Next Article in Journal
Structural Insight into Catalysis by the Flavin-Dependent NADH Oxidase (Pden_5119) of Paracoccus denitrificans
Next Article in Special Issue
APP in the Neuromuscular Junction for the Development of Sarcopenia and Alzheimer’s Disease
Previous Article in Journal
Dimeric Product of Peroxy Radical Self-Reaction Probed with VUV Photoionization Mass Spectrometry and Theoretical Calculations: The Case of C2H5OOC2H5
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Clinical and Pathologic Features of Congenital Myasthenic Syndromes Caused by 35 Genes—A Comprehensive Review

1
Division of Neurogenetics, Center for Neurological Diseases and Cancer, Nagoya University Graduate School of Medicine, Nagoya 466-8550, Japan
2
Department of Neurology and Neuromuscular Research Laboratory, Mayo Clinic, Rochester, MN 55905, USA
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(4), 3730; https://doi.org/10.3390/ijms24043730
Submission received: 28 January 2023 / Revised: 9 February 2023 / Accepted: 9 February 2023 / Published: 13 February 2023
(This article belongs to the Special Issue Neuromuscular Diseases: From Pathogenic Mechanisms to Therapy)

Abstract

:
Congenital myasthenic syndromes (CMS) are a heterogeneous group of disorders characterized by impaired neuromuscular signal transmission due to germline pathogenic variants in genes expressed at the neuromuscular junction (NMJ). A total of 35 genes have been reported in CMS (AGRN, ALG14, ALG2, CHAT, CHD8, CHRNA1, CHRNB1, CHRND, CHRNE, CHRNG, COL13A1, COLQ, DOK7, DPAGT1, GFPT1, GMPPB, LAMA5, LAMB2, LRP4, MUSK, MYO9A, PLEC, PREPL, PURA, RAPSN, RPH3A, SCN4A, SLC18A3, SLC25A1, SLC5A7, SNAP25, SYT2, TOR1AIP1, UNC13A, VAMP1). The 35 genes can be classified into 14 groups according to the pathomechanical, clinical, and therapeutic features of CMS patients. Measurement of compound muscle action potentials elicited by repetitive nerve stimulation is required to diagnose CMS. Clinical and electrophysiological features are not sufficient to identify a defective molecule, and genetic studies are always required for accurate diagnosis. From a pharmacological point of view, cholinesterase inhibitors are effective in most groups of CMS, but are contraindicated in some groups of CMS. Similarly, ephedrine, salbutamol (albuterol), amifampridine are effective in most but not all groups of CMS. This review extensively covers pathomechanical and clinical features of CMS by citing 442 relevant articles.

1. Overview of Congenital Myasthenic Syndromes (CMS)

CMS are caused by defects in molecules expressed at the neuromuscular junction (NMJ) and are characterized by defective neuromuscular signal transduction [1,2,3]. As of January 2023, germline pathogenic variants in 35 genes have been reported (AGRN, ALG14, ALG2, CHAT, CHD8, CHRNA1, CHRNB1, CHRND, CHRNE, CHRNG, COL13A1, COLQ, DOK7, DPAGT1, GFPT1, GMPPB, LAMA5, LAMB2, LRP4, MUSK, MYO9A, PLEC, PREPL, PURA, RAPSN, RPH3A, SCN4A, SLC18A3, SLC25A1, SLC5A7, SNAP25, SYT2, TOR1AIP1, UNC13A, and VAMP1) (Figure 1). The causative genes can be classified into 14 groups depending on the pathomechanical features. Clinical features and therapeutic strategies are shared between some or all groups of CMS, but some features and therapies are unique to specific groups of CMS. Some therapies are ineffective or even contraindicated in some groups. Clinical features and therapeutic responses are generally difficult to predict a defective molecule.
Clinical features of CMS are characterized by muscle fatigue, muscle weakness, muscle hypoplasia, and minor facial anomalies like low-set ears and high-arched palate in some patients. Autoimmune myasthenia gravis (MG) also compromises the NMJ signal transduction but is caused by autoantibodies against the acetylcholine receptor (AChR), muscle-specific receptor tyrosine kinase (MuSK), low-density lipoprotein receptor-related protein 4 (LRP4), or others. In contrast to MG, diurnal fluctuation of muscle strength and muscle fatigue are not always observed in CMS. In some CMS patients, day-to-day fluctuation of muscle strength is prominent. Diurnal fluctuation of external ophthalmoplegia is associated with diplopia in MG, but not always in CMS. This is likely due to the presence of external ophthalmoplegia since infancy, which enables compensation for the fluctuating visual axes. Some lack eye symptoms and are referred to as limb-girdle CMS. Most CMS patients develop the disease before age 2 years, but, in some patients, symptoms develop immediately after birth but temporarily subside thereafter until adolescence or adulthood. Including these neonatal transient patients, CMS can develop at any age including adolescence and adulthood. Some patients with CHAT-CMS, LAMB2-CMS, SLC5A7-CMS, SNAP25-CMS, UNC13A-CMS, DPAGT1-CMS, ALG2-CMS, MYO9A-CMS, SLC25A1-CMS, and PURA-CMS exhibit developmental delay. This can be caused by defective cholinergic neurotransmission in the central nervous system (CNS) or hypoxic brain injury due to repeated apneustic attacks, but the exact mechanisms remain to be elucidated. Episodic apnea is frequently reported in CHAT-CMS, COLQ-CMS, and SCN4A-CMS, but is also observed in other groups of CMS. Siblings of CMS patients with episodic apnea sometimes die with a diagnosis of sudden infantile death syndrome (SIDS) [4,5]. Continuous monitoring of apnea is required for these patients.
Most CMS patients show autosomal recessive inheritance or require biallelic pathogenic variants. An autosomal dominant inheritance or a de novo hemiallelic pathogenic variant is observed in slow-channel CMS (SCCMS), SNAP25-CMS, PURA-CMS, and 4 out of 11 patients with SYT2-CMS. SYT2-CMS, SNAP25-CMS, VAMP1-CMS, UNC13A-CMS, RPH3A-CMS, and LAMA5-CMS are characterized by defects in the SNARE complex, and phenotypically similar to Lambert-Eaton myasthenic syndrome (LEMS). Both LEMS and LEMS-like CMS show an increment of compound muscle action potentials (CMAP) in response to high-frequency repetitive nerve stimulation (RNS) or spontaneous muscle contractions. In addition, in five patients in three pedigrees with AGRN-CMS, which primarily shows endplate AChR deficiency, a marked increment of CMAP after exercise was reported [6], but not in other AGRN-CMS patients. GMPPB-CMS, GFPT1-CMS, and SCCMS show elevated serum creatine kinase (CK) levels up to 24 times the upper limit of the normal range [7,8].

2. Electrophysiology, Muscle Biopsy, Laboratory Examinations, Differential Diagnosis, Epidemiology, Inheritance, and Therapeutic Perspectives

2.1. Electrophysiological Examinations

RNS or single-fiber electromyography (SFEMG) is required to diagnose CMS. However, next-generation sequencing technologies have enabled extensive genetic analysis, and a plethora of CMS patients have been diagnosed and reported in the absence of RNS or SFEMG. RNS at 2–3 Hz shows 10% or more decrements of the compound muscle action potential (CMAP).
A single nerve stimulus elicits a repetitive CMAP (R-CMAP) in some patients with SCCMS, COLQ-CMS, and PURA-CMS. R-CMAP rapidly disappears by RNS or by spontaneous exercise, and a single nerve stimulus after a prolonged rest is required.
In SCN4A-CMS, a decrement of CMAP is not induced by low-frequency RNS but by high-frequency RNS. In CMS caused by defective recycling of acetylcholine (ACh) (CHAT-CMS, SLC18A3-CMS, SLC5A7-CMS, and PREPL-CMS), decremental CMAP by low-frequency RNS is elicited at rest in some patients, but only after exercise or high-frequency RNS in the other patients.
In LEMS-like CMS caused by SYT2-CMS [9], VAMP1-CMS [10], UNC13A-CMS [11], RPH3A-CMS [12], and LAMA5-CMS [13], low-frequency RNS causes a decremental CMAP, whereas high-frequency RNS elicits an incremental CMAP [14]. In another form of LEMS-like CMS of SNAP25-CMS, low-frequency RNS caused a decremental CMAP, but high-frequency RNS was not examined [15].
In performing RNS, it is essential to fix the recording electrodes well. Muscle twitch by the first electrical stimulus in RNS moves the recording electrodes and decreases the height and area of CMAP, which could be misdiagnosed as a decremental response. The movement of the recording electrodes can be easily detected by a change in the shape of CMAP from the second stimulus.
Single fiber electromyography (SFEMG) has a higher sensitivity than RNS to detect defective signal transmission at the NMJ but has a lower specificity. Although SFEMG is technically challenging, some neurophysiologists diagnosed a large number of CMS patients only by SFEMG [16].

2.2. Muscle Biopsy and Creatine Kinase (CK)

Muscle biopsy shows tubular aggregates or rimmed vacuoles in glycosylation defects in GFPT1-CMS [17,18,19,20], DPAGT1-CMS [21,22,23], ALG2-CMS [24], but not in ALG14-CMS [24] or GMPPB-CMS [25]. These muscle pathologies, however, are not always observed. In a case with ALG2-CMS, muscle biopsy at age 14 years showed no tubular aggregates [24]. On the other hand, endplate myopathy in SCCMS was reported as tubular aggregates at the light microscopy level [26] or inclusion body myositis (IBM)-type inclusions at the electron microscopy level [27]. Defects in GMPPB cause muscular dystrophy-dystroglycanopathy (MDDG) type 14 [28]. GMPPB-CMS similarly exhibits hypoglycosylation of α-dystroglycan, as well as muscular dystrophy [25]. GMPPB-CMS also exhibits fibrosis and adiposis of skeletal muscle by MRI [25], as well as centronuclear myopathy [29].
Serum CK levels are normal in most groups of CMS. However, serum CK levels are elevated ~1.5 times the upper limit of normal in endplate myopathies in SCCMS, ~3 times in tubular aggregates in GFPT1-CMS and DOK7-CMS, and 2-to-24 times (average 10.7 times) in GMPPB-CMS [7,8].

2.3. Differential Diagnosis

Abnormal muscle fatigue should be differentiated from MG and LEMS. Genetic analysis of 121 patients with MG with no anti-AChR or anti-MuSK antibodies revealed 9 patients with CHNRA1-, CHRNE-, and RAPSN-CMS [30,31,32]. Muscle hypoplasia should be differentiated from congenital myopathies and limb-girdle muscular dystrophies. As stated above, some CMS patients have elevated serum CK levels. CMS should be considered in patients not only with diurnal fluctuation of muscle weakness, but also with day-to-day fluctuation of muscle weakness as well as continuous muscle weakness. Interestingly, 10 patients with PREPL-CMS were first considered to be Prader-Willi syndrome [33]. Additionally, two pedigrees with SYT2-CMS were initially diagnosed as Charcot-Marie-Tooth disease and distal hereditary motor neuropathy, respectively [9].
In addition, natural and artificial toxins and drugs affect the NMJ signal transmission. For example, (i) a plant toxin, curare, and a snake toxin, α-bungarotoxin, block muscle nicotinic AChR (CRHNA1, CHRNB1, CHRND, CHRNE), (ii) a shell toxin, ⍵-conotoxin, blocks N-type calcium channel (CACNA1B) at the nerve terminal, (iii) a shell toxin, µ-conotoxin, blocks skeletal muscle sodium channel (NaV1.4, SCN4A), (iv) a spider toxin, α-latrotoxin, makes a cation-nonselective ion channel pore at the nerve terminal, which allows excessive influx of calcium ions, (v) a bacterial toxin, botulinum, degrades the SNARE complex at the nerve terminal, (vi) chemical weapons, sarin, soman, tabun, and VX, block acetylcholinesterase (AChE), (vii) a pesticide, organophosphate, also blocks AChE, (viii) an antibiotic, aminoglycoside, inhibits calcium uptake at the nerve terminal, (ix) excessive administration of cholinesterase inhibitor (ChEI) blocks AChE, and (x) spores of Clostridium botulinum in honey cause infantile botulism, which resembles CMS. The diagnosis of infantile botulism is supported by a self-limited course even when there is no apparent history of honey intake.
Arthrogryposis multiplex congenita (AMC) is caused by defects in more than 320 genes [34]. Pathogenic variants in CHRNG show AMC in the lack of myasthenia [35,36,37], and are observed in the largest number of AMC patients [38]. Pathogenic variants in CHRNA1 [39], CHRNB1 [39], CHRND [39], RAPSN [39,40], SLC18A3 [41], SNAP25 [15], and MYO9A [38] also cause AMC in some patients.

2.4. Epidemiology

Analysis of 123 CMS patients in UK showed that the prevalence of CMS under age 18 years largely differ in regions in UK ranging from 2.8 to 14.8 per million with an average of 9.2 per million [42]. This prevalence was about 6 times higher than the prevalence of 1.5 per million of juvenile MG in UK [42]. Similarly, analysis of 22 CMS patients in Brazil [43], 8 CMS patients in Slovenia [44], and 64 CMS patients in Spain [45] showed that the prevalence of CMS under age 18 years were 1.8, 22.2, and 1.8 per million, respectively. All the reports addressed that they underestimated the prevalence because of the presence of undiagnosed CMS patients.
In the 35 causative genes for CMS, pathogenic variants have been frequently observed in genes for AChR ε subunit (CHRNE), collagen Q (COLQ), rapsyn (RAPSN), Dok-7 (DOK7), and glutamine--fructose-6-phosphate transaminase 1 (GFPT1). Founder effects have been reported in RAPSN p.Asn88Lys [46,47,48,49], DOK7 c.1124_1127dupTGCC [50], CHRNE c.1327delG [51], GMPPB c.1000G>A (p.Asp334Asn) [52], and PLEC c.1_9del (p.Met1_Gly3del) [53].

2.5. Inheritance

Autosomal dominant inheritance or hemiallelic pathogenic variants are observed in SCCMS, SNAP25-CMS [15,54], PURA-CMS [55], and some [9,56,57] but not the other [58,59,60] patients of SYT2-CMS. In contrast, other groups of CMS show autosomal recessive inheritance or require pathogenic loss-of-function variants in two alleles. SCCMS is caused by a gain of function missense variant in a single allele, because prolonged AChR channel openings in half of AChRs at the NMJ are sufficient to cause the SCCMS. Both SYT2-CMS and SNAP25-CMS exhibit LEMS-like CMS, and are likely to be caused by dominant negative effects. PURA-CMS is likely to be caused by a hemiallelic loss-of-function.

2.6. Therapeutic Perspectives

Therapeutic strategies for CMS include ChEIs, ephedrine, salbutamol (albuterol in US), amifampridine (3,4-diaminopyridine), quinidine, fluoxetine, and acetazolamide [61]. ChEIs (e.g., pyridostigmine) are effective in many groups of CMS, but are generally ineffective for SCCMS and DOK7-CMS. In addition, ChEIs are contraindicated for COLQ-CMS [62,63,64] and LAMB2-CMS [65], because of severe adverse effects including respiratory arrest in some patients. Although the underlying mechanisms remain unknown, ChEIs sometimes worsen symptoms in DOK7-CMS [64,66,67,68,69], MUSK-CMS [70], and LRP4-CMS [71].
Ephedrine and salbutamol (albuterol) are effective in many groups of CMS including endplate AChR deficiency caused by pathogenic variants in a large number of genes, as well as in DOK7-CMS. Sympathetic nerve innervates the NMJ and facilitates the NMJ signal transmission, which is likely to be a pharmaceutical mechanism of the effects of ephedrine and salbutamol (albuterol) [72]. Ephedrine and salbutamol (albuterol) are also effective in some patients with SCCMS and COLQ-CMS, which is likely to compensate for AChR deficiency due to endplate myopathy.
Amifampridine blocks voltage-gated potassium channel at the nerve terminal to potentiate the action potential of the motor nerve, and enhances calcium entry into the nerve terminal, which subsequently facilitates release of ACh into the synaptic space. Amifampridine is effective for LEMS-like CMS, which is characterized by compromised SNARE complex. In addition, amifampridine is effective for many groups of CMS except for SCCMS, AGRN-CMS, SLC5A7-CMS, SLC25A1-CMS. Amifampridine is also effective in some patients with COLQ-CMS [62,73], although the pharmacological mechanisms remain unknown.
Quinidine [74,75] and fluoxetine [76] ameliorate SCCMS. In a case of RAPSN-CMS, fluoxetine prescribed for depression worsened myasthenia [77]. A marked effect of fluoxetine was reported in a case of COLQ-CMS [78].
Acetazolamide was effective in two patients of SCN4A-CMS [79,80], but was not in another SCN4A-CMS [81].
In 27 pregnancies in 16 CMS patients, all patients continued to take drugs. The symptoms were worsened in 63% of the pregnancies but were subsided after delivery [82].

3. Physiological Aspects of Neuromuscular Signal Transmission

It is essential to understand the physiology of signal transduction at the NMJ to recognize the pathomechanisms of CMS (Figure 1). The action potential of the spinal motor neuron is delivered to the nerve terminal, and activates the P/Q-type calcium channel (CACNA1A). The calcium ions bind to two C2B domains of synaptotagmin 2 (SYT2), and activate the SNARE complex to fuse synaptic vesicles containing ACh to the presynaptic membrane [83]. ACh is then released to the 70-nm synaptic space. ACh released from the nerve terminal is hydrolyzed by AChE in synaptic space, and two molecules of ACh that were not captured by AChE bind to AChR to open a cationic ion channel pore. ACh dissociated from AChR is hydrolyzed to choline by AChE. The generated choline in the synaptic space is up taken by high affinity choline transporter (ChT, SLC5A7) expressed in the membrane of the nerve terminal [84]. Choline acetyltransferase (ChAT, CHAT) in the nerve terminal generates ACh from up taken choline and acetyl-CoA. Vacuolar proton ATPase embedded in the synaptic vesicle generates a proton gradient, which drives an import of ACh into the synaptic vesicle through vesicular acetylcholine transporter (vAChT, SLC18A3) [85].
Adult-type AChR is comprised of the α1 (CHRNA1), β1 (CHRNB1), δ (CHRND), and ε (CHRNE) subunits (Figure 2). Two α1 subunits and each of β1, δ, and ε subunits make a pentameric AChR (α12β1δε). Embryonic AChR of a α12β1δγ pentamer will be addressed in a section for endplate AChR deficiency. AChR subunits have four transmembrane domains (M1, M2, M3, and M4), and their N- and C-terminals are on the extracellular side. The second transmembrane domain, M2, makes an ion channel pore. The N-terminal regions of AChR subunits make a large extracellular complex, and ACh binds to the interfaces between α1–ε subunits and α1–δ subunits. AChR is a cation-nonselective ion channel that can pass through all cationic ions of Na+, Ca2+, and Mg2+. As Na+ is the major cation in the extracellular space, Na+ is the major source to make an endplate potential (EPP). The conductance and the burst duration of fetal AChR are ~70% and ~240% of those of adult-type AChR at the human endplate [86].
Depolarization by EPP elicits the opening of skeletal muscle voltage-gated sodium channel (NaV1.4, SCN4A) to generate a muscle action potential. NaV1.4 is expressed throughout the muscle fiber but is enriched at the motor endplate. Muscle action potential goes into the T tubules, where depolarization is sensed by L-type calcium channel (CaV1.1, CACNA1S), which constitutes dihydropyridine receptor (DHPR) with other molecules. DHPR is coupled to ryanodine receptor (RyR, RYR1), and RyR releases Ca2+ from the sarcoplasmic reticulum (SR). Sarcoplasmic Ca2+ binds to troponin and displaces tropomyosin that covers the binding sites of actin for the myosin head to contract the muscle fibers.
Many molecules drive AChR clustering at the motor endplate to enable finely tuned signal transmission at the NMJ. Agrin (AGRN) released from the nerve terminal binds to LRP4 (LRP4) at the motor endplate [88,89]. Two LRP4 molecules bind to two molecules of MuSK (MUSK) to make a hetero tetrameric receptor complex. Agrin binds to LRP4 and induces MuSK phosphorylation. MuSK phosphorylation is enhanced by Dok-7 (DOK7) [90]. Phosphorylated MuSK then phosphorylates AChR β1 subunit (CHRNB1), which binds to submembranous structural protein, rapsyn (RAPSN), with a stoichiometry of 2:1 or 1:1 to make AChR clusters at the motor endplate [91]. Rapsyn makes membraneless condensates by phase separation to anchor AChR [92]. Rapsyn binds to β catenin (CTNNB1) and chromodomain helicase DNA binding protein 8 (CHD8, CHD8) to reinforce the rapsyn network, which is enhanced by Wnt. Wnt binds to the frizzled-like domain of MuSK and also increases β catenin for the reinforcement of rapsyn. LRP4 is a receptor for agrin on the motor endplate, but also mediates a retrograde signal from the motor endplate to the nerve terminal [93,94]. In addition, we reported that Rspo2 [95,96], Fgf18 [97], and Ctgf/Ccn2 [98] are secreted molecules at the NMJ to enhance the agrin-LRP4-MuSK signaling and the formation of the NMJ [99] (Figure 3).

4. Thirty-Five Genes in 14 Groups of CMS

CMS are caused by 35 genes, which can be grouped into 14 groups, based on pathomechanisms. Pathomechanisms, clinical features, and therapies are widely variable from category to category.

4.1. Endplate AChR Deficiency (CHRNA1, CHRNB1, CHRND, CHRNE, and RAPSN)

4.1.1. Pathomechanisms

Embryonic AChR is composed of α12β1δγ. The embryonic γ subunit is substituted by the adult-type ε subunit after birth to make α12β1δε. The adult-type α12β1δε has a higher conductance and a shorter opening time compared to the embryonic α12β1δγ [86]. Biallelic lack of the ε subunit or (CHRNE) in CMS patients can be compensated for by the γ subunit and is not fatal [16,103,104,105]. The expression of γ-AChR is also observed when the expression of ε-AChR is markedly reduced. In contrast, biallelic lack of the α1, β1, and δ subunits (CHRNA1, CHRNB1, and CHRND, respectively) cannot be compensated for by another subunit, and is fatal. Hemiallelic null variants in CHRNA1, CHRNB1, and CHRND are asymptomatic if the other allele has no pathogenic variant, whereas biallelic null variants are observed only in CHRNE. Missense variants in genes encoding the α1, β1, δ, and ε subunits (CHRNA1, CHRNB1, CHRND, and CHRNE, respectively) that markedly reduce the cell surface expression of AChR cause endplate AChR deficiency [106,107]. Some missense variants in these genes simultaneously cause endplate AChR deficiency, as well as slow or fast channel myasthenic syndrome (SCCMS or FCCMS). Especially, in FCCMS, the reduction in channel opening events to ~50% alone is not pathogenic, but becomes pathogenic when the expression level is also reduced to ~50% [108].
Aberrant splicing of CHRNA1 causes an unusual form of endplate AChR deficiency. CHRNA1 has a 75-nt exon P3A between exons 3 and 4 that is unique to human and anthropoids. An exon P3A-skipped CHRNA1 transcript makes normal AChR, whereas an exon P3A-included CHRNA1 transcript cannot form AChR. In human skeletal muscle, P3A(+) and P3A(-) transcripts are generated at a ratio of 1:1, although the physiological significance remains unknown. Pathogenic variants in exon P3A and its preceding intron exclusively include exon P3A in pre-mRNA splicing, and the generated P3A(+) transcript causes endplate AChR deficiency [109,110,111].
Pathogenic variants in RAPSN encoding rapsyn also cause endplate AChR deficiency. Some missense variants in RAPSN retain self-clustering of rapsyn [112], whereas the others do not [113]. Rapsyn phosphorylated by the agrin-LRP4-MuSK pathway forms submembranous network by self-clustering and also activates the E3 ligase activity, which is compromised by a founder variant, p.N88K in RAPSN [114]. Four siblings born from consanguineous parents carried pathogenic homovariants in RAPSN (c.491G>A, p.R146H) [115]. However, only two of them were affected by CMS, whereas the other two were not. The two affected CMS siblings additionally had homovariants in AK9. AK9 encodes one of nine adenylate kinases, and catalyzes a conversion between nucleotide diphosphate and nucleotide triphosphate. The identified variant in AK9 was a single nucleotide variation (SNV) at 14 nucleotides upstream to the boundary of intron 5 and exon 6. This variant may make a de novo translational start site, but no experimental evidence was provided. In addition, as the same RAPSN variant was reported in another CMS patient [116], lack of phenotypes in the two siblings without AK9 remains unknown.

4.1.2. Clinical Features and Therapies

Endplate AChR deficiency caused by pathogenic variants in CHRNA1, CHRNB1, CHRND, and CHRNE have been repeatedly reported since 1996 [117]. Frameshifting and nonsense variants are recognized to be pathogenic even without expression studies, but pathogenic missense variants in CHRNA1, CHRNB1, CHRND, or CHRNE may cause (i) reduced AChR expression, (ii) SCCMS, or (iii) FCCMS. Except for a dominantly inherited hemiallelic missense variant in a pedigree, which causes SCCMS, the effects of missense variants in CHRNE cannot be differentiated without expression studies. Lack of expression studies in most pathogenic variants prevents us from counting the number of patients or original articles with endplate AChR deficiency. However, endplate AChR deficiency and FCCMS have essentially the same clinical features, which are also similar to myasthenia gravis. In contrast to myasthenia gravis, endplate AChR deficiency is present in embryogenesis in patients with pathogenic variants in CHRNA1, CHRNB1, and CHRND, or is present from birth in patients with pathogenic variants in CHRNE. These are likely to account for minor facial anomalies, muscle hypoplasia, and lack of diplopia.
RAPSN-CMS has been reported in 38 papers [39,40,46,47,49,50,77,112,113,114,115,116,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143]. A review of 10 patients with RAPSN-CMS showed similar clinical features with a neonatal onset, fluctuations of lid ptosis, bulbar signs, neck muscle weakness, mild limb muscle weakness, as well as with episodic worsening of muscle weakness in adults [134]. These symptoms, however, are commonly observed in any groups of CMS, and none is unique to RAPSN-CMS.
ChEIs are generally effective for endplate AChR deficiency irrespective of defective genes. We, however, should be aware that excessive administration of ChEIs causes an iatrogenic pathology similar to endplate AChE deficiency due to pathogenic variants in COLQ stated below. Ephedrine and salbutamol (albuterol) are also generally effective for endplate AChR deficiency [144]. The effects of adrenergic agonists are likely due to the innervation of sympathetic nerve to the NMJ and the facilitation of the NMJ signal transmission by the sympathetic nerve [72]. In addition, amifampridine is also effective for endplate AChR deficiency [144,145].

4.2. Escobar Variant of Multiple Pterygium Syndrome (EVMPS, Escobar Syndrome) (CHRNG) and Lethal Form of Multiple Pterygium Syndrome (LMPS)/Fetal Akinesia Deformation Sequence (FADS) (CHRNA1, CHRND, MUSK, RAPSN, DOK7, and SLC18A3)

4.2.1. Pathomechanisms

Loss-of-function variants of CHRNG cause EVMSP (Escobar syndrome) and LMPS, both of which are characterized by arthrogryposis multiplex congenita (AMC) and pterygium likely due to the embryonic immobility [35,36,37]. Escobar syndrome takes a benign non-progressive course. A case of Escobar syndrome with uniparental disomy, in which a specific region of both alleles arises from a single parent, is reported [146]. FADS and LMPS are spectrum disorders [147]. Pathogenic variants of CHRNA1 [148], CHRND [148], RAPSN [39,40,147,148], DOK7 [147,149], and SLC18A3 [150] also cause LMPS/FADS. The phenotypes are again likely to be caused by embryonic immobility due to defective NMJ signal transmission.

4.2.2. Clinical Features and Therapies

Escobar syndrome has been reported in 101 patients in 72 pedigrees [35,36,38,45,146,151,152,153,154,155,156,157,158,159,160,161]. As the γ subunit is substituted for the ε subunit after birth, patients show no myasthenia or muscle weakness, but is classified into a form of CMS [36]. Some patients with Escobar syndrome have only distal arthrogryposis but no pterygia [37,38,152]. The presence of an incomplete form of Escobar syndrome suggests that pathogenic variants of CHRNG are likely to be undetermined in patients with distal arthrogryposis. More than 220 causative genes have been reported in AMC, and pathogenic variants of CHRNG are the most common with 6 out 17 pedigrees with AMC [38]. In a report from Spain, 5 out 64 genetically identified CMS patients were Escobar syndrome [45]. Surgical corrections are applied to arthrogryposis.

4.3. Slow-Channel CMS (SCCMS) and Fast-Channel CMS (FCCMS) (CHRNA1, CHRNB1, CHRND, and CHRNE)

4.3.1. Pathomechanisms

SCCMS is caused by abnormal prolongation of the opening time of AChR. In contrast, FCCMS is caused by abnormal shortening of the opening time of AChR. Completely oppositive effects on the channel opening times cause defective NMJ signal transmission.
SCCMS is caused by pathogenic missense variants in one allele of CHRNA1, CHRNB1, CHRND, and CHRNE, encoding the AChR ɑ1, β1, δ, and ε subunits, respectively, and shows autosomal dominant inheritance. A case of autosomal recessive SCCMS was also reported [162,163]. Pathogenic missense variants of SCCMS can be classified into two categories. The first category includes pathogenic missense variants at the extracellular domain especially at the ACh-binding site and at the first transmembrane domain, M1. These variants delay the dissociation of ACh from AChR. The second category includes pathogenic missense variants at the second transmembrane domain, M2, that forms the ion channel pore [164,165]. Three mechanisms may result in defective NMJ signal transmission in SCCMS. First, prolonged openings of AChR ion channel increase the intracellular Na+ concentration and depolarize the resting membrane potential, which reduces the amplitude of an endplate potential (EPP) and makes the muscle sodium channel (NaV1.4) difficult to sense an EPP. Second, as AChR is a cation-non-selective ion channel, prolonged openings of AChR allow excessive influx of Ca2+ ions, that cause endplate myopathy [166]. Ca2+ ions constitute 7% of the endplate current of adult-type εAChR, which is higher than fetal γ-AChR. In two pathogenic variants in SCCMS (CHRNE p.T284P [164] and CHRNE p.V279F [27]), the permeability of Ca2+ ions was increased 1.5- to 2.0-folds, which were likely to accelerate endplate myopathy [167]. Third, prolonged openings of AChR desensitize AChR [168]. AChR is physiologically desensitized by prolonged existence of ACh. Desensitized AChR does not respond to ACh and cannot generate EPP anymore. The structure of desensitized Torpedo AChR was recently solved [169]. In the desensitized state, the two agonist-binding sites between the ɑ-δ and ɑ-ε subunits of AChR are rotated counterclockwise perpendicular to the membrane, and the structure of the extracellular end of the M4 helix of the ɑ subunit that interfaces with the δ subunit becomes much different.
FCCMS is kinetically opposite to SCCMS. Pathogenic missense variants can be classified into three categories. The first category includes pathogenic missense variants at the extracellular domain including the ACh-binding site of AChR [170]. Interestingly, detailed kinetic analyses reveal that most of the variants at the ACh-binding site [107,170,171,172] affect the ion channel gating rather than ACh-binding to AChR. However, the other variants affect ACh-binding alone [171] or both ACh-binding and the ion channel gating [173]. The second category is comprised of pathogenic variants in the long cytoplasmic loop between the second and third transmembrane domains (M3 and M4). These variants destabilize the open channel state [86,174,175]. The third category is a pathogenic missense variant at the third transmembrane domain (M3) [176]. The enlarged amino acid in the M3 domain displaces the second transmembrane domain (M2) and narrows the ion channel pore made by the M2 domains of five subunits.

4.3.2. Clinical Features and Therapies

SCCMS has been reported in 34 original articles since 1995 [27,31,162,163,165,168,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204]. As observed in other autosomal dominant disorders, the onset of SCCMS can be in adolescence or adulthood. Adult-onset patients tend to have mild phenotypes. Weakness of the extensor muscles of the upper limbs is frequently observed in SCCMS, although the underlying mechanisms remain unknown. Weakness of the extensor muscles of the upper limbs is also reported in 10 out of 15 patients with DOK7-CMS [66]. R-CMAP in response to a single nerve stimulus is observed in SCCMS, as in COLQ-CMS and PURA-CMS. A review of 60 SCCMS patients showed that R-CMAP was observed when the opening burst durations of mutant AChR were increased 8.68-fold on average compared to those of wild-type AChR, whereas R-CMAP was not observed when they were increased 3.84-fold on average [198]. Based on knowledge that sodium channel blockers also block AChR ion channel to some extent, shortening of abnormally prolonged AChR channel openings by an antiarrhythmic, quinidine [74] and an selective serotonin reuptake inhibitor (SSRI), fluoxetine [76] was reported by a single channel recordings of SCCMS-AChRs. Indeed, both quinidine [75] and fluoxetine [76] are effective for SCCMS. Amelioration of endplate myopathies in SCCMS requires several months, and immediate effects of quinidine and fluoxetine are not usually observed. A review of 15 SCCMS patients showed that most patients improved by quinidine or fluoxetine, but the effects were not observed for respiratory insufficiency and palpebral ptosis [193]. In their report, 2 out of 6 SCCMS patients with quinidine developed adverse reactions of hypersensitivity reaction and impaired liver function. In 10 SCCMS patients treated with fluoxetine, 7 patients showed clear response, whereas 3 patients either showed adverse effects of serotonergic crisis, lethargy, and hypotension, or could not tolerate higher dose (40 mg/day) [193]. Similarly, in the initial report of fluoxetine for SCCMS, one of two patients had insomnia, drowsiness, and anorexia [76]. A SCCMS patient suffered suicidal ideation soon after commencing fluoxetine, which is well-recognized concern with fluoxetine [205]. Another review of 60 SCCMS patients showed that patients showing good responses to quinidine or fluoxetine started treatment at 11.6 years after the onset of symptoms on average, whereas patients without good responses started treatment at 30.7 years after the onset on average [198]. Although both quinidine and fluoxetine minimally shorten the openings of wild-type AChR, worsening of symptoms was reported in a patient with RAPSN-CMS by fluoxetine that was prescribed for depression [77]. ChEIs and amifampridine are ineffective in most SCCMS patients [31,184,188,193], but ChEIs improved the symptoms of a patient with SCCMS [185]. ChEIs presumably enhance the desensitization of AChRs and reduce the number of available AChRs that can respond to ACh. In addition, the effects of ephedrine and salbutamol (albuterol) were reported in SCCMS [193,202,203], as well as in mouse models [206,207].
FCCMS has been reported in 13 original articles since 1996 [64,130,170,170,173,175,208,209,210,211,212,213,214,215]. Biallelic pathogenic missense variants including small indels either cause AChR deficiency or FCCMS. To differentiate the two types of pathologies, microelectrode studies and/or single channel recordings of the biopsied skeletal muscle, and/or single channel recordings of mutant AChR expressed on cultured cells, are required. The unavailability of these techniques is likely to account for the rarity of FCCMS. Indeed, FCCMS has been reported from only three laboratories in the world. Pathogenic variants of FCCMS have been reported in CHRNA1, CHRND, and CHRNE, but not in CHRNB1. Although only the β subunit does not contribute to make a ACh-binding site, missense variants in CHRNB1 can possibly cause FCCMS. Although the differentiation of FCCMS and AChR deficiency is challenging, similar therapies can be applied to both diseases. FCCMS patients respond to ChEIs [2,130,173,200,213], amifampridine [2,213], and salbutamol (albuterol) [216]. Ephedrine is also likely to be effective for FCCMS, but its effect has not been reported. Favorable responses to ChEIs and amifampridine may not necessitate the use of other drugs.

4.4. Synaptic CMS (COLQ, LAMB2, and COL13A1)

4.4.1. Pathomechanisms

One, two, and four molecules of AChE enzyme make globular forms of AChE that are named G1, G2, and G4, respectively. In addition, triple helical collagen Q (ColQ) binds to 4, 8, and 12 molecules of AChE and makes asymmetric forms of AChE named A4, A8, and A12, respectively. Asymmetric forms of AChE are enriched at the NMJ. ColQ has three domains. First, the proline-rich attachment domain (PRAD) at the N terminal end is enriched in prolines. The tetrameric forms of AChE bind to PRAD, and three ColQ strands make A12 -AChE. Second, the collagen domain in the middle of ColQ has prolines at every three residues like other collagens, and makes a stable triple helical structure. The collagen domain has two regions enriched in positively charged basic amino acids, where heparan sulfate proteoglycans (HSP) including perlecan bind [217]. The two regions are named HSP-binding domain (HSPBD) [218]. Third, the C-terminal domain (CTD) of ColQ is enriched in charged amino acids and cysteines, and makes a globular form. CTD of ColQ binds to MuSK [102,219,220]. Asymmetric forms of AChE are generated in the Golgi apparatus, excreted to the synaptic space, and are anchored to the synaptic basal lamina by binding of ColQ to HSP and MuSK.
Loss-of-function variants of COLQ cause endplate AChE deficiency [221,222,223,224,225]. Although the roles of ColQ at the NMJ have been well analyzed, ColQ is also expressed in other tissues including brain, testis, and heart. The roles of ColQ in other tissues, however, remain unknown, and Colq-deficient mice show no phenotypes other than endplate AChE deficiency [226,227]. In contrast to COLQ, no pathogenic variants have been reported in ACHE in any diseases. AChE plays essential roles in the cholinergic synapses in the CNS. Loss-of-function of AChE is thus likely to be fatal in humans. Although there is no relevance to human diseases, p.H322N (rs1799805) in ACHE determines the YT blood group [228]. Pathogenic variants of COLQ are classified into three categories [224]. First, pathogenic variants in PRAD impair the binding of AChE to PRAD. Second, pathogenic variants in the collagen domain impair the formation of the triple helix. Most of them are nonsense or frameshifting variants. Third, pathogenic variants at CTD impair anchoring of ColQ to the NMJ by inhibiting the binding of ColQ to MuSK [225,229].
Although both endplate AChE deficiency and SCCMS are caused by excessive openings of AChR, the mechanisms of defective NMJ signal transmission are not identical. Two mechanisms are similar between the two diseases. First, depolarization of the resting membrane potential reduces the amplitude of EPP, and small EPP cannot activate the skeletal muscle sodium channel. Second, AChRs are desensitized by the prolonged presence of ACh in endplate AChE deficiency and prolonged openings of AChRs in SCCMS. In contrast to SCCMS, however, endplate myopathy due to excessive influx of Ca2+ ions are not observed in endplate AChE deficiency, because the nerve terminal becomes small, and the terminal Schwann cells invaginate into the synaptic space, both of which reduce the number of releasable ACh quanta.
Laminins-221, -421, and -521, all of which include β2-laminin (LAMB2), are expressed at the NMJ. Lamins are heterotrimeric extracellular matrix molecules made of ɑ, β, and γ subunits, and are key molecules constituting the synaptic basal lamina [230]. Laminins play a critical role in maintenance of the NMJ, and organization of synaptic vesicle release sites known as active zones. Laminins-221, -421, and -521 are made of ɑ2-, ɑ4-, and ɑ5-laminins, respectively, as well as of β2-, and γ1-laminins. Laminins at the NMJ play essential roles in the juxtaposition of presynaptic and postsynaptic structures and the placement of the terminal Schwann cells at the NMJ. β2-Laminin directly binds to P/Q- and N-type voltage-gated calcium channel (VGCC) [231,232,233], and is essential for the formation and organization of presynaptic active zones [234]. β2-Laminin is also expressed in renal glomeruli and eyes. Pathogenic variants in LAMB2 cause Pierson syndrome [235] and nephrotic syndrome type 5 [236]. Pierson syndrome is characterized by congenital nephrotic syndrome and a complex maldevelopment of the eye with lens abnormalities, atrophy of the ciliary muscle, corneal changes, and retinal changes. Pathological variants of LAMB2 were reported in a CMS patient with Pierson syndrome [65]. Ultrastructural analysis of the biopsied muscle showed marked reduction in the size of the nerve terminals, invagination of the synaptic space by the processes of Schwann cells, and moderate simplification of postsynaptic folds. Electrophysiological examinations showed marked reduction in quantal release of ACh from the nerve terminal. Lamb2-deficient mice also show similar phenotypes at the NMJ [237].
Collagen 13ɑ1 (COL13A1) enriched at the NMJ has a single transmembrane domain and plays an essential role in the maturation and maintenance of AChR at the NMJ [238]. A frameshifting variant of COL13A1 causes CMS [239]. Introduction of the pathogenic variants into C2C12 myotubes reduced AChR clustering [239]. Col13a1-deficient mice show abnormal formation of the NMJ [238,240], as well as craniofacial malformations and a reduction in cortical bone mass in aged mice [241].

4.4.2. Clinical Features and Therapies

COLQ-CMS has been reported in 30 original articles since 1998 [62,73,78,140,141,221,222,225,242,243,244,245,246,247,248,249,250,251,252,253,254,255,256,257,258,259,260,261,262,263]. Interestingly, a grandmother and a father of two siblings with COLQ-CMS carried a heterozygous truncation variant of COLQ, and showed congenital ptosis [246]. Although the presence of a pathogenic variant on another allele could not be excluded, a heterozygous variant of COLQ might have caused ptosis. Initial symptoms of COLQ-CMS are mostly ophthalmoplegia, respiratory insufficiency, and weak crying at birth. Follow-up of 15 patients with COLQ-CMS aged 3 to 48 years for up to 10 years showed that 80% of patients were still ambulant and 87% had no respiratory difficulties [253]. A report of 22 patients with COLQ-CMS indicated proximal dominant muscle weakness that was characteristic of limb-girdle-type myasthenia as in DOK7-CMS [247]. Fluctuating scoliosis due to truncal muscle weakness is later changed to severe scoliosis, which is uniquely observed in COLQ-CMS and DOK7-CMS [249]. Palpebral ptosis and external ophthalmoplegia are observed in about half of COLQ-CMS patients [247,253]. Diurnal fluctuation and progression of muscle weakness are observed in about half of the patients [253]. Delayed pupillary response is characteristic of COLQ-CMS but is observed in only 25% of the patients [247]. Repetitive CMAP, which is also observed in SCCMS and PURA-CMS, is observed in about half of the patients [253]. In COLQ-CMS, globular forms of AChE and butyrylcholinesterase at the motor endplate catalyze ACh, and blocking of these enzymes by ChEIs can sometimes cause respiratory arrest [62,63,64]. ChEIs had no long-term effects in 22 patients with COLQ-CMS but showed short-time effects in 4 patients [247]. Ephedrine and salbutamol (albuterol) are effective for COLQ-CMS [264,265,266]. Especially, in two patients, ephedrine showed marked effects [264], although the underlying mechanisms remain unknown. The effects of amifampridine are also reported [62,73], but the mechanisms again remain elusive. The effect of fluoxetine that is used for SCCMS was also reported in a patient with COLQ-CMS [78]. Although fluoxetine minimally shortens the channel opening time of wild-type AChR [74], a slight reduction in AChR openings was likely to be sufficient for the patient.
LAMB2-CMS was reported in a 20-year-old female with Pierson syndrome in 2009 [65]. The patient had repeated respiratory distress since birth, miosis, and severe proteinuria. Development of motor functions were delayed, but proteinuria was improved by rental transplantation at age 7 years. Palpebral ptosis, external ophthalmoplegia, and proximal muscle weakness were noted. RNS reduced CMAP by 24%. ChEI worsened her muscle weakness, and respiratory support was required. Ephedrine was effective.
COL13A1-CMS has been reported in 41 patients in 19 pedigrees since 2015 [140,141,239,267,268,269]. All patients developed respiratory distress and weak sucking at birth. In addition to severe palpebral ptosis and mild external ophthalmoplegia, the patients showed facial, bulbar, respiratory, and truncal muscle weakness. Compared to the trunk muscles, limb muscles are spared. ChEIs are ineffective [239,267,268], but salbutamol (albuterol) [239,267,268] and amifampridine [268] are effective.

4.5. Sodium Channel CMS (SCN4A)

4.5.1. Pathomechanisms

Loss-of-function variants of SCN4A cause CMS [79,81,270], whereas gain-of-function variants of SCN4A cause hyperkalemic periodic paralysis [271], hypokalemic periodic paralysis [271], potassium-aggravated myotonia congenita [272], and paramyotonia congenita [273]. Loss-of-function variants of SCN4A in CMS shift the fast inactive curve toward hyperpolarized states and make NaV1.4 inactive even at the resting membrane potential. NaV1.4 ion channel opens in response to the first depolarization stimulus, but not to the second or later depolarization stimuli because of accelerated transition to a fast inactive state. This also causes decremental CMAP response to RNS. In contrast to SCN4A-CMS, gain-of-function variants in hyperkalemic periodic paralysis, hypokalemic periodic paralysis, potassium-aggravated myotonia congenita, and paramyotonia congenita shift the fast inactivation curve toward depolarized states. This allows repeated openings of NaV1.4 or allows leakage of Na+ even in a closed state. In SCN4A-CMS and in some CMS patients with defective recycling of ACh (CHAT-CMS, SLC18A3-CMS, SLC5A7-CMS, and PREPL-CMS), a high-frequency nerve stimulation is required to elicit a decremental CMAP response, and episodic muscle weakness is observed.

4.5.2. Clinical Features and Therapies

SCN4A-CMS has been reported in 6 patients since 2003 [79,80,81,140,270,274]. SCN4A-CMS shows frequent episodes of respiratory arrest, bulbar paralysis, and muscle weakness that persist 30 to 60 min. In the intermittent phase, mild facial, truncal, and limb muscle weakness, as well as external ophthalmoplegia, are observed. Analysis of 278 patients with sudden infantile death syndrome (SIDS) revealed 4 patients with SCN4A-CMS [5]. ChEIs are either effective [79,142,274] or ineffective [270]. In addition, a SCN4A-CMS patient showed marked cholinergic adverse effects with a small amount of ChEI [80]. Salbutamol (albuterol) was effective in a single patient [142]. Similarly, acetazolamide was either effective [79,80] or ineffective [81] to prevent episodic muscle weakness.

4.6. Defective AChR Clustering (AGRN, LRP4, MUSK, and DOK7)

4.6.1. Pathomechanisms

Agrin (AGRN) is a large molecule secreted from the nerve terminal with a molecular weight of ~200 kDa, and carries binding domains for laminins, neural cell adhesion molecule (NCAM), α-dystroglycan, and LRP4. Functionally characterized pathogenic variants of AGRN invariably impair AChR clustering. However, three pathologies exist depending on the affected domains: (i) impairment of MuSK phosphorylation, (ii) accelerated degradation of agrin, (iii) impaired anchoring of agrin to the NMJ [275].
The third β propeller domain of LRP4 binds to agrin. Pathogenic variants in this domain in CMS impair binding of LRP4 to agrin and MuSK, reduce MuSK phosphorylation, and compromise AChR clustering [71]. Pathogenic variants in this domain are also reported in sclerosteosis type 2 (SOST2), which is characterized by cortical hyperostosis [276]. CMS variants affect agrin-LRP4-MuSK signaling but not Wnt signaling, whereas SOST2 variants affect Wnt signaling but not agrin-LRP4-MuSK signaling. Analysis of additional artificial variants revealed that variants at the periphery of the third β propeller domain exclusively affect agrin-LRP4-MuSK signaling, whereas variants at the center of the domain exclusively affect Wnt signaling [71]. Pathogenic variants of the other domains of LRP4 are also reported in another bone disorder, Cenani-Lenz syndactyly syndrome [277]. Thus, pathogenic variants of LRP4 either affect agrin-LRP4-MuSK signaling or Wnt signaling.
Pathogenic variants of MUSK either reduces cell membrane expression of MuSK without affecting agrin-mediated phosphorylation of MuSK [278], or markedly reduces MuSK phosphorylation and AChR clustering [279].
More than 70 missense, truncation, and splicing pathogenic variants have been reported in DOK7 in CMS [280,281,282,283]. Thirteen missense variants have been functionally characterized, and all reduce the phosphorylation of MuSK and AChR β1 subunit [280,282,283,284]. One missense variant in the pleckstrin homology domain markedly reduces DOK7 expression by generating aggresome at the juxtanuclear region [284].

4.6.2. Clinical Features and Therapies

AGRN-CMS has been reported in 13 original articles since 2009 [6,141,142,274,275,285,286,287,288,289,290,291,292]. Two patients with AGRN-CMS reported in 2009 were 42-year-old female and 36-year-old male in a single pedigree, who had mild limb muscle weakness and unilateral ptosis since childhood [285]. ChEIs and amifampridine were ineffective. Most AGRN-CMS patients similarly develop muscle weakness since childhood, and the symptoms range from mild muscle weakness in lower limbs to severe muscle weakness that requires respiratory support. Again, ChEIs and amifampridine are ineffective or mildly effective. On the other hand, salbutamol (albuterol) was effective in 10 out of 12 AGRN-CMS patients [287]. Similarly, ephedrine was effective in a single patient [288]. Biallelic null variants of AGRN caused FADS and gave rise to stillbirth at 30 weeks of gestation [293]. In addition, analysis of 262 patients with autism spectrum disorder (ASD) revealed hemiallelic null variants of AGRN [294]. However, as null variants of AGRN are observed in asymptomatic parents of AGRN-CMS patients, other genetic or environmental factors are likely to be required to be associated with ASD. In addition, pathogenic AGRN variants were identified in hereditary motor neuropathy, in whom jitters by SFEMG were increased [295].
MUSK-CMS has been reported in 15 original articles since 2004 [44,70,141,278,279,296,297,298,299,300,301,302,303,304,305]. A review of 15 MUSK-CMS patients showed that the ages of onset were from birth to 8 years, and most patients had proximal muscle weakness, palpebral ptosis, external ophthalmoplegia, facial weakness, bulbar palsy, and truncal muscle weakness [70]. About half of the patients required respiratory support for respiratory insufficiency. ChEIs were effective or worsened muscle weakness, and amifampridine and salbutamol (albuterol) were mildly or markedly effective [70]. In addition, 19 patients with FADS due to pathogenic variants of MUSK were reported [306,307].
LPR4-CMS with compound heterozygous variants was reported in a single patient in 2014 [71]. The patient had respiratory distress at birth, and was dependent on a respirator up to age 6 years. Evaluations at ages 9 and 14 years showed mild external ophthalmoplegia and severe muscle weakness. ChEIs worsened muscle weakness.
DOK7-CMS has been reported in 34 original articles since 2006 [43,45,50,64,66,67,68,69,139,140,141,142,147,280,281,282,283,308,309,310,311,312,313,314,315,316,317,318,319,320,321,322,323,324]. A review of 15 patients with DOK7-CMS showed that the ages of onset were mostly from birth to infancy, and the oldest age of onset was 13 years [66]. All patients showed proximal and truncal muscle weakness, and scoliosis was frequently observed. In addition, distal muscle weakness, especially finger extensors, was observed in 12 patients. Similar, weakness of finger extensors is also observed in SCCMS. Muscle hypoplasia was present in about half of the patients; palpebral ptosis and external ophthalmoplegia in 11 patients; and facial and bulbar muscle weakness in 8 to 9 patients. DOK7-CMS is recognized as limb-girdle CMS, but ocular, facial, and bulbar weakness is frequently observed. The diagnosis of myasthenia gravis was erroneously given to 4 out of 15 patients, and others were diagnosed as congenital myopathy, metabolic myopathy, or mitochondrial myopathy. Although the causal relation remains unknown, siblings of DOK7-CMS had mitral valve insufficiency [322]. A total of six CMS patients were heterozygous for a truncation variant of DOK7 without any pathogenic variants on another allele [281,325]. Although the presence of a pathogenic variant on another allele could not be excluded, a heterozygous variant of DOK7 might cause CMS when unidentified genetic and/or environmental factor(s) coexisted. More such heterozygous patients may exist, but may be underestimated due to possible publication bias. The effects of ephedrine and salbutamol (albuterol) for DOK7-CMS have been repeatedly reported [64,66,67,68,69,308,310,315,316,317,321,324]. In addition the effect of a patch of tulobuterol, a β2 agonist, was reported in a case with DOK7-CMS [326]. On the other hand, ChEIs are ineffective or worsen muscle weakness [64,66,67,68,69]. The effect of amifampridine was also reported [319]. Fluoxetine was effective in a patient with DOK7-CMS, who was misdiagnosed as SCCMS [327]. Administration of anti-DOK7 antibody that stimulated DOK7 was effective for a mouse model of DOK7-CMS carrying a pathogenic variant of the patient [328]. Although the mechanisms are unknown, a DOK7-CMS patient was responsive to steroid for 40 to 50 years [329]. In addition, 4 patients with FADS due to pathogenic variants of DOK7 were reported [147,149].

4.7. CMS Caused by Defective Structural Molecule at the NMJ (PLEC)

4.7.1. Pathomechanisms

Plectin is a 500-kD intermediate filament-binding protein that provides mechanical strength by acting as a crosslinking element of the cytoskeleton. In the skeletal muscle, plectin is expressed in sarcolemma and Z band. In the skin, plectin makes hemidesmosome. Pathogenic variants of PLEC cause epidermolysis bullosa simplex (EBS) [330] and autosomal recessive limb-girdle muscular dystrophy 17 (LGMD17) [331]. In patients with EBS and LGMD17, endplate AChR deficiency was reported [332,333,334]. A homozygous 9-bp deletion of the translational start site of PLEC caused LGMD17 in 3 patients in 3 pedigrees in Turkey without any evaluation of the NMJ [331]. The same variant, however, caused LGMD17 and endplate AChR deficiency in 4 patients in 4 pedigrees in Turkey [53], indicating that myasthenic symptoms might be masked by muscular dystrophy. Plectin is highly expressed at the NMJ, connects desmin and dystrophin-dystroglycan complex, binds to rapsyn-AChR complex, and stabilizes the NMJ structure [335]. Indeed, ultrastructural analyses show destruction and remodeling of the endplate [332].

4.7.2. Clinical Features and Therapies

PLEC-CMS has been reported in 22 patients since 1999 [53,274,332,334,336,337,338,339]. Although LGMD17 and CMS are always present, EBS may [332,334,336,338] or may not [53,337] be present. Some patients also show mild EBS [338]. A review of 117 PLEC-EBS patients carrying pathogenic variants in PELC showed that 14 patients also had CMS [339]. However, the authors observed the presence of CMS in 7 out of 15 patients in their own cohort of PLEC-EBS [339], indicating that CMS was likely to be underdiagnosed and that the prevalence of CMS was higher than reported. The prevalence of muscular dystrophy in PLEC-EBS was also high. The onsets of PLEC-CMS range from early childhood to age 26 years, and patients show limb muscle weakness, swallowing difficulty, respiratory insufficiency, palpebral ptosis, external ophthalmoplegia [53,332,334,338]. As in most other patients with CMS, low-frequency RNS elicit decremental CMAP responses. ChEIs were ineffective in 3 patients [334], and effective in 3 other patients [338]. A combination of ChEIs and salbutamol (albuterol) was effective in 4 patients [53]. Amifampridine was effective in a case [338], and was ineffective in 2 other patients [334,338]. In addition, a case of PLEC-CHRNE-CMS who had both biallelic insertion of 36 bp in PLEC and biallelic frameshift in CHRNE showed EBS and CMS [333]. ChEIs and ephedrine were mildly effective for this case.

4.8. CMS Caused by Defective Recycling of ACh (CHAT, SLC18A3, SLC5A7, and PREPL)

4.8.1. Pathomechanisms

Choline acetyltransferase (ChAT, CHAT) synthesizes ACh from choline and acetyl-CoA at the nerve terminal. Vesicular acetylcholine transporter (vAChT, SLC18A3) transports synthesized ACh to the synaptic vesicle. SLC18A3 is encoded within the first intron of CHAT. This nested gene structure is conserved from C. elegans. Loss-of-function variants of CHAT cause CMS with episodic apnea [340,341]. ChAT is also expressed at the cholinergic synapses in the CNS, and developmental delay that is observed in about half of CHAT-CMS patients can be accounted for either by defects in the cholinergic synapse in the CNS or by hypoxia due to episodic apnea. Parents of CHAT-CMS who carry a null variant in a single allele are asymptomatic, whereas no CHAT-CMS patients carry biallelic null variants. Thus, the reduction in the enzymatic activity of ChAT to 30-50% is predicted to cause CMS, whereas ChAT activities lower than 30% are lethal and more than 50% are asymptomatic [340,341].
Loss-of-function variants of SLC18A3 also cause CMS. Although the disease mechanisms have not been dissected in detail, failure to pack resynthesized ACh into synaptic vesicles is likely to be the cause of CMS.
A hemiallelic large scale DNA rearrangement at 10q11.2 including CHAT and SLC18A3 was observed in 41 patients with autism, developmental delay and/or intellectual disability, and multiple congenital malformations [342]. Muscle hypotonus, palpebral ptosis, and sleep apnea in these patients may be caused by haploinsufficiency of CHAT and SLC18A3. However, as stated above, as hemiallelic null variants are symptomatic in parents of CHAT-CMS, either complete lack of vAChT (SLC18A3) that is intact in CHAT-CMS or an unidentified pathogenic variant on the other allele may cause the disease. Indeed, in two patients with CMS, a hemiallelic large scale deletion at 10q11.2 region was unmasked by a pathogenic splicing variant in CHAT, or a pathogenic missense variant in SLC18A [343].
High affinity choline transporter (ChT, SLC5A7) expressed at the nerve terminal membrane uptakes choline to the nerve terminal. ChT is a homo-oligomeric membrane transporter. A hemiallelic frameshifting variant of SLC5A7 cause autosomal dominant distal hereditary motor neuropathy type VIIA (DHMN7A) [344,345]. DHMN7A is characterized by teen-age onset progressive distal limb muscle weakness and amyotrophy with vocal paralysis. Later, recessive pathogenic variants of SLC5A7 were reported to cause CMS [41]. Expression studies of cultured cells showed that dominantly inherited variants of SLC5A7 have dominant-negative effects [344], whereas recessively inherited variants have loss-of-function effects [41,346], on choline uptake by ChT. Dominantly inherited variants of SLC5A7 are likely to inhibit the formation of homo-oligomers of ChT, whereas recessively inherited variants do not. However, it remains unknown why similar reductions of the ChT activity give rise to two different phenotypes of DHMN7A and CMS. Mice deficient for Slc5a7 dies in a few minutes after birth probably due to respiratory failure [347], and spinal motor neuron-specific rescue of Slc5a7 prolonged the knockout move and breath for ~24 h after birth [348]. Hemizygous knockout of Slc5a7 in mice decreased cardiac ACh and showed diminished parasympathetic heart effects with basal resting tachycardia [349], which, however, has not been documented in patients with SLC5A7-CMS.
Prolyl endopeptidase-like (PREPL) is one of serine peptidases, and its physiological substrate is unknown. The SLC3A1 gene encoding the cystine, dibasic, and neutral amino acid transporter and the PREPL gene are encoded on the opposite strands each other and have an overlap at their 3′ ends. Deletion of both genes cause hypotonia-cystinuria syndrome (HCS) [350]. Deletion of PREPL causes muscle hypotonia [351], whereas deletion of SLC3A1 causes cystinuria [352]. Loss-of-function variants of PREPL do not cause endplate AChR deficiency, but inhibit refilling of ACh to synaptic vesicles, reduce the number of releasable ACh quanta, and decrease the probability of vesicular release [351].

4.8.2. Clinical Features and Therapies

CHAT-CMS has been reported in 19 original articles since 2001 [45,140,141,142,340,343,353,354,355,356,357,358,359,360,361,362,363,364]. A follow-up study of 11 patients with CHAT-CMS for maximum of 12 years showed two forms of disease [358]. It can present in neonates with episodic apnea, respiratory distress, swallowing difficulty, and limb muscle weakness. It can also start in infancy with episodic apnea, and mild limb muscle weakness. The milder infantile form may show progressive muscle weakness with wheelchair dependency. Episodic apnea is frequently misdiagnosed as epilepsy [358]. Episodic apnea, however, is observed in other groups of CMS, and is not unique to CHAT-CMS. Similar to SCN4A-CMS, about half of CHAT-CMS patients show no decremental response to low-frequency RNS, and require high-frequency RNS at 10 Hz or more. A respiratory monitor is required for neonatal and infantile episodic apnea. ChEIs and amifampridine are effective in most patients with CHAT-CMS [340,341].
SLC18A3-CMS has been reported in seven patients in six pedigrees since 2016 [343,365,366,367]. Patients show severe muscle weakness at birth, muscle hypotonia, arthrogryposis, and respiratory distress. Although not observed in all the patients, palpebral ptosis, external ophthalmoplegia, and episodic apnea are also observed [343,365,367]. In addition, FADS was reported in two patients with biallelic nonsense variants of SLC18A3 [150]. RNS was documented in three patients: two showed decremental CMAP in response to low-frequency RNS [365,367], whereas one showed decremental CMAP only after isometric muscle contractions as observed in SCNA4-CMS and CHAT-CMS [365]. ChEIs [365,366,367], ephedrine [365,367], and amifampridine [365,367] are effective for SLC18A3-CMS.
SLC5A7-CMS has been reported in 12 patients in 10 pedigrees since 2016 [41,346,368,369]. Typical clinical features include neonatal-onset episodic apnea, muscle hypotonia, muscle weakness, and weak crying. Some patients have arthrogryposis and congenital malformations and die in infancy, and some patients show developmental delay. Progressive brain atrophy was reported in a case with SLC5A7-CMS, which was likely due to repeated apneustic attacks [346]. In addition, repeated intestinal perforations were reported in two patients of SLC5A7-CMS in a single pedigree [346]. Among 6 patients with SLC5A7-CMS, in whom RNS results were documented, 5 patients showed decremental CMAP to low-frequency RNS [41,346,368], and a single patient showed decremental CMAP only after RNS at 20 Hz for 10 sec [41,365]. ChEIs are effective [41,346,368], and ephedrine has an additional effect [346]. Amifampridine is ineffective [346].
PREPL-CMS without pathogenic variants SLC3A1 has been reported in 11 patients since 2014 [33,351,370,371,372,373,374,375,376]. Similarly, PREPL-CMS with SLC3A1 deletion, the diagnosis of which is HCS, has been reported in 7 patients since 2014 [33,351]. PREPL-CMS is characterized by fluctuating muscle weakness and feeding difficulty since birth, and sometimes requires respiratory support. Patients show palpebral ptosis, nasal voice, swallowing difficulty and facial muscle weakness, and sometimes proximal limb muscle weakness. Intelligence is normal or slightly affected. CMAP decreases with low-frequency RNS [371,376]. In a patient with PREPL-CMS, however, low-frequency RNS elicited decremental CMAP only after RNS at 20 Hz for 2 min [376]. Ten patients with PREPL-CMS including HCS were initially diagnosed as Prader-Willi syndrome [33]. ChEIs are effective [35,351,372,374]. The first PREPL-CMS patient could discontinue ChEI at age 12 months, although muscle weakness was still present [351].

4.9. Lambert-Eaton Myasthenic Syndrome (LEMS)-Like CMS (SYT2, SNAP25, VAMP1, UNC13A, RPH3A, and LAMA5)

4.9.1. Pathomechanisms

Synaptotagmin 2 (SYT2) senses Ca2+ ions entering into the nerve terminal through P/Q-type calcium channel and triggers the formation of the SNARE complex that releases ACh in synaptic vesicles to the synaptic space. Hemiallelic pathogenic variants were identified in SYT2 in patients with CMS resembling LEMS [56,57]. Functional analysis with Drosophila showed that the variants indeed affected the release of synaptic vesicles [56,57].
The SNARE complex is made of SNAP25, syntaxin, and synaptobrevin (vesicle-associated membrane protein 1, VAMP1). Hemiallelic de novo loss-of-function missense variants of SNAP25 cause CMS with developmental delay and ataxia [15]. SNAP25 has two splicing isoforms: SNAP25A transcript includes 118-bp exon 5A, whereas SNAP25B transcript includes 118-bp exon 5B. Embryonic SNAP25A is switched to adult-type SNAP25B after birth. Pathogenic variants in exon 5B encoding SNAP25B cause CMS [15]. t-SNARE liposome containing a variant SNAP25B failed to properly fuse to v-SNARE liposome induced by calcium ions. In addition, bovine chromaffin cells expressing a variant SNAP5B showed compromised exocytosis in response to depolarization.
Another component of the SNARE complex, syntaxin 1, is bent in the middle and is in a closed conformation at rest. Munc18-1 stabilizes syntaxin 1 in a closed state. In response to the entry of Ca2+ ions to the nerve terminal, Munc13-1 (UNC13A) binds to syntaxin 1 and displaces Munc18-1, which stabilizes the open conformation of syntaxin 1 [377]. Biallelic truncation variants of UNC13A caused severe muscle weakness at birth, microcephaly, hypoplastic corpus callosum, enhanced excitation of cerebral cortex [11]. A microelectrode study of biopsied skeletal muscle showed that UNC13A-CMS decreased the quantal contents of synaptic vesicles but spared the release probability of synaptic vesicles. In contrast to UNC13A-CMS caused by biallelic truncation variants, hemiallelic pathogenic missense variants of UNC13A do not cause CMS but cause dyskinesia, developmental delay, and autism [378].
Biallelic loss-of-function variant of VAMP1 encoding another component of the SNARE complex, synaptobrevin1, cause a neonatal onset CMS [379,380,381]. Vamp1-deficient mice showed marked shrinkage of motor endplate and reduction in endplate potentials, which is electrophysiologically similar to LEMS [379].
Rabphilin 3a (RPH3A) is an effector of a Ras superfamily molecule, Rab3A, and binds to Rab3A at the nerve terminal. In addition, Rabphilin 3a binds to SNAP25 and 14-3-3 proteins. In Drosophila, 14-3-3ζ binds to and regulates potassium channel at the nerve terminal of the NMJ [382]. Although Rph3a knockout shows no phenotypes in mice [383] or Drosophila [384], microinjection of rabphilin 3a into the squid giant axon suppresses release of synaptic vesicles [385]. Two pathogenic missense variants identified in a patient with RPH3A-CMS reduced the binding of rabphilin 3a to 14-3-3, but not to Rab3A or SNAP-25 [12].
Laminin α5 (LAMA5) is highly expressed at the NMJ. Knockout of Lama5 results in embryonic lethality in mice [386], whereas muscle-specific knockout of both Lama4 and Lama5 markedly affect the postsynaptic structure [387]. Muscle-specific knockout of Lama5 alone does not show any motor deficit, but differentiation of the nerve terminal is compromised, and the nerve terminal is juxtaposed to only a part of motor endplate [387]. In LAMA5-CMS, quantal release of ACh is markedly reduced. In addition, although the junctional folds of the motor endplate are spared, the motor endplate is not covered by the nerve terminal or covered by a small nerve terminal [13]. Synaptic vesicle glycoprotein 2A (SV2A) binds to synaptotagmin, and pathogenic variants of LAMA5 impairs binding to SV2A [13].
Some patients with SYT2-CMS [9,56,57,60] and all the two patients with SNAP25-CMS [15,54] are caused by hemiallelic pathogenic variants, whereas the other LEMS-like CMS requires biallelic pathogenic variants.

4.9.2. Clinical Features and Therapies

SYT2-CMS has been reported in 6 original articles since 2014 [9,56,57,58,59,60]. Hemiallelic pathogenic variants caused SYT2-CMS in 15 patients in 4 pedigrees [9,56,57,60], whereas biallelic premature termination codons caused SYT2-CMS in 9 patients in 7 pedigrees [58,59,60]. Patients with hemiallelic variants show childhood onset, whereas those with biallelic variants show severe phenotypes mostly with neonatal or infantile onset. Ten patients in 2 pedigrees in the first report had myasthenia but had no ptosis or external ophthalmoplegia [9,56]. Low-frequency RNS elicited decremental CMAP responses, and short maximal voluntary contractions enhanced CMAP amplitudes as observed in LEMS. They were initially diagnosed as Charcot-Marie-Tooth disease or hereditary distal motor neuropathy [9]. ChEIs [58,59] and amifampridine [9,59] were effective for SYT2-CMS. Amifampridine was more effective than ChEIs [9]. Salbutamol (albuterol) had no effect [59].
SNAP25-CMS was reported in a single patient in 2014 [15] and another in 2022 [54]. Both were caused by hemiallelic pathogenic variants. SNAP25-CMS shows severe muscle hypotonia and muscle weakness at birth, and arthrogryposis multiplex. A case in 2014 could walk with a walker at age 7 years, and sometimes had palpebral ptosis [15]. In a case in 2014, low-frequency RNS caused a decremental CMAP, but high-frequency RNS was not performed [15]. In a case in 2022, no RNS studies were performed [54]. ChEI was ineffective, but amifampridine was effective [15]. A case in 2022 died at age 6 days due to respiratory distress [54].
UNC13A-CMS was reported in a single patient in 2016 [11]. UNC13A-CMS showed severe muscle hypotonia and muscle weakness at birth, and microcephaly and hypoplastic corpus callosum. Low- and high-frequency RNS showed decremental and incremental CMAP responses, respectively. EEG showed sharp waves, but no epileptic attacks were noted. ChEI and amifampridine improved decremental CMAP in response to RNS, but minimally improved clinical symptoms. The patient died at age 50 months due to respiratory failure.
VAMP1-CMS has been reported in 9 patients in 7 pedigrees since 2017 [10,379,380,381]. Patients had muscle hypotonia, muscle weakness, and myasthenia since birth [10,379,381] or age 6 months [380]. CMAPs decremented and incremented in response to low- and high-frequency RNS, respectively [10,379]. External ophthalmoplegia and bulbar palsy were also noted [10,379,381]. ChEIs were effective for VAMP1-CMS [10,379,380,381]. Amifampridine has not been administered to VAMP1-CMS, and the effects remain unknown.
RPH3A-CMS was reported in a 11-year-old female [12]. She had limb muscle weakness, nasal voice, and intolerance to exercises since age 3 years. She had learning disabilities. No palpebral ptosis or external ophthalmoplegia was noted. She had mild proximal limb muscle weakness and cervical muscle weakness. Although the association to CMS is unknown, repeated abdominal pain and hyperglycemia are also noted. CMAP amplitudes were not decreased at 2 Hz RNS, but were increased at 30 Hz RNS, as observed in LEMS. Salbutamol (albuterol) was effective, and other drugs were not used.
LAMA5-CMS was reported in a single patient in 2017 [13]. The patient was noted with weak cry and was dependent on a respirator. A brother died of muscle weakness, but details were unknown. Minor facial anomalies were also noted. Low-frequency RNS decreased CMAP by maximum 55%. Low-frequency RNS after maximum muscle contraction for 30 sec increased CMAP to 250%. Coadministration of ChEI and amifampridine was effective.

4.10. Glycosylation-Deficient CMS (GFPT1, DPAGT1, ALG2, ALG14, GMPPB)

4.10.1. Pathomechanisms

Glutamine-fructose-6-phosphate transaminase 1 (GFPT1) is a rate-limiting enzyme to produce uridine diphosphate N-acetylglucosamine (UDP-GlcNAc) that is an essential source for N- and O-glycosylations (Figure 4). On the other hand, dolichyl-phosphate N-acetylglucosamine phosphotransferase 1 (DPAGT1) and UDP-N-acetylglucosaminyltransferase subunit (asparagine-linked glycosylation 14 homolog, ALG14) work on the first two steps of adding GlcNAc to dolichyl phosphate in N-glycosylation. Pathogenic variants of DPAGT1 [388] and ALG2 [389] were reported in congenital disorder of glycosylation Ij (CDG Ij) with infantile spasms, developmental delay, microcephaly, and finger malformations. Muscle hypotonia and muscle weakness are documented in CDGs, and some CDGs may have CMS. Alpha-1,3/1,6-mannosyltransferase (asparagine-linked glycosylation 2 homolog, ALG2) add mannose in N-glycosylation pathway. GDP-mannose pyrophosphorylase B (GMPPB) makes GDP-mannose from mannose-1-phosphate and GTP. GMPPB is essential for N- and O-mannosylations. Expression of pathogenic variants of GFPT1 causes abnormal structures of muscle fibers and NMJ in zebra fish [390]. In C2C12 myotubes, knockdown of Gfpt1 [391] and Alg14 [24] markedly reduced cell surface expression of AChR. Although pathogenic variants of these genes cause AChR deficiency in cultured cells, the detailed mechanisms remain to be elucidated.

4.10.2. Clinical Features and Therapies

GFPT1-CMS has been reported in 17 original articles since 2011 [18,19,20,45,135,140,274,390,392,393,394]. DPAGT1-CMS [21,395,396,397,398] and GMPPB-CMS [25,29,140,141,274,399,400] have been reported in five and nine original articles, respectively. ALG2-CMS has been reported in 9 patients in 4 pedigrees since 2013 [24,401]. ALG14-CMS has been reported in 12 patients in 7 pedigrees since 2013 [24,402,403,404,405]. In muscle biopsy of a patient with GFPT1-CMS, glycogen storage was observed, and glycogen storage disease was considered [406]. In 12 patients with ALG14-CMS, 10 patients had epilepsy [402,403,404,405], and 2 had severe intellectual disability [403].
Pathogenic variants of GFPT1 [390], DPAGT1 [21], ALG2 [24], ALG14 [24] cause limb-girdle CMS with tubular aggregates. Three patients with GFPT1-CMS had rimmed vacuoles in skeletal muscle [19], and 2 patients with GFPT1-CMS had myofibrillar myopathy with deposition of desmin [394]. Pathogenic variants of GMPPB also cause limb-girdle CMS, but no tubular aggregates [25]. Palpebral ptosis and external ophthalmoplegia are rare in all groups of glycolytic enzyme-deficient CMS.
Deficiency of enzymes in O-mannosylation is observed in congenital muscular dystrophy including Fukuyama-type muscular dystrophy and is called dystroglycanopathy. Pathogenic variants of GMPPB cause muscular dystrophy-dystroglycanopathy (MDDG) type 14 [28] with hypoglycosylation of ɑ-dystroglycan and muscular dystrophy in biopsied muscle [25]. Muscle MRI shows displacement of muscle tissue to fibrous and adipose tissues in paravertebral and proximal skeletal muscles [25,407], as observed in muscular dystrophies. In GMPPB-CMS, serum CK is elevated to 2 to 24 times the upper limit of normal (average 10.7 times) [7,8]. In GFPT1-CMS, serum CK is elevated to about 3 times the upper limit of normal [7,8]. The same variant causes GMPPB-CMS [25] and limb-girdle muscle weakness [408], indicating that myasthenia might not be deeply evaluated.
ChEIs are usually effective for all groups of glycosylation-deficient CMS (GFPT1-CMS [390], DPAGT1-CMS [21], ALG2-CMS [24,401], ALG14-CMS [24,402], GMPPB-CMS [7,25,399]). However, ChEI had no effect on a single patient with ALG2-CMS [409]. Amifampridine was effective in DPAGT1-CMS [21,22]. Salbutamol (albuterol) was effective for GFPT1-CMS [410], DPAGT1-CMS [22], and GMPPB-CMS [25]. Ephedrine was effective for ALG2-CMS [409].

4.11. CMS Caused by Defective Nerve Terminal Formation (MYO9A and SLC25A1)

4.11.1. Pathomechanisms

Myosin 9A (MYO9A) expressed in peripheral nerves is an atypical myosin carrying the Rho GTPase-activating protein (GAP) domain and regulates intracellular transport. Myo9a inhibits RHOA by stimulating its GTPase activity through the GAP domain [411]. Biallelic loss-of-function variants of MYO9A cause CMS [412]. Knockdown of two orthologs, myo9aa/ab, in zebrafish causes shortening and abnormal branching of spinal motor neurons, and defective NMJ signal transmission [412]. Agrin fragment rescued defective neurite elongation and motor deficits in myo9aa/ab-deficient zebrafish [413]. Knockdown of Myo9a in NSC34 cells revealed that Myo9a is essential for the formation and maintenance of neuronal cells, and for the transport of synaptic vesicles and protein secretion [413]. Interestingly, biallelic loss-of-function variants of MYO9A cause AMC [38]. In addition, hemiallelic premature termination codon of MYO9A causes focal segmental glomerulosclerosis [414].
Pathogenic variants of a succinate transporter (SLC25A1) in mitochondrial inner membrane cause combined D-2- and L-2-hydroxyglutaric aciduria (D2L2AD) [415] and CMS. Pathogenic variants of SLC25A1 are predicted to compromise metabolisms of lipid, sterol synthesis, gluconeogenesis, and glycolysis [416], which somehow leads to the development of CMS. Knockdown of Slc25a1 in zebrafish causes aberration in the axonal elongation of spinal motor neurons and compromise the formation of the NMJ [417]. SLC25A1-CMS is thus predicted to be caused by presynaptic defects.

4.11.2. Clinical Features and Therapies

MYO9A-CMS was reported in 3 patients in 2 pedigrees in 2016 [412]. Patients were initially noted with reduced fetal movement before birth and palpebral ptosis at birth. Patients later developed swallowing difficulty, distal and proximal muscle weakness, episodic apnea, respiratory insufficiency, and external ophthalmoplegia. In two patients in a single pedigree, the presence of nystagmus was documented. All patients had developmental delay. ChEI was effective, and combination of ChEI and amifampridine showed marked effects in a patient. However, in a single patient, combination of amifampridine and fluoxetine induced respiratory crisis.
SLC25A1-CMS has been reported in 19 patients in 10 pedigrees since 2014 [417,418,419,420,421]. Limb myasthenia and palpebral ptosis are shared features. External ocular muscles, bulbar muscles, and respiratory muscles are sometimes affected. Developmental delays were also sometimes noted. ChEIs and amifampridine are generally ineffective but are slight effective in some patients.

4.12. CMS Caused by Defective Nuclear Membrane Protein (TOR1AIP1)

4.12.1. Pathomechanisms

LAP1 is a ubiquitously expressed inner nuclear membrane protein. Its N-terminal domain interacts with A-type lamins and emerin in the nucleoplasm [422]. Its C-terminal luminal domain interacts with and activates nucleoplasmic TorsinA, an ATPase for the ATPases associated with diverse cellular activities (AAA+) [423]. Knockout of Tor1aip1 in mouse shows endplate AChR deficiency with markedly increased number of myonuclei at the NMJ. Loss-of-function variants of TOR1AIP1 were previously reported to cause limb-girdle muscular dystrophy or dystonia, with cardiomyopathy or a severe multisystem disorder [424,425,426]. Thus, CMS is a novel phenotype caused by pathogenic variants of TOR1AIP1. It is interesting to note that pathogenic variants of LMNA encoding another nucleolar membrane protein, lamin A, also cause multiple disease phenotypes.

4.12.2. Clinical Features and Therapies

TOR1AIP1-CMS was reported in two adult siblings in 2020 [427] and three adult siblings in 2022 [428]. All patients were noted with mild to moderate muscle weakness and myasthenia in limb muscles and took a slowly progressive or stable course. ChEIs were effective [427,428], and addition of salbutamol (albuterol) had no effect [427].

4.13. CMS Caused by Defective Chromatin Remodeling Protein (CHD8)

4.13.1. Pathomechanisms

CHD8 is one of ATP-dependent chromatin-remodeling enzymes but binds to β-catenin and suppresses the transcription of target genes of β-catenin [429,430,431]. CHD8 is accumulated at the NMJ and binds to rapsyn through β-catenin [432]. Thus, either transcriptional suppression of β-catenin-target genes or suppressed interaction between β-catenin and rapsyn is likely to account for CHD8-CMS [432]. In addition, knockout of Ctnnb1 encoding β-catenin (βCAT) that binds to CHD8 impairs AChR clustering and release of ACh from the nerve terminal [433]. In Drosophila, Kis, a homolog of CHD8, promoted presynaptic endocytosis at the NMJ [434]. Similarly, in C. elegans, a loss-of-function of Chd8 caused reduced synaptic vesicle recycling [435]. As sated below, a marked effect of amifampridine and lack of effects of ChEIs and salbutamol (albuterol) are also consistent with the notion that the major defect in CHD8-CMS is at the motor nerve terminal [432].

4.13.2. Clinical Features and Therapies

Monozygotic female twins with CHD8-CMS were reported in 2020 [432]. Patients showed neonatal onset respiratory distress, palpebral ptosis, and limb muscle weakness. At age 14 years, when the patients were reported, they showed frequent falling attacks and myasthenia, as well as rapidly progressive scoliosis. ChEI and salbutamol (albuterol) showed no effect, but amifampridine was markedly effective [432]. Hemiallelic loss-of-function variants of CHD8 are also reported in intellectual developmental disorder with autism and macrocephaly (IDDAM) [436,437]. The authors of CHD8-CMS stated as personal communications that muscle hypotonia and muscle weakness were observed in 4 out of 66 patients with pathogenic variants of CHD8 in IDDAM [432].

4.14. CMS in PURA Syndrome (PURA)

4.14.1. Pathomechanisms

Purine-rich element-binding protein A (PURA, PURA) is involved in DNA replication, transcription, RNA transport, and mRNA translation, and is conserved across species. PURA plays essential roles in brain development, synapse formation, and proliferation of neuronal and glial cells. Hemiallelic loss-of-function variants of PURA were identified in 11 out of 2117 patients with neurodevelopmental delay in 2014 [438] and thereafter [439,440,441]. Analysis of 32 patients in the authors’ cohort and review of 22 reported patients with PURA syndrome showed that all patients had moderate to severe intellectual disability and neonate-onset symptoms including hypotonia (96%), respiratory problems (57%), feeding difficulties (77%), exaggerated startle response (44%), hypersomnolence (66%), hypothermia (35%), epilepsy (54%), gastrointestinal problems (69%), ophthalmological problems (51%), and endocrine problems (42%) [441]. PURA is expressed in many tissues and has many roles. The exact defects at the NMJ remain undetermined.

4.14.2. Clinical Features and Therapies

Three patients with PURA-CMS showing fluctuating muscle weakness were reported in 2022 [55,442]. One patient showed decremental CAMP, as well as R-CMAP that was much higher than that observed in COLQ-CMS and SCCMS [55]. Another patient showed decremental CAMP followed by incremental CMAP [55]. The third patient showed no decremental CMAP at 3 Hz nerve stimulation but showed non-significant incremental CMAP at 30 Hz stimulation [442]. Two patients were neonates [55,442] and the other was 5 years old [55]. The 5-year-old patient became free of symptoms indicating defective NMJ signal transmission after age 2 years. In a patient, ChEI was ineffective, but salbutamol (albuterol) was effective and unnecessitated non-invasive positive pressure ventilation (NIPPV) because of amelioration of episodic apnea [55]. In another patient, ChEI had markedly ameliorated motor deficits [442].

5. Conclusions

CMS is a group of heterogenous disorders with highly variable clinical phenotypes that require specific treatment for specific pathomechanisms (Table 1). A total of 35 genes have been identified to cause CMS. Clinically overt major phenotypes of recently identified TOR1AIP1-CMS, CHD8-CMS, and PURA-CMS, as well as GMPPB-CMS reported in 2015 and PLEC-CMS in 1989 are not pure CMS. Indeed, pathogenic variants of these genes were initially reported to cause other diseases. Although not all these patients show defective NMJ signal transmission, the presence of defective NMJ was noted by detailed clinical and electrophysiological examinations. Scrutinizing analysis of the NMJ in other diseases may disclose additional groups of CMS in the future.

Author Contributions

Conceptualization, K.O. and A.G.E. Original draft preparation, K.O. Review and editing, B.O., X.-M.S., D.S. and A.G.E. All authors have read and agreed to the published version of the manuscript.

Funding

The works performed in the authors’ laboratories were supported by the National Institute of Neurological Disorders and Stroke (R01 NS109491 to XMS, DS, AGE) and by Grants-in-Aid from the Japan Society for the Promotion of Science (JP20H03561 to KO; JP21K06392 to BO); the Ministry of Health, Labor and Welfare of Japan (20FC1036 to KO); and the National Center of Neurology and Psychiatry (2-5 to KO).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AChacetylcholine
AChEacetylcholinesterase
AChRacetylcholine receptor
ALG14asparagine-linked glycosylation 14 homolog
ALG2asparagine-linked glycosylation 2 homolog
AMCarthrogryposis multiplex congenita
ASDautism spectrum disorder
CaV1.1L-type calcium channel
CDGcongenital disorder of glycosylation
ChATcholine acetyltransferase
CHD8chromatin helicase DNA binding protein 8
ChEIcholinesterase inhibitor
ChThigh affinity choline transporter
CKcreatine kinase
CMAPcompound muscle action potential
CMScongenital myasthenic syndromes
CNScentral nervous system
ColQcollagen Q
CTDC-terminal domain of ColQ
Ctgfconnective tissue growth factor
D2L2ADcombined D-2- and L-2-hydroxyglutaric aciduria
DHMN7Aautosomal dominant distal hereditary motor neuropathy type VIIA
DHPRdihydropyridine receptor
Dok-7docking protein 7
DPAGT1dolichyl-phosphate N-acetylglucosamine phosphotransferase 1
EBSepidermolysis bullosa simplex
EPPendplate potential
EVMPSEscobar variant of multiple pterygium syndrome
FADSfetal akinesia deformation sequence
FCCMSfast-channel congenital myasthenic syndrome
Fgf18fibroblast growth factor 18
GAPRho GTPase-activating protein
GFPT1glutamine--fructose-6-phosphate transaminase 1
GMPPBGDP-mannose pyrophosphorylase B
HCShypotonia-cystinuria syndrome
HSPheparan sulfate proteoglycan
HSPBDHSP-binding domain
IBMinclusion body myositis
IDDAMintellectual developmental disorder with autism and macrocephaly
LAP1lamin-associated protein 1
LEMSLambert-Eaton myasthenic syndrome
LGMD17autosomal recessive limb-girdle muscular dystrophy 17
LMPSlethal form of multiple pterygium syndrome
LRP4low density lipoprotein receptor-related protein 4
MDDGmuscular dystrophy-dystroglycanopathy
MGmyasthenia gravis
MuSKmuscle-specific receptor tyrosine kinase
Myo9amyosin 9A
NaV1.4Skeletal muscle voltage-gated sodium channel
NCAMneuronal cell adhesion molecule
NIPPVnon-invasive positive pressure ventilation
NMJneuromuscular junction
PRADproline-rich attachment domain
PREPLpropyl endopeptidase like
PURApurine-rich element-binding protein A
R-CMAPrepetitive compound muscle action potential
RNSrepetitive nerve stimulation
Rspo2R-spondin 2
RyRryanodine receptor
SCCMSslow-schannel congenital myasthenic syndrome
SFEMGsingle-fiber electromyography
SIDSsudden infantile death syndrome
SNAREsoluble NSF attachment protein receptor
SNVsingle nucleotide variant
SOST2sclerosteosis type 2
SRsarcoplasmic reticulum
SSRIselective serotonin reuptake inhibitor
SV2Asynaptic vesicle glycoprotein 2A
UDP-GlcNAcuridine diphosphate N-acetylglucosamine
vAChTvesicular acetylcholine transporter
VAMP1vesicle-associated membrane protein 1
VGCCvoltage-gated calcium channel

References

  1. Ohno, K.; Ohkawara, B.; Ito, M.; Engel, A.G. Molecular Genetics of Congenital Myasthenic Syndromes. In eLS; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2014. [Google Scholar] [CrossRef]
  2. Engel, A.G.; Shen, X.M.; Selcen, D.; Sine, S.M. Congenital myasthenic syndromes: Pathogenesis, diagnosis, and treatment. Lancet Neurol. 2015, 14, 420–434. [Google Scholar] [CrossRef]
  3. Heintz-Buschart, A.; Pandey, U.; Wicke, T.; Sixel-Doring, F.; Janzen, A.; Sittig-Wiegand, E.; Trenkwalder, C.; Oertel, W.H.; Mollenhauer, B.; Wilmes, P. The nasal and gut microbiome in Parkinson’s disease and idiopathic rapid eye movement sleep behavior disorder. Mov. Disord. 2018, 33, 88–98. [Google Scholar] [CrossRef]
  4. Byring, R.F.; Pihko, H.; Tsujino, A.; Shen, X.M.; Gustafsson, B.; Hackman, P.; Ohno, K.; Engel, A.G.; Udd, B. Congenital myasthenic syndrome associated with episodic apnea and sudden infant death. Neuromuscul. Disord. 2002, 12, 548–553. [Google Scholar] [CrossRef]
  5. Mannikko, R.; Wong, L.; Tester, D.J.; Thor, M.G.; Sud, R.; Kullmann, D.M.; Sweeney, M.G.; Leu, C.; Sisodiya, S.M.; FitzPatrick, D.R.; et al. Dysfunction of NaV1.4, a skeletal muscle voltage-gated sodium channel, in sudden infant death syndrome: A case-control study. Lancet 2018, 391, 1483–1492. [Google Scholar] [CrossRef]
  6. Nicole, S.; Chaouch, A.; Torbergsen, T.; Bauche, S.; de Bruyckere, E.; Fontenille, M.J.; Horn, M.A.; van Ghelue, M.; Loseth, S.; Issop, Y.; et al. Agrin mutations lead to a congenital myasthenic syndrome with distal muscle weakness and atrophy. Brain 2014, 137 Pt 9, 2429–2443. [Google Scholar] [CrossRef]
  7. Rodriguez Cruz, P.M.; Belaya, K.; Basiri, K.; Sedghi, M.; Farrugia, M.E.; Holton, J.L.; Liu, W.W.; Maxwell, S.; Petty, R.; Walls, T.J.; et al. Clinical features of the myasthenic syndrome arising from mutations in GMPPB. J. Neurol. Neurosurg. Psychiatry 2016, 87, 802–809. [Google Scholar] [CrossRef]
  8. Ohno, K. Is the serum creatine kinase level elevated in congenital myasthenic syndrome? J. Neurol. Neurosurg. Psychiatry 2016, 87, 801. [Google Scholar] [CrossRef] [PubMed]
  9. Whittaker, R.G.; Herrmann, D.N.; Bansagi, B.; Hasan, B.A.; Lofra, R.M.; Logigian, E.L.; Sowden, J.E.; Almodovar, J.L.; Littleton, J.T.; Zuchner, S.; et al. Electrophysiologic features of SYT2 mutations causing a treatable neuromuscular syndrome. Neurology 2015, 85, 1964–1971. [Google Scholar] [CrossRef]
  10. Shen, X.M.; Scola, R.H.; Lorenzoni, P.J.; Kay, C.S.; Werneck, L.C.; Brengman, J.; Selcen, D.; Engel, A.G. Novel synaptobrevin-1 mutation causes fatal congenital myasthenic syndrome. Ann. Clin. Transl. Neurol. 2017, 4, 130–138. [Google Scholar] [CrossRef]
  11. Engel, A.G.; Selcen, D.; Shen, X.M.; Milone, M.; Harper, C.M. Loss of MUNC13-1 function causes microcephaly, cortical hyperexcitability, and fatal myasthenia. Neurol. Genet. 2016, 2, e105. [Google Scholar] [CrossRef]
  12. Maselli, R.A.; Vazquez, J.; Schrumpf, L.; Arredondo, J.; Lara, M.; Strober, J.B.; Pytel, P.; Wollmann, R.L.; Ferns, M. Presynaptic congenital myasthenic syndrome with altered synaptic vesicle homeostasis linked to compound heterozygous sequence variants in RPH3A. Mol. Genet. Genom. Med. 2018, 6, 434–440. [Google Scholar] [CrossRef]
  13. Maselli, R.A.; Arredondo, J.; Vazquez, J.; Chong, J.X.; University of Washington Center for Mendelian, G.; Bamshad, M.J.; Nickerson, D.A.; Lara, M.; Ng, F.; Lo, V.L.; et al. Presynaptic congenital myasthenic syndrome with a homozygous sequence variant in LAMA5 combines myopia, facial tics, and failure of neuromuscular transmission. Am. J. Med. Genet. A 2017, 173, 2240–2245. [Google Scholar] [CrossRef] [PubMed]
  14. Nicolau, S.; Milone, M. The Electrophysiology of Presynaptic Congenital Myasthenic Syndromes With and Without Facilitation: From Electrodiagnostic Findings to Molecular Mechanisms. Front. Neurol. 2019, 10, 257. [Google Scholar] [CrossRef]
  15. Shen, X.M.; Selcen, D.; Brengman, J.; Engel, A.G. Mutant SNAP25B causes myasthenia, cortical hyperexcitability, ataxia, and intellectual disability. Neurology 2014, 83, 2247–2255. [Google Scholar] [CrossRef] [PubMed]
  16. Ohno, K.; Anlar, B.; Ozdirim, E.; Brengman, J.M.; DeBleecker, J.L.; Engel, A.G. Myasthenic syndromes in Turkish kinships due to mutations in the acetylcholine receptor. Ann. Neurol. 1998, 44, 234–241. [Google Scholar] [CrossRef] [PubMed]
  17. Guergueltcheva, V.; Muller, J.S.; Dusl, M.; Senderek, J.; Oldfors, A.; Lindbergh, C.; Maxwell, S.; Colomer, J.; Mallebrera, C.J.; Nascimento, A.; et al. Congenital myasthenic syndrome with tubular aggregates caused by GFPT1 mutations. J. Neurol. 2012, 259, 838–850. [Google Scholar] [CrossRef]
  18. Huh, S.Y.; Kim, H.S.; Jang, H.J.; Park, Y.E.; Kim, D.S. Limb-girdle myasthenia with tubular aggregates associated with novel GFPT1 mutations. Muscle Nerve 2012, 46, 600–604. [Google Scholar] [CrossRef]
  19. Selcen, D.; Shen, X.M.; Milone, M.; Brengman, J.; Ohno, K.; Deymeer, F.; Finkel, R.; Rowin, J.; Engel, A.G. GFPT1-myasthenia: Clinical, structural, and electrophysiologic heterogeneity. Neurology 2013, 81, 370–378. [Google Scholar] [CrossRef]
  20. Bauche, S.; Vellieux, G.; Sternberg, D.; Fontenille, M.J.; De Bruyckere, E.; Davoine, C.S.; Brochier, G.; Messeant, J.; Wolf, L.; Fardeau, M.; et al. Mutations in GFPT1-related congenital myasthenic syndromes are associated with synaptic morphological defects and underlie a tubular aggregate myopathy with synaptopathy. J. Neurol. 2017, 264, 1791–1803. [Google Scholar] [CrossRef] [PubMed]
  21. Belaya, K.; Finlayson, S.; Slater, C.R.; Cossins, J.; Liu, W.W.; Maxwell, S.; McGowan, S.J.; Maslau, S.; Twigg, S.R.; Walls, T.J.; et al. Mutations in DPAGT1 cause a limb-girdle congenital myasthenic syndrome with tubular aggregates. Am. J. Hum. Genet. 2012, 91, 193–201. [Google Scholar] [CrossRef]
  22. Finlayson, S.; Palace, J.; Belaya, K.; Walls, T.J.; Norwood, F.; Burke, G.; Holton, J.L.; Pascual-Pascual, S.I.; Cossins, J.; Beeson, D. Clinical features of congenital myasthenic syndrome due to mutations in DPAGT1. J. Neurol. Neurosurg. Psychiatry 2013, 84, 1119–1125. [Google Scholar] [CrossRef]
  23. Selcen, D.; Shen, X.M.; Brengman, J.; Li, Y.; Stans, A.A.; Wieben, E.; Engel, A.G. DPAGT1 myasthenia and myopathy: Genetic, phenotypic, and expression studies. Neurology 2014, 82, 1822–1830. [Google Scholar] [CrossRef]
  24. Cossins, J.; Belaya, K.; Hicks, D.; Salih, M.A.; Finlayson, S.; Carboni, N.; Liu, W.W.; Maxwell, S.; Zoltowska, K.; Farsani, G.T.; et al. Congenital myasthenic syndromes due to mutations in ALG2 and ALG14. Brain 2013, 136 Pt 3, 944–956. [Google Scholar] [CrossRef]
  25. Belaya, K.; Rodriguez Cruz, P.M.; Liu, W.W.; Maxwell, S.; McGowan, S.; Farrugia, M.E.; Petty, R.; Walls, T.J.; Sedghi, M.; Basiri, K.; et al. Mutations in GMPPB cause congenital myasthenic syndrome and bridge myasthenic disorders with dystroglycanopathies. Brain 2015, 138 Pt 9, 2493–2504. [Google Scholar] [CrossRef] [PubMed]
  26. Engel, A.G.; Lambert, E.H.; Mulder, D.M.; Torres, C.F.; Sahashi, K.; Bertorini, T.E.; Whitaker, J.N. A newly recognized congenital myasthenic syndrome attributed to a prolonged open time of the acetylcholine-induced ion channel. Ann. Neurol. 1982, 11, 553–569. [Google Scholar] [CrossRef]
  27. Fidzianska, A.; Ryniewicz, B.; Shen, X.M.; Engel, A.G. IBM-type inclusions in a patient with slow-channel syndrome caused by a mutation in the AChR epsilon subunit. Neuromuscul. Disord. 2005, 15, 753–759. [Google Scholar] [CrossRef] [PubMed]
  28. Carss, K.J.; Stevens, E.; Foley, A.R.; Cirak, S.; Riemersma, M.; Torelli, S.; Hoischen, A.; Willer, T.; van Scherpenzeel, M.; Moore, S.A.; et al. Mutations in GDP-mannose pyrophosphorylase B cause congenital and limb-girdle muscular dystrophies associated with hypoglycosylation of alpha-dystroglycan. Am. J. Hum. Genet. 2013, 93, 29–41. [Google Scholar] [CrossRef]
  29. Nicolau, S.; Liewluck, T.; Shen, X.M.; Selcen, D.; Engel, A.G.; Milone, M. A homozygous mutation in GMPPB leads to centronuclear myopathy with combined pre- and postsynaptic defects of neuromuscular transmission. Neuromuscul. Disord. 2019, 29, 614–617. [Google Scholar] [CrossRef] [PubMed]
  30. Souza, P.V.; Batistella, G.N.; Lino, V.C.; Pinto, W.B.; Annes, M.; Oliveira, A.S. Clinical and genetic basis of congenital myasthenic syndromes. Arq. Neuropsiquiatr. 2016, 74, 750–760. [Google Scholar] [CrossRef]
  31. Kao, J.C.; Milone, M.; Selcen, D.; Shen, X.M.; Engel, A.G.; Liewluck, T. Congenital myasthenic syndromes in adult neurology clinic: A long road to diagnosis and therapy. Neurology 2018, 91, e1770–e1777. [Google Scholar] [CrossRef]
  32. Lorenzoni, P.J.; Ducci, R.D.; Arndt, R.C.; Hrysay, N.M.C.; Fustes, O.J.H.; Topf, A.; Lochmuller, H.; Werneck, L.C.; Kay, C.S.K.; Scola, R.H. Congenital myasthenic syndrome in a cohort of patients with ‘double’ seronegative myasthenia gravis. Arq. Neuropsiquiatr. 2022, 80, 69–74. [Google Scholar] [CrossRef]
  33. Regal, L.; Martensson, E.; Maystadt, I.; Voermans, N.; Lederer, D.; Burlina, A.; Juan Fita, M.J.; Hoogeboom, A.J.M.; Olsson Engman, M.; Hollemans, T.; et al. PREPL deficiency: Delineation of the phenotype and development of a functional blood assay. Genet. Med. 2018, 20, 109–118. [Google Scholar] [CrossRef]
  34. Niles, K.M.; Blaser, S.; Shannon, P.; Chitayat, D. Fetal arthrogryposis multiplex congenita/fetal akinesia deformation sequence (FADS)-Aetiology, diagnosis, and management. Prenat. Diagn. 2019, 39, 720–731. [Google Scholar] [CrossRef] [PubMed]
  35. Morgan, N.V.; Brueton, L.A.; Cox, P.; Greally, M.T.; Tolmie, J.; Pasha, S.; Aligianis, I.A.; van Bokhoven, H.; Marton, T.; Al-Gazali, L.; et al. Mutations in the embryonal subunit of the acetylcholine receptor (CHRNG) cause lethal and Escobar variants of multiple pterygium syndrome. Am. J. Hum. Genet. 2006, 79, 390–395. [Google Scholar] [CrossRef]
  36. Hoffmann, K.; Muller, J.S.; Stricker, S.; Megarbane, A.; Rajab, A.; Lindner, T.H.; Cohen, M.; Chouery, E.; Adaimy, L.; Ghanem, I.; et al. Escobar syndrome is a prenatal myasthenia caused by disruption of the acetylcholine receptor fetal gamma subunit. Am. J. Hum. Genet. 2006, 79, 303–312. [Google Scholar] [CrossRef] [PubMed]
  37. Seo, J.; Choi, I.H.; Lee, J.S.; Yoo, Y.; Kim, N.K.; Choi, M.; Ko, J.M.; Shin, Y.B. Rare cases of congenital arthrogryposis multiplex caused by novel recurrent CHRNG mutations. J. Hum. Genet. 2015, 60, 213–215. [Google Scholar] [CrossRef]
  38. Bayram, Y.; Karaca, E.; Coban Akdemir, Z.; Yilmaz, E.O.; Tayfun, G.A.; Aydin, H.; Torun, D.; Bozdogan, S.T.; Gezdirici, A.; Isikay, S.; et al. Molecular etiology of arthrogryposis in multiple families of mostly Turkish origin. J. Clin. Investig. 2016, 126, 762–778. [Google Scholar] [CrossRef]
  39. Vogt, J.; Harrison, B.J.; Spearman, H.; Cossins, J.; Vermeer, S.; ten Cate, L.N.; Morgan, N.V.; Beeson, D.; Maher, E.R. Mutation analysis of CHRNA1, CHRNB1, CHRND, and RAPSN genes in multiple pterygium syndrome/fetal akinesia patients. Am. J. Hum. Genet. 2008, 82, 222–227. [Google Scholar] [CrossRef] [PubMed]
  40. Winters, L.; Van Hoof, E.; De Catte, L.; Van Den Bogaert, K.; de Ravel, T.; De Waele, L.; Corveleyn, A.; Breckpot, J. Massive parallel sequencing identifies RAPSN and PDHA1 mutations causing fetal akinesia deformation sequence. Eur. J. Paediatr. Neurol. 2017, 21, 745–753. [Google Scholar] [CrossRef]
  41. Bauche, S.; O’Regan, S.; Azuma, Y.; Laffargue, F.; McMacken, G.; Sternberg, D.; Brochier, G.; Buon, C.; Bouzidi, N.; Topf, A.; et al. Impaired Presynaptic High-Affinity Choline Transporter Causes a Congenital Myasthenic Syndrome with Episodic Apnea. Am. J. Hum. Genet. 2016, 99, 753–761. [Google Scholar] [CrossRef]
  42. Parr, J.R.; Andrew, M.J.; Finnis, M.; Beeson, D.; Vincent, A.; Jayawant, S. How common is childhood myasthenia? The UK incidence and prevalence of autoimmune and congenital myasthenia. Arch. Dis. Child. 2014, 99, 539–542. [Google Scholar] [CrossRef]
  43. Mihaylova, V.; Scola, R.H.; Gervini, B.; Lorenzoni, P.J.; Kay, C.K.; Werneck, L.C.; Stucka, R.; Guergueltcheva, V.; von der Hagen, M.; Huebner, A.; et al. Molecular characterisation of congenital myasthenic syndromes in Southern Brazil. J. Neurol. Neurosurg. Psychiatry 2010, 81, 973–977. [Google Scholar] [CrossRef]
  44. Troha Gergeli, A.; Neubauer, D.; Golli, T.; Butenko, T.; Loboda, T.; Maver, A.; Osredkar, D. Prevalence and genetic subtypes of congenital myasthenic syndromes in the pediatric population of Slovenia. Eur. J. Paediatr. Neurol. 2020, 26, 34–38. [Google Scholar] [CrossRef]
  45. Natera-de Benito, D.; Topf, A.; Vilchez, J.J.; Gonzalez-Quereda, L.; Dominguez-Carral, J.; Diaz-Manera, J.; Ortez, C.; Bestue, M.; Gallano, P.; Dusl, M.; et al. Molecular characterization of congenital myasthenic syndromes in Spain. Neuromuscul. Disord. 2017, 27, 1087–1098. [Google Scholar] [CrossRef]
  46. Richard, P.; Gaudon, K.; Andreux, F.; Yasaki, E.; Prioleau, C.; Bauche, S.; Barois, A.; Ioos, C.; Mayer, M.; Routon, M.C.; et al. Possible founder effect of rapsyn N88K mutation and identification of novel rapsyn mutations in congenital myasthenic syndromes. J. Med. Genet. 2003, 40, e81. [Google Scholar] [CrossRef]
  47. Dunne, V.; Maselli, R.A. Common founder effect of rapsyn N88K studied using intragenic markers. J. Hum. Genet. 2004, 49, 366–369. [Google Scholar] [CrossRef]
  48. Muller, J.S.; Abicht, A.; Burke, G.; Cossins, J.; Richard, P.; Baumeister, S.K.; Stucka, R.; Eymard, B.; Hantai, D.; Beeson, D.; et al. The congenital myasthenic syndrome mutation RAPSN N88K derives from an ancient Indo-European founder. J. Med. Genet. 2004, 41, e104. [Google Scholar] [CrossRef]
  49. Ohno, K.; Engel, A.G. Lack of founder haplotype for the rapsyn N88K mutation: N88K is an ancient founder mutation or arises from multiple founders. J. Med. Genet. 2004, 41, e8. [Google Scholar] [CrossRef]
  50. Abicht, A.; Dusl, M.; Gallenmuller, C.; Guergueltcheva, V.; Schara, U.; Della Marina, A.; Wibbeler, E.; Almaras, S.; Mihaylova, V.; von der Hagen, M.; et al. Congenital myasthenic syndromes: Achievements and limitations of phenotype-guided gene-after-gene sequencing in diagnostic practice: A study of 680 patients. Hum. Mutat. 2012, 33, 1474–1484. [Google Scholar] [CrossRef]
  51. Croxen, R.; Newland, C.; Betty, M.; Vincent, A.; Newsom-Davis, J.; Beeson, D. Novel functional epsilon-subunit polypeptide generated by a single nucleotide deletion in acetylcholine receptor deficiency congenital myasthenic syndrome. Ann. Neurol. 1999, 46, 639–647. [Google Scholar] [CrossRef]
  52. Polavarapu, K.; Mathur, A.; Joshi, A.; Nashi, S.; Preethish-Kumar, V.; Bardhan, M.; Sharma, P.; Parveen, S.; Seth, M.; Vengalil, S.; et al. A founder mutation in the GMPPB gene [c.1000G > A (p.Asp334Asn)] causes a mild form of limb-girdle muscular dystrophy/congenital myasthenic syndrome (LGMD/CMS) in South Indian patients. Neurogenetics 2021, 22, 271–285. [Google Scholar] [CrossRef]
  53. Mroczek, M.; Durmus, H.; Topf, A.; Parman, Y.; Straub, V. Four Individuals with a Homozygous Mutation in Exon 1f of the PLEC Gene and Associated Myasthenic Features. Genes 2020, 11, 716. [Google Scholar] [CrossRef] [PubMed]
  54. Reynolds, H.M.; Wen, T.; Farrell, A.; Mao, R.; Moore, B.; Boyden, S.E.; Bayrak-Toydemir, P.; Nicholas, T.J.; Rynearson, S.; Holt, C.; et al. Rapid genome sequencing identifies a novel de novo SNAP25 variant for neonatal congenital myasthenic syndrome. Cold Spring Harb. Mol. Case Stud. 2022, 8, a006242. [Google Scholar] [CrossRef]
  55. Qashqari, H.; McNiven, V.; Gonorazky, H.; Mendoza-Londono, R.; Hassan, A.; Kulkarni, T.; Amburgey, K.; Dowling, J.J. PURA syndrome: Neuromuscular junction manifestations with potential therapeutic implications. Neuromuscul. Disord. 2022, 32, 842–844. [Google Scholar] [CrossRef] [PubMed]
  56. Herrmann, D.N.; Horvath, R.; Sowden, J.E.; Gonzalez, M.; Sanchez-Mejias, A.; Guan, Z.; Whittaker, R.G.; Almodovar, J.L.; Lane, M.; Bansagi, B.; et al. Synaptotagmin 2 mutations cause an autosomal-dominant form of Lambert-Eaton myasthenic syndrome and nonprogressive motor neuropathy. Am. J. Hum. Genet. 2014, 95, 332–339. [Google Scholar] [CrossRef] [PubMed]
  57. Montes-Chinea, N.I.; Guan, Z.; Coutts, M.; Vidal, C.; Courel, S.; Rebelo, A.P.; Abreu, L.; Zuchner, S.; Littleton, J.T.; Saporta, M.A. Identification of a new SYT2 variant validates an unusual distal motor neuropathy phenotype. Neurol. Genet. 2018, 4, e282. [Google Scholar] [CrossRef] [PubMed]
  58. Donkervoort, S.; Mohassel, P.; Laugwitz, L.; Zaki, M.S.; Kamsteeg, E.J.; Maroofian, R.; Chao, K.R.; Verschuuren-Bemelmans, C.C.; Horber, V.; Fock, A.J.M.; et al. Biallelic loss of function variants in SYT2 cause a treatable congenital onset presynaptic myasthenic syndrome. Am. J. Med. Genet. A 2020, 182, 2272–2283. [Google Scholar] [CrossRef]
  59. Maselli, R.A.; van der Linden, H., Jr.; Ferns, M. Recessive congenital myasthenic syndrome caused by a homozygous mutation in SYT2 altering a highly conserved C-terminal amino acid sequence. Am. J. Med. Genet. A 2020, 182, 1744–1749. [Google Scholar] [CrossRef]
  60. Maselli, R.A.; Wei, D.T.; Hodgson, T.S.; Sampson, J.B.; Vazquez, J.; Smith, H.L.; Pytel, P.; Ferns, M. Dominant and recessive congenital myasthenic syndromes caused by SYT2 mutations. Muscle Nerve 2021, 64, 219–224. [Google Scholar] [CrossRef]
  61. Schara, U.; Lochmuller, H. Therapeutic strategies in congenital myasthenic syndromes. Neurotherapeutics 2008, 5, 542–547. [Google Scholar] [CrossRef]
  62. Wargon, I.; Richard, P.; Kuntzer, T.; Sternberg, D.; Nafissi, S.; Gaudon, K.; Lebail, A.; Bauche, S.; Hantai, D.; Fournier, E.; et al. Long-term follow-up of patients with congenital myasthenic syndrome caused by COLQ mutations. Neuromuscul. Disord. 2012, 22, 318–324. [Google Scholar] [CrossRef] [PubMed]
  63. Yis, U.; Becker, K.; Kurul, S.H.; Uyanik, G.; Bayram, E.; Haliloglu, G.; Polat, A.I.; Ayanoglu, M.; Okur, D.; Tosun, A.F.; et al. Genetic Landscape of Congenital Myasthenic Syndromes From Turkey: Novel Mutations and Clinical Insights. J. Child Neurol. 2017, 32, 759–765. [Google Scholar] [CrossRef] [PubMed]
  64. Durmus, H.; Shen, X.M.; Serdaroglu-Oflazer, P.; Kara, B.; Parman-Gulsen, Y.; Ozdemir, C.; Brengman, J.; Deymeer, F.; Engel, A.G. Congenital myasthenic syndromes in Turkey: Clinical clues and prognosis with long term follow-up. Neuromuscul. Disord. 2018, 28, 315–322. [Google Scholar] [CrossRef]
  65. Maselli, R.A.; Ng, J.J.; Anderson, J.A.; Cagney, O.; Arredondo, J.; Williams, C.; Wessel, H.B.; Abdel-Hamid, H.; Wollmann, R.L. Mutations in LAMB2 causing a severe form of synaptic congenital myasthenic syndrome. J. Med. Genet. 2009, 46, 203–208. [Google Scholar] [CrossRef]
  66. Ben Ammar, A.; Petit, F.; Alexandri, N.; Gaudon, K.; Bauche, S.; Rouche, A.; Gras, D.; Fournier, E.; Koenig, J.; Stojkovic, T.; et al. Phenotype genotype analysis in 15 patients presenting a congenital myasthenic syndrome due to mutations in DOK7. J. Neurol. 2010, 257, 754–766. [Google Scholar] [CrossRef] [PubMed]
  67. Lashley, D.; Palace, J.; Jayawant, S.; Robb, S.; Beeson, D. Ephedrine treatment in congenital myasthenic syndrome due to mutations in DOK7. Neurology 2010, 74, 1517–1523. [Google Scholar] [CrossRef] [PubMed]
  68. Khadilkar, S.; Bhutada, A.; Nallamilli, B.; Hegde, M. Limb girdle weakness responding to salbutamol: An Indian family with DOK7 mutation. Indian Pediatr. 2015, 52, 243–244. [Google Scholar] [CrossRef]
  69. Lozowska, D.; Ringel, S.P.; Winder, T.L.; Liu, J.; Liewluck, T. Anticholinesterase Therapy Worsening Head Drop and Limb Weakness Due to a Novel DOK7 Mutation. J. Clin. Neuromuscul. Dis. 2015, 17, 72–77. [Google Scholar] [CrossRef]
  70. Luan, X.; Tian, W.; Cao, L. Limb-girdle congenital myasthenic syndrome in a Chinese family with novel mutations in MUSK gene and literature review. Clin. Neurol. Neurosurg. 2016, 150, 41–45. [Google Scholar] [CrossRef]
  71. Ohkawara, B.; Cabrera-Serrano, M.; Nakata, T.; Milone, M.; Asai, N.; Ito, K.; Ito, M.; Masuda, A.; Ito, Y.; Engel, A.G.; et al. LRP4 third beta-propeller domain mutations cause novel congenital myasthenia by compromising agrin-mediated MuSK signaling in a position-specific manner. Hum. Mol. Genet. 2014, 23, 1856–1868. [Google Scholar] [CrossRef]
  72. Khan, M.M.; Lustrino, D.; Silveira, W.A.; Wild, F.; Straka, T.; Issop, Y.; O’Connor, E.; Cox, D.; Reischl, M.; Marquardt, T.; et al. Sympathetic innervation controls homeostasis of neuromuscular junctions in health and disease. Proc. Natl. Acad. Sci. USA 2016, 113, 746–750. [Google Scholar] [CrossRef]
  73. Guven, A.; Demirci, M.; Anlar, B. Recurrent COLQ mutation in congenital myasthenic syndrome. Pediatr. Neurol. 2012, 46, 253–256. [Google Scholar] [CrossRef] [PubMed]
  74. Fukudome, T.; Ohno, K.; Brengman, J.M.; Engel, A.G. Quinidine normalizes the open duration of slow-channel mutants of the acetylcholine receptor. Neuroreport 1998, 9, 1907–1911. [Google Scholar] [CrossRef]
  75. Harper, C.M.; Engel, A.G. Quinidine sulfate therapy for the slow-channel congenital myasthenic syndrome. Ann. Neurol. 1998, 43, 480–484. [Google Scholar] [CrossRef] [PubMed]
  76. Harper, C.M.; Fukodome, T.; Engel, A.G. Treatment of slow-channel congenital myasthenic syndrome with fluoxetine. Neurology 2003, 60, 1710–1713. [Google Scholar] [CrossRef]
  77. Visser, A.C.; Laughlin, R.S.; Litchy, W.J.; Benarroch, E.E.; Milone, M. Rapsyn congenital myasthenic syndrome worsened by fluoxetine. Muscle Nerve 2017, 55, 131–135. [Google Scholar] [CrossRef]
  78. Vidanagamage, A.; Gooneratne, I.K.; Nandasiri, S.; Gunaratne, K.; Fernando, A.; Maxwell, S.; Cossins, J.; Beeson, D.; Chang, T. A rare mutation in the COLQ gene causing congenital myasthenic syndrome with remarkable improvement to fluoxetine: A case report. Neuromuscul. Disord. 2021, 31, 246–248. [Google Scholar] [CrossRef] [PubMed]
  79. Tsujino, A.; Maertens, C.; Ohno, K.; Shen, X.M.; Fukuda, T.; Harper, C.M.; Cannon, S.C.; Engel, A.G. Myasthenic syndrome caused by mutation of the SCN4A sodium channel. Proc. Natl. Acad. Sci. USA 2003, 100, 7377–7382. [Google Scholar] [CrossRef]
  80. Berghold, V.M.; Koko, M.; Berutti, R.; Plecko, B. Case report: Novel SCN4A variant associated with a severe congenital myasthenic syndrome/myopathy phenotype. Front. Pediatr. 2022, 10, 944784. [Google Scholar] [CrossRef]
  81. Habbout, K.; Poulin, H.; Rivier, F.; Giuliano, S.; Sternberg, D.; Fontaine, B.; Eymard, B.; Morales, R.J.; Echenne, B.; King, L.; et al. A recessive Nav1.4 mutation underlies congenital myasthenic syndrome with periodic paralysis. Neurology 2016, 86, 161–169. [Google Scholar] [CrossRef]
  82. O’Connell, K.; Rooney, T.; Alabaf, S.; Ramdas, S.; Beeson, D.; Palace, J. Pregnancy outcomes in patients with congenital myasthenic syndromes. Muscle Nerve 2022, 66, 345–348. [Google Scholar] [CrossRef] [PubMed]
  83. Desai, R.C.; Vyas, B.; Earles, C.A.; Littleton, J.T.; Kowalchyck, J.A.; Martin, T.F.; Chapman, E.R. The C2B domain of synaptotagmin is a Ca(2+)-sensing module essential for exocytosis. J. Cell Biol. 2000, 150, 1125–1136. [Google Scholar] [CrossRef]
  84. Apparsundaram, S.; Ferguson, S.M.; George, A.L., Jr.; Blakely, R.D. Molecular cloning of a human, hemicholinium-3-sensitive choline transporter. Biochem. Biophys. Res. Commun. 2000, 276, 862–867. [Google Scholar] [CrossRef] [PubMed]
  85. Erickson, J.D.; Varoqui, H.; Schafer, M.K.; Modi, W.; Diebler, M.F.; Weihe, E.; Rand, J.; Eiden, L.E.; Bonner, T.I.; Usdin, T.B. Functional identification of a vesicular acetylcholine transporter and its expression from a “cholinergic” gene locus. J. Biol. Chem. 1994, 269, 21929–21932. [Google Scholar] [CrossRef]
  86. Milone, M.; Wang, H.L.; Ohno, K.; Prince, R.; Fukudome, T.; Shen, X.M.; Brengman, J.M.; Griggs, R.C.; Sine, S.M.; Engel, A.G. Mode switching kinetics produced by a naturally occurring mutation in the cytoplasmic loop of the human acetylcholine receptor epsilon subunit. Neuron 1998, 20, 575–588. [Google Scholar] [CrossRef] [PubMed]
  87. Unwin, N. Refined structure of the nicotinic acetylcholine receptor at 4A resolution. J. Mol. Biol. 2005, 346, 967–989. [Google Scholar] [CrossRef] [PubMed]
  88. Kim, N.; Stiegler, A.L.; Cameron, T.O.; Hallock, P.T.; Gomez, A.M.; Huang, J.H.; Hubbard, S.R.; Dustin, M.L.; Burden, S.J. Lrp4 is a receptor for Agrin and forms a complex with MuSK. Cell 2008, 135, 334–342. [Google Scholar] [CrossRef]
  89. Zhang, B.; Luo, S.; Wang, Q.; Suzuki, T.; Xiong, W.C.; Mei, L. LRP4 serves as a coreceptor of agrin. Neuron 2008, 60, 285–297. [Google Scholar] [CrossRef]
  90. Okada, K.; Inoue, A.; Okada, M.; Murata, Y.; Kakuta, S.; Jigami, T.; Kubo, S.; Shiraishi, H.; Eguchi, K.; Motomura, M.; et al. The muscle protein Dok-7 is essential for neuromuscular synaptogenesis. Science 2006, 312, 1802–1805. [Google Scholar] [CrossRef]
  91. Borges, L.S.; Yechikhov, S.; Lee, Y.I.; Rudell, J.B.; Friese, M.B.; Burden, S.J.; Ferns, M.J. Identification of a motif in the acetylcholine receptor beta subunit whose phosphorylation regulates rapsyn association and postsynaptic receptor localization. J. Neurosci. 2008, 28, 11468–11476. [Google Scholar] [CrossRef]
  92. Xing, G.; Jing, H.; Yu, Z.; Chen, P.; Wang, H.; Xiong, W.C.; Mei, L. Membraneless condensates by Rapsn phase separation as a platform for neuromuscular junction formation. Neuron 2021, 109, 1963–1978 e5. [Google Scholar] [CrossRef]
  93. Wu, H.; Lu, Y.; Shen, C.; Patel, N.; Gan, L.; Xiong, W.C.; Mei, L. Distinct roles of muscle and motoneuron LRP4 in neuromuscular junction formation. Neuron 2012, 75, 94–107. [Google Scholar] [CrossRef] [PubMed]
  94. Yumoto, N.; Kim, N.; Burden, S.J. Lrp4 is a retrograde signal for presynaptic differentiation at neuromuscular synapses. Nature 2012, 489, 438–442. [Google Scholar] [CrossRef]
  95. Nakashima, H.; Ohkawara, B.; Ishigaki, S.; Fukudome, T.; Ito, K.; Tsushima, M.; Konishi, H.; Okuno, T.; Yoshimura, T.; Ito, M.; et al. R-spondin 2 promotes acetylcholine receptor clustering at the neuromuscular junction via Lgr5. Sci. Rep. 2016, 6, 28512. [Google Scholar] [CrossRef] [PubMed]
  96. Li, J.; Ito, M.; Ohkawara, B.; Masuda, A.; Ohno, K. Differential effects of spinal motor neuron-derived and skeletal muscle-derived Rspo2 on acetylcholine receptor clustering at the neuromuscular junction. Sci. Rep. 2018, 8, 13577. [Google Scholar] [CrossRef]
  97. Ito, K.; Ohkawara, B.; Yagi, H.; Nakashima, H.; Tsushima, M.; Ota, K.; Konishi, H.; Masuda, A.; Imagama, S.; Kiyama, H.; et al. Lack of Fgf18 causes abnormal clustering of motor nerve terminals at the neuromuscular junction with reduced acetylcholine receptor clusters. Sci. Rep. 2018, 8, 434. [Google Scholar] [CrossRef]
  98. Ohkawara, B.; Kobayakawa, A.; Kanbara, S.; Hattori, T.; Kubota, S.; Ito, M.; Masuda, A.; Takigawa, M.; Lyons, K.M.; Ishiguro, N.; et al. CTGF/CCN2 facilitates LRP4-mediated formation of the embryonic neuromuscular junction. EMBO Rep. 2020, 21, e48462. [Google Scholar] [CrossRef] [PubMed]
  99. Ohkawara, B.; Ito, M.; Ohno, K. Secreted Signaling Molecules at the Neuromuscular Junction in Physiology and Pathology. Int. J. Mol. Sci. 2021, 22, 2455. [Google Scholar] [CrossRef]
  100. Zhang, W.; Coldefy, A.S.; Hubbard, S.R.; Burden, S.J. Agrin binds to the N-terminal region of Lrp4 protein and stimulates association between Lrp4 and the first immunoglobulin-like domain in muscle-specific kinase (MuSK). J. Biol. Chem. 2011, 286, 40624–40630. [Google Scholar] [CrossRef] [Green Version]
  101. Zong, Y.; Zhang, B.; Gu, S.; Lee, K.; Zhou, J.; Yao, G.; Figueiredo, D.; Perry, K.; Mei, L.; Jin, R. Structural basis of agrin-LRP4-MuSK signaling. Genes Dev. 2012, 26, 247–258. [Google Scholar] [CrossRef]
  102. Otsuka, K.; Ito, M.; Ohkawara, B.; Masuda, A.; Kawakami, Y.; Sahashi, K.; Nishida, H.; Mabuchi, N.; Takano, A.; Engel, A.G.; et al. Collagen Q and anti-MuSK autoantibody competitively suppress agrin/LRP4/MuSK signaling. Sci. Rep. 2015, 5, 13928. [Google Scholar] [CrossRef]
  103. Ohno, K.; Quiram, P.A.; Milone, M.; Wang, H.L.; Harper, M.C.; Pruitt, J.N., 2nd; Brengman, J.M.; Pao, L.; Fischbeck, K.H.; Crawford, T.O.; et al. Congenital myasthenic syndromes due to heteroallelic nonsense/missense mutations in the acetylcholine receptor epsilon subunit gene: Identification and functional characterization of six new mutations. Hum. Mol. Genet. 1997, 6, 753–766. [Google Scholar] [CrossRef]
  104. Ohno, K.; Anlar, B.; Engel, A.G. Congenital myasthenic syndrome caused by a mutation in the Ets-binding site of the promoter region of the acetylcholine receptor epsilon subunit gene. Neuromuscul. Disord. 1999, 9, 131–135. [Google Scholar] [CrossRef] [PubMed]
  105. Ohno, K.; Milone, M.; Shen, X.M.; Engel, A.G. A frameshifting mutation in CHRNE unmasks skipping of the preceding exon. Hum. Mol. Genet. 2003, 12, 3055–3066. [Google Scholar] [CrossRef]
  106. Milone, M.; Shen, X.M.; Ohno, K.; Harper, M.C.; Fukudome, T.; Stilling, G.; Brengman, J.M.; Engel, A.G. Unusual congenital myasthenic syndrome (CMS) with endplate (EP) AChR deficiency caused by alpha subunit mutations and a remitting relapsing clinical course. Neurology 1999, 52, A185–A186. [Google Scholar]
  107. Shen, X.M.; Brengman, J.M.; Sine, S.M.; Engel, A.G. Myasthenic syndrome AChRalpha C-loop mutant disrupts initiation of channel gating. J. Clin. Investig. 2012, 122, 2613–2621. [Google Scholar] [CrossRef] [PubMed]
  108. Azuma, Y.; Nakata, T.; Tanaka, M.; Shen, X.M.; Ito, M.; Iwata, S.; Okuno, T.; Nomura, Y.; Ando, N.; Ishigaki, K.; et al. Congenital myasthenic syndrome in Japan: Ethnically unique mutations in muscle nicotinic acetylcholine receptor subunits. Neuromuscul. Disord. 2015, 25, 60–69. [Google Scholar] [CrossRef]
  109. Masuda, A.; Shen, X.M.; Ito, M.; Matsuura, T.; Engel, A.G.; Ohno, K. hnRNP H enhances skipping of a nonfunctional exon P3A in CHRNA1 and a mutation disrupting its binding causes congenital myasthenic syndrome. Hum. Mol. Genet. 2008, 17, 4022–4035. [Google Scholar] [CrossRef] [PubMed]
  110. Rahman, M.A.; Masuda, A.; Ohe, K.; Ito, M.; Hutchinson, D.O.; Mayeda, A.; Engel, A.G.; Ohno, K. HnRNP L and hnRNP LL antagonistically modulate PTB-mediated splicing suppression of CHRNA1 pre-mRNA. Sci. Rep. 2013, 3, 2931. [Google Scholar] [CrossRef]
  111. Ohno, K.; Rahman, M.A.; Nazim, M.; Nasrin, F.; Lin, Y.; Takeda, J.I.; Masuda, A. Splicing regulation and dysregulation of cholinergic genes expressed at the neuromuscular junction. J. Neurochem. 2017, 142 (Suppl. 2), 64–72. [Google Scholar] [CrossRef] [Green Version]
  112. Ohno, K.; Engel, A.G.; Shen, X.M.; Selcen, D.; Brengman, J.; Harper, C.M.; Tsujino, A.; Milone, M. Rapsyn mutations in humans cause endplate acetylcholine-receptor deficiency and myasthenic syndrome. Am. J. Hum. Genet. 2002, 70, 875–885. [Google Scholar] [CrossRef]
  113. Cossins, J.; Burke, G.; Maxwell, S.; Spearman, H.; Man, S.; Kuks, J.; Vincent, A.; Palace, J.; Fuhrer, C.; Beeson, D. Diverse molecular mechanisms involved in AChR deficiency due to rapsyn mutations. Brain 2006, 129 Pt 10, 2773–2783. [Google Scholar] [CrossRef] [PubMed]
  114. Xing, G.; Jing, H.; Zhang, L.; Cao, Y.; Li, L.; Zhao, K.; Dong, Z.; Chen, W.; Wang, H.; Cao, R.; et al. A mechanism in agrin signaling revealed by a prevalent Rapsyn mutation in congenital myasthenic syndrome. eLife 2019, 8, e49180. [Google Scholar] [CrossRef] [PubMed]
  115. Lam, C.W.; Wong, K.S.; Leung, H.W.; Law, C.Y. Limb girdle myasthenia with digenic RAPSN and a novel disease gene AK9 mutations. Eur. J. Hum. Genet. 2017, 25, 192–199. [Google Scholar] [CrossRef] [PubMed]
  116. Milone, M.; Shen, X.M.; Selcen, D.; Ohno, K.; Brengman, J.; Iannaccone, S.T.; Harper, C.M.; Engel, A.G. Myasthenic syndrome due to defects in rapsyn: Clinical and molecular findings in 39 patients. Neurology 2009, 73, 228–235. [Google Scholar] [CrossRef] [PubMed]
  117. Engel, A.G.; Ohno, K.; Bouzat, C.; Sine, S.M.; Griggs, R.C. End-plate acetylcholine receptor deficiency due to nonsense mutations in the epsilon subunit. Ann. Neurol. 1996, 40, 810–817. [Google Scholar] [CrossRef] [PubMed]
  118. Burke, G.; Cossins, J.; Maxwell, S.; Owens, G.; Vincent, A.; Robb, S.; Nicolle, M.; Hilton-Jones, D.; Newsom-Davis, J.; Palace, J.; et al. Rapsyn mutations in hereditary myasthenia: Distinct early- and late-onset phenotypes. Neurology 2003, 61, 826–828. [Google Scholar] [CrossRef]
  119. Dunne, V.; Maselli, R.A. Identification of pathogenic mutations in the human rapsyn gene. J. Hum. Genet. 2003, 48, 204–207. [Google Scholar] [CrossRef]
  120. Maselli, R.A.; Dunne, V.; Pascual-Pascual, S.I.; Bowe, C.; Agius, M.; Frank, R.; Wollmann, R.L. Rapsyn mutations in myasthenic syndrome due to impaired receptor clustering. Muscle Nerve 2003, 28, 293–301. [Google Scholar] [CrossRef]
  121. Muller, J.S.; Mildner, G.; Muller-Felber, W.; Schara, U.; Krampfl, K.; Petersen, B.; Petrova, S.; Stucka, R.; Mortier, W.; Bufler, J.; et al. Rapsyn N88K is a frequent cause of congenital myasthenic syndromes in European patients. Neurology 2003, 60, 1805–1810. [Google Scholar] [CrossRef]
  122. Ohno, K.; Sadeh, M.; Blatt, I.; Brengman, J.M.; Engel, A.G. E-box mutations in the RAPSN promoter region in eight cases with congenital myasthenic syndrome. Hum. Mol. Genet. 2003, 12, 739–748. [Google Scholar] [CrossRef] [PubMed]
  123. Banwell, B.L.; Ohno, K.; Sieb, J.P.; Engel, A.G. Novel truncating RAPSN mutations causing congenital myasthenic syndrome responsive to 3,4-diaminopyridine. Neuromuscul. Disord. 2004, 14, 202–207. [Google Scholar] [CrossRef] [PubMed]
  124. Muller, J.S.; Abicht, A.; Christen, H.J.; Stucka, R.; Schara, U.; Mortier, W.; Huebner, A.; Lochmuller, H. A newly identified chromosomal microdeletion of the rapsyn gene causes a congenital myasthenic syndrome. Neuromuscul. Disord. 2004, 14, 744–749. [Google Scholar] [CrossRef]
  125. Yasaki, E.; Prioleau, C.; Barbier, J.; Richard, P.; Andreux, F.; Leroy, J.P.; Dartevelle, P.; Koenig, J.; Molgo, J.; Fardeau, M.; et al. Electrophysiological and morphological characterization of a case of autosomal recessive congenital myasthenic syndrome with acetylcholine receptor deficiency due to a N88K rapsyn homozygous mutation. Neuromuscul. Disord. 2004, 14, 24–32. [Google Scholar] [CrossRef] [PubMed]
  126. Ioos, C.; Barois, A.; Richard, P.; Eymard, B.; Hantai, D.; Estournet-Mathiaud, B. Congenital myasthenic syndrome due to rapsyn deficiency: Three cases with arthrogryposis and bulbar symptoms. Neuropediatrics 2004, 35, 246–249. [Google Scholar] [CrossRef]
  127. Muller, J.S.; Baumeister, S.K.; Rasic, V.M.; Krause, S.; Todorovic, S.; Kugler, K.; Muller-Felber, W.; Abicht, A.; Lochmuller, H. Impaired receptor clustering in congenital myasthenic syndrome with novel RAPSN mutations. Neurology 2006, 67, 1159–1164. [Google Scholar] [CrossRef] [PubMed]
  128. Maselli, R.; Dris, H.; Schnier, J.; Cockrell, J.; Wollmann, R. Congenital myasthenic syndrome caused by two non-N88K rapsyn mutations. Clin. Genet. 2007, 72, 63–65. [Google Scholar] [CrossRef]
  129. Gaudon, K.; Penisson-Besnier, I.; Chabrol, B.; Bouhour, F.; Demay, L.; Ben Ammar, A.; Bauche, S.; Vial, C.; Nicolas, G.; Eymard, B.; et al. Multiexon deletions account for 15% of congenital myasthenic syndromes with RAPSN mutations after negative DNA sequencing. J. Med. Genet. 2010, 47, 795–796. [Google Scholar] [CrossRef]
  130. Brugnoni, R.; Maggi, L.; Canioni, E.; Moroni, I.; Pantaleoni, C.; D’Arrigo, S.; Riva, D.; Cornelio, F.; Bernasconi, P.; Mantegazza, R. Identification of previously unreported mutations in CHRNA1, CHRNE and RAPSN genes in three unrelated Italian patients with congenital myasthenic syndromes. J. Neurol. 2010, 257, 1119–1123. [Google Scholar] [CrossRef]
  131. Alseth, E.H.; Maniaol, A.H.; Elsais, A.; Nakkestad, H.L.; Tallaksen, C.; Gilhus, N.E.; Skeie, G.O. Investigation for RAPSN and DOK-7 mutations in a cohort of seronegative myasthenia gravis patients. Muscle Nerve 2011, 43, 574–577. [Google Scholar] [CrossRef]
  132. Leshinsky-Silver, E.; Shapira, D.; Yosovitz, K.; Ginsberg, M.; Lerman-Sagie, T.; Lev, D. A novel mutation in the TPR6 domain of the RAPSN gene associated with congenital myasthenic syndrome. J. Neurol. Sci. 2012, 316, 112–115. [Google Scholar] [CrossRef]
  133. Lee, H.; Deignan, J.L.; Dorrani, N.; Strom, S.P.; Kantarci, S.; Quintero-Rivera, F.; Das, K.; Toy, T.; Harry, B.; Yourshaw, M.; et al. Clinical exome sequencing for genetic identification of rare Mendelian disorders. JAMA 2014, 312, 1880–1887. [Google Scholar] [CrossRef] [PubMed]
  134. Natera-de Benito, D.; Bestue, M.; Vilchez, J.J.; Evangelista, T.; Topf, A.; Garcia-Ribes, A.; Trujillo-Tiebas, M.J.; Garcia-Hoyos, M.; Ortez, C.; Camacho, A.; et al. Long-term follow-up in patients with congenital myasthenic syndrome due to RAPSN mutations. Neuromuscul. Disord. 2016, 26, 153–159. [Google Scholar] [CrossRef] [PubMed]
  135. Aharoni, S.; Sadeh, M.; Shapira, Y.; Edvardson, S.; Daana, M.; Dor-Wollman, T.; Mimouni-Bloch, A.; Halevy, A.; Cohen, R.; Sagie, L.; et al. Congenital myasthenic syndrome in Israel: Genetic and clinical characterization. Neuromuscul. Disord. 2017, 27, 136–140. [Google Scholar] [CrossRef]
  136. Estephan, E.P.; Zambon, A.A.; Marchiori, P.E.; da Silva, A.M.S.; Caldas, V.M.; Moreno, C.A.M.; Reed, U.C.; Horvath, R.; Topf, A.; Lochmuller, H.; et al. Clinical variability of early-onset congenital myasthenic syndrome due to biallelic RAPSN mutations in Brazil. Neuromuscul. Disord. 2018, 28, 961–964. [Google Scholar] [CrossRef] [PubMed]
  137. Espinoza, I.O.; Reynoso, C.; Chavez, G.; Engel, A.G. Congenital myasthenic syndrome due to rapsyn deficiency: A case report with a new mutation and compound heterozygosity. Medwave 2019, 19, e7645. [Google Scholar] [CrossRef]
  138. Liu, P.; Meng, L.; Normand, E.A.; Xia, F.; Song, X.; Ghazi, A.; Rosenfeld, J.; Magoulas, P.L.; Braxton, A.; Ward, P.; et al. Reanalysis of Clinical Exome Sequencing Data. N. Engl. J. Med. 2019, 380, 2478–2480. [Google Scholar] [CrossRef] [PubMed]
  139. Westra, D.; Schouten, M.I.; Stunnenberg, B.C.; Kusters, B.; Saris, C.G.J.; Erasmus, C.E.; van Engelen, B.G.; Bulk, S.; Verschuuren-Bemelmans, C.C.; Gerkes, E.H.; et al. Panel-Based Exome Sequencing for Neuromuscular Disorders as a Diagnostic Service. J. Neuromuscul. Dis. 2019, 6, 241–258. [Google Scholar] [CrossRef] [PubMed]
  140. Estephan, E.P.; Zambon, A.A.; Thompson, R.; Polavarapu, K.; Jomaa, D.; Topf, A.; Helito, P.V.P.; Heise, C.O.; Moreno, C.A.M.; Silva, A.M.S.; et al. Congenital myasthenic syndrome: Correlation between clinical features and molecular diagnosis. Eur. J. Neurol. 2022, 29, 833–842. [Google Scholar] [CrossRef]
  141. Krenn, M.; Sener, M.; Rath, J.; Zulehner, G.; Keritam, O.; Wagner, M.; Laccone, F.; Iglseder, S.; Marte, S.; Baumgartner, M.; et al. The clinical and molecular landscape of congenital myasthenic syndromes in Austria: A nationwide study. J. Neurol. 2022, 270, 909–916. [Google Scholar] [CrossRef]
  142. Ozturk, S.; Gulec, A.; Erdogan, M.; Demir, M.; Canpolat, M.; Gumus, H.; Caglayan, A.O.; Dundar, M.; Per, H. Congenital Myasthenic Syndromes in Turkey: Clinical and Molecular Characterization of 16 Cases With Three Novel Mutations. Pediatr. Neurol. 2022, 136, 43–49. [Google Scholar] [CrossRef]
  143. Saito, M.; Ogasawara, M.; Inaba, Y.; Osawa, Y.; Nishioka, M.; Yamauchi, S.; Atsumi, K.; Takeuchi, S.; Imai, K.; Motobayashi, M.; et al. Successful treatment of congenital myasthenic syndrome caused by a novel compound heterozygous variant in RAPSN. Brain Dev. 2022, 44, 50–55. [Google Scholar] [CrossRef]
  144. Sadeh, M.; Shen, X.M.; Engel, A.G. Beneficial effect of albuterol in congenital myasthenic syndrome with epsilon-subunit mutations. Muscle Nerve 2011, 44, 289–291. [Google Scholar] [CrossRef] [PubMed]
  145. Ishigaki, K.; Murakami, T.; Ito, Y.; Yanagisawa, A.; Kodaira, K.; Shishikura, K.; Suzuki, H.; Hirayama, Y.; Osawa, M. Treatment approach to congenital myasthenic syndrome in a patient with acetylcholine receptor deficiency. No To Hattatsu 2009, 41, 37–42. [Google Scholar]
  146. Erger, F.; Burau, K.; Elsasser, M.; Zimmermann, K.; Moog, U.; Netzer, C. Uniparental isodisomy as a cause of recessive Mendelian disease: A diagnostic pitfall with a quick and easy solution in medium/large NGS analyses. Eur. J. Hum. Genet. 2018, 26, 1392–1395. [Google Scholar] [CrossRef] [PubMed]
  147. Vogt, J.; Morgan, N.V.; Marton, T.; Maxwell, S.; Harrison, B.J.; Beeson, D.; Maher, E.R. Germline mutation in DOK7 associated with fetal akinesia deformation sequence. J. Med. Genet. 2009, 46, 338–340. [Google Scholar] [CrossRef]
  148. Michalk, A.; Stricker, S.; Becker, J.; Rupps, R.; Pantzar, T.; Miertus, J.; Botta, G.; Naretto, V.G.; Janetzki, C.; Yaqoob, N.; et al. Acetylcholine receptor pathway mutations explain various fetal akinesia deformation sequence disorders. Am. J. Hum. Genet. 2008, 82, 464–476. [Google Scholar] [CrossRef]
  149. Radhakrishnan, P.; Shukla, A.; Girisha, K.M.; Nayak, S.S. Biallelic c.1263dupC in DOK7 results in fetal akinesia deformation sequence. Am. J. Med. Genet. A 2020, 182, 804–807. [Google Scholar] [CrossRef] [PubMed]
  150. Hakonen, A.H.; Polvi, A.; Saloranta, C.; Paetau, A.; Heikkila, P.; Almusa, H.; Ellonen, P.; Jakkula, E.; Saarela, J.; Aittomaki, K. SLC18A3 variants lead to fetal akinesia deformation sequence early in pregnancy. Am. J. Med. Genet. A 2019, 179, 1362–1365. [Google Scholar] [CrossRef]
  151. Vogt, J.; Morgan, N.V.; Rehal, P.; Faivre, L.; Brueton, L.A.; Becker, K.; Fryns, J.P.; Holder, S.; Islam, L.; Kivuva, E.; et al. CHRNG genotype-phenotype correlations in the multiple pterygium syndromes. J. Med. Genet. 2012, 49, 21–26. [Google Scholar] [CrossRef]
  152. Al Kaissi, A.; Kenis, V.; Laptiev, S.; Ghachem, M.B.; Klaushofer, K.; Ganger, R.; Grill, F. Is webbing (pterygia) a constant feature in patients with Escobar syndrome? Orthop. Surg. 2013, 5, 297–301. [Google Scholar] [CrossRef]
  153. Laquerriere, A.; Maluenda, J.; Camus, A.; Fontenas, L.; Dieterich, K.; Nolent, F.; Zhou, J.; Monnier, N.; Latour, P.; Gentil, D.; et al. Mutations in CNTNAP1 and ADCY6 are responsible for severe arthrogryposis multiplex congenita with axoglial defects. Hum. Mol. Genet. 2014, 23, 2279–2289. [Google Scholar] [CrossRef] [PubMed]
  154. Sung, K.H.; Lee, S.H.; Kim, N.; Cho, T.J. Orthopaedic manifestations and treatment outcome of two siblings with Escobar syndrome and homozygous mutations in the CHRNG gene. J. Pediatr. Orthop. B 2015, 24, 262–267. [Google Scholar] [CrossRef]
  155. Retterer, K.; Juusola, J.; Cho, M.T.; Vitazka, P.; Millan, F.; Gibellini, F.; Vertino-Bell, A.; Smaoui, N.; Neidich, J.; Monaghan, K.G.; et al. Clinical application of whole-exome sequencing across clinical indications. Genet. Med. 2016, 18, 696–704. [Google Scholar] [CrossRef] [PubMed]
  156. Abouelhoda, M.; Sobahy, T.; El-Kalioby, M.; Patel, N.; Shamseldin, H.; Monies, D.; Al-Tassan, N.; Ramzan, K.; Imtiaz, F.; Shaheen, R.; et al. Clinical genomics can facilitate countrywide estimation of autosomal recessive disease burden. Genet. Med. 2016, 18, 1244–1249. [Google Scholar] [CrossRef] [PubMed]
  157. Kariminejad, A.; Almadani, N.; Khoshaeen, A.; Olsson, B.; Moslemi, A.R.; Tajsharghi, H. Truncating CHRNG mutations associated with interfamilial variability of the severity of the Escobar variant of multiple pterygium syndrome. BMC Genet. 2016, 17, 71. [Google Scholar] [CrossRef] [PubMed]
  158. Monies, D.; Abouelhoda, M.; Assoum, M.; Moghrabi, N.; Rafiullah, R.; Almontashiri, N.; Alowain, M.; Alzaidan, H.; Alsayed, M.; Subhani, S.; et al. Lessons Learned from Large-Scale, First-Tier Clinical Exome Sequencing in a Highly Consanguineous Population. Am. J. Hum. Genet. 2019, 104, 1182–1201. [Google Scholar] [CrossRef]
  159. Pingel, J.; Andersen, J.D.; Christiansen, S.L.; Borsting, C.; Morling, N.; Lorentzen, J.; Kirk, H.; Doessing, S.; Wong, C.; Nielsen, J.B. Sequence variants in muscle tissue-related genes may determine the severity of muscle contractures in cerebral palsy. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2019, 180, 12–24. [Google Scholar] [CrossRef] [PubMed]
  160. Pergande, M.; Motameny, S.; Ozdemir, O.; Kreutzer, M.; Wang, H.; Daimaguler, H.S.; Becker, K.; Karakaya, M.; Ehrhardt, H.; Elcioglu, N.; et al. The genomic and clinical landscape of fetal akinesia. Genet. Med. 2020, 22, 511–523. [Google Scholar] [CrossRef]
  161. Shamseldin, H.E.; Shaheen, R.; Ewida, N.; Bubshait, D.K.; Alkuraya, H.; Almardawi, E.; Howaidi, A.; Sabr, Y.; Abdalla, E.M.; Alfaifi, A.Y.; et al. The morbid genome of ciliopathies: An update. Genet. Med. 2020, 22, 1051–1060. [Google Scholar] [CrossRef]
  162. Croxen, R.; Hatton, C.; Shelley, C.; Brydson, M.; Chauplannaz, G.; Oosterhuis, H.; Vincent, A.; Newsom-Davis, J.; Colquhoun, D.; Beeson, D. Recessive inheritance and variable penetrance of slow-channel congenital myasthenic syndromes. Neurology 2002, 59, 162–168. [Google Scholar] [CrossRef] [PubMed]
  163. Croxen, R.; Hatton, C.; Shelley, C.; Brydson, M.; Chauplannaz, G.; Oosterhuis, H.; Vincent, A.; Newsom-Davis, J.; Colquhoun, D.; Beeson, D. Voluntary partial retraction of: Recessive inheritance and variable penetrance of slow-channel congenital myasthenic syndromes. Neurology 2009, 72, 294. [Google Scholar] [PubMed]
  164. Ohno, K.; Hutchinson, D.O.; Milone, M.; Brengman, J.M.; Bouzat, C.; Sine, S.M.; Engel, A.G. Congenital myasthenic syndrome caused by prolonged acetylcholine receptor channel openings due to a mutation in the M2 domain of the epsilon subunit. Proc. Natl. Acad. Sci. USA 1995, 92, 758–762. [Google Scholar] [CrossRef]
  165. Shen, X.M.; Okuno, T.; Milone, M.; Otsuka, K.; Takahashi, K.; Komaki, H.; Giles, E.; Ohno, K.; Engel, A.G. Mutations Causing Slow-Channel Myasthenia Reveal That a Valine Ring in the Channel Pore of Muscle AChR is Optimized for Stabilizing Channel Gating. Hum. Mutat. 2016, 37, 1051–1059. [Google Scholar] [CrossRef]
  166. Groshong, J.S.; Spencer, M.J.; Bhattacharyya, B.J.; Kudryashova, E.; Vohra, B.P.; Zayas, R.; Wollmann, R.L.; Miller, R.J.; Gomez, C.M. Calpain activation impairs neuromuscular transmission in a mouse model of the slow-channel myasthenic syndrome. J. Clin. Investig. 2007, 117, 2903–2912. [Google Scholar] [CrossRef]
  167. Di Castro, A.; Martinello, K.; Grassi, F.; Eusebi, F.; Engel, A.G. Pathogenic point mutations in a transmembrane domain of the epsilon subunit increase the Ca2+ permeability of the human endplate ACh receptor. J. Physiol. 2007, 579 Pt 3, 671–677. [Google Scholar] [CrossRef] [PubMed]
  168. Milone, M.; Wang, H.L.; Ohno, K.; Fukudome, T.; Pruitt, J.N.; Bren, N.; Sine, S.M.; Engel, A.G. Slow-channel myasthenic syndrome caused by enhanced activation, desensitization, and agonist binding affinity attributable to mutation in the M2 domain of the acetylcholine receptor alpha subunit. J. Neurosci. 1997, 17, 5651–5665. [Google Scholar] [CrossRef] [PubMed]
  169. Rahman, M.M.; Basta, T.; Teng, J.; Lee, M.; Worrell, B.T.; Stowell, M.H.B.; Hibbs, R.E. Structural mechanism of muscle nicotinic receptor desensitization and block by curare. Nat. Struct. Mol. Biol. 2022, 29, 386–394. [Google Scholar] [CrossRef]
  170. Ohno, K.; Wang, H.L.; Milone, M.; Bren, N.; Brengman, J.M.; Nakano, S.; Quiram, P.; Pruitt, J.N.; Sine, S.M.; Engel, A.G. Congenital myasthenic syndrome caused by decreased agonist binding affinity due to a mutation in the acetylcholine receptor epsilon subunit. Neuron 1996, 17, 157–170. [Google Scholar] [CrossRef]
  171. Shen, X.M.; Ohno, K.; Tsujino, A.; Brengman, J.M.; Gingold, M.; Sine, S.M.; Engel, A.G. Mutation causing severe myasthenia reveals functional asymmetry of AChR signature cystine loops in agonist binding and gating. J. Clin. Investig. 2003, 111, 497–505. [Google Scholar] [CrossRef]
  172. Shen, X.M.; Fukuda, T.; Ohno, K.; Sine, S.M.; Engel, A.G. Congenital myasthenia-related AChR delta subunit mutation interferes with intersubunit communication essential for channel gating. J. Clin. Investig. 2008, 118, 1867–1876. [Google Scholar] [CrossRef] [PubMed]
  173. Shen, X.M.; Brengman, J.M.; Edvardson, S.; Sine, S.M.; Engel, A.G. Highly fatal fast-channel syndrome caused by AChR epsilon subunit mutation at the agonist binding site. Neurology 2012, 79, 449–454. [Google Scholar] [CrossRef] [PubMed]
  174. Wang, H.L.; Ohno, K.; Milone, M.; Brengman, J.M.; Evoli, A.; Batocchi, A.P.; Middleton, L.T.; Christodoulou, K.; Engel, A.G.; Sine, S.M. Fundamental gating mechanism of nicotinic receptor channel revealed by mutation causing a congenital myasthenic syndrome. J. Gen. Physiol. 2000, 116, 449–462. [Google Scholar] [CrossRef]
  175. Shen, X.M.; Ohno, K.; Sine, S.M.; Engel, A.G. Subunit-specific contribution to agonist binding and channel gating revealed by inherited mutation in muscle acetylcholine receptor M3-M4 linker. Brain 2005, 128 Pt 2, 345–355. [Google Scholar] [CrossRef]
  176. Wang, H.L.; Milone, M.; Ohno, K.; Shen, X.M.; Tsujino, A.; Batocchi, A.P.; Tonali, P.; Brengman, J.; Engel, A.G.; Sine, S.M. Acetylcholine receptor M3 domain: Stereochemical and volume contributions to channel gating. Nat. Neurosci. 1999, 2, 226–233. [Google Scholar] [CrossRef]
  177. Sine, S.M.; Ohno, K.; Bouzat, C.; Auerbach, A.; Milone, M.; Pruitt, J.N.; Engel, A.G. Mutation of the acetylcholine receptor alpha subunit causes a slow-channel myasthenic syndrome by enhancing agonist binding affinity. Neuron 1995, 15, 229–239. [Google Scholar] [CrossRef] [PubMed]
  178. Gomez, C.M.; Gammack, J.T. A leucine-to-phenylalanine substitution in the acetylcholine receptor ion channel in a family with the slow-channel syndrome. Neurology 1995, 45, 982–985. [Google Scholar] [CrossRef]
  179. Engel, A.G.; Ohno, K.; Milone, M.; Wang, H.L.; Nakano, S.; Bouzat, C.; Pruitt, J.N., 2nd; Hutchinson, D.O.; Brengman, J.M.; Bren, N.; et al. New mutations in acetylcholine receptor subunit genes reveal heterogeneity in the slow-channel congenital myasthenic syndrome. Hum. Mol. Genet. 1996, 5, 1217–1227. [Google Scholar] [CrossRef]
  180. Gomez, C.M.; Maselli, R.; Gammack, J.; Lasalde, J.; Tamamizu, S.; Cornblath, D.R.; Lehar, M.; McNamee, M.; Kuncl, R.W. A beta-subunit mutation in the acetylcholine receptor channel gate causes severe slow-channel syndrome. Ann. Neurol. 1996, 39, 712–723. [Google Scholar] [CrossRef]
  181. Milone, M.; Ohno, K.; Wang, H.L.; Fukudome, T.; Pruitt, J.N.; Sine, S.M.; Engel, A.G. Novel slow-channel syndrome due to mutation in the acetylcholine receptor (AChR) alpha subunit with increased conductance, nanomolar affinity for acetylcholine, and prolonged open durations of the AChR channel. Ann. Neurol. 1996, 40, 9. [Google Scholar]
  182. Croxen, R.; Newland, C.; Beeson, D.; Oosterhuis, H.; Chauplannaz, G.; Vincent, A.; Newsom-Davis, J. Mutations in different functional domains of the human muscle acetylcholine receptor alpha subunit in patients with the slow-channel congenital myasthenic syndrome. Hum. Mol. Genet. 1997, 6, 767–774. [Google Scholar] [CrossRef]
  183. Gomez, C.M.; Maselli, R.; Gundeck, J.E.; Chao, M.; Day, J.W.; Tamamizu, S.; Lasalde, J.A.; McNamee, M.; Wollmann, R.L. Slow-channel transgenic mice: A model of postsynaptic organellar degeneration at the neuromuscular junction. J. Neurosci. 1997, 17, 4170–4179. [Google Scholar] [CrossRef] [Green Version]
  184. Wintzen, A.R.; Plomp, J.J.; Molenaar, P.C.; van Dijk, J.G.; van Kempen, G.T.; Vos, R.M.; Wokke, J.H.; Vincent, A. Acquired slow-channel syndrome: A form of myasthenia gravis with prolonged open time of the acetylcholine receptor channel. Ann. Neurol. 1998, 44, 657–664. [Google Scholar] [CrossRef]
  185. Scola, R.H.; Werneck, L.C.; Iwamoto, F.M.; Comerlato, E.A.; Kay, C.K. Acquired slow-channel syndrome. Muscle Nerve 2000, 23, 1582–1585. [Google Scholar] [CrossRef]
  186. Gomez, C.M.; Maselli, R.A.; Vohra, B.P.; Navedo, M.; Stiles, J.R.; Charnet, P.; Schott, K.; Rojas, L.; Keesey, J.; Verity, A.; et al. Novel delta subunit mutation in slow-channel syndrome causes severe weakness by novel mechanisms. Ann. Neurol. 2002, 51, 102–112. [Google Scholar] [CrossRef]
  187. Hatton, C.J.; Shelley, C.; Brydson, M.; Beeson, D.; Colquhoun, D. Properties of the human muscle nicotinic receptor, and of the slow-channel myasthenic syndrome mutant epsilonL221F, inferred from maximum likelihood fits. J. Physiol. 2003, 547 Pt 3, 729–760. [Google Scholar] [CrossRef] [PubMed]
  188. Colomer, J.; Muller, J.S.; Vernet, A.; Nascimento, A.; Pons, M.; Gonzalez, V.; Abicht, A.; Lochmuller, H. Long-term improvement of slow-channel congenital myasthenic syndrome with fluoxetine. Neuromuscul. Disord. 2006, 16, 329–333. [Google Scholar] [CrossRef] [PubMed]
  189. Lorenzoni, P.J.; Kay, C.S.; Arruda, W.O.; Scola, R.H.; Werneck, L.C. Neurophysiological study in slow-channel congenital myasthenic syndrome: Case report. Arq. Neuropsiquiatr. 2006, 64, 318–321. [Google Scholar] [CrossRef] [PubMed]
  190. Navedo, M.F.; Lasalde-Dominicci, J.A.; Baez-Pagan, C.A.; Diaz-Perez, L.; Rojas, L.V.; Maselli, R.A.; Staub, J.; Schott, K.; Zayas, R.; Gomez, C.M. Novel beta subunit mutation causes a slow-channel syndrome by enhancing activation and decreasing the rate of agonist dissociation. Mol. Cell Neurosci. 2006, 32, 82–90. [Google Scholar] [CrossRef] [PubMed]
  191. Shen, X.M.; Deymeer, F.; Sine, S.M.; Engel, A.G. Slow-channel mutation in acetylcholine receptor alphaM4 domain and its efficient knockdown. Ann. Neurol. 2006, 60, 128–136. [Google Scholar] [CrossRef]
  192. Outteryck, O.; Richard, P.; Lacour, A.; Fournier, E.; Zephir, H.; Gaudon, K.; Eymard, B.; Hantai, D.; Vermersch, P.; Stojkovic, T. Novel epsilon subunit mutation of the muscle acetylcholine receptor causing a slow-channel congenital myasthenic syndrome. J. Neurol. Neurosurg. Psychiatry 2009, 80, 450–451. [Google Scholar] [CrossRef] [PubMed]
  193. Chaouch, A.; Muller, J.S.; Guergueltcheva, V.; Dusl, M.; Schara, U.; Rakocevic-Stojanovic, V.; Lindberg, C.; Scola, R.H.; Werneck, L.C.; Colomer, J.; et al. A retrospective clinical study of the treatment of slow-channel congenital myasthenic syndrome. J. Neurol. 2012, 259, 474–481. [Google Scholar] [CrossRef] [PubMed]
  194. Witoonpanich, R.; Pulkes, T.; Dejthevaporn, C.; Witoonpanich, P.; Yodnopklao, P.; Wetchaphanphesat, S.; Brengman, J.; Engel, A.G. Phenotypic heterogeneity in a large Thai slow-channel congenital myasthenic syndrome kinship: Correction. Neuromuscul. Disord. 2012, 22, 478. [Google Scholar] [CrossRef]
  195. Tan, J.Z.; Man, Y.; Xiao, F. A Missense Mutation in Epsilon-subunit of Acetylcholine Receptor Causing Autosomal Dominant Slow-channel Congenital Myasthenic Syndrome in a Chinese Family. Chin. Med. J. 2016, 129, 2596–2602. [Google Scholar] [CrossRef] [PubMed]
  196. Angelini, C.; Lispi, L.; Salvoro, C.; Mostacciuolo, M.L.; Vazza, G. Clinical and genetic characterization of an Italian family with slow-channel syndrome. Neurol. Sci. 2019, 40, 503–507. [Google Scholar] [CrossRef]
  197. Shen, X.M.; Milone, M.; Wang, H.L.; Banwell, B.; Selcen, D.; Sine, S.M.; Engel, A.G. Slow-channel myasthenia due to novel mutation in M2 domain of AChR delta subunit. Ann. Clin. Transl. Neurol. 2019, 6, 2066–2078. [Google Scholar] [CrossRef]
  198. Di, L.; Chen, H.; Lu, Y.; Selcen, D.; Engel, A.G.; Da, Y.; Shen, X.M. Determinants of the repetitive-CMAP occurrence and therapy efficacy in slow-channel myasthenia. Neurology 2020, 95, e2781–e2793. [Google Scholar] [CrossRef]
  199. Gooneratne, I.K.; Nandasiri, S.; Maxwell, S.; Webster, R.; Cossins, J.; Beeson, D.; Gunaratne, K.; Herath, L.; Senanayake, S.; Chang, T. Slow-Channel Congenital Myasthenic Syndrome due to a Novel Mutation in the Acetylcholine Receptor Alpha Subunit in a South Asian: A Case Report. J. Neuromuscul. Dis. 2021, 8, 163–167. [Google Scholar] [CrossRef]
  200. Huang, K.; Luo, Y.B.; Bi, F.F.; Yang, H. Pharmacological Strategy for Congenital Myasthenic Syndrome with CHRNE Mutations: A Meta-Analysis of Case Reports. Curr. Neuropharmacol. 2021, 19, 718–729. [Google Scholar] [CrossRef]
  201. Kudryavtsev, D.; Isaeva, A.; Barkova, D.; Spirova, E.; Mukhutdinova, R.; Kasheverov, I.; Tsetlin, V. Point Mutations of Nicotinic Receptor alpha1 Subunit Reveal New Molecular Features of G153S Slow-Channel Myasthenia. Molecules 2021, 26, 1278. [Google Scholar] [CrossRef]
  202. Santovito, L.S.; Brugnoni, R.; Banfi, P.; Maggi, L. Salbutamol as effective treatment in slow-channel syndrome- first report. Neurol. Sci. 2021, 42, 1611–1612. [Google Scholar] [CrossRef]
  203. Tawara, N.; Yamashita, S.; Takamatsu, K.; Yamasaki, Y.; Mukaino, A.; Nakane, S.; Farshadyeganeh, P.; Ohno, K.; Ando, Y. Efficacy of salbutamol monotherapy in slow-channel congenital myasthenic syndrome caused by a novel mutation in CHRND. Muscle Nerve 2021, 63, E30–E32. [Google Scholar] [CrossRef]
  204. Dejthevaporn, C.; Wetchaphanphesat, S.; Pulkes, T.; Rattanasiri, S.; Engel, A.G.; Witoonpanich, R. Treatment of slow-channel congenital myasthenic syndrome in a Thai family with fluoxetine. J. Clin. Neurosci. 2022, 96, 85–89. [Google Scholar] [CrossRef]
  205. Kinali, M.; Beeson, D.; Pitt, M.C.; Jungbluth, H.; Simonds, A.K.; Aloysius, A.; Cockerill, H.; Davis, T.; Palace, J.; Manzur, A.Y.; et al. Congenital myasthenic syndromes in childhood: Diagnostic and management challenges. J. Neuroimmunol. 2008, 201–202, 6–12. [Google Scholar] [CrossRef]
  206. Webster, R.G.; Cossins, J.; Lashley, D.; Maxwell, S.; Liu, W.W.; Wickens, J.R.; Martinez-Martinez, P.; de Baets, M.; Beeson, D. A mouse model of the slow channel myasthenic syndrome: Neuromuscular physiology and effects of ephedrine treatment. Exp. Neurol. 2013, 248, 286–298. [Google Scholar] [CrossRef]
  207. Finlayson, S.; Spillane, J.; Kullmann, D.M.; Howard, R.; Webster, R.; Palace, J.; Beeson, D. Slow channel congenital myasthenic syndrome responsive to a combination of fluoxetine and salbutamol. Muscle Nerve 2013, 47, 279–282. [Google Scholar] [CrossRef] [PubMed]
  208. Brownlow, S.; Webster, R.; Croxen, R.; Brydson, M.; Neville, B.; Lin, J.P.; Vincent, A.; Newsom-Davis, J.; Beeson, D. Acetylcholine receptor delta subunit mutations underlie a fast-channel myasthenic syndrome and arthrogryposis multiplex congenita. J. Clin. Investig. 2001, 108, 125–130. [Google Scholar] [CrossRef] [PubMed]
  209. Shen, X.M.; Ohno, K.; Fukudome, T.; Tsujino, A.; Brengman, J.M.; De Vivo, D.C.; Packer, R.J.; Engel, A.G. Congenital myasthenic syndrome caused by low-expressor fast-channel AChR delta subunit mutation. Neurology 2002, 59, 1881–1888. [Google Scholar] [CrossRef]
  210. Sine, S.M.; Shen, X.M.; Wang, H.L.; Ohno, K.; Lee, W.Y.; Tsujino, A.; Brengmann, J.; Bren, N.; Vajsar, J.; Engel, A.G. Naturally occurring mutations at the acetylcholine receptor binding site independently alter ACh binding and channel gating. J. Gen. Physiol. 2002, 120, 483–496. [Google Scholar] [CrossRef]
  211. Sine, S.M.; Wang, H.L.; Ohno, K.; Shen, X.M.; Lee, W.Y.; Engel, A.G. Mechanistic diversity underlying fast channel congenital myasthenic syndromes. Ann. N. Y. Acad. Sci. 2003, 998, 128–137. [Google Scholar] [CrossRef] [PubMed]
  212. Webster, R.; Brydson, M.; Croxen, R.; Newsom-Davis, J.; Vincent, A.; Beeson, D. Mutation in the AChR ion channel gate underlies a fast channel congenital myasthenic syndrome. Neurology 2004, 62, 1090–1096. [Google Scholar] [CrossRef]
  213. Palace, J.; Lashley, D.; Bailey, S.; Jayawant, S.; Carr, A.; McConville, J.; Robb, S.; Beeson, D. Clinical features in a series of fast channel congenital myasthenia syndrome. Neuromuscul. Disord. 2012, 22, 112–117. [Google Scholar] [CrossRef] [PubMed]
  214. Webster, R.; Liu, W.W.; Chaouch, A.; Lochmuller, H.; Beeson, D. Fast-channel congenital myasthenic syndrome with a novel acetylcholine receptor mutation at the alpha-epsilon subunit interface. Neuromuscul. Disord. 2014, 24, 143–147. [Google Scholar] [CrossRef]
  215. Shen, X.M.; Di, L.; Shen, S.; Zhao, Y.; Neumeyer, A.M.; Selcen, D.; Sine, S.M.; Engel, A.G. A novel fast-channel myasthenia caused by mutation in beta subunit of AChR reveals subunit-specific contribution of the intracellular M1-M2 linker to channel gating. Exp. Neurol. 2020, 331, 113375. [Google Scholar] [CrossRef] [PubMed]
  216. Shen, X.M.; Brengman, J.M.; Shen, S.; Durmus, H.; Preethish-Kumar, V.; Yuceyar, N.; Vengalil, S.; Nalini, A.; Deymeer, F.; Sine, S.M.; et al. Mutations causing congenital myasthenia reveal principal coupling pathway in the acetylcholine receptor epsilon-subunit. JCI Insight 2018, 3, e97826. [Google Scholar] [CrossRef] [PubMed]
  217. Deprez, P.; Inestrosa, N.C.; Krejci, E. Two different heparin-binding domains in the triple-helical domain of ColQ, the collagen tail subunit of synaptic acetylcholinesterase. J. Biol. Chem. 2003, 278, 23233–23242. [Google Scholar] [CrossRef]
  218. Peng, H.B.; Xie, H.; Rossi, S.G.; Rotundo, R.L. Acetylcholinesterase clustering at the neuromuscular junction involves perlecan and dystroglycan. J. Cell Biol. 1999, 145, 911–921. [Google Scholar] [CrossRef]
  219. Cartaud, A.; Strochlic, L.; Guerra, M.; Blanchard, B.; Lambergeon, M.; Krejci, E.; Cartaud, J.; Legay, C. MuSK is required for anchoring acetylcholinesterase at the neuromuscular junction. J. Cell Biol. 2004, 165, 505–515. [Google Scholar] [CrossRef]
  220. Kawakami, Y.; Ito, M.; Hirayama, M.; Sahashi, K.; Ohkawara, B.; Masuda, A.; Nishida, H.; Mabuchi, N.; Engel, A.G.; Ohno, K. Anti-MuSK autoantibodies block binding of collagen Q to MuSK. Neurology 2011, 77, 1819–1826. [Google Scholar] [CrossRef]
  221. Ohno, K.; Brengman, J.; Tsujino, A.; Engel, A.G. Human endplate acetylcholinesterase deficiency caused by mutations in the collagen-like tail subunit (ColQ) of the asymmetric enzyme. Proc. Natl. Acad. Sci. USA 1998, 95, 9654–9659. [Google Scholar] [CrossRef]
  222. Donger, C.; Krejci, E.; Serradell, A.P.; Eymard, B.; Bon, S.; Nicole, S.; Chateau, D.; Gary, F.; Fardeau, M.; Massoulie, J.; et al. Mutation in the human acetylcholinesterase-associated collagen gene, COLQ, is responsible for congenital myasthenic syndrome with end-plate acetylcholinesterase deficiency (Type Ic). Am. J. Hum. Genet. 1998, 63, 967–975. [Google Scholar] [CrossRef]
  223. Ohno, K.; Brengman, J.M.; Felice, K.J.; Cornblath, D.R.; Engel, A.G. Congenital end-plate acetylcholinesterase deficiency caused by a nonsense mutation and an A-->G splice-donor-site mutation at position +3 of the collagenlike-tail-subunit gene (COLQ): How does G at position +3 result in aberrant splicing? Am. J. Hum. Genet. 1999, 65, 635–644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Ohno, K.; Engel, A.G.; Brengman, J.M.; Shen, X.-M.; Heidenrich, F.R.; Vincent, A.; Milone, M.; Tan, E.; Demirci, M.; Walsh, P.; et al. The spectrum of mutations causing endplate acetylcholinesterase deficiency. Ann. Neurol. 2000, 47, 162–170. [Google Scholar] [CrossRef] [PubMed]
  225. Nakata, T.; Ito, M.; Azuma, Y.; Otsuka, K.; Noguchi, Y.; Komaki, H.; Okumura, A.; Shiraishi, K.; Masuda, A.; Natsume, J.; et al. Mutations in the C-terminal domain of ColQ in endplate acetylcholinesterase deficiency compromise ColQ-MuSK interaction. Hum. Mutat. 2013, 34, 997–1004. [Google Scholar] [CrossRef] [PubMed]
  226. Feng, G.; Krejci, E.; Molgo, J.; Cunningham, J.M.; Massoulie, J.; Sanes, J.R. Genetic analysis of collagen Q: Roles in acetylcholinesterase and butyrylcholinesterase assembly and in synaptic structure and function. J. Cell Biol. 1999, 144, 1349–1360. [Google Scholar] [CrossRef]
  227. Ito, M.; Suzuki, Y.; Okada, T.; Fukudome, T.; Yoshimura, T.; Masuda, A.; Takeda, S.; Krejci, E.; Ohno, K. Protein-anchoring strategy for delivering acetylcholinesterase to the neuromuscular junction. Mol. Ther. 2012, 20, 1384–1392. [Google Scholar] [CrossRef]
  228. Bartels, C.F.; Zelinski, T.; Lockridge, O. Mutation at Codon-322 in the Human Acetylcholinesterase (Ache) Gene Accounts for Yt Blood-Group Polymorphism. Am. J. Hum. Genet. 1993, 52, 928–936. [Google Scholar]
  229. Kimbell, L.M.; Ohno, K.; Engel, A.G.; Rotundo, R.L. C-terminal and heparin-binding domains of collagenic tail subunit are both essential for anchoring acetylcholinesterase at the synapse. J. Biol. Chem. 2004, 279, 10997–11005. [Google Scholar] [CrossRef]
  230. Rogers, R.S.; Nishimune, H. The role of laminins in the organization and function of neuromuscular junctions. Matrix Biol. 2017, 57–58, 86–105. [Google Scholar] [CrossRef]
  231. Carlson, S.S.; Valdez, G.; Sanes, J.R. Presynaptic calcium channels and alpha3-integrins are complexed with synaptic cleft laminins, cytoskeletal elements and active zone components. J. Neurochem. 2010, 115, 654–666. [Google Scholar] [CrossRef]
  232. Chen, J.; Billings, S.E.; Nishimune, H. Calcium channels link the muscle-derived synapse organizer laminin beta2 to Bassoon and CAST/Erc2 to organize presynaptic active zones. J. Neurosci. 2011, 31, 512–525. [Google Scholar] [CrossRef] [PubMed]
  233. Nishimune, H.; Stanford, J.A.; Mori, Y. Role of exercise in maintaining the integrity of the neuromuscular junction. Muscle Nerve 2014, 49, 315–324. [Google Scholar] [CrossRef]
  234. Chen, F.; Liu, Y.; Sugiura, Y.; Allen, P.D.; Gregg, R.G.; Lin, W. Neuromuscular synaptic patterning requires the function of skeletal muscle dihydropyridine receptors. Nat. Neurosci. 2011, 14, 570–577. [Google Scholar] [CrossRef] [Green Version]
  235. Zenker, M.; Aigner, T.; Wendler, O.; Tralau, T.; Muntefering, H.; Fenski, R.; Pitz, S.; Schumacher, V.; Royer-Pokora, B.; Wuhl, E.; et al. Human laminin beta2 deficiency causes congenital nephrosis with mesangial sclerosis and distinct eye abnormalities. Hum. Mol. Genet. 2004, 13, 2625–2632. [Google Scholar] [CrossRef] [PubMed]
  236. Hasselbacher, K.; Wiggins, R.C.; Matejas, V.; Hinkes, B.G.; Mucha, B.; Hoskins, B.E.; Ozaltin, F.; Nurnberg, G.; Becker, C.; Hangan, D.; et al. Recessive missense mutations in LAMB2 expand the clinical spectrum of LAMB2-associated disorders. Kidney Int. 2006, 70, 1008–1012. [Google Scholar] [CrossRef] [PubMed]
  237. Noakes, P.G.; Gautam, M.; Mudd, J.; Sanes, J.R.; Merlie, J.P. Aberrant differentiation of neuromuscular junctions in mice lacking s-laminin/laminin beta 2. Nature 1995, 374, 258–262. [Google Scholar] [CrossRef]
  238. Latvanlehto, A.; Fox, M.A.; Sormunen, R.; Tu, H.; Oikarainen, T.; Koski, A.; Naumenko, N.; Shakirzyanova, A.; Kallio, M.; Ilves, M.; et al. Muscle-derived collagen XIII regulates maturation of the skeletal neuromuscular junction. J. Neurosci. 2010, 30, 12230–12241. [Google Scholar] [CrossRef]
  239. Logan, C.V.; Cossins, J.; Rodriguez Cruz, P.M.; Parry, D.A.; Maxwell, S.; Martinez-Martinez, P.; Riepsaame, J.; Abdelhamed, Z.A.; Lake, A.V.; Moran, M.; et al. Congenital Myasthenic Syndrome Type 19 Is Caused by Mutations in COL13A1, Encoding the Atypical Non-fibrillar Collagen Type XIII alpha1 Chain. Am. J. Hum. Genet. 2015, 97, 878–885. [Google Scholar] [CrossRef]
  240. Haronen, H.; Zainul, Z.; Tu, H.; Naumenko, N.; Sormunen, R.; Miinalainen, I.; Shakirzyanova, A.; Oikarainen, T.; Abdullin, A.; Martin, P.; et al. Collagen XIII secures pre- and postsynaptic integrity of the neuromuscular synapse. Hum. Mol. Genet. 2017, 26, 2076–2090. [Google Scholar] [CrossRef]
  241. Kemppainen, A.V.; Finnila, M.A.; Heikkinen, A.; Haronen, H.; Izzi, V.; Kauppinen, S.; Saarakkala, S.; Pihlajaniemi, T.; Koivunen, J. The CMS19 disease model specifies a pivotal role for collagen XIII in bone homeostasis. Sci. Rep. 2022, 12, 5866. [Google Scholar] [CrossRef]
  242. Ohno, K.; Brengman, J.; Engel, A. How does an A-to-G splice donor site mutation at position+3 result in aberrant splicing? A lesson learned from a mutation in the COLQ gene. Am. J. Hum. Genet. 1999, 65, A80. [Google Scholar] [CrossRef] [PubMed]
  243. Shapira, Y.A.; Sadeh, M.E.; Bergtraum, M.P.; Tsujino, A.; Ohno, K.; Shen, X.M.; Brengman, J.; Edwardson, S.; Matoth, I.; Engel, A.G. Three novel COLQ mutations and variation of phenotypic expressivity due to G240X. Neurology 2002, 58, 603–609. [Google Scholar] [CrossRef] [PubMed]
  244. Ishigaki, K.; Nicolle, D.; Krejci, E.; Leroy, J.P.; Koenig, J.; Fardeau, M.; Eymard, B.; Hantai, D. Two novel mutations in the COLQ gene cause endplate acetylcholinesterase deficiency. Neuromuscul. Disord. 2003, 13, 236–244. [Google Scholar] [CrossRef]
  245. Muller, J.S.; Petrova, S.; Kiefer, R.; Stucka, R.; Konig, C.; Baumeister, S.K.; Huebner, A.; Lochmuller, H.; Abicht, A. Synaptic congenital myasthenic syndrome in three patients due to a novel missense mutation (T441A) of the COLQ gene. Neuropediatrics 2004, 35, 183–189. [Google Scholar] [CrossRef] [PubMed]
  246. Schreiner, F.; Hoppenz, M.; Klaeren, R.; Reimann, J.; Woelfle, J. Novel COLQ mutation 950delC in synaptic congenital myasthenic syndrome and symptomatic heterozygous relatives. Neuromuscul. Disord. 2007, 17, 262–265. [Google Scholar] [CrossRef] [PubMed]
  247. Mihaylova, V.; Muller, J.S.; Vilchez, J.J.; Salih, M.A.; Kabiraj, M.M.; D’Amico, A.; Bertini, E.; Wolfle, J.; Schreiner, F.; Kurlemann, G.; et al. Clinical and molecular genetic findings in COLQ-mutant congenital myasthenic syndromes. Brain 2008, 131 Pt 3, 747–759. [Google Scholar] [CrossRef] [PubMed]
  248. Yeung, W.L.; Lam, C.W.; Ng, P.C. Intra-familial variation in clinical manifestations and response to ephedrine in siblings with congenital myasthenic syndrome caused by novel COLQ mutations. Dev. Med. Child Neurol. 2010, 52, e243–e244. [Google Scholar] [CrossRef]
  249. Duran, G.S.; Uzunhan, T.A.; Ekici, B.; Citak, A.; Aydinli, N.; Caliskan, M. Severe scoliosis in a patient with COLQ mutation and congenital myasthenic syndrome: A clue for diagnosis. Acta Neurol. Belg. 2013, 113, 531–532. [Google Scholar] [CrossRef]
  250. Arredondo, J.; Lara, M.; Ng, F.; Gochez, D.A.; Lee, D.C.; Logia, S.P.; Nguyen, J.; Maselli, R.A. COOH-terminal collagen Q (COLQ) mutants causing human deficiency of endplate acetylcholinesterase impair the interaction of ColQ with proteins of the basal lamina. Hum. Genet. 2014, 133, 599–616. [Google Scholar] [CrossRef]
  251. Matlik, H.N.; Milhem, R.M.; Saadeldin, I.Y.; Al-Jaibeji, H.S.; Al-Gazali, L.; Ali, B.R. Clinical and molecular analysis of a novel COLQ missense mutation causing congenital myasthenic syndrome in a Syrian family. Pediatr. Neurol. 2014, 51, 165–169. [Google Scholar] [CrossRef]
  252. Wang, W.; Wu, Y.; Wang, C.; Jiao, J.; Klein, C.J. Copy number analysis reveals a novel multiexon deletion of the COLQ gene in congenital myasthenia. Neurol. Genet. 2016, 2, e117. [Google Scholar] [CrossRef]
  253. Al-Muhaizea, M.A.; Al-Mobarak, S.B. COLQ-mutant Congenital Myasthenic Syndrome with Microcephaly: A Unique Case with Literature Review. Transl. Neurosci. 2017, 8, 65–69. [Google Scholar]
  254. Padmanabha, H.; Saini, A.G.; Sankhyan, N.; Singhi, P. COLQ-Related Congenital Myasthenic Syndrome and Response to Salbutamol Therapy. J. Clin. Neuromuscul. Dis. 2017, 18, 162–163. [Google Scholar] [CrossRef]
  255. Zhang, Q.L.; Xu, M.J.; Wang, T.L.; Zhu, Z.Q.; Lai, F.; Zheng, X.C. Newly discovered COLQ gene mutation and its clinical features in patients with acetyl cholinesterase deficiency. J. Integr. Neurosci. 2018, 17, 439–446. [Google Scholar] [CrossRef] [PubMed]
  256. Laforgia, N.; De Cosmo, L.; Palumbo, O.; Ranieri, C.; Sesta, M.; Capodiferro, D.; Pantaleo, A.; Iapicca, P.; Lastella, P.; Capozza, M.; et al. The First Case of Congenital Myasthenic Syndrome Caused by a Large Homozygous Deletion in the C-Terminal Region of COLQ (Collagen Like Tail Subunit of Asymmetric Acetylcholinesterase) Protein. Genes 2020, 11, 1519. [Google Scholar] [CrossRef] [PubMed]
  257. Ren, H.Q.; Zhang, J.W.; Wang, L.Y.; Xue, P.; An, H.B. Congenital myasthenic syndrome with COLQ gene mutation: Report of a case. Zhonghua Bing Li Xue Za Zhi 2020, 49, 269–271. [Google Scholar]
  258. Tay, C.G.; Fong, C.Y.; Li, L.; Ganesan, V.; Teh, C.M.; Gan, C.S.; Thong, M.K. Congenital myasthenic syndrome with novel pathogenic variants in the COLQ gene associated with the presence of antibodies to acetylcholine receptors. J. Clin. Neurosci. 2020, 72, 468–471. [Google Scholar] [CrossRef] [PubMed]
  259. Luo, X.; Wang, C.; Lin, L.; Yuan, F.; Wang, S.; Wang, Y.; Wang, A.; Wang, C.; Wu, S.; Lan, X.; et al. Mechanisms of Congenital Myasthenia Caused by Three Mutations in the COLQ Gene. Front. Pediatr. 2021, 9, 679342. [Google Scholar] [CrossRef]
  260. Pallithanam, J.J.; Prabhudesai, S.P.; Naik, N.; Gauns, S. COLQ-Related Congenital Myasthenic Syndrome in a Child from Western India. Neurol. India 2021, 69, 228–229. [Google Scholar]
  261. Al-Sharif, F.; Alamer, M.F.; Taher, H.O.; Gazzaz, R.Y.; AlRuwaithi, A.O.; Miliany, T.T.; Alrufaihi, M.A.; Al Amer, A.F. Co-occurrence of Glycogen Storage Disease Type 2 and Congenital Myasthenic Syndrome Type 5 in a Pediatric Patient: A Case Report. Cureus 2022, 14, e26345. [Google Scholar] [CrossRef]
  262. El Kadiri, Y.; Ratbi, I.; Sefiani, A.; Lyahyai, J. Novel copy number variation of COLQ gene in a Moroccan patient with congenital myasthenic syndrome: A case report and review of the literature. BMC Neurol. 2022, 22, 292. [Google Scholar] [CrossRef]
  263. Yamashita, A.; Muramatsu, Y.; Matsuda, H.; Okamoto, H. General anesthesia for treating scoliosis with congenital myasthenia syndrome: A case report. JA Clin. Rep. 2022, 8, 70. [Google Scholar] [CrossRef] [PubMed]
  264. Bestue-Cardiel, M.; Saenz de Cabezon-Alvarez, A.; Capablo-Liesa, J.L.; Lopez-Pison, J.; Pena-Segura, J.L.; Martin-Martinez, J.; Engel, A.G. Congenital endplate acetylcholinesterase deficiency responsive to ephedrine. Neurology 2005, 65, 144–146. [Google Scholar] [CrossRef]
  265. Liewluck, T.; Selcen, D.; Engel, A.G. Beneficial effects of albuterol in congenital endplate acetylcholinesterase deficiency and Dok-7 myasthenia. Muscle Nerve 2011, 44, 789–794. [Google Scholar] [CrossRef]
  266. Chan, S.H.; Wong, V.C.; Engel, A.G. Neuromuscular junction acetylcholinesterase deficiency responsive to albuterol. Pediatr. Neurol. 2012, 47, 137–140. [Google Scholar] [CrossRef]
  267. Dusl, M.; Moreno, T.; Munell, F.; Macaya, A.; Gratacos, M.; Abicht, A.; Strom, T.M.; Lochmuller, H.; Senderek, J. Congenital myasthenic syndrome caused by novel COL13A1 mutations. J. Neurol. 2019, 266, 1107–1112. [Google Scholar] [CrossRef] [PubMed]
  268. Rodriguez Cruz, P.M.; Cossins, J.; Estephan, E.P.; Munell, F.; Selby, K.; Hirano, M.; Maroofin, R.; Mehrjardi, M.Y.V.; Chow, G.; Carr, A.; et al. The clinical spectrum of the congenital myasthenic syndrome resulting from COL13A1 mutations. Brain 2019, 142, 1547–1560. [Google Scholar] [CrossRef]
  269. Kediha, M.I.; Tazir, M.; Sternberg, D.; Eymard, B.; Alipacha, L. Moderate phenotype of a congenital myasthenic syndrome type 19 caused by mutation of the COL13A1 gene: A case report. J. Med. Case Rep. 2022, 16, 134. [Google Scholar] [CrossRef]
  270. Arnold, W.D.; Feldman, D.H.; Ramirez, S.; He, L.; Kassar, D.; Quick, A.; Klassen, T.L.; Lara, M.; Nguyen, J.; Kissel, J.T.; et al. Defective fast inactivation recovery of Nav 1.4 in congenital myasthenic syndrome. Ann. Neurol. 2015, 77, 840–850. [Google Scholar] [CrossRef] [PubMed]
  271. Statland, J.M.; Fontaine, B.; Hanna, M.G.; Johnson, N.E.; Kissel, J.T.; Sansone, V.A.; Shieh, P.B.; Tawil, R.N.; Trivedi, J.; Cannon, S.C.; et al. Review of the Diagnosis and Treatment of Periodic Paralysis. Muscle Nerve 2018, 57, 522–530. [Google Scholar] [CrossRef]
  272. Lerche, H.; Heine, R.; Pika, U.; George, A.L., Jr.; Mitrovic, N.; Browatzki, M.; Weiss, T.; Rivet-Bastide, M.; Franke, C.; Lomonaco, M.; et al. Human sodium channel myotonia: Slowed channel inactivation due to substitutions for a glycine within the III-IV linker. J. Physiol. 1993, 470, 13–22. [Google Scholar] [CrossRef] [PubMed]
  273. Ptacek, L.J.; George, A.L., Jr.; Barchi, R.L.; Griggs, R.C.; Riggs, J.E.; Robertson, M.; Leppert, M.F. Mutations in an S4 segment of the adult skeletal muscle sodium channel cause paramyotonia congenita. Neuron 1992, 8, 891–897. [Google Scholar] [CrossRef] [PubMed]
  274. Huang, K.; Duan, H.Q.; Li, Q.X.; Luo, Y.B.; Bi, F.F.; Yang, H. Clinicopathological-genetic features of congenital myasthenic syndrome from a Chinese neuromuscular centre. J. Cell Mol. Med. 2022, 26, 3828–3836. [Google Scholar] [CrossRef]
  275. Ohkawara, B.; Shen, X.; Selcen, D.; Nazim, M.; Bril, V.; Tarnopolsky, M.A.; Brady, L.; Fukami, S.; Amato, A.A.; Yis, U.; et al. Congenital myasthenic syndrome-associated agrin variants affect clustering of acetylcholine receptors in a domain-specific manner. JCI Insight 2020, 5, e132023. [Google Scholar] [CrossRef] [PubMed]
  276. Leupin, O.; Piters, E.; Halleux, C.; Hu, S.; Kramer, I.; Morvan, F.; Bouwmeester, T.; Schirle, M.; Bueno-Lozano, M.; Fuentes, F.J.; et al. Bone overgrowth-associated mutations in the LRP4 gene impair sclerostin facilitator function. J. Biol. Chem. 2011, 286, 19489–19500. [Google Scholar] [CrossRef]
  277. Li, Y.; Pawlik, B.; Elcioglu, N.; Aglan, M.; Kayserili, H.; Yigit, G.; Percin, F.; Goodman, F.; Nurnberg, G.; Cenani, A.; et al. LRP4 mutations alter Wnt/beta-catenin signaling and cause limb and kidney malformations in Cenani-Lenz syndrome. Am. J. Hum. Genet. 2010, 86, 696–706. [Google Scholar] [CrossRef] [Green Version]
  278. Chevessier, F.; Faraut, B.; Ravel-Chapuis, A.; Richard, P.; Gaudon, K.; Bauche, S.; Prioleau, C.; Herbst, R.; Goillot, E.; Ioos, C.; et al. MUSK, a new target for mutations causing congenital myasthenic syndrome. Hum. Mol. Genet. 2004, 13, 3229–3240. [Google Scholar] [CrossRef]
  279. Maselli, R.A.; Arredondo, J.; Cagney, O.; Ng, J.J.; Anderson, J.A.; Williams, C.; Gerke, B.J.; Soliven, B.; Wollmann, R.L. Mutations in MUSK causing congenital myasthenic syndrome impair MuSK-Dok-7 interaction. Hum. Mol. Genet. 2010, 19, 2370–2379. [Google Scholar] [CrossRef]
  280. Beeson, D.; Higuchi, O.; Palace, J.; Cossins, J.; Spearman, H.; Maxwell, S.; Newsom-Davis, J.; Burke, G.; Fawcett, P.; Motomura, M.; et al. Dok-7 mutations underlie a neuromuscular junction synaptopathy. Science 2006, 313, 1975–1978. [Google Scholar] [CrossRef]
  281. Muller, J.S.; Herczegfalvi, A.; Vilchez, J.J.; Colomer, J.; Bachinski, L.L.; Mihaylova, V.; Santos, M.; Schara, U.; Deschauer, M.; Shevell, M.; et al. Phenotypical spectrum of DOK7 mutations in congenital myasthenic syndromes. Brain 2007, 130 Pt 6, 1497–1506. [Google Scholar] [CrossRef]
  282. Hamuro, J.; Higuchi, O.; Okada, K.; Ueno, M.; Iemura, S.; Natsume, T.; Spearman, H.; Beeson, D.; Yamanashi, Y. Mutations causing DOK7 congenital myasthenia ablate functional motifs in Dok-7. J. Biol. Chem. 2008, 283, 5518–5524. [Google Scholar] [CrossRef] [PubMed]
  283. Selcen, D.; Milone, M.; Shen, X.M.; Harper, C.M.; Stans, A.A.; Wieben, E.D.; Engel, A.G. Dok-7 myasthenia: Phenotypic and molecular genetic studies in 16 patients. Ann. Neurol. 2008, 64, 71–87. [Google Scholar] [CrossRef]
  284. Zhang, S.; Ohkawara, B.; Ito, M.; Huang, Z.; Zhao, F.; Nakata, T.; Takeuchi, T.; Sakurai, H.; Komaki, H.; Kamon, M.; et al. A mutation in DOK7 in congenital myasthenic syndrome forms aggresome in cultured cells, and reduces DOK7 expression and MuSK phosphorylation in patient-derived iPS cells. Hum. Mol. Genet. 2022, ddac306. [Google Scholar] [CrossRef] [PubMed]
  285. Huze, C.; Bauche, S.; Richard, P.; Chevessier, F.; Goillot, E.; Gaudon, K.; Ben Ammar, A.; Chaboud, A.; Grosjean, I.; Lecuyer, H.A.; et al. Identification of an agrin mutation that causes congenital myasthenia and affects synapse function. Am. J. Hum. Genet. 2009, 85, 155–167. [Google Scholar] [CrossRef]
  286. Maselli, R.A.; Fernandez, J.M.; Arredondo, J.; Navarro, C.; Ngo, M.; Beeson, D.; Cagney, O.; Williams, D.C.; Wollmann, R.L.; Yarov-Yarovoy, V.; et al. LG2 agrin mutation causing severe congenital myasthenic syndrome mimics functional characteristics of non-neural (z-) agrin. Hum. Genet. 2012, 131, 1123–1135. [Google Scholar] [CrossRef]
  287. Xi, J.; Yan, C.; Liu, W.W.; Qiao, K.; Lin, J.; Tian, X.; Wu, H.; Lu, J.; Wong, L.J.; Beeson, D.; et al. Novel SEA and LG2 Agrin mutations causing congenital Myasthenic syndrome. Orphanet J. Rare Dis. 2017, 12, 182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  288. Zhang, Y.; Dai, Y.; Han, J.N.; Chen, Z.H.; Ling, L.; Pu, C.Q.; Cui, L.Y.; Huang, X.S. A Novel AGRN Mutation Leads to Congenital Myasthenic Syndrome Only Affecting Limb-girdle Muscle. Chin. Med. J. 2017, 130, 2279–2282. [Google Scholar] [PubMed]
  289. Wu, L.; Brady, L.; Shoffner, J.; Tarnopolsky, M.A. Next-Generation Sequencing to Diagnose Muscular Dystrophy, Rhabdomyolysis, and HyperCKemia. Can. J. Neurol. Sci. 2018, 45, 262–268. [Google Scholar] [CrossRef]
  290. Gan, S.; Yang, H.; Xiao, T.; Pan, Z.; Wu, L. AGRN Gene Mutation Leads to Congenital Myasthenia Syndromes: A Pediatric Case Report and Literature Review. Neuropediatrics 2020, 51, 364–367. [Google Scholar] [CrossRef]
  291. Wang, A.; Xiao, Y.; Huang, P.; Liu, L.; Xiong, J.; Li, J.; Mao, D.; Liu, L. Novel NtA and LG1 Mutations in Agrin in a Single Patient Causes Congenital Myasthenic Syndrome. Front. Neurol. 2020, 11, 239. [Google Scholar] [CrossRef]
  292. Xia, P.; Xie, F.; Zhou, Z.J.; Lv, W. Novel LG1 Mutations in Agrin Causing Congenital Myasthenia Syndrome. Intern. Med. 2022, 61, 887–890. [Google Scholar] [CrossRef]
  293. Geremek, M.; Dudarewicz, L.; Obersztyn, E.; Paczkowska, M.; Smyk, M.; Sobecka, K.; Nowakowska, B. Null variants in AGRN cause lethal fetal akinesia deformation sequence. Clin. Genet. 2020, 97, 634–638. [Google Scholar] [CrossRef] [PubMed]
  294. Takata, A.; Miyake, N.; Tsurusaki, Y.; Fukai, R.; Miyatake, S.; Koshimizu, E.; Kushima, I.; Okada, T.; Morikawa, M.; Uno, Y.; et al. Integrative Analyses of De Novo Mutations Provide Deeper Biological Insights into Autism Spectrum Disorder. Cell Rep. 2018, 22, 734–747. [Google Scholar] [CrossRef]
  295. Previtali, S.C.; Zhao, E.; Lazarevic, D.; Pipitone, G.B.; Fabrizi, G.M.; Manganelli, F.; Mazzeo, A.; Pareyson, D.; Schenone, A.; Taroni, F.; et al. Expanding the spectrum of genes responsible for hereditary motor neuropathies. J. Neurol. Neurosurg. Psychiatry 2019, 90, 1171–1179. [Google Scholar] [CrossRef]
  296. Mihaylova, V.; Salih, M.A.; Mukhtar, M.M.; Abuzeid, H.A.; El-Sadig, S.M.; von der Hagen, M.; Huebner, A.; Nurnberg, G.; Abicht, A.; Muller, J.S.; et al. Refinement of the clinical phenotype in musk-related congenital myasthenic syndromes. Neurology 2009, 73, 1926–1928. [Google Scholar] [CrossRef] [PubMed]
  297. Ben Ammar, A.; Soltanzadeh, P.; Bauche, S.; Richard, P.; Goillot, E.; Herbst, R.; Gaudon, K.; Huze, C.; Schaeffer, L.; Yamanashi, Y.; et al. A mutation causes MuSK reduced sensitivity to agrin and congenital myasthenia. PLoS ONE 2013, 8, e53826. [Google Scholar] [CrossRef]
  298. Maggi, L.; Brugnoni, R.; Scaioli, V.; Winden, T.L.; Morandi, L.; Engel, A.G.; Mantegazza, R.; Bernasconi, P. Marked phenotypic variability in two siblings with congenital myasthenic syndrome due to mutations in MUSK. J. Neurol. 2013, 260, 2894–2896. [Google Scholar] [CrossRef]
  299. Gallenmuller, C.; Muller-Felber, W.; Dusl, M.; Stucka, R.; Guergueltcheva, V.; Blaschek, A.; von der Hagen, M.; Huebner, A.; Muller, J.S.; Lochmuller, H.; et al. Salbutamol-responsive limb-girdle congenital myasthenic syndrome due to a novel missense mutation and heteroallelic deletion in MUSK. Neuromuscul. Disord. 2014, 24, 31–35. [Google Scholar] [CrossRef]
  300. Al-Shahoumi, R.; Brady, L.I.; Schwartzentruber, J.; Tarnopolsky, M.A. Two cases of congenital myasthenic syndrome with vocal cord paralysis. Neurology 2015, 84, 1281–1282. [Google Scholar] [CrossRef]
  301. Giarrana, M.L.; Joset, P.; Sticht, H.; Robb, S.; Steindl, K.; Rauch, A.; Klein, A. A severe congenital myasthenic syndrome with “dropped head” caused by novel MUSK mutations. Muscle Nerve 2015, 52, 668–673. [Google Scholar] [CrossRef]
  302. Owen, D.; Topf, A.; Preethish-Kumar, V.; Lorenzoni, P.J.; Vroling, B.; Scola, R.H.; Dias-Tosta, E.; Geraldo, A.; Polavarapu, K.; Nashi, S.; et al. Recessive variants of MuSK are associated with late onset CMS and predominant limb girdle weakness. Am. J. Med. Genet. A 2018, 176, 1594–1601. [Google Scholar] [CrossRef]
  303. Pinto, M.V.; Saw, J.L.; Milone, M. Congenital Vocal Cord Paralysis and Late-Onset Limb-Girdle Weakness in MuSK-Congenital Myasthenic Syndrome. Front. Neurol. 2019, 10, 1300. [Google Scholar] [CrossRef]
  304. Shen, Y.; Wang, B.; Zheng, X.; Zhang, W.; Wu, H.; Hei, M. A Neonate With MuSK Congenital Myasthenic Syndrome Presenting With Refractory Respiratory Failure. Front. Pediatr. 2020, 8, 166. [Google Scholar] [CrossRef] [PubMed]
  305. Rodriguez Cruz, P.M.; Cossins, J.; Cheung, J.; Maxwell, S.; Jayawant, S.; Herbst, R.; Waithe, D.; Kornev, A.P.; Palace, J.; Beeson, D. Congenital myasthenic syndrome due to mutations in MUSK suggests that the level of MuSK phosphorylation is crucial for governing synaptic structure. Hum. Mutat. 2020, 41, 619–631. [Google Scholar] [CrossRef]
  306. Tan-Sindhunata, M.B.; Mathijssen, I.B.; Smit, M.; Baas, F.; de Vries, J.I.; van der Voorn, J.P.; Kluijt, I.; Hagen, M.A.; Blom, E.W.; Sistermans, E.; et al. Identification of a Dutch founder mutation in MUSK causing fetal akinesia deformation sequence. Eur. J. Hum. Genet. 2015, 23, 1151–1157. [Google Scholar] [CrossRef] [PubMed]
  307. Wilbe, M.; Ekvall, S.; Eurenius, K.; Ericson, K.; Casar-Borota, O.; Klar, J.; Dahl, N.; Ameur, A.; Anneren, G.; Bondeson, M.L. MuSK: A new target for lethal fetal akinesia deformation sequence (FADS). J. Med. Genet. 2015, 52, 195–202. [Google Scholar] [CrossRef]
  308. Palace, J.; Lashley, D.; Newsom-Davis, J.; Cossins, J.; Maxwell, S.; Kennett, R.; Jayawant, S.; Yamanashi, Y.; Beeson, D. Clinical features of the DOK7 neuromuscular junction synaptopathy. Brain 2007, 130 Pt 6, 1507–1515. [Google Scholar] [CrossRef]
  309. Anderson, J.A.; Ng, J.J.; Bowe, C.; McDonald, C.; Richman, D.P.; Wollmann, R.L.; Maselli, R.A. Variable phenotypes associated with mutations in DOK7. Muscle Nerve 2008, 37, 448–456. [Google Scholar] [CrossRef]
  310. Schara, U.; Barisic, N.; Deschauer, M.; Lindberg, C.; Straub, V.; Strigl-Pill, N.; Wendt, M.; Abicht, A.; Muller, J.S.; Lochmuller, H. Ephedrine therapy in eight patients with congenital myasthenic syndrome due to DOK7 mutations. Neuromuscul. Disord. 2009, 19, 828–832. [Google Scholar] [CrossRef]
  311. Srour, M.; Bolduc, V.; Guergueltcheva, V.; Lochmuller, H.; Gendron, D.; Shevell, M.I.; Poulin, C.; Mathieu, J.; Bouchard, J.P.; Brais, B. DOK7 mutations presenting as a proximal myopathy in French Canadians. Neuromuscul. Disord. 2010, 20, 453–457. [Google Scholar] [CrossRef] [PubMed]
  312. Jephson, C.G.; Mills, N.A.; Pitt, M.C.; Beeson, D.; Aloysius, A.; Muntoni, F.; Robb, S.A.; Bailey, C.M. Congenital stridor with feeding difficulty as a presenting symptom of Dok7 congenital myasthenic syndrome. Int. J. Pediatr. Otorhinolaryngol. 2010, 74, 991–994. [Google Scholar] [CrossRef] [PubMed]
  313. Cossins, J.; Liu, W.W.; Belaya, K.; Maxwell, S.; Oldridge, M.; Lester, T.; Robb, S.; Beeson, D. The spectrum of mutations that underlie the neuromuscular junction synaptopathy in DOK7 congenital myasthenic syndrome. Hum. Mol. Genet. 2012, 21, 3765–3775. [Google Scholar] [CrossRef] [PubMed]
  314. Irahara, K.; Komaki, H.; Honda, R.; Okumura, A.; Shiraishi, K.; Kobayashi, Y.; Azuma, Y.; Nakata, T.; Ohya, Y.; Sasaki, M. Clinical features of congenital myasthenic syndrome in Japan. No To Hattatsu 2012, 44, 450–454. [Google Scholar]
  315. Mahjneh, I.; Lochmuller, H.; Muntoni, F.; Abicht, A. DOK7 limb-girdle myasthenic syndrome mimicking congenital muscular dystrophy. Neuromuscul. Disord. 2013, 23, 36–42. [Google Scholar] [CrossRef]
  316. Burke, G.; Hiscock, A.; Klein, A.; Niks, E.H.; Main, M.; Manzur, A.Y.; Ng, J.; de Vile, C.; Muntoni, F.; Beeson, D.; et al. Salbutamol benefits children with congenital myasthenic syndrome due to DOK7 mutations. Neuromuscul. Disord. 2013, 23, 170–175. [Google Scholar] [CrossRef]
  317. Lorenzoni, P.J.; Scola, R.H.; Kay, C.S.; Filla, L.; Miranda, A.P.; Pinheiro, J.M.; Chaouch, A.; Lochmuller, H.; Werneck, L.C. Salbutamol therapy in congenital myasthenic syndrome due to DOK7 mutation. J. Neurol. Sci. 2013, 331, 155–157. [Google Scholar] [CrossRef] [PubMed]
  318. Lorenzoni, P.J.; Scola, R.H.; Kay, C.S.; Lochmuller, H.; Werneck, L.C. Congenital myasthenic syndrome and minicore-like myopathy with DOK7 mutation. Muscle Nerve 2013, 48, 151–152. [Google Scholar] [CrossRef]
  319. Nishikawa, A.; Mori-Yoshimura, M.; Okamoto, T.; Oya, Y.; Nakata, T.; Ohno, K.; Murata, M. Beneficial effects of 3,4-diaminopyridine in a 26-year-old woman with DOK7 congenital myasthenic syndrome who was originally diagnosed with facioscapulohumeral dystrophy. Rinsho Shinkeigaku 2014, 54, 561–564. [Google Scholar] [CrossRef] [PubMed]
  320. Witting, N.; Crone, C.; Duno, M.; Vissing, J. Clinical and neurophysiological response to pharmacological treatment of DOK7 congenital myasthenia in an older patient. Clin. Neurol. Neurosurg. 2015, 130, 168–170. [Google Scholar] [CrossRef]
  321. Bevilacqua, J.A.; Lara, M.; Diaz, J.; Campero, M.; Vazquez, J.; Maselli, R.A. Congenital Myasthenic Syndrome due to DOK7 mutations in a family from Chile. Eur. J. Transl. Myol. 2017, 27, 6832. [Google Scholar] [CrossRef]
  322. Gaist, D.; Mogensen, J.; Pedersen, E.G.; Schroder, H.D.; Vissing, J.; Andersen, H.; Hertz, J.M. DOK7 congenital myasthenia may be associated with severe mitral valve insufficiency. J. Neurol. Sci. 2017, 379, 217–218. [Google Scholar] [CrossRef]
  323. Daum, H.; Meiner, V.; Elpeleg, O.; Harel, T.; Collaborating, A. Fetal exome sequencing: Yield and limitations in a tertiary referral center. Ultrasound Obstet. Gynecol. 2019, 53, 80–86. [Google Scholar] [CrossRef]
  324. Tayade, K.; Salunkhe, M.; Agarwal, A.; Radhakrishnan, D.M.; Srivastava, A.K. DOK7 congenital myasthenic syndrome responsive to oral salbutamol. QJM 2022, 115, 323–324. [Google Scholar] [CrossRef]
  325. Bastos, P.; Barbosa, R.; Fernandes, M.; Alonso, I. A late-onset congenital myasthenic syndrome due to a heterozygous DOK7 mutation. Neuromuscul. Disord. 2020, 30, 331–335. [Google Scholar] [CrossRef] [PubMed]
  326. Miyana, K.; Ishiyama, A.; Saito, Y.; Nishino, I. Tulobuterol is a potential therapeutic drug in congenital myasthenic syndrome. Pediatr. Int. 2022, 64, e15115. [Google Scholar] [CrossRef] [PubMed]
  327. Santos, M.; Cruz, S.; Peres, J.; Santos, L.; Tavares, P.; Basto, J.P.; Salgado, V.; Valverde, A.H. DOK7 myasthenic syndrome with subacute adult onset during pregnancy and partial response to fluoxetine. Neuromuscul. Disord. 2018, 28, 278–282. [Google Scholar] [CrossRef] [PubMed]
  328. Oury, J.; Zhang, W.; Leloup, N.; Koide, A.; Corrado, A.D.; Ketavarapu, G.; Hattori, T.; Koide, S.; Burden, S.J. Mechanism of disease and therapeutic rescue of Dok7 congenital myasthenia. Nature 2021, 595, 404–408. [Google Scholar] [CrossRef]
  329. Oh, S.J.; King, P.H.; Schindler, A. Life-Long Steroid Responsive Familial Myopathy With Docking Protein 7 Mutation. J. Clin. Neuromuscul. Dis. 2022, 24, 80–84. [Google Scholar] [CrossRef]
  330. Smith, F.J.; Eady, R.A.; Leigh, I.M.; McMillan, J.R.; Rugg, E.L.; Kelsell, D.P.; Bryant, S.P.; Spurr, N.K.; Geddes, J.F.; Kirtschig, G.; et al. Plectin deficiency results in muscular dystrophy with epidermolysis bullosa. Nat. Genet. 1996, 13, 450–457. [Google Scholar] [CrossRef]
  331. Gundesli, H.; Talim, B.; Korkusuz, P.; Balci-Hayta, B.; Cirak, S.; Akarsu, N.A.; Topaloglu, H.; Dincer, P. Mutation in exon 1f of PLEC, leading to disruption of plectin isoform 1f, causes autosomal-recessive limb-girdle muscular dystrophy. Am. J. Hum. Genet. 2010, 87, 834–841. [Google Scholar] [CrossRef]
  332. Banwell, B.L.; Russel, J.; Fukudome, T.; Shen, X.M.; Stilling, G.; Engel, A.G. Myopathy, myasthenic syndrome, and epidermolysis bullosa simplex due to plectin deficiency. J. Neuropathol. Exp. Neurol. 1999, 58, 832–846. [Google Scholar] [CrossRef] [PubMed]
  333. Maselli, R.A.; Arredondo, J.; Cagney, O.; Mozaffar, T.; Skinner, S.; Yousif, S.; Davis, R.R.; Gregg, J.P.; Sivak, M.; Konia, T.H.; et al. Congenital myasthenic syndrome associated with epidermolysis bullosa caused by homozygous mutations in PLEC1 and CHRNE. Clin. Genet. 2011, 80, 444–451. [Google Scholar] [CrossRef] [PubMed]
  334. Selcen, D.; Juel, V.C.; Hobson-Webb, L.D.; Smith, E.C.; Stickler, D.E.; Bite, A.V.; Ohno, K.; Engel, A.G. Myasthenic syndrome caused by plectinopathy. Neurology 2011, 76, 327–336. [Google Scholar] [CrossRef] [PubMed]
  335. Mihailovska, E.; Raith, M.; Valencia, R.G.; Fischer, I.; Al Banchaabouchi, M.; Herbst, R.; Wiche, G. Neuromuscular synapse integrity requires linkage of acetylcholine receptors to postsynaptic intermediate filament networks via rapsyn-plectin 1f complexes. Mol. Biol. Cell 2014, 25, 4130–4149. [Google Scholar] [CrossRef] [PubMed]
  336. Fine, J.D.; Stenn, J.; Johnson, L.; Wright, T.; Bock, H.G.; Horiguchi, Y. Autosomal recessive epidermolysis bullosa simplex. Generalized phenotypic features suggestive of junctional or dystrophic epidermolysis bullosa, and association with neuromuscular diseases. Arch. Dermatol. 1989, 125, 931–938. [Google Scholar] [CrossRef] [PubMed]
  337. Fattahi, Z.; Kahrizi, K.; Nafissi, S.; Fadaee, M.; Abedini, S.S.; Kariminejad, A.; Akbari, M.R.; Najmabadi, H. Report of a patient with limb-girdle muscular dystrophy, ptosis and ophthalmoparesis caused by plectinopathy. Arch. Iran. Med. 2015, 18, 60–64. [Google Scholar]
  338. Gonzalez Garcia, A.; Tutmaher, M.S.; Upadhyayula, S.R.; Sanchez Russo, R.; Verma, S. Novel PLEC gene variants causing congenital myasthenic syndrome. Muscle Nerve 2019, 60, E40–E43. [Google Scholar] [CrossRef]
  339. Vahidnezhad, H.; Youssefian, L.; Harvey, N.; Tavasoli, A.R.; Saeidian, A.H.; Sotoudeh, S.; Varghaei, A.; Mahmoudi, H.; Mansouri, P.; Mozafari, N.; et al. Mutation update: The spectra of PLEC sequence variants and related plectinopathies. Hum. Mutat. 2022, 43, 1706–1731. [Google Scholar] [CrossRef]
  340. Ohno, K.; Tsujino, A.; Shen, X.-M.; Brengman, J.M.; Harper, C.M.; Bajzer, Z.; Udd, B.; Beyring, R.; Robb, S.; Kirkham, F.J.; et al. Choline acetyltransferase mutations cause myasthenic syndrome associated with episodic apnea in humans. Proc. Natl. Acad. Sci. USA 2001, 98, 2017–2022. [Google Scholar] [CrossRef]
  341. Shen, X.M.; Crawford, T.O.; Brengman, J.; Acsadi, G.; Iannaconne, S.; Karaca, E.; Khoury, C.; Mah, J.K.; Edvardson, S.; Bajzer, Z.; et al. Functional consequences and structural interpretation of mutations of human choline acetyltransferase. Hum. Mutat. 2011, 32, 1259–1267. [Google Scholar] [CrossRef] [PubMed]
  342. Stankiewicz, P.; Kulkarni, S.; Dharmadhikari, A.V.; Sampath, S.; Bhatt, S.S.; Shaikh, T.H.; Xia, Z.; Pursley, A.N.; Cooper, M.L.; Shinawi, M.; et al. Recurrent deletions and reciprocal duplications of 10q11.21q11.23 including CHAT and SLC18A3 are likely mediated by complex low-copy repeats. Hum. Mutat. 2012, 33, 165–179. [Google Scholar] [CrossRef]
  343. Schwartz, M.; Sternberg, D.; Whalen, S.; Afenjar, A.; Isapof, A.; Chabrol, B.; Portnoi, M.F.; Heide, S.; Keren, B.; Chantot-Bastaraud, S.; et al. How chromosomal deletions can unmask recessive mutations? Deletions in 10q11.2 associated with CHAT or SLC18A3 mutations lead to congenital myasthenic syndrome. Am. J. Med. Genet. A 2018, 176, 151–155. [Google Scholar] [CrossRef]
  344. Barwick, K.E.; Wright, J.; Al-Turki, S.; McEntagart, M.M.; Nair, A.; Chioza, B.; Al-Memar, A.; Modarres, H.; Reilly, M.M.; Dick, K.J.; et al. Defective presynaptic choline transport underlies hereditary motor neuropathy. Am. J. Hum. Genet. 2012, 91, 1103–1107. [Google Scholar] [CrossRef] [PubMed]
  345. Ingram, G.; Barwick, K.E.; Hartley, L.; McEntagart, M.; Crosby, A.H.; Llewelyn, G.; Morris, H.R. Distal hereditary motor neuropathy with vocal cord paresis: From difficulty in choral singing to a molecular genetic diagnosis. Pract. Neurol. 2016, 16, 247–251. [Google Scholar] [CrossRef]
  346. Wang, H.; Salter, C.G.; Refai, O.; Hardy, H.; Barwick, K.E.S.; Akpulat, U.; Kvarnung, M.; Chioza, B.A.; Harlalka, G.; Taylan, F.; et al. Choline transporter mutations in severe congenital myasthenic syndrome disrupt transporter localization. Brain 2017, 140, 2838–2850. [Google Scholar] [CrossRef] [PubMed]
  347. Ferguson, S.M.; Bazalakova, M.; Savchenko, V.; Tapia, J.C.; Wright, J.; Blakely, R.D. Lethal impairment of cholinergic neurotransmission in hemicholinium-3-sensitive choline transporter knockout mice. Proc. Natl. Acad. Sci. USA 2004, 101, 8762–8767. [Google Scholar] [CrossRef] [PubMed]
  348. Lund, D.; Ruggiero, A.M.; Ferguson, S.M.; Wright, J.; English, B.A.; Reisz, P.A.; Whitaker, S.M.; Peltier, A.C.; Blakely, R.D. Motor neuron-specific overexpression of the presynaptic choline transporter: Impact on motor endurance and evoked muscle activity. Neuroscience 2010, 171, 1041–1053. [Google Scholar] [CrossRef]
  349. English, B.A.; Appalsamy, M.; Diedrich, A.; Ruggiero, A.M.; Lund, D.; Wright, J.; Keller, N.R.; Louderback, K.M.; Robertson, D.; Blakely, R.D. Tachycardia, reduced vagal capacity, and age-dependent ventricular dysfunction arising from diminished expression of the presynaptic choline transporter. Am. J. Physiol. Heart Circ. Physiol. 2010, 299, H799–H810. [Google Scholar] [CrossRef]
  350. Jaeken, J.; Martens, K.; Francois, I.; Eyskens, F.; Lecointre, C.; Derua, R.; Meulemans, S.; Slootstra, J.W.; Waelkens, E.; de Zegher, F.; et al. Deletion of PREPL, a gene encoding a putative serine oligopeptidase, in patients with hypotonia-cystinuria syndrome. Am. J. Hum. Genet. 2006, 78, 38–51. [Google Scholar] [CrossRef]
  351. Regal, L.; Shen, X.M.; Selcen, D.; Verhille, C.; Meulemans, S.; Creemers, J.W.; Engel, A.G. PREPL deficiency with or without cystinuria causes a novel myasthenic syndrome. Neurology 2014, 82, 1254–1260. [Google Scholar] [CrossRef]
  352. Barbosa, M.; Lopes, A.; Mota, C.; Martins, E.; Oliveira, J.; Alves, S.; De Bonis, P.; Mota Mdo, C.; Dias, C.; Rodrigues-Santos, P.; et al. Clinical, biochemical and molecular characterization of cystinuria in a cohort of 12 patients. Clin. Genet. 2012, 81, 47–55. [Google Scholar] [CrossRef] [PubMed]
  353. Maselli, R.A.; Chen, D.; Mo, D.; Bowe, C.; Fenton, G.; Wollmann, R.L. Choline acetyltransferase mutations in myasthenic syndrome due to deficient acetylcholine resynthesis. Muscle Nerve 2003, 27, 180–187. [Google Scholar] [CrossRef] [PubMed]
  354. Schmidt, C.; Abicht, A.; Krampfl, K.; Voss, W.; Stucka, R.; Mildner, G.; Petrova, S.; Schara, U.; Mortier, W.; Bufler, J.; et al. Congenital myasthenic syndrome due to a novel missense mutation in the gene encoding choline acetyltransferase. Neuromuscul. Disord. 2003, 13, 245–251. [Google Scholar] [CrossRef] [PubMed]
  355. Barisic, N.; Muller, J.S.; Paucic-Kirincic, E.; Gazdik, M.; Lah-Tomulic, K.; Pertl, A.; Sertic, J.; Zurak, N.; Lochmuller, H.; Abicht, A. Clinical variability of CMS-EA (congenital myasthenic syndrome with episodic apnea) due to identical CHAT mutations in two infants. Eur. J. Paediatr. Neurol. 2005, 9, 7–12. [Google Scholar] [CrossRef] [PubMed]
  356. Mallory, L.A.; Shaw, J.G.; Burgess, S.L.; Estrella, E.; Nurko, S.; Burpee, T.M.; Agus, M.S.; Darras, B.T.; Kunkel, L.M.; Kang, P.B. Congenital myasthenic syndrome with episodic apnea. Pediatr. Neurol. 2009, 41, 42–45. [Google Scholar] [CrossRef]
  357. Yeung, W.L.; Lam, C.W.; Fung, L.W.; Hon, K.L.; Ng, P.C. Severe congenital myasthenia gravis of the presynaptic type with choline acetyltransferase mutation in a Chinese infant with respiratory failure. Neonatology 2009, 95, 183–186. [Google Scholar] [CrossRef] [PubMed]
  358. Schara, U.; Christen, H.J.; Durmus, H.; Hietala, M.; Krabetz, K.; Rodolico, C.; Schreiber, G.; Topaloglu, H.; Talim, B.; Voss, W.; et al. Long-term follow-up in patients with congenital myasthenic syndrome due to CHAT mutations. Eur. J. Paediatr. Neurol. 2010, 14, 326–333. [Google Scholar] [CrossRef]
  359. Dilena, R.; Abicht, A.; Sergi, P.; Comi, G.P.; Di Fonzo, A.; Chidini, G.; Natacci, F.; Barbieri, S.; Lochmuller, H. Congenital myasthenic syndrome due to choline acetyltransferase mutations in infants: Clinical suspicion and comprehensive electrophysiological assessment are important for early diagnosis. J. Child Neurol. 2014, 29, 389–393. [Google Scholar] [CrossRef]
  360. Arredondo, J.; Lara, M.; Gospe, S.M., Jr.; Mazia, C.G.; Vaccarezza, M.; Garcia-Erro, M.; Bowe, C.M.; Chang, C.H.; Mezei, M.M.; Maselli, R.A. Choline Acetyltransferase Mutations Causing Congenital Myasthenic Syndrome: Molecular Findings and Genotype-Phenotype Correlations. Hum. Mutat. 2015, 36, 881–893. [Google Scholar] [CrossRef] [PubMed]
  361. Tan, J.S.; Ambang, T.; Ahmad-Annuar, A.; Rajahram, G.S.; Wong, K.T.; Goh, K.J. Congenital myasthenic syndrome due to novel CHAT mutations in an ethnic kadazandusun family. Muscle Nerve 2016, 53, 822–826. [Google Scholar] [CrossRef]
  362. Brunelli, L.; Mao, R.; Jenkins, S.M.; Bleyl, S.B.; Dames, S.A.; Miller, C.E.; Ostrander, B.; Tvrdik, T.; Andrews, S.; Flores, J.; et al. A rapid gene sequencing panel strategy to facilitate precision neonatal medicine. Am. J. Med. Genet. A 2017, 173, 1979–1982. [Google Scholar] [CrossRef] [PubMed]
  363. Liu, Z.M.; Fang, F.; Ding, C.H.; Zhang, W.H.; Deng, J.; Chen, C.H.; Wang, X.; Liu, J.; Li, Z.; Jia, X.L.; et al. Clinical and genetic characteristics of congenital myasthenia syndrome with episodic apnea caused by CHAT gene mutation: A report of 2 cases. Zhonghua Er Ke Za Zhi 2018, 56, 216–220. [Google Scholar] [PubMed]
  364. Zhang, Y.; Cheng, X.; Luo, C.; Lei, M.; Mao, F.; Shi, Z.; Cao, W.; Zhang, J.; Zhang, Q. Congenital Myasthenic Syndrome Caused by a Novel Hemizygous CHAT Mutation. Front. Pediatr. 2020, 8, 185. [Google Scholar] [CrossRef] [PubMed]
  365. O’Grady, G.L.; Verschuuren, C.; Yuen, M.; Webster, R.; Menezes, M.; Fock, J.M.; Pride, N.; Best, H.A.; Benavides Damm, T.; Turner, C.; et al. Variants in SLC18A3, vesicular acetylcholine transporter, cause congenital myasthenic syndrome. Neurology 2016, 87, 1442–1448. [Google Scholar] [CrossRef]
  366. Aran, A.; Segel, R.; Kaneshige, K.; Gulsuner, S.; Renbaum, P.; Oliphant, S.; Meirson, T.; Weinberg-Shukron, A.; Hershkovitz, Y.; Zeligson, S.; et al. Vesicular acetylcholine transporter defect underlies devastating congenital myasthenia syndrome. Neurology 2017, 88, 1021–1028. [Google Scholar] [CrossRef] [PubMed]
  367. Lamond, A.; Buckley, D.; O’Dea, J.; Turner, L. Variants of SLC18A3 leading to congenital myasthenic syndrome in two children with varying presentations. BMJ Case Rep. 2021, 14, e237799. [Google Scholar] [CrossRef] [PubMed]
  368. Pardal-Fernandez, J.M.; Carrascosa-Romero, M.C.; Alvarez, S.; Medina-Monzon, M.C.; Caamano, M.B.; de Cabo, C. A new severe mutation in the SLC5A7 gene related to congenital myasthenic syndrome type 20. Neuromuscul. Disord. 2018, 28, 881–884. [Google Scholar] [CrossRef] [PubMed]
  369. Baker, S.W.; Murrell, J.R.; Nesbitt, A.I.; Pechter, K.B.; Balciuniene, J.; Zhao, X.; Yu, Z.; Denenberg, E.H.; DeChene, E.T.; Wilkens, A.B.; et al. Automated Clinical Exome Reanalysis Reveals Novel Diagnoses. J. Mol. Diagn. 2019, 21, 38–48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  370. Silva, S.; Miyake, N.; Tapia, C.; Matsumoto, N. The second point mutation in PREPL: A case report and literature review. J. Hum. Genet. 2018, 63, 677–681. [Google Scholar] [CrossRef]
  371. Laugwitz, L.; Redler, S.; Buchert, R.; Sturm, M.; Zeile, I.; Schara, U.; Wieczorek, D.; Haack, T.; Distelmaier, F. Isolated PREPL deficiency associated with congenital myasthenic syndrome-22. Klin. Padiatr. 2018, 230, 281–283. [Google Scholar] [CrossRef]
  372. Kim, M.J.; Yum, M.S.; Seo, G.H.; Lee, Y.; Jang, H.N.; Ko, T.S.; Lee, B.H. Clinical Application of Whole Exome Sequencing to Identify Rare but Remediable Neurologic Disorders. J. Clin. Med. 2020, 9, 3724. [Google Scholar] [CrossRef] [PubMed]
  373. Shchagina, O.; Bessonova, L.; Bychkov, I.; Beskorovainaya, T.; Poliakov, A. A Family Case of Congenital Myasthenic Syndrome-22 Induced by Different Combinations of Molecular Causes in Siblings. Genes 2020, 11, 821. [Google Scholar] [CrossRef] [PubMed]
  374. Zhang, P.; Wu, B.; Lu, Y.; Ni, Q.; Liu, R.; Zhou, W.; Wang, H. First maternal uniparental disomy for chromosome 2 with PREPL novel frameshift mutation of congenital myasthenic syndrome 22 in an infant. Mol. Genet. Genom. Med. 2020, 8, e1144. [Google Scholar] [CrossRef]
  375. Yang, Q.; Hua, R.; Qian, J.; Yi, S.; Shen, F.; Zhang, Q.; Li, M.; Yi, S.; Luo, J.; Fan, X. PREPL Deficiency: A Homozygous Splice Site PREPL Mutation in a Patient With Congenital Myasthenic Syndrome and Absence of Ovaries and Hypoplasia of Uterus. Front. Genet. 2020, 11, 198. [Google Scholar] [CrossRef] [PubMed]
  376. Prior, D.E.; Ghosh, P.S. Congenital Myasthenic Syndrome From a Single Center: Phenotypic and Genotypic features. J. Child Neurol. 2021, 36, 610–617. [Google Scholar] [CrossRef]
  377. Rizo, J.; Xu, J. The Synaptic Vesicle Release Machinery. Annu. Rev. Biophys. 2015, 44, 339–367. [Google Scholar] [CrossRef] [PubMed]
  378. Lipstein, N.; Verhoeven-Duif, N.M.; Michelassi, F.E.; Calloway, N.; van Hasselt, P.M.; Pienkowska, K.; van Haaften, G.; van Haelst, M.M.; van Empelen, R.; Cuppen, I.; et al. Synaptic UNC13A protein variant causes increased neurotransmission and dyskinetic movement disorder. J. Clin. Investig. 2017, 127, 1005–1018. [Google Scholar] [CrossRef] [PubMed]
  379. Salpietro, V.; Lin, W.; Delle Vedove, A.; Storbeck, M.; Liu, Y.; Efthymiou, S.; Manole, A.; Wiethoff, S.; Ye, Q.; Saggar, A.; et al. Homozygous mutations in VAMP1 cause a presynaptic congenital myasthenic syndrome. Ann. Neurol. 2017, 81, 597–603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  380. Al-Muhaizea, M.A.; AlQuait, L.; AlRasheed, A.; AlHarbi, S.; Albader, A.A.; AlMass, R.; Albakheet, A.; Alhumaidan, A.; AlRasheed, M.M.; Colak, D.; et al. Pyrostigmine therapy in a patient with VAMP1-related congenital myasthenic syndrome. Neuromuscul. Disord. 2020, 30, 611–615. [Google Scholar] [CrossRef] [PubMed]
  381. Polavarapu, K.; Vengalil, S.; Preethish-Kumar, V.; Arunachal, G.; Nashi, S.; Mohan, D.; Chawla, T.; Bardhan, M.; Nandeesh, B.; Gupta, P.; et al. Recessive VAMP1 mutations associated with severe congenital myasthenic syndromes—A recognizable clinical phenotype. Eur. J. Paediatr. Neurol. 2021, 31, 54–60. [Google Scholar] [CrossRef] [PubMed]
  382. Zhou, Y.; Schopperle, W.M.; Murrey, H.; Jaramillo, A.; Dagan, D.; Griffith, L.C.; Levitan, I.B. A dynamically regulated 14-3-3, Slob, and Slowpoke potassium channel complex in Drosophila presynaptic nerve terminals. Neuron 1999, 22, 809–818. [Google Scholar] [CrossRef] [PubMed]
  383. Schluter, O.M.; Schnell, E.; Verhage, M.; Tzonopoulos, T.; Nicoll, R.A.; Janz, R.; Malenka, R.C.; Geppert, M.; Sudhof, T.C. Rabphilin knock-out mice reveal that rabphilin is not required for rab3 function in regulating neurotransmitter release. J. Neurosci. 1999, 19, 5834–5846. [Google Scholar] [CrossRef]
  384. Staunton, J.; Ganetzky, B.; Nonet, M.L. Rabphilin potentiates soluble N-ethylmaleimide sensitive factor attachment protein receptor function independently of rab3. J. Neurosci. 2001, 21, 9255–9264. [Google Scholar] [CrossRef] [PubMed]
  385. Burns, M.E.; Sasaki, T.; Takai, Y.; Augustine, G.J. Rabphilin-3A: A multifunctional regulator of synaptic vesicle traffic. J. Gen. Physiol. 1998, 111, 243–255. [Google Scholar] [CrossRef]
  386. Miner, J.H.; Cunningham, J.; Sanes, J.R. Roles for laminin in embryogenesis: Exencephaly, syndactyly, and placentopathy in mice lacking the laminin alpha5 chain. J. Cell Biol. 1998, 143, 1713–1723. [Google Scholar] [CrossRef] [PubMed]
  387. Nishimune, H.; Valdez, G.; Jarad, G.; Moulson, C.L.; Muller, U.; Miner, J.H.; Sanes, J.R. Laminins promote postsynaptic maturation by an autocrine mechanism at the neuromuscular junction. J. Cell Biol. 2008, 182, 1201–1215. [Google Scholar] [CrossRef] [PubMed]
  388. Wu, X.; Rush, J.S.; Karaoglu, D.; Krasnewich, D.; Lubinsky, M.S.; Waechter, C.J.; Gilmore, R.; Freeze, H.H. Deficiency of UDP-GlcNAc:Dolichol Phosphate N-Acetylglucosamine-1 Phosphate Transferase (DPAGT1) causes a novel congenital disorder of Glycosylation Type Ij. Hum. Mutat. 2003, 22, 144–150. [Google Scholar] [CrossRef]
  389. Thiel, C.; Schwarz, M.; Peng, J.; Grzmil, M.; Hasilik, M.; Braulke, T.; Kohlschutter, A.; von Figura, K.; Lehle, L.; Korner, C. A new type of congenital disorders of glycosylation (CDG-Ii) provides new insights into the early steps of dolichol-linked oligosaccharide biosynthesis. J. Biol. Chem. 2003, 278, 22498–22505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  390. Senderek, J.; Muller, J.S.; Dusl, M.; Strom, T.M.; Guergueltcheva, V.; Diepolder, I.; Laval, S.H.; Maxwell, S.; Cossins, J.; Krause, S.; et al. Hexosamine biosynthetic pathway mutations cause neuromuscular transmission defect. Am. J. Hum. Genet. 2011, 88, 162–172. [Google Scholar] [CrossRef]
  391. Zoltowska, K.; Webster, R.; Finlayson, S.; Maxwell, S.; Cossins, J.; Muller, J.; Lochmuller, H.; Beeson, D. Mutations in GFPT1 that underlie limb-girdle congenital myasthenic syndrome result in reduced cell-surface expression of muscle AChR. Hum. Mol. Genet. 2013, 22, 2905–2913. [Google Scholar] [CrossRef]
  392. Selvam, P.; Arunachal, G.; Danda, S.; Chapla, A.; Sivadasan, A.; Alexander, M.; Thomas, M.M.; Thomas, N.J. Congenital Myasthenic Syndrome: Spectrum of Mutations in an Indian Cohort. J. Clin. Neuromuscul. Dis. 2018, 20, 14–27. [Google Scholar] [CrossRef] [PubMed]
  393. Matsumoto, C.; Mori-Yoshimura, M.; Noguchi, S.; Endo, Y.; Oya, Y.; Murata, M.; Nishino, I.; Takahashi, Y. Phenotype of a limb-girdle congenital myasthenic syndrome patient carrying a GFPT1 mutation. Brain Dev. 2019, 41, 470–473. [Google Scholar] [CrossRef] [PubMed]
  394. Jiang, K.; Zheng, Y.; Lin, J.; Wu, X.; Yu, Y.; Zhu, M.; Fang, X.; Zhou, M.; Li, X.; Hong, D. Diverse myopathological features in the congenital myasthenia syndrome with GFPT1 mutation. Brain Behav. 2022, 12, e2469. [Google Scholar] [CrossRef] [PubMed]
  395. Basiri, K.; Belaya, K.; Liu, W.W.; Maxwell, S.; Sedghi, M.; Beeson, D. Clinical features in a large Iranian family with a limb-girdle congenital myasthenic syndrome due to a mutation in DPAGT1. Neuromuscul. Disord. 2013, 23, 469–472. [Google Scholar] [CrossRef]
  396. Klein, A.; Robb, S.; Rushing, E.; Liu, W.W.; Belaya, K.; Beeson, D. Congenital myasthenic syndrome caused by mutations in DPAGT. Neuromuscul. Disord. 2015, 25, 253–256. [Google Scholar] [CrossRef]
  397. Bogdanova-Mihaylova, P.; Murphy, R.P.J.; Alexander, M.D.; McHugh, J.C.; Foley, A.R.; Brett, F.; Murphy, S.M. Congenital myasthenic syndrome due to DPAGT1 mutations mimicking congenital myopathy in an Irish family. Eur. J. Neurol. 2018, 25, e22–e23. [Google Scholar] [CrossRef]
  398. Etzel, J.D.; Neely, K.A.; Ely, A.L. Congenital glycosylation disorder: A novel presentation of coexisting anterior and posterior segment pathology and its implications in pediatric cataract management. J. AAPOS 2019, 23, 297–300. [Google Scholar] [CrossRef]
  399. Luo, S.; Cai, S.; Maxwell, S.; Yue, D.; Zhu, W.; Qiao, K.; Zhu, Z.; Zhou, L.; Xi, J.; Lu, J.; et al. Novel mutations in the C-terminal region of GMPPB causing limb-girdle muscular dystrophy overlapping with congenital myasthenic syndrome. Neuromuscul. Disord. 2017, 27, 557–564. [Google Scholar] [CrossRef]
  400. Tian, W.T.; Zhou, H.Y.; Zhan, F.X.; Zhu, Z.Y.; Yang, J.; Chen, S.D.; Luan, X.H.; Cao, L. Lysosomal degradation of GMPPB is associated with limb-girdle muscular dystrophy type 2T. Ann. Clin. Transl. Neurol. 2019, 6, 1062–1071. [Google Scholar] [CrossRef]
  401. Monies, D.M.; Al-Hindi, H.N.; Al-Muhaizea, M.A.; Jaroudi, D.J.; Al-Younes, B.; Naim, E.A.; Wakil, S.M.; Meyer, B.F.; Bohlega, S. Clinical and pathological heterogeneity of a congenital disorder of glycosylation manifesting as a myasthenic/myopathic syndrome. Neuromuscul. Disord. 2014, 24, 353–359. [Google Scholar] [CrossRef]
  402. Schorling, D.C.; Rost, S.; Lefeber, D.J.; Brady, L.; Muller, C.R.; Korinthenberg, R.; Tarnopolsky, M.; Bonnemann, C.G.; Rodenburg, R.J.; Bugiani, M.; et al. Early and lethal neurodegeneration with myasthenic and myopathic features: A new ALG14-CDG. Neurology 2017, 89, 657–664. [Google Scholar] [CrossRef]
  403. Kvarnung, M.; Taylan, F.; Nilsson, D.; Anderlid, B.M.; Malmgren, H.; Lagerstedt-Robinson, K.; Holmberg, E.; Burstedt, M.; Nordenskjold, M.; Nordgren, A.; et al. Genomic screening in rare disorders: New mutations and phenotypes, highlighting ALG14 as a novel cause of severe intellectual disability. Clin. Genet. 2018, 94, 528–537. [Google Scholar] [CrossRef]
  404. Palombo, F.; Piccolo, B.; Saccani, E.; Fiorini, C.; Capristo, M.; Caporali, L.; Pisani, F.; Carelli, V. A novel ALG14 missense variant in an alive child with myopathy, epilepsy, and progressive cerebral atrophy. Am. J. Med. Genet. A 2021, 185, 1918–1921. [Google Scholar] [CrossRef]
  405. Katata, Y.; Uneoka, S.; Saijyo, N.; Aihara, Y.; Miyazoe, T.; Koyamaishi, S.; Oikawa, Y.; Ito, Y.; Abe, Y.; Numata-Uematsu, Y.; et al. The longest reported sibling survivors of a severe form of congenital myasthenic syndrome with the ALG14 pathogenic variant. Am. J. Med. Genet. A 2022, 188, 1293–1298. [Google Scholar] [CrossRef]
  406. Mensch, A.; Cordts, I.; Scholle, L.; Joshi, P.R.; Kleeberg, K.; Emmer, A.; Beck-Woedl, S.; Park, J.; Haack, T.B.; Stoltenburg-Didinger, G.; et al. GFPT1-Associated Congenital Myasthenic Syndrome Mimicking a Glycogen Storage Disease—Diagnostic Pitfalls in Myopathology Solved by Next-Generation-Sequencing. J. Neuromuscul. Dis. 2022, 9, 533–541. [Google Scholar] [CrossRef]
  407. Siddiqui, S.; Polavarapu, K.; Bardhan, M.; Preethish-Kumar, V.; Joshi, A.; Nashi, S.; Vengalil, S.; Raju, S.; Chawla, T.; Leena, S.; et al. Distinct and Recognisable Muscle MRI Pattern in a Series of Adults Harbouring an Identical GMPPB Gene Mutation. J. Neuromuscul. Dis. 2022, 9, 95–109. [Google Scholar] [CrossRef]
  408. Johnson, K.; Bertoli, M.; Phillips, L.; Topf, A.; Van den Bergh, P.; Vissing, J.; Witting, N.; Nafissi, S.; Jamal-Omidi, S.; Lusakowska, A.; et al. Detection of variants in dystroglycanopathy-associated genes through the application of targeted whole-exome sequencing analysis to a large cohort of patients with unexplained limb-girdle muscle weakness. Skelet. Muscle 2018, 8, 23. [Google Scholar] [CrossRef]
  409. Ehrstedt, C.; Liu, W.W.; Frykholm, C.; Beeson, D.; Punga, A.R. Novel pathogenic ALG2 mutation causing congenital myasthenic syndrome: A case report. Neuromuscul. Disord. 2022, 32, 80–83. [Google Scholar] [CrossRef]
  410. Ma, Y.; Xiong, T.; Lei, G.; Ding, J.; Yang, R.; Li, Z.; Guo, J.; Shen, D. Novel compound heterozygous variants in the GFPT1 gene leading to rare limb-girdle congenital myasthenic syndrome with rimmed vacuoles. Neurol. Sci. 2021, 42, 3485–3490. [Google Scholar] [CrossRef]
  411. Liao, W.; Elfrink, K.; Bahler, M. Head of myosin IX binds calmodulin and moves processively toward the plus-end of actin filaments. J. Biol. Chem. 2010, 285, 24933–24942. [Google Scholar] [CrossRef]
  412. O’Connor, E.; Topf, A.; Muller, J.S.; Cox, D.; Evangelista, T.; Colomer, J.; Abicht, A.; Senderek, J.; Hasselmann, O.; Yaramis, A.; et al. Identification of mutations in the MYO9A gene in patients with congenital myasthenic syndrome. Brain 2016, 139 Pt 8, 2143–2153. [Google Scholar] [CrossRef] [PubMed]
  413. O’Connor, E.; Phan, V.; Cordts, I.; Cairns, G.; Hettwer, S.; Cox, D.; Lochmuller, H.; Roos, A. MYO9A deficiency in motor neurons is associated with reduced neuromuscular agrin secretion. Hum. Mol. Genet. 2018, 27, 1434–1446. [Google Scholar] [CrossRef]
  414. Li, Q.; Gulati, A.; Lemaire, M.; Nottoli, T.; Bale, A.; Tufro, A. Rho-GTPase Activating Protein myosin MYO9A identified as a novel candidate gene for monogenic focal segmental glomerulosclerosis. Kidney Int. 2021, 99, 1102–1117. [Google Scholar] [CrossRef]
  415. Nota, B.; Struys, E.A.; Pop, A.; Jansen, E.E.; Fernandez Ojeda, M.R.; Kanhai, W.A.; Kranendijk, M.; van Dooren, S.J.; Bevova, M.R.; Sistermans, E.A.; et al. Deficiency in SLC25A1, encoding the mitochondrial citrate carrier, causes combined D-2- and L-2-hydroxyglutaric aciduria. Am. J. Hum. Genet. 2013, 92, 627–631. [Google Scholar] [CrossRef] [PubMed]
  416. Kaplan, R.S.; Mayor, J.A.; Wood, D.O. The mitochondrial tricarboxylate transport protein. cDNA cloning, primary structure, and comparison with other mitochondrial transport proteins. J. Biol. Chem. 1993, 268, 13682–13690. [Google Scholar] [CrossRef] [PubMed]
  417. Chaouch, A.; Porcelli, V.; Cox, D.; Edvardson, S.; Scarcia, P.; De Grassi, A.; Pierri, C.L.; Cossins, J.; Laval, S.H.; Griffin, H.; et al. Mutations in the Mitochondrial Citrate Carrier SLC25A1 are Associated with Impaired Neuromuscular Transmission. J. Neuromuscul. Dis. 2014, 1, 75–90. [Google Scholar] [CrossRef]
  418. Al-Futaisi, A.; Ahmad, F.; Al-Kasbi, G.; Al-Thihli, K.; Koul, R.; Al-Maawali, A. Missense mutations in SLC25A1 are associated with congenital myasthenic syndrome type 23. Clin. Genet. 2020, 97, 666–667. [Google Scholar] [CrossRef]
  419. Balaraju, S.; Topf, A.; McMacken, G.; Kumar, V.P.; Pechmann, A.; Roper, H.; Vengalil, S.; Polavarapu, K.; Nashi, S.; Mahajan, N.P.; et al. Congenital myasthenic syndrome with mild intellectual disability caused by a recurrent SLC25A1 variant. Eur. J. Hum. Genet. 2020, 28, 373–377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  420. Li, W.; Zhang, M.; Zhang, L.; Shi, Y.; Zhao, L.; Wu, B.; Li, X.; Zhou, S. A case report of an intermediate phenotype between congenital myasthenic syndrome and D-2- and L-2-hydroxyglutaric aciduria due to novel SLC25A1 variants. BMC Neurol. 2020, 20, 278. [Google Scholar] [CrossRef]
  421. Zhao, Y.; Li, Y.; Bian, Y.; Yao, S.; Liu, P.; Yu, M.; Zhang, W.; Wang, Z.; Yuan, Y. Congenital myasthenic syndrome in China: Genetic and myopathological characterization. Ann. Clin. Transl. Neurol. 2021, 8, 898–907. [Google Scholar] [CrossRef] [PubMed]
  422. Senior, A.; Gerace, L. Integral membrane proteins specific to the inner nuclear membrane and associated with the nuclear lamina. J. Cell Biol. 1988, 107 Pt 1, 2029–2036. [Google Scholar] [CrossRef]
  423. Zhao, C.; Brown, R.S.; Chase, A.R.; Eisele, M.R.; Schlieker, C. Regulation of Torsin ATPases by LAP1 and LULL1. Proc. Natl. Acad. Sci. USA 2013, 110, E1545–E1554. [Google Scholar] [CrossRef]
  424. Kayman-Kurekci, G.; Talim, B.; Korkusuz, P.; Sayar, N.; Sarioglu, T.; Oncel, I.; Sharafi, P.; Gundesli, H.; Balci-Hayta, B.; Purali, N.; et al. Mutation in TOR1AIP1 encoding LAP1B in a form of muscular dystrophy: A novel gene related to nuclear envelopathies. Neuromuscul. Disord. 2014, 24, 624–633. [Google Scholar] [CrossRef] [PubMed]
  425. Dorboz, I.; Coutelier, M.; Bertrand, A.T.; Caberg, J.H.; Elmaleh-Berges, M.; Laine, J.; Stevanin, G.; Bonne, G.; Boespflug-Tanguy, O.; Servais, L. Severe dystonia, cerebellar atrophy, and cardiomyopathy likely caused by a missense mutation in TOR1AIP1. Orphanet J. Rare Dis. 2014, 9, 174. [Google Scholar] [CrossRef] [PubMed]
  426. Ghaoui, R.; Benavides, T.; Lek, M.; Waddell, L.B.; Kaur, S.; North, K.N.; MacArthur, D.G.; Clarke, N.F.; Cooper, S.T. TOR1AIP1 as a cause of cardiac failure and recessive limb-girdle muscular dystrophy. Neuromuscul. Disord. 2016, 26, 500–503. [Google Scholar] [CrossRef]
  427. Cossins, J.; Webster, R.; Maxwell, S.; Rodriguez Cruz, P.M.; Knight, R.; Llewelyn, J.G.; Shin, J.Y.; Palace, J.; Beeson, D. Congenital myasthenic syndrome due to a TOR1AIP1 mutation: A new disease pathway for impaired synaptic transmission. Brain Commun. 2020, 2, fcaa174. [Google Scholar] [CrossRef] [PubMed]
  428. Malfatti, E.; Catchpool, T.; Nouioua, S.; Sihem, H.; Fournier, E.; Carlier, R.Y.; Cardone, N.; Davis, M.R.; Laing, N.G.; Sternberg, D.; et al. A TOR1AIP1 variant segregating with an early onset limb girdle myasthenia-Support for the role of LAP1 in NMJ function and disease. Neuropathol. Appl. Neurobiol. 2022, 48, e12743. [Google Scholar] [CrossRef]
  429. Sakamoto, I.; Kishida, S.; Fukui, A.; Kishida, M.; Yamamoto, H.; Hino, S.; Michiue, T.; Takada, S.; Asashima, M.; Kikuchi, A. A novel beta-catenin-binding protein inhibits beta-catenin-dependent Tcf activation and axis formation. J. Biol. Chem. 2000, 275, 32871–32878. [Google Scholar] [CrossRef] [Green Version]
  430. Thompson, B.A.; Tremblay, V.; Lin, G.; Bochar, D.A. CHD8 is an ATP-dependent chromatin remodeling factor that regulates beta-catenin target genes. Mol. Cell Biol. 2008, 28, 3894–3904. [Google Scholar] [CrossRef]
  431. Nishiyama, M.; Skoultchi, A.I.; Nakayama, K.I. Histone H1 recruitment by CHD8 is essential for suppression of the Wnt-beta-catenin signaling pathway. Mol. Cell Biol. 2012, 32, 501–512. [Google Scholar] [CrossRef]
  432. Lee, C.Y.; Petkova, M.; Morales-Gonzalez, S.; Gimber, N.; Schmoranzer, J.; Meisel, A.; Bohmerle, W.; Stenzel, W.; Schuelke, M.; Schwarz, J.M. A spontaneous missense mutation in the chromodomain helicase DNA-binding protein 8 (CHD8) gene: A novel association with congenital myasthenic syndrome. Neuropathol. Appl. Neurobiol. 2020, 46, 588–601. [Google Scholar] [CrossRef] [PubMed]
  433. Li, X.M.; Dong, X.P.; Luo, S.W.; Zhang, B.; Lee, D.H.; Ting, A.K.; Neiswender, H.; Kim, C.H.; Carpenter-Hyland, E.; Gao, T.M.; et al. Retrograde regulation of motoneuron differentiation by muscle beta-catenin. Nat. Neurosci. 2008, 11, 262–268. [Google Scholar] [CrossRef]
  434. Latcheva, N.K.; Delaney, T.L.; Viveiros, J.M.; Smith, R.A.; Bernard, K.M.; Harsin, B.; Marenda, D.R.; Liebl, F.L.W. The CHD Protein, Kismet, is Important for the Recycling of Synaptic Vesicles during Endocytosis. Sci. Rep. 2019, 9, 19368. [Google Scholar] [CrossRef]
  435. McDiarmid, T.A.; Belmadani, M.; Liang, J.; Meili, F.; Mathews, E.A.; Mullen, G.P.; Hendi, A.; Wong, W.R.; Rand, J.B.; Mizumoto, K.; et al. Systematic phenomics analysis of autism-associated genes reveals parallel networks underlying reversible impairments in habituation. Proc. Natl. Acad. Sci. USA 2020, 117, 656–667. [Google Scholar] [CrossRef]
  436. Bernier, R.; Golzio, C.; Xiong, B.; Stessman, H.A.; Coe, B.P.; Penn, O.; Witherspoon, K.; Gerdts, J.; Baker, C.; Vulto-van Silfhout, A.T.; et al. Disruptive CHD8 mutations define a subtype of autism early in development. Cell 2014, 158, 263–276. [Google Scholar] [CrossRef] [PubMed]
  437. Hoffmann, A.; Spengler, D. Chromatin Remodeler CHD8 in Autism and Brain Development. J. Clin. Med. 2021, 10, 366. [Google Scholar] [CrossRef]
  438. Lalani, S.R.; Zhang, J.; Schaaf, C.P.; Brown, C.W.; Magoulas, P.; Tsai, A.C.; El-Gharbawy, A.; Wierenga, K.J.; Bartholomew, D.; Fong, C.T.; et al. Mutations in PURA cause profound neonatal hypotonia, seizures, and encephalopathy in 5q31.3 microdeletion syndrome. Am. J. Hum. Genet. 2014, 95, 579–583. [Google Scholar] [CrossRef]
  439. Hunt, D.; Leventer, R.J.; Simons, C.; Taft, R.; Swoboda, K.J.; Gawne-Cain, M.; the DDD Study; Magee, A.C.; Turnpenny, P.D.; Baralle, D. Whole exome sequencing in family trios reveals de novo mutations in PURA as a cause of severe neurodevelopmental delay and learning disability. J. Med. Genet. 2014, 51, 806–813. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  440. Tanaka, A.J.; Bai, R.; Cho, M.T.; Anyane-Yeboa, K.; Ahimaz, P.; Wilson, A.L.; Kendall, F.; Hay, B.; Moss, T.; Nardini, M.; et al. De novo mutations in PURA are associated with hypotonia and developmental delay. Cold Spring Harb. Mol. Case Stud. 2015, 1, a000356. [Google Scholar] [CrossRef]
  441. Reijnders, M.R.F.; Janowski, R.; Alvi, M.; Self, J.E.; van Essen, T.J.; Vreeburg, M.; Rouhl, R.P.W.; Stevens, S.J.C.; Stegmann, A.P.A.; Schieving, J.; et al. PURA syndrome: Clinical delineation and genotype-phenotype study in 32 individuals with review of published literature. J. Med. Genet. 2018, 55, 104–113. [Google Scholar] [CrossRef]
  442. Wyrebek, R.; DiBartolomeo, M.; Brooks, S.; Geller, T.; Crenshaw, M.; Iyadurai, S. Hypotonic infant with PURA syndrome-related channelopathy successfully treated with pyridostigmine. Neuromuscul. Disord. 2022, 32, 166–169. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Thirty-five genes (red letters) causing CMS.
Figure 1. Thirty-five genes (red letters) causing CMS.
Ijms 24 03730 g001
Figure 2. Crystal structure of AChR viewed from the extracellular side (PDB 2BG9) [87]. (A) Extracellular domains of AChR subunits. Other domains are shown in gray. (B) Transmembrane domains of AChR subunits. αM4 domains are not indicated.
Figure 2. Crystal structure of AChR viewed from the extracellular side (PDB 2BG9) [87]. (A) Extracellular domains of AChR subunits. Other domains are shown in gray. (B) Transmembrane domains of AChR subunits. αM4 domains are not indicated.
Ijms 24 03730 g002
Figure 3. Representative molecules at the nerve terminal and the agrin-LPR4-MuSK signaling pathway to induce AChR clustering. Interactions between binding domains are indicated by double headed arrows [71,98,100,101,102]. Broken arrows in the muscle indicate that the exact signaling molecules are not shown. Diseases other than CMS and toxins affecting the NMJ are indicated in red letters. βCAT, β-catenin; BPD, β-propeller domain; C6, six-cysteine-box; Ctgf, connective tissue growth factor; Fz-CRD, frizzled-like cysteine-rich domain; Ig, immunoglobulin-like domain; LDLR-A, low-density lipoprotein receptor class A repeat; Lgr5, leucine-rich repeat-containing G-protein coupled receptor 5; and Rspo2, R-spondin 2.
Figure 3. Representative molecules at the nerve terminal and the agrin-LPR4-MuSK signaling pathway to induce AChR clustering. Interactions between binding domains are indicated by double headed arrows [71,98,100,101,102]. Broken arrows in the muscle indicate that the exact signaling molecules are not shown. Diseases other than CMS and toxins affecting the NMJ are indicated in red letters. βCAT, β-catenin; BPD, β-propeller domain; C6, six-cysteine-box; Ctgf, connective tissue growth factor; Fz-CRD, frizzled-like cysteine-rich domain; Ig, immunoglobulin-like domain; LDLR-A, low-density lipoprotein receptor class A repeat; Lgr5, leucine-rich repeat-containing G-protein coupled receptor 5; and Rspo2, R-spondin 2.
Ijms 24 03730 g003
Figure 4. Hexosamine biosynthesis pathway to make UDP-GlcNAc and N-glycosylation pathway. Genes causing CMS are shown in red. Figure 4 is shown as an inset in Figure 1.
Figure 4. Hexosamine biosynthesis pathway to make UDP-GlcNAc and N-glycosylation pathway. Genes causing CMS are shown in red. Figure 4 is shown as an inset in Figure 1.
Ijms 24 03730 g004
Table 1. Electrophysiological features and therapies of congenital myasthenic syndromes.
Table 1. Electrophysiological features and therapies of congenital myasthenic syndromes.
SectionPhenotypeGeneOMIM# aInheritanceLow-Frequency RNSHigh-Frequency RNSTreatment
ChEIsEphedrineSalbutamol (albuterol)AmifampridineQuinidineFluoxetineAcetazolamide
4.1Endplate AChR deficiencyCHRNA1 - bARdecrement effectiveeffectiveeffectiveeffective
4.1Endplate AChR deficiencyCHRNB1CMS2C- bARdecrement effectiveeffectiveeffectiveeffective
4.1Endplate AChR deficiencyCHRNDCMS3C- bARdecrement effectiveeffectiveeffectiveeffective
4.1Endplate AChR deficiencyCHRNECMS4C- bARdecrement effectiveeffectiveeffectiveeffective
4.1Endplate AChR deficiencyRAPSNCMS11[38]ARdecrement effectiveeffectiveeffectiveeffective
4.2Escobar syndromeCHRNG 101ARdecrement
4.2FADSCHRNA1 (4)ARdecrement
4.2FADSCHRND (6)ARdecrement
4.2FADSMUSK (6)ARdecrement
4.2FADSRAPSN (8)ARdecrement
4.2FADSDOK7 (2)ARdecrement
4.2FADSSLC18A3 (1)ARdecrement
4.3SCCMSCHRNA1CMS1A(14)ADdecrement, repetitive CMAP mostly ineffectiveeffective in some reportseffective in some reports effectiveeffective
4.3SCCMSCHRNB1CMS2A(5)ADdecrement, repetitive CMAP mostly ineffectiveeffective in some reportseffective in some reports effectiveeffective
4.3SCCMSCHRNDCMS3A(4)ADdecrement, repetitive CMAP mostly ineffectiveeffective in some reportseffective in some reports effectiveeffective
4.3SCCMSCHRNECMS4A(11)AD/ARdecrement, repetitive CMAP mostly ineffectiveeffective in some reportseffective in some reports effectiveeffective
4.3FCCMSCHRNA1CMS1B(3) cARdecrement effectivepresumably effective, but no reporteffective in a reporteffective
4.3FCCMSCHRNB1CMS2B(1) cARdecrement effectivepresumably effective, but no reporteffective in a reporteffective
4.3FCCMSCHRNDCMS3B(1) cARdecrement effectivepresumably effective, but no reporteffective in a reporteffective
4.3FCCMSCHRNECMS4B(6) cARdecrement effectivepresumably effective, but no reporteffective in a reporteffective
4.4Endplate AChE deficiencyCOLQCMS5[30]ARdecrement, repetitive CMAP contraindication, but effective in some reported patientseffective in some reportseffective in some reports effective in a report
4.4Synaptic CMSLAMB2 1ARdecrement contraindicationeffective
4.4Synaptic CMSCOL13A1CMS1941ARdecrement ineffective effectiveeffective
4.5Sodium channel CMSSCN4ACMS166ARno decrementdecrementEffective, ineffective, or marked adverse effects Slightly effective effective or ineffective
4.6CMS caused by defective AChR clusteringAGRNCMS8[13]ARdecrement Ineffective or mildly effectiveeffectiveeffectiveineffective or slightly effective
4.6CMS caused by defective AChR clusteringMUSKCMS9[15]ARdecrement Ineffective or worsened effectiveeffective
4.6CMS caused by defective AChR clusteringLRP4CMS171ARdecrement worsened
4.6CMS caused by defective AChR clusteringDOK7CMS10[34]ARdecrement Combination of ineffective and worseningeffectiveeffectiveeffective in some reports effective in some reports
4.7CMS caused by defective structural moleculesPLEC 22ARdecrement effective or ineffective effective in some reportseffective or ineffective
4.8CMS caused by defective recycling of AChCHATCMS6[19]ARno decrementdecrement in some patientseffective effective
4.8CMS caused by defective recycling of AChSLC18A3CMS217ARdecrement at rest or only after isometric muscle contractiondecrement in some patientseffectiveeffective effective
4.8CMS caused by defective recycling of AChSLC5A7CMS2012ARdecrement at rest or only after isometric muscle contractiondecrement in some patientseffectiveeffective ineffective
4.8CMS caused by defective recycling of AChPREPLCMS2218ARdecrementdecrement in some patientseffective
4.9LEMS-like CMSSYT2CMS7ACMS7B2AD/ARdecrementincrementeffective effectiveeffective
4.9LEMS-like CMSSNAP25CMS182ADdecrement ineffective effective
4.9LEMS-like CMSUNC13A 1ARdecrementincrementminimally effective minimally effective
4.9LEMS-like CMSVAMP1CMS259ARdecrementincrementeffective
4.9LEMS-like CMSRPH3A 1ARno decrementincrement effective
4.9LEMS-like CMSLAMA5 1ARdecrementincrementeffective effective
4.10Glycosylation-deficient CMSGFPT1CMS12[17]ARdecrement effective effective
4.10Glycosylation-deficient CMSDPAGT1CMS13[5]ARdecrement effective effectiveeffective
4.10Glycosylation-deficient CMSALG2CMS149ARdecrement effective or ineffective effectiveeffective
4.10Glycosylation-deficient CMSALG14CMS1512ARdecrement effective
4.10Glycosylation-deficient CMSGMPPB [9]ARdecrement effective effective
4.11CMS caused by defective nerve terminal formationMYO9ACMS243ARdecrement effective effective
4.11CMS caused by defective nerve terminal formationSLC25A1CMS2319ARdecrement ineffective in most patients ineffective in most patients
4.12CMS caused by defective nuclear membrane proteinTOR1AIP1 5ARdecrement effective no additional effect
4.13CMS caused by defective chromatin remodeling proteinCHD8 2ARdecrement ineffective ineffectivemarkedly effective
4.14CMS in PURA syndromePURA 3ADdecrement, repetitive CMAP effective in a patient, but not in another patient effective in a patient
a The number of original reports is shown in square brackets, and the number of pathogenic variants in round brackets. Otherwise, the number of patients is shown. b Differentiation of endplate AChR deficiency and FCCMS requires detailed electrophysiological studies using either intracellular recordings or patch-clamp recordings of biopsied patient’s neuromuscular junction, or patch-clamp recordings of mutant AChRs expressed in culture cells, but most variants are not characterized as such. Thus, the numbers of patients, original articles, pathogenic variants of endplate AChR deficiency were not counted. c For FCCMS, the number of pathogenic variants with electrophysiological analyses was counted.
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

Ohno, K.; Ohkawara, B.; Shen, X.-M.; Selcen, D.; Engel, A.G. Clinical and Pathologic Features of Congenital Myasthenic Syndromes Caused by 35 Genes—A Comprehensive Review. Int. J. Mol. Sci. 2023, 24, 3730. https://doi.org/10.3390/ijms24043730

AMA Style

Ohno K, Ohkawara B, Shen X-M, Selcen D, Engel AG. Clinical and Pathologic Features of Congenital Myasthenic Syndromes Caused by 35 Genes—A Comprehensive Review. International Journal of Molecular Sciences. 2023; 24(4):3730. https://doi.org/10.3390/ijms24043730

Chicago/Turabian Style

Ohno, Kinji, Bisei Ohkawara, Xin-Ming Shen, Duygu Selcen, and Andrew G. Engel. 2023. "Clinical and Pathologic Features of Congenital Myasthenic Syndromes Caused by 35 Genes—A Comprehensive Review" International Journal of Molecular Sciences 24, no. 4: 3730. https://doi.org/10.3390/ijms24043730

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