicles. Int. J. Mol. Sci. 2021, 22, 6072. [Google Scholar] [CrossRef] [PubMed]
  79. Ran, F.A.; Hsu, P.D.; Wright, J.; Agarwala, V.; Scott, D.A.; Zhang, F. Genome engineering using the CRISPR-Cas9 system. Nat. Protoc. 2013, 8, 2281–2308. [Google Scholar] [CrossRef] [PubMed]
  80. Turan, S.; Farruggio, A.P.; Srifa, W.; Day, J.W.; Calos, M.P. Precise Correction of Disease Mutations in Induced Pluripotent Stem Cells Derived From Patients With Limb Girdle Muscular Dystrophy. Mol. Ther. 2016, 24, 685–696. [Google Scholar] [CrossRef] [PubMed]
  81. Mou, H.; Smith, J.L.; Peng, L.; Yin, H.; Moore, J.; Zhang, X.O.; Song, C.Q.; Sheel, A.; Wu, Q.; Ozata, D.M.; et al. CRISPR/Cas9-mediated genome editing induces exon skipping by alternative splicing or exon deletion. Genome Biol. 2017, 18, 108. [Google Scholar] [CrossRef] [PubMed]
  82. Malik, V.; Rodino-Klapac, L.R.; Viollet, L.; Wall, C.; King, W.; Al-Dahhak, R.; Lewis, S.; Shilling, C.J.; Kota, J.; Serrano-Munuera, C.; et al. Gentamicin-induced readthrough of stop codons in Duchenne muscular dystrophy. Ann. Neurol. 2010, 67, 771–780. [Google Scholar] [CrossRef] [PubMed]
  83. Berger, J.; Li, M.; Berger, S.; Meilak, M.; Rientjes, J.; Currie, P.D. Effect of Ataluren on dystrophin mutations. J. Cell Mol. Med. 2020, 24, 6680–6689. [Google Scholar] [CrossRef] [PubMed]
  84. Leung, A.; Sacristan-Reviriego, A.; Perdigao, P.R.L.; Sai, H.; Georgiou, M.; Kalitzeos, A.; Carr, A.F.; Coffey, P.J.; Michaelides, M.; Bainbridge, J.; et al. Investigation of PTC124-mediated translational readthrough in a retinal organoid model of AIPL1-associated Leber congenital amaurosis. Stem Cell Rep. 2022, 17, 2187–2202. [Google Scholar] [CrossRef] [PubMed]
  85. Sermet-Gaudelus, I.; Boeck, K.D.; Casimir, G.J.; Vermeulen, F.; Leal, T.; Mogenet, A.; Roussel, D.; Fritsch, J.; Hanssens, L.; Hirawat, S.; et al. Ataluren (PTC124) induces cystic fibrosis transmembrane conductance regulator protein expression and activity in children with nonsense mutation cystic fibrosis. Am. J. Respir. Crit. Care Med. 2010, 182, 1262–1272. [Google Scholar] [CrossRef]
  86. Samanta, A.; Stingl, K.; Kohl, S.; Ries, J.; Linnert, J.; Nagel-Wolfrum, K. Ataluren for the Treatment of Usher Syndrome 2A Caused by Nonsense Mutations. Int. J. Mol. Sci. 2019, 20, 6274. [Google Scholar] [CrossRef]
  87. Huang, S.; Bhattacharya, A.; Ghelfi, M.D.; Li, H.; Fritsch, C.; Chenoweth, D.M.; Goldman, Y.E.; Cooperman, B.S. Ataluren binds to multiple protein synthesis apparatus sites and competitively inhibits release factor-dependent termination. Nat. Commun. 2022, 13, 2413. [Google Scholar] [CrossRef] [PubMed]
  88. Seo, K.; Kim, E.K.; Choi, J.; Kim, D.S.; Shin, J.H. Functional recovery of a novel knockin mouse model of dysferlinopathy by readthrough of nonsense mutation. Mol. Ther. Methods Clin. Dev. 2021, 21, 702–709. [Google Scholar] [CrossRef] [PubMed]
  89. Tominaga, K.; Tominaga, N.; Williams, E.O.; Rufibach, L.; Schowel, V.; Spuler, S.; Viswanathan, M.; Guarente, L.P. 4-Phenylbutyrate restores localization and membrane repair to human dysferlin mutations. iScience 2022, 25, 103667. [Google Scholar] [CrossRef] [PubMed]
  90. Li, D.H.; Mastaglia, F.L.; Fletcher, S.; Wilton, S.D. Precision Medicine through Antisense Oligonucleotide-Mediated Exon Skipping. Trends Pharmacol. Sci. 2018, 39, 982–994. [Google Scholar] [CrossRef] [PubMed]
  91. Liang, X.H.; Sun, H.; Nichols, J.G.; Crooke, S.T. RNase H1-Dependent Antisense Oligonucleotides Are Robustly Active in Directing RNA Cleavage in Both the Cytoplasm and the Nucleus. Mol. Ther. 2017, 25, 2075–2092. [Google Scholar] [CrossRef]
  92. Kim, J.; Hu, C.; Moufawad El Achkar, C.; Black, L.E.; Douville, J.; Larson, A.; Pendergast, M.K.; Goldkind, S.F.; Lee, E.A.; Kuniholm, A.; et al. Patient-Customized Oligonucleotide Therapy for a Rare Genetic Disease. N. Engl. J. Med. 2019, 381, 1644–1652. [Google Scholar] [CrossRef] [PubMed]
  93. Flynn, L.L.; Mitrpant, C.; Adams, A.; Pitout, I.L.; Stirnweiss, A.; Fletcher, S.; Wilton, S.D. Targeted SMN Exon Skipping: A Useful Control to Assess In Vitro and In Vivo Splice-Switching Studies. Biomedicines 2021, 9, 552. [Google Scholar] [CrossRef] [PubMed]
  94. Zaw, K.; Greer, K.; Aung-Htut, M.T.; Mitrpant, C.; Veedu, R.N.; Fletcher, S.; Wilton, S.D. Consequences of Making the Inactive Active Through Changes in Antisense Oligonucleotide Chemistries. Front. Genet. 2019, 10, 1249. [Google Scholar] [CrossRef]
  95. Hall, J. Future directions for medicinal chemistry in the field of oligonucleotide therapeutics. RNA 2023, 29, 423–433. [Google Scholar] [CrossRef]
  96. Lim, K.R.; Maruyama, R.; Yokota, T. Eteplirsen in the treatment of Duchenne muscular dystrophy. Drug Des. Devel Ther. 2017, 11, 533–545. [Google Scholar] [CrossRef]
  97. Mendell, J.R.; Rodino-Klapac, L.R.; Sahenk, Z.; Roush, K.; Bird, L.; Lowes, L.P.; Alfano, L.; Gomez, A.M.; Lewis, S.; Kota, J.; et al. Eteplirsen for the treatment of Duchenne muscular dystrophy. Ann. Neurol. 2013, 74, 637–647. [Google Scholar] [CrossRef] [PubMed]
  98. Syed, Y.Y. Eteplirsen: First Global Approval. Drugs 2016, 76, 1699–1704. [Google Scholar] [CrossRef] [PubMed]
  99. Heo, Y.A. Golodirsen: First Approval. Drugs 2020, 80, 329–333. [Google Scholar] [CrossRef]
  100. Dhillon, S. Viltolarsen: First Approval. Drugs 2020, 80, 1027–1031. [Google Scholar] [CrossRef] [PubMed]
  101. Wilton-Clark, H.; Yokota, T. Casimersen for Duchenne muscular dystrophy. Drugs Today Barc 2021, 57, 707–717. [Google Scholar] [CrossRef]
  102. Aartsma-Rus, A. FDA Approval of Nusinersen for Spinal Muscular Atrophy Makes 2016 the Year of Splice Modulating Oligonucleotides. Nucleic Acid Ther 2017, 27, 67–69. [Google Scholar] [CrossRef]
  103. Hoy, S.M. Nusinersen: First Global Approval. Drugs 2017, 77, 473–479. [Google Scholar] [CrossRef]
  104. Wurster, C.D.; Ludolph, A.C. Nusinersen for spinal muscular atrophy. Ther. Adv. Neurol. Disord. 2018, 11, 1756285618754459. [Google Scholar] [CrossRef]
  105. Pfaff, A.L.; Singleton, L.M.; Koks, S. Mechanisms of disease-associated SINE-VNTR-Alus. Exp. Biol. Med. Maywood 2022, 247, 756–764. [Google Scholar] [CrossRef]
  106. Barthelemy, F.; Blouin, C.; Wein, N.; Mouly, V.; Courrier, S.; Dionnet, E.; Kergourlay, V.; Mathieu, Y.; Garcia, L.; Butler-Browne, G.; et al. Exon 32 Skipping of Dysferlin Rescues Membrane Repair in Patients’ Cells. J. Neuromuscul. Dis. 2015, 2, 281–290. [Google Scholar] [CrossRef]
  107. Dominov, J.A.; Uyan, O.; Sapp, P.C.; McKenna-Yasek, D.; Nallamilli, B.R.; Hegde, M.; Brown, R.H., Jr. A novel dysferlin mutant pseudoexon bypassed with antisense oligonucleotides. Ann. Clin. Transl. Neurol. 2014, 1, 703–720. [Google Scholar] [CrossRef]
  108. Dominov, J.A.; Uyan, O.; McKenna-Yasek, D.; Nallamilli, B.R.R.; Kergourlay, V.; Bartoli, M.; Levy, N.; Hudson, J.; Evangelista, T.; Lochmuller, H.; et al. Correction of pseudoexon splicing caused by a novel intronic dysferlin mutation. Ann. Clin. Transl. Neur 2019, 6, 642–654. [Google Scholar] [CrossRef] [PubMed]
  109. Takeshima, Y.; Yagi, M.; Okizuka, Y.; Awano, H.; Zhang, Z.; Yamauchi, Y.; Nishio, H.; Matsuo, M. Mutation spectrum of the dystrophin gene in 442 Duchenne/Becker muscular dystrophy cases from one Japanese referral center. J. Hum. Genet. 2010, 55, 379–388. [Google Scholar] [CrossRef] [PubMed]
  110. Kuijper, E.C.; Bergsma, A.J.; Pijnappel, W.; Aartsma-Rus, A. Opportunities and challenges for antisense oligonucleotide therapies. J. Inherit. Metab. Dis. 2021, 44, 72–87. [Google Scholar] [CrossRef] [PubMed]
  111. Anwar, S.; Mir, F.; Yokota, T. Enhancing the Effectiveness of Oligonucleotide Therapeutics Using Cell-Penetrating Peptide Conjugation, Chemical Modification, and Carrier-Based Delivery Strategies. Pharmaceutics 2023, 15, 1130. [Google Scholar] [CrossRef] [PubMed]
  112. Agrawal, S. The Evolution of Antisense Oligonucleotide Chemistry-A Personal Journey. Biomedicines 2021, 9, 503. [Google Scholar] [CrossRef]
  113. Iribe, H.; Miyamoto, K.; Takahashi, T.; Kobayashi, Y.; Leo, J.; Aida, M.; Ui-Tei, K. Chemical Modification of the siRNA Seed Region Suppresses Off-Target Effects by Steric Hindrance to Base-Pairing with Targets. ACS Omega 2017, 2, 2055–2064. [Google Scholar] [CrossRef] [PubMed]
  114. Terada, C.; Oh, K.; Tsubaki, R.; Chan, B.; Aibara, N.; Ohyama, K.; Shibata, M.A.; Wada, T.; Harada-Shiba, M.; Yamayoshi, A.; et al. Dynamic and static control of the off-target interactions of antisense oligonucleotides using toehold chemistry. Nat. Commun. 2023, 14, 7972. [Google Scholar] [CrossRef]
  115. Maksudov, F.; Kliuchnikov, E.; Pierson, D.; Ujwal, M.L.; Marx, K.A.; Chanda, A.; Barsegov, V. Therapeutic phosphorodiamidate morpholino oligonucleotides: Physical properties, solution structures, and folding thermodynamics. Mol. Ther. Nucleic Acids 2023, 31, 631–647. [Google Scholar] [CrossRef]
  116. Langner, H.K.; Jastrzebska, K.; Caruthers, M.H. Synthesis and Characterization of Thiophosphoramidate Morpholino Oligonucleotides and Chimeras. J. Am. Chem. Soc. 2020, 142, 16240–16253. [Google Scholar] [CrossRef]
  117. Le, B.T.; Paul, S.; Jastrzebska, K.; Langer, H.; Caruthers, M.H.; Veedu, R.N. Thiomorpholino oligonucleotides as a robust class of next generation platforms for alternate mRNA splicing. Proc. Natl. Acad. Sci. USA 2022, 119, e2207956119. [Google Scholar] [CrossRef] [PubMed]
  118. Paul, S.; Caruthers, M.H. Synthesis of Backbone-Modified Morpholino Oligonucleotides Using Phosphoramidite Chemistry. Molecules 2023, 28, 5380. [Google Scholar] [CrossRef] [PubMed]
  119. Tran, P.; Weldemichael, T.; Liu, Z.; Li, H.Y. Delivery of Oligonucleotides: Efficiency with Lipid Conjugation and Clinical Outcome. Pharmaceutics 2022, 14, 342. [Google Scholar] [CrossRef] [PubMed]
  120. Ruseska, I.; Zimmer, A. Internalization mechanisms of cell-penetrating peptides. Beilstein J. Nanotechnol. 2020, 11, 101–123. [Google Scholar] [CrossRef] [PubMed]
  121. Liu, B.R.; Chiou, S.H.; Huang, Y.W.; Lee, H.J. Bio-Membrane Internalization Mechanisms of Arginine-Rich Cell-Penetrating Peptides in Various Species. Membranes 2022, 12, 88. [Google Scholar] [CrossRef] [PubMed]
  122. Ait Benichou, S.; Jauvin, D.; De Serres-Berard, T.; Bennett, F.; Rigo, F.; Gourdon, G.; Boutjdir, M.; Chahine, M.; Puymirat, J. Enhanced Delivery of Ligand-Conjugated Antisense Oligonucleotides (C16-HA-ASO) Targeting Dystrophia Myotonica Protein Kinase Transcripts for the Treatment of Myotonic Dystrophy Type 1. Hum. Gene Ther. 2022, 33, 810–820. [Google Scholar] [CrossRef] [PubMed]
  123. Jearawiriyapaisarn, N.; Moulton, H.M.; Buckley, B.; Roberts, J.; Sazani, P.; Fucharoen, S.; Iversen, P.L.; Kole, R. Sustained dystrophin expression induced by peptide-conjugated morpholino oligomers in the muscles of mdx mice. Mol. Ther. 2008, 16, 1624–1629. [Google Scholar] [CrossRef] [PubMed]
  124. Moulton, H.M.; Moulton, J.D. Morpholinos and their peptide conjugates: Therapeutic promise and challenge for Duchenne muscular dystrophy. Biochim. Biophys. Acta 2010, 1798, 2296–2303. [Google Scholar] [CrossRef] [PubMed]
  125. Sheikh, O.; Yokota, T. Pharmacology and toxicology of eteplirsen and SRP-5051 for DMD exon 51 skipping: An update. Arch. Toxicol. 2021, 96, 1–9. [Google Scholar] [CrossRef]
  126. Schoenmaker, L.; Witzigmann, D.; Kulkarni, J.A.; Verbeke, R.; Kersten, G.; Jiskoot, W.; Crommelin, D.J.A. mRNA-lipid nanoparticle COVID-19 vaccines: Structure and stability. Int. J. Pharm. 2021, 601, 120586. [Google Scholar] [CrossRef]
  127. Hald Albertsen, C.; Kulkarni, J.A.; Witzigmann, D.; Lind, M.; Petersson, K.; Simonsen, J.B. The role of lipid components in lipid nanoparticles for vaccines and gene therapy. Adv. Drug Deliv. Rev. 2022, 188, 114416. [Google Scholar] [CrossRef] [PubMed]
  128. Calero, M.; Moleiro, L.H.; Sayd, A.; Dorca, Y.; Miquel-Rio, L.; Paz, V.; Robledo-Montana, J.; Enciso, E.; Accion, F.; Herraez-Aguilar, D.; et al. Lipid nanoparticles for antisense oligonucleotide gene interference into brain border-associated macrophages. Front. Mol. Biosci. 2022, 9, 887678. [Google Scholar] [CrossRef] [PubMed]
  129. Wu, L.; Zhou, W.; Lin, L.; Chen, A.; Feng, J.; Qu, X.; Zhang, H.; Yue, J. Delivery of therapeutic oligonucleotides in nanoscale. Bioact. Mater. 2022, 7, 292–323. [Google Scholar] [CrossRef] [PubMed]
  130. Min, H.S.; Kim, H.J.; Naito, M.; Ogura, S.; Toh, K.; Hayashi, K.; Kim, B.S.; Fukushima, S.; Anraku, Y.; Miyata, K.; et al. Systemic Brain Delivery of Antisense Oligonucleotides across the Blood-Brain Barrier with a Glucose-Coated Polymeric Nanocarrier. Angew. Chem. Int. Ed. Engl. 2020, 59, 8173–8180. [Google Scholar] [CrossRef]
  131. Chu, M.L.; Moran, E. The Limb-Girdle Muscular Dystrophies: Is Treatment on the Horizon? Neurotherapeutics 2018, 15, 849–862. [Google Scholar] [CrossRef]
Figure 1. Architecture of the human dysferlin protein. C2A–C2G; Calcium-binding C2 domains, TM; transmembrane domain, Fer; Ferlin and dysF; dysferlin domains. The amino acid residue positions are indicated above the domain structure, with the encoding exons below. The dysferlin exon map with a reading frame is shown below the domain structure, with the rectangular blocks indicating exons with splice junctions occurring between codons, whereas exons with chevron sides indicate splice junctions that occur within a codon <1:2< or >2:1>. Adapted from Abdullah et al., 2014 [17], Sula et al., 2014 [20], and NM_003494.4.
Figure 1. Architecture of the human dysferlin protein. C2A–C2G; Calcium-binding C2 domains, TM; transmembrane domain, Fer; Ferlin and dysF; dysferlin domains. The amino acid residue positions are indicated above the domain structure, with the encoding exons below. The dysferlin exon map with a reading frame is shown below the domain structure, with the rectangular blocks indicating exons with splice junctions occurring between codons, whereas exons with chevron sides indicate splice junctions that occur within a codon <1:2< or >2:1>. Adapted from Abdullah et al., 2014 [17], Sula et al., 2014 [20], and NM_003494.4.
Ijms 25 05572 g001
Figure 2. The role of dysferlin in membrane repair in a calcium-dependent manner. (a) Normal sarcolemmal membrane showing the location of dysferlin (green), annexin (blue), MG53 (pink), and caveolin (grey). (b) Membrane damage causes high influx of calcium into the cytosol and membrane repair vesicles loaded with mini-dysferlin (dysferlin are cleaved by calpain to form mini-dysferlin) are brought near the site of injury by the interaction of multiple proteins as shown. (c) The membrane is repaired by sealing the lesion by patch formation. Modified from Renzhi Han 2011 and Lek et al., 2013 [24,25].
Figure 2. The role of dysferlin in membrane repair in a calcium-dependent manner. (a) Normal sarcolemmal membrane showing the location of dysferlin (green), annexin (blue), MG53 (pink), and caveolin (grey). (b) Membrane damage causes high influx of calcium into the cytosol and membrane repair vesicles loaded with mini-dysferlin (dysferlin are cleaved by calpain to form mini-dysferlin) are brought near the site of injury by the interaction of multiple proteins as shown. (c) The membrane is repaired by sealing the lesion by patch formation. Modified from Renzhi Han 2011 and Lek et al., 2013 [24,25].
Ijms 25 05572 g002
Figure 3. A schematic showing the protein structure of dysferlin and the mutation spectrum. The positions of each of the seven C2 domains (C2A–C2G), the FerA/B, DYSF, and transmembrane domains, and the exon arrangement are shown on the y-axis. The histogram represents the incidence of pathogenic mutations within each of the 55 exons encoding the canonical skeletal muscle isoform of dysferlin (data derived from the Leiden muscular dystrophy database) [8].
Figure 3. A schematic showing the protein structure of dysferlin and the mutation spectrum. The positions of each of the seven C2 domains (C2A–C2G), the FerA/B, DYSF, and transmembrane domains, and the exon arrangement are shown on the y-axis. The histogram represents the incidence of pathogenic mutations within each of the 55 exons encoding the canonical skeletal muscle isoform of dysferlin (data derived from the Leiden muscular dystrophy database) [8].
Ijms 25 05572 g003
Figure 4. Some possible therapeutic strategies at a glance. CRISPR/Cas9 mediated correction of mutations, antisense oligomer (AO) therapies (exon skipping, exon retention, splice correction), gene therapy, readthrough (red X is blocking of stop codon by Ataluren) and small molecule-based therapies have shown some potential to restore either fully or partially functional dysferlin expression.
Figure 4. Some possible therapeutic strategies at a glance. CRISPR/Cas9 mediated correction of mutations, antisense oligomer (AO) therapies (exon skipping, exon retention, splice correction), gene therapy, readthrough (red X is blocking of stop codon by Ataluren) and small molecule-based therapies have shown some potential to restore either fully or partially functional dysferlin expression.
Ijms 25 05572 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Poudel, B.H.; Fletcher, S.; Wilton, S.D.; Aung-Htut, M. Limb Girdle Muscular Dystrophy Type 2B (LGMD2B): Diagnosis and Therapeutic Possibilities. Int. J. Mol. Sci. 2024, 25, 5572. https://doi.org/10.3390/ijms25115572

AMA Style

Poudel BH, Fletcher S, Wilton SD, Aung-Htut M. Limb Girdle Muscular Dystrophy Type 2B (LGMD2B): Diagnosis and Therapeutic Possibilities. International Journal of Molecular Sciences. 2024; 25(11):5572. https://doi.org/10.3390/ijms25115572

Chicago/Turabian Style

Poudel, Bal Hari, Sue Fletcher, Steve D. Wilton, and May Aung-Htut. 2024. "Limb Girdle Muscular Dystrophy Type 2B (LGMD2B): Diagnosis and Therapeutic Possibilities" International Journal of Molecular Sciences 25, no. 11: 5572. https://doi.org/10.3390/ijms25115572

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

Article Metrics

Back to TopTop