Comprehensive Physiology Wiley Online Library

Immunobiology of Inherited Muscular Dystrophies

Full Article on Wiley Online Library



ABSTRACT

The immune response to acute muscle damage is important for normal repair. However, in chronic diseases such as many muscular dystrophies, the immune response can amplify pathology and play a major role in determining disease severity. Muscular dystrophies are inheritable diseases that vary tremendously in severity, but share the progressive loss of muscle mass and function that can be debilitating and lethal. Mutations in diverse genes cause muscular dystrophy, including genes that encode proteins that maintain membrane strength, participate in membrane repair, or are components of the extracellular matrix or the nuclear envelope. In this article, we explore the hypothesis that an important feature of many muscular dystrophies is an immune response adapted to acute, infrequent muscle damage that is misapplied in the context of chronic injury. We discuss the involvement of the immune system in the most common muscular dystrophy, Duchenne muscular dystrophy, and show that the immune system influences muscle death and fibrosis as disease progresses. We then present information on immune cell function in other muscular dystrophies and show that for many muscular dystrophies, release of cytosolic proteins into the extracellular space may provide an initial signal, leading to an immune response that is typically dominated by macrophages, neutrophils, helper T‐lymphocytes, and cytotoxic T‐lymphocytes. Although those features are similar in many muscular dystrophies, each muscular dystrophy shows distinguishing features in the magnitude and type of inflammatory response. These differences indicate that there are disease‐specific immunomodulatory molecules that determine response to muscle cell damage caused by diverse genetic mutations. © 2018 American Physiological Society. Compr Physiol 8:1313‐1356, 2018.

Comprehensive Physiology offers downloadable PowerPoint presentations of figures for non-profit, educational use, provided the content is not modified and full credit is given to the author and publication.

Download a PowerPoint presentation of all images


Figure 1. Figure 1. Characteristics of Duchenne muscular dystrophy (DMD). (A) Typically, the clinical onset of DMD occurs at about 3 years of age as boys begin to show defects in muscle function. In this image of 5 boys with DMD at progressively older ages, some of the gross features of the disease are apparent. The boys show an increasingly progressive lumbar curvature of the spine that results in postural compensation for increased weakening of paravertebral muscles. There is also a progressive increase of weight bearing on the toes and reduction of weight bearing on the heels, as fibrosis of calf muscles cause contractures that limit plantar flexion. Although gross appearance shows an apparent sparing of calf muscles even in older boys, there is actually an increased replacement of muscle tissue with fibrous and fatty tissue, leading to a pseudohypertrophy of the calves. (B) Magnetic resonance spectroscopy images of the shoulders, upper arms, and forearms of healthy subjects (12, 13, and 13 years of age) and DMD patients (12, 13, and 13 years of age) showing tremendous reductions of muscle mass (white) by fatty tissue (dark). BB (biceps brachii), TB (triceps brachii), DEL (deltoid), subscapularis (SUB), infraspinatus (INF), posterior compartment of the forearm (PF), and anterior compartment of the forearm (AF). [Reproduced, with permission, from reference (361).] (C) Section of muscle biopsy of DMD patient stained with hematoxylin and eosin stain. Large accumulations of connective tissue separate individual muscle fibers and mononucleated leukocytes are present in elevated numbers in the interstitial tissue (arrow) and near blood vessels (brackets). Fiber cross‐sectional area is also highly variable, another characteristic of DMD pathology. Bar = 60 μm. [Reproduced, with permission, from reference (252).]
Figure 2. Figure 2. Schematic representation of dystrophin glycoprotein complex (DGC). Dystrophin provides an important structural link between the thin filaments within skeletal muscle fibers and β‐dystroglycan, which is a transmembrane protein. β‐dystroglycan then binds extracellularly to α‐dystroglycan which is a ligand for extracellular structural proteins, especially laminin‐2 that is present in basal lamina. Several proteins in the basal lamina, including laminin and fibronectin, then provide mechanical links to major connective tissue macromolecules such as collagen type 1. Genetic deletion of dystrophin disrupts this mechanical linkage between the cytoskeleton and extracellular structural proteins, but also leads to reductions in the quantity of other structural proteins in the DGC or associated with the DGC. The secondary loss of those other proteins, including the sarcoglycans, syntrophin and neuronal nitric oxide synthase (nNOS) can contribute to the pathology of dystrophin deficiency through disruption of signaling pathways that are necessary for normal muscle function. [Reproduced, with permission, from reference (319).]
Figure 3. Figure 3. Dystrophin deficiency causes membrane lesions and unregulated influx and efflux of large molecules. (A) Electron microscopy on DMD skeletal muscle shows morphologically detectible lesions in the cell membrane (sarcolemma) of dystrophic fibers. Asterisks (*) indicate basal lamina and other connective tissue associated with the extracellular surface of the muscle fiber. Blue arrows indicate sites of intact sarcolemma associated with dense, subsarcolemmal material. Red arrows indicate sites of lesions in the sarcolemma that would permit unrestricted transit of molecules across the membrane. Scale bar = 150 nm. [Reproduced, with permission, from reference (230).] [(B) and (C)] Intact hind limbs of wild‐type (B) or mdx mice (C) that had been injected with an extracellular marker dye, Evans blue, prior to euthanasia and tissue collection. Skin and fat has been removed from both limbs. The wild‐type muscles show little blue dye, indicating that little dye was able to cross the cell membrane to enter the cytosol of the muscle fibers. The mdx muscle shows some thigh muscles contain high concentrations of blue muscle fibers (red arrow) indicating unregulated entry of dye into the fibers through membrane lesions. In contrast, leg muscle in the same mouse shows little blue dye (blue arrow), illustrating the difference in magnitude of pathology and progression of the disease in different muscles. [Reproduced, with permission, from reference (303).] [(D) and (E)] Cross sections of soleus muscle from wild‐type (D) and mdx (E) mice. The muscles were incubated in a fluorescent, extracellular marker dye after dissection and before freezing the muscle for histology. In the wild‐type muscle, the fluorescent dye remains in the extracellular space because the muscle cell membranes are intact (D). In the mdx muscle, approximately 8% of the muscle fibers showed elevated intracellular fluorescence, indicating the presence of membrane lesions. Bars = 70 µm. [Reproduced, with permission, from reference (351).]
Figure 4. Figure 4. Macrophages are a primary source of muscle fiber damage in the mdx model of DMD. (A) At the early, acute peak of mdx muscle pathology, macrophages dominate large inflammatory lesions in the diseased muscle. Cross section of 4‐week‐old mdx mouse muscle with F4/80 + macrophages immunolabeled to appear red in the sectioned muscle. Bar = 180 µm. (B) Image shows complete, mid‐belly cross section of an entire soleus muscle from 4‐week‐old mdx mice. The mouse from which the muscle on the left was collected received intraperitoneal injections of sterile buffer on 5 days each week from 8 days of age until the mouse was euthanized at 4 weeks of age. The mouse from which the muscle on the right was collected received intraperitoneal injections of antibodies to the F4/80 antigen on the same injection schedule, which produced a reduction in macrophage numbers by over 90%. Both sections were stained with antibodies to neural cell adhesion molecule (NCAM) which is expressed at high levels by muscle fibers that are undergoing repair following injury. Each small, red tile‐like structure in the cross section is a recently injured fiber. Note that the macrophage‐depleted muscle is nearly devoid of NCAM‐expressing muscle fibers. Bar = 1.2 mm. [(C) and (D)] Soleus muscles from nondepleted, 4‐week‐old mdx mice (C) or macrophage‐depleted mdx mice (D) were incubated in Procion red, a fluorescent, extracellular marker dye before muscle sectioning and microscopy. Intracellular Procion red indicates fibers with membrane lesions. The number of Procion‐red‐containing fibers in the 4‐week‐old mdx solei was reduced by more than 75%. Bar = 120 µm. [Reproduced, with permission, from reference (351).]
Figure 5. Figure 5. Inflammation in the mdx mouse model of DMD. (A) Section of a 4‐week‐old mdx mouse muscle immunolabeled with anti‐F4/80, a pan‐macrophage marker (red). Note that some muscle fibers have been invaded and obliterated by large numbers of macrophages (white asterisk) while other, small regenerating fibers are surrounded but not invaded by macrophages (black asterisk). Bar = 50 μm. (B) Section of 4‐week‐old muscle immunolabeled with antibodies to F4/80 (red) and CD206 (green), a marker for M2‐biased macrophages. Blue structures are nuclei binding DAPI reagent. This inflammatory lesion in dystrophic muscle contains proinflammatory, M1‐biased macrophages (red) that can promote muscle damage, as well as antiinflammatory, M2‐biased macrophages (orange), that can affect regeneration and fibrosis. Bar = 50 μm. [Reproduced, with permission, from reference (343).] (C) Characteristics of M1‐biased and M2‐biased macrophages in dystrophic muscle.
Figure 6. Figure 6. Competition for arginine in mdx muscles can affect the pathology of muscular dystrophy. Because arginine is a conditionally essential amino acid in injured and diseased tissues, enzymes that metabolize arginine can compete for substrate. Normally in injured muscle, iNOS and arginase in macrophages and nNOS in muscle fibers compete for arginine. Muscle nNOS transcription is greatly reduced as a consequence of dystrophin deficiency, which increases arginine availability for iNOS and arginase. This amplifies muscle pathology because of muscle fiber lysis by iNOS‐mediated mechanisms and muscle fibrosis mediated by arginase‐mediated mechanisms. [Adapted, with permission, from reference (352).]
Figure 7. Figure 7. Chronic muscle damage in dystrophinopathies causes dysregulation of the immune response that is adapted to acute injuries. Chronic long‐term damage and inflammation of dystrophic muscle can amplify muscle fiber damage and fibrosis. Macrophages increase cytolysis of dystrophic muscle fibers by the release of high levels of NO. Macrophage‐derived NO can also amplify neutrophil‐mediated damage to muscle fibers, which can lyse muscle by producing free radicals including superoxide, hydrogen peroxide (H2O2), hypochlorous acid (HOCl), nitric oxide (NO), and peroxynitrite. CD8+ cytotoxic T‐lymphocytes also contribute to muscle pathology by inducing myonuclear apoptosis through perforin‐mediated mechanisms. The cytotoxicity of CD8+ cells can be diminished by eosinophil MBP, although eosinophils also amplify muscle fiber damage through MBP‐independent mechanisms. M1‐biased macrophages in dystrophic muscle are driven to an M2‐biased phenotype by IL10 and fibrinogen which can subsequently increase muscle fibrosis by providing substrate to fibroblasts for connective tissue production through arginase‐dependent events and by the release of TGFβ, which increases connective tissue production by fibroblasts.
Figure 8. Figure 8. Eosinophils in dystrophinopathy. (A) Inflammatory lesion in 4‐week‐old mdx mouse muscle in section stained with anti‐MBP to indicate locations of MBP‐expressing eosinophils (red). Note the elevated numbers of eosinophils in areas of increased connective tissue accumulation between muscle fibers. Also, note that the cytoplasmic organization of muscle fibers is disrupted in areas enriched with eosinophils, indicating fiber damage, but smooth in areas lacking eosinophils, indicating healthy fibers. Bar = 50 μm. (B) Electron micrograph of a portion of a muscle fiber in 12‐month‐old mdx mouse (left) in close apposition to an eosinophil with a multilobed nucleus (white asterisks). At the pole of the eosinophil closest to the muscle fiber, vesicles containing rods of MBP and other cationic proteins (white arrows) are accumulated. Bar = 1.5 μm. [Reproduced, with permission, from reference (357).] (C) Eosinophil function in dystrophic muscle.
Figure 9. Figure 9. Cytotoxic T‐lymphocytes in dystrophinopathy. (A) Section of 4‐week‐old mdx muscle labeled with antibodies to CD8. Elevated numbers of CD8+ CTLs are present in inflammatory foci in dystrophic muscle. Bar = 50 μm. [Reproduced, with permission, from reference (298).] (B) Schematic of process leading to antigen presentation by muscle fibers leading to activation of CTLs. (C) CTL function in dystrophic muscle.
Figure 10. Figure 10. Activation of the expression of proinflammatory molecules and molecules necessary for antigen presentation by MHC class 1. Both muscle fibers and macrophages can be activated to increase expression of proinflammatory molecules and MHC class 1 via NFκB activation. IκB kinase (IKK) can be activated by TNF binding to its receptor and acting through the adapter protein TRAF2 or by HMGB1 binding its receptor and acting through adapters TRAF6 or MYD88. Activated IKK then phosphorylates IκB, leading to its dissociation from NFκB, causing NFκB activation. Activated NFκB then translocate to the nucleus to bind to the NFκB response element (NRE) of target genes, to activate their expression.
Figure 11. Figure 11. Activation of myogenic precursor cells is essential for repeated muscle regeneration. Satellite cells are myogenic cells that reside in a quiescent state on the surface of muscle fibers and can be identified by their location and by their pattern of expression of myogenic transcription factors (Pax7+/MyoD/MyoG). Following their activation by injury or disease, they enter the cell cycle to give rise to two daughter cells with the same developmental destiny or cells with nonidentical developmental paths. Proliferative daughter cells begin to express MyoD and continue to differentiate while other daughter cells return to the quiescent state and remain Pax7+/ MyoD. As the Pax7+/MyoD+ cells continue to differentiate, they permanently downregulate Pax7 and begin to express the transcription factor myogenin (MyoG) which is required for further differentiation. The Pax7/MyoD+/MyoG+ cells can fuse with other myogenic cells to form multinucleated myotubes. Myotubes then grow and begin to express genes required for terminal differentiation and become nascent muscle fibers. Eventually, nuclei derived from the originally activated satellite cell population become myonuclei that reside within the muscle fiber, in which myogenic regulatory genes are permanently silenced (Pax7/MyoD/MyoG). [Adapted, with permission, from reference (316).]
Figure 12. Figure 12. Inflammation and regeneration are coupled in dystrophic muscle. Damage of dystrophic muscle fibers leads to the release of DAMPs that can recruit and activate leukocytes from the vasculature as well as leukocytes that reside in the muscle. Those leukocytes can further promote recruitment by the release of chemoattractants, such as CCLs. The activated leukocytes then disperse throughout the muscle where they can influence regeneration, in addition to promoting damage. Macrophages that dominate the initial inflammatory infiltrate can be biased toward an M1, cytolytic phenotype by Th1 cytokines such as TNF and IFNγ that can be released by neutrophils and CD8+ cells. In parallel with the recruitment and activation of immune cells, muscle fiber damage also activates satellite cells to begin their program of expansion and differentiation that contributes to muscle regeneration. M1‐biased macrophages release IGF‐1 that stimulates myogenic cell proliferation. M1‐biased macrophages can also switch to an M2‐biased phenotype following stimulation by fibrinogen or IGF‐1 or by IL10 produced by Tregs. M2‐biased macrophages can also contribute to the expansion of myogenic cell populations through the release of Klotho. M2‐biased macrophages can also release TGFβ, that increases the differentiation of FAPs to become fibroblasts. Fibroblast production of IGF‐1 promotes the growth and differentiation of myotubes to become muscle fibers.
Figure 13. Figure 13. Regulatory T‐cells in muscular dystrophy. [(A)‐(C)] Section of DMD muscle biopsy labeled with antibodies to CD3 (A; green) and FoxP3 (B; red) showing that a portion of CD3+ T‐cells in DMD muscle are FoxP3‐expressing Tregs (C; double‐labeled, arrows). [Reproduced, with permission, from reference (344).] (D) Treg function in dystrophic muscle. Bar = 40 μm.
Figure 14. Figure 14. Muscle and cardiac fibrosis are major features of the pathophysiology of dystrophinopathies. (A) Posterior view of thorax of DMD patient showing extreme kyphoscoliosis that is attributable, in part, to fibrosis of paravertebral muscles. (B) Radiograph of 18‐month‐old mdx mouse showing pathological curvature of the spine caused by increased fibrosis of paravertebral muscles. [Reproduced, with permission, from reference (352).] [(C) and (D)] Cross section of quadriceps muscles from 24‐month‐old wild‐type (C) or mdx mouse (D) stained with hematoxylin. Fibers in wild‐type muscle have similar size and are separated by little connective tissue although mdx muscles show large variability in fiber size and extensive accumulation of connective tissue between fibers. Bars = 100 μm. [Reproduced, with permission, from reference (318).] [(E) and (F)] Section through myocardium of 1‐year‐old wild‐type (E) and mdx (F) mice stained for collagen type 1 shows fibrous lesions that occur in mdx myocardia are absent in wild‐type mice. Bars = 50 μm. [Reproduced, with permission, from reference (353).]
Figure 15. Figure 15. Muscle wasting in LGMD1B can be accompanied by extensive immune cell involvement that indicates an acquired immune response. [(A)‐(C)] T1‐weighted magnetic resonance imaging of transverse sections of thighs of three LGMD1B patients. Muscles appearing dark in the image are increasingly replaced by fat and connective tissue, appearing white. Muscles affected include rectus femoris (RF), sartorius (SA), vastus medius (VM), vastus intermedius (VI), vastus lateralis (VL), and gracilis (G). Hematoxylin and eosin‐stained section of biopsied muscle tissue from LGMD1B patient (D) shows greatly elevated numbers of inflammatory cells in lesions within the muscle. (E) Immunohistochemistry of sections of biopsied muscles labeled with antibodies to CD4 (green) and dystrophin (red) shows that many of the leukocytes in inflammatory foci are T‐cells. (F) Immunohistochemistry of sections of biopsied muscles labeled with antibodies to CD20 (green) and dystrophin (red) also shows elevated numbers of B‐cells in the lesions. Bars = 70 μm. [Reproduced, with permission, from reference (175).]
Figure 16. Figure 16. Muscle wasting and inflammation in LGMD2A. (A) LGMD2A is frequently associated with “winged scapulae” attributable to weakening and loss of muscles of the shoulder girdle. [Reproduced, with permission, from reference (7).] (B) Muscle histology shows that muscle fibers in LGMD2A can be replaced by vast numbers of inflammatory cells, that include elevated numbers of eosinophils (arrowheads) in some cases. (C) Muscle histopathology also includes increased deposition of fatty tissue (arrows) and thickening of perimysial and endomysial connective tissue. Bars = 80 μm. [Reproduced, with permission, from reference (276).]
Figure 17. Figure 17. β‐sarcoglycan null mice, a model for LGMD2E, experience extensive muscle fibrosis and inflammation. [(A) and (B)] One‐year old, βSCG mice show large accumulations of connective tissue (blue) between muscle fibers in Masson's trichrome‐stained sections of quadriceps muscles (A) or diaphragm (B). Elevated numbers of macrophages (C), which appear red in acid phosphatase‐stained sections, occur in elevated numbers in the interstitial space between muscle fibers. Bar = 100 μm. [Reproduced, with permission, from reference (108).]
Figure 18. Figure 18. Characteristics of LGMD2B. (A) LGMD2B patient showing wasting of proximal and distal, lower limb musculature. The patient also shows characteristic walking patterns that include reduction of normal heel strike during foot placement. [(Reproduced, with permission, from reference (200).] [(B) and (C)] T1‐weighted MRI of thigh (B) and calf (C) of LGMD2B patient showing extensive replacement of musculature (dark grey) with fat and connective tissue (white). [(Reproduced, with permission, from reference (8).] [(D) and (E)] Anti‐dysferlin staining of healthy, human muscle biopsy (D) and LGMD2B muscle biopsy (E). The normal distribution of dysferlin at the muscle fiber surface is absent in LGMD2B muscles. Bar = 150 μm. [Reproduced, with permission, from reference (172).]
Figure 19. Figure 19. Immune cell infiltrates in LGMD2B muscle. (A) Hematoxylin and eosin‐stained muscle section of LGMD2B biopsy shows large interstitial spaces between muscle fibers that are occupied by inflammatory cells (dark blue) and fibrotic connective tissue. Bar = 100 μm. [Reproduced, with permission, from reference (62).] (B) Anti‐CD3 staining (dark brown) shows that many of the immune cells in LGMD2B muscles are T‐cells. Bar = 100 μm. [Reproduced, with permission, from reference (62).] (C) Indirect immunofluorescence antibody labeling of dystrophin (green) and CD8+ T‐cells in LGMD2B muscle biopsy section show CTLs near or on surface of muscle fibers. Bar = 50 μm. [Reproduced, with permission, from reference (62).] (D) Anti‐CD68 staining (brown) shows that many of the immune cells in inflammatory lesions in LGMD2B muscles are macrophages. Bar = 100 μm. [(Reproduced, with permission, from reference (372).] (E) Anti‐MHC class 1 staining of LGMD2B muscle biopsies shows localization at the surfaces of muscle fibers suggesting antigen presentation by muscle fibers. Bar = 100 μm. [Reproduced, with permission, from reference (62).] (F) Section of LGMD2B muscle biopsy labeled with antibodies to C5b (dark brown) indicates complement activation occurs in dysferlin‐deficient muscles. Bar = 100 μm. [Reproduced, with permission, from reference (372).]
Figure 20. Figure 20. Characteristics of MDC1A. (A) MDC1A patient with characteristic weakness of facial muscles and elbow flexions resulting from increased fibrosis. (B) Increased lumbar lordosis and knee flexions are also attributable to fibrotic contractures. (C) Weakness of neck musculature in MDC1A patient is exhibited when the patient is pulled up from a supine position. [Reproduced, with permission, from reference (105).] [(D) and (E)] T‐weighted MRI of thigh (D) and leg (E) of 1‐year‐old MDC1A patient showing extensive muscle atrophy, fatty infiltration, and edema of lower limb musculature. PT, posterior tibial; SO, soleus; GA, gastrocnemius; AT, anterior tibial; VI, vastus internus; VM, vastus medialis; SA, sartorius; GR, gracilis; AM, adductor magnus; ST, semitendinosus; SM, semimembranosus; BF, biceps femoris; VL, vastus lateralis; RF, rectus femoris. [Reproduced, with permission, from reference (195).] (F) Anti‐merosin immunolabeling of a section of biopsied muscle from healthy subject shows each muscle fiber encircled by a continuous layer of merosin in the basal lamina. Bar = 70 mm. [Reproduced, with permission, from reference (176).] (G) Anti‐merosin labeling of a section of biopsied muscle from an MDC1A shows complete absence of merosin in the basal lamina. Bar = 30 mm. [Reproduced, with permission, from reference (176).] (H) Hematoxylin and eosin‐stained muscle section of MDC1A biopsy shows large interstitial spaces between muscle fibers (dark red) that are occupied by fat (pale spheres) inflammatory cells (dark blue) and fibrotic connective tissue. Bar = 100 mm. [Reproduced, with permission, from reference (176).]
Figure 21. Figure 21. Effects of FSHD on muscle structure and inflammation. (A) FSHD patient showing asymmetric muscle wasting on right side of body, including muscles of shoulder girdle, humeral, and lower limb muscle. (B) T2‐weighted MRI images of right and left, upper, and lower limbs of same patient showing extensive muscle loss and replacement by fat, that is particularly prominent in arm level and crus level images. [Reproduced, with permission, from reference (304).] (C) T1‐weighted (left) and T2‐weighted (right) MRI of FSHD patient. Arrow indicates site of muscle showing hyperintensity in T2 image, that was sampled by biopsy, and found to contain highly elevated numbers of leukocytes. (D) Section of muscle biopsy that is shown in (C), that was immunolabeled with antibodies to CD8 shows large numbers of CD8+ CTLs (red) surrounding muscle fibers (green). Bar = 50 μm. [Reproduced, with permission, from reference (101).] (E) Muscle biopsy from FSHD patient stained with hematoxylin and eosin shows extensive inflammation and fibrosis in the interstitium between muscle fibers. Bar = 70 μm. [Reproduced, with permission, from reference (58).] (F) Quantitative histological analysis of inflammation of muscle biopsies from FSHD patients shows a positive correlation between the number of inflammatory cells (“cell count”) and the number of necrotic fibers (“nec. fibers”). [Reproduced, with permission, from reference (11).]


Figure 1. Characteristics of Duchenne muscular dystrophy (DMD). (A) Typically, the clinical onset of DMD occurs at about 3 years of age as boys begin to show defects in muscle function. In this image of 5 boys with DMD at progressively older ages, some of the gross features of the disease are apparent. The boys show an increasingly progressive lumbar curvature of the spine that results in postural compensation for increased weakening of paravertebral muscles. There is also a progressive increase of weight bearing on the toes and reduction of weight bearing on the heels, as fibrosis of calf muscles cause contractures that limit plantar flexion. Although gross appearance shows an apparent sparing of calf muscles even in older boys, there is actually an increased replacement of muscle tissue with fibrous and fatty tissue, leading to a pseudohypertrophy of the calves. (B) Magnetic resonance spectroscopy images of the shoulders, upper arms, and forearms of healthy subjects (12, 13, and 13 years of age) and DMD patients (12, 13, and 13 years of age) showing tremendous reductions of muscle mass (white) by fatty tissue (dark). BB (biceps brachii), TB (triceps brachii), DEL (deltoid), subscapularis (SUB), infraspinatus (INF), posterior compartment of the forearm (PF), and anterior compartment of the forearm (AF). [Reproduced, with permission, from reference (361).] (C) Section of muscle biopsy of DMD patient stained with hematoxylin and eosin stain. Large accumulations of connective tissue separate individual muscle fibers and mononucleated leukocytes are present in elevated numbers in the interstitial tissue (arrow) and near blood vessels (brackets). Fiber cross‐sectional area is also highly variable, another characteristic of DMD pathology. Bar = 60 μm. [Reproduced, with permission, from reference (252).]


Figure 2. Schematic representation of dystrophin glycoprotein complex (DGC). Dystrophin provides an important structural link between the thin filaments within skeletal muscle fibers and β‐dystroglycan, which is a transmembrane protein. β‐dystroglycan then binds extracellularly to α‐dystroglycan which is a ligand for extracellular structural proteins, especially laminin‐2 that is present in basal lamina. Several proteins in the basal lamina, including laminin and fibronectin, then provide mechanical links to major connective tissue macromolecules such as collagen type 1. Genetic deletion of dystrophin disrupts this mechanical linkage between the cytoskeleton and extracellular structural proteins, but also leads to reductions in the quantity of other structural proteins in the DGC or associated with the DGC. The secondary loss of those other proteins, including the sarcoglycans, syntrophin and neuronal nitric oxide synthase (nNOS) can contribute to the pathology of dystrophin deficiency through disruption of signaling pathways that are necessary for normal muscle function. [Reproduced, with permission, from reference (319).]


Figure 3. Dystrophin deficiency causes membrane lesions and unregulated influx and efflux of large molecules. (A) Electron microscopy on DMD skeletal muscle shows morphologically detectible lesions in the cell membrane (sarcolemma) of dystrophic fibers. Asterisks (*) indicate basal lamina and other connective tissue associated with the extracellular surface of the muscle fiber. Blue arrows indicate sites of intact sarcolemma associated with dense, subsarcolemmal material. Red arrows indicate sites of lesions in the sarcolemma that would permit unrestricted transit of molecules across the membrane. Scale bar = 150 nm. [Reproduced, with permission, from reference (230).] [(B) and (C)] Intact hind limbs of wild‐type (B) or mdx mice (C) that had been injected with an extracellular marker dye, Evans blue, prior to euthanasia and tissue collection. Skin and fat has been removed from both limbs. The wild‐type muscles show little blue dye, indicating that little dye was able to cross the cell membrane to enter the cytosol of the muscle fibers. The mdx muscle shows some thigh muscles contain high concentrations of blue muscle fibers (red arrow) indicating unregulated entry of dye into the fibers through membrane lesions. In contrast, leg muscle in the same mouse shows little blue dye (blue arrow), illustrating the difference in magnitude of pathology and progression of the disease in different muscles. [Reproduced, with permission, from reference (303).] [(D) and (E)] Cross sections of soleus muscle from wild‐type (D) and mdx (E) mice. The muscles were incubated in a fluorescent, extracellular marker dye after dissection and before freezing the muscle for histology. In the wild‐type muscle, the fluorescent dye remains in the extracellular space because the muscle cell membranes are intact (D). In the mdx muscle, approximately 8% of the muscle fibers showed elevated intracellular fluorescence, indicating the presence of membrane lesions. Bars = 70 µm. [Reproduced, with permission, from reference (351).]


Figure 4. Macrophages are a primary source of muscle fiber damage in the mdx model of DMD. (A) At the early, acute peak of mdx muscle pathology, macrophages dominate large inflammatory lesions in the diseased muscle. Cross section of 4‐week‐old mdx mouse muscle with F4/80 + macrophages immunolabeled to appear red in the sectioned muscle. Bar = 180 µm. (B) Image shows complete, mid‐belly cross section of an entire soleus muscle from 4‐week‐old mdx mice. The mouse from which the muscle on the left was collected received intraperitoneal injections of sterile buffer on 5 days each week from 8 days of age until the mouse was euthanized at 4 weeks of age. The mouse from which the muscle on the right was collected received intraperitoneal injections of antibodies to the F4/80 antigen on the same injection schedule, which produced a reduction in macrophage numbers by over 90%. Both sections were stained with antibodies to neural cell adhesion molecule (NCAM) which is expressed at high levels by muscle fibers that are undergoing repair following injury. Each small, red tile‐like structure in the cross section is a recently injured fiber. Note that the macrophage‐depleted muscle is nearly devoid of NCAM‐expressing muscle fibers. Bar = 1.2 mm. [(C) and (D)] Soleus muscles from nondepleted, 4‐week‐old mdx mice (C) or macrophage‐depleted mdx mice (D) were incubated in Procion red, a fluorescent, extracellular marker dye before muscle sectioning and microscopy. Intracellular Procion red indicates fibers with membrane lesions. The number of Procion‐red‐containing fibers in the 4‐week‐old mdx solei was reduced by more than 75%. Bar = 120 µm. [Reproduced, with permission, from reference (351).]


Figure 5. Inflammation in the mdx mouse model of DMD. (A) Section of a 4‐week‐old mdx mouse muscle immunolabeled with anti‐F4/80, a pan‐macrophage marker (red). Note that some muscle fibers have been invaded and obliterated by large numbers of macrophages (white asterisk) while other, small regenerating fibers are surrounded but not invaded by macrophages (black asterisk). Bar = 50 μm. (B) Section of 4‐week‐old muscle immunolabeled with antibodies to F4/80 (red) and CD206 (green), a marker for M2‐biased macrophages. Blue structures are nuclei binding DAPI reagent. This inflammatory lesion in dystrophic muscle contains proinflammatory, M1‐biased macrophages (red) that can promote muscle damage, as well as antiinflammatory, M2‐biased macrophages (orange), that can affect regeneration and fibrosis. Bar = 50 μm. [Reproduced, with permission, from reference (343).] (C) Characteristics of M1‐biased and M2‐biased macrophages in dystrophic muscle.


Figure 6. Competition for arginine in mdx muscles can affect the pathology of muscular dystrophy. Because arginine is a conditionally essential amino acid in injured and diseased tissues, enzymes that metabolize arginine can compete for substrate. Normally in injured muscle, iNOS and arginase in macrophages and nNOS in muscle fibers compete for arginine. Muscle nNOS transcription is greatly reduced as a consequence of dystrophin deficiency, which increases arginine availability for iNOS and arginase. This amplifies muscle pathology because of muscle fiber lysis by iNOS‐mediated mechanisms and muscle fibrosis mediated by arginase‐mediated mechanisms. [Adapted, with permission, from reference (352).]


Figure 7. Chronic muscle damage in dystrophinopathies causes dysregulation of the immune response that is adapted to acute injuries. Chronic long‐term damage and inflammation of dystrophic muscle can amplify muscle fiber damage and fibrosis. Macrophages increase cytolysis of dystrophic muscle fibers by the release of high levels of NO. Macrophage‐derived NO can also amplify neutrophil‐mediated damage to muscle fibers, which can lyse muscle by producing free radicals including superoxide, hydrogen peroxide (H2O2), hypochlorous acid (HOCl), nitric oxide (NO), and peroxynitrite. CD8+ cytotoxic T‐lymphocytes also contribute to muscle pathology by inducing myonuclear apoptosis through perforin‐mediated mechanisms. The cytotoxicity of CD8+ cells can be diminished by eosinophil MBP, although eosinophils also amplify muscle fiber damage through MBP‐independent mechanisms. M1‐biased macrophages in dystrophic muscle are driven to an M2‐biased phenotype by IL10 and fibrinogen which can subsequently increase muscle fibrosis by providing substrate to fibroblasts for connective tissue production through arginase‐dependent events and by the release of TGFβ, which increases connective tissue production by fibroblasts.


Figure 8. Eosinophils in dystrophinopathy. (A) Inflammatory lesion in 4‐week‐old mdx mouse muscle in section stained with anti‐MBP to indicate locations of MBP‐expressing eosinophils (red). Note the elevated numbers of eosinophils in areas of increased connective tissue accumulation between muscle fibers. Also, note that the cytoplasmic organization of muscle fibers is disrupted in areas enriched with eosinophils, indicating fiber damage, but smooth in areas lacking eosinophils, indicating healthy fibers. Bar = 50 μm. (B) Electron micrograph of a portion of a muscle fiber in 12‐month‐old mdx mouse (left) in close apposition to an eosinophil with a multilobed nucleus (white asterisks). At the pole of the eosinophil closest to the muscle fiber, vesicles containing rods of MBP and other cationic proteins (white arrows) are accumulated. Bar = 1.5 μm. [Reproduced, with permission, from reference (357).] (C) Eosinophil function in dystrophic muscle.


Figure 9. Cytotoxic T‐lymphocytes in dystrophinopathy. (A) Section of 4‐week‐old mdx muscle labeled with antibodies to CD8. Elevated numbers of CD8+ CTLs are present in inflammatory foci in dystrophic muscle. Bar = 50 μm. [Reproduced, with permission, from reference (298).] (B) Schematic of process leading to antigen presentation by muscle fibers leading to activation of CTLs. (C) CTL function in dystrophic muscle.


Figure 10. Activation of the expression of proinflammatory molecules and molecules necessary for antigen presentation by MHC class 1. Both muscle fibers and macrophages can be activated to increase expression of proinflammatory molecules and MHC class 1 via NFκB activation. IκB kinase (IKK) can be activated by TNF binding to its receptor and acting through the adapter protein TRAF2 or by HMGB1 binding its receptor and acting through adapters TRAF6 or MYD88. Activated IKK then phosphorylates IκB, leading to its dissociation from NFκB, causing NFκB activation. Activated NFκB then translocate to the nucleus to bind to the NFκB response element (NRE) of target genes, to activate their expression.


Figure 11. Activation of myogenic precursor cells is essential for repeated muscle regeneration. Satellite cells are myogenic cells that reside in a quiescent state on the surface of muscle fibers and can be identified by their location and by their pattern of expression of myogenic transcription factors (Pax7+/MyoD/MyoG). Following their activation by injury or disease, they enter the cell cycle to give rise to two daughter cells with the same developmental destiny or cells with nonidentical developmental paths. Proliferative daughter cells begin to express MyoD and continue to differentiate while other daughter cells return to the quiescent state and remain Pax7+/ MyoD. As the Pax7+/MyoD+ cells continue to differentiate, they permanently downregulate Pax7 and begin to express the transcription factor myogenin (MyoG) which is required for further differentiation. The Pax7/MyoD+/MyoG+ cells can fuse with other myogenic cells to form multinucleated myotubes. Myotubes then grow and begin to express genes required for terminal differentiation and become nascent muscle fibers. Eventually, nuclei derived from the originally activated satellite cell population become myonuclei that reside within the muscle fiber, in which myogenic regulatory genes are permanently silenced (Pax7/MyoD/MyoG). [Adapted, with permission, from reference (316).]


Figure 12. Inflammation and regeneration are coupled in dystrophic muscle. Damage of dystrophic muscle fibers leads to the release of DAMPs that can recruit and activate leukocytes from the vasculature as well as leukocytes that reside in the muscle. Those leukocytes can further promote recruitment by the release of chemoattractants, such as CCLs. The activated leukocytes then disperse throughout the muscle where they can influence regeneration, in addition to promoting damage. Macrophages that dominate the initial inflammatory infiltrate can be biased toward an M1, cytolytic phenotype by Th1 cytokines such as TNF and IFNγ that can be released by neutrophils and CD8+ cells. In parallel with the recruitment and activation of immune cells, muscle fiber damage also activates satellite cells to begin their program of expansion and differentiation that contributes to muscle regeneration. M1‐biased macrophages release IGF‐1 that stimulates myogenic cell proliferation. M1‐biased macrophages can also switch to an M2‐biased phenotype following stimulation by fibrinogen or IGF‐1 or by IL10 produced by Tregs. M2‐biased macrophages can also contribute to the expansion of myogenic cell populations through the release of Klotho. M2‐biased macrophages can also release TGFβ, that increases the differentiation of FAPs to become fibroblasts. Fibroblast production of IGF‐1 promotes the growth and differentiation of myotubes to become muscle fibers.


Figure 13. Regulatory T‐cells in muscular dystrophy. [(A)‐(C)] Section of DMD muscle biopsy labeled with antibodies to CD3 (A; green) and FoxP3 (B; red) showing that a portion of CD3+ T‐cells in DMD muscle are FoxP3‐expressing Tregs (C; double‐labeled, arrows). [Reproduced, with permission, from reference (344).] (D) Treg function in dystrophic muscle. Bar = 40 μm.


Figure 14. Muscle and cardiac fibrosis are major features of the pathophysiology of dystrophinopathies. (A) Posterior view of thorax of DMD patient showing extreme kyphoscoliosis that is attributable, in part, to fibrosis of paravertebral muscles. (B) Radiograph of 18‐month‐old mdx mouse showing pathological curvature of the spine caused by increased fibrosis of paravertebral muscles. [Reproduced, with permission, from reference (352).] [(C) and (D)] Cross section of quadriceps muscles from 24‐month‐old wild‐type (C) or mdx mouse (D) stained with hematoxylin. Fibers in wild‐type muscle have similar size and are separated by little connective tissue although mdx muscles show large variability in fiber size and extensive accumulation of connective tissue between fibers. Bars = 100 μm. [Reproduced, with permission, from reference (318).] [(E) and (F)] Section through myocardium of 1‐year‐old wild‐type (E) and mdx (F) mice stained for collagen type 1 shows fibrous lesions that occur in mdx myocardia are absent in wild‐type mice. Bars = 50 μm. [Reproduced, with permission, from reference (353).]


Figure 15. Muscle wasting in LGMD1B can be accompanied by extensive immune cell involvement that indicates an acquired immune response. [(A)‐(C)] T1‐weighted magnetic resonance imaging of transverse sections of thighs of three LGMD1B patients. Muscles appearing dark in the image are increasingly replaced by fat and connective tissue, appearing white. Muscles affected include rectus femoris (RF), sartorius (SA), vastus medius (VM), vastus intermedius (VI), vastus lateralis (VL), and gracilis (G). Hematoxylin and eosin‐stained section of biopsied muscle tissue from LGMD1B patient (D) shows greatly elevated numbers of inflammatory cells in lesions within the muscle. (E) Immunohistochemistry of sections of biopsied muscles labeled with antibodies to CD4 (green) and dystrophin (red) shows that many of the leukocytes in inflammatory foci are T‐cells. (F) Immunohistochemistry of sections of biopsied muscles labeled with antibodies to CD20 (green) and dystrophin (red) also shows elevated numbers of B‐cells in the lesions. Bars = 70 μm. [Reproduced, with permission, from reference (175).]


Figure 16. Muscle wasting and inflammation in LGMD2A. (A) LGMD2A is frequently associated with “winged scapulae” attributable to weakening and loss of muscles of the shoulder girdle. [Reproduced, with permission, from reference (7).] (B) Muscle histology shows that muscle fibers in LGMD2A can be replaced by vast numbers of inflammatory cells, that include elevated numbers of eosinophils (arrowheads) in some cases. (C) Muscle histopathology also includes increased deposition of fatty tissue (arrows) and thickening of perimysial and endomysial connective tissue. Bars = 80 μm. [Reproduced, with permission, from reference (276).]


Figure 17. β‐sarcoglycan null mice, a model for LGMD2E, experience extensive muscle fibrosis and inflammation. [(A) and (B)] One‐year old, βSCG mice show large accumulations of connective tissue (blue) between muscle fibers in Masson's trichrome‐stained sections of quadriceps muscles (A) or diaphragm (B). Elevated numbers of macrophages (C), which appear red in acid phosphatase‐stained sections, occur in elevated numbers in the interstitial space between muscle fibers. Bar = 100 μm. [Reproduced, with permission, from reference (108).]


Figure 18. Characteristics of LGMD2B. (A) LGMD2B patient showing wasting of proximal and distal, lower limb musculature. The patient also shows characteristic walking patterns that include reduction of normal heel strike during foot placement. [(Reproduced, with permission, from reference (200).] [(B) and (C)] T1‐weighted MRI of thigh (B) and calf (C) of LGMD2B patient showing extensive replacement of musculature (dark grey) with fat and connective tissue (white). [(Reproduced, with permission, from reference (8).] [(D) and (E)] Anti‐dysferlin staining of healthy, human muscle biopsy (D) and LGMD2B muscle biopsy (E). The normal distribution of dysferlin at the muscle fiber surface is absent in LGMD2B muscles. Bar = 150 μm. [Reproduced, with permission, from reference (172).]


Figure 19. Immune cell infiltrates in LGMD2B muscle. (A) Hematoxylin and eosin‐stained muscle section of LGMD2B biopsy shows large interstitial spaces between muscle fibers that are occupied by inflammatory cells (dark blue) and fibrotic connective tissue. Bar = 100 μm. [Reproduced, with permission, from reference (62).] (B) Anti‐CD3 staining (dark brown) shows that many of the immune cells in LGMD2B muscles are T‐cells. Bar = 100 μm. [Reproduced, with permission, from reference (62).] (C) Indirect immunofluorescence antibody labeling of dystrophin (green) and CD8+ T‐cells in LGMD2B muscle biopsy section show CTLs near or on surface of muscle fibers. Bar = 50 μm. [Reproduced, with permission, from reference (62).] (D) Anti‐CD68 staining (brown) shows that many of the immune cells in inflammatory lesions in LGMD2B muscles are macrophages. Bar = 100 μm. [(Reproduced, with permission, from reference (372).] (E) Anti‐MHC class 1 staining of LGMD2B muscle biopsies shows localization at the surfaces of muscle fibers suggesting antigen presentation by muscle fibers. Bar = 100 μm. [Reproduced, with permission, from reference (62).] (F) Section of LGMD2B muscle biopsy labeled with antibodies to C5b (dark brown) indicates complement activation occurs in dysferlin‐deficient muscles. Bar = 100 μm. [Reproduced, with permission, from reference (372).]


Figure 20. Characteristics of MDC1A. (A) MDC1A patient with characteristic weakness of facial muscles and elbow flexions resulting from increased fibrosis. (B) Increased lumbar lordosis and knee flexions are also attributable to fibrotic contractures. (C) Weakness of neck musculature in MDC1A patient is exhibited when the patient is pulled up from a supine position. [Reproduced, with permission, from reference (105).] [(D) and (E)] T‐weighted MRI of thigh (D) and leg (E) of 1‐year‐old MDC1A patient showing extensive muscle atrophy, fatty infiltration, and edema of lower limb musculature. PT, posterior tibial; SO, soleus; GA, gastrocnemius; AT, anterior tibial; VI, vastus internus; VM, vastus medialis; SA, sartorius; GR, gracilis; AM, adductor magnus; ST, semitendinosus; SM, semimembranosus; BF, biceps femoris; VL, vastus lateralis; RF, rectus femoris. [Reproduced, with permission, from reference (195).] (F) Anti‐merosin immunolabeling of a section of biopsied muscle from healthy subject shows each muscle fiber encircled by a continuous layer of merosin in the basal lamina. Bar = 70 mm. [Reproduced, with permission, from reference (176).] (G) Anti‐merosin labeling of a section of biopsied muscle from an MDC1A shows complete absence of merosin in the basal lamina. Bar = 30 mm. [Reproduced, with permission, from reference (176).] (H) Hematoxylin and eosin‐stained muscle section of MDC1A biopsy shows large interstitial spaces between muscle fibers (dark red) that are occupied by fat (pale spheres) inflammatory cells (dark blue) and fibrotic connective tissue. Bar = 100 mm. [Reproduced, with permission, from reference (176).]


Figure 21. Effects of FSHD on muscle structure and inflammation. (A) FSHD patient showing asymmetric muscle wasting on right side of body, including muscles of shoulder girdle, humeral, and lower limb muscle. (B) T2‐weighted MRI images of right and left, upper, and lower limbs of same patient showing extensive muscle loss and replacement by fat, that is particularly prominent in arm level and crus level images. [Reproduced, with permission, from reference (304).] (C) T1‐weighted (left) and T2‐weighted (right) MRI of FSHD patient. Arrow indicates site of muscle showing hyperintensity in T2 image, that was sampled by biopsy, and found to contain highly elevated numbers of leukocytes. (D) Section of muscle biopsy that is shown in (C), that was immunolabeled with antibodies to CD8 shows large numbers of CD8+ CTLs (red) surrounding muscle fibers (green). Bar = 50 μm. [Reproduced, with permission, from reference (101).] (E) Muscle biopsy from FSHD patient stained with hematoxylin and eosin shows extensive inflammation and fibrosis in the interstitium between muscle fibers. Bar = 70 μm. [Reproduced, with permission, from reference (58).] (F) Quantitative histological analysis of inflammation of muscle biopsies from FSHD patients shows a positive correlation between the number of inflammatory cells (“cell count”) and the number of necrotic fibers (“nec. fibers”). [Reproduced, with permission, from reference (11).]
References
 1.Accorsi A, Kumar A, Rhee Y, Miller A, Girgenrath M. IGF‐1/GH axis enhances losartan treatment in Lama2‐related muscular dystrophy. Hum Molec Genet 25: 4624‐4643, 2016. 10.1093/hmg/ddw291.
 2.Acharyya S, Villalta SA, Bakkar N, Bupha‐Intr T, Janssen PM, Carathers M, Li ZW, Beg AA, Ghosh S, Sahenk Z, Weinstein M, Gardner KL, Rafael‐Fortney JA, Karin M, Tidball JG, Baldwin AS, Guttridge DC. Interplay of IKK/NF‐kappaB signaling in macrophages and myofibers promotes muscle degeneration in Duchenne muscular dystrophy. J Clin Invest 117: 889‐901, 2007.
 3.Aderem A, Underhill DM. Mechanisms of phagocytosis in macrophages. Annu Rev Immunol 17: 593‐623, 1999.
 4.Allen RE, Boxhorn LK. Inhibition of skeletal muscle satellite cell differentiation by transforming growth factor‐beta. J Cell Physiol 133: 567‐572, 1987.
 5.Allen RE, Boxhorn LK. Regulation of skeletal muscle satellite cell proliferation and differentiation by transforming growth factor‐beta, insulin‐like growth factor I, and fibroblast growth factor. J Cell Physiol 138: 311‐315, 1989.
 6.Andreetta F, Bernasconi P, Baggi F, Ferro P, Oliva L, Arnoldi E, Cornelio F, Mantegazza R, Confalonieri P. Immunomodulation of TGF‐beta 1 in mdx mouse inhibits connective tissue proliferation in diaphragm but increases inflammatory response: Implications for antifibrotic therapy. J Neuroimmunol 175: 77‐86, 2006.
 7.Angelini C. Limb‐girdle muscular dystrophy type 2A. In: Genetic Neuromuscular Disorders: A Case‐Based Approach. Switzerland: Springer International Publishing, 2014.
 8.Angelini C, Peterle E, Gaiani A, Bortolussi L, Borsato C. Dysferlinopathy course and sportive activity: Clues for possible treatment. Acta Myol 30: 127‐132, 2011.
 9.Appleyard ST, Dunn JJ, Dubowitz V, Rose ML. Increased expression of the HLA ABC class I antigens by muscle fibres in Duchenne muscular dystrophy, inflammatory myopathy, and other neuromuscular disorders. Lancet 1: 361‐363, 1985.
 10.Arahata K, Engel AG. Monoclonal antibody analysis of mononuclear cells in myopathies. IV: Cell‐mediated cytotoxicity and muscle fiber necrosis. Ann Neurol 23: 168‐173, 1988.
 11.Arahata K, Ishihara T, Fukunaga H, Orimo S, Lee JH, Goto K, Nonaka I. Inflammatory response in facioscapulohumeral muscular dystrophy (FSHD): Immunocytochemical and genetic analyses. Muscle Nerve Suppl 2: S56‐S66, 1995.
 12.Arecco N, Clarke CJ, Jones FK, Simpson DM, Mason D, Beynon RJ, Pisconti A. Elastase levels and activity are increased in dystrophic muscle and impair myoblast cell survival, proliferation and differentiation. Sci Rep 6: 24708, 2016. 10.1038/srep24708.
 13.Atkinson JB, Harlan CW, Harlan GC, Virmani R. The association of mast cells and atherosclerosis: A morphologic study of early atherosclerotic lesions in young people. Hum Pathol 25: 154‐159, 1994.
 14.Baek JH, Many GM, Evesson FJ, Kelley VR. Dysferlinopathy promotes an intramuscle expansion of macrophages with a cyto‐destructive phenotype. Am J Pathol 187: 1245‐1257, 2017. 10.1016/j.ajpath.2017.02.011.
 15.Banik NL, Matzelle D, Tery E, Hogan EL. A new mechanism of methylprednisolone and other corticosteroids action demonstrated in vitro: Inhibition of a proteinase (calpain) prevents myelin and cytoskeletal protein degradation. Brain Res 748: 205‐210, 1997.
 16.Bansal D, Miyake K, Vogel SS, Groh S, Chen CC, Williamson R, McNeil PL, Campbell KP. Defective membrane repair in dysferlin‐deficient dystrophy. Nature 423: 168‐172, 2003.
 17.Barbul A. Proline precursors to sustain mammalian collagen synthesis. J Nutrition 138: 2021S‐2024S, 2008.
 18.Barton ER, Morris L, Kawana M, Bish LT, Toursel T. Systemic administration of L‐arginine benefits mdx skeletal muscle function. Muscle Nerve 32: 751‐760, 2005.
 19.Bashir R, Keers S, Strachan T, Passos‐Bueno R, Zatz M, Weissenbach J, Le Paslier D, Meisler M, Bushby K. Genetic and physical mapping at the limb‐girdle muscular dystrophy locus (LGMD 2B) on chromosome 2p. Genomics 33: 46‐52, 1996.
 20.Bashir R, Strachan T, Keers S, Stephenson A, Mahjneh I, Marconi G, Nashef L, Bushby KM. A gene for autosomal recessive limb‐girdle muscular dystrophy maps to chromosome 2p. Hum Mol Genet 3: 455‐457, 1994.
 21.Bautista DS, Xuan JW, Hota C, Chambers AF, Harris JF. Inhibition of Arg‐Gly‐Asp (RGD) mediated cell adhesion to osteopontin by a monoclonal antibody against osteopontin. J Biol Chem 269: 23280‐23285, 1994.
 22.Beckman E, Henriksson KG. Low‐dose prednisolone treatment in Duchenne and Becker muscular dystrophy. Neuromuscul Disord 5: 233‐241, 1995.
 23.Bejaoui K, Hirabayashi K, Henta ti F, Haines JL, Ben Hamida C, Belal S, Miller RG, McKenna‐Yasek D, Weissenbach J, Rowland LP, Griggs RC, Munsat TL, Ben Hamida M, Arahata K, Brown RH, Jr. Linkage of Miyoshi myopathy (distal autosomal recessive muscular dystrophy) locus to chromosome 2p12‐14. Neurology 45: 768‐772, 1995.
 24.Beltran‐Valero de Bernabe D, van Bokhoven E, van Beusekom E, Van Den Akker W, Kant S, Dobyns WB, Cormand B, Currier S, Hamel B, Talim B, Topaloglu H, Brunner HG. A homozygous nonsense mutation in the Fukutin gene causes a Walker‐Warburg syndrome phenotype. J Med Genet 40: 845‐848, 2003.
 25.Beltran‐Valero de Bernabe D, Voit T, Longman C, Steinbrecher A, Straub V, Yuva Y, Herrmann R, Sperner J, Korenke C, Diesen C, Dobyns WB, Brunner HG, van Bokhoven H, Brockington M, Muntoni F. Mutations in the FKRP gene can cause muscle‐eye‐brain disease and Walker‐Warburg syndrome. J Med Genet 41: e61, 2004. 10.1136/jmg.2003.013870.
 26.Benyon RC, Bissonnette EY, Befus AD. Tumor necrosis factor‐alpha dependent cytotoxicity of human skin mast cells is enhanced by anti‐IgE antibodies. J Immunol 147: 2253‐2258, 1991.
 27.Berke G. The CTL's kiss of death. Cell 81: 9‐12, 1995.
 28.Bernasconi P, Torchiana E, Confalonieri P, Brugnoni R, Barresi R, Mora M, Cornelio F, Morandi L, Mantegazza R. Expression of transforming growth factor‐beta 1 in dystrophic patient muscles correlates with fibrosis. Pathogenetic role of a fibrogenic cytokine. J Clin Invest 96: 1137‐1144, 1995.
 29.Biancheri R, Falace A, Tessa A, Pedemonte M, Scapolan S, Cassandrini D, Aiello C, Rossi A, Broda P, Zara F, Santorelli FM, Minetti C, Bruno C. POMT2 gene mutation in limb‐girdle muscular dystrophy with inflammatory changes. Biochem Biophys Res Commun 363: 1033‐1037, 2007.
 30.Bischoff R. Chemotaxis of skeletal muscle satellite cells. Dev Dyn 208: 505‐515, 1997.
 31.Bittner RE, Anderson LV, Burkhardt E, Bashir R, Vafiadaki E, Ivanova S, Raffelsberger T, Maerk I, Hoger H, Jung M, Karbasiyan M, Storch M, Lassmann H, Moss JA, Davison K, Harrison R, Bushby KM, Reis A. Dysferlin deletion in SJL/J mice (SJL‐Dysf) defines a natural model for limb girdle muscular dystrophy 2B. Nat Genet 23: 141‐142, 1999.
 32.Bogdan C, Nathan C. Modulation of macrophage function by transforming growth factor beta, interleukin‐4, and interleukin‐10. Ann N Y Acad Sci 685: 713‐739, 1993.
 33.Bonifati MD, Ruzza G, Bonometto P, Berardinelli A, Gorni K, Orcesi S, Lanzi G, Angelini C. A multicenter, double‐blind, randomized trial of deflazacort versus prednisone in Duchenne muscular dystrophy. Muscle Nerve 23: 1344‐1347, 2000.
 34.Bonuccelli G, Sotgia F, Schubert W, Park DS, Frank PG, Woodman SE, Insabato L, Cammer M, Minetti C, Lisanti MP. Proteasome inhibitor (MG‐132) treatment of mdx mice rescues the expression and membrane localization of dystrophin and dystrophin‐associated proteins. Am J Pathol 163: 1663‐1675, 2003.
 35.Brooke MH, Engel WK. The histologic diagnosis of neuromuscular diseases: Review of 79 biopsies. Arch Phys Med Rehabil 47: 99‐121, 1966.
 36.Brooke MH, Fenichel GM, Griggs RC, Mendell JR, Moxley RC, Miller JP, Kaiser KK, Florence JM, Pandya S, Signore L. Clinical investigation of Duchenne muscular dystrophy. Interesting results in a trial of prednisone. Arch Neurol 44: 812‐817, 1987.
 37.Brown LF, Berse B, Van de Water L, Papadopoulos‐Sergiou A, Perruzzi CA, Manseau EJ, Dvorak HF, Senger DR. Expression and distribution of osteopontin in human tissues: Widespread association with luminal epithelial surfaces. Mol Biol Cell 3: 1169‐1180, 1992.
 38.Brown SC, Torelli S, Brockington M, Yuva Y, Jimenez C, Feng L, Anderson L, Ugo S, Bushby K, Voit T, Sewry C, Muntoni F. Abnormalities in alpha‐dystroglycan expression in MDC1C and LGMD2I muscular dystrophies. Am J Pathol 164: 727‐737, 2004.
 39.Brunelli S, Sciorati C, D'Antona G, Innocenzi A, Covarello D, Galvez BG, Perrotta C, Monopoli A, Sanvito F, Bottinelli R, Ongini E, Cossu G, Clementi E. Nitric oxide release combined with nonsteroidal antiinflammatory activity prevents muscular dystrophy pathology and enhances stem cell therapy. Proc Natl Acad Sci U S A 104: 264‐269, 2007.
 40.Budzynski AZ. Fibrinogen and fibrin: Biochemistry and pathophysiology. Crit Rev Oncol Hematol 6: 97‐146, 1986.
 41.Buechler C, Ritter M, Orso E, Langmann T, Klucken J, Schmitz G. Regulation of scavenger receptor CD163 expression in human monocytes and macrophages by pro‐ and antiinflammatory stimuli. J Leukoc Biol 67: 97‐103, 2000.
 42.Burzyn, D, Kuswanto W, Kolodin D, Shadrach JL, Cerletti M, Jang Y, Sefik E, Tan TG, Wagers AJ, Benoist C, Mathis D. A special population of regulatory T cells potentiates muscle repair. Cell 155: 1282‐1295, 2013. 10.1016/j.cell.2013.10.054.
 43.Buyse GM, Goemans N, Henricson E, Jara A, van den Hauwe M, Leshner R, Florence JM, Mayhew JE, Escolar DM. CINRG pilot trial of oxatomide in steroid‐naïve Duchenne muscular dystrophy. Eur J Paediatr Neurol 11: 337‐340, 2007.
 44.Cai B, Spencer MJ, Nakamura G, Tseng‐Ong L, Tidball JG. Eosinophilia of dystrophin‐deficient muscle is promoted by perforin‐mediated cytotoxicity by T cell effectors. Am J Pathol 156: 1789‐1796, 2000.
 45.Capote J, Kramerova I, Martinez L, Vetrone S, Barton ER, Sweeny HL, Miceli MC, Spencer MJ. Osteopontin ablation ameliorates muscular dystrophy by shifting macrophages to a pro‐regenerative phenotype. J Cell Biol 213: 275‐288, 2016. 10.1083/jcb.201510086.
 46.Carpenter S, Karpati G. Pathology of Skeletal Muscle, (2nd ed). New York, NY: Oxford University Press, 2001.
 47.Caughey GH. Mast cell tryptases and chymases in inflammation and host defense. Immunol Rev 217: 141‐154, 2007.
 48.Cavaillon JM, Fitting C, Haeffner‐Cavaillon N. Recombinant C5a enhances interleukin 1 and tumor necrosis factor release by lipopolysaccharide‐stimulated monocytes and macrophages. Eur J Immunol 20: 253‐257, 1990.
 49.Cenacchi G, Fanin M, Badiali‐De Giorgi L, Angelini C. Ultrastructural changes in dysferlinopathy support defective membrane repair mechanism. J Clin Pathol 58: 190‐195, 2005.
 50.Chang WJ, Iannaccone ST, Lau KS, Masters BS, McCabe TJ, McMillan K, Padre RC, Spencer MJ, Tidball JG, Stull JT. Neuronal nitric oxide synthase and dystrophin‐deficient muscular dystrophy. Proc Natl Acad Sci U S A 93: 9142‐9147, 1996.
 51.Chaussenot R, Edeline JM, Le Bec B, El Massioui N, Laroche S, Vaillend C. Cognitive dysfunction in the dystrophin‐deficient mouse model of Duchenne muscular dystrophy: A reappraisal from sensory to executive processes. Neurobiol Learn Mem 124: 111‐122, 2015.
 52.Chen CN, Graber TG, Bratten WM, Ferrington DA, Thompson LV. Immunoproteasome in animal models of Duchenne muscular dystrophy. J Muscle Res Cell Motil 35: 191‐201, 2014. 10.1007/s10974‐014‐9385‐x.
 53.Chen GY, Nunez G. Sterile inflammation: Sensing and reacting to damage. Nat Rev Immunol 10: 826‐837, 2010.
 54.Chen S, Kammerl IE, Vosyka O, Baumann T, Yu Y, Wu Y, Irmler M, Overkleeft HS, Beckers J, Eickelberg O, Meiner S, Stoeger T. Immunoproteasome dysfunction augments alternative polarization of alveolar macrophages. Cell Death Differ 23: 1026‐1037, 2016. 10.1038/cdd.2016.3.
 55.Chiu YH, Hornsey MA, Klinge L, Jorgensen LH, Laval SH, Charlton R, Barresi R, Straub V, Lochmüller H, Bushby K. Attenuated muscle regeneration is a key factor in dysferlin‐deficient muscular dystrophy. Hum Mol Genet 18: 1976‐1989, 2009. 10.1093/hmg/ddp121.
 56.Cho SH, Anderson AJ, Oh CK. Importance of mast cells in the pathophysiology of asthma. Clin Rev Allergy Immunol 22: 161‐74, 2002.
 57.Choi JH, Park YE, Kim SI, Kim JI, Lee CH, Park KH, Kim DS. Differential immunohistological features of inflammatory myopathies and dysferlinopathy. J Korean Med Sci 24: 1015‐1023, 2009.
 58.Choi JH, Park YE, Shin JH, Lee CH, Kim DS. Extensive inflammatory reaction in facioscapulohumeral muscular dystrophy. Ann Clin Neurophysiol 19: 141‐144, 2017. 10.14253/acn.2017.19.2.141.
 59.Cohen TV, Cohen JE, Partridge TA. Myogenesis in dysferlin‐deficient myoblasts is inhibited by an intrinsic inflammatory response. Neuromuscul Disord 22: 648‐658, 2012. 10.1016/j.nmd.2012.03.002.
 60.Cohen TV, Many GM, Fleming BD, Gnocchi VF, Ghimbovschi S, Mosser DM, Hoffman EP, Partridge TA. Upregulated IL‐1 beta in dysferlin‐deficient muscle attenuates regeneration by blunting the response to pro‐inflammatory macrophages. Skelet Muscle 5: 24, 2015. 10.1186/s13395‐015‐0048‐4.
 61.Collart MA, Baeuerle P, Vassalli P. Regulation of tumor necrosis factor alpha transcription in macrophages: Involvement of four kappa B‐like motifs and of constitutive and inducible forms of NF‐kappa B. Mol Cell Biol 10: 1498‐1506, 1990.
 62.Confalonieri P, Oliva L, Andreetta F, Lorenzoni R, Dassi P, Mariani E, Morandi L, Mora M, Cornelio F, Mantegazza R. Muscle inflammation and MHC class I up‐regulation in muscular dystrophy with lack of dysferlin: An immunopathological study. J Neuroimmunol 142: 130‐136, 2003.
 63.Connolly AM, Keeling RM, Mehta S, Pestronk A, Sanes JR. Three mouse models of muscular dystrophy: The natural history of strength and fatigue in dystrophin‐, dystrophin/utrophin‐, and laminin a‐2 deficient mice. Neuromusc Disord 11: 703‐712, 2001.
 64.Cordier‐Fruh I, Barman L, Briguet A, Meier T. Glucocorticoid‐mediated regulation of utrophin levels in human muscle fibers. Neuromuscul Disord 12 (Suppl 1): S95‐S104, 2002.
 65.Crawford SE, Stellmach V, Murphy‐Ullrich JE, Ribeiro SM, Lawler J, Hynes RO, Boivin GP, Bouck N. Thrombospondin‐1 is a major activator of TGF‐beta1 in vivo. Cell 93: 1159‐1170, 1998.
 66.Curran JN, Winter DC, Bouchier‐Hayes D. Biological fate and clinical implications of arginine metabolism in tissue healing. Wound Repair Regen 14: 376‐386, 2006.
 67.Currier SC, Lee CK, Chang BS, Bodell AL, Pai GS, Job L, Lagae LG, Al‐Gazali LI, Eyaid WM, Enns G, Dobyns WB, Walsh CA. Mutations in POMT1 are found in a minority of patients with Walker‐Warburg syndrome. Am J Med Genet A 133: 53‐57, 2005.
 68.Czaja MJ, Weiner FR, Flanders KC, Giambrone MA, Wind R, Biempica L, Zern MA. In vitro and in vivo association of transforming growth factor‐beta 1 with hepatic fibrosis. J Cell Biol 108: 2477‐2482, 1989.
 69.Dadgar S, Wang Z, Johnston H, Kesari A, Nagaraju K, Chen YW, Hill DA, Partridge TA, Giri M, Freishtat RJ, Nazarian J, Xuan J, Wang Y, Hoffman EP. Asynchronous remodeling is a driver of failed regeneration in Duchenne muscular dystrophy. J Cell Biol 207: 139‐158, 2014. 10.1083/jcb.201402079.
 70.Deconinck N, Tinsley J, De Backer F, Fisher R, Kahn D, Phelps S, Davies K, Gillis JM. Expression of truncated utrophin leads to major functional improvements in dystrophin‐deficient muscles of mice. Nature Med 3: 1216‐1221, 1997.
 71.De Luna N, Freixas A, Gallano P, Caselles L, Rojas‐Garcia R, Paradas C, Nogales G, Domingues‐Perles R, Gonzalez‐Quereda L, Vilchez JJ, Marquez C, Bautista J, Guerrero A, Salazar JA, Pou A, Illa I, Gallardo E. Dysferlin expression in monocytes: A source of mRNA for mutation analysis. Neuromuscul Disord 17: 69‐76, 2007.
 72.De Luna N, Gallardo E, Illa I. In vivo and in vitro dysferlin expression in human muscle satellite cells. J Neuropathol Exp Neurol 63: 1104‐1113, 2004.
 73.De Luna N, Gallardo E, Sonnet C, Chazaud B, Dominguez‐Perles R, Suarez‐Calvet X, Gherardi RK, Illa I. Role of thrombospondin 1 in macrophage inflammation in dysferlin myopathy. J Neuropathol Exp Neurol 69: 643‐653, 2010.
 74.de Morree A, Flix B, Bagaric I, Wang J, van den Boogaard M, Grand Moursel L, Frants RR, Illa I, Gallardo E, Toes R, van der Maarel SM. Dysferlin regulates cell adhesion in human monocytes. J Biol Chem 288: 14147‐14157, 2013. 10.1074/jbc.M112.448589.
 75.Desguerre I, Mayer M, Leturcq F, Barbet JP, Gherardi RK, Christov C. Endomysial fibrosis in Duchenne muscular dystrophy: A marker of poor outcome associated with macrophage alternative activation. J Neuropathol Exp Neurol 68: 762‐773, 2009.
 76.Dillingham BC, Benny Klimek ME, Gernapudi R, Rayavarapu S, Gallardo E, Van der Meulen JH, Jordan S, Ampong B, Gordish‐Dressman H, Spurney CF, Nagaraju K. Inhibition of inflammation with celastrol fails to improve muscle function in dysferlin‐deficient A/J mice. J Neurol Sci 356: 157‐162, 2015. 10.1016/j.jns.2015.06.042.
 77.DiPietro LA, Nissen NN, Gamelli RL, Koch AE, Pyle JM, Polverini PJ. Thrombospondin 1 synthesis and function in wound repair. Am J Pathol 148: 1851‐1860, 1996.
 78.Disatnik MH, Dhawan J, Yu Y, Beal MF, Whirl MM, Franco AA, Rando TA. Evidence of oxidative stress in mdx mouse muscle: Studies of the pre‐necrotic state. J Neurol Sci 161: 77‐84, 1998.
 79.Drachman DB, Toyka KV, Myer E. Prednisone in Duchenne muscular dystrophy. Lancet 2: 1409‐1412, 1974.
 80.Dufaux B, Order U. Complement activation after prolonged exercise. Clin Chim Acta 179: 45‐49, 1989.
 81.Dumont NA, Bentzinger CF, Sincennes MC, Rudnicki MA. Satellite cells and skeletal muscle regeneration. Compr Physiol 5: 1027‐1059, 2015.
 82.Eghtesad S, Jhunjhunwala S, Little SR, Clemens PR. Rapamycin ameliorates dystrophic phenotype in mdx mouse skeletal muscle. Mol Med 17: 917‐924, 2011.
 83.Ember JA, Sanderson SD, Hugli TE, Morgan EL. Induction of interleukin‐8 synthesis from monocytes by human C5a anaphylatoxin. Am J Pathol 144: 393‐403, 1994.
 84.Emslie‐Smith AM, Arahata K, Engel AG. Major histocompatibility complex class I antigen expression, immunolocalization of interferon subtypes, and T cell‐mediated cytotoxicity in myopathies. Hum Pathol 20: 224‐231, 1989.
 85.Erdely A, Kepka‐Lenhart D, Clark M, Zeidler‐Erdely P, Poljakovic M, Calhoun WJ, Morris SM, Jr. Inhibition of phosphodiesterase 4 amplifies cytokine‐dependent induction of arginase in macrophages. Am J Physiol Lung Cell Mol Physiol 290: L534‐L539, 2006.
 86.Ermolova NV, Martinez L, Vetrone SA, Jordan MC, Roos KP, Sweeney HL, Spencer MJ. Long‐term administration of the TNF blocking drug Remicade (cV1q) to mdx mice reduces skeletal and cardiac muscle fibrosis, but negatively impacts cardiac function. Neuromuscul Disord 24: 583‐595, 2014. 10.1016/j.nmd.2014.04.006.
 87.Ervasti JM, Ohlendieck K, Kahl SD, Gaver MG, Campbell KP. Deficiency of a glycoprotein component of the dystrophin complex in dystrophic muscle. Nature 345: 315‐319, 1990.
 88.Escolar DM, O'Carroll P, Leshner R. Treatment and management of muscular dystrophies. In: Neuromuscular Disorders: Management and Treatment, edited by Bertorini TE. Philadelphia, PA: Elsevier Saunders, 2011.
 89.Fantini MC, Becker C, Monteleone G, Pallone F, Galle PR, Neurath MF. Cutting edge: TGF‐beta induces a regulatory phenotype in CD4+CD25‐ T cells through Foxp3 induction and down‐regulation of Smad7. J Immunol 172: 5149‐5153, 2004.
 90.Farini A, Meregalli M, Belicchi M, Battistelli M, Parolini D, D'Antona G, Gavina M, Ottoboni L, Constantin G, Bottinelli R, Torrente Y. T and B lymphocyte depletion has a marked effect on the fibrosis of dystrophic skeletal muscles in the scid/mdx mouse. J Pathol 213: 229‐238, 2007.
 91.Farini A, Sitzia C, Cassani B, Cassinelli L, Rigoni R, Colleoni F, Fusco N, Gatti S, Bella P, Villa C, Napolitano F, Maiavacca R, Bosari S, Villa A, Torrente Y. Therapeutic potential of immunoproteasome inhibition in Duchenne muscular dystrophy. Mol Ther 24: 1898‐1912, 2016. 10.1038/mt.2016.162.
 92.Fenichel GM, Florence JM, Pestronk A, Mendell JR, Moxley RC, Griggs RC, Brooke MH, Miller JP, Robison J, King W. Long‐term benefit from prednisone therapy in Duchenne muscular dystrophy. Neurology 41: 1874‐1877, 1991.
 93.Ferrington DA, Gregerson DS. Immunoproteasomes: Structure, function, and antigen presentation. Prog Mol Biol Transl Sci 109: 75‐112, 2012. 10.1016/b978‐0‐12‐397863‐9.00003‐1.
 94.Ferrington DA, Husom AD, Thompson LV. Altered proteasome structure, function, and oxidation in aged muscle. FASEB J 19: 644‐646, 2005.
 95.Flanigan KM, Campbell K, Viollet L, Wang W, Gomez AM, Walker CM, Mendell JR. Anti‐dystrophin T cell responses in Duchenne muscular dystrophy: Prevalence and a glucocorticoid treatment effect. Hum Gene Ther 24: 797‐806, 2013. 10.1089/hum.2013.092.
 96.Flores ME, Norgard M, Heinegard D, Reinholt FP, Andersson G. RGD‐directed attachment of isolated rat osteoclasts to osteopontin, bone sialoprotein, and fibronectin. Exp Cell Res 201: 526‐530, 1992.
 97.Fong TC, Wu Y, Kipps TJ. Identification of a promoter element that regulates tissue‐specific expression of the human CD80 (B7.1) gene. J Immunol 157: 4442‐4450, 1996.
 98.Frascarelli M, Rocchi L, Feola I. EMG computerized analysis of localized fatigue in Duchenne muscular dystrophy. Muscle Nerve, 11: 757‐761, 1988.
 99.Frenette J, Cai B, Tidball JG. Complement activation promotes muscle inflammation during modified muscle use. Am J Pathol 156: 2103‐2110, 2000.
 100.Frigas E, Loegering DA, Gleich GJ. Cytotoxic effects of guinea pig eosinophil major basic protein on tracheal epithelium. Lab Invest 42: 35‐43, 1980.
 101.Frisullo G, Frusciante R, Nociti V, Tasca G, Renna R, Iorio R, Patanella AK, Iannaccone E, Marti A, Rossi M, Bianco A, Monforte M, Tonali PA, Mirabella M, Batocchi AP, Ricci E. CD8(+) T cells in facioscapulohumeral muscular dystrophy patients with inflammatory features at muscle MRI. J Clin Immunol 31: 155‐166, 2011.
 102.Gallardo E, Rojas‐Garcia R, de Luna N, Pou A, Brown RH, Jr, Illa I. Inflammation in dysferlin myopathy: Immunohistochemical characterization of 13 patients. Neurology 57: 2136‐2138, 2001.
 103.Galli SJ, Kalesnikoff J, Grimbaldeston MA, Piliponsky AM, Williams CM, Tsai M. Mast cells as “tunable” effector and immunoregulatory cells: Recent advances. Annu Rev Immunol 23: 749‐786, 2005.
 104.Ganji A, Roshan HM, Varasteh A, Moghadam M, Sankian M. The effects of WW2/WW3 domains of Smurf2 molecule on TGF‐β signaling and arginase I gene expression. Cell Biol Int 39: 690‐695, 2015.
 105.Gawlik KI, Durbeej M. Skeletal muscle laminin and MDC1A: Pathogenesis and treatment strategies. Skelet Muscle. 1: 9, 2011. 10.1186/2044‐5040‐1‐9.
 106.Geng LN, Yao Z, Snider L, Fong AP, Cech JN, Young JM, van der Maarel SM, Ruzzo WL, Gentleman RC, Tawil R, Tapscott SJ. DUX4 activates germline genes, retroelements, and immune mediators: implications for facioscapulohumeral dystrophy. Dev Cell 22: 38‐51, 2012. 10.1016/j.devcel.2011.11.013.
 107.Giachelli CM, Lombardi D, Johnson RJ, Murry CE, Almeida M. Evidence for a role of osteopontin in macrophage infiltration in response to pathological stimuli in vivo. Am J Pathol 152: 353‐358, 1998.
 108.Gibertini S, Zanotti S, Savadori P, Curcio M, Saredi S, Salerno F, Andreetta F, Bernasconi P, Mantegazza R, Mora M. Fibrosis and inflammation are greater in muscles of beta‐sarcoglycan‐null mouse than mdx mouse. Cell Tissue Res 356: 427‐443, 2014. 10.1007/s00441‐014‐1854‐4.
 109.Giordano C, Mojumdar K, Liang F, Lemaire C, Li T, Richardson J, Divangahi M, Qureshi S, Petrof BJ. Toll‐like receptor 4 ablation in mdx mice reveals innate immunity as a therapeutic target in Duchenne muscular dystrophy. Hum Mol Genet 24: 2147‐2162, 2015. 10.1093/hmg/ddu735.
 110.Giovarelli M, Santoni A, Jemma C, Musso T, Giuffrida AM, Cavallo G, Landolfo S, Forni G. Obligatory role of IFN‐gamma in induction of lymphokine‐activated and T lymphocyte killer activity, but not in boosting of natural cytotoxicity. J Immunol 141: 2831‐2836, 1988.
 111.Gleich GJ, Adolphson CR, Leiferman KM. The biology of the eosinophilic leukocyte. Annu Rev Med 44: 85‐101, 1993.
 112.Gleich GJ, Frigas E, Loegering DA, Wassom DL, Steinmuller D. Cytotoxic properties of the eosinophil major basic protein. J Immunol 123: 2925‐2927, 1979.
 113.Godfrey C, Escolar D, Brockington M, Clement EM, Mein R, Jimenez‐Mallebrera C, Torelli S, Feng L, Brown SC, Sewry CA, Rutherford M, Shapira Y, Abbs S, Muntoni F. Fukutin gene mutations in steroid‐responsive limb girdle muscular dystrophy. Ann Neurol 60: 603‐610, 2006.
 114.Goldspink G, Fernandes K, Williams PE, Wells DJ. Age‐related changes in collagen gene expression in the muscles of mdx dystrophic and normal mice. Neuromuscul Disord 4: 183‐191, 1994.
 115.Goldstein IM, Roos D, Kaplan HB, Weissmann G. Complement and immunoglobulins stimulate superoxide production by human leukocytes independently of phagocytosis. J Clin Invest 56: 1155‐1163, 1975.
 116.Goodman MG, Chenoweth DE, Weigle WO. Induction of interleukin 1 secretion and enhancement of humoral immunity by binding of human C5a to macrophage surface C5a receptors. J Exp Med 156: 912‐917, 1982.
 117.Gordon JR, Galli SJ. Mast cells as a source of both preformed and immunologically inducible TNF‐alpha/cachectin. Nature 346: 274‐276, 1990.
 118.Gordon S. Alternative activation of macrophages. Nature Rev 3: 23‐35, 2003.
 119.Gordon S, Taylor PR. Monocyte and macrophage heterogeneity. Nature Rev 5: 953‐964, 2005.
 120.Gorelik L, Bar‐Dagan Y, Mokyr MB. Insight into the mechanism(s) through which TNF promotes the generation of T cell‐mediated antitumor cytotoxicity by tumor bearer splenic cells. J Immunol 156: 4298‐308, 1996.
 121.Gorospe JR, Nishikawa BK, Hoffman EP. Recruitment of mast cells to muscle after mild damage. J Neurol Sci 135: 10‐17, 1996.
 122.Gosselin LE, Barkley JE, Spencer MJ, McCormick KM, Farkas GA. Ventilatory dysfunction in mdx mice: Impact of tumor necrosis factor‐alpha deletion. Muscle Nerve 28: 336‐343, 2003.
 123.Gosselin LE, McCormick KM. Targeting the immune system to improve ventilatory function in muscular dystrophy. Med Sci Sports Exerc 36: 44‐51, 2004.
 124.Gosselin LE, Williams JE, Deering M, Brazeau D, Koury S, Martinez DA. Localization and early time course of TGF‐beta 1 mRNA expression in dystrophic muscle. Muscle Nerve 30: 645‐653, 2004.
 125.Grewal PK, Hewitt JE. Mutation of Large, which encodes a putative glycosyltransferase, in an animal model of muscular dystrophy. Biochim Biophys Acta 1573: 216‐224, 2002.
 126.Griggs RC, Moxley RT, III, Mendell JR, Fenichel GM, Brooke MH, Pestronk A, Miller JP. Prednisone in Duchenne dystrophy. A randomized, controlled trial defining the time course and dose response. Clinical Investigation of Duchenne Dystrophy Group. Arch Neurol 48: 383‐388, 1991.
 127.Griggs RC, Moxley RT, III, Mendell JR, Fenichel GM, Brooke MH, Pestronk A, Miller JP, Cwik VA, Pandya S, Robison J, King W, Signore L, Schierbecker J, Florence J, Matheson‐Burden N, Wilson B. Duchenne dystrophy: Randomized, controlled trial of prednisone (18 months) and azathioprine (12 months). Neurology 43: 520‐527, 1993.
 128.Grimbaldeston MA, Nakae S, Kalesnikoff J, Tsai M, Galli SJ. Mast‐cell derived interleukin 10 limits skin pathology in contact dermatitis and chronic irradiation with ultraviolet B. Nat Immunol 10: 1095‐1104, 2007.
 129.Grounds MD, Torrisi J. Anti‐TNFalpha (Remicade) therapy protects dystrophic skeletal muscle from necrosis. FASEB J 18: 676‐682, 2004.
 130.Guerron AD, Rawat R, Sali A, Spurney CF, Pistilli E, Cha HJ, Pandey GS, Gernapudi R, Francia D, Farajian V, Escolar DM, Bossi L, Becker M, Zerr P, de la Porte S, Gordish‐Dressman H, Partridge T, Hoffman EP, Nagaraju K. Functional and molecular effects of arginine butyrate and prednisone on muscle and heart in the mdx mouse model of Duchenne muscular dystrophy. PLoS One 5: e11220, 2010. 10.1371/journal.pone.0011220.
 131.Gussoni E, Pavlath GK, Miller RG, Panzara MA, Posell M, Blau HM, Steinman L. Specific T cell receptor gene rearrangements at the site of muscle degeneration in Duchenne muscular dystrophy. J Immunol 153: 4798‐4805, 1994.
 132.Hammadi A, Billard C, Faussat AM, Kolb JP. Stimulation of iNOS expression and apoptosis resistance in B‐cell chronic lymphocytic leukemia (B‐CLL) cells through engagement of Toll‐like receptor 7 (TLR‐7) and NF‐kappaB activation. Nitric Oxide 19: 138‐145, 2008.
 133.Han R, Frett EM, Levy JR, Rader EP, Lueck JD, Bansal D, Moore SA, Ng R, Beltran‐Valero de Bernabe D, Faulkner JA, Campbell KP. Genetic ablation of complement C3 attenuates muscle pathology in dysferlin‐deficient mice. J Clin Invest 120: 4366‐4374, 2010. 10.1172/JCI42390.
 134.Hara Y, Balci‐Hayta B, Yoshida‐Moriguchi T, Kanagawa M, Beltran‐Valero de Bernabe D, Gundesli H, Willer T, Satz JS, Crawford RW, Burden SJ, Kunz S, Oldstone MB, Accardi A, Talim B, Muntoni F, Topaloglu H, Dincer P, Campbell KP. A dystroglycan mutation associated with limb‐girdle muscular dystrophy. N Engl J Med 364: 939‐946, 2011.
 135.Harris E, Bladen CL, Mayhew A, James M, Bettinson K, Moore U, Smith FE, Rufibach L, Cnaan A, Bharucha‐Goebel DX, Blamire AM, Bravver E, Carlier PG, Day JW, Díaz‐Manera J, Eagle M, Grieben U, Harms M, Jones KJ, Lochmuller H, Mendell JR, Mori‐Yoshimura M, Paradas C, Pegoraro E, Pestronk A, Salort‐Campana E, Schreiber‐Katz O, Semplicini C, Spuler S, Stojkovic T, Straub V, Takeda S, Rocha CT, Walter MC, Bushby K; Jain COS Consortium. The Clinical Outcome Study for dysferlinopathy: An international multicenter study. Neurol Genet 2: e89, 2016. 10.1212/NXG.0000000000000089.
 136.Haslett JN, Sanoudou D, Kho AT, Bennett RR, Greenberg SA, Kohane IS, Beggs AH, Kunkel LM. Gene expression comparison of biopsies from Duchenne muscular dystrophy (DMD) and normal skeletal muscle. Proc Natl Acad Sci U S A 99: 15000‐15005, 2002.
 137.Hayashi YK, Tezak Z, Momoi T, Nonaka I, Garcia CA, Hoffman EP, Arahata K. Massive muscle cell degeneration in the early stage of merosin‐deficient congenital muscular dystrophy. Neuromuscul Disord 11: 350‐359, 2001.
 138.Heil F, Hemmi H, Hochrein H, Ampenberger F, Kirschning C, Akira S, Lipford G, Wagner H, Bauer S. Species‐specific recognition of single‐stranded RNA via Toll‐like receptor 7 and 8. Science 303: 1526‐1529, 2004.
 139.Helliwell TR, Gunhan O, Edwards RH. Mast cells in neuromuscular diseases. J Neurol Sci 98: 267‐276, 1990.
 140.Henriques‐Pons A, Yu Q, Rawavarapu S, Cohen TV, Ampong B, Cha HJ, Jahnke V, Van der Muelen J, Wang D, Jiang W, Kandimalia ER, Agrawal S, Spurney CF, Nagaraju K. Role of Toll‐like receptors in the pathogenesis of dystrophin‐deficient skeletal and heart muscle. Hum Mol Genet 23: 2604‐2617, 2014. 10.1093/hmg/ddt656.
 141.Heydemann A. Severe murine limb‐girdle muscular dystrophy type 2C pathology is diminished by FTY720 treatment. Muscle Nerve 56: 486‐494, 2017. 10.1002/mus.25503.
 142.Hinton VJ, De Vivo DC, Nereo NE, Goldstein E, Stern Y. Poor verbal working memory across intellectual level in boys with Duchenne dystrophy. Neurology 54: 2127‐2132, 2000.
 143.Hinton VJ, De Vivo DC, Nereo NE, Goldstein E, Stern Y. Selective deficits in verbal working memory associated with a known genetic etiology: The neuropsychological profile of Duchenne muscular dystrophy. J Int Neuropsychol Soc 7: 45‐54, 2001.
 144.Hirata A, Masuda S, Tamura T, Kai K, Ojima K, Fukase A, Motoyoshi K, Kamakura K, Miyagoe‐Suzuki Y, Takeda S. Expression profiling of cytokines and related genes in regenerating skeletal muscle after cardiotoxin injection: A role for osteopontin. Am J Pathol 163: 203‐215, 2003.
 145.Hnia K, Gayraud J, Hugon G, Ramonatxo M, De La Porte S, Matecki S, Mornet D. L‐arginine decreases inflammation and modulates the nuclear factor‐kappaB/matrix metalloproteinase cascade in mdx muscle fibers. Am J Pathol 172: 1509‐1519, 2008. 10.2353/ajpath.2008.071009.
 146.Ho M, Gallardo E, McKenna‐Yasek D, De Luna N, Illa I, Brown RH, Jr. A novel, blood‐based diagnostic assay for limb girdle muscular dystrophy 2B and Miyoshi myopathy. Ann Neurol 51: 129‐133, 2002.
 147.Hodgetts S, Radley H, Davies M, Grounds MD. Reduced necrosis of dystrophic muscle by depletion of host neutrophils, or blocking TNFalpha function with Etanercept in mdx mice. Neuromuscul Disord 16: 591‐602, 2006.
 148.Hoffman EP, Brown RH, Jr, Kunkel LM. Dystrophin: The protein product of the Duchenne muscular dystrophy locus. Cell 51: 919‐928, 1987
 149.Holmberg J, Alajbegovic A, Gawlik KI, Elowsson L, Durbeej M. Laminin α2 chain‐deficiency is associated with microRNA deregulation in skeletal muscle and plasma. Front Aging Neurosci 6: 155, 2014. 10.3389/fnagi.2014.00155.
 150.Honda H, Mano Y, Takahashi A. Inflammatory changes in affected muscles of facioscapulohumeral dystrophy. J Neurol 234: 408‐411, 1987.
 151.Hori YS, Kuno A, Hosoda R, Tanno M, Miura T, Shimamoto K, Horio Y. Resveratrol ameliorates muscular pathology in the dystrophic mdx mouse, a model for Duchenne muscular dystrophy. J Pharmacol Exp Ther 338: 784‐794, 2011. 10.1124/jpet.111.183210.
 152.Horsley V, Jansen KM, Mills ST, Pavlath GK. IL‐4 acts as a myoblast recruitment factor during mammalian muscle growth. Cell 113: 483‐494, 2003.
 153.Hsieh CC, Kuro‐o M, Rosenblatt KP, Brobey R, Papaconstantinou J. The ASK1‐Signalosome regulates p38 MAPK activity in response to levels of endogenous oxidative stress in the Klotho mouse models of aging. Aging (Albany NY) 2: 597‐611, 2010.
 154.Huang L, Gebreselassie NG, Gagliardo LF, Ruyechan MC, Lee NA, Lee JJ, Appleton JA. Eosinophil‐derived IL‐10 supports chronic nematode infection. J Immunol 193: 4178‐4187, 2014.
 155.Hughes BP. Creatine phosphokinase in facioscapulohumeral muscular dystrophy. Br Med J 3: 464‐465, 1971.
 156.Husom AD, Peters EA, Kolling EA, Fugere NA, Thompson LV, Ferrington DA. Altered proteasome function and subunit composition in aged muscle. Arch Biochem Biophys 421: 67‐76, 2004.
 157.Ibraghimov‐Beskrovnaya O, Ervasti JM, Leveille CJ, Slaughter CA, Sernett SW, Campbell KP. Primary structure of dystrophin‐associated glycoproteins linking dystrophin to the extracellular matrix. Nature 355: 696‐702, 1992.
 158.Iniesta V, Gomez‐Nieto LC, Molano I, Mohedano A, Carcelen J, Miron C, Alonso C, Corraliza I. Arginase I induction in macrophages, triggered by Th2‐type cytokines, supports the growth of intracellular Leishmania parasites. Parasite Immunol 24: 113‐118, 2002.
 159.Ishikawa K. Cardiac involvement in progressive muscular dystrophy of the Duchenne type. Jpn Heart J 38: 163‐180, 1997.
 160.Israel A, Le Bail O, Hatat D, Piette J, Kieran M, Logeat F, Wallach D, Fellous M, Kourilsky P. TNF stimulates expression of mouse MHC class I genes by inducing an NF kappa B‐like enhancer binding activity which displaces constitutive factors. EMBO J 8: 3793‐3800, 1989.
 161.Jaffe EA, Ruggiero JT, Falcone DJ. Monocytes and macrophages synthesize and secrete thrombospondin. Blood 65: 79‐84, 1985.
 162.Jaffe EA, Ruggiero JT, Leung LK, Doyle MJ, McKeown‐Longo PJ, Mosher DF. Cultured human fibroblasts synthesize and secrete thrombospondin and incorporate it into extracellular matrix. Proc Natl Acad Sci U S A 80: 998‐1002, 1983.
 163.Jain A, Sharma MC, Sarkar C, Bhatia R, Singh S, Handa R. Major histocompatibility complex class I and II detection as a diagnostic tool in idiopathic inflammatory myopathies. Arch Pathol Lab Med 131: 1070‐1076, 2007.
 164.Jeudy S, Wardrop KE, Alessi A, Dominov JA. Bcl‐2 inhibits the innate immune response during early pathogenesis of murine congenital muscular dystrophy. PLoS One 6:e22369, 2011. 10.1371/journal.pone.0022369.
 165.Jimenez‐Mallebrera C, Brown SC, Sewry CA, Muntoni F. Congenital muscular dystrophy: Molecular and cellular aspects. Cell Mol Life Sci 62: 809‐823, 2005.
 166.Joe AW, Yi L, Natarajan A, Le Grand F, So L, Wang J, Rudnicki MA, Rossi FM. Muscle injury activates resident fibro/adipogenic progenitors that facilitate myogenesis. Nat Cell Biol 12: 153‐163, 2010.
 167.Johnson DR, Pober JS. HLA class I heavy‐chain gene promoter elements mediating synergy between tumor necrosis factor and interferons. Mol Cell Biol 14:1322‐1332, 1994.
 168.Kang PB, Griggs RC. Advances in muscular dystrophies. JAMA Neurol 72: 741‐742, 2015. 10.1001/jamaneurol.2014.4621.
 169.Karpati G, Carpenter S. Small‐caliber skeletal muscle fibers do not suffer deleterious consequences of dystrophic gene expression. Am J Med Genet 25: 653‐658, 1986.
 170.Kim EY, Lee JW, Suh MR, Choi WA, Kang SW, Oh HJ. Correlation of serum creatine kinase level with pulmonary function in Duchenne muscular dystrophy. Ann Rehabil Med 41: 306‐312, 2017. 10.5535/arm.2017.41.2.306.
 171.Kissel JT, Burrow KL, Rammohan KW, Mendell JR. Mononuclear cell analysis of muscle biopsies in prednisone‐treated and untreated Duchenne muscular dystrophy. CIDD Study Group. Neurology 41: 667‐672, 1991.
 172.Klinge L, Dean AF, Kress W, Dixon P, Charlton R, Muller JS, Anderson LV, Straub V, Barresi R, Lochmuller H, Bushby K. Late onset in dysferlinopathy widens the clinical spectrum. Neuromuscul Disord 18: 288‐290, 2008. 10.1016/j.nmd.2008.01.004.
 173.Koenig M, Monaco AP, Kunkel LM. The complete sequence of dystrophin predicts a rod‐shaped cytoskeletal protein. Cell 53: 219‐228, 1988.
 174.Koh A, da Silva AP, Bansal AK, Bansal M, Sun C, Lee H, Glogauer M, Sodek J, Zohar R. Role of osteopontin in neutrophil function. Immunology 122: 466‐475, 2007.
 175.Komaki H, Hayashi YK, Tsuburaya R, Sugie K, Kato M, Nagai T, Imataka G, Suzuki S, Saitoh S, Asahina N, Honke K, Higuchi Y, Sakuma H, Saito Y, Nakagawa E, Sugai K, Sasaki M, Nonaka I, Nishino I. Inflammatory changes in infantile‐onset LMNA‐associated myopathy. Neuromuscul Disord 21: 563‐568, 2011. 10.1016/j.nmd.2011.04.010.
 176.Konkay K, Kannan MA, Lingappa L, Uppin MS, Challa S. Congenital muscular dystrophy with inflammation: Diagnostic considerations. Ann Indian Acad Neurol 19: 356‐359, 2016. 10.4103/0972‐2327.186814.
 177.Kostek CA, Dominov JA, Miller JB. Up‐regulation of MHC Class I expression accompanies but is not required for spontaneous myopathy in dysferlin‐deficient SJL/J mice. Am J Pathol 160: 833‐839, 2002.
 178.Kowaljow V, Marcowycz A, Ansseau E, Conde CB, Sauvage S, Matteotti C, Arias C, Corona ED, Nunez NG, Leo O, Wattiez R, Figlewicz D, Laoudj‐Chenivesse D, Belayew A, Coppee F, Rosa AL. The DUX4 gene at the FSHD1A locus encodes a pro‐apoptotic protein. Neuromuscul Disord 17: 611‐623, 2007.
 179.Kuhns DB, Priel DA, Gallin JI. Induction of human monocyte interleukin (IL)‐8 by fibrinogen through the toll‐like receptor pathway. Inflammation 30: 178‐188, 2007.
 180.Kuro‐o M. Klotho as a regulator of fibroblast growth factor signaling and phosphate/calcium metabolism. Curr Opin Nephrol Hypertens 15: 437‐441, 2006.
 181.Kuro‐o M. Klotho and aging. Biochim Biophys Acta 1790: 1049‐1058, 2009.
 182.Kuro‐o M, Matsumura Y, Aizawa H, Kawaguchi H, Suga T, Utsugi T, Ohyama Y, Kurabayashi M, Kaname T, Kume E, Iwasaki H, Iida A, Shiraki‐Iida T, Nishikawa S, Nagai R, Nabeshima Y. Mutation of the mouse klotho gene leads to a syndrome resembling ageing. Nature 390: 45‐51, 1997.
 183.Kurosu H, Ogawa Y, Miyoshi M, Yamamoto M, Nandi A, Rosenblatt KP, Baum MG, Schiavi S, Hu MC, Moe OW, Kuro‐o M. Regulation of fibroblast growth factor‐23 signaling by klotho. J Biol Chem 281: 6120‐6123, 2006.
 184.Laitinen LA, Laitinen A. Modulation of bronchial inflammation: Corticosteroids and other therapeutic agents. Am J Respir Crit Care Med 150: S87‐S90, 1994.
 185.Lamperti C, Cagliani R, Ciscato P, Moroni I, Viri M, Romeo A, Fagiolari G, Prelle A, Comi GP, Bresolin N, Moggio M. Congenital muscular dystrophy with muscle inflammation alpha dystroglycan glycosylation defect and no mutation in FKRP gene. J Neurol Sci 243: 47‐51, 2006.
 186.Lanza L, Scudeletti M, Monaco E, Monetti M, Puppo F, Filaci G, Indiveri F. Possible differences in the mechanism(s) of action of different glucocorticoid hormone compounds. Ann N Y Acad Sci 22: 193‐197, 1999.
 187.Lavine KJ, Epelman S, Uchida K, Weber KJ, Nichols CG, Schilling JD, Ornitz DM, Randolph GJ, Mann DL. Distinct macrophage lineages contribute to disparate patterns of cardiac recovery and remodeling in the neonatal and adult heart. Proc Natl Acad Sci U S A 111: 16029‐16034, 2014. 10.1073/pnas.1406508111.
 188.Law DJ, Allen DL, Tidball JG. Talin, vinculin and DRP (utrophin) concentrations are increased at mdx myotendinous junctions following onset of necrosis. J Cell Sci 107: 1477‐1483, 1994.
 189.Lefaucheur JP, Gjata B, Sebille A. Factors inducing mast cell accumulation in skeletal muscle. Neuropathol Appl Neurobiol 22: 248‐255, 1996.
 190.Lemos DR, Babaeijandaghi F, Low M, Chang CK, Lee ST, Fiore D, Zhang RH, Natarajan A, Nedospasov SA, Rossi FM. Nilotinib reduces muscle fibrosis in chronic muscle injury by promoting TNF‐mediated apoptosis of fibro/adipogenic progenitors. Nat Med 21: 786‐794, 2015. 10.1038/nm.3869.
 191.Li D, Shin JH, Duan D. iNOS ablation does not improve specific force of the extensor digitorum longus muscle in dystrophin‐deficient mdx4cv mice. PLoS One 6(6): e21618, 2011. 10.1371/journal.pone.0021618.
 192.Li H, Mittal A, Makonchuk DY, Bhatnagar S, Kumar A. Matrix metalloproteinase‐9 inhibition ameliorates pathogenesis and improves skeletal muscle regeneration in muscular dystrophy. Hum Mol Genet 18: 2584‐2598, 2009.
 193.Li SA, Watanabe M, Yamada H, Nagai A, Kinuta M, Takei K. Immunohistochemical localization of Klotho protein in brain, kidney, and reproductive organs of mice. Cell Struct Funct 29: 91‐99, 2004.
 194.Li Y, Qi X, Tong X, Wang S. Thrombospondin 1 activates the macrophage Toll‐like receptor 4 pathway. Cell Mol Immunol 10: 506‐512, 2013.
 195.Liang Y, Li G, Chen S, He R, Zhou X, Chen Y, Xu X, Zhu R, Zhang C. Muscle MRI findings in a one‐year‐old girl with merosin‐deficient congenital muscular dystrophy type 1A due to LAMA2 mutation: A case report. Biomed Rep 7: 193‐196, 2017. 10.3892/br.2017.935.
 196.Liu J, Aoki M, Illa I, Wu C, Fardeau M, Angelini C, Serrano C, Urtizberea JA, Hentati F, Hamida MB, Bohlega S, Culper EJ, Amato AA, Bossie K, Oeltjen J, Bejaoui K, McKenna‐Yasek D, Hosler BA, Schurr E, Arahata K, de Jong PJ, Brown RH, Jr. Dysferlin, a novel skeletal muscle gene, is mutated in Miyoshi myopathy and limb girdle muscular dystrophy. Nat Genet 20: 31‐36, 1998.
 197.Longman C, Brockington M, Torelli S, Jimenez‐Mallebrera C, Kennedy C, Khalil N, Feng L, Saran RK, Voit T, Merlini L, Sewry CA, Brown SC, Muntoni F. Mutations in the human LARGE gene cause MDC1D, a novel form of congenital muscular dystrophy with severe mental retardation and abnormal glycosylation of alpha‐dystroglycan. Hum Mol Genet 12: 2853‐2861, 2003.
 198.Lopez A, Sanderson C, Gamble J, Campbell H, Young I, Vadas M. Recombinant human interleukin‐5 is a selective activator of human eosinophil function. J Exp Med 167: 219‐224, 1988.
 199.Love DR, Hill DF, Dickson G, Spurr NK, Byth BC, Marsden RF, Walsh FS, Edwards YH, Davies KE. An autosomal transcript in skeletal muscle with homology to dystrophin. Nature 339: 55‐58, 1989.
 200.Mahjneh I, Marconi G, Bushby K, Anderson LV, Tolvanen‐Mahjneh H, Somer H. Dysferlinopathy (LGMD2B): A 23‐year follow‐up study of 10 patients homozygous for the same frameshifting dysferlin mutations. Neuromuscul Disord 11: 20‐26, 2001.
 201.Many GM, Yokosaki Y, Uaesoontrachoon K, Nghiem PP, Bello L, Dadgar S, Yin Y, Damsker JM, Choen HB, Kornegay JN, Bamman MM, Mosser DM, Nagaraju K, Hoffman EP. OPN‐a induces muscle inflammation by increasing recruitment and activation of pro‐inflammatory macrophages. Exp Physiol 101: 1285‐1300, 2016.
 202.Marshall PA, Williams PE, Goldspink G. Accumulation of collagen and altered fiber‐type ratios as indicators of abnormal muscle gene expression in the mdx dystrophic mouse. Muscle Nerve 12: 528‐537, 1989.
 203.Martinez CO, McHale MJ, Wells JT, Ochoa O, Michalek JE, Shireman. Regulation of skeletal muscle regeneration by CCR2 activating chemokines is directly related to macrophage recruitment. Am J Physiol Regul Integr Comp Physiol 299: R832‐R842, 2010.
 204.Mathews KD, Rapisarda D, Bailey HL, Murray JC, Schelper RL, Smith R. Phenotypic and pathologic evaluation of the myd mouse. A candidate model for facioscapulohumeral dystrophy. J Neuropathol Exp Neurol 54: 601‐606, 1995.
 205.Matsuda R, Nishikawa A, Tanaka H. Visualization of dystrophic muscle fibers in mdx mouse by vital staining with Evans blue: Evidence of apoptosis in dystrophin‐deficient muscle. J Biochem 118: 959‐964, 1995.
 206.Matsumura K, Ervasti JM, Ohlendieck K, Kahl SD, Campbell, P. Association of dystrophin‐related protein with dystrophin‐associated proteins in mdx mouse muscle. Nature 360: 588‐591, 1992.
 207.Matsumura Y, Aizawa H, Shiraki‐Iida T, Nagai R, Kuro‐o M, Nabeshima Y. Identification of the human klotho gene and its two transcripts encoding membrane and secreted klotho protein. Biochem Biophys Res Commun 242: 626‐630, 1998.
 208.Mauro A. Satellite cell of skeletal muscle fibers. J Biophys Biochem Cytol 9: 493‐495, 1961.
 209.McDouall RM, Dunn MH, Dubowitz V. Expression of class I and class II MHC antigens in neuromuscular diseases. J Neurol Sci 89: 213‐226, 1989.
 210.McDouall RM, Dunn MJ, Dubowitz, V. Nature of the mononuclear infiltrate and the mechanism of muscle damage in juvenile dermatomyositis and Duchenne muscular dystrophy. J Neurol Sci 99: 199‐217, 1990.
 211.Mehuron T, Kumar A, Duarte L, Yamauchi J, Accorsi A, Girgenrath M. Dysregulation of matricellular proteins is an early signature of pathology in laminin‐deficient muscular dystrophy. Skelet Muscle 4: 14, 2014. 10.1186/2044‐5040‐4‐14.
 212.Meinen S, Lin S, Ruegg MA. Angiotensin II type 1 receptor antagonists alleviate muscle pathology in the mouse model for laminin‐α2‐deficient congenital muscular dystrophy (MDC1A). Skelet Muscle 2: 18, 2012. 10.1186/2044‐5040‐2‐18.
 213.Mestas J, Hughes CC. Of mice and not men: Differences between mouse and human immunology. J Immunol 172: 2731‐2738, 2004.
 214.Metcalfe DD, Baram D, Mekori YA. Mast cells. Physiol Rev 77: 1033‐1079, 1997.
 215.Millay DP, Maillet M, Roche JA, Sargent MA, McNally EM, Bloch RJ, Molkentin JD. Genetic manipulation of dysferlin expression in skeletal muscle: Novel insights into muscular dystrophy. Am J Pathol 175: 1817‐1823, 2009.
 216.Minetti C, Tanji K, Bonilla E. Immunologic study of vinculin in Duchenne muscular dystrophy. Neurology 42: 1751‐1754, 1992.
 217.Miyoshi K. Echocardiographic evaluation of fibrous replacement in the myocardium of patients with Duchenne muscular dystrophy. Br Heart J 66: 452‐455, 1991.
 218.Mojumdar K, Giordano C, Lemaire C, Liang F, Divangahi M, Qureshi ST, Petrof BJ. Divergent impact of Toll‐like receptor 2 deficiency on repair mechanisms in healthy muscle versus Duchenne muscular dystrophy. J Pathol 239: 10‐22, 2016. 10.1002/path.4689.
 219.Mojumdar K, Liang F, Giordano C, Lemaire C, Danialou G, Okazaki T, Bourdon J, Rafei M, Galipeau J, Divangahi M, Petrof BJ. Inflammatory monocytes promote progression of Duchenne muscular dystrophy and can be therapeutically targeted via CCR2. EMBO Mol Med 6: 1476‐1492, 2014. 10.15252/emmm.201403967.
 220.Molnar M, Dioszeghy P, Mechler F. Inflammatory changes in facioscapulohumeral muscular dystrophy. Eur Arch Psychiatry Clin Neurosci 241: 105‐108, 1991.
 221.Moqbel R, Ying S, Barkans J, Newman TM, Kimmitt P, Wakelin M, Taborda‐Barata L, Meng Q, Corrigan CJ, Durham SR, Kay AB. Identification of messenger RNA for IL‐4 in human eosinophils with granule localization and release of the translated product. J Immunol 155: 4939‐4947, 1995.
 222.Morgan BP. Regulation of the complement membrane attack pathway. Crit Rev Immunol 19: 173‐198, 1999.
 223.Mori Y, Ishida W, Bhattacharyya S, Li Y, Platanias LC, Varga J. Selective inhibition of activin receptor‐like kinase 5 signaling blocks profibrotic transforming growth factor beta responses in skin fibroblasts. J Arthritis Rheum 50: 4008‐4021, 2004.
 224.Morrison J, Palmer DB, Cobbold S, Partridge T, Bou‐Gharios G. Effects of T‐lymphocyte depletion on muscle fibrosis in the mdx mouse. Am J Pathol 166: 1701‐1710, 2005.
 225.Motojima M, Matsusaka T, Kon V, Ichikawa I. Fibrinogen that appears in Bowman's space of proteinuric kidneys in vivo activates podocyte Toll‐like receptors 2 and 4 in vitro. Nephron Exp Nephrol 114: e39‐e47, 2010. 10.1159/000254390.
 226.Muller‐Eberhard HJ. The membrane attack complex of complement. Ann Rev Immunol 4: 503‐528, 1986.
 227.Mumby SM, Abbott‐Brown D, Raugi GJ, Bornstein PJ. Regulation of thrombospondin secretion by cells in culture. Cell Physiol 120: 280‐288, 1984.
 228.Munsat TL, Bradley WG. Serum creatine phosphokinase levels and prednisone treated muscle weakness. Neurology 27: 96‐97, 1977.
 229.Munsat TL, Piper D, Cancilla P, Mednick J. Inflammatory myopathy with facioscapulohumeral distribution. Neurology 22: 335‐347, 1972.
 230.Murahashi M, Wakayama Y, Kumagai T, Kobayashi T, Yamashita S, Misugi N, Miyake S, Shibuya S, Jimi T, Oniki H. Observations of muscle plasma membrane undercoats in Duchenne and Fukuyama muscular dystrophies. Med Electron Microsc 2: 102‐110, 1995.
 231.Murry CE, Giachelli CM, Schwartz SM, Vracko R. Macrophages express osteopontin during repair of myocardial necrosis. Am J Pathol 145: 1450‐1462, 1994.
 232.Nagappa M, Nalini A, Narayanappa G. Major histocompatibility complex and inflammatory cell subtype expression in inflammatory myopathies and muscular dystrophies. Neurol India 61: 614‐621, 2013. 10.4103/0028‐3886.125264.
 233.Nagaraju K, Rawat R, Veszelovszky E, Thapliyal R, Kesari A, Sparks S, Raben N, Plotz P, Hoffman EP. Dysferlin deficiency enhances monocyte phagocytosis: A model for the inflammatory onset of limb‐girdle muscular dystrophy 2B. Am J Pathol 172: 774‐785, 2008. 10.2353/ajpath.2008.070327.
 234.Nagel A, Lehmann‐Horn F, Engel AG. Neuromuscular transmission in the mdx mouse. Muscle Nerve 13: 742‐749, 1990.
 235.Nahirney PC, Dow PR, Ovalle WK. Quantitative morphology of mast cells in skeletal muscle of normal and genetically dystrophic mice. Anat Rec 247: 341‐349, 1997.
 236.Nakajima H, Gleich GJ, Kita H. Constitutive production of IL‐4 and IL‐10 stimulated production of IL‐8 by normal peripheral blood eosinophils. J Immunol 156: 4859‐4866, 1996.
 237.Nakano K, Okugawa K, Furuichi H, Matsui Y, Sohmura Y. Augmentation of the generation of cytotoxic T lymphocytes against syngeneic tumor cells by recombinant human tumor necrosis factor. Cell Immunol 120: 154‐164, 1989.
 238.Nelson CA, Hunter RB, Quigley LA, Girgenrath S, Weber WD, McCullough JA, Dinardo CJ, Keefe KA, Ceci L, Clayton NP, McVie‐Wylie A, Cheng SH, Leonard JP, Wentworth BM. Inhibiting TGF‐β activity improves respiratory function in mdx mice. Am J Pathol 178: 2611‐2621, 2011. 10.1016/j.ajpath.2011.02.024.
 239.Nemoto H, Konno S, Nakazora H, Miura H, Kurihara T. Histological and immunohistological changes of the skeletal muscles in older SJL/J mice. Eur Neurol 57: 19‐25, 2007.
 240.Nghiem PP, Kornegay JN, Uaesoontrachoon K, Bello L, Yin Y, Kesari A, Mittal P, Schatzberg SJ, Many GM, Lee NH, Hoffman EP. Osteopontin is linked with AKT, FoxO1, and myostatin in skeletal muscle cells. Muscle Nerve 56(6):1119‐1127, 2017. 10.1002/mus.25752.
 241.Nguyen HX, Lusis AJ, Tidball JG. Null mutation of myeloperoxidase in mice prevents mechanical activation of neutrophil lysis of muscle cell membranes in vitro and in vivo. J Physiol 565: 403‐413, 2005.
 242.Nguyen HX, Tidball JG. Interactions between neutrophils and macrophages promote macrophage killing of rat muscle cells in vitro. J Physiol 547: 125‐132, 2003.
 243.Nguyen TM, Ellis JM, Love DR, Davies KE, Gatter KC, Dickson G, Morris GE. Localization of the DMDL gene‐encoded dystrophin related protein using a panel of nineteen monoclonal antibodies: Presence at neuromuscular junctions, in the sarcolemma of dystrophic skeletal muscle, in vascular and other smooth muscles, and in proliferating brain cell lines. J Cell Biol 115: 1695‐1700, 1991.
 244.NINDS Muscular Dystrophy Information Page. Retrieved from: https://www.ninds.nih.gov/disorders/all‐disorders/muscular‐dystrophy‐ information‐page.
 245.O'Brien ER, Garvin MR, Stewart DK, Hinohara T, Simpson JB, Schwartz SM, Giachelli CM. Osteopontin is synthesized by macrophage, smooth muscle, and endothelial cells in primary and restenotic human coronary atherosclerotic plaques. Arterioscler Thromb 14: 1648‐1656, 1994.
 246.Ohlendieck K, Campbell KP. Dystrophin‐associated proteins are greatly reduced in skeletal muscle from mdx mice. J Cell Biol 115: 1685‐1694, 1991.
 247.O'Regan AW, Chupp GL, Lowry JA, Goetschkes M, Mulligan N, Berman JS. Osteopontin is associated with T cells in sarcoid granulomas and has T cell adhesive and cytokine‐like properties in vitro. J Immunol 162: 1024‐1031, 1999.
 248.Osbourne RJ, Welle S, Venance SL, Thornton CA, Tawil R. Expression profile of FSHD supports a link between retinal vasculopathy and muscular dystrophy. Neurology 68: 569‐77, 2007.
 249.Page‐McCaw A, Ewald AJ, Werb Z. Matrix metalloproteinases and the regulation of tissue remodelling. Nat Rev Mol Cell Biol 8: 221‐233, 2007.
 250.Pan Y, Chen C, Shen Y, Zhu CH, Wang G, Wang XC, Chen HQ, Zhu MS. Curcumin alleviates dystrophic muscle pathology in mdx mice. Mol Cells 25: 531‐537, 2008.
 251.Pangburn MK, Muller‐Eberhard HJ. The alternative pathway of complement. Springer Semin Immunopathol 7: 163‐192, 1984.
 252.Panicker JB, Chacko G, Patil AK, Alexander M, Muliyil J. Immunohistochemical differentiation of inflammatory myopathies. Neurol India 59: 513‐520, 2011. 10.4103/0028‐3886.84329.
 253.Park JS, Gamboni‐Robertson F, He Q, Svetkauskaite D, Kim JY, Strassheim D, Sohn JW, Yamada S, Maruyama I, Banerjee A, Ishizaka A, Abraham E. High mobility group box 1 protein interacts with multiple Toll‐like receptors. Am J Physiol Cell Physiol 290: C917‐C924, 2006.
 254.Passaquin AC, Metzinger L, Léger JJ, Warter JM, Poindron P. Prednisolone enhances myogenesis and dystrophin‐related protein in skeletal muscle cell cultures from mdx mouse. J Neurosci Res 35: 363‐372, 1993.
 255.Passerini L, Bernasconi P, Baggi F, Confalonieri P, Cozzi F, Cornelio F, Mantegazza R. Fibrogenic cytokines and extent of fibrosis in muscle of dogs with X‐linked golden retriever muscular dystrophy. Neuromuscul Disord 12: 828‐835, 2002.
 256.Passos‐Bueno MR, Bashir R, Moreira ES, Vain zof M, Marie SK, Vasquez L, Iughetti P, Bakker E, Keers S, Stephenson A, Strachan T, Mahneh I, Weissenbach J, Bushby K, Mayana Z. Confirmation of the 2p locus for the mild autosomal recessive limb‐girdle muscular dystrophy gene (LGMD2B) in three families allows refinement of the candidate region. Genomics 27: 192‐195, 1995.
 257.Pastoret C, Sebille A. mdx mice show progressive weakness and muscle deterioration with age. J Neurol Sci 129: 97‐105, 1995.
 258.Pegoraro E, Mancias P, Swerdlow SH, Raikow RB, Garcia C, Marks H, Crawford T, Carver V, Di Cianno B, Hoffman EP. Congenital muscular dystrophy with primary laminin alpha2 (merosin) deficiency presenting as inflammatory myopathy. Ann Neurol 40: 782‐791, 1996.
 259.Petrof BJ, Shrager JB, Stedman HH, Kelly AM, Sweeney HL. Dystrophin protects the sarcolemma from stresses developed during muscle contraction. Proc Natl Acad Sci U S A 90: 3710‐3714, 1993.
 260.Philippidis P, Mason JC, Evans BJ, Nadra I, Taylor KM, Haskard DO, Landis RC. Hemoglobin scavenger receptor CD163 mediates interleukin‐10 release and heme oxygenase‐1 synthesis: Antiinflammatory monocyte‐macrophage responses in vitro, in resolving skin blisters in vivo, and after cardiopulmonary bypass surgery. Circ Res 94: 119‐126, 2004.
 261.Porter JD, Khanna S, Kaminski HJ, Rao JS, Merriam AP, Richmonds CR, Leahy P, Li J, Guo W, Andrade FH. A chronic inflammatory response dominates the skeletal muscle molecular signature in dystrophin‐deficient mdx mice. Hum Mol Genet 11: 263‐272, 2002.
 262.Prior TW, Papp AC, Snyder PJ, Burghes AH, Bartolo C, Sedra MS, Western LM, Mendell JR. A missense mutation in the dystrophin gene in a Duchenne muscular dystrophy patient. Nat Genet 4: 357‐360, 1993.
 263.Radley HG, Davies MJ, Grounds MD. Reduced muscle necrosis and long‐term benefits in dystrophic mdx mice after cV1q (blockade of TNF) treatment. Neuromuscul Disord 18: 227‐238, 2008. 10.1016/j.nmd.2007.11.002.
 264.Radley HG, Grounds MD. Cromolyn administration (to block mast cell degranulation) reduces necrosis of dystrophic muscle in mdx mice. Neurobiol Dis 23: 387‐397, 2006.
 265.Rafael JA, Tinsley JM, Potter AC, Deconinck AE, Davies KE. Skeletal muscle‐specific expression of a utrophin transgene rescues utrophin‐dystrophin deficient mice. Nature Genetics 19: 79‐82, 1998.
 266.Raghow B, Irish P, Kang AH. Coordinate regulation of transforming growth factor beta gene expression and cell proliferation in hamster lungs undergoing bleomycin‐induced pulmonary fibrosis. J Clin Invest 84: 1836‐1842, 1989.
 267.Rahimov F, King OD, Leung DG, Bibat GM, Emerson CP, Jr, Kunkel LM, Wagner KR. Transcriptional profiling in facioscapulohumeral muscular dystrophy to identify candidate biomarkers. Proc Natl Acad Sci U S A 109: 16234‐16239, 2012. 10.1073/pnas.1209508109.
 268.Rainbird MA, Macmillan D, Meeusen EN. Eosinophil‐mediated killing of Haemonchus contortus larvae: Effect of eosinophil activation and role of antibody, complement and IL‐5. Parasite Immunol 20: 93‐103, 1998.
 269.Rajakumar D, Senguttuvan S, Alexander M, Oommen A. Involvement of oxidative stress, nuclear factor kappa B and the ubiquitin proteasomal pathway in dysferlinopathy. Life Sci 108: 54‐61, 2014.
 270.Richards DM, Brogden RN, Heel RC, Speight TM, Avery GS. Oxatomide. A review of its pharmacodynamics properties and therapeutic efficacy. Drugs 27: 210‐231, 1984.
 271.Rifai Z, Welle S, Moxley RT, III, Lorenson M, Griggs RC. Effect of prednisone on protein metabolism in Duchenne dystrophy. Am J Physiol 268: E67‐E74, 1995.
 272.Roberts AB, Sporn MB, Assoian RK, Smith JM, Roche NS, Wakefield, LM, Heine UI, Liotta LA, Falanga V, Kehrl JH, Fauci AS. Transforming growth factor type beta: Rapid induction of fibrosis and angiogenesis in vivo and stimulation of collagen formation in vitro. Proc Natl Acad Sci U S A 83: 4167‐4171, 1986.
 273.Roche JA, Lovering RM, Roche R, Ru LW, Reed PW, Bloch RJ. Extensive mononuclear infiltration and myogenesis characterize recovery of dysferlin‐null skeletal muscle from contraction‐induced injuries. Am J Physiol Cell Physiol 298: C298‐C312, 2010.
 274.Roche JA, Tulapurkar ME, Mueller AL, van Rooijen N, Hasday JD, Lovering RM, Bloch RJ. Myofiber damage precedes macrophage infiltration after in vivo injury in dysferlin‐deficient A/J mouse skeletal muscle. Am J Pathol 185: 1686‐1698, 2015.
 275.Roig M, Roma J, Fargas A, Munell F. Longitudinal pathologic study of the gastrocnemius muscle group in mdx mice. Acta Neuropathol 107: 27‐34, 2004.
 276.Rosales XQ, Malik V, Sneh A, Chen L, Lewis S, Kota J, Gastier‐Foster JM, Astbury C, Pyatt R, Reshmi S, Rodino‐Klapac LR, Clark KR, Mendell JR, Sahenk Z. Impaired regeneration in LGMD2A supported by increased PAX7‐positive satellite cell content and muscle‐specific microrna dysregulation. Muscle Nerve 47: 731‐739, 2013.
 277.Rosenberg HF. Recombinant human eosinophil cationic protein. Ribonuclease activity is not essential for cytotoxicity. J Biol Chem 270: 7876‐7881, 1995
 278.Rubin B, Smith A, Romaschin A, Walker P. Participation of the complement system in ischemia/reperfusion injury. Microcirc Endothelium Lymphatics 5: 207‐221, 1989.
 279.Rybalka E, Timpani CA, Cooke MB, Williams AD, Hayes A. Defects in mitochondrial ATP synthesis in dystrophin‐deficient mdx skeletal muscles may be caused by Complex I insufficiency. PLoS One 9(12):e115763, 2014. 10.1371/journal.pone.0115763.
 280.Sander M, Chavoshan B, Harris SA, Iannaccone ST, Stull JT, Thomas GD, Victor RG. Functional muscle ischemia in neuronal nitric oxide synthase‐deficient skeletal muscle of children with Duchenne muscular dystrophy. Proc Natl Acal Sci U S A 97: 13818‐13823, 2000.
 281.Sansome A, Royston P, Dubowitz V. Steroids in Duchenne muscular dystrophy; pilot study of a new low‐dosage schedule. Neuromuscul Disord 3: 567‐569, 1993.
 282.Satriano J. Arginine pathways and the inflammatory response: Interregulation of nitric oxide and polyamines: Review article. Amino Acids 23: 321‐329, 2004.
 283.Schaer DJ, Boretti FS, Hongegger A, Poehler D, Linnscheid P, Staege H, Muller C, Schoedon G, Schaffner A. Molecular cloning and characterization of the mouse CD163 homologue, a highly glucocorticoid‐inducible member of the scavenger receptor cysteine‐rich family. Immunogenetics 53: 170‐177, 2001.
 284.Schafer H, Mathey D, Hugo F, Bhakdi S. Deposition of the terminal C5b‐9 complement complex in infarcted areas of human myocardium. J Immunol 137: 1945‐1949, 1986.
 285.Schultz‐Cherry S, Chen F, Mosher DF, Misenheimer TM, Krutsch HC, Roberts DD, Murphy‐Ullrich JE. Regulation of transforming growth factor‐beta activation by discrete sequences of thrombospondin 1. J Biol Chem 270: 7304‐7310, 1995.
 286.Selcen D, Stilling G, Engel AG. The earliest pathologic alterations in dysferlinopathy. Neurology 56: 1472‐1481, 2001.
 287.Shakhov AN, Collart MA, Vassalli P, Nedospasov SA, Jongeneel CV. Kappa B‐type enhancers are involved in lipopolysaccharide‐mediated transcriptional activation of the tumor necrosis factor alpha gene in primary macrophages. J Exp Med 171: 35‐47, 1990.
 288.Shi HP, Fishel RS, Efron DT, Williams JZ, Fishel MH, Barbul A. Effect of supplemental ornithine on wound healing. J Surg Res 106: 299‐302, 2002.
 289.Shiao T, Fond A, Deng B, Wehling‐Henricks M, Adams ME, Froehner SC, Tidball JG. Defects in neuromuscular junction structure in dystrophic muscle are corrected by expression of a NOS transgene in dystrophin‐deficient muscles, but not in muscles lacking alpha‐ and beta1‐syntrophins. Hum Mol Genet 13: 1873‐1884, 2004.
 290.Shireman PK, Contreras‐Shannon V, Ochoa O, Karla BP, Michalek JE, McManus LM. MCP 1 deficiency causes altered inflammation with impaired skeletal muscle regeneration. J Leukoc Biol 81: 775‐785, 2007.
 291.Sica A, Dorman L, Viggiano V, Cippitelli M, Ghosh P, Rice N, Young HA. Interaction of NF‐kappB and NFAT with the interferon‐gamma promoter. J Biol Chem 272: 30412‐30420, 1997.
 292.Silberstein D. Eosinophil function in health and disease. Crit Rev Oncol Hematol 19: 47‐77, 1994.
 293.Simon MM, Hochgeschwender U, Brugger U, Landolfo S. Monoclonal antibodies to interferon‐gamma inhibit interleukin 2‐dependent induction of growth and maturation in lectin/antigen‐reactive cytolytic T lymphocyte precursors. J Immunol 136: 2755‐2762, 1986.
 294.Singh RP, Patarca R, Schwartz J, Singh P, Cantor, H. Definition of the specific interaction between the early T lymphocyte activation 1 (Eta‐1) protein and the murine macrophages in vitro and its effect upon macrophages in vivo. J Exp Med 171: 1931‐1942, 1990.
 295.Sklar RM, Brown RH, Jr. Methylprednisolone increases dystrophin levels by inhibiting myotube death during myogenesis of normal human muscle in vitro. J Neurol Sci 101: 73‐81, 1991.
 296.Spencer MJ, Marino MW, Winckler WM. Altered pathological progression of diaphragm and quadriceps muscle in TNF‐deficient, dystrophin‐deficient mice. Neuromuscul Disord 10: 612‐619, 2000.
 297.Spencer MJ, Montecino‐Rodriguez E, Dorshkind K, Tidball JG. Helper (CD4(+)) and cytotoxic (CD8(+)) T cells promote the pathology of dystrophin‐deficient muscle. Clin Immunol 98: 235‐243, 2001.
 298.Spencer MJ, Walsh CM, Dorshkind KA, Rodriguez EM, Tidball JG. Myonuclear apoptosis in mdx muscle occurs by perforin‐mediated cytotoxicity. J Clin Invest 99: 2745‐2751, 1997.
 299.Spuler S, Engel AG. Unexpected sarcolemmal complement membrane attack complex deposits on nonnecrotic muscle fibers in muscular dystrophies. Neurology 50: 41‐46, 1988.
 300.Standal T, Borset M, Sundan A. Role of osteopontin in adhesion, migration, cell survival and bone remodeling. Exp Oncol 26: 179‐184, 2004.
 301.Stedman HH, Sweeney HL, Shrager JB, Maguire HC, Panettieri RA, Petrof B, Narusawa M, Leferovich JM, Sladky JT, Kelly AM. The mdx mouse diaphragm reproduces the degenerative changes of Duchenne muscular dystrophy. Nature 352: 536‐539, 1991.
 302.Stewart JD, Masi TL, Cumming AE, Molnar GM, Wentworth BM, Sampath K, McPherson JM, Yaeger PC. Characterization of proliferating human skeletal muscle‐derived cells in vitro: Differential modulation of myoblast markers by TGF‐beta2. J Cell Physiol 196: 70‐78, 2003.
 303.Straub V, Rafael JA, Chamberlain JS, Campbell KP. Animal models for muscular dystrophy show different patterns of sarcolemmal disruption. J Cell Biol 139: 375‐385, 1997.
 304.Sugie K, Hayashi YK, Kin T, Goto K, Nichino I, Ueno S. Teaching NeuroImages: Hemiatrophy as a clinical presentation in facioscapulohumeral muscular dystrophy. Neurology 73: e24, 2009. 10.1212/WNL.0b013e3181b04af9.
 305.Sulahian TH, Hogger P, Wahner AE, Wardwell K, Goulding NJ, Sorg C, Droste A, Stehling M, Wallace PK, Morganelli PM, Guyre PM. Human monocytes express CD163, which is upregulated by IL‐10 and identical to p155. Cytokine 12: 1312‐1321, 2000.
 306.Sun J, Sukhova GK, Yang M, Wolters PJ, MacFarlane LA, Libby P, Sun C, Zhang Y, Liu J, Ennis TL, Knispel R, Xiong W, Thompson RW, Baxter BT, Shi GP. Mast cells modulate the pathogenesis of elastase‐induced abdominal aortic aneurysms in mice. J Clin Invest 117: 3359‐3368, 2007.
 307.Tai P, Hayes DJ, Clark JB, Spry CJ. Toxic effects of human eosinophil products on isolated rat heart cells in vitro. Biochem J 204: 75‐80, 1982.
 308.Talsness DM, Belanto JJ, Ervasti JM. Disease‐proportional proteasomal degradation of missense dystrophins. Proc Natl Acad Sci U S A 112: 12414‐12419, 2015.
 309.Tan K, Lawler J. The interaction of thrombospondins with extracellular matrix proteins. J Cell Commun Signal 3: 177‐187, 2009. 10.1007/s12079‐009‐0074‐2.
 310.Tawil R, McDermott MP, Pandya S, King W, Kissel J, Mendell JR, Griggs RC. A pilot trial of prednisone in facioscapulohumeral muscular dystrophy. FSH‐DY Group. Neurology 48: 46‐49, 1997.
 311.Tawil R, van der Maarel SM, Tapscott SJ. Facioscapulohumeral dystrophy: The path to consensus on pathophysiology. Skelet Muscle 4: 12, 2014. 10.1186/2044‐5040‐4‐12.
 312.Terrill JR, Duong MN, Turner R, Le Guiner C, Boyatzis A, Kettle AJ, Grounds MD, Arthur PG. Levels of inflammation and oxidative stress, and a role for taurine in dystropathology of the Golden Retriever Muscular Dystrophy dog model for Duchenne Muscular Dystrophy. Redox Biol 9: 276‐286, 2016. 10.1016/j.redox.2016.08.016.
 313.Terrill JR, Pinniger GJ, Graves JA, Grounds MD, Arthur PG. Increasing taurine intake and taurine synthesis improves skeletal muscle function in the mdx mouse model for Duchenne muscular dystrophy. J Physiol 594: 3095‐3110, 2016. 10.1113/JP271418.
 314.Terrill JR, Radley‐Crabb HG, Iwasaki T, Lemckert FA, Arthur PG, Grounds MD. Oxidative stress and pathology in muscular dystrophies: Focus on protein thiol oxidation and dysferlinopathies. FEBS J 280: 4149‐4164, 2013. 10.1111/febs.12142.
 315.Thomas GD, Sander M, Lau KS, Huang PL, Stull JT, Victor RG. Impaired metabolic modulation of alpha‐adrenergic vasoconstriction in dystrophin‐deficient skeletal muscle. Proc Natl Acad Sci U S A 95: 15090‐15095, 1998.
 316.Tidball JG. Regulation of muscle growth and regeneration by the immune system. Nat Rev Immunol 17: 165‐178, 2017. 10.1038/nri.2016.150.
 317.Tidball JG, Albrecht DE, Lokensgard BE, Spencer MJ. Apoptosis precedes necrosis of dystrophin‐deficient muscle. J Cell Sci 108: 2197‐2204, 1995.
 318.Tidball JG, Wehling‐Henricks M. Evolving therapeutic strategies for Duchenne muscular dystrophy: Targeting downstream events. Pediatr Res 56: 831‐841, 2004.
 319.Tidball JG, Wehling‐Henricks M. Nitric oxide synthase deficiency and the pathophysiology of muscular dystrophy. J Physiol 592: 4627‐4638, 2014.
 320.Timpani CA, Trewin AJ, Stojanovska V, Robinson A, Goodman CA, Nurgali K, Betik AC, Stepto N, Hayes A, McConell GK, Rybalka E. Attempting to compensate for reduced neuronal nitric oxide synthase protein with nitrate supplementation cannot overcome metabolic dysfunction but rather has detrimental effects in dystrophin‐deficient mdx muscle. Neurotherapeutics 14: 429‐446, 2017. 10.1007/s13311‐016‐0494‐7.
 321.Tinsley JM, Potter AC, Phelps SR, Fisher R, Trickett JI, Davies KE. Amelioration of the dystrophic phenotype of mdx mice using a truncated utrophin transgene. Nature 384: 349‐353, 1996.
 322.Tonkin J, Temmerman L, Sampson RD, Gallego‐Colon E, Barberi L, Bilbao D, Schneider MD, Musaro A, Rosenthal N. Monocyte/macrophage‐derived IGF‐1 orchestrates murine skeletal muscle regeneration and modulates autocrine polarization. Mol Ther 23: 1189‐1200, 2015. 10.1038/mt.2015.66.
 323.Traynor TR, Herring AC, Dorf ME, Kuziel WA, Toews GB, Huffnagle GB. Differential roles of CC chemokine ligand 2/monocyte chemotactic protein‐1 and CCR2 in the development of T1 immunity. J Immunol 168: 4659‐4666, 2002.
 324.Tumurkhuu G, Koide N, Dagvadorj J, Noman AS, Khuda II, Naiki Y, Komatsu T, Yoshida T, Yokochi T. B1 cells produce nitric oxide in response to a series of toll‐like receptor ligands. Cell Immunol 261: 122‐127, 2010.
 325.Turki A, Hayot M, Carnac G, Pillard F, Passerieux E, Bommart S, Raynaud de Mauverger E, Hugon G, Pincemail J, Pietri S, Lambert K, Belayew A, Vassetzky Y, Juntas Morales R, Mercier J, Laoudj‐Chenivesse D. Functional muscle impairment in facioscapulohumeral muscular dystrophy is correlated with oxidative stress and mitochondrial dysfunction. Free Radic Biol Med 53: 1068‐1079, 2012.
 326.Uaesoontrachoon K, Cha HJ, Ampong B, Sali A, Vandermeulen J, Wei B, Creeden B, Huynh T, Quinn J, Tatem K, Rayavarapu S, Hoffman EP, Nagaraju K. The effects of MyD88 deficiency on disease phenotype in dysferlin‐deficient A/J mice: Role of endogenous TLR ligands. J Pathol 231: 199‐209, 2013.
 327.Uaesoontrachoon K, Wasgewatte Wijesinghe DK, Mackie EJ, Pagel CN. Osteopontin deficiency delays inflammatory infiltration and the onset of muscle regeneration in a mouse model of muscle injury. Dis Model Mech 6, 197‐205, 2013.
 328.Uaesoontrachoon K, Yoo HJ, Tudor EM, Pike RN, Mackie EJ, Pagel CN. Osteopontin and skeletal muscle myoblasts: Association with muscle regeneration and regulation of myoblast function in vitro. Int J Biochem Cell Biol 40, 2303‐2314, 2008.
 329.Ueda A, Ishigatsubo Y, Okubo T, Yoshimura T. Transcriptional regulation of the human monocyte chemoattractant protein‐1 gene. Cooperation of two NF‐kappaB sites and NF‐kappaB/Rel subunit specificity. J Biol Chem 272: 31092‐31099, 1997.
 330.Ueda A, Okuda K, Ohno S, Shirai A, Igarashi T, Matsunaga K, Fukushima J, Kawamoto S, Ishigatsubo Y, Okubo T. NF‐kappa B and Sp1 regulate transcription of the human monocyte chemoattractant protein‐1 gene. J Immunol 153: 2052‐2063, 1994.
 331.Uezumi A, Fukada S, Yamamoto N, Takeda S, Tsuchida K. Mesenchymal progenitors distinct from satellite cells contribute to ectopic fat cell formation in skeletal muscle. Nat Cell Biol 12: 143‐152, 2010.
 332.Urakawa I, Yamazaki Y, Shimada T, Iijima K, Hasegawa H, Okawa K, Fujita T, Fukumoto S, Yamashita T. Klotho converts canonical FGF receptor into a specific receptor for FGF23. Nature 444: 770‐774, 2006.
 333.Urao N, Mirza RE, Heydemann A, Garcia J, Koh TJ. Thrombospondin‐1 levels correlate with macrophage activity and disease progression in dysferlin deficient mice. Neuromuscul Disord 26: 240‐251, 2016.
 334.Vallejo AN, Mugge L, Klimiuk PA, Weyand CM, Goronzy JJ. Central role of thrombospondin‐1 in the activation and clonal expansion of inflammatory T cells. J Immunol 164: 2947‐2954, 2000.
 335.van der Pas J, Hengstman GJ, ter Laak HJ, Borm GF, van Engelen BG. Diagnostic value of MHC class I staining in idiopathic inflammatory myopathies. J Neurol Neurosurg Psychiatry 75: 136‐139, 2004.
 336.van der Pijl EM, van Putten M, Niks EH, Verschuuren JJGM, Aartsma‐Rus A, Plomp JJ. Characterization of neuromuscular synapse function abnormalities in multiple Duchenne muscular dystrophy mouse models. Eur J Neurosci 43: 1623‐1635, 2016.
 337.van Dijk S, D'Errico JA, Somerman MJ, Farach‐Carson MC, Butler WT. Evidence that a non‐RGD domain in rat osteopontin is involved in cell attachment. J Bone Miner Res 8: 1499‐1506, 1993
 338.van Reeuwijk J, Brunner HG, Van Bokhoven H. Glyc‐O‐genetics of Walker‐Warburg syndrome. Clin Genet 67: 281‐289, 2005.
 339.van Reeuwijk J, Grewal PK, Salih MA, Beltran‐Valero de Bernabe D, McLaughlan JM, Michielse CB, Herrmann R, Hewitt JE, Steinbrecher A, Seidahmed MZ, Shaheed MM, Abomelha A, Brunner HG, van Bokhoven H, Voit T. Intragenic deletion in the LARGE gene causes Walker‐Warburg syndrome. Hum Genet 121: 685‐690, 2007.
 340.Vetrone SA, Montecino‐Rodriguez E, Kudryashova E, Kramerova I, Hoffman EP, Liu SD, Miceli MC, Spencer MJ. Osteopontin promotes fibrosis in dystrophic mouse muscle by modulating immune cell subsets and intramuscular TGF‐beta. J Clin Invest 119: 1583‐1594, 2009.
 341.Vidal B, Serrano AL, Tjwa M, Suelves M, Ardite E, De Mori R, Baeza‐Raja B, Martínez de Lagran M, Lafuste P, Ruiz‐Bonilla V, Jardí M, Gherardi R, Christov C, Dierssen M, Carmeliet P, Degen JL, Dewerchin M, Munoz‐Canoves P. Fibrinogen drives dystrophic muscle fibrosis via a TGFbeta/alternative macrophage activation pathway. Genes Dev 22: 1747‐1752, 2008.
 342.Villalta SA, Nguyen HX, Deng B, Gotoh T, Tidball JG. Shifts in macrophage phenotypes and macrophage competition for arginine metabolism affect the severity of muscle pathology in muscular dystrophy. Hum Mol Genet 18: 482‐496, 2009.
 343.Villalta SA, Rinaldi C, Deng B, Liu G, Fedor B, Tidball JG. Interleukin‐10 reduces the pathology of mdx muscular dystrophy by deactivating M1 macrophages and modulating macrophage phenotype. Hum Molec Genet 20: 790‐805, 2011.
 344.Villalta SA, Rosenthal W, Martinez L, Kaur A, Sparwasser T, Tidball JG, Margeta M, Spencer MJ, Bluestone JA. Regulatory T cells suppress muscle inflammation and injury in muscular dystrophy. Sci Transl Med 6: 258ra142, 2014. 10.1126/scitranslmed.3009925.
 345.Wallace LM, Garwick SE, Mei W, Belayew A, Coppee F, Ladner KJ, Guttridge D, Yang J, Harper SQ. DUX4, a candidate gene for facioscapulohumeral muscular dystrophy, causes p53‐dependent myopathy in vivo. Ann Neurol 69: 540‐552, 2011.
 346.Walter MC, Reilich P, Thiele S, Schessl J, Schreiber H, Reiners K, Kress W, Muller‐Reible C, Vorgerd M, Urban P, Schrank B, Deschauer M, Schlotter‐Weigel B, Kohnen R, Lockmuller H. Treatment of dysferlinopathy with deflazacort: a double‐blind, placebo‐controlled clinical trial. Orphanet J Rare Dis 8: 26, 2013. 10.1186/1750‐1172‐8‐26.
 347.Wang X, Zhao W, Ransohoff RM, Zhou L. Identification and function of fibrocytes in skeletal muscle injury repair and muscular dystrophy. J Immunol 197: 4750‐4761, 2016.
 348.Wardrop KE, Dominov JA. Proinflammatory signals and the loss of lymphatic vessel hyaluronan receptor‐1 (LYVE‐1) in the early pathogenesis of laminin alpha2‐deficient skeletal muscle. J Histochem Cytochem 59: 167‐179, 2011.
 349.Weber GF, Ashkar S, Glimcher MJ, Cantor H. Receptor‐ligand interaction between CD44 and osteopontin (Eta‐1). Science 271: 509‐512, 1996.
 350.Webster C, Blau HM. Accelerated age‐related decline in replicative life‐span of Duchenne muscular dystrophy myoblasts: Implications for cell and gene therapy. Somat Cell Mol Genet 16: 557‐565, 1990.
 351.Wehling M, Spencer MJ, Tidball JG. A nitric oxide synthase transgene ameliorates muscular dystrophy in mdx mice. J Cell Biol 155: 123‐131, 2001.
 352.Wehling‐Henricks M, Jordan MC, Gotoh T, Grody WW, Roos KP, Tidball JG. Arginine metabolism by macrophages promotes cardiac and muscle fibrosis in mdx muscular dystrophy. PLoS One 5: e10763, 2010. 10.1371/journal.pone.0010763.
 353.Wehling‐Henricks M, Jordan MC, Roos KP, Deng B, Tidball JG. Cardiomyopathy in dystrophin‐deficient hearts is prevented by expression of a neuronal nitric oxide synthase transgene in the myocardium. Hum Mol Genet 14: 1921‐1933, 2005.
 354.Wehling‐Henricks M, Lee JJ, Tidball JG. Prednisolone decreases cellular adhesion molecules required for inflammatory cell infiltration in dystrophin‐deficient skeletal muscle. Neuromuscul Disord 14: 483‐490, 2004.
 355.Wehling‐Henricks M, Li Z, Lindsey C, Wang Y, Welc SS, Ramos JN, Khanlou N, Kuro‐O M, Tidball JG. Klotho gene silencing promotes pathology in the mdx mouse model of Duchenne muscular dystrophy. Hum Mol Genet 25: 2465‐2482, 2016.
 356.Wehling‐Henricks M, Oltmann M, Rinaldi C, Myung KH, Tidball JG. Loss of positive allosteric interactions between neuronal nitric oxide synthase and phosphofructokinase contributes to defects in glycolysis and increased fatigability in muscular dystrophy. Hum Mol Genet 18: 3439‐3451, 2009.
 357.Wehling‐Henricks M, Sokolow S, Lee JJ, Myung KH, Villalta SA, Tidball JG. Major basic protein‐1 promotes fibrosis of dystrophic muscle and attenuates the cellular immune response in muscular dystrophy. Hum Mol Genet 17: 2280‐2292, 2008.
 358.Wehling‐Henricks M, Welc S, Samengo G, Rinaldi C, Lindsey C, Wang Y, Lee J, Kuro‐o M, Tidball JG. Macrophages escape Klotho gene silencing in the mdx mouse model of Duchenne muscular dystrophy and promote muscle growth and increase satellite cell numbers through a Klotho‐mediated pathway. Hum Mol Genet 27: 14‐29, 2018. 10.1093/hmg/ddx380.
 359.Weiner FR, Flanders KC, Giambrone MA, Wind R, Biempica L, Zern MA. In vitro and in vivo association of transforming growth factor‐beta 1 with hepatic fibrosis. J Cell Biol 108: 2477‐2482, 1989.
 360.Wenzel K, Zabojszcza J, Carl M, Taubert S, Lass A, Harris CL, Ho M, Schulz H, Hummel O, Hubner N, Osterziel KJ, Spuler S. Increased susceptibility to complement attack due to down‐regulation of decay‐accelerating factor/CD55 in dysferlin‐deficient muscular dystrophy. J Immunol 175: 6219‐6225, 2005.
 361.Willcocks RJ, Triplett WT, Forbes SC, Arora H, Senesac CR, Lott DJ, Nicholson TR, Rooney WD, Walter GA, Vandenborne K. Magnetic resonance imaging of the proximal upper extremity musculature in boys with Duchenne muscular dystrophy. J Neurol 264: 64‐71, 2017.
 362.Williamson RA, Henry MD, Daniels KJ, Hrstka RF, Lee JC, Sunada Y, Ibraghimov‐Beskrovnaya O, Campbell KP. Dystroglycan is essential for early embryonic development: Disruption of Reichert's membrane in Dag1‐null mice. Hum Mol Genet 6: 831‐841, 1997.
 363.Wiseman DM, Polverini PJ, Kamp DW, Leibovich SJ. Transforming growth factor‐beta (TGF beta) is chemotactic for human monocytes and induces their expression of angiogenic activity. Biochem Biophys Res Commun 157: 793‐800, 1988.
 364.Witte MB, Barbul A. Arginine physiology and its implication for wound healing. Wound Repair Regen 11: 419‐423, 2003.
 365.Witte MB, Vogt N, Stuelten C, Gotoh T, Mori M, Becker HD. Arginase acts as an alternative pathway of L‐arginine metabolism in experimental colon anastomosis. J Gastrointest Surg 7: 378‐385, 2003.
 366.Xu H, Wu XR, Wewer UM, Engvall E. Murine muscular dystrophy caused by a mutation in the laminin alpha 2 (Lama2) gene. Nat Genet 8: 297‐302, 1994.
 367.Yamaguchi Y, Hayashi Y, Sugama Y, Miura Y, Kasahara T, Kitamura S, Torisu M, Mita S, Tominaga A, Takatsu A. Highly purified murine interleukin‐5 stimulates eosinophil function and prolongs survival. IL‐5 as an eosinophil chemotactic factor. J Exp Med 167: 1737‐1742, 1988.
 368.Yamamoto M, Clark JD, Pastor JV, Gurnani P, Nandi A, Kurosu H, Miyoshi M, Ogawa Y, Castrillon DH, Rosenblatt KP, Kuro‐o M. Regulation of oxidative stress by the anti‐aging hormone klotho. J Biol Chem 280: 38029‐38034, 2005.
 369.Yamamoto S, Nasu K, Ishida T, Setoguchi M, Higuchi Y, Hijiya N, Akizuki S. Effect of recombinant osteopotin on adhesion and migration of P388D1 cells. Ann N Y Acad Sci 760: 378‐380, 1995
 370.Yamazaki M, Minota S, Sakurai H, Miyazono K, Yamada A, Kanazawa I, Kawai M. Expression of transforming growth factor‐beta 1 and its relation to endomysial fibrosis in progressive muscular dystrophy. Am J Pathol 144: 221‐226, 1994.
 371.Yilmaz O, Karaduman A, Topaloglu H. Prednisolone therapy in Duchenne muscular dystrophy prolongs ambulation and prevents scoliosis. Eur J Neurol 11: 541‐544, 2004.
 372.Yin X, Wang Q, Chen T, Niu J, Ban R, Liu J, Mao Y, Pu C. CD4+ cells, macrophages, MHC‐I and C5b‐9 involve the pathogenesis of dysferlinopathy. Int J Clin Exp Pathol 8: 3069‐3075, 2015.
 373.Yu M, Wang H, Ding A, Golenbock DT, Latz E, Czura CJ, Fenton MJ, Tracey KJ, Yang H. HMGB1 signals through toll‐like receptor (TLR) 4 and TLR2. Shock 26: 174‐179, 2006.
 374.Yu Q, Stamenkovic I. Cell surface‐localized matrix metalloproteinase‐9 proteolytically activates TGF‐beta and promotes tumor invasion and angiogenesis. Genes Dev 14: 163‐176, 2000.
 375.Zanotti S, Gibertini S, Di Blasi C, Cappelletti C, Bernasconi P, Mantegazza R, Morandi L, Mora M. Osteopontin is highly expressed in severely dystrophic muscle and seems to play a role in muscle regeneration and fibrosis. Histopathology 59: 1215‐1228, 2011.
 376.Zhang J, Xiao Z, Qu C, Cui W, Wang X, Du J. CD8 T cells are involved in skeletal muscle regeneration through facilitating MCP 1 secretion and GR1 high macrophage infiltration. J Immunol 193: 5149‐5160, 2014.
 377.Zhang T, Somasundaram R, Berencsi K, Caputo L, Gimotty P, Rani P, Guerry D, Swoboda R, Herlyn D. Migration of cytotoxic T lymphocytes toward melanoma cells in three‐dimensional organotypic culture is dependent on CCL2 and CCR4. Eur J Immunol 36: 457‐467, 2006.
 378.Zhao J, Freeman GJ, Gray GS, Nadler LM, Glimcher LH. A cell type‐specific enhancer in the human B7.1 gene regulated by NF‐kappaB. J Exp Med 183: 777‐789, 1996.
 379.Zhou L, Porter JD, Cheng G, Gong B, Hatala DA, Merriam AP, Zhou X, Rafael JA, Kaminski HJ. Temporal and spatial mRNA expression patterns of TGF‐beta1, 2, 3 and TbetaRI, II, III in skeletal muscles of mdx mice. Neuromuscul Disord 16: 32‐38, 2006.
 380.Zhu B, Suzuki K, Goldberg HA, Rittling, SR, Denhardt DT, McCulloch CA, Sodek J. Osteopontin modulates CD44‐dependent chemotaxis of peritoneal macrophages through G‐protein‐coupled receptors: Evidence of a role for an intracellular form of osteopontin. J Cell Physiol 198: 155‐167, 2004.

 

Teaching Material

J. G. Tidball, S. S. Welc, M. Wehling-Henricks. Immunobiology of Inherited Muscular Dystrophies. Compr Physiol 8: 2018, 1313-1356.

Didactic Synopsis

Major Teaching Points:

  1. Inherited muscular dystrophies arise from diverse mutations that lead to pathologies that have the shared characteristic of muscle wasting.
  2. Many of muscular dystrophies, especially the most commonly occurring muscular dystrophy called Duchenne muscular dystrophy (DMD), cause a more easily damaged cell membrane that contributes to muscle death.
  3. Damage muscle fibers in DMD and other muscular dystrophies can produce an immune response that can amplify muscle pathology.
  4. The initial immune response to DMD muscle entails an ancient response called innate immunity, which is adaptive for acute injuries but is maladaptive for chronic muscle damage that occurrence over the lifetime of the affected individual.
  5. Perturbations in the expression or activity of endogenous immunomodulators can significantly influence interactions between muscle and the immune system that are specific to different muscular dystrophies.
  6. The immune response to dystrophic muscle extends beyond innate immunity in which myeloid cells are the primary effector population, to include components of the acquired immune system, in which the actions of lymphoid cells are of central importance.

Didactic Legends

The figures—in a freely downloadable PowerPoint format—can be found on the Images tab along with the formal legends published in the article. The following legends to the same figures are written to be useful for teaching.

Figure 1 Teaching points: In Duchenne muscular dystrophy (DMD), progressive muscle wasting leads to musculoskeletal defects that are apparent on the gross level in young boys, in whom muscle tissue is replaced by fat and fibrotic connective tissue and invading immune cells.

Figure 2 Teaching points: DMD is caused by mutation of the dystrophin gene, leading to loss of the membrane-associated protein, dystrophin. Loss of dystrophin leads to a secondary reduction in the quantity of proteins in the dystrophin glycoprotein complex, which contributes to further pathological defects.

Figure 3 Teaching points: Loss of dystrophin produces a weakened cell membrane, which can cause membrane tears large enough to see by electron microscopy and large enough to allow large molecules of dye to enter the muscle fibers. This causes huge disruptions in cellular homeostasis and can eventually lead to cell death.

Figure 4 Teaching points: In the mdx mouse model of DMD, large numbers of macrophages invade the dystrophic muscle. If macrophage invasion is prevented or greatly reduced, muscle fiber damage can be reduced by as much as 80%,

Figure 5 Teaching points: Macrophages are a diverse population of leukocytes that vary tremendously in the regulatory roles they play in injured and diseased tissue. At opposite ends of the broad range of phenotypes are macrophages that are biased toward the M1 or the M2 ends of the spectrum of macrophage phenotypes, which differ in the pathways through which they are activated, the immunomodulatory molecules they produce and the processes they mediate in diseased and injured tissues.

Figure 6 Teaching points: Several enzymes compete for the substrate arginine in injured or rapidly growing tissue. In dystrophic muscle, iNOS in M1-biased macrophages, arginase in M2-biased macrophages, and nNOS in muscle fibers can all compete for arginine. Increased arginine availability for iNOS can increase muscle fiber lysis and increased arginine availability arginase can increase muscle fibrosis.

Figure 7 Teaching points: Complex interactions between populations of immune cells and muscle fibers in dystrophic muscle influence the extent of muscle fiber damage and fibrosis.

Figure 8 Teaching points: Eosinophils in dystrophic muscle can increase muscle fibrosis, cause muscle fiber damage, and modulate the immune response to dystrophic muscle.

Figure 9 Teaching points: Cytotoxic T-lymphocytes increase death of dystrophin-deficient muscles. They are potentially activated to a cytolytic state by antigens that are presented to them by the dystrophic muscle fibers in association with MHC class I molecules.

Figure 10 Teaching points: Two important immunomodulatory molecules, TNF and HMGB1, can activate signaling pathways in dystrophic muscle that lead to increased expression of molecules that amplify the immune response and that are necessary for the presentation of antigen to immune cells to contribute to their activation.

Figure 11 Teaching points: The regeneration of skeletal muscle following injury or disease relies on the activation of a population of quiescent, muscle stem cells, called satellite cells. Following activation, satellite cells can either return to the quiescent stem cell pool or they can undergo terminal differentiation to become muscle fibers. The pathway for myogenic differentiation is regulated by a series of myogenic transcription factors, including Pax7, MyoD, and myogenin (MyoG).

Figure 12 Teaching points: The process of satellite cell activation, proliferation, and differentiation that is required for regeneration of dystrophin-deficient muscle is regulated by the immune system. Cytokines and growth factors released by macrophages, neutrophils CD8 + cells, and eosinophils modulate muscle regeneration by direct actions on myogenic cells or by influencing the activities of other immune cells or fibrogenic cells.

Figure 13 Teaching points: Regulatory T-lymphocytes (Tregs) can promote repair of dystrophin-deficient muscle by attenuating the activation of cytolytic macrophages and shifting macrophages toward a population that supports muscle repair.

Figure 14 Teaching points: Muscle fibrosis that results from dystrophin mutation can cause musculoskeletal deformities that are caused by contractures of skeletal muscle. Fibrosis of the myocardium also results from dystrophin-deficiency and cardiac functional defects that are caused by fibrosis are a major cause of death in DMD.

Figure 15 Teaching points: Muscle wasting in LGMD1B can be accompanied by extensive immune cell involvement that indicates an acquired immune response. Unlike DMD muscle LGMD1B shows elevated numbers of B-lymphocytes, suggesting that LGMD1B pathology involves a humoral immune response that does not occur in DMD.

Figure 16 Teaching points: Muscle wasting and inflammation are prominent components of the pathology in LGMD2A. LGMD2A is frequently associated with “winged scapulae” attributable to weakening and loss of muscles of the shoulder girdle. Muscle histology shows that muscle fibers in LGMD2A can be replaced by vast numbers of inflammatory cells, that include elevated numbers of eosinophils and increased deposition of fatty tissue.

Figure 17 Teaching points: Muscles of β-sarcoglycan null mice, a model for LGMD2E, experience extensive muscle fibrosis and an inflammatory response that is dominated by macrophages.

Figure 18 Teaching points: LGMD2B is characterized by wasting of proximal and distal, lower limb musculature. Affected muscles of LGMD2B patients show extensive replacement of musculature with fat and connective tissue.

Figure 19 Teaching points: The immune cell infiltrate in LGMD2B muscles shows large numbers of T-cells, many of which are cytotoxic T-cells, and numerous macrophages. In addition, MHC class 1 distributed at the surfaces of LGMD2B muscle fibers suggesting antigen presentation by muscle fibers.

Figure 20 Teaching points: MDC1A patients show a characteristic weakness of facial muscles and elbow flexions that result from increased fibrosis. Increased lumbar lordosis and knee flexions are also attributable to fibrotic contractures. Muscles of MDC1A patients can show extensive atrophy, fatty infiltration, edema, and inflammation.

Figure 21 Teaching points: This FSHD patient shows asymmetric muscle wasting on right side of body, including muscles of shoulder girdle, humeral, and lower limb muscle. The extensive muscle loss was replaced by fat, that is particularly prominent in arm-level and crus-level images. Muscle biopsies show large numbers of CD8 + cytotoxic T-cells that surround muscle fibers.

 


Related Articles:

The Dystrophin Complex: Structure, Function, and Implications for Therapy
Satellite Cells and Skeletal Muscle Regeneration
Mechanisms of Muscle Injury, Repair, and Regeneration
Reactive Oxygen Species: Impact on Skeletal Muscle
Diseases of Skeletal Muscle
Top downloaded articles of 2019

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

James G. Tidball, Steven S. Welc, Michelle Wehling‐Henricks. Immunobiology of Inherited Muscular Dystrophies. Compr Physiol 2018, 8: 1313-1356. doi: 10.1002/cphy.c170052