ﺑﺎﺯﮔﺸﺖ ﺑﻪ ﺻﻔﺤﻪ ﻗﺒﻠﯽ
خرید پکیج
تعداد آیتم قابل مشاهده باقیمانده : 3 مورد
نسخه الکترونیک
medimedia.ir

Duchenne and Becker muscular dystrophy: Genetics and pathogenesis

Duchenne and Becker muscular dystrophy: Genetics and pathogenesis
Literature review current through: Jan 2024.
This topic last updated: Jan 27, 2023.

INTRODUCTION — The Duchenne and Becker muscular dystrophies (as well as a third intermediate form) are caused by pathogenic variants of the dystrophin gene and are therefore named dystrophinopathies. Weakness is the principal symptom as muscle fiber degeneration is the primary pathologic process.

The genetics and pathogenesis of Duchenne and Becker muscular dystrophy are reviewed here. Other clinical aspects of these conditions are discussed elsewhere. (See "Duchenne and Becker muscular dystrophy: Clinical features and diagnosis" and "Duchenne and Becker muscular dystrophy: Management and prognosis".)

TERMINOLOGY — The dystrophinopathies are inherited as X-linked recessive traits and have varying clinical characteristics:

Duchenne muscular dystrophy (DMD) is associated with the most severe clinical symptoms (see "Duchenne and Becker muscular dystrophy: Clinical features and diagnosis", section on 'Duchenne muscular dystrophy')

Becker muscular dystrophy (BMD) has a similar presentation to DMD, but typically has a later onset and a milder clinical course (see "Duchenne and Becker muscular dystrophy: Clinical features and diagnosis", section on 'Becker muscular dystrophy')

Patients with an intermediate phenotype (outliers) may be classified clinically as having either mild DMD or severe BMD (see "Duchenne and Becker muscular dystrophy: Clinical features and diagnosis", section on 'Intermediate phenotype')

GENETICS AND PATHOGENESIS — DMD and BMD are caused by a defective gene located on the X chromosome that is responsible for the production of dystrophin [1-3]. The dystrophin (DMD) gene is the largest gene yet identified in humans, spanning approximately 2.3 megabases at chromosome Xp21.2. The protein product is also extremely large, weighing 427 kilodaltons (kD) [4].

Pathogenic variants in the DMD gene — The majority of pathogenic variants of the DMD gene are deletions of one or more exons, which are found in approximately 68 to 77 and 65 to 70 percent of patients with DMD and BMD, respectively [5-9]. Partial gene duplications have also been reported in approximately 11 to 13 percent of affected individuals [7,10]; duplication of DMD exon 2 is the most common, accounting for approximately 6 to 10 percent of all DMD duplications [11,12].

The DMD pathogenic variants in the remaining patients (ie, without detectable deletions or duplications) are single nucleotide variants, small deletions or insertions in the coding sequence, or splice site variants. Patients with clinical phenotypes suggestive of DMD or BMD, but without a clear X-linked pattern of inheritance, may have defects in other genes, including those encoding the dystrophin-associated glycoproteins.

Dystrophin — The phenotype of dystrophinopathy is primarily dependent upon the quantity of residual dystrophin in muscle, with absent or minimal dystrophin (eg, 0 to 5 percent of normal) correlating with a more severe phenotype (DMD) and higher levels of dystrophin (5 to 50 percent of normal) correlating with less severe phenotypes (intermediate or BMD) [10]; other reports suggest that very low dystrophin levels (<5 percent) may be associated with a milder phenotype in a subset of patients [13]. However, these correlations are imprecise and confounded primarily by variations in Western blot quantification methods and the effect of genetic modifiers on disease course.

Dystrophin is located on the cytoplasmic face of the plasma membrane of muscle fibers, functioning as a component of a large, tightly associated glycoprotein complex (figure 1) [3]. Dystrophin normally provides mechanical reinforcement to the sarcolemma and stabilizes the glycoprotein complex, thereby shielding it from degradation. In its absence, the glycoprotein complex is digested by proteases. Loss of these membrane proteins may initiate the degeneration of muscle fibers, resulting in muscle weakness.

The loss of dystrophin in mdx mice leads to myofibril membrane instability [14]. However, genetic disruption of the dystrophin gene in mdx mice is associated with only a mild dystrophy [15]. Although the reduced disease severity in mdx mice compared with human DMD is not fully understood, several possible explanations are postulated. As an example, a homolog of dystrophin called utrophin is present in mice and humans. Its expression in muscle can compensate physiologically for the absence of dystrophin in mice, but this compensation does not occur in humans. This hypothesis is supported by the finding that mice lacking both dystrophin and utrophin have a severe dystrophy that phenotypically resembles DMD [15]. Further, the selective expression in skeletal muscle of utrophin via the use of a transgene completely rescues these double-knockout mice from early death and the DMD phenotype [16].

As part of the glycoprotein complex, dystrophin secures a number of dystrophin-associated proteins (figure 2), including neuronal nitric oxide synthase (nNOS), to the sarcolemma [17]. Sarcolemmal nNOS is necessary for the production of nitric oxide, which regulates vasodilation and increased blood flow into muscle, and is important for the prevention of early muscle fatigue with exercise [18-20]. The absence of dystrophin in humans with DMD or mdx mice is associated with a loss of muscle nNOS [17,21], resulting in exercise-induced muscle fatigue [18-20,22].

Disruption of calcium regulation may also play a role in the pathogenesis of DMD [23-26]. Muscle cell membrane damage related to the loss of dystrophin may permit the pathologic entry of extracellular calcium into muscle fibers. In addition, inflammatory mediators in dystrophic muscle may increase the expression of inducible nitric oxide synthase (iNOS), which binds to and destabilizes ryanodine receptors of the sarcoplasmic reticulum that regulate calcium ion flow [27,28]. The result is calcium leakage from the sarcoplasmic reticulum into the cytosol. The excess cytosolic calcium can activate calpains, which promote muscle proteolysis [29,30].

Genotype-phenotype correlations — There is no clear correlation between the size of DMD gene deletions and the severity and progression of the DMD/BMD phenotype [31,32]. The molecular basis of Duchenne (more severe) versus Becker (less severe) muscular dystrophy is related, at least in part, to whether the amino acid translational reading frame is disrupted or preserved by the deletion [5,10,33]. This "reading frame rule," and important exceptions to it, are reviewed in the sections that follow.

Reading frame rule — During the synthesis of mature mRNA, the joining ends of the exons (after splicing of introns) must be in phase to maintain the correct translational open reading frame.

Out-of-frame deletions – A deletion that juxtaposes exons and shifts the translational reading frame (ie, an out-of-frame deletion) most often leads to unstable messenger RNA (mRNA) and the production of a severely truncated dystrophin molecule (figure 3). Truncated protein is rapidly degraded in the cell by nonsense-mediated mRNA decay and results in a DMD (severe) phenotype [33].

In-frame deletions – A deletion that juxtaposes exons but preserves the translational reading frame (eg, an in-frame deletion) usually results in an internally deleted but semifunctional dystrophin protein with intact amino-terminal and carboxy-terminal domains (figure 3). Semi-functional dystrophin protein can persist in some quantity in the cell, and results in a BMD (mild) phenotype.

Exceptions to the reading frame rule — Exceptions to the reading frame rule as a predictor phenotype occur in approximately 8 percent of DMD cases [34-40] and up to 34 percent of BMD cases [7,41]. These exceptions may occur because of variants in critical domains, exon skipping events, post-translational mechanisms, and other factors. Thus, affected males and their families and caregivers should be informed that phenotype prediction may not be accurate when based on the reading frame rule [10].

Some regions of the dystrophin protein are more critical to function than others. Deletions in the regions coding for the amino-terminal or carboxy-terminal domains of dystrophin often result in more severe phenotypes than pathogenic variants affecting the rod domain [32,36]. However, exceptions occur [40]. Very large deletions or deletions in protein-binding domains (eg, cysteine-rich domain), even if in-frame, can result in a severe phenotype [37,38].

Out-of-frame deletions of exons 3 to 7 or exon 45 can result in BMD, DMD, or an intermediate phenotype [39,42]. Exon-skipping events or activation of new cryptic translational start sites may create situations in which apparently out-of-frame deletions behave as in-frame deletions or vice versa [43]. Alternatively, post-transcriptional mechanisms (eg, ribosomal frame shifting or reinitiation) may help restore the mRNA reading frame in patients with BMD who have out-of-frame deletions involving exons 3 to 7 [39,44]. Variations in the severity of the phenotype among patients with similar pathogenic variants (eg, deletion of exon 45) have been reported, supporting the contribution of other factors in the phenotypic expression of these variants [36,45].

Certain missense or nonsense variants may occur within exonic splicing enhancers (ESEs) and induce exon skipping with variable results on the phenotype [46]. For example, a nonsense variant within exon 27 disrupts an ESE sequence and induces partial skipping of the exon, which results in a BMD phenotype [47]. Nonsense variants within DMD muscle isoform exon 1 have been associated with an extremely mild BMD phenotype with loss of ambulation in the seventh decade. It seems that internal initiation of translation at two AUG codons within exon 6 ameliorates the phenotype [48].

Specific examples follow:

DMD phenotype with in-frame deletions – Large in-frame deletions in the 5' region that extend into the mid-rod domain (eg, exons 3 to 31, 3 to 25, 4 to 41, or 4 to 18) result in severe phenotype [49]. By contrast, large deletions in the rod domain that do not involve the 5' putative actin binding site of dystrophin result in a mild BMD phenotype [50-53]. However, small in-frame deletions, such as the deletion of exons 3 to 13, which disrupt the 5' actin binding domain, are associated with a DMD phenotype [54]. These exceptions underscore the functional relevance of the 5' actin binding domain.

BMD phenotype with out-of-frame deletions Out-of-frame deletions or duplications involving the 5' end of the gene (exons 3 to 7, 5 to 7, 3 to 6) or further downstream (exons 51, 49 to 50, 47 to 52, 44 or 45) have been described in patients with BMD. The carboxy-terminal domain is not disrupted in these patients and some dystrophin is produced because of exon skipping that occurs via alternative splicing. The exon skipping events probably create larger mRNA deletions, which presumably are in-frame and hence functional [55-57]. In patients with BMD and out-of-frame deletions involving exons 3 to 7, it is hypothesized that an alternative translation initiation site within exon 8 becomes activated [58].

The DMD exon 2 duplication is an out-of-frame pathogenic variant and, as such, is expected to result in a frameshift and thus a severe DMD phenotype. However, milder phenotypes (intermediate and BMD) have been described in cohorts of individuals with DMD exon 2 duplications, possibly related to activation of an internal ribosome entry site in DMD exon 5, which can drive production of mildly truncated but functional dystrophin from an AUG codon within exon 6 [11,12].

In men with BMD, deletions involving the amino-terminal domain (5' end) have been correlated with an earlier-onset (third decade) dilated cardiomyopathy, whereas deletions affecting part of the rod domain and hinge 3 have been correlated with a later-onset (fifth decade) dilated cardiomyopathy [10,59].

Other genetic modifiers — Some data suggest that genes other than dystrophin may affect disease severity and response to treatment [60-64]. As an example, variants of the LTBP4 gene may influence the age at loss of ambulation in DMD [61-63]. In addition, a few reports suggested that a variant of the SPP1 gene promoter region (the G allele of the polymorphism rs28357094) was associated with decreased muscle strength and younger age at loss of ambulation in patients with DMD, and appeared to modulate glucocorticoid treatment response [60,62]. However, another study found no significant association of the SPP1 rs28357094 polymorphism with age of ambulation loss [63]. Further studies are needed to clarify these relationships.

SOCIETY GUIDELINE LINKS — Links to society and government-sponsored guidelines from selected countries and regions around the world are provided separately. (See "Society guideline links: Muscular dystrophy".)

SUMMARY

Duchenne and Becker muscular dystrophies (DMD and BMD) are caused by a defective gene located on the X chromosome that is responsible for the production of dystrophin. (See 'Genetics and pathogenesis' above.)

Dystrophin is located on the cytoplasmic face of the plasma membrane of muscle fibers, functioning as a component of a large, tightly associated glycoprotein complex (figure 1). Dystrophin normally provides mechanical reinforcement to the sarcolemma and stabilizes the glycoprotein complex, thereby shielding it from degradation. In its absence, the glycoprotein complex is digested by proteases. Loss of these membrane proteins may initiate the degeneration of muscle fibers, resulting in muscle weakness. (See 'Dystrophin' above.)

There is no clear correlation between the size of dystrophin (DMD) gene deletions and the severity and progression of the DMD/BMD phenotype. The molecular basis of DMD (more severe) versus BMD (less severe) is partly related to whether the amino acid translational reading frame is disrupted or preserved by the deletion. However, there are exceptions to this general rule. (See 'Genotype-phenotype correlations' above.)

  1. Worton R. Muscular dystrophies: diseases of the dystrophin-glycoprotein complex. Science 1995; 270:755.
  2. Kunkel LM, Hejtmancik JF, Caskey CT, et al. Analysis of deletions in DNA from patients with Becker and Duchenne muscular dystrophy. Nature 1986; 322:73.
  3. Ervasti JM, Ohlendieck K, Kahl SD, et al. Deficiency of a glycoprotein component of the dystrophin complex in dystrophic muscle. Nature 1990; 345:315.
  4. Hoffman EP, Brown RH Jr, Kunkel LM. Dystrophin: the protein product of the Duchenne muscular dystrophy locus. Cell 1987; 51:919.
  5. Darras BT, Menache-Stroninki CC, Hinton V, Kunkel LM. Dystrophinopathies. In: Neuromuscular Disorders of Infancy, Childhood and Adolescence: A Clinician’s Approach, 2nd ed, Darras BT, Jones HR Jr, Ryan MM, De Vivo DC (Eds), Academic Press, San Diego 2015. p.551.
  6. Aartsma-Rus A, Van Deutekom JC, Fokkema IF, et al. Entries in the Leiden Duchenne muscular dystrophy mutation database: an overview of mutation types and paradoxical cases that confirm the reading-frame rule. Muscle Nerve 2006; 34:135.
  7. Takeshima Y, Yagi M, Okizuka Y, et al. Mutation spectrum of the dystrophin gene in 442 Duchenne/Becker muscular dystrophy cases from one Japanese referral center. J Hum Genet 2010; 55:379.
  8. Tuffery-Giraud S, Béroud C, Leturcq F, et al. Genotype-phenotype analysis in 2,405 patients with a dystrophinopathy using the UMD-DMD database: a model of nationwide knowledgebase. Hum Mutat 2009; 30:934.
  9. Bladen CL, Salgado D, Monges S, et al. The TREAT-NMD DMD Global Database: analysis of more than 7,000 Duchenne muscular dystrophy mutations. Hum Mutat 2015; 36:395.
  10. Darras BT, Urion DK, Ghosh PS. Dystrophinopathies. GeneReviews. https://www.ncbi.nlm.nih.gov/books/NBK1119/ (Accessed on February 03, 2021).
  11. Zambon AA, Waldrop MA, Alles R, et al. Phenotypic Spectrum of Dystrophinopathy Due to Duchenne Muscular Dystrophy Exon 2 Duplications. Neurology 2022; 98:e730.
  12. Juan-Mateu J, Gonzalez-Quereda L, Rodriguez MJ, et al. DMD Mutations in 576 Dystrophinopathy Families: A Step Forward in Genotype-Phenotype Correlations. PLoS One 2015; 10:e0135189.
  13. de Feraudy Y, Ben Yaou R, Wahbi K, et al. Very Low Residual Dystrophin Quantity Is Associated with Milder Dystrophinopathy. Ann Neurol 2021; 89:280.
  14. Petrof BJ, Shrager JB, Stedman HH, et al. Dystrophin protects the sarcolemma from stresses developed during muscle contraction. Proc Natl Acad Sci U S A 1993; 90:3710.
  15. Grady RM, Teng H, Nichol MC, et al. Skeletal and cardiac myopathies in mice lacking utrophin and dystrophin: a model for Duchenne muscular dystrophy. Cell 1997; 90:729.
  16. Rafael JA, Tinsley JM, Potter AC, et al. Skeletal muscle-specific expression of a utrophin transgene rescues utrophin-dystrophin deficient mice. Nat Genet 1998; 19:79.
  17. Lai Y, Thomas GD, Yue Y, et al. Dystrophins carrying spectrin-like repeats 16 and 17 anchor nNOS to the sarcolemma and enhance exercise performance in a mouse model of muscular dystrophy. J Clin Invest 2009; 119:624.
  18. Sander M, Chavoshan B, Harris SA, et al. Functional muscle ischemia in neuronal nitric oxide synthase-deficient skeletal muscle of children with Duchenne muscular dystrophy. Proc Natl Acad Sci U S A 2000; 97:13818.
  19. Percival JM, Anderson KN, Gregorevic P, et al. Functional deficits in nNOSmu-deficient skeletal muscle: myopathy in nNOS knockout mice. PLoS One 2008; 3:e3387.
  20. Kobayashi YM, Rader EP, Crawford RW, et al. Sarcolemma-localized nNOS is required to maintain activity after mild exercise. Nature 2008; 456:511.
  21. Brenman JE, Chao DS, Xia H, et al. Nitric oxide synthase complexed with dystrophin and absent from skeletal muscle sarcolemma in Duchenne muscular dystrophy. Cell 1995; 82:743.
  22. Heydemann A, McNally E. NO more muscle fatigue. J Clin Invest 2009; 119:448.
  23. Bodensteiner JB, Engel AG. Intracellular calcium accumulation in Duchenne dystrophy and other myopathies: a study of 567,000 muscle fibers in 114 biopsies. Neurology 1978; 28:439.
  24. Fong PY, Turner PR, Denetclaw WF, Steinhardt RA. Increased activity of calcium leak channels in myotubes of Duchenne human and mdx mouse origin. Science 1990; 250:673.
  25. Robert V, Massimino ML, Tosello V, et al. Alteration in calcium handling at the subcellular level in mdx myotubes. J Biol Chem 2001; 276:4647.
  26. Blake DJ, Weir A, Newey SE, Davies KE. Function and genetics of dystrophin and dystrophin-related proteins in muscle. Physiol Rev 2002; 82:291.
  27. Bellinger AM, Reiken S, Carlson C, et al. Hypernitrosylated ryanodine receptor calcium release channels are leaky in dystrophic muscle. Nat Med 2009; 15:325.
  28. Tidball JG, Villalta SA. NO may prompt calcium leakage in dystrophic muscle. Nat Med 2009; 15:243.
  29. Turner PR, Westwood T, Regen CM, Steinhardt RA. Increased protein degradation results from elevated free calcium levels found in muscle from mdx mice. Nature 1988; 335:735.
  30. Spencer MJ, Croall DE, Tidball JG. Calpains are activated in necrotic fibers from mdx dystrophic mice. J Biol Chem 1995; 270:10909.
  31. Darras BT, Blattner P, Harper JF, et al. Intragenic deletions in 21 Duchenne muscular dystrophy (DMD)/Becker muscular dystrophy (BMD) families studied with the dystrophin cDNA: location of breakpoints on HindIII and BglII exon-containing fragment maps, meiotic and mitotic origin of the mutations. Am J Hum Genet 1988; 43:620.
  32. Baumbach LL, Chamberlain JS, Ward PA, et al. Molecular and clinical correlations of deletions leading to Duchenne and Becker muscular dystrophies. Neurology 1989; 39:465.
  33. Monaco AP, Bertelson CJ, Liechti-Gallati S, et al. An explanation for the phenotypic differences between patients bearing partial deletions of the DMD locus. Genomics 1988; 2:90.
  34. Koenig M, Beggs AH, Moyer M, et al. The molecular basis for Duchenne versus Becker muscular dystrophy: correlation of severity with type of deletion. Am J Hum Genet 1989; 45:498.
  35. Malhotra SB, Hart KA, Klamut HJ, et al. Frame-shift deletions in patients with Duchenne and Becker muscular dystrophy. Science 1988; 242:755.
  36. Beggs AH, Hoffman EP, Snyder JR, et al. Exploring the molecular basis for variability among patients with Becker muscular dystrophy: dystrophin gene and protein studies. Am J Hum Genet 1991; 49:54.
  37. Hoffman EP, Garcia CA, Chamberlain JS, et al. Is the carboxyl-terminus of dystrophin required for membrane association? A novel, severe case of Duchenne muscular dystrophy. Ann Neurol 1991; 30:605.
  38. Bies RD, Caskey CT, Fenwick R. An intact cysteine-rich domain is required for dystrophin function. J Clin Invest 1992; 90:666.
  39. Gangopadhyay SB, Sherratt TG, Heckmatt JZ, et al. Dystrophin in frameshift deletion patients with Becker muscular dystrophy. Am J Hum Genet 1992; 51:562.
  40. Helliwell TR, Ellis JM, Mountford RC, et al. A truncated dystrophin lacking the C-terminal domains is localized at the muscle membrane. Am J Hum Genet 1992; 50:508.
  41. Kesari A, Pirra LN, Bremadesam L, et al. Integrated DNA, cDNA, and protein studies in Becker muscular dystrophy show high exception to the reading frame rule. Hum Mutat 2008; 29:728.
  42. Winnard AV, Klein CJ, Coovert DD, et al. Characterization of translational frame exception patients in Duchenne/Becker muscular dystrophy. Hum Mol Genet 1993; 2:737.
  43. Chelly J, Gilgenkrantz H, Lambert M, et al. Effect of dystrophin gene deletions on mRNA levels and processing in Duchenne and Becker muscular dystrophies. Cell 1990; 63:1239.
  44. Bushby KM. Genetic and clinical correlations of Xp21 muscular dystrophy. J Inherit Metab Dis 1992; 15:551.
  45. Prior TW, Bartolo C, Papp AC, et al. Dystrophin expression in a Duchenne muscular dystrophy patient with a frame shift deletion. Neurology 1997; 48:486.
  46. Ginjaar IB, Kneppers AL, v d Meulen JD, et al. Dystrophin nonsense mutation induces different levels of exon 29 skipping and leads to variable phenotypes within one BMD family. Eur J Hum Genet 2000; 8:793.
  47. Shiga N, Takeshima Y, Sakamoto H, et al. Disruption of the splicing enhancer sequence within exon 27 of the dystrophin gene by a nonsense mutation induces partial skipping of the exon and is responsible for Becker muscular dystrophy. J Clin Invest 1997; 100:2204.
  48. Gurvich OL, Maiti B, Weiss RB, et al. DMD exon 1 truncating point mutations: amelioration of phenotype by alternative translation initiation in exon 6. Hum Mutat 2009; 30:633.
  49. Nevo Y, Muntoni F, Sewry C, et al. Large in-frame deletions of the rod-shaped domain of the dystrophin gene resulting in severe phenotype. Isr Med Assoc J 2003; 5:94.
  50. Love DR, Flint TJ, Marsden RF, et al. Characterization of deletions in the dystrophin gene giving mild phenotypes. Am J Med Genet 1990; 37:136.
  51. Bushby KM, Gardner-Medwin D, Nicholson LV, et al. The clinical, genetic and dystrophin characteristics of Becker muscular dystrophy. II. Correlation of phenotype with genetic and protein abnormalities. J Neurol 1993; 240:105.
  52. England SB, Nicholson LV, Johnson MA, et al. Very mild muscular dystrophy associated with the deletion of 46% of dystrophin. Nature 1990; 343:180.
  53. Love DR, Flint TJ, Genet SA, et al. Becker muscular dystrophy patient with a large intragenic dystrophin deletion: implications for functional minigenes and gene therapy. J Med Genet 1991; 28:860.
  54. Muntoni F, Gobbi P, Sewry C, et al. Deletions in the 5' region of dystrophin and resulting phenotypes. J Med Genet 1994; 31:843.
  55. Arahata K, Beggs AH, Honda H, et al. Preservation of the C-terminus of dystrophin molecule in the skeletal muscle from Becker muscular dystrophy. J Neurol Sci 1991; 101:148.
  56. Nicholson LV, Bushby KM, Johnson MA, et al. Predicted and observed sizes of dystrophin in some patients with gene deletions that disrupt the open reading frame. J Med Genet 1992; 29:892.
  57. Patria SY, Alimsardjono H, Nishio H, et al. A case of Becker muscular dystrophy resulting from the skipping of four contiguous exons (71-74) of the dystrophin gene during mRNA maturation. Proc Assoc Am Physicians 1996; 108:308.
  58. Winnard AV, Mendell JR, Prior TW, et al. Frameshift deletions of exons 3-7 and revertant fibers in Duchenne muscular dystrophy: mechanisms of dystrophin production. Am J Hum Genet 1995; 56:158.
  59. Kaspar RW, Allen HD, Ray WC, et al. Analysis of dystrophin deletion mutations predicts age of cardiomyopathy onset in becker muscular dystrophy. Circ Cardiovasc Genet 2009; 2:544.
  60. Pegoraro E, Hoffman EP, Piva L, et al. SPP1 genotype is a determinant of disease severity in Duchenne muscular dystrophy. Neurology 2011; 76:219.
  61. Flanigan KM, Ceco E, Lamar KM, et al. LTBP4 genotype predicts age of ambulatory loss in Duchenne muscular dystrophy. Ann Neurol 2013; 73:481.
  62. Bello L, Kesari A, Gordish-Dressman H, et al. Genetic modifiers of ambulation in the Cooperative International Neuromuscular Research Group Duchenne Natural History Study. Ann Neurol 2015; 77:684.
  63. van den Bergen JC, Hiller M, Böhringer S, et al. Validation of genetic modifiers for Duchenne muscular dystrophy: a multicentre study assessing SPP1 and LTBP4 variants. J Neurol Neurosurg Psychiatry 2015; 86:1060.
  64. Hightower RM, Alexander MS. Genetic modifiers of Duchenne and facioscapulohumeral muscular dystrophies. Muscle Nerve 2018; 57:6.
Topic 116654 Version 9.0

References

آیا می خواهید مدیلیب را به صفحه اصلی خود اضافه کنید؟