Entry - *118425 - CHLORIDE CHANNEL 1, SKELETAL MUSCLE; CLCN1 - OMIM
* 118425

CHLORIDE CHANNEL 1, SKELETAL MUSCLE; CLCN1


Alternative titles; symbols

CHLORIDE CHANNEL, MUSCLE; CLC1


HGNC Approved Gene Symbol: CLCN1

Cytogenetic location: 7q34     Genomic coordinates (GRCh38): 7:143,316,111-143,352,083 (from NCBI)


Gene-Phenotype Relationships
Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
7q34 Myotonia congenita, dominant 160800 AD 3
Myotonia congenita, recessive 255700 AR 3
Myotonia levior 160800 AD 3

TEXT

Description

The muscle chloride channel CLCN1 regulates the electric excitability of the skeletal muscle membrane. Skeletal muscle has an unusually high resting Cl(-) conductance and in vitro studies suggest that reduction of this conductance causes electrical instability and resulting myotonia in both humans and animal models. Muscle Cl(-) conductance is predominantly mediated by the CLCN1 chloride channel (summary by Steinmeyer et al., 1994).


Cloning and Expression

By homology screening with the major rat skeletal muscle chloride channel CLCN1, Koch et al. (1992) cloned a partial human CLCN1 cDNA that covered about 80% of the coding sequence. This region was 88% identical to the rat channel in amino acid sequence.

Chen et al. (2013) found expression of the CLCN1 gene in various human brain regions, including cerebellum, hippocampus, spinal cord, and cerebral cortex, as well as in heart. Clcn1 was also expressed in the brain and heart of developing and adult mice. In mouse brain, neuronal expression of Clcn1 was detected in the pyramidal and dentate granule cells of the hippocampus, cerebellar Purkinje cells, scattered brainstem nuclei, frontal neocortex, and thalamus. The results suggested that CLCN1 has a role in neuronal function and excitability in addition to its known role in skeletal muscle.


Gene Function

Steinmeyer et al. (1994) determined that CLCN1 functions as a homooligomer, most likely with 4 subunits.

Pusch et al. (1995) used a Xenopus transfection to demonstrate shifting of the gating of CLCN1 toward positive voltages by 4 different mutations identified in patients with myotonia congenita (160800). When these mutant cDNAs were coexpressed with wildtype subunits, they imposed altered voltage dependence on the heteromeric channels which would then open only in a voltage range where they could not contribute significantly to the repolarization of action potentials. Without such repolarizations, sodium channels have enough time to recover from inactivation leading to typical myotonic runs, which are a series of repetitive action potentials.


Biochemical Features

Cryoelectron Microscopy

Mindell et al. (2001) reported the formation of 2-dimensional crystals of the prokaryotic CLC channel homolog EriC reconstituted into phospholipid bilayer membranes. Cryoelectron microscopic analysis of these crystals yielded a projection structure at 6.5-angstrom resolution that showed off-axis water-filled pores within the dimeric channel complex.

Crystal Structure

Dutzler et al. (2002) presented the x-ray structures of 2 prokaryotic CLC chloride channels, from Salmonella typhimurium and E. coli, at 3.0 and 3.5 angstroms, respectively. Both structures revealed 2 identical pores, each pore being formed by a separate subunit contained within a homodimeric membrane protein.

Using the 3-dimensional crystal structure of a bacterial CLC protein to predict residues involved in chloride channel inhibitor binding, Estevez et al. (2003) identified an inhibitor-binding site in human CLCN1. The binding site is localized close to the chloride-binding site and is accessible only from the intracellular side. Estevez et al. (2003) concluded that the structures of bacterial CLCs can be extrapolated with fidelity to mammalian chloride channels.


Gene Structure

Lorenz et al. (1994) showed that the protein coding sequence of the CLCN1 gene is organized into 23 exons. Its upstream region contains a canonical TATA box, several consensus binding sites for myogenic transcription factors, and 2 other putative regulatory elements.


Mapping

By blot hybridization to a panel of chromosome 7-specific, human-mouse somatic cell hybrids, Koch et al. (1992) mapped the CLCN1 gene to 7q32-qter. With RFLPs in the CLCN1 gene, they demonstrated that the locus is linked to the T-cell receptor beta locus (see 186930) at 7q35 (maximum lod = 5.23 at theta = 0.0). In the mouse, it had previously been demonstrated that the corresponding loci are linked on chromosome 6, which shows other evidence of homology of synteny to human 7q (Steinmeyer et al., 1991).


Molecular Genetics

Myotonia Congenita

In 3 brothers, born of consanguineous parents, with autosomal recessive myotonia congenita (Becker disease; 255700), Koch et al. (1992) identified a homozygous mutation in the CLCN1 gene (F413C; 118425.0001). In affected members of 3 unrelated families with autosomal dominant myotonia congenita (Thomsen disease; 160800), George et al. (1993) identified a heterozygous mutation in the CLCN1 gene (G230E; 118425.0002). The findings indicated that the 2 disorders are allelic.

Meyer-Kleine et al. (1995) identified 15 different mutations in the CLCN1 gene, of which 10 were novel, in a total of 17 unrelated families and 13 single patients with Becker-type myotonia congenita. One additional family had the dominant Thomsen form. Three mutations accounted for 32% of the Becker chromosomes in the German population; F413C, R894X (118425.0015), and a 14-bp deletion in exon 13 (118425.0009). Although a 437A-T transversion had been described as a disease-causing mutation (Koty et al., 1994), Meyer-Kleine et al. (1995) observed it in 3 myotonia families and in 5 of 200 control chromosomes, consistent with a polymorphism. Moreover, mutant 437A-T cRNA was functionally expressed in Xenopus oocytes and found to induce currents that were indistinguishable from wildtype currents.

In a screening of 6 unrelated patients with recessive Becker-type myotonia, Mailander et al. (1996) identified 4 novel CLCN1 mutations and a previously reported 14-bp deletion. Five patients were homozygous and the sixth patient was compound heterozygous. Heterozygous carriers of the Becker mutation did not display any clinical symptoms of myotonia; however, all heterozygous males, but none of the heterozygous females, exhibited myotonic discharges on EMG, suggesting a gene-dosage effect of the mutations on chloride conductance and a male predominance of subclinical myotonia.

Wollnik et al. (1997) analyzed the effect of 1 dominant and 3 recessive mutations in the CLCN1 gene after functional expression in Xenopus oocytes.

Esteban et al. (1998) stated that 31 specific mutations in the CLCN1 gene had been related to myotonia congenita in humans and 3 in mice. They described a homozygous arg317-to-gln (R317Q; 118425.0011) mutation in affected members of a family with autosomal recessive myotonia congenita. A heterozygous brother had mild muscle stiffness consistent with latent myotonia. The parents were unaffected, although only the mother, who was heterozygous, was available for DNA study. The same R317Q mutation had previously been identified in a family with dominant Thomsen myotonia congenita. At least 2 other mutations, G230E (118425.0002) and R894X, had been found in both dominant and recessive myotonia congenita, depending on the particular family.

Kubisch et al. (1998) identified 4 novel missense mutations in the CLCN gene, 2 in the dominant form and 2 in the recessive form of myotonia. The first 2 displayed a classic dominant phenotype with a dominant-negative effect by significantly imparting a voltage shift on mutant/wildtype heteromeric channels as found in heterozygous patients. One of the recessive mutations also shifted the voltage dependence to positive values, but coexpression with wildtype CLCN1 gave almost wildtype currents. The voltage dependence of mutant heteromeric channels was not always intermediate between those of the constituent homomeric channel subunits. These complex interactions correlated clinically with various inheritance patterns, ranging from autosomal dominant with various degrees of penetrance to autosomal recessive.

In affected members of 18 unrelated families from Norway and Sweden with both autosomal dominant (5 families) and autosomal recessive (13 families) inheritance of myotonia congenita, Sun et al. (2001) identified 8 different mutations (1 nonsense, 4 missense, and 3 splice mutations) in the CLCN1 gene; 3 mutations were novel. Fifteen patients had mutations on both alleles, consistent with the recessive disorder; 2 probands had mutations in a single allele; and 2 probands had no CLCN1 mutations. In 2 families, 3 CLCN1 mutations were found in the proband, and Sun et al. (2001) suspected that this phenomenon may be underestimated because mutation search in a disease gene usually ends by the identification of 2 mutations in a family with recessive inheritance. Families with apparently dominant segregation of myotonia congenita may actually represent recessive inheritance with undetected heterozygous individuals married-in as a consequence of a high population carrier frequency of some mutations. The findings, together with the very variable clinical presentation, challenged the classification into dominant Thomsen or recessive Becker disease. Sun et al. (2001) suggested that most cases of myotonia congenita show recessive inheritance with some modifying factors or genetic heterogeneity.

In a review, Pusch (2002) stated that more than 60 CLCN1 mutations had been identified as causing myotonia, with only a few of them being dominant. A dominant-negative effect of mutant subunits in mutant-wildtype heterodimers was suggested as the usual mechanism for dominant mutations.

Duno et al. (2004) reported 4 unrelated families with myotonia congenita and the R894X mutation: 2 families had a single mutant allele, showing dominant inheritance, and 2 families were compound heterozygous for R894X and another mutation, showing recessive inheritance. RT-PCR did not reveal any association between total CLCN1 mRNA in muscle and the mode of inheritance, but the dominant family with the most severe phenotype expressed twice the expected amount of the R894X mRNA allele, even compared to the recessive families. Duno et al. (2004) suggested that variation in allelic expression may be a modifier of disease expression and progression in myotonia congenita.

Raja Rayan et al. (2012) performed multiplex ligation-dependent probe amplification (MLPA) specific to the CLCN1 gene in 60 families with recessive myotonia congenita in whom either no mutations or only a single pathogenic CLCN1 mutation had been identified. The results were positive in 4 (6.7%) patients: 2 unrelated patients were found to have 2 different multiexon deletions within the CLCN1 gene on the second allele, and 2 additional patients had a homozygous duplication of exons 8 through 14 of the CLCN1 gene (118425.0020). The 2 patients with the duplication were both of Iraqi origin, but were unrelated. Both Iraqi patients had a severe form of the disorder with onset in infancy. Haplotype analysis suggested a founder effect for this duplication mutation. Raja Rayan et al. (2012) concluded that copy number variation involving the CLCN1 gene is an important genetic mechanism in patients with recessive myotonia congenita, and that MLPA analysis may aid in genetic counseling.

In 4 Costa Rican families with myotonia congenita, Vindas-Smith et al. (2016) identified mutations in the CLCN1 gene by bidirectional sequencing of the CLCN1 gene, with confirmation by RFLP-PCR. In 2 families in which the proband had Thomsen disease (families 1 and 4), heterozygous mutations were identified (F167L; Q412P, 118425.0022). In another family in which the proband had Thomsen disease (family 2), compound heterozygous mutations were identified (R105C and F167L). In a proband (family 3) with a myositis-like disorder, a variant of unknown significance (Q154R) was identified. Functional studies of CLCN1 with the F167L, R105C, or Q154R mutation did not show alterations of gating parameters or channel conductance. Functional studies of CLCN1 with the Q412P mutation expressed in Xenopus oocytes showed reduced surface expression and reduced current density. Vindas-Smith et al. (2016) concluded that the Q412P mutation induces a severe folding defect that leads to its degradation before it can dimerize with the wildtype subunit.

Altamura et al. (2018) evaluated the functional significance of 7 mutations in the C-terminal region of the CLCN1 gene associated with either autosomal dominant Thomsen disease or autosomal recessive Becker disease. CLCN1 with each mutation was transfected into HEK293 cells and analyzed with patch-clamp analysis. Five of the mutations were in the CBS2 domain (V829M, T832I, V851M, G859V, L861P) and 2 of the mutations were in the C-terminal peptide (P883T, V947E). Mutations located between residues 829 and 835 and in residue 883 resulted in alteration of voltage dependence. Mutations between residues 851 and 859 and in residue 947 resulted in a reduction of chloride currents. The results were consistent with a role for CBS2 in protein channel gating and demonstrated the importance of the C-peptide region in protein function and expression.

Suetterlin et al. (2022) assessed the function of 95 CLCN1 mutations, including 34 novel mutations, identified in 233 patients with myotonia congenita. Mutations in CLCN1 were assessed by transfection of cDNA with each mutation into Xenopus laevis oocytes and analyzed with 2-electrode voltage clamp assays. From a functional standpoint, mutations that altered voltage dependence of activation clustered in the first half of the transmembrane domains and mutations resulting in absent currents clustered in the second half of the transmembrane domain. In terms of assessment of clinical significance and inheritance patterns, mutations that resulted in dominant functional features clustered in the TM1 domain, and variants associated with recessive functional features and without a shift in voltage dependence of activation were clustered in the TM2 domain. Mutations in the intracellular domain were not associated with a dominant inheritance pattern. Suetterlin et al. (2022) concluded that functional characterization of CLCN1 mutations improves the assessment of their clinical implications.

Myotonic Dystrophy

Myotonic dystrophy (DM1; 160900) is caused by a trinucleotide repeat expansion of the DMPK gene (605377.0001). In DM, expression of RNAs that contain expanded CUG or CCUG repeats is associated with degeneration and repetitive action potentials (myotonia) in skeletal muscle. Using skeletal muscle from a transgenic mouse model of DM, Mankodi et al. (2002) showed that expression of expanded CUG repeats reduced the transmembrane chloride conductance to levels well below those expected to cause myotonia. The expanded CUG repeats triggered aberrant splicing of pre-mRNA for CLCN1, resulting in loss of the CLCN1 protein from the surface membrane. Mankodi et al. (2002) identified a similar defect in CLCN1 splicing and expression in DM1 and DM2 (602668). The authors proposed that a transdominant effect of mutant RNA on CLCN1 RNA processing leads to chloride channelopathy and membrane hyperexcitability in DM.

Charlet-B et al. (2002) demonstrated loss of CLCN1 mRNA and protein in DM1 skeletal muscle tissue due to aberrant splicing of the CLCN1 pre-mRNA. They showed that the splicing regulator, CUG-binding protein (CUGBP; 601074), which is elevated in DM1 striated muscle, bound to the CLCN1 pre-mRNA, and that overexpression of CUGBP in normal cells reproduced the aberrant pattern of CLCN1 splicing observed in DM1 skeletal muscle. Charlet-B et al. (2002) proposed that disruption of alternative splicing regulation causes a predominant pathologic feature of DM1.

Associations Pending Confirmation

See 118425.0021 for discussion of a possible association between variation in the CLCN1 gene and idiopathic generalized epilepsy (EIG; see 600669).


Genotype/Phenotype Correlations

Aminoff et al. (1977) found that a subset of patients with myotonia congenita showed a marked decrement of the compound motor action potential (CMAP) with repeated stimulation compared to controls. Although the presence of decrement was not related to the degree of myotonia, it was usually observed in patients with autosomal recessive disease. Colding-Jorgensen et al. (2003) found that all 6 patients with the dominant P480L (118425.0006) mutation had CMAP decrements above 30%. In patients with the R894X (118425.0015) mutation, some had a large decrement, some had a slight decrement, and some had no change. Two patients with 2 CLCN1 mutations, 1 of whom carried an R894X mutation, had large decrements above 80%. Presence of decrement did not correlate with disease severity. Colding-Jorgensen et al. (2003) concluded that CMAP decrement may occur in dominant myotonia congenita and likely reflects the degree of chloride conduction reduction caused by particular mutations.


Animal Model

The Adr ('arrested development of righting') mouse, a model of autosomal recessive myotonia congenita, was found to be due to a mutation on mouse chromosome 6 (Rudel, 1990) in a region of the genome with homology of synteny to human chromosome 7q31-q35. The locus in the mouse is between those for Tcrb and Hox1. Steinmeyer et al. (1991) found that the Adr mouse phenotype was caused by a transposon insertion mutation that altered and inactivated the muscle-membrane chloride channel gene. The findings confirmed that the Adr mouse is an authentic model of Becker disease in the human.

Wu and Olson (2002) determined that Adr mice, which have an inactivated Clcn1 gene, showed an increased number of oxidative fibers, lacked glycolytic fibers, and showed muscle hypertrophy, similar to patients with Becker syndrome. By breeding Clcn1-null mice with mice harboring an Mef2 (600660)-dependent reporter gene, they found that the transcriptional activity of Mef2 was dramatically enhanced in myotonic muscle. Induction of Mef2 did not correlate with enhanced DNA-binding activity, but did correlate with activation of p38 Mapk (see 600289), an activator of Mef2, and with reduced expression of class II histone deacetylases (HDACs; see 605314), which repress Mef2 activity. Wu and Olson (2002) hypothesized that mutations in Clcn1 alter intracellular calcium levels, leading to the combined effects of class II Hdac deficiency and p38 Mapk activation, resulting in upregulation of Mef2 transcriptional activity followed by long-term changes in gene expression and to fiber-type transformation.

Beck et al. (1996) noted that the current hypotheses regarding the pathophysiology of autosomal dominant myotonia congenita, or Thomsen disease, were initially formulated from studies of the myotonic goat, an unusual breed afflicted with severe autosomal dominant congenital myotonia that closely resembles the human disease clinically and in its mode of inheritance. These animals are often referred to as 'fainting,' 'nervous,' 'stiff-legged,' or 'epileptic' goats because of their tendency to develop severe acute muscle stiffness and become immobile and often fall when attempting to make sudden forceful movements or when startled. The pathogenesis of myotonia in the goat was elucidated by Bryant and colleagues (Bryant, 1962, Lipicky and Bryant, 1966) who first described a severely diminished resting chloride conductance in muscle fibers from affected animals. The same group (Adrian and Bryant, 1974) also demonstrated that myotonia could be produced in normal skeletal muscle fibers bathed in a chloride-free solution. Beck et al. (1996) demonstrated the molecular basis for the decreased muscle chloride conductance in this historically important animal model. They found a single nucleotide change (GCC to CCC), resulting in an ala885-to-pro (A885P) substitution in a conserved residue in the C terminus of the goat muscle chloride channel, 104 residues from the termination codon. Heterologous expression of the mutation demonstrated a substantial (+47 mV) shift in the midpoint of steady-state activation of the channel, resulting in a diminished channel open probability at voltages near the resting membrane potential of skeletal muscle.


ALLELIC VARIANTS ( 22 Selected Examples):

.0001 MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, PHE413CYS
  
RCV000019083...

In 3 brothers with autosomal recessive generalized myotonia congenita (Becker disease; 255700), born of consanguineous parents, Koch et al. (1992) identified a homozygous T-to-G transversion in the CLCN1 gene, resulting in a phe413-to-cys (F413C) substitution toward the end of putative membrane span D8. This residue lies in a highly conserved region of the channel protein that is predicted to form a membrane-spanning alpha helix and possibly a component of the permeation pore (George et al., 1993).

Koch et al. (1993) found the F413C missense mutation in 15% of chromosomes carrying a gene for recessive myotonia congenita.


.0002 MYOTONIA CONGENITA, AUTOSOMAL DOMINANT

CLCN1, GLY230GLU
  
RCV000019084...

In affected members of 3 unrelated families with autosomal dominant myotonia congenita (Thomsen disease; 160800), George et al. (1993) identified a G-to-A transition in the CLCN1 gene, resulting in a gly230-to-glu (G230E) substitution between the third and fourth predicted membrane-spanning segments. This glycine residue is conserved in all known members of this class of chloride channel proteins. The codon number used for this mutation was based on the available partial-length cDNA of CLCN1; when the full-length cDNA became known, the codon number was changed from 180 to 230 (George, 1997).

In functional expression studies, Fahlke et al. (1997) found that the G230E mutation caused substantial changes in anion and cation selectivity, as well as a fundamental change in rectification of the current-voltage relationship. Whereas wildtype channels were characterized by pronounced inward rectification and a characteristic pattern of selectivity, G230E exhibited outward rectification at positive potentials and a different pattern of selectivity. Furthermore, the cation-to-anion permeability ratio of the mutant was much greater than that of the wildtype channel.


.0003 MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, IVSDS, G-A, +1
  
RCV000019086

In affected members of a German family with recessive myotonia (255700), Lorenz et al. (1994) identified compound heterozygosity for 2 mutations in the CLCN1 gene: a 979G-A transition in a splice consensus site at the end of exon 8, and a 1488G-T transversion in exon 14, resulting in an arg496-to-ser (R496S; 118425.0004). Functional expression of R496S cRNA in Xenopus oocytes yielded no detectable currents. Furthermore, it did not suppress wildtype currents in coexpression assay, confirming it as a recessive mutation. The G-to-A transition in exon 8 was stated to affect the last nucleotide of the exon. If this interfered with mRNA splicing at that exon/intron boundary, the translation product would be terminated by a stop codon after 51 additional amino acids or other splice sites in the intron might be used. Alternatively, if splicing were normal, this mutation would lead to a substitution of isoleucine for valine at position 327 (V327I). Since this residue is not conserved among the members of this gene family and most members have negatively charged glutamate residues at this position, it is unlikely that such a substitution would have a dramatic effect on channel function. This would argue for an aberrant splicing as the effect of the G979A mutation.


.0004 MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, ARG496SER
  
RCV000019087...

For discussion of the arg496-to-ser (R496S) mutation in the CLCN1 gene that was found in compound heterozygous state in affected members of a family with recessive myotonia (255700) by Lorenz et al. (1994), see 118425.0003.


.0005 MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, GLY482ARG
  
RCV000019088...

In affected members of a family with Becker myotonia congenita (255700), Koch et al. (1994) identified a mutation in the CLCN1 gene, resulting in a gly482-to-arg (G482R) substitution. Interestingly, this mutation producing a recessive phenotype is only 2 codons removed from a pro480-to-leu (P480L; 118425.0006) mutation which results in the dominant Thomsen type of myotonia congenita.


.0006 MYOTONIA CONGENITA, AUTOSOMAL DOMINANT

CLCN1, PRO480LEU
  
RCV000019089...

In affected members of Thomsen's own family (Thomsen, 1876; Thomasen, 1948) with autosomal dominant myotonia congenita (160800), Steinmeyer et al. (1994) identified a heterozygous mutation in the CLCN1 gene, resulting in a pro480-to-leu (P480L) substitution. Functional expression studies showed that the P480L mutation dramatically inhibited normal CLCN1 channel function via a dominant-negative effect.

Koch et al. (1994) also reported the P480L mutation as causative of Thomsen disease.

Pusch et al. (1995) transfected cDNA bearing the P480L mutation into Xenopus oocytes, demonstrating a large 90-mV shift of the gating toward positive voltages. In further structure studies, they replaced isoleucine 290 by 18 different amino acids. Substitution with valine shifted the gating by -17 millivolts. In all other replacements, the gating was either shifted to more positive voltages or resulted in no current above background.

Colding-Jorgensen et al. (2003) found that all 6 patients with the dominant P480L mutation had CMAP decrements above 30%, which the authors suggested was correlated with the large voltage shift conferred by the mutation.


.0007 MYOTONIA LEVIOR

CLCN1, GLN552ARG
  
RCV000019090...

In 2 brothers with a mild form of dominant myotonia, referred to as myotonia levior (see 160800), Lehmann-Horn et al. (1995) identified a 1655A-G transition in exon 15 of the CLCN1 gene, resulting in a gln552-to-arg (Q552R) substitution. Their affected mother also had the mutation. The findings indicated that myotonia levior is a variant or allelic form of Thomsen disease.

By functional expression studies, Ryan et al. (2002) demonstrated that CLCN1 channels with the Q552R mutation were formed normally, but had altered gating properties. Voltage dependence of the mutant channels was shifted by more than +90 mV compared to wildtype channels, resulting in a reduction in the open probability of the channel.


.0008 MYOTONIA CONGENITA, AUTOSOMAL DOMINANT

CLCN1, ILE290MET
  
RCV000019091...

In affected members of a family with Thomsen myotonia (160800), Lehmann-Horn et al. (1995) identified an 870C-G transversion in the CLCN1 gene, resulting in an ile290-to-met (I290M) substitution.


.0009 MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, 14-BP DEL
  
RCV000395981...

In patients with Becker myotonia (255700), Meyer-Kleine et al. (1994) identified a 14-bp deletion in exon 13 of the CLCN1 gene.

In a family in which the index patient and his mother had been examined by Becker (1977), who classified their disorder as dominant myotonia congenita, Lehmann-Horn et al. (1995) found that the proband was homozygous for a 14-bp deletion (involving nucleotides 1437-1450) in exon 13 of the CLCN1 gene. Both nonmyotonic sons of the index patient were heterozygous for the deletion.

In French Canadian patients with recessive myotonia, Dupre et al. (2009) found that homozygosity for the 14-bp deletion resulted in a severe phenotype with generalized hypertrophy, moderate myotonia, and transient weakness.


.0010 MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, GLU291LYS
  
RCV000019093...

In 2 sibs with autosomal recessive myotonia congenita (255700), Pusch et al. (1995) identified compound heterozygosity for 2 mutations in the CLCN1 gene: glu291-to-lys (E291K) and arg894-to-ter (R894X; 118425.0015). Functional expression studies showed that the E291K channels did not yield currents between -140 and 100 millivolts, indicating that this mutation totally abolished channel activity. In contrast to the mutation in the neighboring amino acid (I290M; 118425.0008), which acts as a dominant as a result of interactions with wildtype monomers, the E291K mutation is recessive. Whereas the I290M mutant shifts the voltage dependence of chloride channels positive (via homomers or heteromers with wildtype subunits), the E291K mutation shows no evidence of interaction and does not shift the voltage dependence.


.0011 MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

MYOTONIA CONGENITA, AUTOSOMAL DOMINANT, INCLUDED
CLCN1, ARG317GLN
  
RCV000019094...

The arg317-to-gln (R317Q) mutation illustrates the resultant occurrence of either autosomal dominant myotonia congenita (160800) or autosomal recessive myotonia congenita (255700), depending on the family background. Meyer-Kleine et al. (1995) observed the R317Q mutation in dominant Thomsen myotonia congenita; Esteban et al. (1998) observed it in Becker myotonia congenita and pointed to other mutations that had been observed as a recessive or a dominant, depending on the particular family.


.0012 MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, GLY499ARG
  
RCV000019096...

In a boy with Becker myotonia congenita (255700), Zhang et al. (2000) identified a C-to-T transition at codon 1495 in exon 14 of the CLCN1 gene, resulting in a gly499-to-arg (G499R) substitution in the putative transmembrane domain 10 of the CLCN1 protein. In contrast to normal CLCN1 channels that deactivate upon hyperpolarization, functional expression of G499R CLCN1 yielded a hyperpolarization-activated chloride current when measured in the presence of a high intracellular chloride concentration. Current was abolished when measured with a physiologic chloride transmembrane gradient. Electrophysiologic analysis of other gly449 mutants suggested that the positive charge introduced by the G499R mutation may be responsible for this unique gating behavior.


.0013 MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, 1-BP INS, 831G
  
RCV000019097...

In 2 African American sibs with autosomal recessive myotonia congenita (255700), Nagamitsu et al. (2000) identified compound heterozygosity for 2 mutations in the CLCN1 gene: a 1-bp insertion (831insG) in exon 7, resulting in a premature termination codon at codon 289, and a C-to-T transition in exon 23, resulting in a pro932-to-leu (P932L; 118425.0014) substitution. In addition to generalized myotonia and proximal muscle hypertrophy, the sibs also displayed progressive muscle weakness, joint contractures, and dystrophic muscle pathology.


.0014 MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, PRO932LEU
  
RCV000019085...

For discussion of the pro932-to-leu (P932L) mutation in the CLCN1 gene that was found in compound heterozygous state in sibs with autosomal recessive myotonia congenita (255700) by Nagamitsu et al. (2000), see 118425.0013.


.0015 MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

MYOTONIA CONGENITA, AUTOSOMAL DOMINANT, INCLUDED
CLCN1, ARG894TER
  
RCV000019098...

For discussion of the arg894-to-ter (R894X) mutation in the CLCN1 gene that was found in compound heterozygous state in sibs with autosomal recessive myotonia congenita (255700) by Pusch et al. (1995), see 118425.0010.

In patients with both autosomal recessive and autosomal dominant (160800) myotonia congenita, Meyer-Kleine et al. (1995) identified a mutation in the CLCN1 gene, resulting in an R894X substitution. Functional expression of the R894X mutant in Xenopus oocytes revealed a large reduction, but not complete abolition, of chloride currents. Further, it had a weak dominant-negative effect on wildtype currents in coexpression studies. Reduction of currents predicted for heterozygous carriers were close to the borderline value, sufficient to elicit myotonia.

Sun et al. (2001) found a carrier frequency of 0.87% for the R894X mutation in the northern Scandinavian population.

In a French Canadian family with myotonia, Dupre et al. (2009) found that the R894X mutation could be expressed in a semidominant or recessive manner. The proband, who was heterozygous for the R894X mutation, had muscle stiffness and mild warm-up phenomenon, but no significant percussion myotonia and no myotonia on EMG. In contrast, her daughters, who were compound heterozygous for the R894X and R300X (118425.0017) mutations, showed a moderately severe phenotype with generalized hypertrophy consistent with autosomal recessive Becker myotonia. This compound heterozygous genotype showed resistance to phenytoin, mexiletine, and quinine.


.0016 MYOTONIA CONGENITA, AUTOSOMAL DOMINANT

CLCN1, MET128VAL
  
RCV000019100...

In affected members of a family with autosomal dominant myotonia congenita (160800), Colding-Jorgensen et al. (2003) identified a heterozygous A-to-G transition in the CLCN1 gene, resulting in a met128-to-val (M128V) substitution.


.0017 MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, ARG300TER
  
RCV000019101

For discussion of the arg300-to-ter (R300X) mutation in the CLCN1 gene that was found in compound heterozygous state in patients with autosomal recessive myotonia congenita (255700) by Dupre et al. (2009), see 118425.0015.


.0018 MYOTONIA CONGENITA, AUTOSOMAL DOMINANT

CLCN1, SER189PHE
  
RCV000019102...

In 2 French Canadian women with autosomal dominant myotonia congenita (160800), Dupre et al. (2009) identified a heterozygous ser189-to-phe (S189F) substitution in the CLCN1 gene. Both women had mild fluctuating myotonia and mild muscle hypertrophy and reported aggravation of symptoms with menstrual periods and pregnancy.


.0019 MYOTONIA CONGENITA, AUTOSOMAL DOMINANT

MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE, INCLUDED
CLCN1, TRP433ARG
   RCV000019103...

Dupre et al. (2009) identified a heterozygous or homozygous trp433-to-arg (W433R) substitution in the CLCN1 gene in French Canadian patients with autosomal dominant (160800) and recessive (255700) myotonia, respectively. The recessive phenotype was characterized by severe myotonia, dysphagia, generalized hypertrophy predominant in lower limb muscles, and the warm-up phenomenon.


.0020 MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, DUP, EX8-14
   RCV000033240

In 2 unrelated Iraqi women with severe infantile-onset myotonia congenita (255700), Raja Rayan et al. (2012) identified a homozygous duplication of exons 8 through 14 of the CLCN1 gene, predicted to result in a frameshift. Both patients came from consanguineous families and had multiple affected family members. The duplication was identified by multiplex ligation-dependent probe amplification (MLPA) specific to the CLCN1 gene, and was not found in 124 control chromosomes or in a variant database. Haplotype analysis suggested a founder effect for this duplication mutation. Raja Rayan et al. (2012) concluded that copy number variation involving the CLCN1 gene is an important genetic mechanism in patients with recessive myotonia congenita, and that MLPA analysis may aid in genetic counseling.


.0021 VARIANT OF UNKNOWN SIGNIFICANCE

CLCN1, ARG976TER
  
RCV000180791...

This variant is classified as a variant of unknown significance because its contribution to idiopathic generalized epilepsy (EIG; see 600669) has not been confirmed.

In a 26-year-old Italian woman with EIG, Chen et al. (2013) identified a de novo heterozygous c.2926C-T transition (c.2926C-T, NM_000083) in exon 23 of the CLCN1 gene, resulting in an arg976-to-ter (R976X) substitution and truncation of the distal C terminus by 12 residues. The mutation, which was found by exome sequencing of 291 patients with epilepsy, was not found in the dbSNP or 1000 Genomes Project databases. The patient had generalized pharmacoresistant epilepsy with mixed seizure types, including generalized tonic-clonic and absence seizures. Her first seizure occurred during a febrile illness at age 11 months. EEG repeatedly showed generalized spike-wave discharges. In addition, she presented with myotonic writer's cramp at age 9 years, which subsequently resolved. EMG as an adult showed no evidence of myotonic discharges. Brain MRI and cognitive function were normal. Functional studies of the variant were not performed, but Chen et al. (2013) found expression of the CLCN1 gene in multiple human brain regions, and suggested that the mutation identified in this patient could result in neuronal hyperexcitability, thus contributing to the seizure phenotype.


.0022 MYOTONIA CONGENITA, AUTOSOMAL DOMINANT

CLCN1, GLN412PRO
  
RCV001382412...

In a Costa Rican patient (family 4-CR14) with autosomal dominant myotonia congenita (160800), Vindas-Smith et al. (2016) identified a heterozygous c.1235A-C transversion in exon 11 of the CLCN1 gene, resulting in a gln412-to-pro (Q412P) substitution. The mutation, which was identified by bidirectional sequencing of the CLCN1 gene and confirmed by RFLP-PCR, was also found in the patient's asymptomatic (self-reported) mother and 2 sibs. Functional studies of CLCN1 with the Q412P mutation expressed in Xenopus oocytes showed reduced surface expression and reduced current density but did not display a dominant-negative effect when coexpressed with wildtype CLCN1.


REFERENCES

  1. Adrian, R. H., Bryant, S. H. On the repetitive discharge in myotonic muscle fibres. J. Physiol. 240: 505-515, 1974. [PubMed: 4420758, related citations] [Full Text]

  2. Altamura, C., Lucchiari, S., Sahbani, D., Ulzi, G., Comi, G. P., D'Ambrosio, P., Petillo, R., Politano, L., Vercelli, L., Mongini, T., Dotti, M. T., Cardani, R., Meola, G., Lo Monaco, M., Matthews, E., Hanna, M. G., Carratu, M. R., Conte, D., Imbrici, P., Desaphy, J.-F. The analysis of myotonia congenita mutations discloses functional clusters of amino acids within the CBS2 domain and the C-terminal peptide of the CLC-1 channel. Hum. Mutat. 39: 1273-1283, 2018. [PubMed: 29935101, related citations] [Full Text]

  3. Aminoff, M. J., Layzer, R. B., Satya-Murti, S., Faden, A. I. The declining electrical response of muscle to repetitive nerve stimulation in myotonia. Neurology 27: 812-816, 1977. [PubMed: 561337, related citations] [Full Text]

  4. Beck, C. L., Fahlke, C., George, A. L., Jr. Molecular basis for decreased muscle chloride conductance in the myotonic goat. Proc. Nat. Acad. Sci. 93: 11248-11252, 1996. [PubMed: 8855341, related citations] [Full Text]

  5. Becker, P. E. Myotonia Congenita and Syndromes associated with Myotonia: Clinical-Genetic Studies of the Nondystrophic Myotonias. Stuttgart: Georg Thieme (pub.) 1977.

  6. Bryant, S. H. Muscle membrane of normal and myotonic goats in normal and low external chloride. (Abstract) Fed. Proc. 21: 312 only, 1962.

  7. Charlet-B, N., Savkur, R. S., Singh, G., Philips, A. V., Grice, E. A., Cooper, T. A. Loss of the muscle-specific chloride channel in type 1 myotonic dystrophy due to misregulated alternative splicing. Molec. Cell 10: 45-53, 2002. [PubMed: 12150906, related citations] [Full Text]

  8. Chen, T. T., Klassen, T. L., Goldman, A. M., Marini, C., Guerrini, R., Noebels, J. L. Novel brain expression of ClC-1 chloride channels and enrichment of CLCN1 variants in epilepsy. Neurology 80: 1078-1085, 2013. [PubMed: 23408874, images, related citations] [Full Text]

  9. Colding-Jorgensen, E., Duno, M., Schwartz, M., Vissing, J. Decrement of compound muscle action potential is related to mutation type in myotonia congenita. Muscle Nerve 27: 449-455, 2003. [PubMed: 12661046, related citations] [Full Text]

  10. Duno, M., Colding-Jorgensen, E., Grunnet, M., Jespersen, T., Vissing, J., Schwartz, M. Difference in allelic expression of the CLCN1 gene and the possible influence on the myotonia congenita phenotype. Europ. J. Hum. Genet. 12: 738-743, 2004. [PubMed: 15162127, related citations] [Full Text]

  11. Dupre, N., Chrestian, N., Bouchard, J.-P., Rossignol, E., Brunet, D., Sternberg, D., Brias, B., Mathieu, J., Puymirat, J. Clinical, electrophysiologic, and genetic study of non-dystrophic myotonia in French-Canadians. Neuromusc. Disord. 19: 330-334, 2009. [PubMed: 18337100, related citations] [Full Text]

  12. Dutzler, R., Campbell, E. B., Cadene, M., Chait, B. T., MacKinnon, R. X-ray structure of a CIC chloride channel at 3.0 angstrom reveals the molecular basis of anion selectivity. Nature 415: 287-294, 2002. [PubMed: 11796999, related citations] [Full Text]

  13. Esteban, J., Neumeyer, A. M., McKenna-Yasek, D., Brown, R. H. Identification of two mutations and a polymorphism in the chloride channel CLCN-1 in patients with Becker's generalized myotonia. Neurogenetics 1: 185-188, 1998. [PubMed: 10737121, related citations] [Full Text]

  14. Estevez, R., Schroeder, B. C., Accardi, A., Jentsch, T. J., Pusch, M. Conservation of chloride channel structure revealed by an inhibitor binding site in CIC-1. Neuron 38: 47-59, 2003. [PubMed: 12691663, related citations] [Full Text]

  15. Fahlke, C., Beck, C. L., George, A. L., Jr. A mutation in autosomal dominant myotonia congenita affects pore properties of the muscle chloride channel. Proc. Nat. Acad. Sci. 94: 2729-2734, 1997. [PubMed: 9122265, images, related citations] [Full Text]

  16. George, A. L., Jr., Crackower, M. A., Abdalla, J. A., Hudson, A. J., Ebers, G. C. Molecular basis of Thomsen's disease (autosomal dominant myotonia congenita). Nature Genet. 3: 305-310, 1993. [PubMed: 7981750, related citations] [Full Text]

  17. George, A. L., Jr. Personal Communication. Nashville, Tenn. 4/14/1997.

  18. Koch, M. C., Meyer-Kleine, C., Otto, M., Ricker, K., Lorenz, C., Steinmeyer, K., Jentsch, T. J. Mutations in the CLCN1 gene leading to myotonia congenita Thomsen and generalized myotonia Becker. (Abstract) Am. J. Hum. Genet. 55 (suppl.): A226, 1994.

  19. Koch, M. C., Ricker, K., Otto, M., Wolf, F., Zoll, B., Lorenz, C., Steinmeyer, K., Jentsch, T. J. Evidence for genetic homogeneity in autosomal recessive generalised myotonia (Becker). J. Med. Genet. 30: 914-917, 1993. [PubMed: 8301644, related citations] [Full Text]

  20. Koch, M. C., Steinmeyer, K., Lorenz, C., Ricker, K., Wolf, F., Otto, M., Zoll, B., Lehmann-Horn, F., Grzeschik, K.-H., Jentsch, T. J. The skeletal muscle chloride channel in dominant and recessive human myotonia. Science 257: 797-800, 1992. [PubMed: 1379744, related citations] [Full Text]

  21. Koty, P. P., Marks, H. G., Turel, A., Flagler, D., Angelini, C., Pegoraro, E., Vancott, A. C., Manchester, D., Zonana, J., Bird, T. D., Hoffman, E. P. Linkage analysis of Thomsen and Becker myotonia families. (Abstract) Am. J. Hum. Genet. 55 (suppl.): A227, 1994.

  22. Kubisch, C., Schmidt-Rose, T., Fontaine, B., Bretag, A. H., Jentsch, T. J. ClC-1 chloride channel mutations in myotonia congenita: variable penetrance of mutations shifting the voltage dependence. Hum. Molec. Genet. 7: 1753-1760, 1998. [PubMed: 9736777, related citations] [Full Text]

  23. Lehmann-Horn, F., Mailander, V., Heine, R., George, A. L. Myotonia levior is a chloride channel disorder. Hum. Molec. Genet. 4: 1397-1402, 1995. [PubMed: 7581380, related citations] [Full Text]

  24. Lipicky, R. J., Bryant, S. H. Sodium, potassium, and chloride fluxes in intercostal muscle from normal goats and goats with hereditary myotonia. J. Gen. Physiol. 50: 89-111, 1966. [PubMed: 5971035, related citations] [Full Text]

  25. Lorenz, C., Meyer-Kleine, C., Steinmeyer, K., Koch, M. C., Jentsch, T. J. Genomic organization of the human muscle chloride channel CLC-1 and analysis of novel mutations leading to Becker-type myotonia. Hum. Molec. Genet. 3: 941-946, 1994. [PubMed: 7951242, related citations] [Full Text]

  26. Mailander, V., Heine, R., Deymeer, F., Lehmann-Horn, F. Novel muscle chloride channel mutations and their effects on heterozygous carriers. Am. J. Hum. Genet. 58: 317-324, 1996. [PubMed: 8571958, related citations]

  27. Mankodi, A., Takahashi, M. P., Jiang, H., Beck, C. L., Bowers, W. J., Moxley, R. T., Cannon, S. C., Thornton, C. A. Expanded CUG repeats trigger aberrant splicing of ClC-1 chloride channel pre-mRNA and hyperexcitability of skeletal muscle in myotonic dystrophy. Molec. Cell 10: 35-44, 2002. [PubMed: 12150905, related citations] [Full Text]

  28. Meyer-Kleine, C., Ricker, K., Otto, M., Koch, M. C. A recurrent 14 bp deletion in the CLCN1 gene associated with generalized myotonia (Becker). Hum. Molec. Genet. 3: 1015-1016, 1994. [PubMed: 7951215, related citations] [Full Text]

  29. Meyer-Kleine, C., Steinmeyer, K., Ricker, K., Jentsch, T. J., Koch, M. C. Spectrum of mutations in the major human skeletal muscle chloride channel gene (CLCN1) leading to myotonia. Am. J. Hum. Genet. 57: 1325-1334, 1995. [PubMed: 8533761, related citations]

  30. Mindell, J. A., Maduke, M., Miller, C., Grigorieff, N. Projection structure of a CIC-type chloride channel at 6.5-angstrom resolution. Nature 409: 219-223, 2001. [PubMed: 11196649, related citations] [Full Text]

  31. Nagamitsu, S., Matsuura, T., Khajavi, M., Armstrong, R., Gooch, C., Harati, Y., Ashizawa, T. A 'dystrophic' variant of autosomal recessive myotonia congenita caused by novel mutations in the CLCN1 gene. Neurology 55: 1697-1703, 2000. [PubMed: 11113225, related citations] [Full Text]

  32. Pusch, M., Steinmeyer, K., Koch, M. C., Jentsch, T. J. Mutations in dominant human myotonia congenita drastically alter the voltage dependence of the CIC-1 chloride channel. Neuron 15: 1455-1463, 1995. [PubMed: 8845168, related citations] [Full Text]

  33. Pusch, M. Myotonia caused by mutations in the muscle chloride channel gene CLCN1. Hum. Mutat. 19: 423-434, 2002. [PubMed: 11933197, related citations] [Full Text]

  34. Raja Rayan, D. L., Haworth, A., Sud, R., Matthews, E., Fialho, D., Burge, J., Portaro, S., Schorge, S., Tuin, K., Lunt, P., McEntagart, M., Toscano, A., Davis, M. B., Hanna, M. G. A new explanation for recessive myotonia congenita: exon deletions and duplications in CLCN1. Neurology 78: 1953-1958, 2012. [PubMed: 22649220, images, related citations] [Full Text]

  35. Rudel, R. The myotonic mouse--a realistic model for the study of human recessive generalized myotonia. Trends Neurosci. 13: 1-3, 1990. [PubMed: 1688667, related citations] [Full Text]

  36. Ryan, A., Rudel, R., Kuchenbecker, M., Fahlke, C. A novel alteration of muscle chloride channel gating in myotonia levior. J. Physiol. 545: 345-354, 2002. [PubMed: 12456816, images, related citations] [Full Text]

  37. Steinmeyer, K., Klocke, R., Ortland, C., Gronemeier, M., Jockusch, H., Grunder, S., Jentsch, T. J. Inactivation of muscle chloride channel by transposon insertion in myotonic mice. Nature 354: 304-308, 1991. [PubMed: 1659665, related citations] [Full Text]

  38. Steinmeyer, K., Lorenz, C., Pusch, M., Koch, M. C., Jentsch, T. J. Multimeric structure of ClC-1 chloride channel revealed by mutations in dominant myotonia congenita (Thomsen). EMBO J. 13: 737-743, 1994. [PubMed: 8112288, related citations] [Full Text]

  39. Suetterlin, K., Matthews, E., Sud, R., McCall, S., Fialho, D., Burge, J., Jayaseelan, D., Haworth, A., Sweeney, M. G., Kullmann, D. M., Schorge, S., Hanna, M. G., Mannikko, R. Translating genetic and functional data into clinical practice: a series of 223 families with myotonia. Brain 145: 607-620, 2022. [PubMed: 34529042, images, related citations] [Full Text]

  40. Sun, C., Tranebjaerg, L., Torbergsen, T., Holmgren, G., Van Ghelue, M. Spectrum of CLCN1 mutations in patients with myotonia congenita in northern Scandinavia. Europ. J. Hum. Genet. 9: 903-909, 2001. Note: Erratum: Europ. J. Hum. Genet. 18: 264 only, 2010. [PubMed: 11840191, related citations] [Full Text]

  41. Thomasen, E. Myotonia, Thomsen's disease. Paramyotonia, and dystrophia myotonica. Op. Ex Domo Biol. Hered. Hum. U. Hafniensis 17: 11-251, 1948.

  42. Thomsen, J. Tonische Kraempfe in willkuerlich beweglichen Muskeln in Folge von ererbter psychischer Disposition: Ataxia muscularis? Arch. Psychiat. Nervenkr. 6: 702-718, 1876.

  43. Vindas-Smith, R., Fiore, M., Vasquez, M., Cuenca, P., del Valle, G., Lagostena, L., Gaitan-Penas, H., Estevez, R., Pusch, M., Morales, F. Identification and functional characterization of CLCN1 mutations found in nondystrophic myotonia patients. Hum. Mutat. 37: 74-83, 2016. [PubMed: 26510092, related citations] [Full Text]

  44. Wollnik, B., Kubisch, C., Steinmeyer, K., Pusch, M. Identification of functionally important regions of the muscular chloride channel CIC-1 by analysis of recessive and dominant myotonic mutations. Hum. Molec. Genet. 6: 805-811, 1997. [PubMed: 9158157, related citations] [Full Text]

  45. Wu, H., Olson, E. N. Activation of the MEF2 transcription factor in skeletal muscles from myotonic mice. J. Clin. Invest. 109: 1327-1333, 2002. [PubMed: 12021248, images, related citations] [Full Text]

  46. Zhang, J., Sanguinetti, M. C., Kwiecinski, H., Ptacek, L. J. Mechanism of inverted activation of ClC-1 channels caused by a novel myotonia congenita mutation. J. Biol. Chem. 275: 2999-3005, 2000. [PubMed: 10644771, related citations] [Full Text]


Hilary J. Vernon - updated : 05/27/2022
Cassandra L. Kniffin - updated : 6/16/2015
Cassandra L. Kniffin - updated : 3/11/2013
Cassandra L. Kniffin - updated : 10/27/2009
Cassandra L. Kniffin - updated : 6/20/2005
Patricia A. Hartz - updated : 5/4/2004
Victor A. McKusick - updated : 2/6/2003
Patricia A. Hartz - updated : 10/8/2002
Stylianos E. Antonarakis - updated : 9/10/2002
Michael B. Petersen - updated : 8/19/2002
Victor A. McKusick - updated : 4/12/2002
Ada Hamosh - updated : 1/17/2002
Kathryn R. Wagner - updated : 8/6/2001
Ada Hamosh - updated : 1/9/2001
Ada Hamosh - updated : 2/11/2000
Victor A. McKusick - updated : 5/6/1998
Victor A. McKusick - updated : 6/23/1997
Victor A. McKusick - updated : 5/16/1997
Victor A. McKusick - updated : 4/21/1997
Orest Hurko - updated : 3/9/1996
Creation Date:
Victor A. McKusick : 9/29/1992
carol : 06/01/2022
carol : 05/27/2022
carol : 11/22/2019
carol : 11/21/2019
carol : 06/21/2019
alopez : 10/12/2016
alopez : 02/24/2016
carol : 6/26/2015
mcolton : 6/17/2015
ckniffin : 6/16/2015
mgross : 10/7/2013
alopez : 3/12/2013
ckniffin : 3/11/2013
ckniffin : 10/31/2011
wwang : 11/16/2009
ckniffin : 10/27/2009
terry : 6/23/2006
carol : 6/23/2006
carol : 6/29/2005
ckniffin : 6/20/2005
terry : 3/14/2005
mgross : 5/4/2004
carol : 2/11/2003
tkritzer : 2/11/2003
terry : 2/6/2003
mgross : 10/8/2002
mgross : 9/10/2002
mgross : 9/10/2002
mgross : 9/10/2002
alopez : 8/19/2002
alopez : 8/19/2002
alopez : 4/25/2002
cwells : 4/22/2002
terry : 4/12/2002
alopez : 1/22/2002
terry : 1/17/2002
carol : 8/6/2001
carol : 8/6/2001
mgross : 1/10/2001
terry : 1/9/2001
alopez : 2/15/2000
terry : 2/11/2000
carol : 11/9/1999
carol : 7/7/1999
carol : 6/8/1998
carol : 5/12/1998
terry : 5/6/1998
mark : 7/9/1997
alopez : 6/27/1997
terry : 6/23/1997
terry : 6/18/1997
mark : 6/9/1997
terry : 6/5/1997
terry : 6/5/1997
mark : 5/16/1997
terry : 5/10/1997
terry : 5/6/1997
jenny : 4/21/1997
terry : 4/14/1997
mark : 2/23/1997
mark : 12/18/1996
jamie : 12/6/1996
terry : 12/4/1996
terry : 4/15/1996
mark : 3/9/1996
terry : 2/23/1996
mark : 2/22/1996
terry : 2/19/1996
mark : 1/19/1996
mark : 12/18/1995
joanna : 12/15/1995
mark : 12/15/1995
mark : 12/15/1995
terry : 12/14/1995
terry : 12/13/1995
terry : 10/30/1995
mark : 9/7/1995
carol : 1/23/1995
jason : 7/27/1994
carol : 12/20/1993
carol : 4/29/1993

* 118425

CHLORIDE CHANNEL 1, SKELETAL MUSCLE; CLCN1


Alternative titles; symbols

CHLORIDE CHANNEL, MUSCLE; CLC1


HGNC Approved Gene Symbol: CLCN1

SNOMEDCT: 20305008, 57938005, 726051002, 8960007;   ICD10CM: G71.12;  


Cytogenetic location: 7q34     Genomic coordinates (GRCh38): 7:143,316,111-143,352,083 (from NCBI)


Gene-Phenotype Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
7q34 Myotonia congenita, dominant 160800 Autosomal dominant 3
Myotonia congenita, recessive 255700 Autosomal recessive 3
Myotonia levior 160800 Autosomal dominant 3

TEXT

Description

The muscle chloride channel CLCN1 regulates the electric excitability of the skeletal muscle membrane. Skeletal muscle has an unusually high resting Cl(-) conductance and in vitro studies suggest that reduction of this conductance causes electrical instability and resulting myotonia in both humans and animal models. Muscle Cl(-) conductance is predominantly mediated by the CLCN1 chloride channel (summary by Steinmeyer et al., 1994).


Cloning and Expression

By homology screening with the major rat skeletal muscle chloride channel CLCN1, Koch et al. (1992) cloned a partial human CLCN1 cDNA that covered about 80% of the coding sequence. This region was 88% identical to the rat channel in amino acid sequence.

Chen et al. (2013) found expression of the CLCN1 gene in various human brain regions, including cerebellum, hippocampus, spinal cord, and cerebral cortex, as well as in heart. Clcn1 was also expressed in the brain and heart of developing and adult mice. In mouse brain, neuronal expression of Clcn1 was detected in the pyramidal and dentate granule cells of the hippocampus, cerebellar Purkinje cells, scattered brainstem nuclei, frontal neocortex, and thalamus. The results suggested that CLCN1 has a role in neuronal function and excitability in addition to its known role in skeletal muscle.


Gene Function

Steinmeyer et al. (1994) determined that CLCN1 functions as a homooligomer, most likely with 4 subunits.

Pusch et al. (1995) used a Xenopus transfection to demonstrate shifting of the gating of CLCN1 toward positive voltages by 4 different mutations identified in patients with myotonia congenita (160800). When these mutant cDNAs were coexpressed with wildtype subunits, they imposed altered voltage dependence on the heteromeric channels which would then open only in a voltage range where they could not contribute significantly to the repolarization of action potentials. Without such repolarizations, sodium channels have enough time to recover from inactivation leading to typical myotonic runs, which are a series of repetitive action potentials.


Biochemical Features

Cryoelectron Microscopy

Mindell et al. (2001) reported the formation of 2-dimensional crystals of the prokaryotic CLC channel homolog EriC reconstituted into phospholipid bilayer membranes. Cryoelectron microscopic analysis of these crystals yielded a projection structure at 6.5-angstrom resolution that showed off-axis water-filled pores within the dimeric channel complex.

Crystal Structure

Dutzler et al. (2002) presented the x-ray structures of 2 prokaryotic CLC chloride channels, from Salmonella typhimurium and E. coli, at 3.0 and 3.5 angstroms, respectively. Both structures revealed 2 identical pores, each pore being formed by a separate subunit contained within a homodimeric membrane protein.

Using the 3-dimensional crystal structure of a bacterial CLC protein to predict residues involved in chloride channel inhibitor binding, Estevez et al. (2003) identified an inhibitor-binding site in human CLCN1. The binding site is localized close to the chloride-binding site and is accessible only from the intracellular side. Estevez et al. (2003) concluded that the structures of bacterial CLCs can be extrapolated with fidelity to mammalian chloride channels.


Gene Structure

Lorenz et al. (1994) showed that the protein coding sequence of the CLCN1 gene is organized into 23 exons. Its upstream region contains a canonical TATA box, several consensus binding sites for myogenic transcription factors, and 2 other putative regulatory elements.


Mapping

By blot hybridization to a panel of chromosome 7-specific, human-mouse somatic cell hybrids, Koch et al. (1992) mapped the CLCN1 gene to 7q32-qter. With RFLPs in the CLCN1 gene, they demonstrated that the locus is linked to the T-cell receptor beta locus (see 186930) at 7q35 (maximum lod = 5.23 at theta = 0.0). In the mouse, it had previously been demonstrated that the corresponding loci are linked on chromosome 6, which shows other evidence of homology of synteny to human 7q (Steinmeyer et al., 1991).


Molecular Genetics

Myotonia Congenita

In 3 brothers, born of consanguineous parents, with autosomal recessive myotonia congenita (Becker disease; 255700), Koch et al. (1992) identified a homozygous mutation in the CLCN1 gene (F413C; 118425.0001). In affected members of 3 unrelated families with autosomal dominant myotonia congenita (Thomsen disease; 160800), George et al. (1993) identified a heterozygous mutation in the CLCN1 gene (G230E; 118425.0002). The findings indicated that the 2 disorders are allelic.

Meyer-Kleine et al. (1995) identified 15 different mutations in the CLCN1 gene, of which 10 were novel, in a total of 17 unrelated families and 13 single patients with Becker-type myotonia congenita. One additional family had the dominant Thomsen form. Three mutations accounted for 32% of the Becker chromosomes in the German population; F413C, R894X (118425.0015), and a 14-bp deletion in exon 13 (118425.0009). Although a 437A-T transversion had been described as a disease-causing mutation (Koty et al., 1994), Meyer-Kleine et al. (1995) observed it in 3 myotonia families and in 5 of 200 control chromosomes, consistent with a polymorphism. Moreover, mutant 437A-T cRNA was functionally expressed in Xenopus oocytes and found to induce currents that were indistinguishable from wildtype currents.

In a screening of 6 unrelated patients with recessive Becker-type myotonia, Mailander et al. (1996) identified 4 novel CLCN1 mutations and a previously reported 14-bp deletion. Five patients were homozygous and the sixth patient was compound heterozygous. Heterozygous carriers of the Becker mutation did not display any clinical symptoms of myotonia; however, all heterozygous males, but none of the heterozygous females, exhibited myotonic discharges on EMG, suggesting a gene-dosage effect of the mutations on chloride conductance and a male predominance of subclinical myotonia.

Wollnik et al. (1997) analyzed the effect of 1 dominant and 3 recessive mutations in the CLCN1 gene after functional expression in Xenopus oocytes.

Esteban et al. (1998) stated that 31 specific mutations in the CLCN1 gene had been related to myotonia congenita in humans and 3 in mice. They described a homozygous arg317-to-gln (R317Q; 118425.0011) mutation in affected members of a family with autosomal recessive myotonia congenita. A heterozygous brother had mild muscle stiffness consistent with latent myotonia. The parents were unaffected, although only the mother, who was heterozygous, was available for DNA study. The same R317Q mutation had previously been identified in a family with dominant Thomsen myotonia congenita. At least 2 other mutations, G230E (118425.0002) and R894X, had been found in both dominant and recessive myotonia congenita, depending on the particular family.

Kubisch et al. (1998) identified 4 novel missense mutations in the CLCN gene, 2 in the dominant form and 2 in the recessive form of myotonia. The first 2 displayed a classic dominant phenotype with a dominant-negative effect by significantly imparting a voltage shift on mutant/wildtype heteromeric channels as found in heterozygous patients. One of the recessive mutations also shifted the voltage dependence to positive values, but coexpression with wildtype CLCN1 gave almost wildtype currents. The voltage dependence of mutant heteromeric channels was not always intermediate between those of the constituent homomeric channel subunits. These complex interactions correlated clinically with various inheritance patterns, ranging from autosomal dominant with various degrees of penetrance to autosomal recessive.

In affected members of 18 unrelated families from Norway and Sweden with both autosomal dominant (5 families) and autosomal recessive (13 families) inheritance of myotonia congenita, Sun et al. (2001) identified 8 different mutations (1 nonsense, 4 missense, and 3 splice mutations) in the CLCN1 gene; 3 mutations were novel. Fifteen patients had mutations on both alleles, consistent with the recessive disorder; 2 probands had mutations in a single allele; and 2 probands had no CLCN1 mutations. In 2 families, 3 CLCN1 mutations were found in the proband, and Sun et al. (2001) suspected that this phenomenon may be underestimated because mutation search in a disease gene usually ends by the identification of 2 mutations in a family with recessive inheritance. Families with apparently dominant segregation of myotonia congenita may actually represent recessive inheritance with undetected heterozygous individuals married-in as a consequence of a high population carrier frequency of some mutations. The findings, together with the very variable clinical presentation, challenged the classification into dominant Thomsen or recessive Becker disease. Sun et al. (2001) suggested that most cases of myotonia congenita show recessive inheritance with some modifying factors or genetic heterogeneity.

In a review, Pusch (2002) stated that more than 60 CLCN1 mutations had been identified as causing myotonia, with only a few of them being dominant. A dominant-negative effect of mutant subunits in mutant-wildtype heterodimers was suggested as the usual mechanism for dominant mutations.

Duno et al. (2004) reported 4 unrelated families with myotonia congenita and the R894X mutation: 2 families had a single mutant allele, showing dominant inheritance, and 2 families were compound heterozygous for R894X and another mutation, showing recessive inheritance. RT-PCR did not reveal any association between total CLCN1 mRNA in muscle and the mode of inheritance, but the dominant family with the most severe phenotype expressed twice the expected amount of the R894X mRNA allele, even compared to the recessive families. Duno et al. (2004) suggested that variation in allelic expression may be a modifier of disease expression and progression in myotonia congenita.

Raja Rayan et al. (2012) performed multiplex ligation-dependent probe amplification (MLPA) specific to the CLCN1 gene in 60 families with recessive myotonia congenita in whom either no mutations or only a single pathogenic CLCN1 mutation had been identified. The results were positive in 4 (6.7%) patients: 2 unrelated patients were found to have 2 different multiexon deletions within the CLCN1 gene on the second allele, and 2 additional patients had a homozygous duplication of exons 8 through 14 of the CLCN1 gene (118425.0020). The 2 patients with the duplication were both of Iraqi origin, but were unrelated. Both Iraqi patients had a severe form of the disorder with onset in infancy. Haplotype analysis suggested a founder effect for this duplication mutation. Raja Rayan et al. (2012) concluded that copy number variation involving the CLCN1 gene is an important genetic mechanism in patients with recessive myotonia congenita, and that MLPA analysis may aid in genetic counseling.

In 4 Costa Rican families with myotonia congenita, Vindas-Smith et al. (2016) identified mutations in the CLCN1 gene by bidirectional sequencing of the CLCN1 gene, with confirmation by RFLP-PCR. In 2 families in which the proband had Thomsen disease (families 1 and 4), heterozygous mutations were identified (F167L; Q412P, 118425.0022). In another family in which the proband had Thomsen disease (family 2), compound heterozygous mutations were identified (R105C and F167L). In a proband (family 3) with a myositis-like disorder, a variant of unknown significance (Q154R) was identified. Functional studies of CLCN1 with the F167L, R105C, or Q154R mutation did not show alterations of gating parameters or channel conductance. Functional studies of CLCN1 with the Q412P mutation expressed in Xenopus oocytes showed reduced surface expression and reduced current density. Vindas-Smith et al. (2016) concluded that the Q412P mutation induces a severe folding defect that leads to its degradation before it can dimerize with the wildtype subunit.

Altamura et al. (2018) evaluated the functional significance of 7 mutations in the C-terminal region of the CLCN1 gene associated with either autosomal dominant Thomsen disease or autosomal recessive Becker disease. CLCN1 with each mutation was transfected into HEK293 cells and analyzed with patch-clamp analysis. Five of the mutations were in the CBS2 domain (V829M, T832I, V851M, G859V, L861P) and 2 of the mutations were in the C-terminal peptide (P883T, V947E). Mutations located between residues 829 and 835 and in residue 883 resulted in alteration of voltage dependence. Mutations between residues 851 and 859 and in residue 947 resulted in a reduction of chloride currents. The results were consistent with a role for CBS2 in protein channel gating and demonstrated the importance of the C-peptide region in protein function and expression.

Suetterlin et al. (2022) assessed the function of 95 CLCN1 mutations, including 34 novel mutations, identified in 233 patients with myotonia congenita. Mutations in CLCN1 were assessed by transfection of cDNA with each mutation into Xenopus laevis oocytes and analyzed with 2-electrode voltage clamp assays. From a functional standpoint, mutations that altered voltage dependence of activation clustered in the first half of the transmembrane domains and mutations resulting in absent currents clustered in the second half of the transmembrane domain. In terms of assessment of clinical significance and inheritance patterns, mutations that resulted in dominant functional features clustered in the TM1 domain, and variants associated with recessive functional features and without a shift in voltage dependence of activation were clustered in the TM2 domain. Mutations in the intracellular domain were not associated with a dominant inheritance pattern. Suetterlin et al. (2022) concluded that functional characterization of CLCN1 mutations improves the assessment of their clinical implications.

Myotonic Dystrophy

Myotonic dystrophy (DM1; 160900) is caused by a trinucleotide repeat expansion of the DMPK gene (605377.0001). In DM, expression of RNAs that contain expanded CUG or CCUG repeats is associated with degeneration and repetitive action potentials (myotonia) in skeletal muscle. Using skeletal muscle from a transgenic mouse model of DM, Mankodi et al. (2002) showed that expression of expanded CUG repeats reduced the transmembrane chloride conductance to levels well below those expected to cause myotonia. The expanded CUG repeats triggered aberrant splicing of pre-mRNA for CLCN1, resulting in loss of the CLCN1 protein from the surface membrane. Mankodi et al. (2002) identified a similar defect in CLCN1 splicing and expression in DM1 and DM2 (602668). The authors proposed that a transdominant effect of mutant RNA on CLCN1 RNA processing leads to chloride channelopathy and membrane hyperexcitability in DM.

Charlet-B et al. (2002) demonstrated loss of CLCN1 mRNA and protein in DM1 skeletal muscle tissue due to aberrant splicing of the CLCN1 pre-mRNA. They showed that the splicing regulator, CUG-binding protein (CUGBP; 601074), which is elevated in DM1 striated muscle, bound to the CLCN1 pre-mRNA, and that overexpression of CUGBP in normal cells reproduced the aberrant pattern of CLCN1 splicing observed in DM1 skeletal muscle. Charlet-B et al. (2002) proposed that disruption of alternative splicing regulation causes a predominant pathologic feature of DM1.

Associations Pending Confirmation

See 118425.0021 for discussion of a possible association between variation in the CLCN1 gene and idiopathic generalized epilepsy (EIG; see 600669).


Genotype/Phenotype Correlations

Aminoff et al. (1977) found that a subset of patients with myotonia congenita showed a marked decrement of the compound motor action potential (CMAP) with repeated stimulation compared to controls. Although the presence of decrement was not related to the degree of myotonia, it was usually observed in patients with autosomal recessive disease. Colding-Jorgensen et al. (2003) found that all 6 patients with the dominant P480L (118425.0006) mutation had CMAP decrements above 30%. In patients with the R894X (118425.0015) mutation, some had a large decrement, some had a slight decrement, and some had no change. Two patients with 2 CLCN1 mutations, 1 of whom carried an R894X mutation, had large decrements above 80%. Presence of decrement did not correlate with disease severity. Colding-Jorgensen et al. (2003) concluded that CMAP decrement may occur in dominant myotonia congenita and likely reflects the degree of chloride conduction reduction caused by particular mutations.


Animal Model

The Adr ('arrested development of righting') mouse, a model of autosomal recessive myotonia congenita, was found to be due to a mutation on mouse chromosome 6 (Rudel, 1990) in a region of the genome with homology of synteny to human chromosome 7q31-q35. The locus in the mouse is between those for Tcrb and Hox1. Steinmeyer et al. (1991) found that the Adr mouse phenotype was caused by a transposon insertion mutation that altered and inactivated the muscle-membrane chloride channel gene. The findings confirmed that the Adr mouse is an authentic model of Becker disease in the human.

Wu and Olson (2002) determined that Adr mice, which have an inactivated Clcn1 gene, showed an increased number of oxidative fibers, lacked glycolytic fibers, and showed muscle hypertrophy, similar to patients with Becker syndrome. By breeding Clcn1-null mice with mice harboring an Mef2 (600660)-dependent reporter gene, they found that the transcriptional activity of Mef2 was dramatically enhanced in myotonic muscle. Induction of Mef2 did not correlate with enhanced DNA-binding activity, but did correlate with activation of p38 Mapk (see 600289), an activator of Mef2, and with reduced expression of class II histone deacetylases (HDACs; see 605314), which repress Mef2 activity. Wu and Olson (2002) hypothesized that mutations in Clcn1 alter intracellular calcium levels, leading to the combined effects of class II Hdac deficiency and p38 Mapk activation, resulting in upregulation of Mef2 transcriptional activity followed by long-term changes in gene expression and to fiber-type transformation.

Beck et al. (1996) noted that the current hypotheses regarding the pathophysiology of autosomal dominant myotonia congenita, or Thomsen disease, were initially formulated from studies of the myotonic goat, an unusual breed afflicted with severe autosomal dominant congenital myotonia that closely resembles the human disease clinically and in its mode of inheritance. These animals are often referred to as 'fainting,' 'nervous,' 'stiff-legged,' or 'epileptic' goats because of their tendency to develop severe acute muscle stiffness and become immobile and often fall when attempting to make sudden forceful movements or when startled. The pathogenesis of myotonia in the goat was elucidated by Bryant and colleagues (Bryant, 1962, Lipicky and Bryant, 1966) who first described a severely diminished resting chloride conductance in muscle fibers from affected animals. The same group (Adrian and Bryant, 1974) also demonstrated that myotonia could be produced in normal skeletal muscle fibers bathed in a chloride-free solution. Beck et al. (1996) demonstrated the molecular basis for the decreased muscle chloride conductance in this historically important animal model. They found a single nucleotide change (GCC to CCC), resulting in an ala885-to-pro (A885P) substitution in a conserved residue in the C terminus of the goat muscle chloride channel, 104 residues from the termination codon. Heterologous expression of the mutation demonstrated a substantial (+47 mV) shift in the midpoint of steady-state activation of the channel, resulting in a diminished channel open probability at voltages near the resting membrane potential of skeletal muscle.


ALLELIC VARIANTS 22 Selected Examples):

.0001   MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, PHE413CYS
SNP: rs121912799, gnomAD: rs121912799, ClinVar: RCV000019083, RCV000184008, RCV000346725, RCV000638232, RCV001548747, RCV002291268

In 3 brothers with autosomal recessive generalized myotonia congenita (Becker disease; 255700), born of consanguineous parents, Koch et al. (1992) identified a homozygous T-to-G transversion in the CLCN1 gene, resulting in a phe413-to-cys (F413C) substitution toward the end of putative membrane span D8. This residue lies in a highly conserved region of the channel protein that is predicted to form a membrane-spanning alpha helix and possibly a component of the permeation pore (George et al., 1993).

Koch et al. (1993) found the F413C missense mutation in 15% of chromosomes carrying a gene for recessive myotonia congenita.


.0002   MYOTONIA CONGENITA, AUTOSOMAL DOMINANT

CLCN1, GLY230GLU
SNP: rs80356700, gnomAD: rs80356700, ClinVar: RCV000019084, RCV000020113, RCV000291823, RCV000627758, RCV003317041, RCV003904849

In affected members of 3 unrelated families with autosomal dominant myotonia congenita (Thomsen disease; 160800), George et al. (1993) identified a G-to-A transition in the CLCN1 gene, resulting in a gly230-to-glu (G230E) substitution between the third and fourth predicted membrane-spanning segments. This glycine residue is conserved in all known members of this class of chloride channel proteins. The codon number used for this mutation was based on the available partial-length cDNA of CLCN1; when the full-length cDNA became known, the codon number was changed from 180 to 230 (George, 1997).

In functional expression studies, Fahlke et al. (1997) found that the G230E mutation caused substantial changes in anion and cation selectivity, as well as a fundamental change in rectification of the current-voltage relationship. Whereas wildtype channels were characterized by pronounced inward rectification and a characteristic pattern of selectivity, G230E exhibited outward rectification at positive potentials and a different pattern of selectivity. Furthermore, the cation-to-anion permeability ratio of the mutant was much greater than that of the wildtype channel.


.0003   MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, IVSDS, G-A, +1
SNP: rs1563078827, ClinVar: RCV000019086

In affected members of a German family with recessive myotonia (255700), Lorenz et al. (1994) identified compound heterozygosity for 2 mutations in the CLCN1 gene: a 979G-A transition in a splice consensus site at the end of exon 8, and a 1488G-T transversion in exon 14, resulting in an arg496-to-ser (R496S; 118425.0004). Functional expression of R496S cRNA in Xenopus oocytes yielded no detectable currents. Furthermore, it did not suppress wildtype currents in coexpression assay, confirming it as a recessive mutation. The G-to-A transition in exon 8 was stated to affect the last nucleotide of the exon. If this interfered with mRNA splicing at that exon/intron boundary, the translation product would be terminated by a stop codon after 51 additional amino acids or other splice sites in the intron might be used. Alternatively, if splicing were normal, this mutation would lead to a substitution of isoleucine for valine at position 327 (V327I). Since this residue is not conserved among the members of this gene family and most members have negatively charged glutamate residues at this position, it is unlikely that such a substitution would have a dramatic effect on channel function. This would argue for an aberrant splicing as the effect of the G979A mutation.


.0004   MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, ARG496SER
SNP: rs121912801, gnomAD: rs121912801, ClinVar: RCV000019087, RCV000692794, RCV001781281, RCV003415718

For discussion of the arg496-to-ser (R496S) mutation in the CLCN1 gene that was found in compound heterozygous state in affected members of a family with recessive myotonia (255700) by Lorenz et al. (1994), see 118425.0003.


.0005   MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, GLY482ARG
SNP: rs746125212, gnomAD: rs746125212, ClinVar: RCV000019088, RCV000657922, RCV000810078, RCV003764473, RCV003989574

In affected members of a family with Becker myotonia congenita (255700), Koch et al. (1994) identified a mutation in the CLCN1 gene, resulting in a gly482-to-arg (G482R) substitution. Interestingly, this mutation producing a recessive phenotype is only 2 codons removed from a pro480-to-leu (P480L; 118425.0006) mutation which results in the dominant Thomsen type of myotonia congenita.


.0006   MYOTONIA CONGENITA, AUTOSOMAL DOMINANT

CLCN1, PRO480LEU
SNP: rs80356694, gnomAD: rs80356694, ClinVar: RCV000019089, RCV000020101, RCV001237767

In affected members of Thomsen's own family (Thomsen, 1876; Thomasen, 1948) with autosomal dominant myotonia congenita (160800), Steinmeyer et al. (1994) identified a heterozygous mutation in the CLCN1 gene, resulting in a pro480-to-leu (P480L) substitution. Functional expression studies showed that the P480L mutation dramatically inhibited normal CLCN1 channel function via a dominant-negative effect.

Koch et al. (1994) also reported the P480L mutation as causative of Thomsen disease.

Pusch et al. (1995) transfected cDNA bearing the P480L mutation into Xenopus oocytes, demonstrating a large 90-mV shift of the gating toward positive voltages. In further structure studies, they replaced isoleucine 290 by 18 different amino acids. Substitution with valine shifted the gating by -17 millivolts. In all other replacements, the gating was either shifted to more positive voltages or resulted in no current above background.

Colding-Jorgensen et al. (2003) found that all 6 patients with the dominant P480L mutation had CMAP decrements above 30%, which the authors suggested was correlated with the large voltage shift conferred by the mutation.


.0007   MYOTONIA LEVIOR

CLCN1, GLN552ARG
SNP: rs80356696, gnomAD: rs80356696, ClinVar: RCV000019090, RCV000020103, RCV000498537, RCV000685420, RCV001253100

In 2 brothers with a mild form of dominant myotonia, referred to as myotonia levior (see 160800), Lehmann-Horn et al. (1995) identified a 1655A-G transition in exon 15 of the CLCN1 gene, resulting in a gln552-to-arg (Q552R) substitution. Their affected mother also had the mutation. The findings indicated that myotonia levior is a variant or allelic form of Thomsen disease.

By functional expression studies, Ryan et al. (2002) demonstrated that CLCN1 channels with the Q552R mutation were formed normally, but had altered gating properties. Voltage dependence of the mutant channels was shifted by more than +90 mV compared to wildtype channels, resulting in a reduction in the open probability of the channel.


.0008   MYOTONIA CONGENITA, AUTOSOMAL DOMINANT

CLCN1, ILE290MET
SNP: rs80356690, gnomAD: rs80356690, ClinVar: RCV000019091, RCV000020117, RCV000626584, RCV000690053, RCV000711241, RCV001196224, RCV003398545

In affected members of a family with Thomsen myotonia (160800), Lehmann-Horn et al. (1995) identified an 870C-G transversion in the CLCN1 gene, resulting in an ile290-to-met (I290M) substitution.


.0009   MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, 14-BP DEL
SNP: rs768119034, gnomAD: rs768119034, ClinVar: RCV000395981, RCV000552780, RCV000664241, RCV000778142, RCV001262336, RCV001753742, RCV002259330, RCV003319191

In patients with Becker myotonia (255700), Meyer-Kleine et al. (1994) identified a 14-bp deletion in exon 13 of the CLCN1 gene.

In a family in which the index patient and his mother had been examined by Becker (1977), who classified their disorder as dominant myotonia congenita, Lehmann-Horn et al. (1995) found that the proband was homozygous for a 14-bp deletion (involving nucleotides 1437-1450) in exon 13 of the CLCN1 gene. Both nonmyotonic sons of the index patient were heterozygous for the deletion.

In French Canadian patients with recessive myotonia, Dupre et al. (2009) found that homozygosity for the 14-bp deletion resulted in a severe phenotype with generalized hypertrophy, moderate myotonia, and transient weakness.


.0010   MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, GLU291LYS
SNP: rs121912805, gnomAD: rs121912805, ClinVar: RCV000019093, RCV001041229, RCV001781282, RCV002468557

In 2 sibs with autosomal recessive myotonia congenita (255700), Pusch et al. (1995) identified compound heterozygosity for 2 mutations in the CLCN1 gene: glu291-to-lys (E291K) and arg894-to-ter (R894X; 118425.0015). Functional expression studies showed that the E291K channels did not yield currents between -140 and 100 millivolts, indicating that this mutation totally abolished channel activity. In contrast to the mutation in the neighboring amino acid (I290M; 118425.0008), which acts as a dominant as a result of interactions with wildtype monomers, the E291K mutation is recessive. Whereas the I290M mutant shifts the voltage dependence of chloride channels positive (via homomers or heteromers with wildtype subunits), the E291K mutation shows no evidence of interaction and does not shift the voltage dependence.


.0011   MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

MYOTONIA CONGENITA, AUTOSOMAL DOMINANT, INCLUDED
CLCN1, ARG317GLN
SNP: rs80356702, gnomAD: rs80356702, ClinVar: RCV000019094, RCV000019095, RCV000020121, RCV000516960, RCV000626585, RCV000763169

The arg317-to-gln (R317Q) mutation illustrates the resultant occurrence of either autosomal dominant myotonia congenita (160800) or autosomal recessive myotonia congenita (255700), depending on the family background. Meyer-Kleine et al. (1995) observed the R317Q mutation in dominant Thomsen myotonia congenita; Esteban et al. (1998) observed it in Becker myotonia congenita and pointed to other mutations that had been observed as a recessive or a dominant, depending on the particular family.


.0012   MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, GLY499ARG
SNP: rs121912807, gnomAD: rs121912807, ClinVar: RCV000019096, RCV001382414

In a boy with Becker myotonia congenita (255700), Zhang et al. (2000) identified a C-to-T transition at codon 1495 in exon 14 of the CLCN1 gene, resulting in a gly499-to-arg (G499R) substitution in the putative transmembrane domain 10 of the CLCN1 protein. In contrast to normal CLCN1 channels that deactivate upon hyperpolarization, functional expression of G499R CLCN1 yielded a hyperpolarization-activated chloride current when measured in the presence of a high intracellular chloride concentration. Current was abolished when measured with a physiologic chloride transmembrane gradient. Electrophysiologic analysis of other gly449 mutants suggested that the positive charge introduced by the G499R mutation may be responsible for this unique gating behavior.


.0013   MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, 1-BP INS, 831G
SNP: rs140026363, gnomAD: rs140026363, ClinVar: RCV000019097, RCV001382411

In 2 African American sibs with autosomal recessive myotonia congenita (255700), Nagamitsu et al. (2000) identified compound heterozygosity for 2 mutations in the CLCN1 gene: a 1-bp insertion (831insG) in exon 7, resulting in a premature termination codon at codon 289, and a C-to-T transition in exon 23, resulting in a pro932-to-leu (P932L; 118425.0014) substitution. In addition to generalized myotonia and proximal muscle hypertrophy, the sibs also displayed progressive muscle weakness, joint contractures, and dystrophic muscle pathology.


.0014   MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, PRO932LEU
SNP: rs80356706, gnomAD: rs80356706, ClinVar: RCV000019085, RCV000020108, RCV000478940, RCV000638250

For discussion of the pro932-to-leu (P932L) mutation in the CLCN1 gene that was found in compound heterozygous state in sibs with autosomal recessive myotonia congenita (255700) by Nagamitsu et al. (2000), see 118425.0013.


.0015   MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

MYOTONIA CONGENITA, AUTOSOMAL DOMINANT, INCLUDED
CLCN1, ARG894TER
SNP: rs55960271, gnomAD: rs55960271, ClinVar: RCV000019098, RCV000019099, RCV000020107, RCV000292791, RCV000626582, RCV000627759, RCV001564017, RCV001794458, RCV001813999, RCV004515784

For discussion of the arg894-to-ter (R894X) mutation in the CLCN1 gene that was found in compound heterozygous state in sibs with autosomal recessive myotonia congenita (255700) by Pusch et al. (1995), see 118425.0010.

In patients with both autosomal recessive and autosomal dominant (160800) myotonia congenita, Meyer-Kleine et al. (1995) identified a mutation in the CLCN1 gene, resulting in an R894X substitution. Functional expression of the R894X mutant in Xenopus oocytes revealed a large reduction, but not complete abolition, of chloride currents. Further, it had a weak dominant-negative effect on wildtype currents in coexpression studies. Reduction of currents predicted for heterozygous carriers were close to the borderline value, sufficient to elicit myotonia.

Sun et al. (2001) found a carrier frequency of 0.87% for the R894X mutation in the northern Scandinavian population.

In a French Canadian family with myotonia, Dupre et al. (2009) found that the R894X mutation could be expressed in a semidominant or recessive manner. The proband, who was heterozygous for the R894X mutation, had muscle stiffness and mild warm-up phenomenon, but no significant percussion myotonia and no myotonia on EMG. In contrast, her daughters, who were compound heterozygous for the R894X and R300X (118425.0017) mutations, showed a moderately severe phenotype with generalized hypertrophy consistent with autosomal recessive Becker myotonia. This compound heterozygous genotype showed resistance to phenytoin, mexiletine, and quinine.


.0016   MYOTONIA CONGENITA, AUTOSOMAL DOMINANT

CLCN1, MET128VAL
SNP: rs80356699, ClinVar: RCV000019100, RCV000020109, RCV001049292

In affected members of a family with autosomal dominant myotonia congenita (160800), Colding-Jorgensen et al. (2003) identified a heterozygous A-to-G transition in the CLCN1 gene, resulting in a met128-to-val (M128V) substitution.


.0017   MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, ARG300TER
SNP: rs1586496726, ClinVar: RCV000019101

For discussion of the arg300-to-ter (R300X) mutation in the CLCN1 gene that was found in compound heterozygous state in patients with autosomal recessive myotonia congenita (255700) by Dupre et al. (2009), see 118425.0015.


.0018   MYOTONIA CONGENITA, AUTOSOMAL DOMINANT

CLCN1, SER189PHE
SNP: rs121912810, ClinVar: RCV000019102, RCV003387727

In 2 French Canadian women with autosomal dominant myotonia congenita (160800), Dupre et al. (2009) identified a heterozygous ser189-to-phe (S189F) substitution in the CLCN1 gene. Both women had mild fluctuating myotonia and mild muscle hypertrophy and reported aggravation of symptoms with menstrual periods and pregnancy.


.0019   MYOTONIA CONGENITA, AUTOSOMAL DOMINANT

MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE, INCLUDED
CLCN1, TRP433ARG
ClinVar: RCV000019103, RCV000019104

Dupre et al. (2009) identified a heterozygous or homozygous trp433-to-arg (W433R) substitution in the CLCN1 gene in French Canadian patients with autosomal dominant (160800) and recessive (255700) myotonia, respectively. The recessive phenotype was characterized by severe myotonia, dysphagia, generalized hypertrophy predominant in lower limb muscles, and the warm-up phenomenon.


.0020   MYOTONIA CONGENITA, AUTOSOMAL RECESSIVE

CLCN1, DUP, EX8-14
ClinVar: RCV000033240

In 2 unrelated Iraqi women with severe infantile-onset myotonia congenita (255700), Raja Rayan et al. (2012) identified a homozygous duplication of exons 8 through 14 of the CLCN1 gene, predicted to result in a frameshift. Both patients came from consanguineous families and had multiple affected family members. The duplication was identified by multiplex ligation-dependent probe amplification (MLPA) specific to the CLCN1 gene, and was not found in 124 control chromosomes or in a variant database. Haplotype analysis suggested a founder effect for this duplication mutation. Raja Rayan et al. (2012) concluded that copy number variation involving the CLCN1 gene is an important genetic mechanism in patients with recessive myotonia congenita, and that MLPA analysis may aid in genetic counseling.


.0021   VARIANT OF UNKNOWN SIGNIFICANCE

CLCN1, ARG976TER
SNP: rs142539932, gnomAD: rs142539932, ClinVar: RCV000180791, RCV000195160, RCV000535831, RCV003488433

This variant is classified as a variant of unknown significance because its contribution to idiopathic generalized epilepsy (EIG; see 600669) has not been confirmed.

In a 26-year-old Italian woman with EIG, Chen et al. (2013) identified a de novo heterozygous c.2926C-T transition (c.2926C-T, NM_000083) in exon 23 of the CLCN1 gene, resulting in an arg976-to-ter (R976X) substitution and truncation of the distal C terminus by 12 residues. The mutation, which was found by exome sequencing of 291 patients with epilepsy, was not found in the dbSNP or 1000 Genomes Project databases. The patient had generalized pharmacoresistant epilepsy with mixed seizure types, including generalized tonic-clonic and absence seizures. Her first seizure occurred during a febrile illness at age 11 months. EEG repeatedly showed generalized spike-wave discharges. In addition, she presented with myotonic writer's cramp at age 9 years, which subsequently resolved. EMG as an adult showed no evidence of myotonic discharges. Brain MRI and cognitive function were normal. Functional studies of the variant were not performed, but Chen et al. (2013) found expression of the CLCN1 gene in multiple human brain regions, and suggested that the mutation identified in this patient could result in neuronal hyperexcitability, thus contributing to the seizure phenotype.


.0022   MYOTONIA CONGENITA, AUTOSOMAL DOMINANT

CLCN1, GLN412PRO
SNP: rs1279658001, gnomAD: rs1279658001, ClinVar: RCV001382412, RCV002251767

In a Costa Rican patient (family 4-CR14) with autosomal dominant myotonia congenita (160800), Vindas-Smith et al. (2016) identified a heterozygous c.1235A-C transversion in exon 11 of the CLCN1 gene, resulting in a gln412-to-pro (Q412P) substitution. The mutation, which was identified by bidirectional sequencing of the CLCN1 gene and confirmed by RFLP-PCR, was also found in the patient's asymptomatic (self-reported) mother and 2 sibs. Functional studies of CLCN1 with the Q412P mutation expressed in Xenopus oocytes showed reduced surface expression and reduced current density but did not display a dominant-negative effect when coexpressed with wildtype CLCN1.


REFERENCES

  1. Adrian, R. H., Bryant, S. H. On the repetitive discharge in myotonic muscle fibres. J. Physiol. 240: 505-515, 1974. [PubMed: 4420758] [Full Text: https://doi.org/10.1113/jphysiol.1974.sp010620]

  2. Altamura, C., Lucchiari, S., Sahbani, D., Ulzi, G., Comi, G. P., D'Ambrosio, P., Petillo, R., Politano, L., Vercelli, L., Mongini, T., Dotti, M. T., Cardani, R., Meola, G., Lo Monaco, M., Matthews, E., Hanna, M. G., Carratu, M. R., Conte, D., Imbrici, P., Desaphy, J.-F. The analysis of myotonia congenita mutations discloses functional clusters of amino acids within the CBS2 domain and the C-terminal peptide of the CLC-1 channel. Hum. Mutat. 39: 1273-1283, 2018. [PubMed: 29935101] [Full Text: https://doi.org/10.1002/humu.23581]

  3. Aminoff, M. J., Layzer, R. B., Satya-Murti, S., Faden, A. I. The declining electrical response of muscle to repetitive nerve stimulation in myotonia. Neurology 27: 812-816, 1977. [PubMed: 561337] [Full Text: https://doi.org/10.1212/wnl.27.9.812]

  4. Beck, C. L., Fahlke, C., George, A. L., Jr. Molecular basis for decreased muscle chloride conductance in the myotonic goat. Proc. Nat. Acad. Sci. 93: 11248-11252, 1996. [PubMed: 8855341] [Full Text: https://doi.org/10.1073/pnas.93.20.11248]

  5. Becker, P. E. Myotonia Congenita and Syndromes associated with Myotonia: Clinical-Genetic Studies of the Nondystrophic Myotonias. Stuttgart: Georg Thieme (pub.) 1977.

  6. Bryant, S. H. Muscle membrane of normal and myotonic goats in normal and low external chloride. (Abstract) Fed. Proc. 21: 312 only, 1962.

  7. Charlet-B, N., Savkur, R. S., Singh, G., Philips, A. V., Grice, E. A., Cooper, T. A. Loss of the muscle-specific chloride channel in type 1 myotonic dystrophy due to misregulated alternative splicing. Molec. Cell 10: 45-53, 2002. [PubMed: 12150906] [Full Text: https://doi.org/10.1016/s1097-2765(02)00572-5]

  8. Chen, T. T., Klassen, T. L., Goldman, A. M., Marini, C., Guerrini, R., Noebels, J. L. Novel brain expression of ClC-1 chloride channels and enrichment of CLCN1 variants in epilepsy. Neurology 80: 1078-1085, 2013. [PubMed: 23408874] [Full Text: https://doi.org/10.1212/WNL.0b013e31828868e7]

  9. Colding-Jorgensen, E., Duno, M., Schwartz, M., Vissing, J. Decrement of compound muscle action potential is related to mutation type in myotonia congenita. Muscle Nerve 27: 449-455, 2003. [PubMed: 12661046] [Full Text: https://doi.org/10.1002/mus.10347]

  10. Duno, M., Colding-Jorgensen, E., Grunnet, M., Jespersen, T., Vissing, J., Schwartz, M. Difference in allelic expression of the CLCN1 gene and the possible influence on the myotonia congenita phenotype. Europ. J. Hum. Genet. 12: 738-743, 2004. [PubMed: 15162127] [Full Text: https://doi.org/10.1038/sj.ejhg.5201218]

  11. Dupre, N., Chrestian, N., Bouchard, J.-P., Rossignol, E., Brunet, D., Sternberg, D., Brias, B., Mathieu, J., Puymirat, J. Clinical, electrophysiologic, and genetic study of non-dystrophic myotonia in French-Canadians. Neuromusc. Disord. 19: 330-334, 2009. [PubMed: 18337100] [Full Text: https://doi.org/10.1016/j.nmd.2008.01.007]

  12. Dutzler, R., Campbell, E. B., Cadene, M., Chait, B. T., MacKinnon, R. X-ray structure of a CIC chloride channel at 3.0 angstrom reveals the molecular basis of anion selectivity. Nature 415: 287-294, 2002. [PubMed: 11796999] [Full Text: https://doi.org/10.1038/415287a]

  13. Esteban, J., Neumeyer, A. M., McKenna-Yasek, D., Brown, R. H. Identification of two mutations and a polymorphism in the chloride channel CLCN-1 in patients with Becker's generalized myotonia. Neurogenetics 1: 185-188, 1998. [PubMed: 10737121] [Full Text: https://doi.org/10.1007/s100480050027]

  14. Estevez, R., Schroeder, B. C., Accardi, A., Jentsch, T. J., Pusch, M. Conservation of chloride channel structure revealed by an inhibitor binding site in CIC-1. Neuron 38: 47-59, 2003. [PubMed: 12691663] [Full Text: https://doi.org/10.1016/s0896-6273(03)00168-5]

  15. Fahlke, C., Beck, C. L., George, A. L., Jr. A mutation in autosomal dominant myotonia congenita affects pore properties of the muscle chloride channel. Proc. Nat. Acad. Sci. 94: 2729-2734, 1997. [PubMed: 9122265] [Full Text: https://doi.org/10.1073/pnas.94.6.2729]

  16. George, A. L., Jr., Crackower, M. A., Abdalla, J. A., Hudson, A. J., Ebers, G. C. Molecular basis of Thomsen's disease (autosomal dominant myotonia congenita). Nature Genet. 3: 305-310, 1993. [PubMed: 7981750] [Full Text: https://doi.org/10.1038/ng0493-305]

  17. George, A. L., Jr. Personal Communication. Nashville, Tenn. 4/14/1997.

  18. Koch, M. C., Meyer-Kleine, C., Otto, M., Ricker, K., Lorenz, C., Steinmeyer, K., Jentsch, T. J. Mutations in the CLCN1 gene leading to myotonia congenita Thomsen and generalized myotonia Becker. (Abstract) Am. J. Hum. Genet. 55 (suppl.): A226, 1994.

  19. Koch, M. C., Ricker, K., Otto, M., Wolf, F., Zoll, B., Lorenz, C., Steinmeyer, K., Jentsch, T. J. Evidence for genetic homogeneity in autosomal recessive generalised myotonia (Becker). J. Med. Genet. 30: 914-917, 1993. [PubMed: 8301644] [Full Text: https://doi.org/10.1136/jmg.30.11.914]

  20. Koch, M. C., Steinmeyer, K., Lorenz, C., Ricker, K., Wolf, F., Otto, M., Zoll, B., Lehmann-Horn, F., Grzeschik, K.-H., Jentsch, T. J. The skeletal muscle chloride channel in dominant and recessive human myotonia. Science 257: 797-800, 1992. [PubMed: 1379744] [Full Text: https://doi.org/10.1126/science.1379744]

  21. Koty, P. P., Marks, H. G., Turel, A., Flagler, D., Angelini, C., Pegoraro, E., Vancott, A. C., Manchester, D., Zonana, J., Bird, T. D., Hoffman, E. P. Linkage analysis of Thomsen and Becker myotonia families. (Abstract) Am. J. Hum. Genet. 55 (suppl.): A227, 1994.

  22. Kubisch, C., Schmidt-Rose, T., Fontaine, B., Bretag, A. H., Jentsch, T. J. ClC-1 chloride channel mutations in myotonia congenita: variable penetrance of mutations shifting the voltage dependence. Hum. Molec. Genet. 7: 1753-1760, 1998. [PubMed: 9736777] [Full Text: https://doi.org/10.1093/hmg/7.11.1753]

  23. Lehmann-Horn, F., Mailander, V., Heine, R., George, A. L. Myotonia levior is a chloride channel disorder. Hum. Molec. Genet. 4: 1397-1402, 1995. [PubMed: 7581380] [Full Text: https://doi.org/10.1093/hmg/4.8.1397]

  24. Lipicky, R. J., Bryant, S. H. Sodium, potassium, and chloride fluxes in intercostal muscle from normal goats and goats with hereditary myotonia. J. Gen. Physiol. 50: 89-111, 1966. [PubMed: 5971035] [Full Text: https://doi.org/10.1085/jgp.50.1.89]

  25. Lorenz, C., Meyer-Kleine, C., Steinmeyer, K., Koch, M. C., Jentsch, T. J. Genomic organization of the human muscle chloride channel CLC-1 and analysis of novel mutations leading to Becker-type myotonia. Hum. Molec. Genet. 3: 941-946, 1994. [PubMed: 7951242] [Full Text: https://doi.org/10.1093/hmg/3.6.941]

  26. Mailander, V., Heine, R., Deymeer, F., Lehmann-Horn, F. Novel muscle chloride channel mutations and their effects on heterozygous carriers. Am. J. Hum. Genet. 58: 317-324, 1996. [PubMed: 8571958]

  27. Mankodi, A., Takahashi, M. P., Jiang, H., Beck, C. L., Bowers, W. J., Moxley, R. T., Cannon, S. C., Thornton, C. A. Expanded CUG repeats trigger aberrant splicing of ClC-1 chloride channel pre-mRNA and hyperexcitability of skeletal muscle in myotonic dystrophy. Molec. Cell 10: 35-44, 2002. [PubMed: 12150905] [Full Text: https://doi.org/10.1016/s1097-2765(02)00563-4]

  28. Meyer-Kleine, C., Ricker, K., Otto, M., Koch, M. C. A recurrent 14 bp deletion in the CLCN1 gene associated with generalized myotonia (Becker). Hum. Molec. Genet. 3: 1015-1016, 1994. [PubMed: 7951215] [Full Text: https://doi.org/10.1093/hmg/3.6.1015]

  29. Meyer-Kleine, C., Steinmeyer, K., Ricker, K., Jentsch, T. J., Koch, M. C. Spectrum of mutations in the major human skeletal muscle chloride channel gene (CLCN1) leading to myotonia. Am. J. Hum. Genet. 57: 1325-1334, 1995. [PubMed: 8533761]

  30. Mindell, J. A., Maduke, M., Miller, C., Grigorieff, N. Projection structure of a CIC-type chloride channel at 6.5-angstrom resolution. Nature 409: 219-223, 2001. [PubMed: 11196649] [Full Text: https://doi.org/10.1038/35051631]

  31. Nagamitsu, S., Matsuura, T., Khajavi, M., Armstrong, R., Gooch, C., Harati, Y., Ashizawa, T. A 'dystrophic' variant of autosomal recessive myotonia congenita caused by novel mutations in the CLCN1 gene. Neurology 55: 1697-1703, 2000. [PubMed: 11113225] [Full Text: https://doi.org/10.1212/wnl.55.11.1697]

  32. Pusch, M., Steinmeyer, K., Koch, M. C., Jentsch, T. J. Mutations in dominant human myotonia congenita drastically alter the voltage dependence of the CIC-1 chloride channel. Neuron 15: 1455-1463, 1995. [PubMed: 8845168] [Full Text: https://doi.org/10.1016/0896-6273(95)90023-3]

  33. Pusch, M. Myotonia caused by mutations in the muscle chloride channel gene CLCN1. Hum. Mutat. 19: 423-434, 2002. [PubMed: 11933197] [Full Text: https://doi.org/10.1002/humu.10063]

  34. Raja Rayan, D. L., Haworth, A., Sud, R., Matthews, E., Fialho, D., Burge, J., Portaro, S., Schorge, S., Tuin, K., Lunt, P., McEntagart, M., Toscano, A., Davis, M. B., Hanna, M. G. A new explanation for recessive myotonia congenita: exon deletions and duplications in CLCN1. Neurology 78: 1953-1958, 2012. [PubMed: 22649220] [Full Text: https://doi.org/10.1212/WNL.0b013e318259e19c]

  35. Rudel, R. The myotonic mouse--a realistic model for the study of human recessive generalized myotonia. Trends Neurosci. 13: 1-3, 1990. [PubMed: 1688667] [Full Text: https://doi.org/10.1016/0166-2236(90)90049-g]

  36. Ryan, A., Rudel, R., Kuchenbecker, M., Fahlke, C. A novel alteration of muscle chloride channel gating in myotonia levior. J. Physiol. 545: 345-354, 2002. [PubMed: 12456816] [Full Text: https://doi.org/10.1113/jphysiol.2002.027037]

  37. Steinmeyer, K., Klocke, R., Ortland, C., Gronemeier, M., Jockusch, H., Grunder, S., Jentsch, T. J. Inactivation of muscle chloride channel by transposon insertion in myotonic mice. Nature 354: 304-308, 1991. [PubMed: 1659665] [Full Text: https://doi.org/10.1038/354304a0]

  38. Steinmeyer, K., Lorenz, C., Pusch, M., Koch, M. C., Jentsch, T. J. Multimeric structure of ClC-1 chloride channel revealed by mutations in dominant myotonia congenita (Thomsen). EMBO J. 13: 737-743, 1994. [PubMed: 8112288] [Full Text: https://doi.org/10.1002/j.1460-2075.1994.tb06315.x]

  39. Suetterlin, K., Matthews, E., Sud, R., McCall, S., Fialho, D., Burge, J., Jayaseelan, D., Haworth, A., Sweeney, M. G., Kullmann, D. M., Schorge, S., Hanna, M. G., Mannikko, R. Translating genetic and functional data into clinical practice: a series of 223 families with myotonia. Brain 145: 607-620, 2022. [PubMed: 34529042] [Full Text: https://doi.org/10.1093/brain/awab344]

  40. Sun, C., Tranebjaerg, L., Torbergsen, T., Holmgren, G., Van Ghelue, M. Spectrum of CLCN1 mutations in patients with myotonia congenita in northern Scandinavia. Europ. J. Hum. Genet. 9: 903-909, 2001. Note: Erratum: Europ. J. Hum. Genet. 18: 264 only, 2010. [PubMed: 11840191] [Full Text: https://doi.org/10.1038/sj.ejhg.5200736]

  41. Thomasen, E. Myotonia, Thomsen's disease. Paramyotonia, and dystrophia myotonica. Op. Ex Domo Biol. Hered. Hum. U. Hafniensis 17: 11-251, 1948.

  42. Thomsen, J. Tonische Kraempfe in willkuerlich beweglichen Muskeln in Folge von ererbter psychischer Disposition: Ataxia muscularis? Arch. Psychiat. Nervenkr. 6: 702-718, 1876.

  43. Vindas-Smith, R., Fiore, M., Vasquez, M., Cuenca, P., del Valle, G., Lagostena, L., Gaitan-Penas, H., Estevez, R., Pusch, M., Morales, F. Identification and functional characterization of CLCN1 mutations found in nondystrophic myotonia patients. Hum. Mutat. 37: 74-83, 2016. [PubMed: 26510092] [Full Text: https://doi.org/10.1002/humu.22916]

  44. Wollnik, B., Kubisch, C., Steinmeyer, K., Pusch, M. Identification of functionally important regions of the muscular chloride channel CIC-1 by analysis of recessive and dominant myotonic mutations. Hum. Molec. Genet. 6: 805-811, 1997. [PubMed: 9158157] [Full Text: https://doi.org/10.1093/hmg/6.5.805]

  45. Wu, H., Olson, E. N. Activation of the MEF2 transcription factor in skeletal muscles from myotonic mice. J. Clin. Invest. 109: 1327-1333, 2002. [PubMed: 12021248] [Full Text: https://doi.org/10.1172/JCI15417]

  46. Zhang, J., Sanguinetti, M. C., Kwiecinski, H., Ptacek, L. J. Mechanism of inverted activation of ClC-1 channels caused by a novel myotonia congenita mutation. J. Biol. Chem. 275: 2999-3005, 2000. [PubMed: 10644771] [Full Text: https://doi.org/10.1074/jbc.275.4.2999]


Contributors:
Hilary J. Vernon - updated : 05/27/2022
Cassandra L. Kniffin - updated : 6/16/2015
Cassandra L. Kniffin - updated : 3/11/2013
Cassandra L. Kniffin - updated : 10/27/2009
Cassandra L. Kniffin - updated : 6/20/2005
Patricia A. Hartz - updated : 5/4/2004
Victor A. McKusick - updated : 2/6/2003
Patricia A. Hartz - updated : 10/8/2002
Stylianos E. Antonarakis - updated : 9/10/2002
Michael B. Petersen - updated : 8/19/2002
Victor A. McKusick - updated : 4/12/2002
Ada Hamosh - updated : 1/17/2002
Kathryn R. Wagner - updated : 8/6/2001
Ada Hamosh - updated : 1/9/2001
Ada Hamosh - updated : 2/11/2000
Victor A. McKusick - updated : 5/6/1998
Victor A. McKusick - updated : 6/23/1997
Victor A. McKusick - updated : 5/16/1997
Victor A. McKusick - updated : 4/21/1997
Orest Hurko - updated : 3/9/1996

Creation Date:
Victor A. McKusick : 9/29/1992

Edit History:
carol : 06/01/2022
carol : 05/27/2022
carol : 11/22/2019
carol : 11/21/2019
carol : 06/21/2019
alopez : 10/12/2016
alopez : 02/24/2016
carol : 6/26/2015
mcolton : 6/17/2015
ckniffin : 6/16/2015
mgross : 10/7/2013
alopez : 3/12/2013
ckniffin : 3/11/2013
ckniffin : 10/31/2011
wwang : 11/16/2009
ckniffin : 10/27/2009
terry : 6/23/2006
carol : 6/23/2006
carol : 6/29/2005
ckniffin : 6/20/2005
terry : 3/14/2005
mgross : 5/4/2004
carol : 2/11/2003
tkritzer : 2/11/2003
terry : 2/6/2003
mgross : 10/8/2002
mgross : 9/10/2002
mgross : 9/10/2002
mgross : 9/10/2002
alopez : 8/19/2002
alopez : 8/19/2002
alopez : 4/25/2002
cwells : 4/22/2002
terry : 4/12/2002
alopez : 1/22/2002
terry : 1/17/2002
carol : 8/6/2001
carol : 8/6/2001
mgross : 1/10/2001
terry : 1/9/2001
alopez : 2/15/2000
terry : 2/11/2000
carol : 11/9/1999
carol : 7/7/1999
carol : 6/8/1998
carol : 5/12/1998
terry : 5/6/1998
mark : 7/9/1997
alopez : 6/27/1997
terry : 6/23/1997
terry : 6/18/1997
mark : 6/9/1997
terry : 6/5/1997
terry : 6/5/1997
mark : 5/16/1997
terry : 5/10/1997
terry : 5/6/1997
jenny : 4/21/1997
terry : 4/14/1997
mark : 2/23/1997
mark : 12/18/1996
jamie : 12/6/1996
terry : 12/4/1996
terry : 4/15/1996
mark : 3/9/1996
terry : 2/23/1996
mark : 2/22/1996
terry : 2/19/1996
mark : 1/19/1996
mark : 12/18/1995
joanna : 12/15/1995
mark : 12/15/1995
mark : 12/15/1995
terry : 12/14/1995
terry : 12/13/1995
terry : 10/30/1995
mark : 9/7/1995
carol : 1/23/1995
jason : 7/27/1994
carol : 12/20/1993
carol : 4/29/1993