Next Article in Journal
Cisd2 Protects the Liver from Oxidative Stress and Ameliorates Western Diet-Induced Nonalcoholic Fatty Liver Disease
Next Article in Special Issue
Skeletal Lipocalin-2 Is Associated with Iron-Related Oxidative Stress in ob/ob Mice with Sarcopenia
Previous Article in Journal
The Impact of a Plant-Based Diet on Gestational Diabetes: A Review
Previous Article in Special Issue
To Synthesize Hydroxyapatite by Modified Low Temperature Method Loaded with Bletilla striata Polysaccharide as Antioxidant for the Prevention of Sarcopenia by Intramuscular Administration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Beneficial Role of Exercise in the Modulation of mdx Muscle Plastic Remodeling and Oxidative Stress

by
Monica Frinchi
1,
Giuseppe Morici
1,2,
Giuseppa Mudó
1,*,
Maria R. Bonsignore
2,3 and
Valentina Di Liberto
1,*
1
Department of Biomedicine, Neuroscience and Advanced Diagnostic (BIND), University of Palermo, 90134 Palermo, Italy
2
Institute for Biomedical Research and Innovation, National Research Council, 90146 Palermo, Italy
3
Department of Health Promotion Sciences Maternal and Infantile Care, Internal Medicine and Medical Specialties (PROMISE), University of Palermo, 90127 Palermo, Italy
*
Authors to whom correspondence should be addressed.
Antioxidants 2021, 10(4), 558; https://doi.org/10.3390/antiox10040558
Submission received: 28 January 2021 / Revised: 2 March 2021 / Accepted: 15 March 2021 / Published: 3 April 2021
(This article belongs to the Special Issue Oxidative Stress in Skeletal Muscle)

Abstract

:
Duchenne muscular dystrophy (DMD) is an X-linked recessive progressive lethal disorder caused by the lack of dystrophin, which determines myofibers mechanical instability, oxidative stress, inflammation, and susceptibility to contraction-induced injuries. Unfortunately, at present, there is no efficient therapy for DMD. Beyond several promising gene- and stem cells-based strategies under investigation, physical activity may represent a valid noninvasive therapeutic approach to slow down the progression of the pathology. However, ethical issues, the limited number of studies in humans and the lack of consistency of the investigated training interventions generate loss of consensus regarding their efficacy, leaving exercise prescription still questionable. By an accurate analysis of data about the effects of different protocol of exercise on muscles of mdx mice, the most widely-used pre-clinical model for DMD research, we found that low intensity exercise, especially in the form of low speed treadmill running, likely represents the most suitable exercise modality associated to beneficial effects on mdx muscle. This protocol of training reduces muscle oxidative stress, inflammation, and fibrosis process, and enhances muscle functionality, muscle regeneration, and hypertrophy. These conclusions can guide the design of appropriate studies on human, thereby providing new insights to translational therapeutic application of exercise to DMD patients.

Graphical Abstract

1. Introduction

This review examines the role of exercise in the modulation of muscle plasticity and oxidative stress in the mdx mouse, the most widely used pre-clinical animal model for Duchenne muscular dystrophy (DMD), a fatal X-linked disorders characterized by progressive muscle weakness. Beyond the current medication, such as steroids, which show many side effects and can only slow down disease progression, and new gene- and stem cells-based strategies, still under experimentation [1,2], physical activity may represent an effective noninvasive therapeutic approach for DMD. However, if on one hand a moderate intensity exercise is usually associated to beneficial effects in healthy muscles [3], on the other hand the limited number of investigations in DMD patients and the lack of consensus regarding the fine balance between benefits of training and DMD muscle overuse and damage, impede the generation of definitive guidelines for exercise prescription [4,5].
It is well known that healthy skeletal muscle displays the ability to easily adapt to environmental stimuli, such as exercise and physical activity, through a plastic remodeling of the phenotype, strictly dependent on different properties of the applied stimuli. This adaptive remodeling, which plays an essential role in improving neuromuscular performance and/or enhancing endurance capabilities, includes metabolism modifications, structural changes, such as muscle pennation and fiber type switch, as well as size changes due to phenomena of hypertrophy [3,6,7]. The effects of exercise on muscle tissue affected by chronic disorders, such as DMD, have been extensively studied and reviewed [4,5,8,9,10,11,12]. However, since the beneficial effects of physical exercise depend on many parameters, such as type, duration, frequency, and intensity of training, and considering the large variation among individuals with regard to the magnitude of muscle remodeling in response to exercise, the lack of uniform and standardized protocols of training impairs comparisons between studies and translation of results obtained in animals to patients.
Therefore, in order to compare and shed some light on the many existing protocols of exercise, with the aim of identifying the best suited modality of training associated to beneficial effect on DMD muscle, here we present an updated data report about the effects of different protocols of physical exercise on muscles of mdx mice (Figure 1). The main modifications induced by exercise on mdx muscles, including the mechanism of fibrosis, the ability of satellite cells to regenerate damaged muscle tissue, the modifications of muscle performance (in term of endurance and strength) and structure (fiber type and muscle volume), as well as the metabolic processes associated with the control of fibers redox state and inflammation, will be reviewed in the next paragraphs. Furthermore, a difference in the functional response of mdx skeletal, cardiac and diaphragm muscles will be discussed, taking into considerations that the three types of muscle are not equally affected by the absence of dystrophin and exercise [13].
At the end of this discussion, we propose an optimized exercise protocol able to induce beneficial adaptive remodeling in mdx muscles, thereby providing new insights to translational therapeutic application of exercise to DMD patients.

2. Duchenne Muscular Dystrophy

DMD is an X-linked recessive progressive lethal disorder with a worldwide incidence of one in 5000 live male births [14]. It is caused by the lack of dystrophin, a critical muscle protein that connects the cytoskeleton of muscle fibers to the extracellular matrix, acting as an essential stabilizer of muscle fibers during contraction. Deletion of dystrophin, in both mature muscle fibers and myogenic stem cells, results in myofibers mechanical instability and susceptibility to contraction-induced injuries [15,16], as well as weakness, muscle loss [10], oxidative stress and inflammation [17,18]. The clinical course of the pathology is progressive and associated to life-long debilitation and shorter longevity: Patients with DMD first develop skeletal muscle weakness in early childhood, which quickly progresses to loss of muscle tissue, spinal curvature, paralysis, and premature death because of cardiorespiratory failure, usually in the third or fourth decade of life [1,14].
Muscle mechanical instability in DMD is associated with degeneration and regeneration of myofibers and activation of satellite muscle stem cells. Indeed, muscle injury activates satellite cells, which start to proliferate and to differentiate, leading, through a multistep process, to the formation of new regenerating fibers with centrally located nuclei, which represent a distinctive histological marker of DMD muscle [19,20]. The chronic nature of the pathology is associated at later stages to development of chronic inflammation, increased oxidative stress, inhibition of muscle fiber regeneration, depletion of the satellite cells pool and replacement of muscle by fibrotic and adipose tissue, and generating weakness in the diaphragm and limb muscles [21,22,23]. Importantly, dystrophic muscle fibers are more vulnerable to exercise-induced damage [24,25].
Unfortunately, at present, there is no efficient therapy for DMD. The classically available treatments, such as steroids, mostly interfere with inflammatory processes related to the pathology, thus reducing the immunological responses involved in the progression of the disease. However, these treatments show many adverse effects and can only slow down disease progression [1,26,27,28]. Indeed, research is still ongoing, and important developments have been achieved in the field, with the discovery of emerging therapies that are in the clinical trial phase or have already been approved. Briefly, gene-addition, exon-skipping, stop codon readthrough, and genome-editing-based therapies can restore the expression and the function of dystrophin, while stem cells-based therapy can partially replace damaged muscle tissue. Finally, other therapeutic approaches can improve muscle functionality by targeting pathways involved in the pathogenesis of DMD, such as inflammation and oxidative stress [1,2,29,30]. Although genome-editing based- and stem cells based-approaches have the potential to lead to meaningful life-changing therapeutic interventions, they show some limitations, especially in terms of cost burden and accessibility, and their therapeutic impact will be fully understood only in the next decades. Furthermore, genetic therapy for DMD patients remains an issue, taking into consideration the differences in mutations in the DMD gene and the complex mechanisms of gene expression regulation [31]. In addition to pharmacological, gene- and stem cells-based therapies, physical exercise, by inducing muscle plastic remodeling, represents a potential therapeutic approach for improving DMD patient outcomes and quality of life [5,9,32].

3. The mdx Model

The mdx mouse represents the most widely used pre-clinical animal model for DMD research [33,34,35]. mdx mouse was discovered in the early 1980s in a colony of C57BL/10ScSn, in which a spontaneous nonsense point mutation, consisting in a C-to-T transition at exon 23, determined loss of dystrophin, elevated serum creatine kinase (CK) and muscle damage [36]. Specifically, mdx skeletal muscle exhibits elevated myofiber necrosis, cellular infiltration, a wide range of myofiber sizes and several centrally nucleated regenerating myofibers. This phenotype is particularly evident in the diaphragm, which presents progressive degeneration, fibrosis and myofiber loss, resulting in significant strength reduction, thus closely reproducing the degenerative changes found in DMD muscles [33,34,35,37]. However, in general, although both mdx and DMD patients are genetically homologous and united by a complete absence of dystrophin, loss of dystrophin in mdx mice leads to a less severe phenotype, with minimal clinical symptoms. Indeed, the lifespan of mdx mice is only reduced by ~25% as compared to wild type (wt) animals, probably due to the activation of compensatory mechanisms, and/or to a species-specific feature of the muscle [38,39] and/or to the influence of the animal house environment, which hinders active movement and may therefore protect muscle from exercise-induced damage [40]. This experience has driven the development of new mouse models showing increased severity of the phenotype which better recapitulates the disease [33]. Nonetheless, the mdx mouse is still by far the most widely used animal model for DMD.
Several distinctive phases can be recognized in the pathology development in mdx mice [34,41]. In the first 2 weeks, even if higher mortality in mdx litters has been described, mdx muscle is still close to that of wt mice [42]. Muscle pathology becomes more pronounced between 2 and 8 weeks of age, when a clear presence of necrotic areas, newly regenerated centrally nucleated myofibers and high plasma concentrations of CK can be observed. In this phase, chronic inflammation, as evidenced by infiltration of inflammatory cells and high levels of inflammatory cytokines, can be also detected, together with fibrosis development [43,44,45]. However, the severity of fibrosis and loss of function in limb muscles is less in mdx mice than in human [46,47,48]. Severe dystrophic phenotypes, such as muscle loss, scoliosis, and heart failure, become evident in mice 15 months old or older, when a second robust phase of necrosis is not compensated by a regeneration process [49,50,51,52,53,54].

4. Effects of Exercise on Plastic Remodeling of mdx Muscle

Healthy striated muscles are characterized by a remarkable plasticity, as they can adjust the metabolic and contractile status in response to changes in functional demands. This adaptive remodeling, which plays an essential role in improving muscle performance, depends on the properties of the stimulus, especially in terms of intensity and duration of training [6,55]. Exercise intensity is classified accordingly to the physical exertion that the body uses when performing the activity [56] and is typically measured in metabolic equivalent task (MET), defined as the rate of energy expenditure at rest [57]. Exercise Activities with METs below 3 are considered as low intensity, with METs between 3 and 6 as moderate intensity, and with METs higher than 6 as high intensity [58].
Aerobic endurance exercise, characterized by repeated, sustained, low intensity contractions (i.e., long distance running, cycling and swimming) is generally associated to the switch of muscle fibers towards the type I, slow-twitch and fatigue-resistant [59], while resistance training, whose intensity is above 75% of the maximal capacity, characterized by low-frequency, high intensity contractions (i.e., body building and weight lifting), usually determines, accordingly to the protocol used, an increase in the proportion of type II fast-twitch myofibers and in force generation via muscle hypertrophy [55,60,61,62]. Generally, exercise provides numerous beneficial effects on skeletal muscle. However, since the effects of exercise engage several molecular and metabolic players, whose pattern of activation depends on the modalities of training, exercise might be also associated to detrimental effects in vulnerable DMD muscles, especially in terms of oxidative stress generation and fiber necrosis [3].
Similar to wt muscles, plastic remodeling of mdx muscle in response to exercise depends on many variables, including modality, intensity and duration of training. In rodent models, in order to study physiological adaptation associated with exercise, treadmill, swimming and wheel running exercise modalities have become quite popular [63]. While wheel running rely upon voluntary exercise, treadmill running and swimming can be controlled according to standardized protocols, especially in terms of duration and intensity set by the operator, thus allowing the investigator to track changes in the muscle response in relation to exercise parameters’ modifications. The adaptive remodeling induced by different typology of exercise in mdx muscles will be discussed in the next paragraphs.

4.1. Effects of Forced Running on Plastic Remodeling of mdx Muscle

Treadmill running represents the most common and effective modality of training used to investigate the effects of exercise on mouse muscles. Treadmill exercise intensity, expressed as the speed of running, can be set to low (<12 m/min), moderate (12–15 m/min), or high (>15 m/min) levels. Regarding exercise duration, although mice can run continuously for up to 2 h at a time, most studies used a 30–60 min/day training protocol, with a frequency ranging between 2 and 7 days/week, and a total duration lasting up to 24 weeks. When treadmill running is performed on an inclined platform, according to the negative or positive slope, it is referred as downhill or uphill running, respectively. Uphill running determines concentric muscle contraction and increases cardiovascular response, while downhill running represents a physiologically relevant model of eccentric loading associated to fiber membrane damage and subsequent metabolic adaptations [64,65]. Both training are associated muscle-specific increased workload as compared with running on a horizontal treadmill [8,66,67,68,69].
The main plastic remodeling induced by horizontal treadmill and downhill/uphill running exercise in mdx muscles have been summarized in Appendix A Table A1 and Appendix A Table A2, respectively. This remodeling modulates muscle performance (in term of endurance and strength) and structure (fiber type and muscle volume), the mechanism of fibrosis, the degeneration/regeneration process, muscle metabolism together with muscle redox and inflammatory state. These last two parameters will be reviewed later.
Fibrosis, referred as the process of aberrant replacement of normal muscle tissue with connective tissue as response of reactive or reparative process, is a prominent pathological feature of skeletal muscle in patients with DMD, associated with severe muscle wasting and impairment of the muscle contractile functions and muscle regeneration ability [22,70,71,72]. Several studies have shown that treadmill exercise can efficiently modulate this process. Indeed, a low intensity exercise, at a speed of 9 m/min for two months, on a horizontal platform reduces collagen deposition in limb muscles of mdx mice [73,74,75]. In addition, low intensity running induced a strong beneficial effect on the degeneration-regeneration process of mdx muscle and a shift in favor of regeneration, both in limb muscles and in diaphragm [76,77], and reduced the percentage of necrotic area in mdx soleus and gastrocnemius muscles, while it seemed to exert opposite effects on mdx plantaris muscle [78]. These structural improvements are associated to the enhancement of muscle performances, such as increased grip strength, tetanic and specific force and resistance to fatigue [79].
On the contrary, higher intensity exercise, in the form of treadmill running at a speed equal or greater than 12 m/min or downhill/uphill running, independently of training duration, mostly enhanced collagen deposition and fibrosis area [15,16,80,81,82,83], and determined a series of detrimental effects on mdx muscles. Indeed, this type of exercise aberrantly up-regulated the phosphorylated form of extracellular signal-regulated kinase 1/2 (ERK1/2), p38 mitogen-activated protein kinases (p38 MAPK) and c-Jun N-terminal kinases 2 (JNK2), which might play a key role in the degeneration and regeneration process of mdx muscles [84]. The same type of exercise enhanced Ca2+ influx and sarcolemmal permeability [25,85], induced an increase in muscle injury/necrosis area [82,83,86,87,88,89,90,91,92,93,94], and promoted sarcoplasm fragmentation, oxidative stress and muscle inflammation [15,82,87,92]. Interestingly, a proteomic analysis revealed that 12 m/min treadmill running for 4 weeks failed to stimulate the metabolic changes associated to fast-to-slow transition, usually observed in aerobically trained muscles, decreased the expression of myosin regulatory light chain 2 and enhanced specific protein degradation, thus further reducing the amount of sarcomeres’ proteins in mdx mice [95]. All these molecular and histological changes are associated to deficit in muscle performances, such as the increase in fatigability and the reduction of forelimb strength [82,83,85,87,88,90,92,96,97,98,99], twitch and tetanic force [86]. In disagreement with all the other data about the effects of running on an inclined platform on muscle performance, 3 weeks of uphill running at a speed of 4 m/min initiated in three weeks old mdx mice increased soleus twitch tension and decreased muscle necrosis [100], likely due to the combination of low speed running and age of exercise initiation.
The equilibrium between degeneration and regeneration depends not only on the deposition of fibrotic tissue, but also on muscle stem cells functionality. Indeed, satellite cells impairment and, in particular, exhaustion and loss of stem cell properties seem to promote failure of the regeneration process and the subsequent degeneration overcoming [101,102]. Accordingly, stem cell-based therapies can represent a promising tool to induce DMD muscle regeneration [103,104]. Treadmill running seems to influence satellite cells properties of mdx muscles. Indeed, low speed treadmill running is associated to the increase in the regeneration area and to the reduction of connexin 39 (Cx39), a specific marker of injured muscle, in hind limb muscles [77], thus indirectly suggesting a beneficial effect of exercise on muscle stem cells functionality. Accordingly, 4 weeks of noninjurious isometric strength training, improved mdx phenotype and muscle performance, by inducing fiber hypertrophy, reducing fibrosis and increasing the number of satellite cells [105]. Furthermore, it has been demonstrated that treadmill running induced a telomere shortening in limb muscles in wt, but not in mdx mice, thereby suggesting that exercise might efficiently activate in dystrophic muscles compensatory mechanisms for this process [106], which has been hypothesized to contribute to the deficit of regenerative activity by promoting the premature senescence of satellite cells [107].
Based on the above reported experimental evidences, we can conclude that beneficial effects on mdx limb muscles induced by horizontal treadmill running exercise depend primarily on training intensity, being mainly associated to running speed ranging between 4 and 9 m/min. These positive effects, including fibrosis reduction, increased regeneration, improvement of muscle performances and structure, such as muscle hypertrophy and fiber type switch, seem to be not influenced by age, since they can be found in mdx mice 4–5 weeks old [108,109], 8 weeks old [74,76,77,110], and up to 20 weeks old [73,75,79], or by the duration of training, given the persistence of positive effects for exercise duration ranging between 1 and 6 months. On the contrary, horizontal treadmill exercise at higher intensity, with speed equal or greater than 12 m/min, or downhill/uphill running, generally exert detrimental effects on mdx muscles, and are often used to worsen the mdx phenotype before assessing the efficacy of drug treatment [8,81,82,84,87,92,99].
Interestingly, another key factor that influences exercise outcome is the number of training sessions. Indeed, while generally chronic exercise is associated to muscle beneficial effects, a single bout of treadmill running at a speed of 12 m/min or downhill running induced detrimental effects on mdx limb muscles, by generating membrane breakdown, myofibers necrosis, inflammation and oxidative stress [87,111,112,113,114,115,116]. Interestingly and consistently with mdx data, acute or single bouts of exercise resulted in increased serum CK activity and circulating myoglobin in DMD patients [117], likely due to the fact that the beneficial plastic muscle remodeling requires longer time to overwhelm muscle stress induced by acute exercise. In this context, another critical factor that seems to influence the outcome of exercise in mdx muscles is the sampling time following exercise. Indeed, it seems that tissue samples collected immediately after an acute o chronic high intensity treadmill running protocol is associated to increased damage, oxidative stress and inflammation as compared to samples collected 24 h or 96 h after the completion of the last training session [87], which may reflect a muscle adaptive response to exercise-dependent stress over time.
It is evident that exercise protocol might discordantly affect different muscles in mdx mice, mainly according to different muscle workload in response to training, and to specific muscle morphometric features. For example, as general principle, compared to the diaphragm, mdx limb muscles present larger necrosis area, but reduced fibrosis, following regeneration [13]. Furthermore, even skeletal limb muscles can differently respond to exercise, as, for example, low intensity training induced a decrease in gastrocnemius and soleus muscles necrotic area, while it exerted opposite effects on plantaris mdx muscle [78].
Similar to limb muscles, treadmill intensity strongly influences adaptive remodeling of cardiac muscle and diaphragm. Indeed, low intensity exercise, at a speed of 4–8 m/min on a treadmill, besides enhancing cardiac function, improved respiratory capacity [79] and increased the ratio between diaphragm regeneration and necrosis areas [76] without worsening cardiac and diaphragmatic fibrosis [79]. Conversely, higher intensity exercise seems to be associated to deleterious effects and enhancement of fibrotic tissue deposition in cardiac muscle and diaphragm. Indeed, running on a horizontal treadmill at 12 m/min induced a reduction in diaphragm twitch and tetanic tension [86], an increase in diaphragm necrosis [87] and an increase in cardiac collagen deposition and fibrosis [15]. Uphill running induced upregulation of phosphorylated p38 MAPK, phosphorylated ERK1/2 and calcineurin, extensive infiltration of inflammatory cells, together with increased interstitial fibrosis and adipose tissue in heart [118]. Similarly, downhill running increased the expression of Transforming Growth Factor Beta 1 (TGFβ1), a key modulatory of fibrosis deposition, on biceps brachii and heart [83], and enhanced diaphragm necrosis [25,93]. However, in contrast to these data, other papers have rather reported no detrimental effects to diaphragm muscle associated to running on a horizontal treadmill at 12 m/min [80,92,99].
Finally, diaphragm response to exercise seems to depend also on the sex of the subjects. Indeed, male mdx mice submitted to a 4-week period of 15–30 min running on a treadmill at a speed of 6 m/min, showed increased CK levels and inflammatory area as compared to trained females, suggesting that estrogens, whose receptor expression was increased following exercise in female mdx mice, may have contributed to the prevention of increased inflammatory process and diaphragm injury [119]. Accordingly, some investigations suggest that there are differences between female and male muscles, such as energy metabolism, fiber type composition, and contractile speed, greatly dependent on estrogen levels present in females [120]. Interestingly, estrogens may play an important protective role against muscle damage, by attenuating the inflammatory and oxidative process [121,122,123] and/or influencing muscle regeneration [124,125], thus affecting all the parameters modulated by exercise.
However, a similar protocol of training, consisting in 4 weeks of 30–60 min running on a treadmill at a speed of 4–4.8 m/min, did not seem to induce detrimental effects on male mdx diaphragm [76], thus suggesting the need for more specific comparative studies in order to address the impact of animal sex on training-induced muscle plastic remodeling. Nonetheless, it is worth emphasizing that the pre-training and the warm-up phase consisting of progressive increasing in muscle workload (in terms of intensity and duration of a single session of training) employed by Morici et al. [76] might have limited exercise-induced detrimental effects associated to diaphragm overload, promoting gradual plastic remodeling associated to morphological and functional improvement of male dystrophic muscle [76]. In support to the beneficial role of pre-training, it seems that even a treadmill exercise at a speed of 12 m/min was able to induce a reduction of serum CK levels if preceded by a warm-up period and a gradual increase of running speed [126], which is also responsible for the increased ability of mdx mice to complete a 30 min treadmill exercise session [87].
In conclusion, despite the limited available data on heart and diaphragm plastic remodeling in response to treadmill running complicate the definition of a reliable overview, they suggest that, similarly to limb muscles, beneficial effects on these muscles are mostly associated to low intensity treadmill training.

4.2. Effects of Swimming Exercise on Plastic Remodeling of mdx Muscle

Swimming training, although less extensively used in mice exercise interventions, presents some advantages compared to treadmill training. Indeed, swimming recruits all body muscles and ligaments, thus representing an effective form of aerobic endurance training. On the other hand, swimming represents a quite stressful modality of training, and requires careful and constant monitoring during the entire experiment to prevent mice from drowning or floating [67]. Swimming training intensity can be classified, according to the amount of swimming that takes place each day, as low intensity (20–59 min/day), moderate intensity (60–89 min/day), and high intensity (≥90 min/day) [67].
The effects of swimming on mdx muscle plastic remodeling have been summarized in Appendix A Table A3. Swimming, with variable duration up to 10 weeks, seems to produce adaptation to the functional demand and beneficial effects on mdx limb muscles, independently on intensity of exercise. Indeed, low intensity swimming for 30 min once a day, for 4 weeks, increased forelimb grip strength and decreased inflammation [127] and oxidative stress [128]. Interestingly, swimming exercise protracted for as long as possible, but never exceeding 25 min of training, once a day, for 10 weeks, exerted no detrimental effect even on 24 months old mdx mice, and increased relative tetanic tensions [129]. Even a more intense swimming exercise, executed for 2 h once a day for 15 weeks, but associated to a pre-training protocol, produced beneficial effects on limb muscles, by reducing muscle fatigability [130] and by increasing the proportion of oxidative fibers and twitch tension [130,131], confirming the beneficial effects of pre-training on mdx muscle adaptive response. Surprisingly, low intensity swimming consisting of 30 min bouts of training per day conducted on 4 consecutive days, seems to rather increase muscle necrosis and induce an increase in serum CK [132], maybe due to the short training duration which might prevent long-term beneficial muscle adaptation to training. Furthermore, a moderate protocol of swimming for 60 min once a day, for 2 months, produced detrimental effects on heart and diaphragm, exacerbating muscle degeneration, inflammation, and fibrosis in 11 months old mdx mice [133]. These data suggest that, similar to the effects of treadmill training, limb muscles, heart and diaphragm do not show the same adaptive remodeling in response to swimming training. Interestingly, a single 20 min swimming session seems to be detrimental for mdx muscle by inducing a membrane breakdown in tibialis anterior muscle [134], thereby confirming that acute exercise is generally associated to deleterious muscle remodeling.

4.3. Effects of Voluntary Running on Plastic Remodeling of mdx Muscle

Different to forced running, voluntary wheel running allows mice to exercise at a lower intensity and to freely run. Although this training may generate individual differences in the amount of exercise among mice, the animals are subjected to a lower stress [67]. The effects of voluntary running on mdx muscle plastic remodeling have been summarized in Appendix A Table A4. Voluntary exercise, independently of training duration, seems to be generally associated to beneficial effects on the functional properties of mdx muscles. Indeed, short-term (1 week) voluntary wheel running, initiated when mice were 2-4 months old, by activating calcineurin pathway and modifying the expression program of genes involved in excitability and slower contractile phenotype, preserved muscle excitability and improved tibialis anterior muscle fragility in mdx mice, without worsening weakness [135]. Longer periods of voluntary exercise seem to positively affect mdx muscle physiology as well. For example, 3 weeks of voluntary wheel running in maturing 21 days old mdx mice was not detrimental and enhanced endurance capacity by inducing molecular adaptions in both skeletal and cardiac muscle, such as increased total contractile protein content and markers of aerobic metabolism [136]. Four weeks of voluntary wheel running initiated when mice were 2 months old, increased cross sectional area and reduced protein ubiquitination in triceps brachialis muscle [137], while 8–9 weeks of voluntary wheel running, initiated at the age of 3–4 weeks, improved muscle performance and induced adaptive response in mdx muscles, such as increased muscle mass in soleus and the shift of fibers toward a less fatigable phenotype [138,139], without improving the activity of mitochondrial enzymes of the Krebs cycle, β-oxidation, and the electron transport chain [138]. This nonadaptive mitochondrial response is unusual considering that adaptive myosin heavy chain isoform response and the shift from IIb toward IIa fibers typically occurs together with the enhancement of mitochondrial enzyme activities.
Continuing with the review of positive effects associated to voluntary exercise, 12–18 weeks of low resistance wheel running initiated in 4 weeks old mdx mice led to increased utrophin, a dystrophin protein homologue that enhances dystrophic muscle function [140], improved contractile function and reduced fatigability, and increased aerobic metabolism along with a trend of fiber type transformation toward a slower-contracting and less fatigable fiber type [141,142,143]. Beneficial results were also obtained with 12 weeks of voluntary progressive resistance wheel running, which was able to increase forelimb muscles performance in 4–5 weeks old mdx mice [144]. Noteworthy, 3 or 11 months of wheel running seem to exert positive effects on mdx limb muscles even when initiated at an age of 6 or 7 months, by ameliorating the age-associated loss in tension production and fatigability [145,146]. Finally, 1 year of voluntary wheel running induced positive exercise-induced remodeling of limb muscles, such as the increase in the muscle mass and tetanic force [147].
In addition to this large body of evidence, mostly demonstrating beneficial effects of voluntary exercise on dystrophic limb muscle, other studies have rather described detrimental roles associated to this type of training. Indeed, 2–3 weeks of voluntary wheel running started in 9–12 weeks old mice exacerbated limb mdx phenotype and increased the amount of damaged necrotic tissue and macrophage infiltration, associated to a more pronounced inflammation and fibrosis genes dysregulation [148,149,150]. Moreover, wheel running for 8 weeks increased skeletal muscle fibers and endothelial cells apoptosis in young mdx mice [151], although it remains controversial whether this process is related to repair/regeneration mechanisms or may contribute to the pathogenesis of muscle damage [152].
Based on the above reported literature data, it is possible to conclude that voluntary wheel running seems to be definitely associated to beneficial effects on mdx limb muscles especially if initiated early, at an age ranging between 1 and 6 weeks, and independently on the duration of training. On the contrary, under the same conditions of training duration, a later initiation of voluntary exercise is associated to discordant outcomes, either beneficial [135,145,146] or detrimental, such as the enhancement of inflammation, necrosis and fibrosis [148,149,150,153]. These differences, in addition to be influenced by changes in the age and the subsequent stage of the dystrophic process, which may affect muscle regeneration and remodeling ability [154], are likely dependent on the large individual variations associated to this typology of exercise. Furthermore, as already pointed in relation to other modalities of training, sex of studied animals and the presence of estrogen receptors may contribute to the prevention of muscle injury in mdx mice and affect muscle response to exercise [119]. Accordingly, voluntary activity was not detrimental for mdx limb and cardiac muscles of either sex but increased absolute maximal force and muscle weight only in female mice [146].
Independently of age of exercise initiation, another factor that strongly influences voluntary exercise outcome is the number of training session. Indeed, a single bout of training seems to be detrimental and induces muscle damage, fiber apoptosis [152,155,156,157], and an increase in membrane permeability [157], thereby confirming acute deleterious effects associate to training.
Like treadmill running and swimming, cardiac and diaphragm muscles, as compared to limb muscles, need different generalizations. Indeed, for example, while long term wheel running, initiated when mice were 4 weeks of age, improved cardiac and plantarflexor function in the mdx mouse, it greatly impaired diaphragm function, likely due to increased respiratory muscle workload and frequency of eccentric contractions during exercise training [147]. However, other studies have rather demonstrated that 1 year of wheel running in young mdx mice increased diaphragm active tension [139], while lifetime wheel running increased its contraction time [158], thereby decreasing progression of muscular dystrophy. A large variation in muscle response to voluntary training can be also observed in cardiac muscle. For example, 12 weeks of low resistance wheel running initiated in 4 weeks old mdx mice led to the increase in heart mass [141], while a similar protocol of training in age-matched mdx mice did not seem to produce variation in cardiac relative mass [144]. Furthermore, 3 or 8 weeks of voluntary wheel running did not exacerbate heart mdx phenotype in aged [159] and maturing mdx mice [136], respectively, and, in the latter case, enhanced the expressions of aerobic metabolism markers, such as citrate synthase (CS) and β-hydroxyacyl-CoA dehydrogenase activities [136]. On the contrary, 4 to 4.5 months of voluntary wheel running, initiated when mice were 4–5 weeks old, reduced left ventricular function [143]. Finally, 4 weeks of voluntary wheel running accelerated the progression of ventricular dilation and fibrosis in 7 weeks old mdx mice [160], while a longer training period in 7 months old mice, did not affect left ventricular function, structural heart dimensions, cardiac gene expression of inflammation, fibrosis, or remodeling markers [146].
In conclusion, although experimental evidence has demonstrated muscle beneficial effects induced by voluntary exercise, especially in limb muscles, the large individual variation in terms of effective muscle load associated to this modality of training impedes a clear correlation between exercise and observed effects and hinders an exercise protocol standardization [67,148], thus suggesting that this modality of training might not represent the most efficient therapeutic approach for DMD patients.

5. Mitochondria Impairment and Redox Equilibrium in DMD Muscles

Loss of dystrophin in DMD muscles generates a systemic metabolic impairment, which is a key contributor to the etiology of the disease. Indeed, DMD is characterized by a significant dysregulation of intracellular Ca2+ homeostasis and by a deficiency in glycolysis, purine nucleotide, and tricarboxylic acid cycle, as well as in oxidative phosphorylation [161,162,163,164,165,166,167,168,169], which result in bioenergetic impairment and reduced ATP levels [170,171]. This, in turn, leads to the impairment of muscle contraction and, ultimately, to cell necrosis.
Muscle necrosis is closely associated with increased inflammation and oxidative stress [172], and both these processes can be efficiently modulated by exercise.
Oxidative stress is a cell condition caused by an excess of intracellular free radicals, such as reactive oxygen species (ROS) and radical nitrogen species (RNS). Normally, the levels of free radicals inside the muscle fibers are strictly balanced by controlling the rate of their synthesis and removal by antioxidants defenses. Usually ROS are physiologically produced at low amount in healthy cells and an accurate and well organized control of their levels guarantees the maintenance of physiological cell functions involving ROS as cell signaling molecules, i.e., control of gene expression, regulation of cell signaling pathways and modulation of skeletal muscle force production [173,174,175]. On the other hand, excessive ROS/RNS production and/or defects of antioxidant systems may impair the cellular redox balance, inducing oxidative stress in cells, subsequent damage to biologic macromolecules, and cell death [176,177].
A large body of evidence shows that oxidative stress, driven by free radicals produced in damaged myofibers or released by inflammatory infiltrating cells, represents a key pathogenic event in DMD [18,172,178,179,180,181]. Accordingly, antioxidant drugs (coenzyme Q10, green tea extracts, resveratrol, N-acetylcysteine), although with several caveats, are considered a promising treatment strategy in muscular dystrophy [182,183]. The prominent role of oxidative stress in the pathology has been suggested quite early by observing that muscles from DMD patients showed higher level of lipid peroxidation and enhanced antioxidant enzymes activity [184]. In general, abnormal enzymatic antioxidant responses and increased levels of oxidant molecules and markers of oxidative damage to macromolecules, such as 8-hydroxy-2′-deoxyguanosine, oxidized proteins, and lipids, have been found in DMD muscles and blood [185,186,187,188,189]. Interestingly, the extent of oxidative damage further increased with pathology progression and aging of affected subjects [189].
Mitochondria represent a crucial site of ROS production, and impairment of mitochondrial respiration strongly contributes to DMD pathogenesis in both patients and animal models [164,190,191]. Specifically, DMD-induced perturbation of intracellular Ca2+ homeostasis, metabolism and ATP production seems to be bidirectionally associated with mitochondria impairment [165,179,190]. Under physiological conditions, the mitochondrial electron transport chain transfers a single electron to molecular oxygen leading to the synthesis of ATP and superoxide ions that can be neutralized by antioxidant systems [192]. Loss of dystrophin in DMD induces destabilization of cell membrane and disruption of cytoskeleton organization, and the subsequent abnormal Ca2+ influx, which activates proteases and leads to mitochondrial Ca2+ overload and dysfunction [193,194]. Specifically, Ca2+ overload triggers mitochondrial permeability transition pore (PTP) opening, alteration in mitochondrial membrane potential and swelling of the organelles, ROS overproduction, and oxidative stress generation, which result in the subsequent exacerbation of mitochondrial dysfunction that further impairs cell bioenergetics, Ca2+ homeostasis and cell viability [185,195,196]. In other words, loss of dystrophin generates a feed-forward loop in which mitochondria impairment induces muscle ROS increase and oxidative stress, which further compromise mitochondria functions, leading to even greater production of ROS and muscle necrosis [197]. Furthermore, monoamine oxidase (MAO), a mitochondria enzyme catalyzing the oxidative deamination of biogenic amines to generate H2O2, is increased in mdx muscles [198], contributing to the overall oxidative imbalance by generating higher levels of H2O2, which in turn alter the redox homeostasis and causes myofibrillar protein oxidation, muscle damage and impairment of contractile function [198]. In addition to the alteration in mitochondria enzymatic activity, reduced density of sub-sarcolemmal mitochondria and abnormal localization of inter-myofibrillar mitochondria contribute to the overall mitochondria dysfunction in mdx mice [191]. Notable, elevated levels of mitochondria H2O2 in mdx mice, due to impaired oxidative phosphorylation, aberrant MAO expression and activity and altered mitochondria biogenesis and dynamics, occur very early in the disease process in 4 weeks old mice, before the onset of the disease and myofiber necrosis, and are associated with the induction of mitochondrial-linked caspase 9, necrosis, and severe myopathy in skeletal muscles and diaphragm [166,199,200]. These pieces of evidence indicate that mitochondria dysfunction is likely involved in the initial phases of pathology, being an essential contributor to DMD pathogenesis [178]. Accordingly, genetic or pharmacological overexpression of Peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC-1α), a master regulator of mitochondria functions, ameliorates mdx mice phenotype [201,202,203].
Besides mitochondria, NADPH oxidase (NOX), phospholipase A2 (PLA2), and xanthine oxidase (XO) are other important muscle sources of ROS in DMD. NOX enzymes transport electrons across biological membranes to reduce oxygen to superoxide or H2O2 [204,205]. Skeletal muscles express three isoforms of NOX (NOX1, NOX2, and NOX4) that have been described as key regulators of redox homeostasis [206,207]. In particular, NOX2 represents the major source of skeletal muscle ROS during contractions, while NOX4 is mostly involved in skeletal muscle hypertrophy induced by muscle overload [206]. Indeed, muscle contraction induces the association of NOX2 regulatory subunits, leading to its activation and the subsequent ROS generation [208]. NOX2 is also involved in the regulation of excitation-contraction coupling, since moderate levels of superoxide anion produced by its activity in the sarcoplasmic reticulum (SR) can activate ryanodine receptors and regulate the proper release of calcium from intracellular stores [209,210]. PLA2 stimulates NOX activity [211] and cleaves arachidonic acid, a lipoxygenases substrate for ROS production, from phospholipids contained in membranes, both in resting conditions and during contraction [212]. XO is a cytosolic enzyme that catalyzes the hydroxylation of hypoxanthine to xanthine and of xanthine to uric acid, generating ROS [213]. Several pieces of evidence have demonstrated a large enhancement of NOX-dependent ROS production in dystrophin-deficient heart and skeletal muscle [214], together with a hyper-activation of XO enzyme [215]. Accordingly, pharmacological inhibition of NOX and XO activity significantly ameliorated mdx muscle function by protecting from tissue damage and inflammation [215,216].
Importantly, membrane damage during DMD and subsequent muscle necrosis activate resident mast cells, which in turn secrete inflammatory mediators to recruit circulating inflammatory cells from the surrounding vasculature, like neutrophils and macrophages, which contribute to ROS generation to promote phagocytosis [217]. However, the chronic inflammatory state of DMD muscle leads to excessive ROS generation, oxidative stress exacerbation and to secondary damage of previously uninjured fibers [218,219].
In summary, inflammatory cells, mitochondria, NOX, and XO are the main sources of ROS in DMD [219,220]. Although mitochondria in muscle fibers have traditionally been considered the major intracellular generator of ROS, many recent evidences have suggested NOXs as the most significant source in DMD muscles, especially during contractions, since the increase in cytosolic ROS is quicker and greater than the rise in mitochondrial ROS [207,219,221]. In support to this thesis, pharmacologic o genetic blockade of NOX activity inhibits or attenuates contraction-induced increase in ROS [222,223,224].
In order to counteract the excessive increase in radicals and oxidative stress, muscle cells arrange enzymatic and non-enzymatic defenses. The primary enzymes in cells are represented by superoxide dismutase (SOD), glutathione peroxidase (GPX), and catalase. Other antioxidant enzymes, such as peroxiredoxins, glutaredoxins, and thioredoxin reductases, and non-enzymatic antioxidants (e.g., glutathione, uric acid, and bilirubin) also contribute to cellular protection against oxidation [225]. SODs catalyze the dismutation of superoxide in molecular oxygen and hydrogen peroxide [226], which, in turn, is converted into water by catalase or GPX. Three isoforms of SOD (SOD1, SOD2, and SOD3) were identified in mammalian cells. SOD1, which uses copper-zinc as a cofactor, is mainly located in the cytosol. SOD2 requires manganese as a cofactor and is localized in the mitochondria, while SOD3, which contains copper-zinc as a cofactor, is mainly located in the extracellular space [174]. In skeletal muscle fibers, SOD isoforms are involved in contraction of myogenic fibre [219] and their activity is greatly dependent on muscle fiber type and modality of exercise. Specifically, slow twitch and Type IIa fibers have higher levels of SOD2 and SOD1 than Type IIb and IIx fibers [226,227].
Similar to SOD, Activity of GPX and catalase changes according to fiber types, being highest in skeletal muscles composed of highly oxidative fibers compared to fibers with lower mitochondrial content [174,175]. Interestingly, muscles from mdx mice exhibit increased expression of antioxidant enzymes, such as SOD1, SOD2, GPX, and catalase, in the pre-necrotic state, indicative of a cellular response to oxidative stress [178,219], although other studies have provided contrasting results about the levels of antioxidant enzyme activities in dystrophin-deficient muscle, likely due to the different stage progression of the pathology [184,228,229,230]. On the contrary, decreased levels of glutathione, accompanied by a concomitant increase in the activity of glutathione metabolizing enzymes (GPX and glutathione reductase), have been detected in DMD muscles, contributing to the generation of oxidative stress in these muscles [181,231].
Finally, the overall control of fibers redox state is also dependent on SR and on its role in protein-folding homeostasis [232,233]. Indeed, the accumulation of unfolded or misfolded proteins within the SR leads to SR stress and to the activation of an adaptive complex intracellular signal transduction pathway aimed to resolve protein misfolding and re-establish SR proteostasis through autophagy and enhancement of nuclear factor E2-related factor 2 (Nrf2)-dependent gene transcription [17,234,235], involved in mitochondrial biogenesis [236] and in the upregulation of antioxidant enzymes in response to oxidative stress [237,238]. However, depending on the severity of SR stress, activation of alternative pathways can even enhance ROS production, leading to further oxidative stress, and, eventually, apoptosis [239]. It has been reported that skeletal muscles in mdx mice are more susceptible to oxidative stress compared to those of wt animals [240], and that this difference might be, at least in part, attributed to the deregulation of Nrf2 signaling in mdx muscles [241]. Furthermore, SR stress markers, such as caspase-12, are increased in dystrophic muscle of mdx mice and DMD patients, suggesting that dystrophin deficiency leads to disruption of muscle SR homeostasis, which contributes to the worsening of oxidative stress, inflammation and DMD phenotype [242]. Accordingly, SR stress inhibitors restore mitochondria functions, mitochondrial Ca2+ uptake, and improve contractility of the diaphragm in mdx mice [243].

6. Exercise Modulation of mdx Muscle Oxidative Stress and Inflammation

It is known that muscle cells can generate ROS in resting conditions, and even more during exercise [244,245]. Indeed, during exercise, the increased energy demand leads to the boost in mitochondrial activity and oxygen consumption and subsequent muscle ROS generation [212]. The exercise-induced increase in ROS levels is essential to modulate muscle adaptation and force production, and exerts beneficial or detrimental effects according to ROS concentration, which mostly depends on duration and intensity of exposure and training status of the subject [174,205,245].
Usually, endurance training produces low concentrations of ROS, which in turn induce the expression of antioxidant enzymes and other defense mechanisms [226,246] and beneficial muscle adaptations [247,248]. Noteworthy, the beneficial effects of exercise are inhibited by administration of antioxidant compounds, such as vitamin C and E [249,250,251] thus suggesting that ROS, at moderate levels, act as critical signals in exercise promoting useful plastic adaptations in muscle. Endurance training compensates also for the decrease, during aging, PGC-1α pathway [252], the master transcriptional coactivator involved in the regulation of mitochondrial biogenesis and metabolism. Specifically, exercise-induced PGC-1α expression strongly improves the control of energy metabolism by promoting mitochondrial oxidative metabolism and mitochondrial biogenesis, protects against development of sarcopenia, muscle oxidative stress, and inflammation, and drives angiogenesis [203,253]. In addition, exercise, by activating several interconnected intracellular signaling, including PGC-1α and AMP-activated protein kinase (AMPK) pathways, in aged muscles, ameliorates mitochondrial quality by altering mitochondrial dynamics [254,255,256] and promoting mitophagy of damaged or dysfunctional mitochondria [257,258]. All the above mentioned processes contribute to guarantee muscle mass and strength preservation in senior sportsmen compared to age-matched sedentary subjects [259]. These beneficial effects of exercise on aged muscles strongly support the potential positive effect of low intensity training on DMD patients, which show progressive muscle weakness and deterioration with age [1,54]. Indeed, although the shorter longevity, DMD muscles share several features with aged muscles, such as increased oxidative stress, fibrosis, sarcopenia, altered properties of satellite cells and impaired regeneration [260], which might be efficiently compensated by low intensity training. Accordingly, increasing, for example, PGC-1α expression (by exercise, pharmacologic agents, genetic manipulation, etc.) exerts beneficial effects on aged muscles as well as on DMD muscles [203,261].
In order to regulate the overall muscle redox state, exercise can also influence activity and expression of antioxidants systems. In details, regular/moderate exercise enhances antioxidant defenses by inducing the activity of endogenous antioxidant enzymes such as SOD, GPX and catalase [262,263]. Moderate exercise, indeed, promotes the activation of Mitogen-activated protein kinase (MAPK) pathway, which in turn increases the expression of antioxidant enzymes, thus counteracting excessive ROS generation and stimulating adaptation to exercise [264]. Specifically, although some studies reported that prolonged endurance exercise training does not enhance muscle SOD activity [225,246], most investigations have shown that endurance exercise increases SOD muscle activity [227,265,266,267,268,269] proportionally to the intensity and duration of exercise [227,267]. Specifically, endurance training, or moderate exercise, seem to primarily induce the enhancement of SOD2 protein and activity in type IIa fibers [226,246,270,271], even if several reports have also described SOD1 activation [272]. More recently, the increased expression of SOD3 mRNA has been also reported in skeletal muscle after acute exercise [273].
While is still controversial whether catalase expression in skeletal muscle is enhanced by chronic exercise [225,227,268], GPX, similarly to SOD, is strongly increased in muscle fibers during training [246,265,274,275,276,277], according to both intensity and duration of exercise. Specifically, high intensity exercise induces a greater increase in muscle GPX activity if compared to low intensity exercise [227].
Similarly to wt muscle, low intensity exercise seems to be able to counteract oxidative stress in mdx muscles. Indeed, 6 months of low intensity treadmill running increased PGC-1α expression in mdx cardiac muscle [79], which acts as a central inducer of mitochondrial biogenesis and a primary regulator of redox balance inside contracting muscle [253,278]. 8 weeks of low intensity treadmill running, at a speed of 9 m/min, started when mice were 4 weeks old, reversed lipid and protein oxidation in the gastrocnemius of mdx mice [108,109], increased the antioxidant activity of glutathione, free thiols concentrations and reversed the changes in mitochondrial respiratory chain complex activity associated to the disease, thus improving energy metabolism [108]. The same exercise protocol, started when mice were 11 weeks old, although associated with higher levels of lipid peroxidation and protein carbonylation in the tibial anterior and gastrocnemius muscles, was also able to enhance SOD activity and total antioxidant capacity [73]. These observations suggest that the exercise-dependent increase in antioxidant enzymes is not sufficient to reverse oxidative damage to macromolecules, likely due to the age of treatment initiation, but can sustain the attenuation of intramuscular collagen fibers deposition [73], which occurs in mdx mice after the tenth week of age [8]. A modulation of antioxidant defenses was also observed following 30 days of low intensity endurance exercise on rotating treadmill, which produced an increase in quadriceps SOD1 levels [110]. In agreement with the beneficial effects associated to exercise and dependent on the up-regulation of antioxidant systems, the catalytic mimetic of SOD and catalase or catalase overexpression have been shown to reduce muscle damage and markers of oxidative stress [279] and ameliorate functions of skeletal muscles in mdx mice [280].
Different protocols of low intensity training, such as 30 min of swimming per day for 4 weeks, corroborated the positive effects of training on mdx oxidative stress, showing a decrease in protein carbonylation in the gastrocnemius muscle when compared with the non-exercised mdx group [128]. The only exception to positive effect associated to low intensity training is represented by the study of Faist et al., in which running on a treadmill at a speed of 8 m/min, even in presence of a pre-training, decreased oxygen consumption, respiratory control indices and GPX activity and increased lipid peroxidation in young (4 weeks) mdx muscle [281]. Surprisingly, no effects or even opposite effects were obtained by the same protocol of exercise in adult (16 weeks), suggesting that young mdx muscle is more vulnerable to exercise-induced to oxidative stress as compared to adult muscle, likely due to the higher degree of regenerating fibers in adult mdx mice [281]. However, as already described, other studies have shown beneficial effects associated to a similar protocol of exercise in young mdx mice [108,109], thus suggesting the need for further investigations finalized at a more reliable standardization of a training protocol associated to a reduction of DMD muscle oxidative stress.
The role of voluntary exercise on mdx oxidative stress regulation is controversial. On one hand, 7 weeks of voluntary run in 6 to 7 weeks old mdx mice enhanced autophagy and reversed the oxidative phenotype by counteracting the decrease of PGC1-α content [282], whose increased activation has been shown to ameliorate DMD pathology [202,283]. On the other hand, the same protocol of exercise, performed by age-matched mdx mice, increased oxidized glutathione and protein carbonyl levels in gastrocnemius, without increasing ER stress [284]. One week of voluntary wheel running promoted the slower contractile phenotype switch in tibialis anterior muscle of 2–4 months old mdx mice and up-regulated the expression of Mitochondrial transcription factor A (Tfam), Nuclear respiratory factor 1 (Nrf1), and peroxisome-proliferator-activated receptor-gamma co-activator 1 beta (PGC1-β) [135], which are important regulators of the oxidative metabolism [285,286]. Three weeks of voluntary wheel running in 21 days old mdx mice increased serum antioxidant capacity and markers of oxidative metabolism, such as CS and β-oxidation activity, necessary to determine metabolic adaptations to endurance exercise, in both quadriceps and heart muscles [136]. Similar metabolic modifications were also observed in the gastrocnemius of 4 weeks old mdx mice submitted to 12 weeks of voluntary wheel running together with the up-regulation in cytochrome c oxidase subunit 4 levels and a trend PGC-1α increase [141], which suggest enhanced mitochondria metabolism and biogenesis in response to training. However, given the limited amount of available data and the large variability associated to voluntary training, further studies are necessary to clarify the role of this type of exercise on mdx oxidative stress.
In healthy muscle cells, when ROS amount reaches a maximum peak, the linear relationship between ROS and muscle generated force fails and a further increase in ROS induces a strong reduction in the force and muscle oxidative damage, such as protein carbonylation, DNA damage, and RNA oxidation [287]. Usually, oxidative damage and inflammatory processes in active myofibers are associated to prolonged and intense exercise [288,289,290]. One of the main mechanism involved in oxidative damage to muscle fibers following strenuous exercise, which is associated to ROS production, proinflammatory cytokines and increased Nuclear factor-kB (NF-kB) activation, is the impairment of pathways involved in maintaining mitochondrial integrity and morphology [291,292]. Moreover, excessive ROS production is associated with the impairment of PGC-1α pathway and alters calcium homeostasis, leading to aberrant NF-kB transcriptional activity, activation of proteolytic systems and muscle wasting [293]. Finally, during intense prolonged exercise, XO activity is significantly increased leading to the excessive ROS generation, oxidative stress and muscle damage [294], while a single session of exhaustive exercise is able to determine oxidative damage only in untrained persons or senescent muscle, due to a higher vulnerability to oxidative stress [295]. In addition to these intracellular sources, ROS can also be produced from non-muscle sources during training, especially by immune cells activated by exercise-elicited muscle injuries [296,297], and by oxidation of catecholamines, whose plasmatic concentrations increase in response to exercise [298].
Similar to wt muscle, high intensity exercise is mostly associated to worsening of mdx phenotype and muscle redox status. Indeed, acute downhill treadmill running at a speed of 15 m/min caused significant increases of XO activity and the subsequent systemic ROS generation as indicated by elevated urinary isoxanthopterin in mdx mice [215]. Four weeks of moderate intensity treadmill running at 12 m/min speed increased ROS [92] and oxidized glutathione levels [15], and further worsened calcium homeostasis and sarcolemmal permeability [85]. Similarly, 6 weeks of the same protocol of exercise increased protein thiol oxidation [111], which was also significantly enhanced even after a single treadmill session [87]. Finally, 12 weeks of treadmill running at 12 m/min induced a significant reduction in PGC-1α expression, accompanied by a marked downregulation of Sirtuin 1 (Sirt1) and Peroxisome proliferator-activated receptor γ (Pparγ) [88], which play key roles in exercise-induced adaptation, via stimulation of the fast-to-slow transition, mitochondrial biogenesis and cellular antioxidant response [283,299]. The same exercise is associated also to a reduction of the autophagy marker BCL2 Interacting Protein 3 (Bnip3) [88]. Interestingly, a single session of the same protocol of exercise executed by 16 weeks mdx mice did not influence the expression of PGC-1α, Sirt1 and Pparγ [88], in contrast to all the other data on muscle plastic remodeling demonstrating mostly detrimental effects associated to acute exercise [87,111,112,113,114,115,116,134,152,157].
It is well known that inflammation process is strictly linked to oxidative stress. Indeed, oxidant species can mediate activation of NF-kB [300] which in turn regulates the expression of genes involved in the inflammatory and stress response [301]. Conversely, inflammation leads to immune cells infiltration in damaged muscled, which in turn determine the release of free radicals, generation of oxidative stress and chronic inflammation, thereby forming a self-propelled vicious cycle [302].
Inflammation is an early event in the pathological process of DMD, closely associated with myonecrosis, already evident in 3–4 weeks old mdx mice [303]. Indeed, loss of dystrophin leads to a fragility of plasma membrane that is easily injured during muscle contraction, inducing extracellular calcium influx, which in turn activates the NF-kB inflammatory pathway and enhance inflammatory cytokine release, e.g., Tumor Necrosis Factor α (TNFα) and Interleukin (IL)-1β, which further impair muscle regeneration [304]. Hence, activation of NF-kB is an important mediator of muscle damage and DMD pathology in both humans [305] and mdx mice [303,305,306]. Accordingly, calcium influx is sufficient to induce muscular dystrophy [307], while genetic ablation [306] or pharmacological blockade [305,308,309] of NF-kB alleviate the severity of the disease and improves muscle phenotype of mdx mice. Glucocorticoid therapy interferes with inflammatory processes related to the pathology and mitigates elevation of NF-kB and oxidative stress, determining the subsequent reduction of fiber damage and the enhancement of muscle functional properties [151,310]. However, this treatment is associated to severe side effects such as weight gain, behavioral changes, growth issue, late puberty, bone demineralization, and gastroesophageal reflux, and can only slow down disease progression [26,27,28,311].
Interestingly, emerging evidence described the beneficial effects induced by exercise on mdx inflammatory processes. In details, low intensity swimming training decreased Monocyte chemoattractant protein-1 (a chemoattracting agent playing a crucial role in the coordination of the inflammatory response) levels and, although increased CD68+ macrophage numbers, shifted the macrophage population from to M1 pro-inflammatory type to the regenerative M2 type, thus inducing a modulation of inflammatory state of gastrocnemius, a switch of fibers to the oxidative slow type and a grip strength increase [127]. A low intensity running exercise on treadmill, preceded by a pre-training, exerted anti-inflammatory effects on mdx quadriceps muscles by reducing NF-kB expression [76] and increased serum adiponectin, which shows anti-inflammatory properties as well as modulatory effects on oxidative stress [79].
On the contrary, acute or chronic treadmill running at higher intensity (12 m/min speed) or downhill/uphill running generally worsened the inflammatory status of mdx mice [98], leading to increased infiltration of inflammatory cells [90,118] and expression for pro-inflammatory cytokines IL-1β, IL-6 [87], and TGFβ1 [81,82,83], and decreased expression of protective factors such as insulin-like growth factor 1 [89], PGC-1α and Pparγ [88], which exert anti-inflammatory actions by inhibition of NF-kB [299,312]. Moderate exercise, even in the swimming modality, is also associated to the worsening of inflammatory status, leading to the increase of inflammatory cells in cardiac muscle of 11 months old mdx male mice [133]. Increased inflammation and macrophages number were also observed after 3 weeks of voluntary running initiated when mice were 9 weeks old [149].
In summary, although the few available data prevent definitive conclusions, low intensity exercise and voluntary wheel running, with only few exceptions, besides improving structural and metabolic adaptions to exercise, seem to reverse oxidative stress and inflammation associated to DMD pathology, while higher intensity exercise worsens these processes. This conclusion mostly applies to limb muscles, whereas few data are available for cardiac and diaphragm muscles. The lack of an adequate amount of experimental data affects also the evaluation of mdx redox and inflammatory muscle state after acute training, which is usually associated to the worsening of mdx phenotype.
In addition to training modality, duration, intensity, and frequency as major determinants of muscle adaptation to exercise, other parameters, including age, sex, training status of the animal and time point of sacrifice following exercise, may strongly affect mdx muscle response to exercise. Age of exercise initiation represents a key factor in influencing muscle plastic remodeling, in both healthy and diseased subjects. DMD is a progressive chronic muscle degenerative disorder, whose symptomatology worsens with age [5]. Consequently, timing of therapy initiation, including exercise, can strongly affect the severity of clinical outcome. Accordingly, voluntary exercise is mostly associated to beneficial effects if initiated in mdx mice not older than 6 weeks.
DMD mainly affects boys, although female symptomatic carriers, with a different severity of DMD phenotypes, have been frequently described [313,314]. Gender, and the related hormonal variations, may greatly influence muscle physiology [315,316], disease progression and therapeutic outcome, even in response to exercise. Accordingly, female gender is associated to muscle beneficial effects in response to exercise [146] and seems to protect diaphragm from exercise-induced damage [119].
Training status of the subject, and the presence of a pre-training period and a warm-up phase associated to the main protocol of exercise, by limiting exercise-induced detrimental effects associated to sudden muscle overload, represent another key factor influencing muscle plastic remodeling in response to exercise. Indeed, pre-training and warm up, by inducing gradual metabolic adaptation to exercise and increased blood flow and temperature into the involved muscles, can decrease the risk of muscle injuries, as well as reduce heavy loads on the heart, which can occur when high intensity exercises are suddenly applied [317]. Accordingly, the presence of a pre-training phase seems to contribute to the enhancement of mdx muscle performance [87] and to preserve mdx muscles from high intensity exercise-induced damage [126,130,131].
Finally, the time of muscle sampling following exercise seems to influence the results of post-mortem analysis. Indeed, a later sampling following the last session of exercise is associated to decreased muscle damage and oxidative stress as compared to samples collected immediately after the last training session [87,113], likely due to the activation of muscle compensatory and adaptive response to exercise-dependent stress, including the activation of muscle regeneration and antioxidant defense.
The investigation of the burden and the combination of all these parameters, not enough explored in comparative studies, may strongly contribute to explain the differences in mdx response to similar protocols of training and to a better understanding of pathways involved in DMD pathophysiology, as well as testing of therapeutics hypotheses, including exercise.

7. Conclusions and Future Perspectives

Our comparative analysis of the existing literature clearly suggests that physical exercise, by inducing beneficial muscle plastic adaptations, represents a potential not invasive therapeutic approach for improving DMD patient outcomes and quality of life.
Specifically, chronic low intensity treadmill running, at a speed less than or equal to 9 m/min, might represent the most suitable exercise modality associated to beneficial effects on mdx muscle, especially hind limb muscles. This typology of exercise is mostly associated to the improvement of mdx phenotype, including the reduction of muscle oxidative stress, inflammation and fibrosis process, and the increase in muscle functionality (force, fatigue resistance), muscle regeneration and hypertrophy (Appendix A Table A5 and Figure 2). In addition, this modality of training generates reproducible results that seem to be little or not influenced by other variables, such as duration of training and age of subjects, which, on the contrary, appear to affect the outcomes of voluntary exercise, per se associated to larger individual variability.
However, when considering exercise as a therapy for DMD patients, it is essential to take into account the overall effect on patients and particularly to heart and respiratory muscles, whose failure is typically responsible of DMD patient death. Accordingly, an exercise protocol that protects or improves limb muscle function while impairing respiratory or cardiac function provides no or little therapeutic benefit to DMD patients. With respect to this key point, our analysis highlights the huge necessity for conducting new studies to deeply assess the benefits of exercise on all mdx muscles, including heart and diaphragm, whose functionality is poorly investigated in preclinical research. Moreover, we have shown that differences in the age, sex and training status of animals, time point of sacrifice following exercise, number of training sessions, exercise modality, intensity and frequency, might affect mdx muscle response to exercise, thereby pointing out the importance of new comparative studies finalized to investigate the burden of all these variables on muscle functional response to training. Given the ethical and practical issues of determining exercise prescription for DMD patients, these studies should be conducted first in dystrophic animals in order to determine functional thresholds to maximize the benefits of training and minimize the potential for exacerbating muscle injury.
These preclinical data can guide the design of appropriate studies on DMD patients. The main limitations in translational therapeutic application of mdx results to humans are the differences in phenotypic expression and biomechanics between humans and genetically homologous animal models. However, several studies performed in DMD patients seems to corroborate our conclusions, suggesting that submaximal low intensity exercise may be beneficial, especially if performed early in the course of the disease, while eccentric high intensity exercise is mostly associated to muscle injury.

Author Contributions

Conceptualization, G.M. (Giuseppe Morici), V.D.L. and G.M. (Giuseppa Mudó); writing—original draft preparation, V.D.L. and M.F.; writing—review and editing, V.D.L., G.M. (Giuseppe Morici), G.M. (Giuseppa Mudó), and M.R.B.; visualization, V.D.L., M.F., and G.M. (Giuseppa Mudó); funding acquisition, G.M. (Giuseppe Morici). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by University of Palermo, grant number PJ_RIC_FFABR_2017_ 008104.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Appendix A

Table A1. Effects of treadmill running on mdx mouse muscles.
Table A1. Effects of treadmill running on mdx mouse muscles.
A. Chronic Exercise
ReferenceSex/AgeProtocolMuscleHindlimb Muscle (Morphological and Functional Changes)Respiratory FunctionCardiac FunctionDegeneration/
Regeneration
FibrosisMetabolic Adaptive ChangesInflammatory MarkersOxidant/
Antioxidant Markers
Exercise Muscle Effects
Faist, V., et al. (2001)
[281]
4 weeks (young mice)
16 weeks
(adult mice)
Pre-training + LIT
Pre-training:
4 to 8.3 m/min,
2 × 5 to 2 × 30 min/day, 8 days
Training:
8.3 m/min,
2 × 30 min/day,
6 weeks
Femoris GAST
QUAD
Young mice:
= body weight
Adult mice:
= body weight
Young mice:
↑ CK
Adult mice:
= CK
Young mice:
↓ oxygen consumption
↓ RCI
= pADP/OC
Adult mice:
↑ oxygen consumption
= RCI
= pADP/OC
Young mice:
↑ TBARS
↑ Lipofuscin
= peroxyl radicals
↓ GSH
= SOD activity
= α-tocopherol
Adult mice:
= TBARS
= Lipofuscin
= peroxyl radicals
= GSH
= SOD activity
= α-tocopherol
(−) Young
(±) Adult
Fernandes, D.C., et al. (2019)
[73]
11 weeksLIT
9 m/min,
30 min/ day,
3 days/week,
60 days
GAST
TA
↓ collagen fiber area ↑ TBARS
↑ carbonylated protein in GAST and TA
↑ SOD activity
↑ FRAP activity
in TA
(±)
Fontana, S., et al. (2015)
[110]
Male
8 weeks
Pre-training + LIT
Pre-training:
3.2 m/min,
15 to 30 min/day,
5 days/week, 2 weeks.
Training: workload progressive increase,
4 to 4.8 m/min,
30 to 60 min/day,
5 days/week, 4 weeks
QUAD ↑ SOD1
↓ CA3
(+)
Frinchi, M., et al. (2013)
[77]
Male
8 weeks
Pre-training + LIT
Pre-training:
3.2 m/min,
15 to 30 min/day,
5 days/week, 2 weeks.
Training: workload progressive increase,
4 to 4.8 m/min,
30 to 60 min/day,
5 days/week, 4 weeks
GAST
QUAD
↓ degenerative myofibers
↓ Cx39
(+)
Gaiad, T.P., et al. (2017)
[74]
Male
8 weeks
Pre-training + LIT
Pre-training: Speed progressive increase.
Training: 9 m/min,
30 min/day, 3 days/
week, 60 days
TA = CLN↓ fibrosis deposition
(↑ Col 3 and
= Col 1)
(+)
Hermes, T.A., et al. (2018)
[119]
Male and female
2 months
LIT
6 m/min,
15 to 30 min/day,
2 days/week,
4 weeks
DIA ↑ CK (in males)
= CNF
↑ inflammatory area (in males) (−) males
(±) females
Hoepers, A., et al. (2020)
[108]
Male
28 days
Pre-training + LIT
Pre-training:
7 days, 4 m/min.
Training:
6 and 9 m/min,
30 min/day,
2 days/week,
8 weeks
GAST ↓ activity of complex II, II-III and IV ↓ lipid peroxidation
↓ carbonylated proteins
↑ free thiols
(+)
Kaczor, J.J., et al.
(2007)
[109]
28 daysLIT
9 m/min,
30 min/day,
2 days/week,
8 weeks
DIA, EDL, GAST
SOL
= CLN in DIA ↓ MDA in white GAST
↓ carbonylated protein
= OGDH activity in GAST
= COX activity in GAST
= SOD1 and SOD2 activity in GAST
= CAT activity in GAST
= GPX in EDL and SOL
(+) GAST
(±) EDL, SOL, DIA
Morici, G., et al. (2017)
[76]
Male
8 weeks
Pre-training + LIT
Pre-training:
3.2 m/min,
15 to 30 min/day,
5 days/week, 2 weeks.
Training: workload progressive increase,
4 to 4.8 m/min,
30 to 60 min/day,
5 days/week, 4 weeks
DIA, GAST QUAD ↑ regeneration area
↓ necrotic area in DIA
↓ NF-kB
in QUAD
(+)
Pinto, P.A.F., et al. (2018)
[75]
Male
11 weeks
LIT
9 m/min,
30 min/day,
3 days/week,
8 weeks
Lateral GAST = CLN↓ Col fibers type I
= Col fibers type III
(+)
Zelikovich, A.S., et al. (2019)
[79]
16–20 weeksLIT
4 m/min or
8 m/min,
30 min/day,
3 days/week,
6 months
(Each session consisted of 7’30’’ warm up followed by 22’30’’ training at target speed)
DIA, GAST
Heart
TA
↑ grip strength
↑ tetanic force in TA
↑ specific force in TA
↑ resistance to fatigue in TA
↑ type IIa fibers in TA
↓ Adipocyte CSA in GAST
↑ MV
↓ Ti
↓ Te
↓ LVIDD and LVPWT thickness= CNF in GAST, DIA and Heart
= serum CK
= fibrosis in GAST, DIA and Heart↑ PGC-1α in Heart
Trend ↑ PGC-1α in GAST
↑ Adipoq serum and in GAST
(+)
Zeman, J.R.,
et al. (2000)
[78]
10 weeksLIT
9 m/min,
1 h/day,
5 days/week,
10 weeks
GAST
Plantaris SOL
↓ necrotic area in GAST and SOL
↑ necrotic area in Plantaris
(+) GAST, SOL
(−) Plantaris
Burdi, R.,
et al. (2009)
[92]
Male
4–5 weeks
MIT
12 m/min,
30 min/day,
2 days/week,
4–8 weeks
DIA, EDL
GAST
TA
↓ forelimb strength= isometric tetanic tension in DIA ↑ CK = antioxidant activity
↑ plasma ROS level
↑ O-2 in TA
(−) EDL, GAST, TA
(±) DIA
Burdi, R.,
et al. (2006)
[99]
Male
4–5 weeks
MIT
12 m/min,
30 min/day,
2 days/week,
4–8 weeks
DIA, EDL
GAST
TA
↓ forelimb strength = serum CK
↓ gCl in EDL
= gCl in DIA
= gK
↓ TGFβ1 in DIA
Trend ↑ TGFβ1 in TA and GAST
(−) EDL, GAST, TA
(±) DIA
Camerino, G.M., et al. (2014)
[88]
Male
4–5 weeks
MIT
A:
12 m/min,
30 min/day,
2 days/week,
4 weeks.
B:
12 m/min,
30 min/day,
2 days/week,
12 weeks
EDL
GAST
SOL
A:
↓ forelimbs strength
↑ fatigue
↓ MHC2b
= SERCA1
= SERCA2
↓ HDAC5
= MHC isoforms
= Utrn
= Cn
= MEF2 isoforms
↓ HDAC5
B:
↓ forelimbs strength
= MHC isoforms
= SERCA1
↓ SERCA2
↑ Cn
= MEF2 isoforms
↓ HDAC5
A:
= Fst
= Myog
= IGF1
= Svil
↓ BNIP3
↓ LC3
B:
↓ Fst
↓ Myog
= IGF1
= Svil
↓ BNIP3
↓ LC3
A:
= TGFβ
= Mstn
= Atrogin1
B:
= TGFβ
= Mstn
= Atrogin1
A:
= PGC-1α
= Sirt1
= Pparγ
= COX4
= CS
= Utrn
= VEGF-a and –b
B:
↓ PGC-1α
↓ Sirt1
↓ Pparγ
= COX4
= CS
= Utrn
= VEGF-a and -b
A:
= TNFα
= Adipoq
= Adipor1
B:
= TNFα
= Adipoq
= Adipor1
A:
= Tubα
= NOX2
= c-Src
B:
= Tubα
= NOX2
= c-Src
A: (−)
B: (−)
Capogrosso, R.F., et al. (2017)
[86]
Male
4–5 weeks
MIT
A:
12 m/min,
30 min/day,
2 days/week,
4 weeks
B:
12 m/min,
30 min/day,
2 days/week,
12 to 20 weeks
DIA
EDL
GAST
TA
A:
↓ forelimbs strength
= twitch and tetanic tension
= torque
= resistance to eccentric contractions
= fatigue
in EDL
B:
↓ forelimbs strength
↓ twitch and tetanic tension
= torque
= resistance to eccentric contractions
= fatigue
↑ AChR1
↓ MHC1
in EDL
A:
= twitch and tetanic tension
= fatigue
B:
= twitch and tetanic tension
= fatigue
B:
↓ Fst in EDL
= pAMPK/ AMPK in TA
= CNF in GAST and DIA
= TGFβ1 in GAST
= MMP9 serum
B:
= Col 1
= ITGA7
= Eln
in EDL
B:
↑ PGC-1α
↓ Sirt1
= FIH1
= CD31
in EDL
B:
= Tubα
= NOX2
in EDL
A: (−) EDL, TA
(±) DIA
B: (−) EDL, TA
(±) DIA, GAST
De Luca, A., et al. (2003)
[90]
Male
3–4 weeks
MIT
12 m/min,
30 min/day,
2 days/week,
4–8 weeks
DIA
EDL
TA
↓ forelimb strength
↑ RM in EDL
= RM in DIA
↓ gCl in EDL
= gCl in DIA
↑ necrotic cell in TA ↑ inflammatory infiltrated cells in TA (−) EDL, TA
(±) DIA
De Luca, A., et al. (2005)
[96]
Male
4–5 weeks
MIT
12 m/min,
30 min/day,
2 days/week,
4–8 weeks
DIA
EDL, GAST
= body weight
↓ forelimb strength
↓ % fiber expressing slow MHC in EDL
↑ serum CK
= Utrn level in DIA
Trend ↑ TGFβ1 in GAST
Trend ↑ % area of connective tissue in GAST
(−) EDL, GAST
(±) DIA
Fraysse, B.,
et al. (2004)
[85]
3–4 weeksMIT
12 m/min,
30 min/day,
2 days/week,
4–8 weeks
EDL↓ forelimb strength
Any fast-to-slow fiber transition
↑ resting [Ca2+]i
↑ Sarcolemmal permeability
= CK
(−)
Gamberi, T.,
et al. (2018)
[95]
Male
4–5 weeks
MIT
12 m/min,
30 min/day,
2 days/week,
4 weeks
TA↑ Tnnt3
↑ Tnnt2
↑ MYOZ1
↑ Actin
↑ LDB3
↓ MYLPF
↓ TPM1
↓ MYL1
↑ CK ↑ ALDOA
↑ TPI1
↑ Eno3
↑ UGPR
↑ FH
↑ MDH2
↑ ALDH4A1
↑ complex II
↑ AK1
↓ Complex I, III
↓ NDUFB7
↓ ATP5a1
= Complex IV, V
= PGC1-α
= Sirt1
(−)
Hall, J.E.,
et al. (2007)
[126]
Male
4 weeks
Pre-training + MIT
Pre-training: workload progressive increase 8 to 12 m/min, 4 weeks.
Training: 12 m/min,
30 min/day, 6 weeks
EDL
GAST
QUAD
↓ serum CK (+)
Morales, M.G., et al. (2013)
[82]
Male
8 weeks
MIT
12 m/min,
30 min/day,
3 days/week,
6 months
GAST↓ Net Force ↑ degeneration↑ Col 3
↑ Fn
↑ CTGF
↑ TGFβ
↑ SMAD3 phosphorylation
(−)
Pessina, P., et al. (2014)
[81]
3–4–5 monthsMIT
12 m/min,
30 min/day
(with a rest of
5 minutes every 10 minutes of exercise),
3 days/week,
1, 2 or 3 months
GAST↓ tetanic force ↑ fibrosis
(↑ Col 1,
↑ Fn deposition, ↑ TGFβ1 and CTGF mRNA,
↑ SMAD2/3 protein level)
(−)
Radley-Crabb, H., et al. (2012)
[87]
Male
8–12 weeks
MIT
12 m/min,
30 min/day,
2 days/week,
4 weeks
DIA
EDL
GAST
QUAD
TA
Triceps
↓ forelimb grip strenght Necrosis:
DIA (↑
(0 min, ↑ 24 h,
= 96 h)
GAST (trend ↑ 0 min, trend ↑ 24 h, = 96 h)
QUAD (trend ↑ 0 min,
↑ 24 h,
= 96 h)
TA (= 0 min, ↑ 24 h, ↓ 96 h)
Triceps (= 0 min, ↑ 24 h, = 96 h)
↑ CK (24 h)
= CK (96 h)
↓ IL-1β (96 h)
↓ IL-6 (96 h)
↓ TNF (0 min)
= TNF (24 h)
in QUAD
↑ thiol oxidation
(0 min)
= thiol oxidation (24 h)
= MDA
in QUAD
(−) 0 min, 24 h
(±) 96 h
Rocco, A.B., et al. (2014)
[80]
Male
8 weeks
Pre-training +MIT
Pre-training:
5 m/min,
5 min/day, 3 days.
Training:
(5 min at 0 m/min + 2 min at 5 m/min + 8 min at 8 m/min + 30 min at 12 m/min)/day,
2 days/week, 4 weeks
DIA, GAST
Pectoralis
QUAD
TA
TB
= muscle mass ↑ Col in all muscles except DIA= VO2max
= time to exhaustion, = speed at
exhaustion
= total energy = VO2cv
= VO2min
(−) GAST,
TA,
QUAD,
TB,
Pectoralis
(±) DIA
Schill, K.E., et al. (2016)
[15]
Male and female
4 weeks
MIT
12 m/min,
30 min/day,
2 days/week,
4 weeks
Abdominal
Heart QUAD
↓ exhaustion times ↑ intracellular IgG in Heart
= intracellular IgG in QUAD
↑ hydroxyproline in QUAD and Heart
= hydroxyproline in Abdominal
↑ Col 1 in Heart
↓ basal oxygen consumption
= maximal
oxygen
consumption
↑ GSH oxidized by ROS in QUAD and Abdominal
Trend ↑ GSH oxidized by ROS in Heart
(−)
Van Putten, M., et al. (2012)
[16]
Female
4 weeks
MIT
12 m/min,
30 min/day,
3 days/week,
12 weeks
QUAD↑ CK↑ fibrosis
in QUAD
(−)
Vita, G.L., et al. (2020)
[106]
4–5 weeksMIT
12 m/min,
30 min/day,
2 days/week,
8–10 weeks
DIA, GAST
TA
= Minimum
telomere length
↑ TRF1 in TA
↓ TRF1 in DIA
= TRF1 in GAST
= PARP1 in GAST, TA and DIA
↑ MTERT protein level and activity in DIA
= MTERT in GAST and TA
(±)
B. Acute exercise
ReferenceSex/AgeProtocolMuscleHindlimb Muscle (Morphological and Functional Changes)Respiratory FunctionCardiac FunctionDegeneration
/Regeneration
FibrosisMetabolic Adaptive ChangesInflammatory MarkersOxidant
/Antioxidant Markers
Exercise Muscle Effects
Camerino, G.M.,
et al. (2014)
[88]
Male
16 weeks
MIT
12 m/min, 30 min
GAST= Mhc1
= Cn
= Serca2
= Fst = PGC-1α
= Sirt1
= PPARγ
(±)
Capogrosso, R.F., et al. (2017)
[86]
Male
16 weeks
MIT
12 m/min, 30 min
DIA
EDL
↓ twitch and tetanic tension
↓ TTP
↓ HRT
↑ resistance to eccentric contractions
= fatigue
↓ twitch and tetanic tension
= fatigue
(−)
Radley-Crabb, H.,
et al. (2012)
[87]
Male
8–12 weeks
MIT
12 m/min, 30 min
EDL
GAST
QUAD
TA
Triceps
↑ necrosis in QUAD (24 h, 48 h)
= necrosis in TA, EDL, Triceps and GAST
↑ CK (0 h)
= CK (24 h)
↑ IL-1β (2 h)
↑ IL-6
(0 min, 2 h)
= IL1β and
IL-6 (24 h),
↓ TNF
(0 min, 2 h, 24 h)
in QUAD
↑ thiol oxidation (0 min, 2 h)
= thiol oxidation (24 h)
in QUAD
(−) QUAD
(±) EDL, GAST, TA, Triceps
Terrill, J.R.,
et al. (2012)
[111]
Male
12 weeks
MIT
12 m/min, 30 min
GAST
QUAD
↑ necrosis
↑ CK
↑ thiol oxidation(−)
Abbreviations: (+)—Beneficial effects, (±)—No effects or discordant effects; (−)—Detrimental effects; AChR1—Acetylcholine receptor 1; Adipoq—Adiponectin; Adipor1—Adiponectin receptor 1; ALDH4A1—Delta 1 pyrroline 5 carboxilate dehydrogenase; ALDOA—Fructose-bisphosphate aldolase A; AK1—Adenilate kinase 1; ATP5a1—ATP synthase subunit alpha; BNIP3—BCL2/adenovirus E1B 19 kDa protein-interacting protein 3; CA3—Carbonic anhydrase 3; CCL2 (Monocyte chemoattractant protein-1); CD31—Cluster of differentiation 31; CK—Creatine kinase ; CLN—Centrally located nuclei; Cn—Calcineurin; CAT—Catalase; CNF—Centrally nucleated fibers, Col—Collagen; COX—Cytochrome c oxidase; CS—Citrate synthase; CSA—Cross sectional area; c-Src—cellular Src; CTGF—Connective tissue growth factor; Cx39—Connexin 39; DIA—Diaphragm; EDL—Extensor digitorum longus; Eln—Elastin; Eno3—Enolase 3; FH—Fumarate hydratase; FIH1—Hypoxia-inducible factor-1; Fn—Fibronectin; FRAP—Ferric reducing ability of plasma; Fst—Follistatin; GAST—Gastrocnemius; gCl—Chloride conductance; gK—Potassium conductance; GPX—glutathione peroxidase; GSH—Glutathione; HDAC5—Histone deacetylase 5; HRT—Half relaxation time; IGF1—Insuline growth factor 1; IL—Interleukin; ITGA7—Integrin Subunit Alpha 7; LC3—Microtubule-associated protein 1 light chain 3 alpha; LDB3—LIM domain binding protein 3; LIT—Low intensity training (speed of running <12 m/min); LVIDD—Left ventricle internal diameter during diastole; LVPWT—Left ventricle posterior wall tickness; MDA—Malondialdehyde; MDH2—Malate dehydrogenase; MEF2—myocyte enhancer factor 2; MHC—Myosin heavy chain; MIT—Moderate intensity training (speed of running ranging between 12 m/min and 15 m/min); MMP9—Matrix metallopeptidase 9; Mstn—Myostatin; MTERT—Mouse telomerase reverse transcriptase; MYL1—Myosin light chain 1/3; MYLPF– Myosin light chain, phosphorylatable, fast skeletal muscle; Myog—Myogenin; MYOZ1—Myozenin 1; MV—Minute Volume; NDUFB7—NADH dehydrogenase [ubiquinone] 1 beta subcomplex subunit 7; NF-kB- Nuclear factor-κB; NOX2—NADPH oxidase 2; OGDH—2-oxoglutarate dehydrogenase; pADP/OC—Phosphorilated Adenosine di-phosphate/oxigen consumption; PARP1—Poly [ADP-ribose] polymerase 1; PGC-1α- Peroxisome proliferator-activated receptor gamma coactivator 1-alpha; Pparγ—Peroxisome proliferator-activated receptor gamma; QUAD—Quadriceps; RCI—Respiratory control index; RM—Membrane resistance; ROS—Reactive oxygen species; SERCA—Sarco/endoplasmic reticulum Ca2+-ATPase; Sirt1—Sirtuin 1; SMAD- Small mother against decapentaplegic; SOD—Superoxide dismutase; SOL—Soleus; Svil—Supervillin; TA—Tibial Anterior; TB—Triceps brachii; TBARS—Thiobarbituric acid reactive substance; TGFβ1—Transforming growth factor beta 1; Ti—Inspiratory time; Te—Expiratory time; TNF—Tumor necrosis factor; Tnnt—Troponin; TPI1—Triosophosphate isomerase; TPM1—Tropomyosin alpha 1 chain; TRF1—Telomeric Repeat Factor 1; TTP—time-to-peak; Tubα—Tubulin α; UGPR—UTP glucose-1 phosphate uridyl transferase; Utrn—Utrophin; VEGF—Vascular endothelial growth factor; VO2cv—coefficient of variation for VO2; VO2max—maximum oxygen consumption; VO2min—Minimum oxygen consumption.
Table A2. Effects of downhill and uphill running on mdx mouse muscles.
Table A2. Effects of downhill and uphill running on mdx mouse muscles.
A. Chronic Exercise
ReferenceSex/AgeProtocolMuscleHindlimb Muscle (Morphological and Functional Changes)Respiratory
Function
Cardiac FunctionDegeneration/

Regeneration
FibrosisMetabolic Adaptive ChangesInflammatory MarkersOxidant/

Antioxidant Markers
Exercise
Muscle
Effects
Anderson, C.L.,
et al. (2006)
[93]
Male
7 weeks
−15°,
10 m/min,
10 min/day,
3 days
BB
DIA
EDL
GAST
SOL
TA
↑ damaged fibers in TA, BB and DIA (−) TA,
BB, DIA
Brussee, V.,
et al. (1997)
[25]
Male
8 weeks
−15°,
10 m/min,
10 min/day,
3 days
BB
DIA
EDL, GAST, SOL
TA
TB
↑ degenerating muscle fibers (−)
Cerri, D.G.,
et al. (2009)
[94]
Male
4 weeks
Downhill,
12 m/min,
20 min/day,
3 days/week,
5 weeks
DIA, GAST ↑ Gal-1 (−)
Fowler, W.M.,
et al. (1990)
[100]
3 weeks+18°,
4 m/min,
3 weeks
EDL, SOL↑ twitch tension
↑ rate of twitch tension development
↑ rate of twitch tension relaxation
↑ twitch/tetanus in SOL
↓ CLN
↓ necrosis
↓ fibers splitting
↓ moth-eaten fibers
(+)
Kobayashi, Y.M., et al. (2012)
[98]
Male
10 weeks
−15°,
5 min at
3 m/min +
10 min at
15 m/min,
2 weeks
EDL
GAST
QUAD
↓ strength in EDL ↑ myoglobinuria
= CK
↑ oedema and inflammation in QUAD and GAST (−)
Mokhtarian, A., et al. (1995)
[91]
14 days+15°,
4 m/min,
30 min/days,
7 days
EDL, SOL= ratio of type I and type II myofibers in SOL
= ratio of type IIa and type IIb
myofibers in EDL
↑ necrosis (−)
Nakamura, A., et al. (2002)
[118]
Male
6 weeks
+7°,
15 m/min,
60 min/day,
2 days/week,
5 weeks
+
23 m/min,
60 min/day,
2 days/week,
5 weeks
Heart ↑ heart/body weight↑ % area of dystrophic lesion
↑ pERK1/2
↑ p38 MAPK
↑ calcineurin
= pJNK1
↑ fibrotic area ↑ inflammatory cells (−)
Nakamura, A., et al. (2005)
[84]
Male
6 weeks
+7°,
15 m/min,
60 min/day,
2 days/week,
5 weeks
+
23 m/min,
60 min/day,
2 days/week,
5 weeks
GAST ↑ pERK1/2
↑ p38 MAPK
↑ JNK2
= JNK1
↑ MMP9 (−)
Okano, T.,
et al. (2005)
[89]
Male
6 weeks
+7°,
15 m/min,
60 min/day,
2 days/week,
5 weeks
+
23 m/min,
60 min/day,
2 days/week,
5 weeks
GAST
QUAD
SOL
TA
= body weight
= Serum albumin levels
↓ large-sized DRGs in SOL, GAST and QUAD
= small-sized DRGs
Trend ↓ medium-sized DRGs
↓ IGF1 mRNA in SOL, GAST and TA
↓ MyoD mRNA in SOL, GAST and TA
= Col 3 in SOL and GAST (−)
Taniguti, A.P.,
et al. (2011)
[83]
Male and female
6 months
−17°,
17 m/min,
60 min/day,
7 weeks
BB
DIA
Heart, TA
↓ grip strength and normalized
forelimb muscle
strength
= body mass
↑ serum CK↑ TGFβ1 in
BB and Heart
↑ fibrosis in TA and BB
= fibrosis in DIA and Heart
(−) TA,
BB, Heart
(±) DIA
B. Acute exercise
ReferenceSex/AgeProtocolMuscleHindlimb Muscle (Morphological and Functional Changes)Respiratory
Function
Cardiac FunctionDegeneration/

Regeneration
FibrosisMetabolic Adaptive ChangesInflammatory MarkersOxidant/

Antioxidant Markers
Exercise
Muscle
Effects
Clarke, M.S.F., et al. (1993)
[116]
12 weeks−16°,
0.6 m/min,
5 min
Triceps ↑ damaged myofibers ↓ bFGF (−)
Lindsay, A.,
et al. (2018)
[215]
Male
3 months
−10°,
workload progressive increase
10 to 15 m/min,
5 min
+
15 m/min,
30 min
↑ Isoxan
thopterin
(−)
Mathur, S.,
et al. (2011)
[115]
Male
5–12 months
−14°,
workload progressive increase
8 to 10 m/min,
45 min
Lower hindlimbs ↑ membrane breakdown
↑ T2
(−)
Quinlan, J.G.,
et al. (2006)
[112]
5–7 weeks−10°,
8 m/min,
10 min
+
workload progressive increase
8 to 16 m/min,
up to 90 min or fatigue
QUAD
SOL
TB
↑ membrane breakdown in QUAD (−)
Vilquin, J.T.,
et al.
(1998)
[113]
Male and female
9–14 months
−16°,
10 m/min,
5 min
↑ serum CK (1 h)
= serum CK
(3 days)
(−) 1 h
(±) 3 days
Whitehead, N.P., et al. (2006)
[114]
Male
7–10 weeks
−17°,
10 m/min,
45 min
EDL↓ isometric force ↑ membrane breakdown (−)
Abbreviations: (+)—Beneficial effects, (±)—No effects or discordant effects; (−)—Detrimental effects; BB—Biceps brachii; bFGF—basic fibroblast growth factor; CK—Creatine kinase; CLN—Centrally located nuclei; Col—Collagen; DIA—Diaphragm; DRG—Degenerative regenerative group; EDL—Extensor digitorum longus; Gal-1—Galectin 1; GAST—Gastrocnemius; IGF1—Insuline growth factor 1; MMP9—Matrix metallopeptidase 9; MyoD—Myoblast determination protein 1; pERK1/2—Phosphorylated extracellular signal-regulated kinase 1/2; pJKN—Phosphorylated c-Jun N-terminal kinase; p38 MAPK—p38 mitogen-activated protein kinases; QUAD—Quadriceps; SOL—Soleus; T2—Transverse relaxation time constant; TA—Tibial Anterior; TB—Triceps brachii; TGFβ1—Transforming growth factor beta 1.
Table A3. Effects of swimming on mdx mouse muscles.
Table A3. Effects of swimming on mdx mouse muscles.
A. Chronic Exercise
ReferenceSex/AgeProtocolMuscleHindlimb Muscle (Morphological and Functional Changes)Respiratory
Function
Cardiac FunctionDegeneration/

Regeneration
FibrosisMetabolic Adaptive ChangesInflammatory MarkersOxidant/

Antioxidant Markers
Exercise
Muscle
Effects
Hayes, A.,
et al. (1998)
[129]
Male
24 months
LIT
25 min/day,
4 or 5 days/week,
10 weeks
EDL
SOL
↑ relative tetanic tension
= absolute force
↓ RT1/2 in SOL
= % of type I, type IIa and type IIb fibers
(+)
Hyzewicz, J.,
et al. (2015)
[128]
Male
4 weeks
LIT
30 min/day,
4 days/week,
4 weeks
GAST↑ grip strength
↑ fast isoform of Tnnt
MyBP-C↑ CK (1 h)
= CK(2 h)
↑ respiratory chain protein
↑ UDPGP
↑ CA3
↓ carbonylated proteins (VDAC1,
fast isoform of Tnnt,
MyBP-C and
PGM1)
(+)
Hyzewicz, J.,
et al. (2017)
[127]
Male
4 weeks
LIT
30 min/day,
4 days/week,
4 weeks
GAST↑ grip strength = CK
= CNF
= IgG+ cells
↓ CCL2
↓ TIMP-1
= C5α
↑ CD68+/
CD11b+ cells
↑ F4/80+/
CD11+cells
Trend ↓ Ly6C+ / CD11b+ cells
Trend ↑ Ly6C+/CD11b-cells
↓ iNOS
↑ CD68
= CD206
= CD163
(+)
Matsakas, A.,
et al. (2013)
[132]
Male
6–8 weeks
LIT
30 min/day,
4 days
GAST
TA
CK Hypoxia in GAST and TA (−)
Barbin, I.C.C., et al. (2016)
[133]
Male
11 months
MIT
60 min/day,
6 days/week,
2 months
DIA
Heart
↑ wall thickness/lumen diameter of the pulmonary trunk degenerating cardio-
myocytes
fibrosis in DIA and Heart
↑ pro-MMP2, pre-MMP2 and active MMP2 in DIA and Heart
↑ TGFβ in Heart
(−)
Hayes, A.,
et al. (1993)
[130]
5 weeksPre-training + HIT
Pre-training:
workload progressive increase
5 min to 2 h/day
Training:
2 h/day,
5 days/week,
5 weeks
EDL
SOL
↓ muscle mass in EDL (↓ type IIa fiber area) and SOL (↓ type I and type IIa fiber area)
↑ RT1/2 in EDL
↑ twitch tension in SOL
↑ resistance to fatigue in EDL and SOL
(+)
Lynch, G.S.,
et al. (1993)
[131]
5 weeksPre training + HIT
Pre-training:
workload progressive increase
5 min to 2 h/day
Training:
2 h/day,
5 days/week,
5 weeks
EDL
SOL
↑ intermediate type fiber in SOL
↑ sensitivity to Sr2* in type IIa fiber in EDL
↓ sensitivity to Ca2+ and Sr2+ in EDL
type IIB fibers
↓ sensitivity to Ca2+ and Sr2+ in
SOL type IIA fibers
(+)
B. Acute exercise
ReferenceSex/AgeProtocolMuscleHindlimb muscle (Morphological and Functional Changes)Respiratory
Function
Cardiac FunctionDegeneration/

Regeneration
FibrosisMetabolic Adaptive ChangesInflammatory MarkersOxidant/

Antioxidant Markers
Exercise
Muscle
Effects
Bouchentouf, M., et al. (2006)
[134]
Male
2 months
LIT
20 min
TA fiber damage (−)
AbbAbreviations: (+)—Beneficial effects, (−)—Detrimental effects; C5α—Complement component C5 alpha; CA3—Carbonic Anhydrase 3; CCL2—chemokine ligand 2; CD—Clusters of differentiation; CK—creatine kinase; CNF—Centrally nucleated fibers; DIA—Diaphragm; EDL—Extensor digitorum longus; GAST—Gastrocnemius; HIT—High intensity training (≥90 min/day); IgG—immunoglobulin G; iNOS—Inducible Nitric Oxide Synthase; LIT—Low intensity training (20–59 min/day); Ly6—lymphocytes antigen 6; MyBP-C—myosin-binding protein C; MIT—Moderate intensity training (60–89 min/day); MMP2—Matrix metallopeptidase 2; PGM1—phosphoglucomutase-1; RT1/2—one-half relaxation time; SOL—Soleus; TA—Tibial Anterior; TGFβ—Transforming growth factor beta; TIMP1—tissue inhibitor of metalloproteinases-1; Tnnt—troponin; UDPGP—UTP-glucose-1-phosphate uridylyltransferase; VDAC1—voltage-dependent anion-selective channel protein 1.
Table A4. Effects of voluntary wheel running on mdx mouse muscles.
Table A4. Effects of voluntary wheel running on mdx mouse muscles.
A. Chronic Exercise
ReferenceSex/AgeProtocolMuscleHindlimb Muscle (Morphological and Functional Changes)Respiratory
Function
Cardiac FunctionDegeneration/
Regeneration
FibrosisMetabolic Adaptive ChangesInflammatory MarkersOxidant/
Antioxidant Markers
Exercise
Muscle
Effects
Baltgalvis, K.A.,
et al. (2012)
[141]
Male and female
4 weeks
12 weeks low-resistance wheel runningDIA
GAST
Heart, TA
= muscle mass
↑ maximal isometric torque in GAST
= concentric and passive torque in GAST
↑ fatigue resistance
↑ type IIx fiber in GAST
↑ Heart mass= CK
= CNF in TA and DIA
↓ CNF in GAST
↑ CS, β-HAD in GAST
= CS, β-HAD in TA
↑ COXIV in GAST
(+) GAST, Heart
(±) TA, DIA
Brereton, D., et al. (2012)
[153]
4 months4 weekserector spinae muscle ↑ Fibrosis
↑ Kyphosis
(−)
Bueno, Jr. C.R., et al. (2012)
[137]
Male
2 months
1 monthTB↑ aerobic capacity
↑ CSA
= Myoglobin
↓ ubiquitinated proteins
↑ CK
↑ p-AMPKα
(Thr172)
= AMPKα
↑ p-ACC/ACC
= Pparδ
= SCD
(+)
Call, J.A.,
et al. (2008)
[136]
Male
21 days
3 weeksEDL
GAST
Heart
SOL
↑ muscle mass in GAST
↑ weekly distance
↑ tetanic stress output in EDL
↑ active stifness in EDL
↑ total contractile protein in EDL
↑ actin in EDL
↑ myosin in EDL
shift to fatigue-resistant fibers
(↑ Type I fibers and ↓ Type IIb fibers) in EDL
↑ muscle mass in Heart ↑ CS in Heart
= CS in GAST and SOL
↑ β-HAD in GAST and Heart
= β-HAD in SOL
(+) EDL, GAST, Heart
(±) SOL
Call, J.A.,
et al. (2010)
[144]
Male
4–5 weeks
12 weeks of voluntary free wheel or resistance wheel runningEDL
GAST
Heart QUAD
SOL
TA
Triceps
↑ grip strength
= dorsiflexion torque and whole body tension
↑ specific tetanic force in SOL
↑ specific eccentric force (only free wheel) in SOL
= absolute twitch, tetanic or eccentric force in SOL
↑ relative mass in SOL and TA (only free wheel)
↑ relative mass in Triceps (only in Resist Wheel)
= relative mass in GAST and EDL
↑ β-dystroglycan in GAST
↑ Vinculin in SOL
= α-7 integrin in SOL, GAST and Triceps
= Talin in SOL, GAST and Triceps
= relative mass in Heart= CK (+)
(±) Heart
Coles, C.A.,
et al. (2020)
[149]
Male
9 weeks
3 weeksQUAD ↑ necrosis
↑ CNF
↑ Extracelluar matrix genes (i.e. Col 1a2, Col 3a1, Fn1, PCOLCE, etc.)
↓ EGLN1
↑ F4/80+ cell
↑ M1 macrophages (F4/80+/CD206-)
Trend ↑ M2a macrophages (CD206+/CD163-)
= M2c (CD206+/CD163+)
↑ inflammation related genes (i.e. CD14, IFi30, IL10rα, etc.)
↓ Pde4d
↑ Receptors recognising extracellular DAMPs (TLR2, TLR4, TLR6, P2XR4, etc.)
(+)
Costas, J.M.,
et al. (2010)
[160]
Female
7 weeks
4 weeksHeart = body mass
↓ LVLWT
↑ LVSLD
↑ fibrosis (−)
Delacroix, C.,
et al. (2018)
[135]
2–4 months1 weekEDL
GAST
Plantaris TA
↑muscle excitability in TA and EDL (↑ CMAP
↑ SLC8A1
↓ SCN4A
↓ Cacna1s)
↑ calcineurine pathway in TA and EDL (↑Rcan3,
↓ Cmya5,
↓ Myoz1
= Rcan1)
↑ fast to slow transition in TA
(↑ MHC2, MHC7, Tnni1)
= MHC3
↓ MHC4, Tnni2
↓ muscle fragility in TA and EDL
No effect in GAST and Plantaris
= Myog
in TA
= TGFβ
= Col 1a1 in TA
↑ TFAM
↑ Nrf1
↑ PGC-1β
= PGC-1α
= SDHB
= NDUF5
in TA
= TNF in TA (+) TA, EDL
(±) GAST, Plantaris
Dupont-Versteegden, E.E.,
et al. (1994)
[158]
3 weeks37–49 weeksDIA
SOL
↑ muscle/body weight in SOL
= fatigue in SOL
↑ active tension in DIA
↑ CT in DIA
= fatigue in DIA
(+)
Dupont-Versteegden, E.E.,
et al. (1996)
[139]
Male
3 weeks
9 (A) or 49 (B) weeksDIA
SOL
↑ muscle/body weight in SOL
(A and B)
↑ Myosin in SOL (B)
↑ active tension of DIA (B)
↑ contraction time of DIA (A)
↑ Fatigability of DIA (A)
= CK A: (+)
B: (+)
Ferry, A.,
et al. (2015)
[146]
Male and female
7 months
3 monthsHeart
TA
= force drop
= specific maximal force
↑ absolute maximal force (only in female)
↑ muscle weight (only in female)
= ejection and shortening fraction
= BNP
= MHC7
↑ ANF
= fibrosis in Heart
= Col 1a1
= Col 3a1
= CTFG
= TGFβ1
(+) TA
in female
(±) Heart
Gordon, B.S.,
et al. (2014)
[140]
Male and female
4 weeks
12 weeksQUAD
SOL
= CSA in QUAD ↑ CLN in QUAD
↑ Utrn in QUAD
Trend ↑ Utrn in SOL
Trend ↓ CD45+ (+)
Hayes, A.,
et al. (1996)
[142]
Male
4 weeks
16 weeksEDL
SOL
↑ absolute and relative muscle mass of SOL
↑ absolute twitch of SOL
↑ tetanic tension of SOL
↑ Fatigue resistance of EDL and SOL
↑ Type IIa fibers in EDL
↓ Type IIb fibers in EDL
↑ Type I fibers in SOL
↓ Type IIa fibers in SOL
(+) SOL, EDL
Hourdé, C.,
et al. (2013)
[143]
Female
4–5 weeks
4.5 monthsHeart
TA
↑ Absolute maximal force
↑ Muscle weight
↓ susceptibility to contraction induced-injury
= MHC1, MHC2x, MHC2b
↑ MHC2a mRNA but not protein
↓ ACTG1
ejection and shortening fraction
↓ LVPWT
↑ LVEDD
= regeneration
(= MHC, eMHC, IL6, Myog, MyoD, CyclinD1)
= fibrosis
(= Col 1α1, TGFβ1, PKC)
= BNIP3, LC3, Atrogin
= Mstn, Fst, REDD1, REDD2
= PGC1-α
= Sirt1
= SERCA2
= NFAT (+) TA
(−) Heart
Hulmi, J.J.,
et al. (2013)
[282]
Male
6–7 weeks
7 weeksEDL, GAST
MQF
SOL
TA
= weight mass = CK
in GAST
↑ PGC-1α
↑ LC3
↑ CS
↑ Cyt-C
↑ SDH
in GAST
= TNFα
in GAST
(+) GAST
Hulmi, J.J.,
et al. (2016)
[284]
Male
6–7 weeks
7 weeksGAST= grip strenght,
Shift to smaller fibers
↑ p-Sirt1
= pAMPKα at Thr172
= AMPK
= p-AMPK/AMPK
= PDI
= IRE1α
= pERK
= eIF2α
= p-eIF2α
= GRP78
= TxNIP
= TRX
= Hsp70
= Hsp60
= Hsp90
= Hsp47
= Hsp25
↑ protein carbonyl levels
↑ GSSG
↑ GSSG/GSH
↑ TPOR
= GSH, GRD, GST
= GPX
(−)
Hunt, L.C.,
et al. (2011)
[150]
Male
10 weeks
2 weeksQUAD ↑ necrosis
= LIFR
= LIF
= IL-6
(−)
Kogelman, B.,
et al. (2018)
[159]
Female
15 months
3–4 days/week,
8 weeks
Anterior and posterior hindlimb
Heart
QUAD
Left ventricular functions:
= Heart/body mass
= ejection fraction
= end systolic volume
= stroke volume
= cardiac output
Right ventricular functions:
↑ end systolic volume
Trend ↓ ejection fraction
= stroke volume
= cardiac output
= fibrosis
= Col 1a1
= Lox
= Lgals3
= CD68
(±)
Landisch, M.R.,
et al. (2008)
[138]
Male and female
4 weeks
8 weeksEDL
SOL
TA
↑ muscle mass in SOL
↑ type IIa fiber and ↓ type IIb fibers in EDL
↓ large fibers in SOL and EDL
= CNF in EDL and SOL
= eMHC in SOL
↑ CS activity in EDL
= CS activity in
SOL and TA
= CCO activity in EDL, SOL and TA
= β-HAD activity in TA
(+) EDL, SOL
(±) TA
Lim, J.H.,
et al. (2004)
[151]
28 days12 h/day,
8 weeks
SOL ↑ degeneration (↑ apoptotic myonuclei,
↑ Bax, ↓ Bcl2)
(−)
Selsby, J.T., et al. (2013)
[147]
Male
4 weeks
1 yearDIA
EDL
GAST
Heart
QUAD SOL
TA
↑ absolute mass of GAST and SOL
= absolute mass of EDL, TA, QUAD
= specific tension in EDL and soleus
↑ CSA and tetanic force in SOL
↓ specific tension↑ Heart mass
↑ LVDd and LVDs
↑ LVEDV
↑ LVESV
↑stroke volume
(+) SOL, GAST, Heart
(±) EDL, TA, QUAD
(−) DIA
Smythe, G.M.,
et al. (2012)
[148]
Male
10–12 weeks
2 weeksGAST QUAD= CSA ↑ necrosis in QUAD
↑ CNF in QUAD
(−)
Wineinger, M.A.,
et al. (1998)
[145]
6 months11 monthsEDL, SOL↑ muscle mass in SOL
↑ fatigue resistance in EDL
(+)
B. Acute exercise
ReferenceSex/AgeProtocolMuscleHindlimb Muscle (Morphological and Functional Changes)Respiratory
Function
Cardiac FunctionDegeneration/

Regeneration
FibrosisMetabolic Adaptive ChangesInflammatory MarkersOxidant/

Antioxidant Markers
Exercise
Muscle
Effects
Archer, J.D.,
et al. (2006)
[157]
4 weeks24 hDIA
GAST
QUAD
TA
↑ fiber damage in DIA, QUAD, GAST, TA
↑ Myf5 in QUAD
(−)
Sandri, M.,
et al. (1995)
[155]
4 weeks16 hEDL
SOL
TA
↑ DNA Fragmentation
↑ apoptotic myonuclei
↑ ubiquitinated protein
(−)
Sandri, M.,
et al. (1997)
[156]
4 weeks16 hTA ↑ DNA Fragmentation
↑ apoptotic myonuclei
↑ ubiquitinated protein
↓ Bcl2
= BAG1
= FASL
(−)
Abbreviations: (+)—Beneficial effects, (±)—No effects or discordant effects; (−)—Detrimental effects; ACC—Acetyl-CoA carboxylase ; ACTG1—Actin Gamma 1; AMPK—AMP-activated protein kinase; ANF—Atrial natriuretic factor; BAG1—BAG family molecular chaperone regulator 1; Bax—Bcl2 Associated X; Bcl2—B-cell lymphoma 2; β-HAD—Beta-hydroxy acyl-CoA dehydrogenase; BNIP3—BCL2/adenovirus E1B 19 kDa protein-interacting protein 3; BNP—Brain natriuretic peptide; CA3—Carbonic Anhydrase 3; Cacna1s—Calcium voltage-gated channel subunit alpha1 S; CCO—Cytochrome c oxidase; CD—Clusters of differentiation; CK—creatine kinase; CLN—Centrally located nuclei; CMAP—Compound muscle action potential; Cmya5—Cardiomyopathy Associated 5; CNF—Centrally nucleated fibers; Col—Collagen; COXIV—Mitochondrial cytochrome c oxidase subunit IV; CS—Citrate synthase; CSA—Cross-sectional area; CT—Contraction time; CTGF—Connective tissue growth factor; DAMPs—Damage-associated Molecular Patterns; Cyt-C—Cytochrome C; DIA—Diaphragm; EDL—Extensor digitorum longus; EGLN1—EGL nine homolog 1; EI2α—Eukaryotic initiation factor 2 subunit α; eMHC—Embryonic myosin heavy chain; FASL—FAS ligand; Fn1—Fibronectin 1; Fst—Follistatin; GAST—Gastrocnemius; GPX—Glutathione peroxidase; GRP78—Glucoseregulated protein 78; GRD—Glutathione reductase; GSH—Glutathione; GSSG—Oxidized glutathione; GST—Glutathione S-transferase; Hsp—Heat shock protein; IFi30—Interferon gamma inducible protein 30; IL-6—Interleukin 6; IL-10rα—Interleukin 10 receptor alpha; IRE1α—Inositol-requiring enzyme 1α; Lc3—Microtubule-associated protein 1 light chain 3 alpha; Lgals3—Galectin 3; LIF—Leukemia inhibitory factor; LIFR—Leukemia inhibitory factor receptor; Lox—Lysyl Oxidase; LVDd—Left Ventricle end-diastolic dimension; LVDs—Left Ventricle end-systolic dimension; LVEDD—Left Ventricle end systolic diameter; LVEDV—Left Ventricle end-diastolic volume; LVESV—Left Ventricle end-sistolic volume; LVLWT—Left ventricle lateral wall thickness; LVPWT—Left Ventricle posterior wall Thickness; LVSLD—Left ventricle Septolateral diameter; MQF—Musculus quadriceps femoris; MHC—Myosin heavy chain; MMP—Matrix metallopeptidase; Mstn—Myostatin; MyBP-C—Myosin-binding protein C; Myf5—Myogenic factor 5; MyoD—Myoblast determination protein; Myog—Myogenin; Myoz1—Myozenin 1; Nduf5—NADH dehydrogenase [ubiquinone] 1 alpha subcomplex assembly factor 5; NFAT—Nuclear Factor of Activated T-Cells; Nrf—Nuclear Respiratory Factor; P2XR4—Purinergic receptor P2X, ligand-gated ion channel 4; p-ACC—Phosphorilated Acetyl-CoA carboxylase; p-AMPK—Phosphorilated AMP-activated protein kinase; PCOLCE—Procollagen C-endopeptidase enhancer protein; Pde4d—Phosphodiesterase 4D, CAMP specific; PDI—Protein disulfide isomerase; p-ERK—Phosphorylated extracellular signal-regulated kinase; PGC-1—Peroxisome proliferator-activated receptor gamma coactivator 1; PGM1—Phosphoglucomutase-1; RT1/2—one-half relaxation time; PKC—Protein kinase C; Pparδ—Peroxisome proliferator-activated receptor delta; p-Sirt—Phosphorilated Sirtuin 1; Rcan—Calcipressina; REDD—Regulated in development and DNA damage responses; SCD—Stearoyl-CoA desaturase; Scn4α—Sodium Voltage-Gated Channel Alpha Subunit 4; SDH—succinate dehydrogenase; SDHB—Succinate dehydrogenase complex iron sulfur subunit B; SERCA—Sarco/endoplasmic reticulum Ca2+-ATPase; Sirt—Sirtuin 1; Slca81—Solute Carrier Family 8 Member A1; SOL—Soleus; TA—Tibial Anterior; TB—Triceps brachii; Tfam—Transcription Factor A, Mitochondrial; TGFβ—Transforming growth factor beta; TLR—Toll like receptor; TNF—Tumor necrosis factor; Tnn—Troponin I; Tnnt—Troponin T; TPOR1—Thiol protein oxidoreductase; TRX—Thioredoxin; TxNIP—thioredoxin-interacting protein; Utrn—Utrophin; VDAC1—Voltage-dependent anion-selective channel protein 1.
Table A5. Beneficial effects of low intensity treadmill running on mdx mice.
Table A5. Beneficial effects of low intensity treadmill running on mdx mice.
ReferenceSex/AgeProtocolMuscleLow Intensity Treadmill Outcomes
Fernandes, D.C., et al.
(2019)
[73]
11 weeks9 m/min,
30 min/day,
3 days/week,
60 days
GAST
TA
Reduces collagen deposition and induces a modulation of the redox status.
Fontana, S., et al.
(2015)
[110]
Male
8 weeks
Pre-training:
3.2 m/min,
15 to 30 min/day,
5 days/week,
2 weeks
Training:
workload progressive increase,
4 to 4.8 m/min,
30 to 60 min/day,
5 days/week,
4 weeks
QUADContributes to reduce cell degeneration process, by counteracting oxidative stress.
Frinchi, M., et al.
(2013)
[77]
Male
8 weeks
Pre-training:
3.2 m/min,
15 to 30 min/day,
5 days/week,
2 weeks
Training:
workload progressive increase,
4 to 4.8 m/min,
30 to 60 min/day,
5 days/week,
4 weeks
GAST
QUAD
Induces a strong beneficial effect on the degeneration-regeneration process.
Gaiad, T.P., et al
(2017)
[74]
Male
8 weeks
9 m/min,
30 min/day, 3 days/week,
60 days
TAInduces adaptations in extracellular matrix, increasing elasticity of dystrophic muscle tissue and delaying fibrosis deposition.
Hoepers, A., et al.
(2020)
[108]
Male
28 days
Pre training:
7 days at 4 m/min
Training:
6 or 9 m/min, 30 min/day,
2 days/week,
8 weeks
GASTReverses lipid and protein damage and increases the antioxidant activity.
Kaczor, J.J., et al.
(2007)
[109]
28 days9 m/min,
30 min/day,
2 days/week,
8 weeks
GAST
EDL
SOL
Decreases oxidative stress markers in white muscle.
Morici, G., et al.
(2017)
[76]
Male
8 weeks
Pre-training:
3.2 m/min,
15 to 30 min/day,
5 days/week,
2 weeks
Training:
workload progressive increase,
4 to 4.8 m/min,
30 to 60 min/day,
5 days/week,
4 weeks
DIA
GAST
QUAD
Induces a trend for regeneration areas to be larger than necrosis areas in diaphragm and modulates the inflammatory status of hindlimb muscle.
Pinto, P.A.F., et al.
(2018)
[75]
Male
11 weeks
9 m/min,
30 min/day, 3 days/week,
8 weeks
Lateral GASTReduces intramuscular fibrosis deposition and does not exacerbates markers of muscle injury.
Zelikovich, A.S., et al. (2019)
[79]
4–5 months4 m/min or
8 m/min, 30 min/day,
3 days/week,
6 months
(Each exercise session consisted of 7’30’’ warm up followed by a 22’30’’ training at target speed)
DIA
GAST
Heart
QUAD
TA
Improves tetanic and specific force in TA muscle, increases respiratory capacity, attenuates cardiac decline associated with disease progression, increases adiponectin and reduces adipocyte cross sectional area, and induces a modest increase in expression of the PGC-1α gene in the gastrocnemius muscle.
Abbreviations: DIA-Diaphragm; EDL-Extensor digitorum longus; GAST-Gastrocnemius; PGC-1α-Peroxisome proliferator-activated receptor gamma coactivator 1-alpha; QUAD-Quadriceps; SOL-Soleus; TA-Tibial Anterior.

References

  1. Verhaart, I.E.C.; Aartsma-Rus, A. Therapeutic developments for Duchenne muscular dystrophy. Nat. Rev. Neurol. 2019, 15, 373–386. [Google Scholar] [CrossRef] [PubMed]
  2. Shimizu-Motohashi, Y.; Miyatake, S.; Komaki, H.; Takeda, S.; Aoki, Y. Recent advances in innovative therapeutic approaches for Duchenne muscular dystrophy: From discovery to clinical trials. Am. J. Transl. Res. 2016, 8, 2471–2489. [Google Scholar] [PubMed]
  3. Ferraro, E.; Giammarioli, A.M.; Chiandotto, S.; Spoletini, I.; Rosano, G. Exercise-induced skeletal muscle remodeling and metabolic adaptation: Redox signaling and role of autophagy. Antioxid. Redox Signal. 2014, 21, 154–176. [Google Scholar] [CrossRef] [PubMed]
  4. Grange, R.W.; Call, J.A. Recommendations to define exercise prescription for Duchenne muscular dystrophy. Exerc. Sport Sci. Rev. 2007, 35, 12–17. [Google Scholar] [CrossRef]
  5. Markert, C.D.; Case, L.E.; Carter, G.T.; Furlong, P.A.; Grange, R.W. Exercise and Duchenne muscular dystrophy: Where we have been and where we need to go. Muscle Nerve 2012, 45, 746–751. [Google Scholar] [CrossRef]
  6. Franchi, M.V.; Reeves, N.D.; Narici, M.V. Skeletal Muscle Remodeling in Response to Eccentric vs. Concentric Loading: Morphological, Molecular, and Metabolic Adaptations. Front. Physiol. 2017, 8, 447. [Google Scholar] [CrossRef] [Green Version]
  7. McGlory, C.; Devries, M.C.; Phillips, S.M. Skeletal muscle and resistance exercise training; the role of protein synthesis in recovery and remodeling. J. Appl. Physiol. 2017, 122, 541–548. [Google Scholar] [CrossRef]
  8. Hyzewicz, J.; Ruegg, U.T.; Takeda, S. Comparison of Experimental Protocols of Physical Exercise for mdx Mice and Duchenne Muscular Dystrophy Patients. J. Neuromuscul. Dis. 2015, 2, 325–342. [Google Scholar] [CrossRef] [Green Version]
  9. Spaulding, H.R.; Selsby, J.T. Is Exercise the Right Medicine for Dystrophic Muscle? Med. Sci. Sports Exerc. 2018, 50, 1723–1732. [Google Scholar] [CrossRef]
  10. Gianola, S.; Pecoraro, V.; Lambiase, S.; Gatti, R.; Banfi, G.; Moja, L. Efficacy of muscle exercise in patients with muscular dystrophy: A systematic review showing a missed opportunity to improve outcomes. PLoS ONE 2013, 8, e65414. [Google Scholar] [CrossRef]
  11. Kostek, M. Precision Medicine and Exercise Therapy in Duchenne Muscular Dystrophy. Sports 2019, 7, 64. [Google Scholar] [CrossRef] [Green Version]
  12. Kostek, M.C.; Gordon, B. Exercise Is an Adjuvant to Contemporary Dystrophy Treatments. Exerc. Sport Sci. Rev. 2018, 46, 34–41. [Google Scholar] [CrossRef] [PubMed]
  13. Louboutin, J.P.; Fichter-Gagnepain, V.; Thaon, E.; Fardeau, M. Morphometric analysis of mdx diaphragm muscle fibres. Comparison with hindlimb muscles. Neuromuscul. Disord. 1993, 3, 463–469. [Google Scholar] [CrossRef]
  14. Crisafulli, S.; Sultana, J.; Fontana, A.; Salvo, F.; Messina, S.; Trifiro, G. Global epidemiology of Duchenne muscular dystrophy: An updated systematic review and meta-analysis. Orphanet J. Rare Dis. 2020, 15, 141. [Google Scholar] [CrossRef] [PubMed]
  15. Schill, K.E.; Altenberger, A.R.; Lowe, J.; Periasamy, M.; Villamena, F.A.; Rafael-Fortney, J.A.; Devor, S.T. Muscle damage, metabolism, and oxidative stress in mdx mice: Impact of aerobic running. Muscle Nerve 2016, 54, 110–117. [Google Scholar] [CrossRef] [PubMed]
  16. Van Putten, M.; Hulsker, M.; Nadarajah, V.D.; van Heiningen, S.H.; van Huizen, E.; van Iterson, M.; Admiraal, P.; Messemaker, T.; den Dunnen, J.T.; ’t Hoen, P.A.; et al. The effects of low levels of dystrophin on mouse muscle function and pathology. PLoS ONE 2012, 7, e31937. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Petrillo, S.; Pelosi, L.; Piemonte, F.; Travaglini, L.; Forcina, L.; Catteruccia, M.; Petrini, S.; Verardo, M.; D’Amico, A.; Musaro, A.; et al. Oxidative stress in Duchenne muscular dystrophy: Focus on the NRF2 redox pathway. Hum. Mol. Genet. 2017, 26, 2781–2790. [Google Scholar] [CrossRef]
  18. Grounds, M.D.; Terrill, J.R.; Al-Mshhdani, B.A.; Duong, M.N.; Radley-Crabb, H.G.; Arthur, P.G. Biomarkers for Duchenne muscular dystrophy: Myonecrosis, inflammation and oxidative stress. Dis. Model. Mech. 2020, 13. [Google Scholar] [CrossRef]
  19. Dhawan, J.; Rando, T.A. Stem cells in postnatal myogenesis: Molecular mechanisms of satellite cell quiescence, activation and replenishment. Trends Cell Biol. 2005, 15, 666–673. [Google Scholar] [CrossRef] [PubMed]
  20. Snijders, T.; Nederveen, J.P.; McKay, B.R.; Joanisse, S.; Verdijk, L.B.; van Loon, L.J.; Parise, G. Satellite cells in human skeletal muscle plasticity. Front. Physiol. 2015, 6, 283. [Google Scholar] [CrossRef] [Green Version]
  21. Kim, J.H.; Kwak, H.B.; Thompson, L.V.; Lawler, J.M. Contribution of oxidative stress to pathology in diaphragm and limb muscles with Duchenne muscular dystrophy. J. Muscle Res. Cell Motil. 2013, 34, 1–13. [Google Scholar] [CrossRef] [PubMed]
  22. Klingler, W.; Jurkat-Rott, K.; Lehmann-Horn, F.; Schleip, R. The role of fibrosis in Duchenne muscular dystrophy. Acta Myol. 2012, 31, 184–195. [Google Scholar] [PubMed]
  23. Smith, L.R.; Hammers, D.W.; Sweeney, H.L.; Barton, E.R. Increased collagen cross-linking is a signature of dystrophin-deficient muscle. Muscle Nerve 2016, 54, 71–78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Weller, B.; Karpati, G.; Carpenter, S. Dystrophin-deficient mdx muscle fibers are preferentially vulnerable to necrosis induced by experimental lengthening contractions. J. Neurol. Sci. 1990, 100, 9–13. [Google Scholar] [CrossRef]
  25. Brussee, V.; Tardif, F.; Tremblay, J.P. Muscle fibers of mdx mice are more vulnerable to exercise than those of normal mice. Neuromuscul. Disord. 1997, 7, 487–492. [Google Scholar] [CrossRef]
  26. Ricotti, V.; Ridout, D.A.; Muntoni, F. Steroids in Duchenne muscular dystrophy. Neuromuscul. Disord. 2013, 23, 696–697. [Google Scholar] [CrossRef] [PubMed]
  27. Falzarano, M.S.; Scotton, C.; Passarelli, C.; Ferlini, A. Duchenne Muscular Dystrophy: From Diagnosis to Therapy. Molecules 2015, 20, 18168–18184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Van Westering, T.L.; Betts, C.A.; Wood, M.J. Current understanding of molecular pathology and treatment of cardiomyopathy in duchenne muscular dystrophy. Molecules 2015, 20, 8823–8855. [Google Scholar] [CrossRef]
  29. Sun, C.; Shen, L.; Zhang, Z.; Xie, X. Therapeutic Strategies for Duchenne Muscular Dystrophy: An Update. Genes 2020, 11, 837. [Google Scholar] [CrossRef]
  30. Fairclough, R.J.; Wood, M.J.; Davies, K.E. Therapy for Duchenne muscular dystrophy: Renewed optimism from genetic approaches. Nat. Rev. Genet. 2013, 14, 373–378. [Google Scholar] [CrossRef]
  31. Iftikhar, M.; Frey, J.; Shohan, M.J.; Malek, S.; Mousa, S.A. Current and emerging therapies for Duchenne muscular dystrophy and spinal muscular atrophy. Pharmacol. Ther. 2020, 107719. [Google Scholar] [CrossRef]
  32. Voet, N.B.; van der Kooi, E.L.; van Engelen, B.G.; Geurts, A.C. Strength training and aerobic exercise training for muscle disease. Cochrane Database Syst. Rev. 2019, 12, CD003907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Yucel, N.; Chang, A.C.; Day, J.W.; Rosenthal, N.; Blau, H.M. Humanizing the mdx mouse model of DMD: The long and the short of it. NPJ Regen. Med. 2018, 3, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. McGreevy, J.W.; Hakim, C.H.; McIntosh, M.A.; Duan, D. Animal models of Duchenne muscular dystrophy: From basic mechanisms to gene therapy. Dis. Model. Mech. 2015, 8, 195–213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Van Putten, M.; Lloyd, E.M.; de Greef, J.C.; Raz, V.; Willmann, R.; Grounds, M.D. Mouse models for muscular dystrophies: An overview. Dis. Model. Mech. 2020, 13. [Google Scholar] [CrossRef] [Green Version]
  36. Bulfield, G.; Siller, W.G.; Wight, P.A.; Moore, K.J. X chromosome-linked muscular dystrophy (mdx) in the mouse. Proc. Natl. Acad. Sci. USA 1984, 81, 1189–1192. [Google Scholar] [CrossRef] [Green Version]
  37. Stedman, H.H.; Sweeney, H.L.; Shrager, J.B.; Maguire, H.C.; Panettieri, R.A.; Petrof, B.; Narusawa, M.; Leferovich, J.M.; Sladky, J.T.; Kelly, A.M. The mdx mouse diaphragm reproduces the degenerative changes of Duchenne muscular dystrophy. Nature 1991, 352, 536–539. [Google Scholar] [CrossRef]
  38. Chamberlain, J.S.; Metzger, J.; Reyes, M.; Townsend, D.; Faulkner, J.A. Dystrophin-deficient mdx mice display a reduced life span and are susceptible to spontaneous rhabdomyosarcoma. FASEB J. 2007, 21, 2195–2204. [Google Scholar] [CrossRef] [Green Version]
  39. Li, D.; Long, C.; Yue, Y.; Duan, D. Sub-physiological sarcoglycan expression contributes to compensatory muscle protection in mdx mice. Hum. Mol. Genet. 2009, 18, 1209–1220. [Google Scholar] [CrossRef] [Green Version]
  40. Collins, C.A.; Morgan, J.E. Duchenne’s muscular dystrophy: Animal models used to investigate pathogenesis and develop therapeutic strategies. Int. J. Exp. Pathol. 2003, 84, 165–172. [Google Scholar] [CrossRef]
  41. Duddy, W.; Duguez, S.; Johnston, H.; Cohen, T.V.; Phadke, A.; Gordish-Dressman, H.; Nagaraju, K.; Gnocchi, V.; Low, S.; Partridge, T. Muscular dystrophy in the mdx mouse is a severe myopathy compounded by hypotrophy, hypertrophy and hyperplasia. Skelet. Muscle 2015, 5, 16. [Google Scholar] [CrossRef] [Green Version]
  42. Torres, L.F.; Duchen, L.W. The mutant mdx: Inherited myopathy in the mouse: Morphological studies of nerves, muscles and end-plates. Brain 1987, 110(Pt. 2), 269–299. [Google Scholar] [CrossRef]
  43. Spencer, M.J.; Montecino-Rodriguez, E.; Dorshkind, K.; Tidball, J.G. Helper (CD4(+)) and cytotoxic (CD8(+)) T cells promote the pathology of dystrophin-deficient muscle. Clin. Immunol. 2001, 98, 235–243. [Google Scholar] [CrossRef] [PubMed]
  44. Sussman, M. Duchenne muscular dystrophy. J. Am. Acad. Orthop. Surg. 2002, 10, 138–151. [Google Scholar] [CrossRef] [PubMed]
  45. Hartel, J.V.; Granchelli, J.A.; Hudecki, M.S.; Pollina, C.M.; Gosselin, L.E. Impact of prednisone on TGF-beta1 and collagen in diaphragm muscle from mdx mice. Muscle Nerve 2001, 24, 428–432. [Google Scholar] [CrossRef]
  46. Dangain, J.; Vrbova, G. Muscle development in mdx mutant mice. Muscle Nerve 1984, 7, 700–704. [Google Scholar] [CrossRef]
  47. Coulton, G.R.; Curtin, N.A.; Morgan, J.E.; Partridge, T.A. The mdx mouse skeletal muscle myopathy: II. Contractile properties. Neuropathol. Appl. Neurobiol. 1988, 14, 299–314. [Google Scholar] [CrossRef]
  48. Coulton, G.R.; Morgan, J.E.; Partridge, T.A.; Sloper, J.C. The mdx mouse skeletal muscle myopathy: I. A histological, morphometric and biochemical investigation. Neuropathol. Appl. Neurobiol. 1988, 14, 53–70. [Google Scholar] [CrossRef]
  49. Bostick, B.; Yue, Y.; Long, C.; Duan, D. Prevention of dystrophin-deficient cardiomyopathy in twenty-one-month-old carrier mice by mosaic dystrophin expression or complementary dystrophin/utrophin expression. Circ. Res. 2008, 102, 121–130. [Google Scholar] [CrossRef] [Green Version]
  50. Bostick, B.; Yue, Y.; Long, C.; Marschalk, N.; Fine, D.M.; Chen, J.; Duan, D. Cardiac expression of a mini-dystrophin that normalizes skeletal muscle force only partially restores heart function in aged Mdx mice. Mol. Ther. 2009, 17, 253–261. [Google Scholar] [CrossRef]
  51. Hakim, C.H.; Grange, R.W.; Duan, D. The passive mechanical properties of the extensor digitorum longus muscle are compromised in 2- to 20-mo-old mdx mice. J. Appl. Physiol. 2011, 110, 1656–1663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Lefaucheur, J.P.; Pastoret, C.; Sebille, A. Phenotype of dystrophinopathy in old mdx mice. Anat. Rec. 1995, 242, 70–76. [Google Scholar] [CrossRef] [PubMed]
  53. Lynch, G.S.; Hinkle, R.T.; Chamberlain, J.S.; Brooks, S.V.; Faulkner, J.A. Force and power output of fast and slow skeletal muscles from mdx mice 6-28 months old. J. Physiol. 2001, 535, 591–600. [Google Scholar] [CrossRef] [PubMed]
  54. Pastoret, C.; Sebille, A. mdx mice show progressive weakness and muscle deterioration with age. J. Neurol. Sci. 1995, 129, 97–105. [Google Scholar] [CrossRef]
  55. Fluck, M. Functional, structural and molecular plasticity of mammalian skeletal muscle in response to exercise stimuli. J. Exp. Biol. 2006, 209, 2239–2248. [Google Scholar] [CrossRef] [Green Version]
  56. Liguori, G. ACSM’s Guidelines for Exercise Testing and Prescription; ACSM: Indianapolis, IN, USA, 2021. [Google Scholar]
  57. Jetté, M.; Sidney, K.; Blümchen, G. Metabolic equivalents (METS) in exercise testing, exercise prescription, and evaluation of functional capacity. Clin. Cardiol. 1990, 13, 555–565. [Google Scholar] [CrossRef]
  58. Ainsworth, B.E.; Haskell, W.L.; Herrmann, S.D.; Meckes, N.; Bassett, D.R., Jr.; Tudor-Locke, C.; Greer, J.L.; Vezina, J.; Whitt-Glover, M.C.; Leon, A.S. 2011 Compendium of Physical Activities: A second update of codes and MET values. Med. Sci. Sports Exerc. 2011, 43, 1575–1581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Pette, D. The adaptive potential of skeletal muscle fibers. Can. J. Appl. Physiol. 2002, 27, 423–448. [Google Scholar] [CrossRef]
  60. Fluck, M.; Hoppeler, H. Molecular basis of skeletal muscle plasticity—From gene to form and function. Rev. Physiol. Biochem. Pharmacol. 2003, 146, 159–216. [Google Scholar]
  61. Qaisar, R.; Bhaskaran, S.; Van Remmen, H. Muscle fiber type diversification during exercise and regeneration. Free Radic. Biol. Med. 2016, 98, 56–67. [Google Scholar] [CrossRef]
  62. Egan, B.; Zierath, J.R. Exercise metabolism and the molecular regulation of skeletal muscle adaptation. Cell Metab. 2013, 17, 162–184. [Google Scholar] [CrossRef] [Green Version]
  63. Kim, Y.J.; Kim, H.J.; Lee, W.J.; Seong, J.K. A comparison of the metabolic effects of treadmill and wheel running exercise in mouse model. Lab. Anim. Res. 2020, 36, 3. [Google Scholar] [CrossRef] [Green Version]
  64. Dehghani, M.; Kargarfard, M.; Rabiee, F.; Nasr-Esfahani, M.H.; Ghaedi, K. A comparative study on the effects of acute and chronic downhill running vs uphill running exercise on the RNA levels of the skeletal muscles PGC1-α, FNDC5 and the adipose UCP1 in BALB/c mice. Gene 2018, 679, 369–376. [Google Scholar] [CrossRef]
  65. Vernillo, G.; Giandolini, M.; Edwards, W.B.; Morin, J.B.; Samozino, P.; Horvais, N.; Millet, G.Y. Biomechanics and Physiology of Uphill and Downhill Running. Sports Med. 2017, 47, 615–629. [Google Scholar] [CrossRef]
  66. Kemi, O.J.; Loennechen, J.P.; Wisloff, U.; Ellingsen, O. Intensity-controlled treadmill running in mice: Cardiac and skeletal muscle hypertrophy. J. Appl. Physiol. 2002, 93, 1301–1309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Guo, S.; Huang, Y.; Zhang, Y.; Huang, H.; Hong, S.; Liu, T. Impacts of exercise interventions on different diseases and organ functions in mice. J. Sport Health Sci. 2020, 9, 53–73. [Google Scholar] [CrossRef]
  68. Yin, X.; Zhao, Y.; Zheng, Y.L.; Wang, J.Z.; Li, W.; Lu, Q.J.; Huang, Q.N.; Zhang, C.Y.; Chen, X.; Ma, J.Z. Time-Course Responses of Muscle-Specific MicroRNAs Following Acute Uphill or Downhill Exercise in Sprague-Dawley Rats. Front. Physiol. 2019, 10, 1275. [Google Scholar] [CrossRef] [PubMed]
  69. Cai, Z.-Y.; Hsu, C.-C.; Su, C.-P.; Lin, C.-F.; Lin, Y.-A.; Lin, C.-L.; Hsu, M.-C. Comparison of lower limb muscle activation during downhill, level and uphill running. Isokinet. Exerc. Sci. 2010, 18, 163–168. [Google Scholar] [CrossRef]
  70. Zhou, L.; Lu, H. Targeting fibrosis in Duchenne muscular dystrophy. J. Neuropathol. Exp. Neurol. 2010, 69, 771–776. [Google Scholar] [CrossRef]
  71. Kharraz, Y.; Guerra, J.; Pessina, P.; Serrano, A.L.; Munoz-Canoves, P. Understanding the process of fibrosis in Duchenne muscular dystrophy. Biomed. Res. Int. 2014, 2014, 965631. [Google Scholar] [CrossRef]
  72. Mahdy, M.A.A. Skeletal muscle fibrosis: An overview. Cell Tissue Res. 2019, 375, 575–588. [Google Scholar] [CrossRef]
  73. Fernandes, D.C.; Cardoso-Nascimento, J.J.A.; Garcia, B.C.C.; Costa, K.B.; Rocha-Vieira, E.; Oliveira, M.X.; Machado, A.S.D.; Santos, A.P.; Gaiad, T.P. Low intensity training improves redox status and reduces collagen fibers on dystrophic muscle. J. Exerc. Rehabil. 2019, 15, 213–223. [Google Scholar] [CrossRef]
  74. Gaiad, T.P.; Oliveira, M.X.; Lobo, A.R., Jr.; Liborio, L.R.; Pinto, P.A.F.; Fernandes, D.C.; Santos, A.P.; Ambrosio, C.E.; Machado, A.S.D. Low-intensity training provokes adaptive extracellular matrix turnover of a muscular dystrophy model. J. Exerc. Rehabil. 2017, 13, 693–703. [Google Scholar] [CrossRef] [Green Version]
  75. Pinto, P.A.F.; Machado, A.S.D.; Libório, L.R.; Santos, A.P.; Oliveira, M.X.; Gaiad, T.P. Low Intensity Training Provokes Adaptations on Muscle Fibrosis of a Muscular Dystrophy Model. Int. J. Morphol. 2018, 36, 471–477. [Google Scholar] [CrossRef] [Green Version]
  76. Morici, G.; Frinchi, M.; Pitruzzella, A.; Di Liberto, V.; Barone, R.; Pace, A.; Di Felice, V.; Belluardo, N.; Cappello, F.; Mudo, G.; et al. Mild Aerobic Exercise Training Hardly Affects the Diaphragm of mdx Mice. J. Cell. Physiol. 2017, 232, 2044–2052. [Google Scholar] [CrossRef]
  77. Frinchi, M.; Macaluso, F.; Licciardi, A.; Perciavalle, V.; Coco, M.; Belluardo, N.; Morici, G.; Mudo, G. Recovery of damaged skeletal muscle in mdx mice through low-intensity endurance exercise. Int. J. Sports Med. 2014, 35, 19–27. [Google Scholar] [CrossRef] [PubMed]
  78. Zeman, R.J.; Peng, H.; Danon, M.J.; Etlinger, J.D. Clenbuterol reduces degeneration of exercised or aged dystrophic (mdx) muscle. Muscle Nerve 2000, 23, 521–528. [Google Scholar] [CrossRef]
  79. Zelikovich, A.S.; Quattrocelli, M.; Salamone, I.M.; Kuntz, N.L.; McNally, E.M. Moderate exercise improves function and increases adiponectin in the mdx mouse model of muscular dystrophy. Sci. Rep. 2019, 9, 5770. [Google Scholar] [CrossRef]
  80. Rocco, A.B.; Levalley, J.C.; Eldridge, J.A.; Marsh, S.A.; Rodgers, B.D. A novel protocol for assessing exercise performance and dystropathophysiology in the mdx mouse. Muscle Nerve 2014, 50, 541–548. [Google Scholar] [CrossRef]
  81. Pessina, P.; Cabrera, D.; Morales, M.G.; Riquelme, C.A.; Gutierrez, J.; Serrano, A.L.; Brandan, E.; Munoz-Canoves, P. Novel and optimized strategies for inducing fibrosis in vivo: Focus on Duchenne Muscular Dystrophy. Skelet. Muscle 2014, 4, 7. [Google Scholar] [CrossRef] [Green Version]
  82. Morales, M.G.; Cabrera, D.; Cespedes, C.; Vio, C.P.; Vazquez, Y.; Brandan, E.; Cabello-Verrugio, C. Inhibition of the angiotensin-converting enzyme decreases skeletal muscle fibrosis in dystrophic mice by a diminution in the expression and activity of connective tissue growth factor (CTGF/CCN-2). Cell Tissue Res. 2013, 353, 173–187. [Google Scholar] [CrossRef]
  83. Taniguti, A.P.; Pertille, A.; Matsumura, C.Y.; Santo Neto, H.; Marques, M.J. Prevention of muscle fibrosis and myonecrosis in mdx mice by suramin, a TGF-beta1 blocker. Muscle Nerve 2011, 43, 82–87. [Google Scholar] [CrossRef] [PubMed]
  84. Nakamura, A.; Yoshida, K.; Ueda, H.; Takeda, S.; Ikeda, S. Up-regulation of mitogen activated protein kinases in mdx skeletal muscle following chronic treadmill exercise. Biochim. Biophys. Acta 2005, 1740, 326–331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Fraysse, B.; Liantonio, A.; Cetrone, M.; Burdi, R.; Pierno, S.; Frigeri, A.; Pisoni, M.; Camerino, C.; De Luca, A. The alteration of calcium homeostasis in adult dystrophic mdx muscle fibers is worsened by a chronic exercise in vivo. Neurobiol. Dis. 2004, 17, 144–154. [Google Scholar] [CrossRef]
  86. Capogrosso, R.F.; Mantuano, P.; Cozzoli, A.; Sanarica, F.; Massari, A.M.; Conte, E.; Fonzino, A.; Giustino, A.; Rolland, J.F.; Quaranta, A.; et al. Contractile efficiency of dystrophic mdx mouse muscle: In vivo and ex vivo assessment of adaptation to exercise of functional end points. J. Appl. Physiol. 2017, 122, 828–843. [Google Scholar] [CrossRef]
  87. Radley-Crabb, H.; Terrill, J.; Shavlakadze, T.; Tonkin, J.; Arthur, P.; Grounds, M. A single 30 min treadmill exercise session is suitable for ‘proof-of concept studies’ in adult mdx mice: A comparison of the early consequences of two different treadmill protocols. Neuromuscul. Disord. 2012, 22, 170–182. [Google Scholar] [CrossRef]
  88. Camerino, G.M.; Cannone, M.; Giustino, A.; Massari, A.M.; Capogrosso, R.F.; Cozzoli, A.; De Luca, A. Gene expression in mdx mouse muscle in relation to age and exercise: Aberrant mechanical-metabolic coupling and implications for pre-clinical studies in Duchenne muscular dystrophy. Hum. Mol. Genet. 2014, 23, 5720–5732. [Google Scholar] [CrossRef] [Green Version]
  89. Okano, T.; Yoshida, K.; Nakamura, A.; Sasazawa, F.; Oide, T.; Takeda, S.; Ikeda, S. Chronic exercise accelerates the degeneration-regeneration cycle and downregulates insulin-like growth factor-1 in muscle of mdx mice. Muscle Nerve 2005, 32, 191–199. [Google Scholar] [CrossRef] [PubMed]
  90. De Luca, A.; Pierno, S.; Liantonio, A.; Cetrone, M.; Camerino, C.; Fraysse, B.; Mirabella, M.; Servidei, S.; Ruegg, U.T.; Conte Camerino, D. Enhanced dystrophic progression in mdx mice by exercise and beneficial effects of taurine and insulin-like growth factor-1. J. Pharmacol. Exp. Ther. 2003, 304, 453–463. [Google Scholar] [CrossRef]
  91. Mokhtarian, A.; Lefaucheur, J.P.; Even, P.C.; Sebille, A. Effects of treadmill exercise and high-fat feeding on muscle degeneration in mdx mice at the time of weaning. Clin. Sci. 1995, 89, 447–452. [Google Scholar] [CrossRef]
  92. Burdi, R.; Rolland, J.F.; Fraysse, B.; Litvinova, K.; Cozzoli, A.; Giannuzzi, V.; Liantonio, A.; Camerino, G.M.; Sblendorio, V.; Capogrosso, R.F.; et al. Multiple pathological events in exercised dystrophic mdx mice are targeted by pentoxifylline: Outcome of a large array of in vivo and ex vivo tests. J. Appl. Physiol. 2009, 106, 1311–1324. [Google Scholar] [CrossRef] [Green Version]
  93. Anderson, C.L.; De Repentigny, Y.; Cifelli, C.; Marshall, P.; Renaud, J.M.; Worton, R.G.; Kothary, R. The mouse dystrophin muscle promoter/enhancer drives expression of mini-dystrophin in transgenic mdx mice and rescues the dystrophy in these mice. Mol. Ther. 2006, 14, 724–734. [Google Scholar] [CrossRef]
  94. Cerri, D.G.; Rodrigues, L.C.; Stowell, S.R.; Araujo, D.D.; Coelho, M.C.; Oliveira, S.R.; Bizario, J.C.; Cummings, R.D.; Dias-Baruffi, M.; Costa, M.C. Degeneration of dystrophic or injured skeletal muscles induces high expression of Galectin-1. Glycobiology 2008, 18, 842–850. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Gamberi, T.; Fiaschi, T.; Valocchia, E.; Modesti, A.; Mantuano, P.; Rolland, J.F.; Sanarica, F.; De Luca, A.; Magherini, F. Proteome analysis in dystrophic mdx mouse muscle reveals a drastic alteration of key metabolic and contractile proteins after chronic exercise and the potential modulation by anti-oxidant compounds. J. Proteom. 2018, 170, 43–58. [Google Scholar] [CrossRef] [PubMed]
  96. De Luca, A.; Nico, B.; Liantonio, A.; Didonna, M.P.; Fraysse, B.; Pierno, S.; Burdi, R.; Mangieri, D.; Rolland, J.F.; Camerino, C.; et al. A multidisciplinary evaluation of the effectiveness of cyclosporine a in dystrophic mdx mice. Am. J. Pathol. 2005, 166, 477–489. [Google Scholar] [CrossRef] [Green Version]
  97. Bizario, J.C.; Cerri, D.G.; Rodrigues, L.C.; Oliveira, G.L.; Nomizo, A.; de Araujo, D.D.; Fukuhara, P.S.; Ribeiro, J.C.; de Castro, F.A.; Costa, M.C. Imatinib mesylate ameliorates the dystrophic phenotype in exercised mdx mice. J. Neuroimmunol. 2009, 212, 93–101. [Google Scholar] [CrossRef]
  98. Kobayashi, Y.M.; Rader, E.P.; Crawford, R.W.; Campbell, K.P. Endpoint measures in the mdx mouse relevant for muscular dystrophy pre-clinical studies. Neuromuscul. Disord. 2012, 22, 34–42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Burdi, R.; Didonna, M.P.; Pignol, B.; Nico, B.; Mangieri, D.; Rolland, J.F.; Camerino, C.; Zallone, A.; Ferro, P.; Andreetta, F.; et al. First evaluation of the potential effectiveness in muscular dystrophy of a novel chimeric compound, BN 82270, acting as calpain-inhibitor and anti-oxidant. Neuromuscul. Disord. 2006, 16, 237–248. [Google Scholar] [CrossRef]
  100. Fowler, W.M., Jr.; Abresch, R.T.; Larson, D.B.; Sharman, R.B.; Entrikin, R.K. High-repetitive submaximal treadmill exercise training: Effect on normal and dystrophic mice. Arch. Phys. Med. Rehabil. 1990, 71, 552–557. [Google Scholar]
  101. Shadrach, J.L.; Wagers, A.J. Stem cells for skeletal muscle repair. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2011, 366, 2297–2306. [Google Scholar] [CrossRef]
  102. Schmidt, M.; Schuler, S.C.; Huttner, S.S.; von Eyss, B.; von Maltzahn, J. Adult stem cells at work: Regenerating skeletal muscle. Cell. Mol. Life Sci. 2019, 76, 2559–2570. [Google Scholar] [CrossRef] [Green Version]
  103. Sun, C.; Serra, C.; Lee, G.; Wagner, K.R. Stem cell-based therapies for Duchenne muscular dystrophy. Exp. Neurol. 2020, 323, 113086. [Google Scholar] [CrossRef]
  104. Judson, R.N.; Rossi, F.M.V. Towards stem cell therapies for skeletal muscle repair. NPJ Regen. Med. 2020, 5, 10. [Google Scholar] [CrossRef]
  105. Lindsay, A.; Larson, A.A.; Verma, M.; Ervasti, J.M.; Lowe, D.A. Isometric resistance training increases strength and alters histopathology of dystrophin-deficient mouse skeletal muscle. J. Appl. Physiol. 2019, 126, 363–375. [Google Scholar] [CrossRef]
  106. Vita, G.L.; Aguennouz, M.; Sframeli, M.; Sanarica, F.; Mantuano, P.; Oteri, R.; Polito, F.; Licata, N.; Romeo, S.; Distefano, M.G.; et al. Effect of exercise on telomere length and telomere proteins expression in mdx mice. Mol. Cell. Biochem. 2020, 470, 189–197. [Google Scholar] [CrossRef]
  107. Lund, T.C.; Grange, R.W.; Lowe, D.A. Telomere shortening in diaphragm and tibialis anterior muscles of aged mdx mice. Muscle Nerve 2007, 36, 387–390. [Google Scholar] [CrossRef] [PubMed]
  108. Hoepers, A.; Alberti, A.; Freiberger, V.; Ventura, L.; Grigollo, L.R.; Andreu, C.S.; da Silva, B.B.; Martins, D.F.; Junior, R.J.N.; Streck, E.L.; et al. Effect of Aerobic Physical Exercise in an Animal Model of Duchenne Muscular Dystrophy. J. Mol. Neurosci. 2020, 70, 1552–1564. [Google Scholar] [CrossRef]
  109. Kaczor, J.J.; Hall, J.E.; Payne, E.; Tarnopolsky, M.A. Low intensity training decreases markers of oxidative stress in skeletal muscle of mdx mice. Free Radic. Biol. Med. 2007, 43, 145–154. [Google Scholar] [CrossRef] [PubMed]
  110. Fontana, S.; Schillaci, O.; Frinchi, M.; Giallombardo, M.; Morici, G.; Di Liberto, V.; Alessandro, R.; De Leo, G.; Perciavalle, V.; Belluardo, N.; et al. Reduction in mdx mouse muscle degeneration by low-intensity endurance exercise: A proteomic analysis in quadriceps muscle of exercised compared with sedentary mdx mice. Biosci. Rep. 2015, 35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Terrill, J.R.; Radley-Crabb, H.G.; Grounds, M.D.; Arthur, P.G. N-Acetylcysteine treatment of dystrophic mdx mice results in protein thiol modifications and inhibition of exercise induced myofibre necrosis. Neuromuscul. Disord. 2012, 22, 427–434. [Google Scholar] [CrossRef]
  112. Quinlan, J.G.; Wong, B.L.; Niemeier, R.T.; McCullough, A.S.; Levin, L.; Emanuele, M. Poloxamer 188 failed to prevent exercise-induced membrane breakdown in mdx skeletal muscle fibers. Neuromuscul. Disord. 2006, 16, 855–864. [Google Scholar] [CrossRef]
  113. Vilquin, J.T.; Brussee, V.; Asselin, I.; Kinoshita, I.; Gingras, M.; Tremblay, J.P. Evidence of mdx mouse skeletal muscle fragility in vivo by eccentric running exercise. Muscle Nerve 1998, 21, 567–576. [Google Scholar] [CrossRef]
  114. Whitehead, N.P.; Streamer, M.; Lusambili, L.I.; Sachs, F.; Allen, D.G. Streptomycin reduces stretch-induced membrane permeability in muscles from mdx mice. Neuromuscul. Disord. 2006, 16, 845–854. [Google Scholar] [CrossRef]
  115. Mathur, S.; Vohra, R.S.; Germain, S.A.; Forbes, S.; Bryant, N.D.; Vandenborne, K.; Walter, G.A. Changes in muscle T2 and tissue damage after downhill running in mdx mice. Muscle Nerve 2011, 43, 878–886. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Clarke, M.S.; Khakee, R.; McNeil, P.L. Loss of cytoplasmic basic fibroblast growth factor from physiologically wounded myofibers of normal and dystrophic muscle. J. Cell Sci. 1993, 106(Pt. 1), 121–133. [Google Scholar]
  117. Poche, H.; Hopfenmuller, W.; Hoffmann, M. Detection and identification of myoglobin in serum by immunoblotting. Effect of exercise on patients with Duchenne muscular dystrophy. Clin. Physiol. Biochem. 1987, 5, 103–111. [Google Scholar]
  118. Nakamura, A.; Yoshida, K.; Takeda, S.; Dohi, N.; Ikeda, S. Progression of dystrophic features and activation of mitogen-activated protein kinases and calcineurin by physical exercise, in hearts of mdx mice. FEBS Lett. 2002, 520, 18–24. [Google Scholar] [CrossRef] [Green Version]
  119. Hermes, T.A.; Kido, L.A.; Macedo, A.B.; Mizobuti, D.S.; Moraes, L.H.R.; Somazz, M.C.; Cagnon, V.H.A.; Minatel, E. Sex influences diaphragm muscle response in exercised mdx mice. Cell Biol. Int. 2018, 42, 1611–1621. [Google Scholar] [CrossRef]
  120. Grounds, M.D.; Radley, H.G.; Lynch, G.S.; Nagaraju, K.; De Luca, A. Towards developing standard operating procedures for pre-clinical testing in the mdx mouse model of Duchenne muscular dystrophy. Neurobiol. Dis. 2008, 31, 1–19. [Google Scholar] [CrossRef] [Green Version]
  121. Tiidus, P.M. Estrogen and gender effects on muscle damage, inflammation, and oxidative stress. Can. J. Appl. Physiol. 2000, 25, 274–287. [Google Scholar] [CrossRef]
  122. Tiidus, P.M. Oestrogen and sex influence on muscle damage and inflammation: Evidence from animal models. Curr. Opin. Clin. Nutr. Metab. Care 2001, 4, 509–513. [Google Scholar] [CrossRef]
  123. Tiidus, P.M. Influence of estrogen on skeletal muscle damage, inflammation, and repair. Exerc. Sport Sci. Rev. 2003, 31, 40–44. [Google Scholar] [CrossRef] [PubMed]
  124. Enns, D.L.; Tiidus, P.M. Estrogen influences satellite cell activation and proliferation following downhill running in rats. J. Appl. Physiol. 2008, 104, 347–353. [Google Scholar] [CrossRef] [Green Version]
  125. Velders, M.; Schleipen, B.; Fritzemeier, K.H.; Zierau, O.; Diel, P. Selective estrogen receptor-beta activation stimulates skeletal muscle growth and regeneration. FASEB J. 2012, 26, 1909–1920. [Google Scholar] [CrossRef]
  126. Hall, J.E.; Kaczor, J.J.; Hettinga, B.P.; Isfort, R.J.; Tarnopolsky, M.A. Effects of a CRF2R agonist and exercise on mdx and wildtype skeletal muscle. Muscle Nerve 2007, 36, 336–341. [Google Scholar] [CrossRef]
  127. Hyzewicz, J.; Tanihata, J.; Kuraoka, M.; Nitahara-Kasahara, Y.; Beylier, T.; Ruegg, U.T.; Vater, A.; Takeda, S. Low-Intensity Training and the C5a Complement Antagonist NOX-D21 Rescue the mdx Phenotype through Modulation of Inflammation. Am. J. Pathol. 2017, 187, 1147–1161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Hyzewicz, J.; Tanihata, J.; Kuraoka, M.; Ito, N.; Miyagoe-Suzuki, Y.; Takeda, S. Low intensity training of mdx mice reduces carbonylation and increases expression levels of proteins involved in energy metabolism and muscle contraction. Free Radic. Biol. Med. 2015, 82, 122–136. [Google Scholar] [CrossRef] [Green Version]
  129. Hayes, A.; Williams, D.A. Contractile function and low-intensity exercise effects of old dystrophic (mdx) mice. Am. J. Physiol. 1998, 274, C1138–C1144. [Google Scholar] [CrossRef]
  130. Hayes, A.; Lynch, G.S.; Williams, D.A. The effects of endurance exercise on dystrophic mdx mice: I. Contractile and histochemical properties of intact muscles. Proc. Biol. Sci. 1993, 253, 19–25. [Google Scholar]
  131. Lynch, G.S.; Hayes, A.; Lam, M.H.; Williams, D.A. The effects of endurance exercise on dystrophic mdx mice: II. Contractile properties of skinned muscle fibres. Proc. Biol. Sci. 1993, 253, 27–33. [Google Scholar]
  132. Matsakas, A.; Yadav, V.; Lorca, S.; Narkar, V. Muscle ERRgamma mitigates Duchenne muscular dystrophy via metabolic and angiogenic reprogramming. FASEB J. 2013, 27, 4004–4016. [Google Scholar] [CrossRef] [PubMed]
  133. Barbin, I.C.; Pereira, J.A.; Bersan Rovere, M.; de Oliveira Moreira, D.; Marques, M.J.; Santo Neto, H. Diaphragm degeneration and cardiac structure in mdx mouse: Potential clinical implications for Duchenne muscular dystrophy. J. Anat. 2016, 228, 784–791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Bouchentouf, M.; Benabdallah, B.F.; Mills, P.; Tremblay, J.P. Exercise improves the success of myoblast transplantation in mdx mice. Neuromuscul. Disord. 2006, 16, 518–529. [Google Scholar] [CrossRef] [PubMed]
  135. Delacroix, C.; Hyzewicz, J.; Lemaitre, M.; Friguet, B.; Li, Z.; Klein, A.; Furling, D.; Agbulut, O.; Ferry, A. Improvement of Dystrophic Muscle Fragility by Short-Term Voluntary Exercise through Activation of Calcineurin Pathway in mdx Mice. Am. J. Pathol. 2018, 188, 2662–2673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Call, J.A.; Voelker, K.A.; Wolff, A.V.; McMillan, R.P.; Evans, N.P.; Hulver, M.W.; Talmadge, R.J.; Grange, R.W. Endurance capacity in maturing mdx mice is markedly enhanced by combined voluntary wheel running and green tea extract. J. Appl. Physiol. 2008, 105, 923–932. [Google Scholar] [CrossRef] [Green Version]
  137. Bueno Jr, C.R.; Pantaleao, L.C.; Voltarelli, V.A.; Bozi, L.H.; Brum, P.C.; Zatz, M. Combined effect of AMPK/PPAR agonists and exercise training in mdx mice functional performance. PLoS ONE 2012, 7, e45699. [Google Scholar] [CrossRef] [Green Version]
  138. Landisch, R.M.; Kosir, A.M.; Nelson, S.A.; Baltgalvis, K.A.; Lowe, D.A. Adaptive and nonadaptive responses to voluntary wheel running by mdx mice. Muscle Nerve 2008, 38, 1290–1303. [Google Scholar] [CrossRef] [Green Version]
  139. Dupont-Versteegden, E.E. Exercise and clenbuterol as strategies to decrease the progression of muscular dystrophy in mdx mice. J. Appl. Physiol. 1996, 80, 734–741. [Google Scholar] [CrossRef]
  140. Gordon, B.S.; Lowe, D.A.; Kostek, M.C. Exercise increases utrophin protein expression in the mdx mouse model of Duchenne muscular dystrophy. Muscle Nerve 2014, 49, 915–918. [Google Scholar] [CrossRef]
  141. Baltgalvis, K.A.; Call, J.A.; Cochrane, G.D.; Laker, R.C.; Yan, Z.; Lowe, D.A. Exercise training improves plantar flexor muscle function in mdx mice. Med. Sci. Sports Exerc. 2012, 44, 1671–1679. [Google Scholar] [CrossRef] [Green Version]
  142. Hayes, A.; Williams, D.A. Beneficial effects of voluntary wheel running on the properties of dystrophic mouse muscle. J. Appl. Physiol. 1996, 80, 670–679. [Google Scholar] [CrossRef]
  143. Hourde, C.; Joanne, P.; Medja, F.; Mougenot, N.; Jacquet, A.; Mouisel, E.; Pannerec, A.; Hatem, S.; Butler-Browne, G.; Agbulut, O.; et al. Voluntary physical activity protects from susceptibility to skeletal muscle contraction-induced injury but worsens heart function in mdx mice. Am. J. Pathol. 2013, 182, 1509–1518. [Google Scholar] [CrossRef]
  144. Call, J.A.; McKeehen, J.N.; Novotny, S.A.; Lowe, D.A. Progressive resistance voluntary wheel running in the mdx mouse. Muscle Nerve 2010, 42, 871–880. [Google Scholar] [CrossRef] [Green Version]
  145. Wineinger, M.A.; Abresch, R.T.; Walsh, S.A.; Carter, G.T. Effects of aging and voluntary exercise on the function of dystrophic muscle from mdx mice. Am. J. Phys. Med. Rehabil. 1998, 77, 20–27. [Google Scholar] [CrossRef]
  146. Ferry, A.; Benchaouir, R.; Joanne, P.; Peat, R.A.; Mougenot, N.; Agbulut, O.; Butler-Browne, G. Effect of voluntary physical activity initiated at age 7 months on skeletal hindlimb and cardiac muscle function in mdx mice of both genders. Muscle Nerve 2015, 52, 788–794. [Google Scholar] [CrossRef]
  147. Selsby, J.T.; Acosta, P.; Sleeper, M.M.; Barton, E.R.; Sweeney, H.L. Long-term wheel running compromises diaphragm function but improves cardiac and plantarflexor function in the mdx mouse. J. Appl. Physiol. 2013, 115, 660–666. [Google Scholar] [CrossRef] [Green Version]
  148. Smythe, G.M.; White, J.D. Voluntary wheel running in dystrophin-deficient (mdx) mice: Relationships between exercise parameters and exacerbation of the dystrophic phenotype. PLoS Curr. 2011, 3, RRN1295. [Google Scholar] [CrossRef]
  149. Coles, C.A.; Gordon, L.; Hunt, L.C.; Webster, T.; Piers, A.T.; Kintakas, C.; Woodman, K.; Touslon, S.L.; Smythe, G.M.; White, J.D.; et al. Expression profiling in exercised mdx suggests a role for extracellular proteins in the dystrophic muscle immune response. Hum. Mol. Genet. 2020, 29, 353–368. [Google Scholar] [CrossRef]
  150. Hunt, L.C.; Coles, C.A.; Gorman, C.M.; Tudor, E.M.; Smythe, G.M.; White, J.D. Alterations in the expression of leukemia inhibitory factor following exercise: Comparisons between wild-type and mdx muscles. PLoS Curr. 2011, 3, RRN1277. [Google Scholar] [CrossRef]
  151. Lim, J.H.; Kim, D.Y.; Bang, M.S. Effects of exercise and steroid on skeletal muscle apoptosis in the mdx mouse. Muscle Nerve 2004, 30, 456–462. [Google Scholar] [CrossRef]
  152. Podhorska-Okolow, M.; Sandri, M.; Zampieri, S.; Brun, B.; Rossini, K.; Carraro, U. Apoptosis of myofibres and satellite cells: Exercise-induced damage in skeletal muscle of the mouse. Neuropathol. Appl. Neurobiol. 1998, 24, 518–531. [Google Scholar] [CrossRef] [PubMed]
  153. Brereton, D.; Plochocki, J.; An, D.; Costas, J.; Simons, E. The effects of glucocorticoid and voluntary exercise treatment on the development of thoracolumbar kyphosis in dystrophin-deficient mice. PLoS Curr. 2012, 4, e4ffdff160de168b. [Google Scholar] [CrossRef] [PubMed]
  154. DiMario, J.X.; Uzman, A.; Strohman, R.C. Fiber regeneration is not persistent in dystrophic (MDX) mouse skeletal muscle. Dev. Biol. 1991, 148, 314–321. [Google Scholar] [CrossRef]
  155. Sandri, M.; Carraro, U.; Podhorska-Okolov, M.; Rizzi, C.; Arslan, P.; Monti, D.; Franceschi, C. Apoptosis, DNA damage and ubiquitin expression in normal and mdx muscle fibers after exercise. FEBS Lett. 1995, 373, 291–295. [Google Scholar] [CrossRef]
  156. Sandri, M.; Podhorska-Okolow, M.; Geromel, V.; Rizzi, C.; Arslan, P.; Franceschi, C.; Carraro, U. Exercise induces myonuclear ubiquitination and apoptosis in dystrophin-deficient muscle of mice. J. Neuropathol. Exp. Neurol. 1997, 56, 45–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Archer, J.D.; Vargas, C.C.; Anderson, J.E. Persistent and improved functional gain in mdx dystrophic mice after treatment with L-arginine and deflazacort. FASEB J. 2006, 20, 738–740. [Google Scholar] [CrossRef]
  158. Dupont-Versteegden, E.E.; McCarter, R.J.; Katz, M.S. Voluntary exercise decreases progression of muscular dystrophy in diaphragm of mdx mice. J. Appl. Physiol. 1994, 77, 1736–1741. [Google Scholar] [CrossRef] [PubMed]
  159. Kogelman, B.; Putker, K.; Hulsker, M.; Tanganyika-de Winter, C.; van der Weerd, L.; Aartsma-Rus, A.; van Putten, M. Voluntary exercise improves muscle function and does not exacerbate muscle and heart pathology in aged Duchenne muscular dystrophy mice. J. Mol. Cell. Cardiol. 2018, 125, 29–38. [Google Scholar] [CrossRef] [Green Version]
  160. Costas, J.M.; Nye, D.J.; Henley, J.B.; Plochocki, J.H. Voluntary exercise induces structural remodeling in the hearts of dystrophin-deficient mice. Muscle Nerve 2010, 42, 881–885. [Google Scholar] [CrossRef]
  161. Chinet, A.E.; Even, P.C.; Decrouy, A. Dystrophin-dependent efficiency of metabolic pathways in mouse skeletal muscles. Experientia 1994, 50, 602–605. [Google Scholar] [CrossRef]
  162. Van Bennekom, C.A.; Oerlemans, F.T.; Kulakowski, S.; De Bruyn, C.H. Enzymes of purine metabolism in muscle specimens from patients with Duchenne-type muscular dystrophy. Adv. Exp. Med. Biol. 1984, 165(Pt. B), 447–450. [Google Scholar]
  163. Ge, Y.; Molloy, M.P.; Chamberlain, J.S.; Andrews, P.C. Proteomic analysis of mdx skeletal muscle: Great reduction of adenylate kinase 1 expression and enzymatic activity. Proteomics 2003, 3, 1895–1903. [Google Scholar] [CrossRef]
  164. Kuznetsov, A.V.; Winkler, K.; Wiedemann, F.R.; von Bossanyi, P.; Dietzmann, K.; Kunz, W.S. Impaired mitochondrial oxidative phosphorylation in skeletal muscle of the dystrophin-deficient mdx mouse. Mol. Cell. Biochem. 1998, 183, 87–96. [Google Scholar] [CrossRef]
  165. Rybalka, E.; Timpani, C.A.; Cooke, M.B.; Williams, A.D.; Hayes, A. Defects in mitochondrial ATP synthesis in dystrophin-deficient mdx skeletal muscles may be caused by complex I insufficiency. PLoS ONE 2014, 9, e115763. [Google Scholar] [CrossRef] [Green Version]
  166. Hughes, M.C.; Ramos, S.V.; Turnbull, P.C.; Rebalka, I.A.; Cao, A.; Monaco, C.M.F.; Varah, N.E.; Edgett, B.A.; Huber, J.S.; Tadi, P.; et al. Early myopathy in Duchenne muscular dystrophy is associated with elevated mitochondrial H2O2 emission during impaired oxidative phosphorylation. J. Cachexia Sarcopenia Muscle 2019, 10, 643–661. [Google Scholar] [CrossRef] [Green Version]
  167. Gaglianone, R.B.; Bloise, F.F.; Ortiga-Carvalho, T.M.; Quirico-Santos, T.; Costa, M.L.; Mermelstein, C. Comparative study of calcium and calcium-related enzymes with differentiation markers in different ages and muscle types in mdx mice. Histol. Histopathol. 2020, 35, 203–216. [Google Scholar]
  168. Blanchet, E.; Annicotte, J.S.; Pradelli, L.A.; Hugon, G.; Matecki, S.; Mornet, D.; Rivier, F.; Fajas, L. E2F transcription factor-1 deficiency reduces pathophysiology in the mouse model of Duchenne muscular dystrophy through increased muscle oxidative metabolism. Hum. Mol. Genet. 2012, 21, 3910–3917. [Google Scholar] [CrossRef]
  169. Strakova, J.; Kamdar, F.; Kulhanek, D.; Razzoli, M.; Garry, D.J.; Ervasti, J.M.; Bartolomucci, A.; Townsend, D. Integrative effects of dystrophin loss on metabolic function of the mdx mouse. Sci. Rep. 2018, 8, 13624. [Google Scholar] [CrossRef] [Green Version]
  170. Cole, M.A.; Rafael, J.A.; Taylor, D.J.; Lodi, R.; Davies, K.E.; Styles, P. A quantitative study of bioenergetics in skeletal muscle lacking utrophin and dystrophin. Neuromuscul. Disord. 2002, 12, 247–257. [Google Scholar] [CrossRef]
  171. Percival, J.M.; Whitehead, N.P.; Adams, M.E.; Adamo, C.M.; Beavo, J.A.; Froehner, S.C. Sildenafil reduces respiratory muscle weakness and fibrosis in the mdx mouse model of Duchenne muscular dystrophy. J. Pathol. 2012, 228, 77–87. [Google Scholar] [CrossRef] [Green Version]
  172. Tidball, J.G.; Welc, S.S.; Wehling-Henricks, M. Immunobiology of Inherited Muscular Dystrophies. Compr. Physiol. 2018, 8, 1313–1356. [Google Scholar] [PubMed]
  173. Droge, W. Free radicals in the physiological control of cell function. Physiol. Rev. 2002, 82, 47–95. [Google Scholar] [CrossRef] [PubMed]
  174. Powers, S.K.; Ji, L.L.; Kavazis, A.N.; Jackson, M.J. Reactive oxygen species: Impact on skeletal muscle. Compr. Physiol. 2011, 1, 941–969. [Google Scholar] [PubMed] [Green Version]
  175. Powers, S.K.; Talbert, E.E.; Adhihetty, P.J. Reactive oxygen and nitrogen species as intracellular signals in skeletal muscle. J. Physiol. 2011, 589, 2129–2138. [Google Scholar] [CrossRef] [PubMed]
  176. Jackson, M.J. On the mechanisms underlying attenuated redox responses to exercise in older individuals: A hypothesis. Free Radic. Biol. Med. 2020, 161, 326–338. [Google Scholar] [CrossRef]
  177. Scicchitano, B.M.; Pelosi, L.; Sica, G.; Musaro, A. The physiopathologic role of oxidative stress in skeletal muscle. Mech. Ageing Dev. 2018, 170, 37–44. [Google Scholar] [CrossRef]
  178. Disatnik, M.H.; Dhawan, J.; Yu, Y.; Beal, M.F.; Whirl, M.M.; Franco, A.A.; Rando, T.A. Evidence of oxidative stress in mdx mouse muscle: Studies of the pre-necrotic state. J. Neurol. Sci. 1998, 161, 77–84. [Google Scholar] [CrossRef]
  179. Whitehead, N.P.; Yeung, E.W.; Allen, D.G. Muscle damage in mdx (dystrophic) mice: Role of calcium and reactive oxygen species. Clin. Exp. Pharmacol. Physiol. 2006, 33, 657–662. [Google Scholar] [CrossRef]
  180. Shkryl, V.M.; Martins, A.S.; Ullrich, N.D.; Nowycky, M.C.; Niggli, E.; Shirokova, N. Reciprocal amplification of ROS and Ca(2+) signals in stressed mdx dystrophic skeletal muscle fibers. Pflügers Arch. 2009, 458, 915–928. [Google Scholar] [CrossRef]
  181. Renjini, R.; Gayathri, N.; Nalini, A.; Srinivas Bharath, M.M. Oxidative damage in muscular dystrophy correlates with the severity of the pathology: Role of glutathione metabolism. Neurochem. Res. 2012, 37, 885–898. [Google Scholar] [CrossRef]
  182. Guiraud, S.; Davies, K.E. Pharmacological advances for treatment in Duchenne muscular dystrophy. Curr. Opin. Pharmacol. 2017, 34, 36–48. [Google Scholar] [CrossRef] [PubMed]
  183. Choi, M.H.; Ow, J.R.; Yang, N.D.; Taneja, R. Oxidative Stress-Mediated Skeletal Muscle Degeneration: Molecules, Mechanisms, and Therapies. Oxid. Med. Cell. Longev. 2016, 2016, 6842568. [Google Scholar] [CrossRef] [Green Version]
  184. Kar, N.C.; Pearson, C.M. Catalase, superoxide dismutase, glutathione reductase and thiobarbituric acid-reactive products in normal and dystrophic human muscle. Clin. Chim. Acta 1979, 94, 277–280. [Google Scholar] [CrossRef]
  185. Rodriguez, M.C.; Tarnopolsky, M.A. Patients with dystrophinopathy show evidence of increased oxidative stress. Free Radic. Biol. Med. 2003, 34, 1217–1220. [Google Scholar] [CrossRef]
  186. Almeida-Becerril, T.; Rodriguez-Cruz, M.; Sanchez-Gonzalez, J.R.; Villaldama-Soriano, M.A.; Atilano-Miguel, S.; Villa-Morales, J.; Cardenas-Conejo, A.; Cardenas-Vazquez, R. Circulating markers of oxidative stress are associated with a muscle injury in patients with muscular dystrophy Duchenne. Brain Dev. 2021, 43, 111–120. [Google Scholar] [CrossRef] [PubMed]
  187. Souza, L.B.; Maziero, C.; Lazzarin, M.C.; Quintana, H.T.; Tome, T.C.; Baptista, V.I.A.; de Oliveira, F. Presence of metalloproteinases 2 and 9 and 8-OHdG in the fibrotic process in skeletal muscle of Mdx mice. Acta Histochem. 2020, 122, 151458. [Google Scholar] [CrossRef]
  188. El-Shafey, A.F.; Armstrong, A.E.; Terrill, J.R.; Grounds, M.D.; Arthur, P.G. Screening for increased protein thiol oxidation in oxidatively stressed muscle tissue. Free Radic. Res. 2011, 45, 991–999. [Google Scholar] [CrossRef]
  189. Terrill, J.R.; Radley-Crabb, H.G.; Iwasaki, T.; Lemckert, F.A.; Arthur, P.G.; Grounds, M.D. Oxidative stress and pathology in muscular dystrophies: Focus on protein thiol oxidation and dysferlinopathies. FEBS J. 2013, 280, 4149–4164. [Google Scholar] [CrossRef]
  190. Timpani, C.A.; Hayes, A.; Rybalka, E. Revisiting the dystrophin-ATP connection: How half a century of research still implicates mitochondrial dysfunction in Duchenne Muscular Dystrophy aetiology. Med. Hypotheses 2015, 85, 1021–1033. [Google Scholar] [CrossRef]
  191. Percival, J.M.; Siegel, M.P.; Knowels, G.; Marcinek, D.J. Defects in mitochondrial localization and ATP synthesis in the mdx mouse model of Duchenne muscular dystrophy are not alleviated by PDE5 inhibition. Hum. Mol. Genet. 2013, 22, 153–167. [Google Scholar] [CrossRef] [Green Version]
  192. Zhao, R.Z.; Jiang, S.; Zhang, L.; Yu, Z.B. Mitochondrial electron transport chain, ROS generation and uncoupling (Review). Int. J. Mol. Med. 2019, 44, 3–15. [Google Scholar] [CrossRef] [Green Version]
  193. Allen, D.G.; Whitehead, N.P.; Froehner, S.C. Absence of Dystrophin Disrupts Skeletal Muscle Signaling: Roles of Ca2+, Reactive Oxygen Species, and Nitric Oxide in the Development of Muscular Dystrophy. Physiol. Rev. 2016, 96, 253–305. [Google Scholar] [CrossRef] [Green Version]
  194. Vila, M.C.; Rayavarapu, S.; Hogarth, M.W.; Van der Meulen, J.H.; Horn, A.; Defour, A.; Takeda, S.; Brown, K.J.; Hathout, Y.; Nagaraju, K.; et al. Mitochondria mediate cell membrane repair and contribute to Duchenne muscular dystrophy. Cell Death Differ. 2017, 24, 330–342. [Google Scholar] [CrossRef] [Green Version]
  195. Dubinin, M.V.; Talanov, E.Y.; Tenkov, K.S.; Starinets, V.S.; Mikheeva, I.B.; Sharapov, M.G.; Belosludtsev, K.N. Duchenne muscular dystrophy is associated with the inhibition of calcium uniport in mitochondria and an increased sensitivity of the organelles to the calcium-induced permeability transition. Biochim. Biophys. Acta Mol. Basis Dis. 2020, 1866, 165674. [Google Scholar] [CrossRef]
  196. Ramadasan-Nair, R.; Gayathri, N.; Mishra, S.; Sunitha, B.; Mythri, R.B.; Nalini, A.; Subbannayya, Y.; Harsha, H.C.; Kolthur-Seetharam, U.; Srinivas Bharath, M.M. Mitochondrial alterations and oxidative stress in an acute transient mouse model of muscle degeneration: Implications for muscular dystrophy and related muscle pathologies. J. Biol. Chem. 2014, 289, 485–509. [Google Scholar] [CrossRef] [Green Version]
  197. Zorov, D.B.; Juhaszova, M.; Sollott, S.J. Mitochondrial reactive oxygen species (ROS) and ROS-induced ROS release. Physiol. Rev. 2014, 94, 909–950. [Google Scholar] [CrossRef] [Green Version]
  198. Menazza, S.; Blaauw, B.; Tiepolo, T.; Toniolo, L.; Braghetta, P.; Spolaore, B.; Reggiani, C.; Di Lisa, F.; Bonaldo, P.; Canton, M. Oxidative stress by monoamine oxidases is causally involved in myofiber damage in muscular dystrophy. Hum. Mol. Genet. 2010, 19, 4207–4215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Moore, T.M.; Lin, A.J.; Strumwasser, A.R.; Cory, K.; Whitney, K.; Ho, T.; Ho, T.; Lee, J.L.; Rucker, D.H.; Nguyen, C.Q.; et al. Mitochondrial Dysfunction Is an Early Consequence of Partial or Complete Dystrophin Loss in mdx Mice. Front. Physiol. 2020, 11, 690. [Google Scholar] [CrossRef] [PubMed]
  200. Hughes, M.C.; Ramos, S.V.; Turnbull, P.C.; Edgett, B.A.; Huber, J.S.; Polidovitch, N.; Schlattner, U.; Backx, P.H.; Simpson, J.A.; Perry, C.G.R. Impairments in left ventricular mitochondrial bioenergetics precede overt cardiac dysfunction and remodelling in Duchenne muscular dystrophy. J. Physiol. 2020, 598, 1377–1392. [Google Scholar] [CrossRef] [PubMed]
  201. Jahnke, V.E.; Van Der Meulen, J.H.; Johnston, H.K.; Ghimbovschi, S.; Partridge, T.; Hoffman, E.P.; Nagaraju, K. Metabolic remodeling agents show beneficial effects in the dystrophin-deficient mdx mouse model. Skelet Muscle 2012, 2, 16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Selsby, J.T.; Morine, K.J.; Pendrak, K.; Barton, E.R.; Sweeney, H.L. Rescue of dystrophic skeletal muscle by PGC-1alpha involves a fast to slow fiber type shift in the mdx mouse. PLoS ONE 2012, 7, e30063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Chan, M.C.; Rowe, G.C.; Raghuram, S.; Patten, I.S.; Farrell, C.; Arany, Z. Post-natal induction of PGC-1alpha protects against severe muscle dystrophy independently of utrophin. Skelet Muscle 2014, 4, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Brandes, R.P.; Weissmann, N.; Schroder, K. Nox family NADPH oxidases in mechano-transduction: Mechanisms and consequences. Antioxid. Redox Signal. 2014, 20, 887–898. [Google Scholar] [CrossRef] [Green Version]
  205. Powers, S.K.; Jackson, M.J. Exercise-induced oxidative stress: Cellular mechanisms and impact on muscle force production. Physiol. Rev. 2008, 88, 1243–1276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Ferreira, L.F.; Laitano, O. Regulation of NADPH oxidases in skeletal muscle. Free Radic. Biol. Med. 2016, 98, 18–28. [Google Scholar] [CrossRef] [Green Version]
  207. Sakellariou, G.K.; Jackson, M.J.; Vasilaki, A. Redefining the major contributors to superoxide production in contracting skeletal muscle. The role of NAD(P)H oxidases. Free Radic. Res. 2014, 48, 12–29. [Google Scholar] [CrossRef]
  208. Ward, C.W.; Prosser, B.L.; Lederer, W.J. Mechanical stretch-induced activation of ROS/RNS signaling in striated muscle. Antioxid. Redox Signal. 2014, 20, 929–936. [Google Scholar] [CrossRef]
  209. Cherednichenko, G.; Zima, A.V.; Feng, W.; Schaefer, S.; Blatter, L.A.; Pessah, I.N. NADH oxidase activity of rat cardiac sarcoplasmic reticulum regulates calcium-induced calcium release. Circ. Res. 2004, 94, 478–486. [Google Scholar] [CrossRef] [Green Version]
  210. Hidalgo, C.; Sanchez, G.; Barrientos, G.; Aracena-Parks, P. A transverse tubule NADPH oxidase activity stimulates calcium release from isolated triads via ryanodine receptor type 1 S -glutathionylation. J. Biol. Chem. 2006, 281, 26473–26482. [Google Scholar] [CrossRef] [Green Version]
  211. Zhao, X.; Bey, E.A.; Wientjes, F.B.; Cathcart, M.K. Cytosolic phospholipase A2 (cPLA2) regulation of human monocyte NADPH oxidase activity. cPLA2 affects translocation but not phosphorylation of p67(phox) and p47(phox). J. Biol. Chem. 2002, 277, 25385–25392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Barbieri, E.; Sestili, P. Reactive oxygen species in skeletal muscle signaling. J. Signal Transduct. 2012, 2012, 982794. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Harrison, R. Structure and function of xanthine oxidoreductase: Where are we now? Free Radic. Biol. Med. 2002, 33, 774–797. [Google Scholar] [CrossRef]
  214. Prosser, B.L.; Khairallah, R.J.; Ziman, A.P.; Ward, C.W.; Lederer, W.J. X-ROS signaling in the heart and skeletal muscle: Stretch-dependent local ROS regulates [Ca(2+)]i. J. Mol. Cell. Cardiol. 2013, 58, 172–181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Lindsay, A.; McCourt, P.M.; Karachunski, P.; Lowe, D.A.; Ervasti, J.M. Xanthine oxidase is hyper-active in Duchenne muscular dystrophy. Free Radic. Biol. Med. 2018, 129, 364–371. [Google Scholar] [CrossRef] [PubMed]
  216. Whitehead, N.P.; Kim, M.J.; Bible, K.L.; Adams, M.E.; Froehner, S.C. A new therapeutic effect of simvastatin revealed by functional improvement in muscular dystrophy. Proc. Natl. Acad. Sci. USA 2015, 112, 12864–12869. [Google Scholar] [CrossRef] [Green Version]
  217. Rosenberg, A.S.; Puig, M.; Nagaraju, K.; Hoffman, E.P.; Villalta, S.A.; Rao, V.A.; Wakefield, L.M.; Woodcock, J. Immune-mediated pathology in Duchenne muscular dystrophy. Sci. Transl. Med. 2015, 7, 299rv294. [Google Scholar] [CrossRef] [Green Version]
  218. Tidball, J.G. Mechanisms of muscle injury, repair, and regeneration. Compr. Physiol. 2011, 1, 2029–2062. [Google Scholar]
  219. Kozakowska, M.; Pietraszek-Gremplewicz, K.; Jozkowicz, A.; Dulak, J. The role of oxidative stress in skeletal muscle injury and regeneration: Focus on antioxidant enzymes. J. Muscle Res. Cell Motil. 2015, 36, 377–393. [Google Scholar] [CrossRef] [Green Version]
  220. Williams, I.A.; Allen, D.G. The role of reactive oxygen species in the hearts of dystrophin-deficient mdx mice. Am. J. Physiol. Heart Circ. Physiol. 2007, 293, H1969–H1977. [Google Scholar] [CrossRef] [Green Version]
  221. Damiano, S.; Muscariello, E.; La Rosa, G.; Di Maro, M.; Mondola, P.; Santillo, M. Dual Role of Reactive Oxygen Species in Muscle Function: Can Antioxidant Dietary Supplements Counteract Age-Related Sarcopenia? Int. J. Mol. Sci. 2019, 20, 3815. [Google Scholar] [CrossRef] [Green Version]
  222. Sakellariou, G.K.; Vasilaki, A.; Palomero, J.; Kayani, A.; Zibrik, L.; McArdle, A.; Jackson, M.J. Studies of mitochondrial and nonmitochondrial sources implicate nicotinamide adenine dinucleotide phosphate oxidase(s) in the increased skeletal muscle superoxide generation that occurs during contractile activity. Antioxid. Redox Signal. 2013, 18, 603–621. [Google Scholar] [CrossRef] [Green Version]
  223. Loehr, J.A.; Abo-Zahrah, R.; Pal, R.; Rodney, G.G. Sphingomyelinase promotes oxidant production and skeletal muscle contractile dysfunction through activation of NADPH oxidase. Front. Physiol. 2014, 5, 530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Pal, R.; Basu Thakur, P.; Li, S.; Minard, C.; Rodney, G.G. Real-time imaging of NADPH oxidase activity in living cells using a novel fluorescent protein reporter. PLoS ONE 2013, 8, e63989. [Google Scholar] [CrossRef]
  225. Powers, S.K. Can antioxidants protect against disuse muscle atrophy? Sports Med. 2014, 44 (Suppl. 2), S155–S165. [Google Scholar] [CrossRef] [Green Version]
  226. Hollander, J.; Fiebig, R.; Gore, M.; Bejma, J.; Ookawara, T.; Ohno, H.; Ji, L.L. Superoxide dismutase gene expression in skeletal muscle: Fiber-specific adaptation to endurance training. Am. J. Physiol. 1999, 277, R856–R862. [Google Scholar] [CrossRef]
  227. Powers, S.K.; Criswell, D.; Lawler, J.; Ji, L.L.; Martin, D.; Herb, R.A.; Dudley, G. Influence of exercise and fiber type on antioxidant enzyme activity in rat skeletal muscle. Am. J. Physiol. 1994, 266, R375–R380. [Google Scholar] [CrossRef] [PubMed]
  228. Burr, I.M.; Asayama, K.; Fenichel, G.M. Superoxide dismutases, glutathione peroxidase, and catalase in neuromuscular disease. Muscle Nerve 1987, 10, 150–154. [Google Scholar] [CrossRef] [PubMed]
  229. Mechler, F.; Imre, S.; Dioszeghy, P. Lipid peroxidation and superoxide dismutase activity in muscle and erythrocytes in Duchenne muscular dystrophy. J. Neurol. Sci. 1984, 63, 279–283. [Google Scholar] [CrossRef]
  230. Ragusa, R.J.; Chow, C.K.; St Clair, D.K.; Porter, J.D. Extraocular, limb and diaphragm muscle group-specific antioxidant enzyme activity patterns in control and mdx mice. J. Neurol. Sci. 1996, 139, 180–186. [Google Scholar] [CrossRef]
  231. Dudley, R.W.; Khairallah, M.; Mohammed, S.; Lands, L.; Des Rosiers, C.; Petrof, B.J. Dynamic responses of the glutathione system to acute oxidative stress in dystrophic mouse (mdx) muscles. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2006, 291, R704–R710. [Google Scholar] [CrossRef] [Green Version]
  232. Cao, S.S.; Kaufman, R.J. Endoplasmic reticulum stress and oxidative stress in cell fate decision and human disease. Antioxid. Redox Signal. 2014, 21, 396–413. [Google Scholar] [CrossRef] [PubMed]
  233. Hetz, C.; Papa, F.R. The Unfolded Protein Response and Cell Fate Control. Mol. Cell 2018, 69, 169–181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Sun, C.; Yang, C.; Xue, R.; Li, S.; Zhang, T.; Pan, L.; Ma, X.; Wang, L.; Li, D. Sulforaphane alleviates muscular dystrophy in mdx mice by activation of Nrf2. J. Appl. Physiol. 2015, 118, 224–237. [Google Scholar] [CrossRef]
  235. Sun, C.; Li, S.; Li, D. Sulforaphane mitigates muscle fibrosis in mdx mice via Nrf2-mediated inhibition of TGF-beta/Smad signaling. J. Appl. Physiol. 2016, 120, 377–390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Gureev, A.P.; Shaforostova, E.A.; Popov, V.N. Regulation of Mitochondrial Biogenesis as a Way for Active Longevity: Interaction Between the Nrf2 and PGC-1alpha Signaling Pathways. Front. Genet. 2019, 10, 435. [Google Scholar] [CrossRef] [Green Version]
  237. Nguyen, T.; Sherratt, P.J.; Nioi, P.; Yang, C.S.; Pickett, C.B. Nrf2 controls constitutive and inducible expression of ARE-driven genes through a dynamic pathway involving nucleocytoplasmic shuttling by Keap1. J. Biol. Chem. 2005, 280, 32485–32492. [Google Scholar] [CrossRef] [Green Version]
  238. Vargas-Mendoza, N.; Morales-Gonzalez, A.; Madrigal-Santillan, E.O.; Madrigal-Bujaidar, E.; Alvarez-Gonzalez, I.; Garcia-Melo, L.F.; Anguiano-Robledo, L.; Fregoso-Aguilar, T.; Morales-Gonzalez, J.A. Antioxidant and Adaptative Response Mediated by Nrf2 during Physical Exercise. Antioxidants 2019, 8, 196. [Google Scholar] [CrossRef] [Green Version]
  239. Redza-Dutordoir, M.; Averill-Bates, D.A. Activation of apoptosis signalling pathways by reactive oxygen species. Biochim. Biophys. Acta 2016, 1863, 2977–2992. [Google Scholar] [CrossRef]
  240. Rando, T.A.; Disatnik, M.H.; Yu, Y.; Franco, A. Muscle cells from mdx mice have an increased susceptibility to oxidative stress. Neuromuscul. Disord. 1998, 8, 14–21. [Google Scholar] [CrossRef]
  241. Choi, S.J.; Kim, H.S. Deregulation of Nrf2/ARE signaling pathway causes susceptibility of dystrophin-deficient myotubes to menadione-induced oxidative stress. Exp. Cell Res. 2018, 364, 224–233. [Google Scholar] [CrossRef]
  242. Afroze, D.; Kumar, A. ER stress in skeletal muscle remodeling and myopathies. FEBS J. 2019, 286, 379–398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Pauly, M.; Angebault-Prouteau, C.; Dridi, H.; Notarnicola, C.; Scheuermann, V.; Lacampagne, A.; Matecki, S.; Fauconnier, J. ER stress disturbs SR/ER-mitochondria Ca(2+) transfer: Implications in Duchenne muscular dystrophy. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 2229–2239. [Google Scholar] [CrossRef] [PubMed]
  244. Di Meo, S.; Reed, T.T.; Venditti, P.; Victor, V.M. Role of ROS and RNS Sources in Physiological and Pathological Conditions. Oxid. Med. Cell. Longev. 2016, 2016, 1245049. [Google Scholar] [CrossRef] [PubMed]
  245. Steinbacher, P.; Eckl, P. Impact of oxidative stress on exercising skeletal muscle. Biomolecules 2015, 5, 356–377. [Google Scholar] [CrossRef] [PubMed]
  246. Lambertucci, R.H.; Levada-Pires, A.C.; Rossoni, L.V.; Curi, R.; Pithon-Curi, T.C. Effects of aerobic exercise training on antioxidant enzyme activities and mRNA levels in soleus muscle from young and aged rats. Mech. Ageing Dev. 2007, 128, 267–275. [Google Scholar] [CrossRef]
  247. Simioni, C.; Zauli, G.; Martelli, A.M.; Vitale, M.; Sacchetti, G.; Gonelli, A.; Neri, L.M. Oxidative stress: Role of physical exercise and antioxidant nutraceuticals in adulthood and aging. Oncotarget 2018, 9, 17181–17198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Finaud, J.; Lac, G.; Filaire, E. Oxidative stress: Relationship with exercise and training. Sports Med. 2006, 36, 327–358. [Google Scholar] [CrossRef]
  249. Gomez-Cabrera, M.C.; Domenech, E.; Romagnoli, M.; Arduini, A.; Borras, C.; Pallardo, F.V.; Sastre, J.; Vina, J. Oral administration of vitamin C decreases muscle mitochondrial biogenesis and hampers training-induced adaptations in endurance performance. Am. J. Clin. Nutr. 2008, 87, 142–149. [Google Scholar] [CrossRef] [Green Version]
  250. Ristow, M.; Zarse, K.; Oberbach, A.; Kloting, N.; Birringer, M.; Kiehntopf, M.; Stumvoll, M.; Kahn, C.R.; Bluher, M. Antioxidants prevent health-promoting effects of physical exercise in humans. Proc. Natl. Acad. Sci. USA 2009, 106, 8665–8670. [Google Scholar] [CrossRef] [Green Version]
  251. Paulsen, G.; Cumming, K.T.; Holden, G.; Hallen, J.; Ronnestad, B.R.; Sveen, O.; Skaug, A.; Paur, I.; Bastani, N.E.; Ostgaard, H.N.; et al. Vitamin C and E supplementation hampers cellular adaptation to endurance training in humans: A double-blind, randomised, controlled trial. J. Physiol. 2014, 592, 1887–1901. [Google Scholar] [CrossRef]
  252. Kang, C.; Chung, E.; Diffee, G.; Ji, L.L. Exercise training attenuates aging-associated mitochondrial dysfunction in rat skeletal muscle: Role of PGC-1alpha. Exp. Gerontol. 2013, 48, 1343–1350. [Google Scholar] [CrossRef]
  253. Kang, C.; Li Ji, L. Role of PGC-1alpha signaling in skeletal muscle health and disease. Ann. N. Y. Acad. Sci. 2012, 1271, 110–117. [Google Scholar] [CrossRef]
  254. Tanaka, T.; Nishimura, A.; Nishiyama, K.; Goto, T.; Numaga-Tomita, T.; Nishida, M. Mitochondrial dynamics in exercise physiology. Pflügers Arch. 2020, 472, 137–153. [Google Scholar] [CrossRef]
  255. Moore, T.M.; Zhou, Z.; Cohn, W.; Norheim, F.; Lin, A.J.; Kalajian, N.; Strumwasser, A.R.; Cory, K.; Whitney, K.; Ho, T.; et al. The impact of exercise on mitochondrial dynamics and the role of Drp1 in exercise performance and training adaptations in skeletal muscle. Mol. Metab. 2019, 21, 51–67. [Google Scholar] [CrossRef]
  256. Trewin, A.J.; Berry, B.J.; Wojtovich, A.P. Exercise and Mitochondrial Dynamics: Keeping in Shape with ROS and AMPK. Antioxidants 2018, 7, 7. [Google Scholar] [CrossRef] [Green Version]
  257. Balan, E.; Schwalm, C.; Naslain, D.; Nielens, H.; Francaux, M.; Deldicque, L. Regular Endurance Exercise Promotes Fission, Mitophagy, and Oxidative Phosphorylation in Human Skeletal Muscle Independently of Age. Front. Physiol. 2019, 10, 1088. [Google Scholar] [CrossRef] [Green Version]
  258. Vainshtein, A.; Tryon, L.D.; Pauly, M.; Hood, D.A. Role of PGC-1alpha during acute exercise-induced autophagy and mitophagy in skeletal muscle. Am. J. Physiol. Cell Physiol. 2015, 308, C710–C719. [Google Scholar] [CrossRef] [Green Version]
  259. Carnio, S.; LoVerso, F.; Baraibar, M.A.; Longa, E.; Khan, M.M.; Maffei, M.; Reischl, M.; Canepari, M.; Loefler, S.; Kern, H.; et al. Autophagy impairment in muscle induces neuromuscular junction degeneration and precocious aging. Cell Rep. 2014, 8, 1509–1521. [Google Scholar] [CrossRef] [PubMed]
  260. McCormick, R.; Vasilaki, A. Age-related changes in skeletal muscle: Changes to life-style as a therapy. Biogerontology 2018, 19, 519–536. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  261. Gordon, B.S.; Delgado Diaz, D.C.; Kostek, M.C. Resveratrol decreases inflammation and increases utrophin gene expression in the mdx mouse model of Duchenne muscular dystrophy. Clin. Nutr. 2013, 32, 104–111. [Google Scholar] [CrossRef] [PubMed]
  262. Mastaloudis, A.; Leonard, S.W.; Traber, M.G. Oxidative stress in athletes during extreme endurance exercise. Free Radic. Biol. Med. 2001, 31, 911–922. [Google Scholar] [CrossRef]
  263. Knez, W.L.; Jenkins, D.G.; Coombes, J.S. The effect of an increased training volume on oxidative stress. Int. J. Sports Med. 2014, 35, 8–13. [Google Scholar] [CrossRef] [PubMed]
  264. Gomez-Cabrera, M.C.; Domenech, E.; Vina, J. Moderate exercise is an antioxidant: Upregulation of antioxidant genes by training. Free Radic. Biol. Med. 2008, 44, 126–131. [Google Scholar] [CrossRef] [PubMed]
  265. Leeuwenburgh, C.; Fiebig, R.; Chandwaney, R.; Ji, L.L. Aging and exercise training in skeletal muscle: Responses of glutathione and antioxidant enzyme systems. Am. J. Physiol. 1994, 267, R439–R445. [Google Scholar] [CrossRef] [PubMed]
  266. Oh-ishi, S.; Kizaki, T.; Nagasawa, J.; Izawa, T.; Komabayashi, T.; Nagata, N.; Suzuki, K.; Taniguchi, N.; Ohno, H. Effects of endurance training on superoxide dismutase activity, content and mRNA expression in rat muscle. Clin. Exp. Pharmacol. Physiol. 1997, 24, 326–332. [Google Scholar] [CrossRef] [PubMed]
  267. Powers, S.K.; Criswell, D.; Lawler, J.; Martin, D.; Ji, L.L.; Herb, R.A.; Dudley, G. Regional training-induced alterations in diaphragmatic oxidative and antioxidant enzymes. Respir. Physiol. 1994, 95, 227–237. [Google Scholar] [CrossRef]
  268. Vincent, H.K.; Powers, S.K.; Stewart, D.J.; Demirel, H.A.; Shanely, R.A.; Naito, H. Short-term exercise training improves diaphragm antioxidant capacity and endurance. Eur. J. Appl. Physiol. 2000, 81, 67–74. [Google Scholar] [CrossRef]
  269. Ji, L.L. Modulation of skeletal muscle antioxidant defense by exercise: Role of redox signaling. Free Radic. Biol. Med. 2008, 44, 142–152. [Google Scholar] [CrossRef]
  270. Pereira, B.; Costa Rosa, L.F.; Safi, D.A.; Medeiros, M.H.; Curi, R.; Bechara, E.J. Superoxide dismutase, catalase, and glutathione peroxidase activities in muscle and lymphoid organs of sedentary and exercise-trained rats. Physiol. Behav. 1994, 56, 1095–1099. [Google Scholar] [CrossRef]
  271. Nakao, C.; Ookawara, T.; Kizaki, T.; Oh-Ishi, S.; Miyazaki, H.; Haga, S.; Sato, Y.; Ji, L.L.; Ohno, H. Effects of swimming training on three superoxide dismutase isoenzymes in mouse tissues. J. Appl. Physiol. 2000, 88, 649–654. [Google Scholar] [CrossRef] [PubMed]
  272. Suzuki, K.; Ohno, H.; Oh-ishi, S.; Kizaki, T.; Ookawara, T.; Fujii, J.; Radák, Z.; Taniguchi, N. Chapter 11—Superoxide dismutases in exercise and disease. In Handbook of Oxidants and Antioxidants in Exercise; Part IV; Sen, C.K., Packer, L., Hänninen, O.O.P., Eds.; Elsevier Science B.V.: Amsterdam, The Netherlands, 2000; pp. 243–295. [Google Scholar] [CrossRef]
  273. Hitomi, Y.; Watanabe, S.; Kizaki, T.; Sakurai, T.; Takemasa, T.; Haga, S.; Ookawara, T.; Suzuki, K.; Ohno, H. Acute exercise increases expression of extracellular superoxide dismutase in skeletal muscle and the aorta. Redox Rep. 2008, 13, 213–216. [Google Scholar] [CrossRef] [Green Version]
  274. Criswell, D.; Powers, S.; Dodd, S.; Lawler, J.; Edwards, W.; Renshler, K.; Grinton, S. High intensity training-induced changes in skeletal muscle antioxidant enzyme activity. Med. Sci. Sports Exerc. 1993, 25, 1135–1140. [Google Scholar] [CrossRef]
  275. Hellsten, Y.; Apple, F.S.; Sjodin, B. Effect of sprint cycle training on activities of antioxidant enzymes in human skeletal muscle. J. Appl. Physiol. 1996, 81, 1484–1487. [Google Scholar] [CrossRef]
  276. Leeuwenburgh, C.; Hollander, J.; Leichtweis, S.; Griffiths, M.; Gore, M.; Ji, L.L. Adaptations of glutathione antioxidant system to endurance training are tissue and muscle fiber specific. Am. J. Physiol. 1997, 272, R363–R369. [Google Scholar] [CrossRef]
  277. Venditti, P.; Di Meo, S. Effect of training on antioxidant capacity, tissue damage, and endurance of adult male rats. Int. J. Sports Med. 1997, 18, 497–502. [Google Scholar] [CrossRef] [PubMed]
  278. Rius-Perez, S.; Torres-Cuevas, I.; Millan, I.; Ortega, A.L.; Perez, S. PGC-1alpha, Inflammation, and Oxidative Stress: An Integrative View in Metabolism. Oxid. Med. Cell. Longev. 2020, 2020, 1452696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  279. Kim, J.H.; Lawler, J.M. Amplification of proinflammatory phenotype, damage, and weakness by oxidative stress in the diaphragm muscle of mdx mice. Free Radic. Biol. Med. 2012, 52, 1597–1606. [Google Scholar] [CrossRef] [PubMed]
  280. Selsby, J.T. Increased catalase expression improves muscle function in mdx mice. Exp. Physiol. 2011, 96, 194–202. [Google Scholar] [CrossRef] [Green Version]
  281. Faist, V.; Konig, J.; Hoger, H.; Elmadfa, I. Decreased mitochondrial oxygen consumption and antioxidant enzyme activities in skeletal muscle of dystrophic mice after low-intensity exercise. Ann. Nutr. Metab. 2001, 45, 58–66. [Google Scholar] [CrossRef]
  282. Hulmi, J.J.; Oliveira, B.M.; Silvennoinen, M.; Hoogaars, W.M.; Pasternack, A.; Kainulainen, H.; Ritvos, O. Exercise restores decreased physical activity levels and increases markers of autophagy and oxidative capacity in myostatin/activin-blocked mdx mice. Am. J. Physiol. Endocrinol. Metab. 2013, 305, E171–E182. [Google Scholar] [CrossRef]
  283. Handschin, C.; Kobayashi, Y.M.; Chin, S.; Seale, P.; Campbell, K.P.; Spiegelman, B.M. PGC-1alpha regulates the neuromuscular junction program and ameliorates Duchenne muscular dystrophy. Genes Dev. 2007, 21, 770–783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Hulmi, J.J.; Hentila, J.; DeRuisseau, K.C.; Oliveira, B.M.; Papaioannou, K.G.; Autio, R.; Kujala, U.M.; Ritvos, O.; Kainulainen, H.; Korkmaz, A.; et al. Effects of muscular dystrophy, exercise and blocking activin receptor IIB ligands on the unfolded protein response and oxidative stress. Free Radic. Biol. Med. 2016, 99, 308–322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Gali Ramamoorthy, T.; Laverny, G.; Schlagowski, A.I.; Zoll, J.; Messaddeq, N.; Bornert, J.M.; Panza, S.; Ferry, A.; Geny, B. The transcriptional coregulator PGC-1beta controls mitochondrial function and anti-oxidant defence in skeletal muscles. Nat. Commun. 2015, 6, 10210. [Google Scholar] [CrossRef] [Green Version]
  286. Booth, F.W.; Ruegsegger, G.N.; Toedebusch, R.G.; Yan, Z. Endurance Exercise and the Regulation of Skeletal Muscle Metabolism. Prog. Mol. Biol. Transl. Sci. 2015, 135, 129–151. [Google Scholar] [PubMed]
  287. Debold, E.P. Potential molecular mechanisms underlying muscle fatigue mediated by reactive oxygen and nitrogen species. Front. Physiol. 2015, 6, 239. [Google Scholar] [CrossRef]
  288. Ostrowski, K.; Rohde, T.; Asp, S.; Schjerling, P.; Pedersen, B.K. Pro- and anti-inflammatory cytokine balance in strenuous exercise in humans. J. Physiol. 1999, 515(Pt. 1), 287–291. [Google Scholar] [CrossRef]
  289. Cerqueira, E.; Marinho, D.A.; Neiva, H.P.; Lourenco, O. Inflammatory Effects of High and Moderate Intensity Exercise—A Systematic Review. Front. Physiol. 2019, 10, 1550. [Google Scholar] [CrossRef]
  290. Rose, G.L.; Skinner, T.L.; Mielke, G.I.; Schaumberg, M.A. The effect of exercise intensity on chronic inflammation: A systematic review and meta-analysis. J. Sci. Med. Sport 2020, 24, 345–351. [Google Scholar] [CrossRef]
  291. Kramer, H.F.; Goodyear, L.J. Exercise, MAPK, and NF-kappaB signaling in skeletal muscle. J. Appl. Physiol. 2007, 103, 388–395. [Google Scholar] [CrossRef]
  292. Cuevas, M.J.; Almar, M.; Garcia-Glez, J.C.; Garcia-Lopez, D.; De Paz, J.A.; Alvear-Ordenes, I.; Gonzalez-Gallego, J. Changes in oxidative stress markers and NF-kappaB activation induced by sprint exercise. Free Radic. Res. 2005, 39, 431–439. [Google Scholar] [CrossRef]
  293. Scicchitano, B.M.; Faraldi, M.; Musaro, A. The Proteolytic Systems of Muscle Wasting. Recent Adv. DNA Gene Seq. 2015, 9, 26–35. [Google Scholar] [CrossRef] [Green Version]
  294. Judge, A.R.; Dodd, S.L. Xanthine oxidase and activated neutrophils cause oxidative damage to skeletal muscle after contractile claudication. Am. J. Physiol. Heart Circ. Physiol. 2004, 286, H252–H256. [Google Scholar] [CrossRef] [Green Version]
  295. Radak, Z.; Taylor, A.W.; Ohno, H.; Goto, S. Adaptation to exercise-induced oxidative stress: From muscle to brain. Exerc. Immunol. Rev. 2001, 7, 90–107. [Google Scholar]
  296. Peake, J.; Suzuki, K. Neutrophil activation, antioxidant supplements and exercise-induced oxidative stress. Exerc. Immunol. Rev. 2004, 10, 129–141. [Google Scholar]
  297. Moylan, J.S.; Reid, M.B. Oxidative stress, chronic disease, and muscle wasting. Muscle Nerve 2007, 35, 411–429. [Google Scholar] [CrossRef]
  298. Gomes, E.C.; Silva, A.N.; de Oliveira, M.R. Oxidants, antioxidants, and the beneficial roles of exercise-induced production of reactive species. Oxid. Med. Cell. Longev. 2012, 2012, 756132. [Google Scholar] [CrossRef]
  299. Handschin, C.; Spiegelman, B.M. The role of exercise and PGC1alpha in inflammation and chronic disease. Nature 2008, 454, 463–469. [Google Scholar] [CrossRef] [Green Version]
  300. Morgan, M.J.; Liu, Z.G. Crosstalk of reactive oxygen species and NF-kappaB signaling. Cell Res. 2011, 21, 103–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  301. Lawrence, T. The nuclear factor NF-kappaB pathway in inflammation. Cold Spring Harb. Perspect. Biol. 2009, 1, a001651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  302. Mittal, M.; Siddiqui, M.R.; Tran, K.; Reddy, S.P.; Malik, A.B. Reactive oxygen species in inflammation and tissue injury. Antioxid. Redox Signal. 2014, 20, 1126–1167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Kumar, A.; Boriek, A.M. Mechanical stress activates the nuclear factor-kappaB pathway in skeletal muscle fibers: A possible role in Duchenne muscular dystrophy. FASEB J. 2003, 17, 386–396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Miyatake, S.; Shimizu-Motohashi, Y.; Takeda, S.; Aoki, Y. Anti-inflammatory drugs for Duchenne muscular dystrophy: Focus on skeletal muscle-releasing factors. Drug Des. Devel. Ther. 2016, 10, 2745–2758. [Google Scholar] [PubMed] [Green Version]
  305. Monici, M.C.; Aguennouz, M.; Mazzeo, A.; Messina, C.; Vita, G. Activation of nuclear factor-kappaB in inflammatory myopathies and Duchenne muscular dystrophy. Neurology 2003, 60, 993–997. [Google Scholar] [CrossRef] [PubMed]
  306. Acharyya, S.; Villalta, S.A.; Bakkar, N.; Bupha-Intr, T.; Janssen, P.M.; Carathers, M.; Li, Z.W.; Beg, A.A.; Ghosh, S.; Sahenk, Z.; et al. Interplay of IKK/NF-kappaB signaling in macrophages and myofibers promotes muscle degeneration in Duchenne muscular dystrophy. J. Clin. Invest. 2007, 117, 889–901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  307. Millay, D.P.; Goonasekera, S.A.; Sargent, M.A.; Maillet, M.; Aronow, B.J.; Molkentin, J.D. Calcium influx is sufficient to induce muscular dystrophy through a TRPC-dependent mechanism. Proc. Natl. Acad. Sci. USA 2009, 106, 19023–19028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  308. Carlson, C.G.; Samadi, A.; Siegel, A. Chronic treatment with agents that stabilize cytosolic IkappaB-alpha enhances survival and improves resting membrane potential in MDX muscle fibers subjected to chronic passive stretch. Neurobiol. Dis. 2005, 20, 719–730. [Google Scholar] [CrossRef] [PubMed]
  309. Pan, Y.; Chen, C.; Shen, Y.; Zhu, C.H.; Wang, G.; Wang, X.C.; Chen, H.Q.; Zhu, M.S. Curcumin alleviates dystrophic muscle pathology in mdx mice. Mol. Cells 2008, 25, 531–537. [Google Scholar] [PubMed]
  310. Manzur, A.Y.; Kuntzer, T.; Pike, M.; Swan, A. Glucocorticoid corticosteroids for Duchenne muscular dystrophy. Cochrane Database Syst. Rev. 2004, CD003725. [Google Scholar] [CrossRef]
  311. Andrews, J.G.; Wahl, R.A. Duchenne and Becker muscular dystrophy in adolescents: Current perspectives. Adolesc. Health Med. Ther. 2018, 9, 53–63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  312. Iwabu, M.; Yamauchi, T.; Okada-Iwabu, M.; Sato, K.; Nakagawa, T.; Funata, M.; Yamaguchi, M.; Namiki, S.; Nakayama, R.; Tabata, M.; et al. Adiponectin and AdipoR1 regulate PGC-1alpha and mitochondria by Ca(2+) and AMPK/SIRT1. Nature 2010, 464, 1313–1319. [Google Scholar] [CrossRef]
  313. Song, T.J.; Lee, K.A.; Kang, S.W.; Cho, H.; Choi, Y.C. Three cases of manifesting female carriers in patients with Duchenne muscular dystrophy. Yonsei Med. J. 2011, 52, 192–195. [Google Scholar] [CrossRef]
  314. Finsterer, J.; Stollberger, C.; Freudenthaler, B.; Simoni, D.; Hoftberger, R.; Wagner, K. Muscular and cardiac manifestations in a Duchenne-carrier harboring a dystrophin deletion of exons 12-29. Intractable Rare Dis. Res. 2018, 7, 120–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  315. Hakim, C.H.; Duan, D. Gender differences in contractile and passive properties of mdx extensor digitorum longus muscle. Muscle Nerve 2012, 45, 250–256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  316. Bostick, B.; Yue, Y.; Duan, D. Gender influences cardiac function in the mdx model of Duchenne cardiomyopathy. Muscle Nerve 2010, 42, 600–603. [Google Scholar] [CrossRef] [PubMed]
  317. Powers, S.K.; Dodd, S.L.; Jackson, E.M. Total Fitness & Wellness, 6th ed.; Pearson: Boston, MA, USA, 2013. [Google Scholar]
Figure 1. Search strategy and selection criteria of papers used in the review.
Figure 1. Search strategy and selection criteria of papers used in the review.
Antioxidants 10 00558 g001
Figure 2. Effects of low intensity treadmill running on mdx muscle pathological changes. Exercise can reverse: (a) Inflammatory response, by inducing a shift of macrophage population from to M1 pro-inflammatory type to the regenerative M2 type, and by modulating inflammatory pathways and molecules, including NF-kB (Nuclear factor-kB), CCL2 (Monocyte chemoattractant protein-1), Adiponectin and PGC-1α (Peroxisome proliferator-activated receptor gamma coactivator 1-alpha); (b) Fibrosis, by reducing collagen deposition; (c) Muscle degeneration, likely by activating satellite cells; (d) Oxidative stress and metabolic impairment, and the related mitochondria and SR (sarcoplasmic reticulum) stress, NOX (NADPH oxidase) and XO (xanthine oxidase) hyperactivation, by inducing activation of antioxidants, mitochondria metabolism and PGC-1α pathway. Created with BioRender.com.
Figure 2. Effects of low intensity treadmill running on mdx muscle pathological changes. Exercise can reverse: (a) Inflammatory response, by inducing a shift of macrophage population from to M1 pro-inflammatory type to the regenerative M2 type, and by modulating inflammatory pathways and molecules, including NF-kB (Nuclear factor-kB), CCL2 (Monocyte chemoattractant protein-1), Adiponectin and PGC-1α (Peroxisome proliferator-activated receptor gamma coactivator 1-alpha); (b) Fibrosis, by reducing collagen deposition; (c) Muscle degeneration, likely by activating satellite cells; (d) Oxidative stress and metabolic impairment, and the related mitochondria and SR (sarcoplasmic reticulum) stress, NOX (NADPH oxidase) and XO (xanthine oxidase) hyperactivation, by inducing activation of antioxidants, mitochondria metabolism and PGC-1α pathway. Created with BioRender.com.
Antioxidants 10 00558 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Frinchi, M.; Morici, G.; Mudó, G.; Bonsignore, M.R.; Di Liberto, V. Beneficial Role of Exercise in the Modulation of mdx Muscle Plastic Remodeling and Oxidative Stress. Antioxidants 2021, 10, 558. https://doi.org/10.3390/antiox10040558

AMA Style

Frinchi M, Morici G, Mudó G, Bonsignore MR, Di Liberto V. Beneficial Role of Exercise in the Modulation of mdx Muscle Plastic Remodeling and Oxidative Stress. Antioxidants. 2021; 10(4):558. https://doi.org/10.3390/antiox10040558

Chicago/Turabian Style

Frinchi, Monica, Giuseppe Morici, Giuseppa Mudó, Maria R. Bonsignore, and Valentina Di Liberto. 2021. "Beneficial Role of Exercise in the Modulation of mdx Muscle Plastic Remodeling and Oxidative Stress" Antioxidants 10, no. 4: 558. https://doi.org/10.3390/antiox10040558

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop