Back To Search Results

Charcot-Marie-Tooth Disease

Editor: Arun B Taly Updated: 6/22/2024 1:33:54 AM

Introduction

Charcot-Marie-Tooth (CMT) disease, also known as hereditary motor and sensory neuropathies, is a group of inherited motor and sensory peripheral neuropathies and the most common inherited neuromuscular disorder. CMT was first described as a clinical entity in 1886 by physicians Jean-Marie Charcot, Pierre Marie, and Howard Henry Tooth, who referred to it as "progressive muscular atrophy."[1] CMT is heterogeneous in its clinical, electrophysiological, genetic, and pathological features. Based on neurophysiological findings, CMT was previously classified into the demyelinating form, CMT type 1 (CMT1), and the axonal form, CMT type 2 (CMT2). However, the increasing use of next-generation gene sequencing (NGS) technologies has altered the classification and diagnosis of CMT.[2][3][4]

CMT is a nerve-length-dependent disorder characterized by slowly progressive foot deformities (most often pes cavus), sensory loss, weakness in the lower extremities, and reduced or absent deep tendon reflexes. Most individuals with CMT exhibit symptoms in the first or second decade of life, with an insidious onset of weakness that begins in the lower extremities and later involves the upper extremities.[5][6] The presentation of CMT can overlap with other neurodegenerative disorders.

Etiology

Register For Free And Read The Full Article
Get the answers you need instantly with the StatPearls Clinical Decision Support tool. StatPearls spent the last decade developing the largest and most updated Point-of Care resource ever developed. Earn CME/CE by searching and reading articles.
  • Dropdown arrow Search engine and full access to all medical articles
  • Dropdown arrow 10 free questions in your specialty
  • Dropdown arrow Free CME/CE Activities
  • Dropdown arrow Free daily question in your email
  • Dropdown arrow Save favorite articles to your dashboard
  • Dropdown arrow Emails offering discounts

Learn more about a Subscription to StatPearls Point-of-Care

Etiology

CMT and related disorders are genetically determined, involving more than 100 genes. About 80% to 90% of the genetic abnormalities are due to copy number variations (CNVs) in peripheral myelin protein 22 (PMP22) and mutations in GJB1, MPZ, and MFN2 genes. Only a few families are affected by mutations in other genes, and 18% to 50% of cases remain unsolved, possibly due to mutations in noncoding DNA or structural variations. Structural variations describe genomic rearrangements that disrupt chromosomal organization and architecture, including duplications and deletions (also referred to as CNV), insertions, inversions, and translocations.[7][8][9][10]

Autosomal dominant CMT is the most common genetic subtype, followed by X-linked CMT, while autosomal recessive forms are rare.

About 90% of mutations in individuals with CMT type 1 are caused by 3 genetic defects—PMP22 duplication, GJB1 point mutations, and MPZ point mutations—which cause damage to Schwann cells and the myelin sheath of the axon. CMT type 1A, the most common CMT neuropathy, is caused by a 1.5-Mb duplication of chromosome 17p11.2, resulting in trisomy of PMP22. Due to its large size, this region is susceptible to frequent genomic rearrangements. Duplication or deletion in PMP22 leads to disease via a gene dosage effect, given the sensitivity of nervous tissue to the gain or loss of a copy of the PMP22 gene.[10] 

Duplication in the heterozygous state results in 1.5-fold overexpression of PMP22 in Schwann cells, while homozygous duplication causes 2-fold overexpression. This genomic duplication results from an unequal meiotic crossover facilitated in the male germline by flanking homologous repeat sequences. In laboratory models, this overexpression overloads the proteasome system, leading to cytoplasmic aggregation of the ubiquitinated PMP22 protein, recruitment of autophagosomes and lysosomes, and increased autophagy.[11][12][13][14] 

Frameshift mutations in MPZ lead to mutant proteins aggregating in the endoplasmic reticulum of the neuron and consequent apoptosis. Other variations of CMT are usually the result of loss-of-function mutations in different genes and are uncommonly due to a toxic gain of function.[15]

Epidemiology

CMT is a clinically and genetically heterogeneous neuropathic disorder that occurs worldwide and affects individuals from all ethnic groups. Overall, CMT has an estimated frequency of 1 in 2500 people.[16] Epidemiological data from adult neuropathy clinics indicate that CMT is more common than inflammatory or paraneoplastic neuropathies.[17][18] The prevalence of CMT ranges from 9.7 per 100,000 population in Serbia to 82.3 per 100,000 population in Norway.[19] Published studies vary in quality and methodology, making it challenging to accurately assess the true prevalence of CMT due to the diverse clinical symptoms and presentations of the disease. This variability is evident in older population-based surveys, which reported a wide range of CMT frequencies, from 8 per 100,000 individuals in Libya to 28 per 100,000 individuals in northern Spain.[20] 

Based on median motor nerve conduction velocity (NCV), CMT is broadly categorized into demyelinating (CMT1) and axonal (CMT2) types. The frequency ranges from 37.5% to 84% for CMT1 and 12% to 35.9% for CMT2. However, prevalence studies focusing solely on CMT2 are rare, making the identification of affected individuals and families more challenging.[19] 

Pathophysiology

The normal structure and functions of peripheral nerves depend upon the close anatomical and physiological interaction between Schwann cells and axons. Axons regulate the survival, proliferation, and differentiation of Schwann cells, which in turn have a crucial role in regulating ion channels and their maintenance, supporting survival, and facilitating regeneration of axons. Abnormalities in genes involved in myelin assembly and axonal transport can lead to primary demyelination and axonopathy.

Some of the molecular and cellular mechanisms thought to be involved in inducing CMT are mentioned below.[21]

  • Myelin assembly: Genes such as MPZ (involved in myelin compaction), GJB1 (gap junction formation), and PMP22 (synthesis and maintenance of myelin) lead to impaired myelin sheath formation, the primary cause of demyelinating CMT.
  • Cytoskeletal structure: These genes are involved in actin polymerization (INF2), stabilization of the myelin sheath through membrane-protein interactions (PRX), intermediate filaments (NEFL), cell signaling (FGD4), and axonal transport (DYNC1H1).
  • Myelin-specific transcription factor: EGR2 is part of gene expression of myelin component proteins and Schwann cell differentiation.
  • Nuclear-related genes: These genes include those encoding nuclear envelope proteins (LMNA), nucleotide biosynthesis enzymes, DNA repair factors, and proteins involved in cell survival.
  • Endosomal sorting and cell signaling: These genes regulate vesicular transport, membrane trafficking, transport of intracellular organelles, and cell signaling. They include LITAF, MTMR2, SBF1, SBF2, SH3TC2, NDRG1, FIG4, RAB7, TFG, DNM2, and SIMPLE.
  • Proteasome and protein aggregation: These genes regulate microtubules (HSPB1 and HSPB8), cell adhesion (LRSAM1), and ubiquitin ligase (TRIM2).
  • Mitochondria: These genes are related to energy metabolism, mitochondrial dynamics, and ATP synthesis. MFN2 and GDAP1 are major contributors to CMT.
  • Gene mutations: These genes are responsible for motor proteins and axonal transport, tRNA synthetases, RNA metabolism, and ion channel-related abnormalities.

Due to the close functional interaction between Schwann cells and axons, demyelinating neuropathies in CMT often progress to functional axonopathies, leading to secondary axonal degeneration.[22][23] Common secondary phenomena include axonal loss, secondary Schwann cell proliferation, and accelerated pathology due to immune-mediated mechanisms.[24]

Evolving Classifications of Charcot-Marie-Tooth Disease

CMT is categorized based on the age of onset into early infantile (age less than 2), childhood (age 2-10), juvenile (age 10-20), adult (age 20-50), and late adult (age 50 or older) forms.[8] Additionally, based on electrophysiological findings, CMT is classified as either a demyelinating or axonal neuropathy. CMT exhibits autosomal dominant, autosomal recessive, or X-linked patterns of inheritance, with autosomal dominant being the most common.[25] In the 1990s, reflecting evolving electrophysiological and pathological findings alongside key clinical features, CMT was renamed hereditary motor and sensory neuropathy.

With the wider availability of genetic testing, the classification of CMT has evolved, as listed below.

  • CMT1: CMT1 accounts for 50% of cases and is characterized by demyelinating pathology, an autosomal dominant mode of inheritance, early onset, distal motor weakness, and moderate slowing of NCVs (>15 to ≤35 m/s). The most common subtype, CMT1A, results from a mutation in the PMP22 gene, while CMT1B is caused by mutations in MPZ
  • CMT2: CMT2 represents 15% to 30% of cases and is characterized by axonal pathology, an autosomal dominant mode of inheritance, onset typically in the second or third decade, distal motor weakness, and normal or slightly slowed NCVs (>40 m/s). The most common gene mutation associated with CMT2 is in MFN2.
  • CMTX: CMTX accounts for 10% to 15% of cases and is characterized by both demyelinating and axonal pathology. CMTX follows an X-linked mode of inheritance with onset typically in the first or second decade. Symptoms are generally more severe in males than females, featuring gait impairment and mild slowing of NCVs (>35 to ≤45 m/s). CMT1X is caused by mutations in the Cx32 gene.
  • CMT3: CMT3 is uncommon and characterized by demyelinating pathology with autosomal dominant inheritance. CMT3 has its onset in infancy with symptoms such as hypotonia, feeding difficulties, severe sensory and motor impairments, and profound slowing of NCVs (≤15 m/s). Most cases of CMT3 are caused by point mutations in the genes encoding PMP­22, MPZ, or ERG­2.
  • CMT4: CMT4 accounts for less than 10% of cases and is characterized by demyelinating pathology with an autosomal recessive mode of inheritance. CMT4 presents with progressively severe sensory and motor symptoms and moderate slowing of NCVs (>15 to ≤35 m/s). CMT4 is genetically heterogeneous.

Significant phenotypic heterogeneity exists with mutations that can cause both demyelinating and axonal forms of CMT. Similarly, mutations in MFN2, MPZ, GDAP1, and EGR2 may have CMTs with autosomal dominant or recessive inheritance patterns.[26][27][2]

Histopathology

After obtaining informed consent, a nerve biopsy can help identify the underlying genetic etiology in sporadic cases and also help distinguish CMT from acquired disorders such as chronic inflammatory demyelinating polyradiculoneuropathy (CIDP). Nerve biopsy may also provide a functional association when genetic tests detect "variants of uncertain significance" or a novel variant. Morphological and ultrastructural changes in axons, myelin, nodes of Ranvier, and mitochondria provide insights into the functions of mutated genes and the pathways contributing to disease pathogenesis.

Nonspecific histological changes may include axonal loss, demyelination or remyelination, and features indicative of inflammation. Chronic demyelination and remyelination can lead to concentric Schwann cell cytoplasm or basal lamina proliferation, forming multilayered "onion bulbs."[28] In CMT1, demyelination remains stable while axonal loss progresses over time. Occasionally, nerve biopsy in CMT1A may reveal "tomacula"—a characteristic feature of hereditary neuropathy with liability to pressure palsies (HNPP).[28] Tomacula are multifocal hypermyelinating processes that appear longitudinally like a "chain of sausages."[29] The presence of infiltration by inflammatory cells (lymphocytes and macrophages) may lead to a misdiagnosis of CIDP, particularly in cases associated with PMP22, MPZ, GJB1, and GDAP1 neuropathies. Some hypotheses explaining these inflammatory infiltrates include greater susceptibility to inflammation in CMT1A, the involvement of immune cells in genetically mediated demyelination, and superimposed CIDP.[30][31][32]

MPZ-associated neuropathy shows loss of compaction of myelin sheath layers, dissociation of the paired intra-period lines, regular widening between major dense lines, and irregularly uncompacted myelin sheaths.[29][33] Mutations in MPZ, MTMR2, MTMR12, MTMR5, SBF2, FGD4, SH3TC2, PRX, and NEFL can cause unusually thickened and folded myelin with redundant myelin loops.[29][34][35][36][37][38] These foldings differ from typical tomacula, characterized by focal hypermyelination and smooth external contours.[39][40][37][41] 

SH3TC2-associated neuropathy exhibits abnormal extensions of Schwann cell cytoplasm and the formation of onion bulbs characterized by concentric proliferations of the basal lamina around both myelinated and unmyelinated fibers.[42][38] INF2-neuropathy shows unusual whorl-like proliferation and supernumerary elongated extensions of Schwann cell cytoplasm resembling filopodia, along with prominent axoplasmic reticulum and nodal widening.[43][28] Disruption of Cajal bands with focal hypermyelination is noted in PRX mutations.[44][45]

Mitochondria in MFN2 and GDAP1 mutations exhibit abnormal swelling, rounding, and vacuolation, often with accumulation of amorphous material and loss of cristae.[29] Uncommonly, ultrastructural observations may show an absence of transverse bands and the widening of the paranodal junctional gap between myelin loops and axolemma in CNTNAP1 mutations.[46][47][48] In NDRG1-associated neuropathy (CMT4D), a characteristic feature is the presence of pleomorphic granular deposits, sometimes containing filaments or vesicles filled with glycogen, in the adaxonal space of myelinated fibers. NEFL-associated neuropathies exhibit abnormally condensed unmyelinated fibers with aggregates containing glycogen granules and dense microtubules. Mutations in SH3TC2 and NEFL may lead to the formation of "giant axons."[29] In congenital amyelinating neuropathy due to EGR2 mutations, there is a total absence of myelin with normal axons.[49]

History and Physical

History of Present Illness

The history of the current illness includes assessments of pain, weakness, loss of sensation, foot deformities, and muscle cramps. Symptoms include difficulty walking fast or running, tripping, falls, and twisting or spraining ankles. Young patients with CMT may experience delayed motor milestones. During childhood, these patients may appear clumsy and struggle with athletic activities.

Since the onset of CMT is insidious and progression is usually slow, determining the exact age of onset is sometimes difficult. Affected patients may have foot deformities, delayed motor milestones, difficulty walking, and impaired sensation.[50][51] Severe phenotypes such as Dejerine-Sottas syndrome may manifest with neonatal hypotonia, difficulty in feeding, hip dysplasia, and respiratory compromise.[52][53]

Patients have varying severities of CMT, spanning a spectrum from minimally symptomatic to severe.[54] Concomitant illnesses such as hypothyroidism, diabetes, obesity, and toxin exposure increase impairment in these patients.[55] Early or infantile-onset CMT usually correlates with worse disability.[56]

Family History 

The family history should include inquiries about symptoms of neuropathy in relatives. Patients exhibit intra-familial and interfamilial phenotypic heterogeneity in age at onset, motor deficits, and sensory loss. This variability is observed even in families sharing the same genetic abnormalities.[57] For example, point mutations in PMP22 may manifest as classical CMT1A, HNPP, Dejerine-Sottas syndrome, or congenital hypomyelinating neuropathy. Alternatively, they may remain asymptomatic, with only minor abnormalities detected on electrophysiological testing.[58][59] 

Physical Examination 

The physical examination must include a complete neurological assessment. Common clinical features across all variations of CMT are distal symmetrical weakness of the feet and legs, atrophy of weak muscles, reduction or loss of tendon reflexes, and skeletal deformities. Assessing gait is crucial, as weakness progression in CMT patients may lead to foot drop and a high-stepping gait. Onset in the upper limbs with proximal muscle weakness is rare. Hand weakness manifests as difficulty in buttoning, zipping, and writing. About 20% to 30% of patients with CMT complain of musculoskeletal pain rather than neuropathic pain. Paresthesias and positive sensory symptoms are infrequent. 

Physical examination may reveal foot deformities such as pes cavus (high arches) or hammer toes and subtle wasting of hypothenar muscles in patients with long-standing disease due to the weakness of the intrinsic muscles of the hand. Atrophy of the lower legs and distal thighs may cause the "stork leg deformity." Spinal deformities (scoliosis) may also occur. Early and severe scoliosis is suggestive of, but is not an exclusive feature of, SH3TC2 mutations.[38] Abnormal gait and instability occur due to weakness, proprioceptive loss, and skeletal deformities such as pes cavus and hammer toes.[25] These clinical manifestations result from axonal loss, even in demyelinating CMT, as these patients have slowed NCVs before the onset of clinical manifestations.[60][61][62][54] Pyramidal signs may occur, which is associated with MFN2 mutations.[63]

In a study involving 49 patients with genetically confirmed CMT, 88% had at least 1 additional feature, and 65% had 2 or more additional symptoms. This may provide clues for underlying genetic variations.[64] Patients with CNVs in PMP22 and mutations in MPZ and GJB1 may exhibit pupillary abnormalities, including tonic pupils, miosis, mydriasis, anisocoria, and impaired pupillary reaction.[65][66][64] 

Additional ocular manifestations of CMT include ptosis, optic atrophy, early cataract, glaucoma, and age-related macular degeneration. Cranial neuropathies, such as deafness, vocal cord palsy, facial or bulbar weakness, and tongue atrophy, can also occur in CMT, although rarely. Patients with PMP22 and MPZ-associated neuropathies may report pain and paresthesias. Patients with Thr124Met mutations in MPZ may present with a late-onset phenotype with pupillary abnormalities, shooting pains, disturbing paresthesias, and sometimes rapid progression.[65] In Hungarian patients, deafness and autoimmune disorders were often associated with PMP22 duplication.[9]

Features of dysautonomia, including urinary urgency and incontinence, orthostatic hypotension, and hyperhidrosis, have also been reported in CMT.[64] Patients with mutations in INF2 may develop focal segmental glomerulosclerosis, which can progress rapidly to end-stage renal disease.[67][68] Restless legs syndrome is more prevalent in patients with CMT as compared to the general population.[69][64] Axonal loss, irrespective of the CMT subtype, increases axonal excitability in the primary sensory units of leg muscles and leads to creeping sensations.[70] Other rare features include hyperkeratosis, skin hyperlaxity, arthrogryposis, impaired cognition, learning disability, lip or chin myokymia, respiratory insufficiency, tremor, nystagmus, ataxia, fasciculations, cold-induced cramps, hip dysplasia, and Ehlers-Danlos syndrome.[66][64][71] An occasional patient with genetically confirmed CMT may have acute worsening resembling inflammatory polyradiculoneuropathies.[72]

Evaluation

The evaluation of CMT has advanced with the development of more sophisticated tests. However, this is tempered by the fact that not all tests are uniformly available and may be expensive. In the past, diagnosis was solely based on the patient's clinical presentation. Presently, diagnosis involves a combination of clinical presentation, electrodiagnostic features, and genetic testing. Most experts in CMT advocate a stepwise process of testing when the disorder is suspected, starting with motor nerve conduction studies and then proceeding to sequential gene testing based on motor NCVs.

Electrophysiology

Electromyography (EMG) and NCV testing are crucial for confirming the diagnosis of neuropathy and distinguishing between demyelinating and axonal types of CMT. These minimally invasive tests offer rapid results and aid in excluding alternative causes of neuropathy. Additionally, EMG and NCV are valuable for screening asymptomatic relatives of the index patient. 

The key parameters measured are distal latencies, amplitudes, and velocities of motor and sensory nerves. Slowing of conduction velocities is an indirect measure of myelin dysfunction. Reduced amplitude of compound muscle action potential with preserved conduction velocity indicates axonopathy. In some cases, both demyelinating and axonal neuropathies may coexist. Evidence of denervation detected on concentric needle EMG helps establish axonal pathology.

The median NCV of 38 m/s is commonly used to differentiate demyelinating from axonal types of CMT. However, an intermediate form of CMT exists where NCV ranges from 25 to 45 m/s.[25] Based on NCVs, CMT is categorized as very slow (<15 m/s), slow (15-35 m/s), and intermediate (35-45 m/s), aiding in the selection of genetic testing for patients with inherited neuropathies.[73]

Slowing is usually uniform and diffuse in inherited neuropathies. However, in GJB1-associated CMT, motor conduction slowing is nonuniform and heterogeneous within a single nerve and between multiple nerves, with greater velocity reduction in the median compared to the ulnar nerve. This inter-nerve variability is particularly prominent in females.[74][75] Demyelination is more severe and uniform in affected males, leading to a more pronounced reduction in conduction velocities.[74]

Temporal dispersion and conduction blocks are distinguishing features of acquired demyelinating neuropathies compared to inherited ones.[76] The absence of dispersion and conduction blocks, along with slowed conduction velocities typically ranging from 20 to 30 m/s, are indicative of an underlying hereditary neuropathy. However, in cases involving GJB1 mutations, temporal dispersion can still occur despite the genetic association.[77][78][74][79]

Reduction in the amplitude of evoked motor responses and sensory potentials indicates the extent of axonal loss, which correlates with clinical disability, especially in patients presenting late with demyelinating CMTs.[76] Distinguishing primary axonal from demyelinating forms of CMT becomes challenging when the motor and sensory potentials are unrecordable. Proximal nerve studies, including the musculocutaneous and radial nerves, provide valuable insights in such cases.[79]

Genetic Testing

Genetic testing is the standard for establishing a conclusive diagnosis, providing medical counseling, aiding reproductive planning, and selecting patients for therapeutic trials and research.[80] All patients suspected of having CMT should have phenotypic characterization, detailed and accurate pedigree analysis, and proper pretest counseling. 

If there is a family history of CMT with a known variant, genetic testing should be performed for that variant. If the results are not diagnostic and NGS is available and affordable, CMT large-panel genetic testing should be performed. If not, sequential genetic testing based on motor NCVs should be done.

The most common genetic abnormality in CMT is CNV in PMP22. Molecular techniques that are available to detect CNVs in PMP22 include polymerase chain reaction (PCR), restriction fragment length polymorphism (RFLP), denaturing gradient gel electrophoresis (DGGE), single-strand conformational polymorphism (SSCP), array comparative genomic hybridization (aCGH), and fluorescence in-situ hybridization (FISH). These techniques are time-consuming, labor-intensive, expensive, and have limited sensitivity. The multiplex ligation-dependant probe amplification (MLPA) follows the principle of comparative quantification of specifically bound probes amplified by PCR using universal primers.[81] This test is simple, quick, sequence-specific, sensitive, and efficient, and thus MLPAt is the recommended analytical platform.[82]

Sanger sequencing is the standard in molecular genetic diagnosis. However, it has limitations in diagnosing CMT due to the large number of possible genetic variations. Mutations in genes other than PMP22, MPZ, GJB1, and MFN2 are rare, making it challenging to prioritize and conduct sequential analysis of all the possible CMT genes by Sanger sequencing. This approach is both time-consuming and expensive.

Genetic testing has evolved with the development of NGS technology, which is capable of sequencing the entire human genome, its protein-coding sequences, or specific genes of interest in parallel.[83] The yield of NGS depends on the gene panel used. The exome refers to the complete set of protein-coding regions in the human genome. Although it constitutes only 1% of the total genomic content, it contains more than 85% of the functional variations. Whole-exome sequencing evaluates the entire protein-coding region in an unselected or unbiased manner. Whole-genome sequencing evaluates both coding and noncoding regions. Some laboratories distinguish between clinical exome and whole-exome sequencing; the former refers to a set of genes implicated in most human diseases.[84] 

NGS of the entire exome may find the majority of the disease-causing genes. Targeted panel gene testing typically includes a limited number of genes based on the patient's phenotype. It has better technical performance, easier data analysis, fewer incidental findings, lower cost, and is easier to adopt in small laboratories. However, targeted panels may fail to identify novel gene associations.[85] NGS is the preferred technique for establishing the genetic diagnosis in CMT once CNVs or mutations in common genes are excluded. Even with whole-exome sequencing, about 40% of patients remain without a genetic diagnosis. NGS is negative in cases of mutations in genes that are unidentified or not described in the setting of CMT. NGS may miss mutations in the untranslated region (UTR), promoter, and other non-exonic regions. NGS is not sensitive to identifying CNVs or epigenetic, posttranscriptional, and posttranslational changes. Functional analysis is essential to ascribe pathogenicity when NGS identifies novel variants.[86]

Radiographic Imaging

Advances in imaging have enabled visualization of peripheral nerves along their entire length. Nerve ultrasound and magnetic resonance neurography are increasingly used to evaluate neuropathies. In CMT, diffuse enlargement of nerves exists, including roots, plexuses, and peripheral nerves, without variation between entrapment and nonentrapment sites.[87] Enlarged cranial nerves have also been described. The enlargement is more pronounced in upper limbs and CMT1A compared to other forms of CMT. In CMT2, there is no significant increase in the cross-sectional area (CSA) of peripheral nerves. An increase in CSA correlates with disability and disease progression.[88] Postcontrast enhancement, vascularity, altered signal characteristics within the nerve, and fascicular architecture differentiate CMT from other diagnoses, such as CIDP and leprosy.[89] 

In children and adolescents with CMT, magnetic resonance imaging (MRI) can be used to evaluate the relationship between muscle volume and intramuscular fat accumulation (IMFA) in the lower legs and impaired gait, weakness, and disability.[90]

Evaluation of Systemic Involvement

Patients with CMT need evaluation for concomitant illnesses, including diabetes mellitus, thyroid dysfunction, and nutritional deficiencies that affect peripheral nerve function. They may also need slit-lamp examination (for cataracts), intraocular tension recording (for glaucoma), laryngoscopy (for vocal cord mobility), and audiometry (for hearing loss). Additionally, spirometry and polysomnography may be necessary for evaluating restrictive lung disease, obstructive sleep apnea, and restless leg syndrome. Patients with CMT and INF2 mutations need evaluation for proteinuria and nephrotic syndrome due to the risk of focal segmental glomerulosclerosis, as these patients are often asymptomatic and their renal function may worsen rapidly, necessitating a renal transplant.[67]

Treatment / Management

Treatment of CMT is predominantly rehabilitative and symptomatic, as there are currently no effective disease-modifying therapies to alter its natural progression. Management typically involves a multidisciplinary healthcare team comprising neurologists, physiatrists, orthopedic surgeons, podiatrists, and physical and occupational therapists.

Pharmacotherapy

Treatment remains largely symptomatic. Musculoskeletal pain may be alleviated with acetaminophen or nonsteroidal anti-inflammatory drugs. Neuropathic pain may respond to tricyclic antidepressants, carbamazepine, or gabapentin. Fatigue can be managed with modafinil.[91] Additional modalities rest on the current understanding of the underlying genetic abnormalities and pathophysiology of CMT, coupled with newer drug development techniques such as systemic biology-based modeling, anti-sense oligonucleotides, adenoviral vector-based drug delivery, and RNA interference technology. In CMT1A, efforts to target PMP22 overexpression—using agents such as ascorbic acid, onapristone, geldanamycin, and rapamycin—showed promise in animal and cell studies, but have not demonstrated effectiveness in human clinical trials.[92] (B3)

PXT3003 (a combination of baclofen, naltrexone, and d-sorbitol) has demonstrated efficacy in reducing the toxic effects of PMP22 overexpression in both mice and humans. Subjects treated with PXT3003 showed no deterioration or improvement in CMT Neuropathy score (CMTNS), Overall Neuropathy Limitations Scale (ONLS), the 10-m walk test, and conduction velocities compared to placebo. PXT3003 was well tolerated and safe.[93] Additionally, curcumin has shown promise in alleviating endoplasmic reticulum stress and improving MPZ-associated neuropathy in murine models.[94](A1)

Other agents investigated for their therapeutic potential in CMTs, albeit with limited success, include promoters of axonal regeneration (neurotrophin-3 and neuregulins), gene expression regulators (histone deacetylases), chaperones and heat shock protein inducers (arimoclomol and celastrol), calcium homeostasis modulators (P2X7 antagonists-adenosine homodinucleotide P18), neuroprotective and antioxidant drugs (purified polyols-resveratrol), and potassium channel blockers (3,4-diaminopyridine).[92] Supplementation with essential fatty acids, phospholipids, vitamin E, creatine, and bovine-derived ganglioside mixture has not been effective.[95](B3)

Rehabilitation

An interprofessional healthcare team is crucial for treating patients with CMT who are at risk of developing reduced range of motion, contractures, and musculoskeletal deformities. The important components of physical therapy include stretching, gripping exercises, aerobics, resistance training, and timely use of orthotic devices such as special shoes with good ankle support to correct foot drop and facilitate walking. Some individuals may require canes or forearm crutches for gait stability. Physical therapy improves and maintains muscle strength and function, enhances joint flexibility and range of movement, and helps improve balance and cardiorespiratory fitness. Physical therapy also addresses fatigue and pain and prevents stiffness and deformities.[92][96] (A1)

Patients with CMT may face challenges participating in exercise programs due to heightened energy demands and altered gait kinetics. The concept of "overuse weakness," where exercise potentially exacerbates muscle loss and weakness, remains debated in CMT.[97] Orthoses improve posture and walking balance, particularly in cases of ankle weakness and deformity. Supervised exercise programs are necessary for managing hand muscle weakness, complemented by occupational therapy to aid daily activities. Splinting is effective in enhancing hand dexterity.[96](A1)

Additional Treatment Measures

Symptomatic treatment of fatigue, depression, neuropathic pain, and restless legs syndrome is similar to other neuropathies. Foot deformities, scoliosis, and hip dysplasia may need corrective surgeries and tendon transfers. Restrictive lung disease, sleep apnea, and vocal cord palsy need suitable intervention in collaboration with surgeons, pulmonologists, otorhinolaryngologists, physiatrists, and related specialties.[92] Close monitoring and management are recommended for pregnant women with CMT.[98](A1)

Pediatrics

Progressive resistance exercise of the ankle dorsiflexors is recommended to improve muscle strength and slow the progression of weakness. Strength training of core muscles is encouraged, as are joint stretching, balance training, and the use of ankle-foot orthoses. Referral to an orthopedic surgeon may be necessary for addressing progressive pes cavus, ankle contractures, hip dysplasia, and scoliosis. Adaptive equipment should be utilized to improve daily living activities. Respiratory deficits should be assessed, and a referral to a pulmonologist or sleep physician is advised for recurrent lower respiratory tract infections or sleep-disordered breathing.[99] (A1)

Differential Diagnosis

CMT must be differentiated from other conditions characterized by predominant distal weakness, muscle wasting, foot deformities, and a progressive course. The most critical differential diagnosis involves distinguishing between acquired dysimmune neuropathies and demyelinating CMT.[100]

Infantile spinal muscular atrophy with respiratory distress type 1 (SMARD1) due to mutations in the IGHMBP2 gene is an important differential diagnosis, as motor NCVs are very low in this condition. However, patients with SMARD1 have normal sensory nerve action potentials.[101] Distal myopathies can manifest with foot drop, but the pattern of weakness and electrophysiological studies (nerve conduction studies and concentric needle EMG) can distinguish them from CMT. In children, distinguishing CIDP from CMT can be challenging. Genetic testing, along with clinical and electrophysiological examinations of asymptomatic family members, and nerve biopsies, may assist in making an accurate diagnosis.[100]

Other differential diagnoses include acquired neuropathies due to diabetes mellitus, nutritional deficiencies, vasculitis, and heavy metal intoxication. Inherited conditions where neuropathy is part of a complex multisystem disorder, such as autosomal-recessive spastic ataxia of Charlevoix-Saguenay (ARSACS), Friedreich’s ataxia, metachromatic leukodystrophy, the syndrome of neurogenic weakness, ataxia, and retinitis pigmentosa (NARP), or Refsum’s disease, should also be considered.[100][102]

Staging

Patients with CMT have a progressive clinical course and need periodic monitoring. Composite scoring systems are available to assess longitudinal changes in function and disability, including natural history and outcome. The CMTNS is a reliable composite of 9 items that incorporates motor and sensory signs and symptoms, as well as electrophysiological parameters. The CMTNS score ranges from 0 (best) to 36 (worst) and is used for classifying patients as having mild (≤10), moderate (11-20), or severe (≥20) disability.[103] 

A similar score exists for use in the pediatric population (CMTPedS).[104] However, the annual change in this score is insufficient to assess therapeutic response in clinical trials. Other clinical and functional outcome measures include handgrip myometry, ankle dorsiflexion myometry, the overall neuropathy limitations scale, and the 9-hole peg test.[105] 

Alternate sensitive, reliable, reproducible, and clinically relevant markers that reflect degeneration of Schwann cells or axons and act as measures of disease burden and progression have been developed. These markers include IMFA in calf muscles, serum neurofilament light chain, transmembrane protease serine 5 (TMPRSS5), and cutaneous mRNA profiling.[105][106][107]

Prognosis

CMT is a group of insidiously progressive disorders characterized by increasing weakness and wasting of the extremities, which interfere with mobility and activities of daily living. The rate of progression varies among different forms of CMT. Although life expectancy is usually unaffected, the condition can be severe if early onset occurs. Periodic evaluation and interventions by an interprofessional rehabilitation team are essential for maintaining independence, safe ambulation, and functional activities.

Complications

Foot deformities are a major source of disability in CMT. These deformities develop over time and may worsen despite physiotherapy and the use of orthoses. Early surgical intervention to correct muscle imbalance and establish a plantigrade foot position may prevent further deformity and reduce the need for future joint fusions.[108][109] Corns and calluses may develop at sites of increased pressure and lead to skin ulceration. Respiratory insufficiency may arise due to severe neuropathy, diaphragmatic palsy, or concomitant spinal deformity.[64] 

Worsening of CMT-related symptoms may occur during pregnancy, increasing the risk of abnormal fetal lie, postpartum hemorrhage, and the need for emergency intervention during delivery. Therefore, regular maternal-fetal prenatal care is recommended. Additionally, neurotoxic drugs, chemotherapeutic agents, and comorbid neuropathic diseases can exacerbate neuropathy in CMT.[80]

Deterrence and Patient Education

Individuals with CMT and their families should be educated about the disease and its symptoms, such as weakness and foot problems. Additional information should be provided about the available treatments and the overall progression of the disease. The likely mode of inheritance should be explained to the patient based on pedigree chart analysis and genetic testing. Patients with CMT are encouraged to remain active, maintain a healthy weight, follow up with the interprofessional team, and avoid neurotoxic agents and drugs that can worsen neuropathy. Referral to genetic counseling is essential to address questions about family planning.

Pearls and Other Issues

  • CMT refers to a group of inherited neuropathies characterized by progressive weakness, muscle wasting, and foot deformities.
  • Electrophysiological tests help confirm the diagnosis of neuropathy and exclude alternative conditions such as distal myopathies, muscular dystrophies, and idiopathic pes cavus that also present with foot drop and foot deformities.
  • Electrophysiological tests are valuable for screening family members for asymptomatic neuropathy.
  • Patients with demyelinating CMT typically exhibit slowed NCVs in early childhood. Over time, distal axonal loss leads to weakness, muscle wasting, and foot deformities.
  • All patients suspected of having CMT should undergo genetic counseling before undergoing genetic testing.
  • In demyelinating neuropathies, patients should first undergo testing for PMP22 duplication, as it is the most common genetic abnormality in CMT. Following the exclusion of CNVs in PMP22, targeted gene sequencing or whole-exome sequencing should be considered.
  • In axonal neuropathies, the patient should first be tested for mutations in MFN2. Alternatively, they can undergo targeted gene sequencing or whole-exome sequencing.
  • Establishing the genetic diagnosis is crucial for ongoing genetic counseling, reproductive planning, and considering the patient for potential new therapies in research.
  • Patients with CMT should undergo monitoring for comorbid conditions.
  • Patients with CMT should avoid using drugs that worsen neuropathy.
  • The interprofessional healthcare team should offer patient education and counseling, regular follow-up, emphasize rehabilitation measures, and provide access to therapeutic trials.

Enhancing Healthcare Team Outcomes

Patients benefit most from an interprofessional healthcare team with a patient-centric approach to comprehensive care. The interprofessional healthcare team should include specialists in the fields of neurology, primary care, physiatry, physiotherapy, nursing, occupational therapy, social work, and neurogenetic counseling. Having experts in orthopedics, pulmonology, speech therapy, sleep medicine, and nephrology on the team is also essential. Healthcare providers should coordinate with these team members to optimize the patient's function, prevent injuries and complications, and improve the overall quality of life for individuals with CMT.

The team should include expertise in recognizing the varied clinical presentations of CMT and understanding the nuances of diagnostic techniques such as EMG, NCV testing, and genetic testing. 

References


[1]

Kazamel M, Boes CJ. Charcot Marie Tooth disease (CMT): historical perspectives and evolution. Journal of neurology. 2015:262(4):801-5. doi: 10.1007/s00415-014-7490-9. Epub 2014 Sep 9     [PubMed PMID: 25201224]

Level 3 (low-level) evidence

[2]

Magy L, Mathis S, Le Masson G, Goizet C, Tazir M, Vallat JM. Updating the classification of inherited neuropathies: Results of an international survey. Neurology. 2018 Mar 6:90(10):e870-e876. doi: 10.1212/WNL.0000000000005074. Epub 2018 Feb 2     [PubMed PMID: 29429969]

Level 3 (low-level) evidence

[3]

Pareyson D, Saveri P, Pisciotta C. New developments in Charcot-Marie-Tooth neuropathy and related diseases. Current opinion in neurology. 2017 Oct:30(5):471-480. doi: 10.1097/WCO.0000000000000474. Epub     [PubMed PMID: 28678038]

Level 3 (low-level) evidence

[4]

Laurá M, Pipis M, Rossor AM, Reilly MM. Charcot-Marie-Tooth disease and related disorders: an evolving landscape. Current opinion in neurology. 2019 Oct:32(5):641-650. doi: 10.1097/WCO.0000000000000735. Epub     [PubMed PMID: 31343428]

Level 3 (low-level) evidence

[5]

Fridman V, Bundy B, Reilly MM, Pareyson D, Bacon C, Burns J, Day J, Feely S, Finkel RS, Grider T, Kirk CA, Herrmann DN, Laurá M, Li J, Lloyd T, Sumner CJ, Muntoni F, Piscosquito G, Ramchandren S, Shy R, Siskind CE, Yum SW, Moroni I, Pagliano E, Zuchner S, Scherer SS, Shy ME, Inherited Neuropathies Consortium. CMT subtypes and disease burden in patients enrolled in the Inherited Neuropathies Consortium natural history study: a cross-sectional analysis. Journal of neurology, neurosurgery, and psychiatry. 2015 Aug:86(8):873-8. doi: 10.1136/jnnp-2014-308826. Epub 2014 Nov 27     [PubMed PMID: 25430934]

Level 2 (mid-level) evidence

[6]

Hoebeke C, Bonello-Palot N, Audic F, Boulay C, Tufod D, Attarian S, Chabrol B. Retrospective study of 75 children with peripheral inherited neuropathy: Genotype-phenotype correlations. Archives de pediatrie : organe officiel de la Societe francaise de pediatrie. 2018 Nov:25(8):452-458. doi: 10.1016/j.arcped.2018.09.006. Epub 2018 Oct 16     [PubMed PMID: 30340945]

Level 2 (mid-level) evidence

[7]

Gess B, Schirmacher A, Boentert M, Young P. Charcot-Marie-Tooth disease: frequency of genetic subtypes in a German neuromuscular center population. Neuromuscular disorders : NMD. 2013 Aug:23(8):647-51. doi: 10.1016/j.nmd.2013.05.005. Epub 2013 Jun 3     [PubMed PMID: 23743332]

Level 2 (mid-level) evidence

[8]

Rudnik-Schöneborn S, Tölle D, Senderek J, Eggermann K, Elbracht M, Kornak U, von der Hagen M, Kirschner J, Leube B, Müller-Felber W, Schara U, von Au K, Wieczorek D, Bußmann C, Zerres K. Diagnostic algorithms in Charcot-Marie-Tooth neuropathies: experiences from a German genetic laboratory on the basis of 1206 index patients. Clinical genetics. 2016 Jan:89(1):34-43. doi: 10.1111/cge.12594. Epub 2015 Apr 29     [PubMed PMID: 25850958]


[9]

Milley GM, Varga ET, Grosz Z, Nemes C, Arányi Z, Boczán J, Diószeghy P, Molnár MJ, Gál A. Genotypic and phenotypic spectrum of the most common causative genes of Charcot-Marie-Tooth disease in Hungarian patients. Neuromuscular disorders : NMD. 2018 Jan:28(1):38-43. doi: 10.1016/j.nmd.2017.08.007. Epub 2017 Sep 8     [PubMed PMID: 29174527]


[10]

Cutrupi AN, Brewer MH, Nicholson GA, Kennerson ML. Structural variations causing inherited peripheral neuropathies: A paradigm for understanding genomic organization, chromatin interactions, and gene dysregulation. Molecular genetics & genomic medicine. 2018 May:6(3):422-433. doi: 10.1002/mgg3.390. Epub 2018 Mar 23     [PubMed PMID: 29573232]

Level 3 (low-level) evidence

[11]

Fortun J, Li J, Go J, Fenstermaker A, Fletcher BS, Notterpek L. Impaired proteasome activity and accumulation of ubiquitinated substrates in a hereditary neuropathy model. Journal of neurochemistry. 2005 Mar:92(6):1531-41     [PubMed PMID: 15748170]

Level 3 (low-level) evidence

[12]

Ryan MC, Shooter EM, Notterpek L. Aggresome formation in neuropathy models based on peripheral myelin protein 22 mutations. Neurobiology of disease. 2002 Jul:10(2):109-18     [PubMed PMID: 12127149]

Level 3 (low-level) evidence

[13]

Chies R, Nobbio L, Edomi P, Schenone A, Schneider C, Brancolini C. Alterations in the Arf6-regulated plasma membrane endosomal recycling pathway in cells overexpressing the tetraspan protein Gas3/PMP22. Journal of cell science. 2003 Mar 15:116(Pt 6):987-99     [PubMed PMID: 12584243]

Level 3 (low-level) evidence

[14]

Isaacs AM, Jeans A, Oliver PL, Vizor L, Brown SD, Hunter AJ, Davies KE. Identification of a new Pmp22 mouse mutant and trafficking analysis of a Pmp22 allelic series suggesting that protein aggregates may be protective in Pmp22-associated peripheral neuropathy. Molecular and cellular neurosciences. 2002 Sep:21(1):114-25     [PubMed PMID: 12359155]

Level 3 (low-level) evidence

[15]

Chrast R, Saher G, Nave KA, Verheijen MH. Lipid metabolism in myelinating glial cells: lessons from human inherited disorders and mouse models. Journal of lipid research. 2011 Mar:52(3):419-34. doi: 10.1194/jlr.R009761. Epub 2010 Nov 9     [PubMed PMID: 21062955]

Level 3 (low-level) evidence

[16]

Skre H. Genetic and clinical aspects of Charcot-Marie-Tooth's disease. Clinical genetics. 1974:6(2):98-118     [PubMed PMID: 4430158]


[17]

Barohn RJ. Approach to peripheral neuropathy and neuronopathy. Seminars in neurology. 1998:18(1):7-18     [PubMed PMID: 9562663]


[18]

Dyck PJ, Oviatt KF, Lambert EH. Intensive evaluation of referred unclassified neuropathies yields improved diagnosis. Annals of neurology. 1981 Sep:10(3):222-6     [PubMed PMID: 7294727]


[19]

Barreto LC, Oliveira FS, Nunes PS, de França Costa IM, Garcez CA, Goes GM, Neves EL, de Souza Siqueira Quintans J, de Souza Araújo AA. Epidemiologic Study of Charcot-Marie-Tooth Disease: A Systematic Review. Neuroepidemiology. 2016:46(3):157-65. doi: 10.1159/000443706. Epub 2016 Feb 6     [PubMed PMID: 26849231]

Level 1 (high-level) evidence

[20]

Martyn CN, Hughes RA. Epidemiology of peripheral neuropathy. Journal of neurology, neurosurgery, and psychiatry. 1997 Apr:62(4):310-8     [PubMed PMID: 9120441]


[21]

Higuchi Y, Takashima H. Clinical genetics of Charcot-Marie-Tooth disease. Journal of human genetics. 2023 Mar:68(3):199-214. doi: 10.1038/s10038-022-01031-2. Epub 2022 Mar 18     [PubMed PMID: 35304567]


[22]

Berger P, Niemann A, Suter U. Schwann cells and the pathogenesis of inherited motor and sensory neuropathies (Charcot-Marie-Tooth disease). Glia. 2006 Sep:54(4):243-57     [PubMed PMID: 16856148]

Level 3 (low-level) evidence

[23]

Jerath NU, Shy ME. Hereditary motor and sensory neuropathies: Understanding molecular pathogenesis could lead to future treatment strategies. Biochimica et biophysica acta. 2015 Apr:1852(4):667-78. doi: 10.1016/j.bbadis.2014.07.031. Epub 2014 Aug 6     [PubMed PMID: 25108281]

Level 3 (low-level) evidence

[24]

Hanemann CO. Hereditary demyelinating neuropathies: from gene to disease. Neurogenetics. 2001 Mar:3(2):53-7     [PubMed PMID: 11354826]

Level 3 (low-level) evidence

[25]

Tazir M, Hamadouche T, Nouioua S, Mathis S, Vallat JM. Hereditary motor and sensory neuropathies or Charcot-Marie-Tooth diseases: an update. Journal of the neurological sciences. 2014 Dec 15:347(1-2):14-22. doi: 10.1016/j.jns.2014.10.013. Epub 2014 Oct 16     [PubMed PMID: 25454638]


[26]

Vallat JM, Mathis S, Funalot B. The various Charcot-Marie-Tooth diseases. Current opinion in neurology. 2013 Oct:26(5):473-80. doi: 10.1097/WCO.0b013e328364c04b. Epub     [PubMed PMID: 23945280]

Level 3 (low-level) evidence

[27]

Mathis S, Goizet C, Tazir M, Magdelaine C, Lia AS, Magy L, Vallat JM. Charcot-Marie-Tooth diseases: an update and some new proposals for the classification. Journal of medical genetics. 2015 Oct:52(10):681-90. doi: 10.1136/jmedgenet-2015-103272. Epub 2015 Aug 5     [PubMed PMID: 26246519]


[28]

Mathis S, Funalot B, Boyer O, Lacroix C, Marcorelles P, Magy L, Richard L, Antignac C, Vallat JM. Neuropathologic characterization of INF2-related Charcot-Marie-Tooth disease: evidence for a Schwann cell actinopathy. Journal of neuropathology and experimental neurology. 2014 Mar:73(3):223-33. doi: 10.1097/NEN.0000000000000047. Epub     [PubMed PMID: 24487800]


[29]

Duchesne M, Mathis S, Richard L, Magdelaine C, Corcia P, Nouioua S, Tazir M, Magy L, Vallat JM. Nerve Biopsy Is Still Useful in Some Inherited Neuropathies. Journal of neuropathology and experimental neurology. 2018 Feb 1:77(2):88-99. doi: 10.1093/jnen/nlx111. Epub     [PubMed PMID: 29300988]


[30]

Gabriel CM, Gregson NA, Wood NW, Hughes RA. Immunological study of hereditary motor and sensory neuropathy type 1a (HMSN1a). Journal of neurology, neurosurgery, and psychiatry. 2002 Feb:72(2):230-5     [PubMed PMID: 11796774]


[31]

Martini R, Toyka KV. Immune-mediated components of hereditary demyelinating neuropathies: lessons from animal models and patients. The Lancet. Neurology. 2004 Aug:3(8):457-65     [PubMed PMID: 15261606]

Level 3 (low-level) evidence

[32]

Rajabally YA, Adams D, Latour P, Attarian S. Hereditary and inflammatory neuropathies: a review of reported associations, mimics and misdiagnoses. Journal of neurology, neurosurgery, and psychiatry. 2016 Oct:87(10):1051-60. doi: 10.1136/jnnp-2015-310835. Epub 2016 Mar 23     [PubMed PMID: 27010614]


[33]

Xu W, Manichella D, Jiang H, Vallat JM, Lilien J, Baron P, Scarlato G, Kamholz J, Shy ME. Absence of P0 leads to the dysregulation of myelin gene expression and myelin morphogenesis. Journal of neuroscience research. 2000 Jun 15:60(6):714-24     [PubMed PMID: 10861783]

Level 3 (low-level) evidence

[34]

Nelis E, Erdem S, Tan E, Löfgren A, Ceuterick C, De Jonghe P, Van Broeckhoven C, Timmerman V, Topaloglu H. A novel homozygous missense mutation in the myotubularin-related protein 2 gene associated with recessive Charcot-Marie-Tooth disease with irregularly folded myelin sheaths. Neuromuscular disorders : NMD. 2002 Nov:12(9):869-73     [PubMed PMID: 12398840]

Level 3 (low-level) evidence

[35]

Takashima H, Boerkoel CF, De Jonghe P, Ceuterick C, Martin JJ, Voit T, Schröder JM, Williams A, Brophy PJ, Timmerman V, Lupski JR. Periaxin mutations cause a broad spectrum of demyelinating neuropathies. Annals of neurology. 2002 Jun:51(6):709-15     [PubMed PMID: 12112076]


[36]

Kochanski A, Drac H, Kabzińska D, Hausmanowa-Petrusewicz I. A novel mutation, Thr65Ala, in the MPZ gene in a patient with Charcot-Marie-Tooth type 1B disease with focally folded myelin. Neuromuscular disorders : NMD. 2004 Mar:14(3):229-32     [PubMed PMID: 15036333]

Level 3 (low-level) evidence

[37]

Fabrizi GM, Taioli F, Cavallaro T, Ferrari S, Bertolasi L, Casarotto M, Rizzuto N, Deconinck T, Timmerman V, De Jonghe P. Further evidence that mutations in FGD4/frabin cause Charcot-Marie-Tooth disease type 4H. Neurology. 2009 Mar 31:72(13):1160-4. doi: 10.1212/01.wnl.0000345373.58618.b6. Epub     [PubMed PMID: 19332693]

Level 3 (low-level) evidence

[38]

Houlden H, Laura M, Ginsberg L, Jungbluth H, Robb SA, Blake J, Robinson S, King RH, Reilly MM. The phenotype of Charcot-Marie-Tooth disease type 4C due to SH3TC2 mutations and possible predisposition to an inflammatory neuropathy. Neuromuscular disorders : NMD. 2009 Apr:19(4):264-9. doi: 10.1016/j.nmd.2009.01.006. Epub 2009 Mar 9     [PubMed PMID: 19272779]


[39]

Vallat JM. Dominantly inherited peripheral neuropathies. Journal of neuropathology and experimental neurology. 2003 Jul:62(7):699-714     [PubMed PMID: 12901697]


[40]

Nouioua S, Hamadouche T, Funalot B, Bernard R, Bellatache N, Bouderba R, Grid D, Assami S, Benhassine T, Levy N, Vallat JM, Tazir M. Novel mutations in the PRX and the MTMR2 genes are responsible for unusual Charcot-Marie-Tooth disease phenotypes. Neuromuscular disorders : NMD. 2011 Aug:21(8):543-50. doi: 10.1016/j.nmd.2011.04.013. Epub 2011 Jul 7     [PubMed PMID: 21741241]

Level 2 (mid-level) evidence

[41]

Bonneick S, Boentert M, Berger P, Atanasoski S, Mantei N, Wessig C, Toyka KV, Young P, Suter U. An animal model for Charcot-Marie-Tooth disease type 4B1. Human molecular genetics. 2005 Dec 1:14(23):3685-95     [PubMed PMID: 16249189]

Level 3 (low-level) evidence

[42]

Arnaud E, Zenker J, de Preux Charles AS, Stendel C, Roos A, Médard JJ, Tricaud N, Kleine H, Luscher B, Weis J, Suter U, Senderek J, Chrast R. SH3TC2/KIAA1985 protein is required for proper myelination and the integrity of the node of Ranvier in the peripheral nervous system. Proceedings of the National Academy of Sciences of the United States of America. 2009 Oct 13:106(41):17528-33. doi: 10.1073/pnas.0905523106. Epub 2009 Sep 29     [PubMed PMID: 19805030]

Level 3 (low-level) evidence

[43]

Roos A, Weis J, Korinthenberg R, Fehrenbach H, Häusler M, Züchner S, Mache C, Hubmann H, Auer-Grumbach M, Senderek J. Inverted formin 2-related Charcot-Marie-Tooth disease: extension of the mutational spectrum and pathological findings in Schwann cells and axons. Journal of the peripheral nervous system : JPNS. 2015 Mar:20(1):52-9. doi: 10.1111/jns.12106. Epub     [PubMed PMID: 25676889]

Level 3 (low-level) evidence

[44]

Brennan KM, Bai Y, Pisciotta C, Wang S, Feely SM, Hoegger M, Gutmann L, Moore SA, Gonzalez M, Sherman DL, Brophy PJ, Züchner S, Shy ME. Absence of Dystrophin Related Protein-2 disrupts Cajal bands in a patient with Charcot-Marie-Tooth disease. Neuromuscular disorders : NMD. 2015 Oct:25(10):786-93. doi: 10.1016/j.nmd.2015.07.001. Epub 2015 Jul 7     [PubMed PMID: 26227883]


[45]

Sherman DL, Wu LM, Grove M, Gillespie CS, Brophy PJ. Drp2 and periaxin form Cajal bands with dystroglycan but have distinct roles in Schwann cell growth. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2012 Jul 4:32(27):9419-28. doi: 10.1523/JNEUROSCI.1220-12.2012. Epub     [PubMed PMID: 22764250]

Level 3 (low-level) evidence

[46]

Laquérriere A, Maluenda J, Camus A, Fontenas L, Dieterich K, Nolent F, Zhou J, Monnier N, Latour P, Gentil D, Héron D, Desguerres I, Landrieu P, Beneteau C, Delaporte B, Bellesme C, Baumann C, Capri Y, Goldenberg A, Lyonnet S, Bonneau D, Estournet B, Quijano-Roy S, Francannet C, Odent S, Saint-Frison MH, Sigaudy S, Figarella-Branger D, Gelot A, Mussini JM, Lacroix C, Drouin-Garraud V, Malinge MC, Attié-Bitach T, Bessieres B, Bonniere M, Encha-Razavi F, Beaufrère AM, Khung-Savatovsky S, Perez MJ, Vasiljevic A, Mercier S, Roume J, Trestard L, Saugier-Veber P, Cordier MP, Layet V, Legendre M, Vigouroux-Castera A, Lunardi J, Bayes M, Jouk PS, Rigonnot L, Granier M, Sternberg D, Warszawski J, Gut I, Gonzales M, Tawk M, Melki J. Mutations in CNTNAP1 and ADCY6 are responsible for severe arthrogryposis multiplex congenita with axoglial defects. Human molecular genetics. 2014 May 1:23(9):2279-89. doi: 10.1093/hmg/ddt618. Epub 2013 Dec 6     [PubMed PMID: 24319099]


[47]

Vallat JM, Nizon M, Magee A, Isidor B, Magy L, Péréon Y, Richard L, Ouvrier R, Cogné B, Devaux J, Zuchner S, Mathis S. Contactin-Associated Protein 1 (CNTNAP1) Mutations Induce Characteristic Lesions of the Paranodal Region. Journal of neuropathology and experimental neurology. 2016 Dec 1:75(12):1155-1159. doi: 10.1093/jnen/nlw093. Epub     [PubMed PMID: 27818385]


[48]

Hengel H, Magee A, Mahanjah M, Vallat JM, Ouvrier R, Abu-Rashid M, Mahamid J, Schüle R, Schulze M, Krägeloh-Mann I, Bauer P, Züchner S, Sharkia R, Schöls L. CNTNAP1 mutations cause CNS hypomyelination and neuropathy with or without arthrogryposis. Neurology. Genetics. 2017 Apr:3(2):e144. doi: 10.1212/NXG.0000000000000144. Epub 2017 Mar 22     [PubMed PMID: 28374019]


[49]

Funalot B, Topilko P, Arroyo MA, Sefiani A, Hedley-Whyte ET, Yoldi ME, Richard L, Touraille E, Laurichesse M, Khalifa E, Chauzeix J, Ouedraogo A, Cros D, Magdelaine C, Sturtz FG, Urtizberea JA, Charnay P, Bragado FG, Vallat JM. Homozygous deletion of an EGR2 enhancer in congenital amyelinating neuropathy. Annals of neurology. 2012 May:71(5):719-23. doi: 10.1002/ana.23527. Epub     [PubMed PMID: 22522483]

Level 3 (low-level) evidence

[50]

Wilmshurst JM, Pollard JD, Nicholson G, Antony J, Ouvrier R. Peripheral neuropathies of infancy. Developmental medicine and child neurology. 2003 Jun:45(6):408-14     [PubMed PMID: 12785442]


[51]

Burns J, Ryan MM, Ouvrier RA. Evolution of foot and ankle manifestations in children with CMT1A. Muscle & nerve. 2009 Feb:39(2):158-66. doi: 10.1002/mus.21140. Epub     [PubMed PMID: 19145658]


[52]

Smit LS, Roofthooft D, van Ruissen F, Baas F, van Doorn PA. Congenital hypomyelinating neuropathy, a long term follow-up study in an affected family. Neuromuscular disorders : NMD. 2008 Jan:18(1):59-62     [PubMed PMID: 17825553]

Level 3 (low-level) evidence

[53]

Dierick I, Baets J, Irobi J, Jacobs A, De Vriendt E, Deconinck T, Merlini L, Van den Bergh P, Rasic VM, Robberecht W, Fischer D, Morales RJ, Mitrovic Z, Seeman P, Mazanec R, Kochanski A, Jordanova A, Auer-Grumbach M, Helderman-van den Enden AT, Wokke JH, Nelis E, De Jonghe P, Timmerman V. Relative contribution of mutations in genes for autosomal dominant distal hereditary motor neuropathies: a genotype-phenotype correlation study. Brain : a journal of neurology. 2008 May:131(Pt 5):1217-27. doi: 10.1093/brain/awn029. Epub 2008 Mar 5     [PubMed PMID: 18325928]


[54]

Nave KA, Sereda MW, Ehrenreich H. Mechanisms of disease: inherited demyelinating neuropathies--from basic to clinical research. Nature clinical practice. Neurology. 2007 Aug:3(8):453-64     [PubMed PMID: 17671523]

Level 3 (low-level) evidence

[55]

Ursino G, Alberti MA, Grandis M, Reni L, Pareyson D, Bellone E, Gemelli C, Sabatelli M, Pisciotta C, Luigetti M, Santoro L, Massollo L, Schenone A. Influence of comorbidities on the phenotype of patients affected by Charcot-Marie-Tooth neuropathy type 1A. Neuromuscular disorders : NMD. 2013 Nov:23(11):902-6. doi: 10.1016/j.nmd.2013.07.002. Epub 2013 Jul 23     [PubMed PMID: 23891256]

Level 2 (mid-level) evidence

[56]

Sanmaneechai O, Feely S, Scherer SS, Herrmann DN, Burns J, Muntoni F, Li J, Siskind CE, Day JW, Laura M, Sumner CJ, Lloyd TE, Ramchandren S, Shy RR, Grider T, Bacon C, Finkel RS, Yum SW, Moroni I, Piscosquito G, Pareyson D, Reilly MM, Shy ME, Inherited Neuropathies Consortium - Rare Disease Clinical Research Consortium (INC-RDCRC). Genotype-phenotype characteristics and baseline natural history of heritable neuropathies caused by mutations in the MPZ gene. Brain : a journal of neurology. 2015 Nov:138(Pt 11):3180-92. doi: 10.1093/brain/awv241. Epub 2015 Aug 25     [PubMed PMID: 26310628]


[57]

Birouk N, Gouider R, Le Guern E, Gugenheim M, Tardieu S, Maisonobe T, Le Forestier N, Agid Y, Brice A, Bouche P. Charcot-Marie-Tooth disease type 1A with 17p11.2 duplication. Clinical and electrophysiological phenotype study and factors influencing disease severity in 119 cases. Brain : a journal of neurology. 1997 May:120 ( Pt 5)():813-23     [PubMed PMID: 9183252]

Level 3 (low-level) evidence

[58]

Scherer SS, Wrabetz L. Molecular mechanisms of inherited demyelinating neuropathies. Glia. 2008 Nov 1:56(14):1578-1589. doi: 10.1002/glia.20751. Epub     [PubMed PMID: 18803325]

Level 3 (low-level) evidence

[59]

Taioli F, Cabrini I, Cavallaro T, Acler M, Fabrizi GM. Inherited demyelinating neuropathies with micromutations of peripheral myelin protein 22 gene. Brain : a journal of neurology. 2011 Feb:134(Pt 2):608-17. doi: 10.1093/brain/awq374. Epub 2011 Jan 19     [PubMed PMID: 21252112]


[60]

Krajewski KM, Lewis RA, Fuerst DR, Turansky C, Hinderer SR, Garbern J, Kamholz J, Shy ME. Neurological dysfunction and axonal degeneration in Charcot-Marie-Tooth disease type 1A. Brain : a journal of neurology. 2000 Jul:123 ( Pt 7)():1516-27     [PubMed PMID: 10869062]


[61]

Lewis RA, Li J, Fuerst DR, Shy ME, Krajewski K. Motor unit number estimate of distal and proximal muscles in Charcot-Marie-Tooth disease. Muscle & nerve. 2003 Aug:28(2):161-7     [PubMed PMID: 12872319]


[62]

Lawson VH, Gordon Smith A, Bromberg MB. Assessment of axonal loss in Charcot-Marie-Tooth neuropathies. Experimental neurology. 2003 Dec:184(2):753-7     [PubMed PMID: 14769367]


[63]

Stuppia G, Rizzo F, Riboldi G, Del Bo R, Nizzardo M, Simone C, Comi GP, Bresolin N, Corti S. MFN2-related neuropathies: Clinical features, molecular pathogenesis and therapeutic perspectives. Journal of the neurological sciences. 2015 Sep 15:356(1-2):7-18. doi: 10.1016/j.jns.2015.05.033. Epub 2015 May 29     [PubMed PMID: 26143526]

Level 3 (low-level) evidence

[64]

Werheid F, Azzedine H, Zwerenz E, Bozkurt A, Moeller MJ, Lin L, Mull M, Häusler M, Schulz JB, Weis J, Claeys KG. Underestimated associated features in CMT neuropathies: clinical indicators for the causative gene? Brain and behavior. 2016 Apr:6(4):e00451. doi: 10.1002/brb3.451. Epub 2016 Mar 4     [PubMed PMID: 27088055]


[65]

Stojkovic T, de Seze J, Dubourg O, Arne-Bes MC, Tardieu S, Hache JC, Vermersch P. Autonomic and respiratory dysfunction in Charcot-Marie-Tooth disease due to Thr124Met mutation in the myelin protein zero gene. Clinical neurophysiology : official journal of the International Federation of Clinical Neurophysiology. 2003 Sep:114(9):1609-14     [PubMed PMID: 12948789]


[66]

Hattori N, Yamamoto M, Yoshihara T, Koike H, Nakagawa M, Yoshikawa H, Ohnishi A, Hayasaka K, Onodera O, Baba M, Yasuda H, Saito T, Nakashima K, Kira J, Kaji R, Oka N, Sobue G, Study Group for Hereditary Neuropathy in Japan. Demyelinating and axonal features of Charcot-Marie-Tooth disease with mutations of myelin-related proteins (PMP22, MPZ and Cx32): a clinicopathological study of 205 Japanese patients. Brain : a journal of neurology. 2003 Jan:126(Pt 1):134-51     [PubMed PMID: 12477701]


[67]

Boyer O, Nevo F, Plaisier E, Funalot B, Gribouval O, Benoit G, Huynh Cong E, Arrondel C, Tête MJ, Montjean R, Richard L, Karras A, Pouteil-Noble C, Balafrej L, Bonnardeaux A, Canaud G, Charasse C, Dantal J, Deschenes G, Deteix P, Dubourg O, Petiot P, Pouthier D, Leguern E, Guiochon-Mantel A, Broutin I, Gubler MC, Saunier S, Ronco P, Vallat JM, Alonso MA, Antignac C, Mollet G. INF2 mutations in Charcot-Marie-Tooth disease with glomerulopathy. The New England journal of medicine. 2011 Dec 22:365(25):2377-88. doi: 10.1056/NEJMoa1109122. Epub     [PubMed PMID: 22187985]

Level 3 (low-level) evidence

[68]

Mademan I, Deconinck T, Dinopoulos A, Voit T, Schara U, Devriendt K, Meijers B, Lerut E, De Jonghe P, Baets J. De novo INF2 mutations expand the genetic spectrum of hereditary neuropathy with glomerulopathy. Neurology. 2013 Nov 26:81(22):1953-8. doi: 10.1212/01.wnl.0000436615.58705.c9. Epub 2013 Oct 30     [PubMed PMID: 24174593]


[69]

Hattan E, Chalk C, Postuma RB. Is there a higher risk of restless legs syndrome in peripheral neuropathy? Neurology. 2009 Mar 17:72(11):955-60. doi: 10.1212/01.wnl.0000336341.72621.db. Epub 2008 Nov 26     [PubMed PMID: 19038854]

Level 2 (mid-level) evidence

[70]

Boentert M, Dziewas R, Heidbreder A, Happe S, Kleffner I, Evers S, Young P. Fatigue, reduced sleep quality and restless legs syndrome in Charcot-Marie-Tooth disease: a web-based survey. Journal of neurology. 2010 Apr:257(4):646-52. doi: 10.1007/s00415-009-5390-1. Epub 2009 Nov 24     [PubMed PMID: 19937049]

Level 2 (mid-level) evidence

[71]

Russo M, Laurá M, Polke JM, Davis MB, Blake J, Brandner S, Hughes RA, Houlden H, Bennett DL, Lunn MP, Reilly MM. Variable phenotypes are associated with PMP22 missense mutations. Neuromuscular disorders : NMD. 2011 Feb:21(2):106-14. doi: 10.1016/j.nmd.2010.11.011. Epub 2010 Dec 30     [PubMed PMID: 21194947]

Level 3 (low-level) evidence

[72]

Bansagi B, Antoniadi T, Burton-Jones S, Murphy SM, McHugh J, Alexander M, Wells R, Davies J, Hilton-Jones D, Lochmüller H, Chinnery P, Horvath R. Genotype/phenotype correlations in AARS-related neuropathy in a cohort of patients from the United Kingdom and Ireland. Journal of neurology. 2015 Aug:262(8):1899-908. doi: 10.1007/s00415-015-7778-4. Epub 2015 Jun 2     [PubMed PMID: 26032230]


[73]

Saporta AS, Sottile SL, Miller LJ, Feely SM, Siskind CE, Shy ME. Charcot-Marie-Tooth disease subtypes and genetic testing strategies. Annals of neurology. 2011 Jan:69(1):22-33. doi: 10.1002/ana.22166. Epub     [PubMed PMID: 21280073]


[74]

Dubourg O, Tardieu S, Birouk N, Gouider R, Léger JM, Maisonobe T, Brice A, Bouche P, LeGuern E. Clinical, electrophysiological and molecular genetic characteristics of 93 patients with X-linked Charcot-Marie-Tooth disease. Brain : a journal of neurology. 2001 Oct:124(Pt 10):1958-67     [PubMed PMID: 11571214]

Level 2 (mid-level) evidence

[75]

Siskind CE, Murphy SM, Ovens R, Polke J, Reilly MM, Shy ME. Phenotype expression in women with CMT1X. Journal of the peripheral nervous system : JPNS. 2011 Jun:16(2):102-7. doi: 10.1111/j.1529-8027.2011.00332.x. Epub     [PubMed PMID: 21692908]


[76]

Pareyson D, Scaioli V, Laurà M. Clinical and electrophysiological aspects of Charcot-Marie-Tooth disease. Neuromolecular medicine. 2006:8(1-2):3-22     [PubMed PMID: 16775364]


[77]

Tabaraud F, Lagrange E, Sindou P, Vandenberghe A, Levy N, Vallat JM. Demyelinating X-linked Charcot-Marie-Tooth disease: unusual electrophysiological findings. Muscle & nerve. 1999 Oct:22(10):1442-7     [PubMed PMID: 10487913]

Level 3 (low-level) evidence

[78]

Gutierrez A, England JD, Sumner AJ, Ferer S, Warner LE, Lupski JR, Garcia CA. Unusual electrophysiological findings in X-linked dominant Charcot-Marie-Tooth disease. Muscle & nerve. 2000 Feb:23(2):182-8     [PubMed PMID: 10639608]


[79]

Klein CJ, Duan X, Shy ME. Inherited neuropathies: clinical overview and update. Muscle & nerve. 2013 Oct:48(4):604-22. doi: 10.1002/mus.23775. Epub 2013 Jun 26     [PubMed PMID: 23801417]

Level 3 (low-level) evidence

[80]

McCorquodale D, Pucillo EM, Johnson NE. Management of Charcot-Marie-Tooth disease: improving long-term care with a multidisciplinary approach. Journal of multidisciplinary healthcare. 2016:9():7-19. doi: 10.2147/JMDH.S69979. Epub 2016 Jan 19     [PubMed PMID: 26855581]


[81]

Stangler Herodez S, Zagradisnik B, Erjavec Skerget A, Zagorac A, Kokalj Vokac N. Molecular diagnosis of PMP22 gene duplications and deletions: comparison of different methods. The Journal of international medical research. 2009 Sep-Oct:37(5):1626-31     [PubMed PMID: 19930872]


[82]

Hung CC, Lee CN, Lin CY, Cheng WF, Chen CA, Hsieh ST, Yang CC, Jong YJ, Su YN, Lin WL. Identification of deletion and duplication genotypes of the PMP22 gene using PCR-RFLP, competitive multiplex PCR, and multiplex ligation-dependent probe amplification: a comparison. Electrophoresis. 2008 Feb:29(3):618-25. doi: 10.1002/elps.200700340. Epub     [PubMed PMID: 18200636]

Level 2 (mid-level) evidence

[83]

Record CJ, Pipis M, Skorupinska M, Blake J, Poh R, Polke JM, Eggleton K, Nanji T, Zuchner S, Cortese A, Houlden H, Rossor AM, Laura M, Reilly MM. Whole genome sequencing increases the diagnostic rate in Charcot-Marie-Tooth disease. Brain : a journal of neurology. 2024 Mar 14:():. pii: awae064. doi: 10.1093/brain/awae064. Epub 2024 Mar 14     [PubMed PMID: 38481354]


[84]

Fogel BL, Satya-Murti S, Cohen BH. Clinical exome sequencing in neurologic disease. Neurology. Clinical practice. 2016 Apr:6(2):164-176     [PubMed PMID: 27104068]


[85]

Høyer H, Braathen GJ, Busk ØL, Holla ØL, Svendsen M, Hilmarsen HT, Strand L, Skjelbred CF, Russell MB. Genetic diagnosis of Charcot-Marie-Tooth disease in a population by next-generation sequencing. BioMed research international. 2014:2014():210401. doi: 10.1155/2014/210401. Epub 2014 Jun 16     [PubMed PMID: 25025039]


[86]

Li LX, Zhao SY, Liu ZJ, Ni W, Li HF, Xiao BG, Wu ZY. Improving molecular diagnosis of Chinese patients with Charcot-Marie-Tooth by targeted next-generation sequencing and functional analysis. Oncotarget. 2016 May 10:7(19):27655-64. doi: 10.18632/oncotarget.8377. Epub     [PubMed PMID: 27027447]


[87]

Padua L, Coraci D, Lucchetta M, Paolasso I, Pazzaglia C, Granata G, Cacciavillani M, Luigetti M, Manganelli F, Pisciotta C, Piscosquito G, Pareyson D, Briani C. Different nerve ultrasound patterns in charcot-marie-tooth types and hereditary neuropathy with liability to pressure palsies. Muscle & nerve. 2018 Jan:57(1):E18-E23. doi: 10.1002/mus.25766. Epub 2017 Aug 24     [PubMed PMID: 28802056]


[88]

Noto Y, Shiga K, Tsuji Y, Mizuta I, Higuchi Y, Hashiguchi A, Takashima H, Nakagawa M, Mizuno T. Nerve ultrasound depicts peripheral nerve enlargement in patients with genetically distinct Charcot-Marie-Tooth disease. Journal of neurology, neurosurgery, and psychiatry. 2015 Apr:86(4):378-84. doi: 10.1136/jnnp-2014-308211. Epub 2014 Aug 4     [PubMed PMID: 25091364]

Level 2 (mid-level) evidence

[89]

Zanette G, Tamburin S, Taioli F, Lauriola MF, Badari A, Ferrarini M, Cavallaro T, Fabrizi GM. Nerve size correlates with clinical severity in Charcot-Marie-Tooth disease 1A. Muscle & nerve. 2019 Dec:60(6):744-748. doi: 10.1002/mus.26688. Epub 2019 Sep 10     [PubMed PMID: 31469427]


[90]

Cornett KMD, Wojciechowski E, Sman AD, Walker T, Menezes MP, Bray P, Halaki M, Burns J, FAST Study Group. Magnetic resonance imaging of the anterior compartment of the lower leg is a biomarker for weakness, disability, and impaired gait in childhood Charcot-Marie-Tooth disease. Muscle & nerve. 2019 Feb:59(2):213-217. doi: 10.1002/mus.26352. Epub 2018 Dec 19     [PubMed PMID: 30265406]


[91]

Adam MP, Feldman J, Mirzaa GM, Pagon RA, Wallace SE, Bean LJH, Gripp KW, Amemiya A, Bird TD. Charcot-Marie-Tooth Hereditary Neuropathy Overview. GeneReviews(®). 1993:():     [PubMed PMID: 20301532]

Level 3 (low-level) evidence

[92]

Mathis S, Magy L, Vallat JM. Therapeutic options in Charcot-Marie-Tooth diseases. Expert review of neurotherapeutics. 2015 Apr:15(4):355-66. doi: 10.1586/14737175.2015.1017471. Epub 2015 Feb 21     [PubMed PMID: 25703094]

Level 3 (low-level) evidence

[93]

Attarian S, Vallat JM, Magy L, Funalot B, Gonnaud PM, Lacour A, Péréon Y, Dubourg O, Pouget J, Micallef J, Franques J, Lefebvre MN, Ghorab K, Al-Moussawi M, Tiffreau V, Preudhomme M, Magot A, Leclair-Visonneau L, Stojkovic T, Bossi L, Lehert P, Gilbert W, Bertrand V, Mandel J, Milet A, Hajj R, Boudiaf L, Scart-Grès C, Nabirotchkin S, Guedj M, Chumakov I, Cohen D. An exploratory randomised double-blind and placebo-controlled phase 2 study of a combination of baclofen, naltrexone and sorbitol (PXT3003) in patients with Charcot-Marie-Tooth disease type 1A. Orphanet journal of rare diseases. 2014 Dec 18:9():199. doi: 10.1186/s13023-014-0199-0. Epub 2014 Dec 18     [PubMed PMID: 25519680]

Level 1 (high-level) evidence

[94]

Patzkó A, Bai Y, Saporta MA, Katona I, Wu X, Vizzuso D, Feltri ML, Wang S, Dillon LM, Kamholz J, Kirschner D, Sarkar FH, Wrabetz L, Shy ME. Curcumin derivatives promote Schwann cell differentiation and improve neuropathy in R98C CMT1B mice. Brain : a journal of neurology. 2012 Dec:135(Pt 12):3551-66. doi: 10.1093/brain/aws299. Epub     [PubMed PMID: 23250879]

Level 3 (low-level) evidence

[95]

Morena J, Gupta A, Hoyle JC. Charcot-Marie-Tooth: From Molecules to Therapy. International journal of molecular sciences. 2019 Jul 12:20(14):. doi: 10.3390/ijms20143419. Epub 2019 Jul 12     [PubMed PMID: 31336816]


[96]

Corrado B, Ciardi G, Bargigli C. Rehabilitation Management of the Charcot-Marie-Tooth Syndrome: A Systematic Review of the Literature. Medicine. 2016 Apr:95(17):e3278. doi: 10.1097/MD.0000000000003278. Epub     [PubMed PMID: 27124017]

Level 1 (high-level) evidence

[97]

Sman AD, Hackett D, Fiatarone Singh M, Fornusek C, Menezes MP, Burns J. Systematic review of exercise for Charcot-Marie-Tooth disease. Journal of the peripheral nervous system : JPNS. 2015 Dec:20(4):347-62. doi: 10.1111/jns.12116. Epub     [PubMed PMID: 26010435]

Level 1 (high-level) evidence

[98]

Sivera Mascaró R, García Sobrino T, Horga Hernández A, Pelayo Negro AL, Alonso Jiménez A, Antelo Pose A, Calabria Gallego MD, Casasnovas C, Cemillán Fernández CA, Esteban Pérez J, Fenollar Cortés M, Frasquet Carrera M, Gallano Petit MP, Giménez Muñoz A, Gutiérrez Gutiérrez G, Gutiérrez Martínez A, Juntas Morales R, Ciano-Petersen NL, Martínez Ulloa PL, Mederer Hengstl S, Millet Sancho E, Navacerrada Barrero FJ, Navarrete Faubel FE, Pardo Fernández J, Pascual Pascual SI, Pérez Lucas J, Pino Mínguez J, Rabasa Pérez M, Sánchez González M, Sotoca J, Rodríguez Santiago B, Rojas García R, Turon-Sans J, Vicent Carsí V, Sevilla Mantecón T. Clinical practice guidelines for the diagnosis and management of Charcot-Marie-Tooth disease. Neurologia. 2024 Mar 1:():. pii: S2173-5808(24)00047-6. doi: 10.1016/j.nrleng.2024.02.008. Epub 2024 Mar 1     [PubMed PMID: 38431252]

Level 1 (high-level) evidence

[99]

Yiu EM, Bray P, Baets J, Baker SK, Barisic N, de Valle K, Estilow T, Farrar MA, Finkel RS, Haberlová J, Kennedy RA, Moroni I, Nicholson GA, Ramchandren S, Reilly MM, Rose K, Shy ME, Siskind CE, Yum SW, Menezes MP, Ryan MM, Burns J. Clinical practice guideline for the management of paediatric Charcot-Marie-Tooth disease. Journal of neurology, neurosurgery, and psychiatry. 2022 May:93(5):530-538. doi: 10.1136/jnnp-2021-328483. Epub 2022 Feb 9     [PubMed PMID: 35140138]

Level 1 (high-level) evidence

[100]

Pareyson D. Differential diagnosis of Charcot-Marie-Tooth disease and related neuropathies. Neurological sciences : official journal of the Italian Neurological Society and of the Italian Society of Clinical Neurophysiology. 2004 Jun:25(2):72-82     [PubMed PMID: 15221625]


[101]

Pitt M, Houlden H, Jacobs J, Mok Q, Harding B, Reilly M, Surtees R. Severe infantile neuropathy with diaphragmatic weakness and its relationship to SMARD1. Brain : a journal of neurology. 2003 Dec:126(Pt 12):2682-92     [PubMed PMID: 14506069]

Level 3 (low-level) evidence

[102]

Rossor AM, Carr AS, Devine H, Chandrashekar H, Pelayo-Negro AL, Pareyson D, Shy ME, Scherer SS, Reilly MM. Peripheral neuropathy in complex inherited diseases: an approach to diagnosis. Journal of neurology, neurosurgery, and psychiatry. 2017 Oct:88(10):846-863. doi: 10.1136/jnnp-2016-313960. Epub 2017 Aug 9     [PubMed PMID: 28794150]


[103]

Murphy SM, Herrmann DN, McDermott MP, Scherer SS, Shy ME, Reilly MM, Pareyson D. Reliability of the CMT neuropathy score (second version) in Charcot-Marie-Tooth disease. Journal of the peripheral nervous system : JPNS. 2011 Sep:16(3):191-8. doi: 10.1111/j.1529-8027.2011.00350.x. Epub     [PubMed PMID: 22003934]


[104]

Burns J, Ouvrier R, Estilow T, Shy R, Laurá M, Pallant JF, Lek M, Muntoni F, Reilly MM, Pareyson D, Acsadi G, Shy ME, Finkel RS. Validation of the Charcot-Marie-Tooth disease pediatric scale as an outcome measure of disability. Annals of neurology. 2012 May:71(5):642-52. doi: 10.1002/ana.23572. Epub     [PubMed PMID: 22522479]

Level 1 (high-level) evidence

[105]

Rossor AM, Shy ME, Reilly MM. Are we prepared for clinical trials in Charcot-Marie-Tooth disease? Brain research. 2020 Feb 15:1729():146625. doi: 10.1016/j.brainres.2019.146625. Epub 2019 Dec 30     [PubMed PMID: 31899213]


[106]

Wang H, Davison M, Wang K, Xia TH, Kramer M, Call K, Luo J, Wu X, Zuccarino R, Bacon C, Bai Y, Moran JJ, Gutmann L, Feely SME, Grider T, Rossor AM, Reilly MM, Svaren J, Shy ME. Transmembrane protease serine 5: a novel Schwann cell plasma marker for CMT1A. Annals of clinical and translational neurology. 2020 Jan:7(1):69-82. doi: 10.1002/acn3.50965. Epub 2019 Dec 12     [PubMed PMID: 31833243]


[107]

Fledrich R, Mannil M, Leha A, Ehbrecht C, Solari A, Pelayo-Negro AL, Berciano J, Schlotter-Weigel B, Schnizer TJ, Prukop T, Garcia-Angarita N, Czesnik D, Haberlová J, Mazanec R, Paulus W, Beissbarth T, Walter MC, Triaal C, Hogrel JY, Dubourg O, Schenone A, Baets J, De Jonghe P, Shy ME, Horvath R, Pareyson D, Seeman P, Young P, Sereda MW. Biomarkers predict outcome in Charcot-Marie-Tooth disease 1A. Journal of neurology, neurosurgery, and psychiatry. 2017 Nov:88(11):941-952. doi: 10.1136/jnnp-2017-315721. Epub 2017 Aug 31     [PubMed PMID: 28860329]


[108]

Laurá M, Singh D, Ramdharry G, Morrow J, Skorupinska M, Pareyson D, Burns J, Lewis RA, Scherer SS, Herrmann DN, Cullen N, Bradish C, Gaiani L, Martinelli N, Gibbons P, Pfeffer G, Phisitkul P, Wapner K, Sanders J, Flemister S, Shy ME, Reilly MM, Inherited Neuropathies Consortium. Prevalence and orthopedic management of foot and ankle deformities in Charcot-Marie-Tooth disease. Muscle & nerve. 2018 Feb:57(2):255-259. doi: 10.1002/mus.25724. Epub 2017 Jul 7     [PubMed PMID: 28632967]


[109]

Pfeffer GB, Gonzalez T, Brodsky J, Campbell J, Coetzee C, Conti S, Guyton G, Herrmann DN, Hunt K, Johnson J, McGarvey W, Pinzur M, Raikin S, Sangeorzan B, Younger A, Michalski M, An T, Noori N. A Consensus Statement on the Surgical Treatment of Charcot-Marie-Tooth Disease. Foot & ankle international. 2020 Jul:41(7):870-880. doi: 10.1177/1071100720922220. Epub 2020 Jun 1     [PubMed PMID: 32478578]

Level 3 (low-level) evidence