Next Article in Journal
Chimeric Antigen Receptor (CAR) T Cell Therapy for Metastatic Melanoma: Challenges and Road Ahead
Next Article in Special Issue
Circulating myomiRs in Muscle Denervation: From Surgical to ALS Pathological Condition
Previous Article in Journal
Systemic Administration of Insulin Receptor Antagonist Results in Endothelial and Perivascular Adipose Tissue Dysfunction in Mice
Previous Article in Special Issue
Redox Homeostasis in Muscular Dystrophies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Skeletal Muscle Metabolism: Origin or Prognostic Factor for Amyotrophic Lateral Sclerosis (ALS) Development?

by
Cyril Quessada
1,†,
Alexandra Bouscary
1,2,†,
Frédérique René
1,
Cristiana Valle
3,
Alberto Ferri
3,
Shyuan T. Ngo
4,5,6 and
Jean-Philippe Loeffler
1,*
1
Mécanismes Centraux et Périphériques de la Neurodégénérescence, INSERM, UMR_S 1118, Centre de Recherche de Biomédecine de Strasbourg (CRBS), Université de Strasbourg, F-67000 Strasbourg, France
2
Neuro-Sys SAS, F-13120 Gardanne, France
3
IRCCS Fondazione Santa Lucia, Institute of Translational Pharmacology, CNR, 00133 Roma, Italy
4
Australian Institute for Bioengineering and Nanotechnology, The University of Queensland, Brisbane, QLD 4072, Australia
5
Centre for Clinical Research, The University of Queensland, Herston, QLD 4006, Australia
6
Queensland Brain Institute, The University of Queensland, St. Lucia, QLD 4072, Australia
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Cells 2021, 10(6), 1449; https://doi.org/10.3390/cells10061449
Submission received: 2 May 2021 / Revised: 3 June 2021 / Accepted: 7 June 2021 / Published: 9 June 2021

Abstract

:
Amyotrophic lateral sclerosis (ALS) is a fatal neurodegenerative disease characterized by progressive and selective loss of motor neurons, amyotrophy and skeletal muscle paralysis usually leading to death due to respiratory failure. While generally considered an intrinsic motor neuron disease, data obtained in recent years, including our own, suggest that motor neuron protection is not sufficient to counter the disease. The dismantling of the neuromuscular junction is closely linked to chronic energy deficit found throughout the body. Metabolic (hypermetabolism and dyslipidemia) and mitochondrial alterations described in patients and murine models of ALS are associated with the development and progression of disease pathology and they appear long before motor neurons die. It is clear that these metabolic changes participate in the pathology of the disease. In this review, we summarize these changes seen throughout the course of the disease, and the subsequent impact of glucose–fatty acid oxidation imbalance on disease progression. We also highlight studies that show that correcting this loss of metabolic flexibility should now be considered a major goal for the treatment of ALS.

1. Introduction

In humans, the muscular system is divided into two distinct categories: smooth and striated muscles. The number, organization, and function of striated muscle require considerable energy consumption when compared to the entire human body. In this review, we first present a summary of amyotrophic lateral sclerosis (ALS). Second, we examine the role of muscle energy metabolism in the pathophysiology of ALS and we further discuss how targeting muscle offers an avenue for treating the disease.

1.1. Amyotrophic Lateral Sclerosis

Amyotrophic lateral sclerosis (ALS) is a fatal and incurable neurodegenerative disease. The gradual loss of weight and muscle strength and the onset and progression of muscle paralysis are the main visible presentations of this disease. From a cellular perspective, the loss of cortical motor neurons and spinal and bulbar motor neurons is clearly established [1]. ALS typically affects patients between the ages of 50 and 60 and invariably causes death 3 to 5 years after onset [2]. Worldwide, 60,000 people die of ALS each year [3] with an incidence of 1.75 to 3 per 100,000 per year [4]. In France, there are 3 to 4 new cases per year per 100,000 individuals and a prevalence of 6000 [5]. Many genes have been directly linked with the sporadic (90% of cases) and familial forms (10% of cases) of ALS [6]. In the latter, the pattern of inheritance is generally autosomal dominant. Among the main major susceptibility genes, C9ORF72, SOD1, FUS and TARDBP account for 48%, 12%, 5% and 5% in familial forms, respectively [5,6,7,8,9,10,11]. Currently, only two FDA-approved treatments are indicated for the treatment of ALS: Rilutek and Edaravone. These molecules increase the life expectancy of some patients by a few months [12,13]. The pathogenic mechanisms that are proposed to contribute to the degeneration of motor neurons include excitotoxicity, oxidative stress, protein aggregation, alteration of RNA metabolism and mitochondrial dysfunction [14,15,16,17,18,19,20] (Figure 1). Two hypotheses integrating these pathogenic mechanisms leading to ALS are currently proposed: the corticofugal dying-forward hypothesis which describes the progressive descending neurodegeneration, initiated in the motor cortex, spreading to the motor neuron and ultimately affecting the neuromuscular junctions (NMJ) [21,22,23] and the dying back hypothesis which is initiated at the NMJ with a retrograde progression of degeneration [24]. These two hypotheses are not mutually exclusive and may coexist to initiate ALS. The topic of this review is focused on the dying-back process of the motor neurons initiated at the NMJ and how skeletal muscle can be involved in ALS.

1.2. Neuromuscular Junction

While ALS is clearly a complex disease, skeletal muscle is now being considered a key player in its pathogenesis. Although recent neurophysiological data obtained in ALS patients support an early hyperexcitability of cortical motor neurons, the dismantling of the NMJ is one of the first anatomical pathogenic events in ALS [25]. In both ALS patients and in SOD1 mouse models of ALS [26,27,28,29,30], the dismantling of the NMJ takes place before the degeneration of motor neurons [31,32,33,34], when no clinical motor signs are visible [35] (see Table 1). This supports the idea that motor neuron death is not the only cause of NMJ dismantlement but that degeneration begins in the most distal portion of the axon, namely the synapse. In addition, the safeguard of soma does not prevent the loss of NMJs [36,37]. Indeed, the destabilization of NMJs precedes motor neuron death and induces the dying-back phenomenon of motor neurons [24]. To clarify the reasons of this dismantling, several studies have been carried out. It was shown that the expression of mutant SOD1G93A and SOD1G37R specifically in muscle led to the loss of NMJ before the loss of motor neuron [38,39,40]. These data demonstrate that expression of SOD1 mutant limited to skeletal muscle is sufficient by itself to induce NMJ destabilization and may lead to motor neuron death.
In the skeletal muscle of ALS patients [41,42] and in presymptomatic Sod1G86R mice, expression of Nogo-A (neurite outgrowth inhibitor) was significantly increased and was correlated with the severity of disease [43,44]. Nogo-A could therefore participate in the destabilization of the NMJ and therefore in the degeneration of axon terminals of motor neurons. Ablation of Nogo-A in the muscle of Sod1G86R mice was shown to prevent muscle atrophy and denervation and to prolong survival by 10%. Conversely, muscle overexpression of Nogo-A induced muscle atrophy and denervation, and significantly reduced the size of the NMJ [45]. Thus, these data suggest that Nogo-A plays a role in maintaining the integrity and stabilization of neuromuscular synapses in SOD1G93A mice. However, none of these processes would, by itself, be the sole cause of ALS but they could all contribute together to induce ALS.

2. Mitochondrial Failure and Oxidative Stress in ALS

2.1. Muscle Mitochondria and Respiratory Complexes

One of the main features of ALS is mitochondrial dysfunction [19,46,47,48,49,50] (Figure 1). Defects in mitochondrial structure [51,52,53] and function have been observed in the skeletal muscle of sporadic ALS patients [54,55] and animal models of ALS [56,57] (Table 2).
Moreover, early abnormalities in mitochondrial dynamics contribute to the degeneration of motor neurons in culture [61] and may contribute to the pathophysiology of ALS [56]. In addition, mitochondrial calcium overload occurs in the nerve endings of ALS patients [29] and disruption in calcium homeostasis was seen in different cell models expressing mutant SOD1 [62,63], and in the CNS [62,64] and the skeletal muscles of SOD1G93A mice [57]. Moreover, in asymptomatic ALS mice, mitochondria of muscle cells are no longer able to regulate calcium signaling around NMJs, and an excessively high concentration of calcium in the cytosol may contribute to the progression of muscle atrophy in ALS [57,65]. These data confirm a close link between mitochondrial dysfunction and calcium deregulation, where the latter would consequently cause a defect in the mitochondrial respiratory chain, triggering a vicious cycle.
The main mitochondrial abnormalities found in ALS concern respiratory complexes. Wiedemann and colleagues [58] reported severe deregulation of the respiratory chain complex I, and a decrease in the activity of respiratory complexes I and IV in the muscle of sporadic ALS patients as well as in the muscles of the SOD1G93A mouse model from a presymptomatic stage [51,54,55,59,66,67]. Mitochondrial functions also become progressively impaired when disease progresses [55] and abnormalities in mitochondrial DNA result in decreased activity of certain enzymes (e.g., NADH, COX) [59]. In addition, a significant induction of UCP3 protein has been observed in the muscles of ALS patients and ALS mice [68]. UCP3 is an uncoupling protein mainly expressed in the mitochondria of skeletal muscles, and overexpression of UCP3 in this tissue would induce an increase in lipid oxidation (β-oxidation) and energy expenditure [69,70]. Finally, the overexpression of UCP1 in Sod1G86R mice leads to the degeneration of motor neurons, dismantling of the NMJ and decreased survival [71]. Importantly, the alterations observed in skeletal muscle can be detrimental to the integrity of the NMJ [39] without necessarily representing a causal link [72]. Despite data indicating a strong relationship between mitochondrial abnormalities and the progression of ALS, it is still impossible to establish a causal link between these two phenomena [39,71,72,73,74].

2.2. ROS and Oxidative Stress

Reactive oxygen species (ROS) are very short-lived metabolites produced during oxidative phosphorylation. Under normal physiological conditions, a cell consumes oxygen to produce energy, and at the same time must eliminate the ROS produced via defense mechanisms such as superoxide dismutase (SOD) and antioxidant metalloenzymes [75]. Under conditions of oxidative stress and reduced mitochondrial respiration, large amounts of ROS are produced and lead to cellular damage such as inflammatory response, excitotoxicity, protein aggregation and apoptosis [76,77]. Furthermore, increased β-oxidation of fatty acids leads to the generation of lipid by-products which contribute to lipotoxicity and to ROS production [78,79]. Several studies have already demonstrated the implication of oxidative stress in aging and in ALS [14,80,81]. Abnormally high levels of ROS markers were observed in fluids [82,83] and post-mortem tissues from sporadic ALS patients [84,85,86] (Figure 1). In muscle of Sod1G86R mice, oxidative stress was observed even before the onset of motor symptoms and obvious signs of denervation [87]. Moreover, in muscle of SOD1G93A, SOD1 activity was increased throughout ALS progression, indicating the presence of oxidative stress in muscle [81,88]. Dobrowolny and colleagues demonstrated that muscle expression of mutant SOD1G93A was sufficient to induce oxidative damage, muscle atrophy and dismantlement of the NMJs [38,73]. Recently, an increased production of ROS was shown in the muscle of SOD1G93A mice and in muscle of wild-type mice with transient overexpression of the SOD1G93A mutation [89]. Changes in mitochondrial functions were dependent on the progression of pathology, and the SOD1G93A mutation was found to directly contribute to mitochondrial dysfunction long before the death of motor neurons. Although the induction of oxidative stress was not sufficient to cause motor neuron death, the above evidence supports the contribution of uncontrolled ROS production in skeletal muscle to ALS development. However, other studies suggest that mitochondrial disorders in ALS are minor [66,90,91,92] but are increased as ALS escalates [55]. Altogether, these data support the hypothesis of a major involvement of oxidative stress and mitochondrial alterations in ALS progression.

3. Metabolic Alterations in Amyotrophic Lateral Sclerosis

3.1. Discovery of Hypermetabolism

Classically presented as a strict disease of the cortical, bulbar and spinal motor neurons, the alterations of skeletal muscle observed in ALS are often considered to reflect the loss of these neurons. However, many studies examining the involvement of altered energy metabolism in ALS are starting to challenge this dogma. One of the major symptoms of ALS is weight loss, which is often studied through the measurement of the body mass index (BMI). As ALS progresses, a reduction in BMI and body fat is reported in patients with ALS [93]. Early insulin resistance [94] and glucose intolerance [95] have also been reported in ALS patients. This insulin resistance, which leads to a decreased sensitivity of the peripheral tissues to insulin and limits nutrients entry into the cells, could participate to the reduced BMI seen in patients. The loss of BMI in ALS is also associated with malnutrition due to dysphagia, and a worse survival outcome [96,97], whereas a high BMI is linked to a lower risk of developing ALS [98,99]. BMI is therefore a prognostic factor for ALS [96,100,101,102,103].
In 2001, Desport et al. [104] identified an abnormal increase by 10% in resting energy expenditure in patients with ALS (n = 62; sex ratio M/F = 1.07) compared to healthy control group (n = 31). Defined as hypermetabolism, this phenomenon was subsequently confirmed, in a larger study, and was shown to be significantly increased by 14% and affected 62.3% of the 168 ALS patients (sex ratio M/F = 0.97) [105]. Data from the literature report that hypermetabolism affects up to 66% of ALS patients [104,105,106,107,108,109] and is an early event that persists throughout the course of disease [108,110,111]. Weight loss [112], hypermetabolism, and dyslipidemia are now considered as three major risk factors for ALS [107,109,113,114], and are associated with the severity of disease [108] (Figure 1). Remarkably, similar to what was seen in ALS patients, Sod1G86R and SOD1G93A mice were hypermetabolic and this metabolic change was already detectable at the clinically asymptomatic stage of the disease [115]. Moreover, an experimental induction of muscle hypermetabolism was sufficient to cause muscle denervation and motor neurons loss [71]. As such, it appears that hypermetabolism negatively impacts the progression of the disease. ALS mice also develop a loss of metabolic flexibility before any motor symptoms. This results in the inability to use glucose for energy production, leading to a decreased glycolysis and an increased β-oxidation in skeletal muscle [116] (Figure 2).
In addition, this loss of metabolic flexibility preceded hypermetabolism in SOD1G93A mice [60]. Interestingly, both Sod1G86R [116] and SOD1G93A [60] mice had marked glucose intolerance from an early stage of ALS, and glucose intolerance is, with dyslipidemia, one of the key features of metabolic dysregulations in ALS patients [117]. One of the major players in this energy imbalance is pyruvate dehydrogenase kinase 4 (PDK4), the major muscle isoform of pyruvate dehydrogenase kinase [118,119], which plays a crucial role in balancing glucose–fatty acid flux. A significant induction of this metabolic marker was found in the skeletal muscles of Sod1G86R and SOD1G93A mice but also in the muscle of ALS patients [116]. The main metabolic changes are summarized below in Table 3.
According to these observations, it can therefore be hypothesized that correcting the energy balance (glucose vs. fatty acid) in skeletal muscle would reduce or slow down the development and/or progression of ALS.

3.2. Impairment of Skeletal Muscle Metabolism by Physical Activity

The muscular system requires a considerable supply of energy to handle a variety of physical challenges. To do this, skeletal muscle adapts its energy needs continuously, according to its environment. The selection of fuel source is based on the interaction between the metabolism of glucose and that of fatty acids and is controlled by the Randle cycle [127]. During short and intense exercise (e.g., sprinting), the involvement of fast-type glycolytic fibers necessitates that glucose oxidation (glycolysis) is favored as the primary fuel pathway. Conversely, during sustained moderate exercise (e.g., a marathon, jogging), which mobilizes slow type oxidative fibers, the oxidation of fatty acids (β-oxidation) is preferred [128]. Subjected to different conditions (e.g., intensive and repeated sports activity), muscle fibers adapt and change their phenotypic profile. For example, endurance training is correlated with a strong release of fatty acids and an improved fatty acid muscle absorption [129], thus reflecting improved β-oxidation and changes in muscle fiber type composition [130]. Intense physical exercise is increasingly studied in ALS due to the large number of patients diagnosed with ALS having had a sustained athletic career (Figure 1). Several studies reported a high risk of ALS for athletes such as soccer players, baseballers or tennis players [131,132,133,134]. However, one cannot exclude that other associated factors such as exposure to pesticides, doping agents or repeated injuries could be the cause or could participate in the development of the disease [135,136,137].
However, this link between high level athletes [133] or people with an intense lifestyle [138,139] and ALS is still debated. Indeed, some studies do not observe any link between physical activity and ALS [140,141] nor report activity as being a risk factor for the disease [142] given that ALS patients who perform moderate exercise in the clinic have improved ALSFRS scores [143]. As such, considering physical activity as a risk factor is highly controversial. Based on published data, it appears that if physical activity is not directly causative of ALS, it might in some cases worsen disease progression. Interestingly, in ALS mice, the type of physical exercise performed appeared to be either deleterious or protective. Indeed, Mahoney and his colleagues showed that high intensity exercise was detrimental to motor performance and survival in male SOD1G93A mice [144]. In contrast, in the same ALS mouse line, high frequency and large amplitude exercise, such as swimming, improved motor functions, delayed the loss of motor neurons, and significantly lengthened survival [125]. In addition, swimming had significant benefits on energy metabolism in the muscle, allowing it to reuse glucose as an energy source at the expense of lipids [120], while improving muscle strength [126]. Moreover, moderate exercise (low-speed treadmill running or free access to running wheel) significantly preserved motor performance as well as motor neuron density [145,146], unlike intense exercise (high-speed treadmill running) which slightly accelerated the onset of motor disorders [145]. The differences in effects between these two types of exercise result from the documented fact that swimming causes fast twitch fiber type transition and lactate production which promotes glucose metabolism [147]. Overall, the beneficial effects of swimming can be explained by the type of fibers recruited during this type of physical exercise. Swimming preferentially solicits fast fibers while endurance exercise recruits slow fibers [147]. These data demonstrate the benefit of preserving and/or stimulating glycolytic metabolism in skeletal muscle which is compromised in ALS.

3.3. The Metabolic Switch of Muscle Fiber Types in ALS

In ALS, fast-type synaptic connections are more vulnerable as the disease progresses, while slow-type synapses are relatively spared until the end stage of the disease [31]. In 2007, Hegedus and colleagues proposed that the loss of motor units occurred before the detection of key motor symptoms and loss of motor neurons [122]. In addition, in the SOD1G93A mouse model, they showed that contraction force of the tibialis anterior (TA), a glycolytic muscle, was reduced when compared to the gastrocnemius, an oxidative muscle. This observation was correlated with a selective and progressive degeneration of motor neurons innervating glycolytic fibers twitch (especially IIB fibers) [123]. It is now accepted that there is a change in muscle fiber types from glycolytic to oxidative in muscle of ALS patients [121], of SOD1G93A mice [123,124,125] and of mice expressing SOD1 mutant specifically in skeletal muscle [38,39,73]. Interestingly, at the onset of the disease, the entry of glucose into muscle fibers was not affected [115,148], suggesting that glucose is rerouted to glycogen stores, rather than being immediately used as a source of energy. Indeed, a decrease in glycogen synthase was observed at the presymptomatic stage of disease in Sod1G86R mice, and this was associated with the deterioration of glycolysis as well as unused glycogen stores in muscle [116]. In SOD1G93A, the proportion of glycolytic fibers was reduced when compared to oxidative fibers. This was consistent with the induction of oxidative myosin heavy chains and the repression of glycolytic myosin heavy chains [60,125]. Additionally, a muscle transition from a glycolytic to oxidative phenotype was described in Sod1G86R mice throughout the development of ALS [116]. In short, these studies demonstrate that metabolic changes in skeletal muscle are a hallmark of ALS, appear before the motor symptoms in mouse models and can have consequences at the NMJ. They further show that altered energy balance plays a role in the progression of ALS.

3.4. Main Actors of the Randle Cycle

To best ensure energy homeostasis, the Randle cycle continuously adapts specific fuel usage (glycolysis vs. ß-oxidation) to cellular demand by modifying blood glucose and free fatty acid concentrations via their respective GLUT4 and FAT/CD36 transporters (Figure 2). GLUT4 expression is drastically reduced in patients with ALS [149] and SOD1G93A mice [60,120] while FAT/CD36 is significantly increased in presymptomatic Sod1G86R mice [116]. Thus, the supply of glucose to muscle fibers is no longer assured, leading to insulin resistance [150,151,152] and glucose intolerance [94,95,153,154] in ALS. At the same time, the CPT1 transporter which allows the entry of fatty acids into the mitochondria was also overexpressed in the SOD1G93A [60] and Sod1G86R models [115,116]. These data are indicative of a disturbance in the assimilation of energy substrates and highlight metabolic imbalance at the level of the muscle.
Once in the cell, glucose enters glycolysis or is stored as glycogen through the action of glycogen synthase. In Sod1G86R mice, at the asymptomatic stage, glycogen synthase activity and glycogen accumulation are significantly increased in skeletal muscle, suggesting that muscle cells are no longer able to use glycogen to produce energy, reflecting a problem with carbohydrate metabolism [116]. Regarding glycolysis, one of the first enzymes to be affected in ALS is phosphofructokinase 1 (PFK1). Indeed, from a presymptomatic stage, the expression of PFK1 and its activity were significantly reduced in the muscle of Sod1G86R mice [116]. The decrease in PFK1 expression in Sod1G86R mice could be a consequence of an increased uptake of fatty acid, enabled by induction of FAT/CD36 and CPT1 transporters, known to strongly inhibit PFK1 [155]. The reduction in PFK1 expression and activity occurs in response to an overexpression of PDK4 and high levels of PDK4 have been observed in the muscle of ALS patients, and in SOD1G93A and Sod1G86R mice even before any detectable sign of denervation [60,120]. While suggesting that the overexpression of PDK4 is not specific for mutations in the SOD1 gene, these data highlight that Randle cycle intermediates become altered in response to changes in glucose–fatty acid flux.
The expression of PDK4 depends on several transcription factors including Foxo1, PGC1α and PPARβ/δ. Foxo1 is an ubiquitous transcription factor that was strongly induced in muscles of SOD1G93A and Sod1G86R mice, as well as non-transgenic animals after sciatic nerve injury [116]. PGC1α plays an essential role in regulating the expression of genes involved in energy metabolism, lipid metabolism and in mitochondrial biogenesis [156,157,158,159,160]. Muscle overexpression of PGC1α leads to an induction of genes involved in the oxidative pathway, causing repression of glycolytic enzymes and glucose intolerance [161]. In mutant SOD1G37R mice, skeletal muscle induction of PGC1α led to maintenance of mitochondrial biogenesis, improved muscle function at the latter stages of disease [162], and was associated with an increase in oxidative type IIA fibers [163]. Alternatively, overexpression of PPARβ/δ in skeletal muscle increased the proportion of type 1 oxidative fibers [164], and constitutive overexpression of PPARβ/δ increases mitochondrial biogenesis and caused a switch from fast to slow fiber type [165].
Taken together, these studies demonstrate the crucial role of muscle energy balance in the pathogenesis of ALS, and are in agreement with the idea that correcting for, or preventing the loss of, metabolic flexibility could be a promising therapeutic approach in ALS.

4. Pharmacological Strategies for Targeting Energetic Imbalance in ALS

Below, we present a brief overview of two pharmacological modulators of energy balance that were studied for their possible repositioning as ALS treatments and a third one which is currently under investigation.

4.1. Dichloroacetate (DCA)

DCA is a drug that restores optimal glucose oxidation while inhibiting the oxidation of fatty acids [166]. DCA inhibits the activity of PDK and thus stimulates the activity of PDH [167] (Figure 3). In 2012, Miquel and colleagues found that DCA improved mitochondrial function in astrocytes expressing the SOD1G93A mutation, while preventing their toxicity on cultured motor neurons. Additionally, they found that DCA improved muscle strength, preserved the integrity of NMJs, reduced motor neuron loss and prolonged survival in SOD1G93A mice [168]. In addition, DCA also delayed the onset of motor symptoms in Sod1G86R mice by limiting denervation and muscle atrophy [116]. The benefits observed in Sod1G86R mice could be explained by the protective effect of DCA on muscle fibers by restoring the energy balance or preventing its imbalance. Treatment with DCA limited the expression of genes controlling the oxidative pathway (e.g., PDK4, Foxo1 and PPARβ/δ) and increased the expression of those involved in the glycolytic pathway (e.g., PFK1). Moreover, DCA treatment improved motor functions, reduced the expression of denervation and atrophy markers, and limited oxidative stress [116]. These data therefore prove that restoring or preserving the metabolic balance can prevent oxidative stress, protect mitochondria, and prevent denervation and muscle atrophy in Sod1G86R mice. Although DCA exerts a number of beneficial effects, translation into the clinic is limited as long-term use of DCA leads to hepatotoxicity [169].
In ALS skeletal muscle, energy metabolism is disturbed: glycolysis is decreased while β-oxidation is increased. The three presented drugs could help normalize this energy imbalance. DCA can restore glycolysis by inhibiting PDK4 activity. RAN and TMZ inhibit β-oxidation. A correction of the glucose–fatty acid balance could explain the beneficial effects of these molecules found in various studies on ALS.

4.2. Ranolazine (RAN)

RAN is a β-oxidation inhibitor approved by the FDA for the treatment of angina pectoris [170,171,172,173]. RAN increases the oxidation of glucose [174,175,176] in patients with symptoms of chronic angina pectoris [177] or congenital myotonia [178] (Figure 3). Several studies have shown its efficacy against insulin resistance in animals [179] and patients with type 2 diabetes [180,181]. Despite its beneficial effects on energy metabolism, only one group of researchers working on ALS have exploited RAN efficacy. In 2020, RAN was shown to significantly improve motor functions, restore metabolic homeostasis of skeletal muscle, and prevent hypermetabolism in SOD1G93A mice. Due to the extreme severity of ALS, the positive effects of RAN could not be sustained until the final stage of the disease and could not affect the survival of animals [60]. Therefore, the putative utility of RAN remains an open question. Further preclinical studies, followed by clinical trials, are still needed at this stage to clarify whether RAN can be used as a clinically relevant drug to cure ALS.

4.3. Trimetazidine (TMZ)

TMZ is an anti-anginal and anti-ischemic agent [182,183] that inhibits the oxidation of fatty acids and promotes the oxidation of glucose [175,184,185]. The target of TMZ is the 3-ketoacyl thiolase, an enzyme involved in the last step of β-oxidation [184] (Figure 3). Recently, Ferraro and colleagues showed improved motor performance in a mouse model of sarcopenia after TMZ treatment [186]. In addition, this molecule increased the differentiation of C2C12 myoblasts and induced myogenesis in a tumor-bearing mouse model [187]. A recent study investigating the effects of TMZ on the peripheral nervous system demonstrated an antioxidant effect of TMZ, which resulted in a microenvironment conducive to nerve regeneration and increased remyelination [188]. Based on its pharmacological properties, TMZ appears as an interesting drug to test on ALS models. In preliminary experiments, TMZ significantly increased the motor functions of Sod1G86R mice. We are currently pursuing these experiments to decipher the molecular mechanisms by which TMZ exerts its positive effects on ALS mice.

5. Open Questions and Future Directions

From recent works, it is now clear that ALS is not solely a neuronal disease but that the target of motor neurons, namely the skeletal muscle, is also a major player in disease initiation and progression. Further, the type of metabolism (e.g., glycolysis versus β-oxidation) is also decisive. There is clearly a shift in metabolism from glycolytic toward β-oxidation when disease progresses. This might account for the increased oxidative stress since, at a given ATP demand, producing the cellular fuel from β-oxidation consumes more oxygen and subsequently increases ROS production. We saw above that several molecules such as DCA, but also FDA-approved drugs such RAN or TMZ, can be repositioned for treating ALS to restore glycolysis in ALS mouse models and exert positive effects on muscle strength. Future studies should now be aimed at studying the effects of these drugs in ALS patients.

6. Conclusions

In conclusion, ALS is a highly complex disease. ALS etiology and the multiple pathophysiological mechanisms that trigger the disease remain poorly understood. However, it is clear that skeletal muscle and its bioenergetic disturbances are involved in the development of the disease. Metabolic alterations observed in skeletal muscle in patients with ALS, and in mouse models of the disease prior to motor neuron degeneration, challenge the idea of ALS being a disease that originates from the neuron. Further investigations of muscle energy metabolism are essential and necessary to define new therapeutic approaches and to develop drug candidates for treating ALS.

Funding

Not applicable.

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.

References

  1. Rowland, L.P.; Shneider, N.A. Amyotrophic Lateral Sclerosis. N. Engl. J. Med. 2001, 13. [Google Scholar] [CrossRef] [PubMed]
  2. van Es, M.A.; Hardiman, O.; Chio, A.; Al-Chalabi, A.; Pasterkamp, R.J.; Veldink, J.H.; van den Berg, L.H. Amyotrophic Lateral Sclerosis. Lancet 2017, 390, 2084–2098. [Google Scholar] [CrossRef]
  3. Petrov, D.; Mansfield, C.; Moussy, A.; Hermine, O. ALS Clinical Trials Review: 20 Years of Failure. Are We Any Closer to Registering a New Treatment? Front. Aging Neurosci. 2017, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Masrori, P.; Van Damme, P. Amyotrophic Lateral Sclerosis: A Clinical Review. Eur. J. Neurol. 2020, 27, 1918–1929. [Google Scholar] [CrossRef]
  5. Couratier, P.; Corcia, P.; Lautrette, G.; Nicol, M.; Preux, P.-M.; Marin, B. Epidemiology of Amyotrophic Lateral Sclerosis: A Review of Literature. Rev. Neurol. 2016, 172, 37–45. [Google Scholar] [CrossRef]
  6. Renton, A.E.; Chiò, A.; Traynor, B.J. State of Play in Amyotrophic Lateral Sclerosis Genetics. Nat. Neurosci. 2014, 17, 17–23. [Google Scholar] [CrossRef]
  7. Renton, A.E.; Majounie, E.; Waite, A.; Simón-Sánchez, J.; Rollinson, S.; Gibbs, J.R.; Schymick, J.C.; Laaksovirta, H.; van Swieten, J.C.; Myllykangas, L.; et al. A Hexanucleotide Repeat Expansion in C9ORF72 Is the Cause of Chromosome 9p21-Linked ALS-FTD. Neuron 2011, 72, 257–268. [Google Scholar] [CrossRef] [Green Version]
  8. Millecamps, S.; Salachas, F.; Cazeneuve, C.; Gordon, P.; Bricka, B.; Camuzat, A.; Guillot-Noel, L.; Russaouen, O.; Bruneteau, G.; Pradat, P.-F.; et al. SOD1, ANG, VAPB, TARDBP, and FUS Mutations in Familial Amyotrophic Lateral Sclerosis: Genotype-Phenotype Correlations. J. Med. Genet. 2010, 47, 554–560. [Google Scholar] [CrossRef]
  9. Millecamps, S.; Boillée, S.; Le Ber, I.; Seilhean, D.; Teyssou, E.; Giraudeau, M.; Moigneu, C.; Vandenberghe, N.; Danel-Brunaud, V.; Corcia, P.; et al. Phenotype Difference between ALS Patients with Expanded Repeats in C9ORF72 and Patients with Mutations in Other ALS-Related Genes. J. Med. Genet. 2012, 49, 258–263. [Google Scholar] [CrossRef]
  10. Kwiatkowski, T.J.; Bosco, D.A.; LeClerc, A.L.; Tamrazian, E.; Vanderburg, C.R.; Russ, C.; Davis, A.; Gilchrist, J.; Kasarskis, E.J.; Munsat, T.; et al. Mutations in the FUS/TLS Gene on Chromosome 16 Cause Familial Amyotrophic Lateral Sclerosis. Science 2009, 323, 1205–1208. [Google Scholar] [CrossRef] [Green Version]
  11. Vance, C.; Rogelj, B.; Hortobagyi, T.; De Vos, K.J.; Nishimura, A.L.; Sreedharan, J.; Hu, X.; Smith, B.; Ruddy, D.; Wright, P.; et al. Mutations in FUS, an RNA Processing Protein, Cause Familial Amyotrophic Lateral Sclerosis Type 6. Science 2009, 323, 1208–1211. [Google Scholar] [CrossRef] [Green Version]
  12. Abe, K. Safety and Efficacy of Edaravone in Well Defined Patients with Amyotrophic Lateral Sclerosis: A Randomised, Double-Blind, Placebo-Controlled Trial. Lancet Neurol. 2017, 16, 8. [Google Scholar] [CrossRef]
  13. Dorst, J.; Ludolph, A.C.; Huebers, A. Disease-Modifying and Symptomatic Treatment of Amyotrophic Lateral Sclerosis. Adv. Neurol. Disord. 2018, 11, 1756285617734734. [Google Scholar] [CrossRef]
  14. Barber, S.C.; Shaw, P.J. Oxidative Stress in ALS: Key Role in Motor Neuron Injury and Therapeutic Target. Free Radic. Biol. Med. 2010, 48, 629–641. [Google Scholar] [CrossRef]
  15. Fischer-Hayes, L.R.; Brotherton, T.; Glass, J.D. Axonal Degeneration in the Peripheral Nervous System: Implications for the Pathogenesis of Amyotrophic Lateral Sclerosis. Exp. Neurol. 2013, 246, 6–13. [Google Scholar] [CrossRef]
  16. Ling, S.-C.; Polymenidou, M.; Cleveland, D.W. Converging Mechanisms in ALS and FTD: Disrupted RNA and Protein Homeostasis. Neuron 2013, 79, 416–438. [Google Scholar] [CrossRef] [Green Version]
  17. Blokhuis, A.M.; Groen, E.J.N.; Koppers, M.; van den Berg, L.H.; Pasterkamp, R.J. Protein Aggregation in Amyotrophic Lateral Sclerosis. Acta. Neuropathol. 2013, 125, 777–794. [Google Scholar] [CrossRef] [Green Version]
  18. Bogaert, E.; d’Ydewalle, C.; Van Den Bosch, L. Amyotrophic Lateral Sclerosis and Excitotoxicity: From Pathological Mechanism to Therapeutic Target. CNSNDDT 2010, 9, 297–304. [Google Scholar] [CrossRef]
  19. Cozzolino, M.; Carrì, M.T. Mitochondrial Dysfunction in ALS. Prog. Neurobiol. 2012, 97, 54–66. [Google Scholar] [CrossRef]
  20. Mejzini, R.; Flynn, L.L.; Pitout, I.L.; Fletcher, S.; Wilton, S.D.; Akkari, P.A. ALS Genetics, Mechanisms, and Therapeutics: Where Are We Now? Front. Neurosci. 2019, 13, 1310. [Google Scholar] [CrossRef] [Green Version]
  21. Eisen, A.; Kim, S.; Pant, B. Amyotrophic Lateral Sclerosis (ALS): A Phylogenetic Disease of the Corticomotoneuron? Muscle Nerve 1992, 15, 219–224. [Google Scholar] [CrossRef]
  22. Eisen, A.; Braak, H.; Del Tredici, K.; Lemon, R.; Ludolph, A.C.; Kiernan, M.C. Cortical Influences Drive Amyotrophic Lateral Sclerosis. J. Neurol. Neurosurg. Psychiatry 2017, 88, 917–924. [Google Scholar] [CrossRef] [PubMed]
  23. Brunet, A.; Stuart-Lopez, G.; Burg, T.; Scekic-Zahirovic, J.; Rouaux, C. Cortical Circuit Dysfunction as a Potential Driver of Amyotrophic Lateral Sclerosis. Front. Neurosci. 2020, 14, 363. [Google Scholar] [CrossRef] [PubMed]
  24. Dadon-Nachum, M.; Melamed, E.; Offen, D. The “Dying-Back” Phenomenon of Motor Neurons in ALS. J. Mol. Neurosci. 2011, 43, 470–477. [Google Scholar] [CrossRef] [PubMed]
  25. Moloney, E.B.; de Winter, F.; Verhaagen, J. ALS as a Distal Axonopathy: Molecular Mechanisms Affecting Neuromuscular Junction Stability in the Presymptomatic Stages of the Disease. Front. Neurosci. 2014, 8. [Google Scholar] [CrossRef] [Green Version]
  26. Fischer, L.R.; Culver, D.G.; Tennant, P.; Davis, A.A.; Wang, M.; Castellano-Sanchez, A.; Khan, J.; Polak, M.A.; Glass, J.D. Amyotrophic Lateral Sclerosis Is a Distal Axonopathy: Evidence in Mice and Man. Exp. Neurol. 2004, 185, 232–240. [Google Scholar] [CrossRef]
  27. Gonzalez de Aguilar, J.-L.; Niederhauser-Wiederkehr, C.; Halter, B.; De Tapia, M.; Di Scala, F.; Demougin, P.; Dupuis, L.; Primig, M.; Meininger, V.; Loeffler, J.-P. Gene Profiling of Skeletal Muscle in an Amyotrophic Lateral Sclerosis Mouse Model. Physiol. Genom. 2008, 32, 207–218. [Google Scholar] [CrossRef] [Green Version]
  28. Tsujihata, M.; Hazama, R.; Yoshimura, T.; Satoh, A.; Mori, M.; Nagataki, S. The Motor End-Plate Fine Structure and Ultrastructural Localization of Acetylcholine Receptors in Amyotrophic Lateral Sclerosis. Muscle Nerve. 1984, 7, 243–249. [Google Scholar] [CrossRef]
  29. Siklós, L.; Engelhardt, J.; Harati, Y.; Smith, R.G.; Joó, F.; Appel, S.H. Ultrastructural Evidence for Altered Calcium in Motor Nerve Terminals in Amyotrophc Lateral Sclerosis: Calcium in ALS Motor Nerve Terminals. Ann. Neurol. 1996, 39, 203–216. [Google Scholar] [CrossRef]
  30. Aggarwal, A. Detection of Preclinical Motor Neurone Loss in SOD1 Mutation Carriers Using Motor Unit Number Estimation. J. Neurol. Neurosurg. Psychiatry 2002, 73, 199–201. [Google Scholar] [CrossRef] [Green Version]
  31. Frey, D.; Schneider, C.; Xu, L.; Borg, J.; Spooren, W.; Caroni, P. Early and Selective Loss of Neuromuscular Synapse Subtypes with Low Sprouting Competence in Motoneuron Diseases. J. Neurosci. 2000, 20, 2534–2542. [Google Scholar] [CrossRef]
  32. Pun, S.; Santos, A.F.; Saxena, S.; Xu, L.; Caroni, P. Selective Vulnerability and Pruning of Phasic Motoneuron Axons in Motoneuron Disease Alleviated by CNTF. Nat. Neurosci. 2006, 9, 408–419. [Google Scholar] [CrossRef]
  33. Vinsant, S.; Mansfield, C.; Jimenez-Moreno, R.; Moore, V.D.G.; Yoshikawa, M.; Hampton, T.G.; Prevette, D.; Caress, J.; Oppenheim, R.W.; Milligan, C. Characterization of Early Pathogenesis in the SOD1 G93A Mouse Model of ALS: Part II, Results and Discussion. Brain Behav. 2013, 3, 431–457. [Google Scholar] [CrossRef]
  34. Ngo, S.T.; Baumann, F.; Ridall, P.G.; Pettitt, A.N.; Henderson, R.D.; Bellingham, M.C.; McCombe, P.A. The Relationship between Bayesian Motor Unit Number Estimation and Histological Measurements of Motor Neurons in Wild-Type and SOD1G93A Mice. Clin. Neurophysiol. 2012, 123, 2080–2091. [Google Scholar] [CrossRef]
  35. Rocha, M.C.; Pousinha, P.A.; Correia, A.M.; Sebastião, A.M.; Ribeiro, J.A. Early Changes of Neuromuscular Transmission in the SOD1(G93A) Mice Model of ALS Start Long before Motor Symptoms Onset. PLoS ONE 2013, 8, e73846. [Google Scholar] [CrossRef] [Green Version]
  36. Gould, T.W.; Buss, R.R.; Vinsant, S.; Prevette, D.; Sun, W.; Knudson, C.M.; Milligan, C.E.; Oppenheim, R.W. Complete Dissociation of Motor Neuron Death from Motor Dysfunction by Bax Deletion in a Mouse Model of ALS. J. Neurosci. 2006, 26, 8774–8786. [Google Scholar] [CrossRef]
  37. Rouaux, C.; Panteleeva, I.; Rene, F.; Gonzalez de Aguilar, J.-L.; Echaniz-Laguna, A.; Dupuis, L.; Menger, Y.; Boutillier, A.-L.; Loeffler, J.-P. Sodium Valproate Exerts Neuroprotective Effects In Vivo through CREB-Binding Protein-Dependent Mechanisms But Does Not Improve Survival in an Amyotrophic Lateral Sclerosis Mouse Model. J. Neurosci. 2007, 27, 5535–5545. [Google Scholar] [CrossRef]
  38. Dobrowolny, G.; Lepore, E.; Martini, M.; Barberi, L.; Nunn, A.; Scicchitano, B.M.; Musarò, A. Metabolic Changes Associated With Muscle Expression of SOD1G93A. Front. Physiol. 2018, 9, 831. [Google Scholar] [CrossRef] [Green Version]
  39. Wong, M.; Martin, L.J. Skeletal Muscle-Restricted Expression of Human SOD1 Causes Motor Neuron Degeneration in Transgenic Mice. Hum. Mol. Genet. 2010, 19, 2284–2302. [Google Scholar] [CrossRef] [Green Version]
  40. Martin, L.J.; Wong, M. Skeletal Muscle-Restricted Expression of Human SOD1 in Transgenic Mice Causes a Fatal ALS-Like Syndrome. Front. Neurol. 2020, 11. [Google Scholar] [CrossRef]
  41. Pradat, P.-F.; Dubourg, O.; de Tapia, M.; di Scala, F.; Dupuis, L.; Lenglet, T.; Bruneteau, G.; Salachas, F.; Lacomblez, L.; Corvol, J.-C.; et al. Muscle Gene Expression Is a Marker of Amyotrophic Lateral Sclerosis Severity. Neurodegener. Dis. 2012, 9, 38–52. [Google Scholar] [CrossRef] [PubMed]
  42. Pradat, P.-F.; Bruneteau, G.; Gonzalez de Aguilar, J.-L.; Dupuis, L.; Jokic, N.; Salachas, F.; Le Forestier, N.; Echaniz-Laguna, A.; Dubourg, O.; Hauw, J.-J.; et al. Muscle Nogo-a Expression Is a Prognostic Marker in Lower Motor Neuron Syndromes. Ann. Neurol. 2007, 62, 15–20. [Google Scholar] [CrossRef] [PubMed]
  43. Jokic, N.; Gonzalez de Aguilar, J.-L.; Pradat, P.-F.; Dupuis, L.; Echaniz-Laguna, A.; Muller, A.; Dubourg, O.; Seilhean, D.; Hauw, J.-J.; Loeffler, J.-P.; et al. Nogo Expression in Muscle Correlates with Amyotrophic Lateral Sclerosis Severity. Ann. Neurol. 2005, 57, 553–556. [Google Scholar] [CrossRef] [PubMed]
  44. Dupuis, L.; Gonzalez de Aguilar, J.-L.; di Scala, F.; Rene, F.; de Tapia, M.; Pradat, P.-F.; Lacomblez, L.; Seihlan, D.; Prinjha, R.; Walsh, F.S.; et al. Nogo Provides a Molecular Marker for Diagnosis of Amyotrophic Lateral Sclerosis. Neurobiol. Dis. 2002, 10, 358–365. [Google Scholar] [CrossRef] [Green Version]
  45. Jokic, N.; Gonzalez de Aguilar, J.; Dimou, L.; Lin, S.; Fergani, A.; Ruegg, M.A.; Schwab, M.E.; Dupuis, L.; Loeffler, J. The Neurite Outgrowth Inhibitor Nogo-A Promotes Denervation in an Amyotrophic Lateral Sclerosis Model. Embo. Rep. 2006, 7, 1162–1167. [Google Scholar] [CrossRef] [Green Version]
  46. Bowling, A.C.; Schulz, J.B.; Brown, R.H.; Beal, M.F. Superoxide Dismutase Activity, Oxidative Damage, and Mitochondrial Energy Metabolism in Familial and Sporadic Amyotrophic Lateral Sclerosis. J. Neurochem. 1993, 61, 2322–2325. [Google Scholar] [CrossRef]
  47. Dupuis, L.; Gonzalez de Aguilar, J.-L.; Oudart, H.; de Tapia, M.; Barbeito, L.; Loeffler, J.-P. Mitochondria in Amyotrophic Lateral Sclerosis: A Trigger and a Target. Neurodegener. Dis. 2004, 1, 245–254. [Google Scholar] [CrossRef]
  48. Sasaki, S.; Horie, Y.; Iwata, M. Mitochondrial Alterations in Dorsal Root Ganglion Cells in Sporadic Amyotrophic Lateral Sclerosis. Acta Neuropathol. 2007, 114, 633–639. [Google Scholar] [CrossRef]
  49. Vandoorne, T.; De Bock, K.; Van Den Bosch, L. Energy Metabolism in ALS: An Underappreciated Opportunity? Acta. Neuropathol. 2018, 135, 489–509. [Google Scholar] [CrossRef] [Green Version]
  50. Carrì, M.T.; D’Ambrosi, N.; Cozzolino, M. Pathways to Mitochondrial Dysfunction in ALS Pathogenesis. Biochem. Biophys. Res. Commun. 2017, 483, 1187–1193. [Google Scholar] [CrossRef]
  51. Vielhaber, S.; Winkler, K.; Kirches, E.; Kunz, D.; Büchner, M.; Feistner, H.; Elger, C.E.; Ludolph, A.C.; Riepe, M.W.; Kunz, W.S. Visualization of Defective Mitochondrial Function in Skeletal Muscle Fibers of Patients with Sporadic Amyotrophic Lateral Sclerosis. J. Neurol. Sci. 1999, 169, 133–139. [Google Scholar] [CrossRef]
  52. Chung, M.J.; Suh, Y.-L. Ultrastructural Changes of Mitochondria in the Skeletal Muscle of Patients with Amyotrophic Lateral Sclerosis. Ultrastruct. Pathol. 2002, 26, 3–7. [Google Scholar] [CrossRef]
  53. Menzies, F.M. Mitochondrial Dysfunction in a Cell Culture Model of Familial Amyotrophic Lateral Sclerosis. Brain 2002, 125, 1522–1533. [Google Scholar] [CrossRef] [Green Version]
  54. Crugnola, V.; Lamperti, C.; Lucchini, V.; Ronchi, D.; Peverelli, L.; Prelle, A.; Sciacco, M.; Bordoni, A.; Fassone, E.; Fortunato, F.; et al. Mitochondrial Respiratory Chain Dysfunction in Muscle from Patients with Amyotrophic Lateral Sclerosis. Arch. Neurol. 2010, 67, 6. [Google Scholar] [CrossRef] [Green Version]
  55. Echaniz-Laguna, A.; Zoll, J.; Ponsot, E.; N’Guessan, B.; Tranchant, C.; Loeffler, J.-P.; Lampert, E. Muscular Mitochondrial Function in Amyotrophic Lateral Sclerosis Is Progressively Altered as the Disease Develops: A Temporal Study in Man. Exp. Neurol. 2006, 198, 25–30. [Google Scholar] [CrossRef]
  56. Luo, G.; Yi, J.; Ma, C.; Xiao, Y.; Yi, F.; Yu, T.; Zhou, J. Defective Mitochondrial Dynamics Is an Early Event in Skeletal Muscle of an Amyotrophic Lateral Sclerosis Mouse Model. PLoS ONE 2013, 8, e82112. [Google Scholar] [CrossRef]
  57. Zhou, J.; Yi, J.; Fu, R.; Liu, E.; Siddique, T.; Ríos, E.; Deng, H.-X. Hyperactive Intracellular Calcium Signaling Associated with Localized Mitochondrial Defects in Skeletal Muscle of an Animal Model of Amyotrophic Lateral Sclerosis*. J. Biol. Chem. 2010, 285, 705–712. [Google Scholar] [CrossRef] [Green Version]
  58. Wiedemann, F.R.; Winkler, K.; Kuznetsov, A.V.; Bartels, C.; Vielhaber, S.; Feistner, H.; Kunz, W.S. Impairment of Mitochondrial Function in Skeletal Muscle of Patients with Amyotrophic Lateral Sclerosis. J. Neurol. Sci. 1998, 156, 65–72. [Google Scholar] [CrossRef]
  59. Vielhaber, S. Mitochondrial DNA Abnormalities in Skeletal Muscle of Patients with Sporadic Amyotrophic Lateral Sclerosis. Brain 2000, 123, 1339–1348. [Google Scholar] [CrossRef] [Green Version]
  60. Scaricamazza, S.; Salvatori, I.; Giacovazzo, G.; Loeffler, J.P.; Renè, F.; Rosina, M.; Quessada, C.; Proietti, D.; Heil, C.; Rossi, S.; et al. Skeletal-Muscle Metabolic Reprogramming in ALS-SOD1G93A Mice Predates Disease Onset and Is A Promising Therapeutic Target. iScience 2020, 23, 101087. [Google Scholar] [CrossRef]
  61. Magrané, J.; Manfredi, G. Mitochondrial Function, Morphology, and Axonal Transport in Amyotrophic Lateral Sclerosis. Antioxid. Redox Signal. 2009, 11, 1615–1626. [Google Scholar] [CrossRef] [PubMed]
  62. Jaiswal, M.; Zech, W.-D.; Goos, M.; Leutbecher, C.; Ferri, A.; Zippelius, A.; Carrì, M.; Nau, R.; Keller, B.U. Impairment of Mitochondrial Calcium Handling in a MtSOD1 Cell Culture Model of Motoneuron Disease. BMC Neurosci. 2009, 10, 64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Carrì, M.T. Expression of a Cu, Zn Superoxide Dismutase Typical of Familial Amyotrophic Lateral Sclerosis Induces Mitochondrial Alteration and Increase of Cytosolic Ca 2+ Concentration in Transfected Neuroblastoma SH-SY5Y Cells. FEBS Lett. 1997, 414, 365–368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Damiano, M.; Starkov, A.A.; Petri, S.; Kipiani, K.; Kiaei, M.; Mattiazzi, M.; Flint Beal, M.; Manfredi, G. Neural Mitochondrial Ca 2+ Capacity Impairment Precedes the Onset of Motor Symptoms in G93A Cu/Zn-Superoxide Dismutase Mutant Mice: Mitochondrial Ca 2+ Capacity in Mutant SOD1 Mice. J. Neurochem. 2006, 96, 1349–1361. [Google Scholar] [CrossRef]
  65. Rossi, A.E.; Boncompagni, S.; Dirksen, R.T. Sarcoplasmic Reticulum-Mitochondrial Symbiosis: Bidirectional Signaling in Skeletal Muscle. Exerc. Sport Sci. Rev. 2009, 37, 29–35. [Google Scholar] [CrossRef] [Green Version]
  66. Al-Sarraj, S.; King, A.; Cleveland, M.; Pradat, P.-F.; Corse, A.; Rothstein, J.D.; Leigh, P.N.; Abila, B.; Bates, S.; Wurthner, J.; et al. Mitochondrial Abnormalities and Low Grade Inflammation Are Present in the Skeletal Muscle of a Minority of Patients with Amyotrophic Lateral Sclerosis; an Observational Myopathology Study. Acta Neuropathol. Commun. 2014, 2, 165. [Google Scholar] [CrossRef] [Green Version]
  67. Capitanio, D.; Vasso, M.; Ratti, A.; Grignaschi, G.; Volta, M.; Moriggi, M.; Daleno, C.; Bendotti, C.; Silani, V.; Gelfi, C. Molecular Signatures of Amyotrophic Lateral Sclerosis Disease Progression in Hind and Forelimb Muscles of an SOD1 G93A Mouse Model. Antioxid. Redox Signal. 2012, 17, 1333–1350. [Google Scholar] [CrossRef] [Green Version]
  68. Dupuis, L.; Scala, F.; Rene, F.; Tapia, M.; Oudart, H.; Pradat, P.-F.; Meininger, V.; Loeffler, J.-P. Up-regulation of Mitochondrial Uncoupling Protein 3 Reveals an Early Muscular Metabolic Defect in Amyotrophic Lateral Sclerosis. FASEB J. 2003, 17, 1–19. [Google Scholar] [CrossRef]
  69. Nabben, M.; Hoeks, J. Mitochondrial Uncoupling Protein 3 and Its Role in Cardiac- and Skeletal Muscle Metabolism. Physiol. Behav. 2008, 94, 259–269. [Google Scholar] [CrossRef]
  70. Aguer, C.; Fiehn, O.; Seifert, E.L.; Bézaire, V.; Meissen, J.K.; Daniels, A.; Scott, K.; Renaud, J.; Padilla, M.; Bickel, D.R.; et al. Muscle Uncoupling Protein 3 Overexpression Mimics Endurance Training and Reduces Circulating Biomarkers of Incomplete Β-oxidation. FASEB J. 2013, 27, 4213–4225. [Google Scholar] [CrossRef] [Green Version]
  71. Dupuis, L.; Gonzalez de Aguilar, J.-L.; Echaniz-Laguna, A.; Eschbach, J.; Rene, F.; Oudart, H.; Halter, B.; Huze, C.; Schaeffer, L.; Bouillaud, F.; et al. Muscle Mitochondrial Uncoupling Dismantles Neuromuscular Junction and Triggers Distal Degeneration of Motor Neurons. PLoS ONE 2009, 4, e5390. [Google Scholar] [CrossRef] [Green Version]
  72. Loeffler, J.; Picchiarelli, G.; Dupuis, L.; Gonzalez De Aguilar, J. The Role of Skeletal Muscle in Amyotrophic Lateral Sclerosis. Brain Pathol. 2016, 26, 227–236. [Google Scholar] [CrossRef]
  73. Dobrowolny, G.; Aucello, M.; Rizzuto, E.; Beccafico, S.; Mammucari, C.; Bonconpagni, S.; Belia, S.; Wannenes, F.; Nicoletti, C.; Del Prete, Z.; et al. Skeletal Muscle Is a Primary Target of SOD1G93A-Mediated Toxicity. Cell Metab. 2008, 8, 425–436. [Google Scholar] [CrossRef]
  74. Ioannides, Z.A.; Ngo, S.T.; Henderson, R.D.; McCombe, P.A.; Steyn, F.J. Altered Metabolic Homeostasis in Amyotrophic Lateral Sclerosis: Mechanisms of Energy Imbalance and Contribution to Disease Progression. Neurodegener. Dis. 2016, 16, 382–397. [Google Scholar] [CrossRef]
  75. Smith, E.F.; Shaw, P.J.; De Vos, K.J. The Role of Mitochondria in Amyotrophic Lateral Sclerosis. Neurosci. Lett. 2019, 710, 132933. [Google Scholar] [CrossRef]
  76. Kim, G.H.; Kim, J.E.; Rhie, S.J.; Yoon, S. The Role of Oxidative Stress in Neurodegenerative Diseases. Exp. Neurobiol. 2015, 24, 325–340. [Google Scholar] [CrossRef]
  77. Gandhi, S.; Abramov, A.Y. Mechanism of Oxidative Stress in Neurodegeneration. Oxidative Med. Cell. Longev. 2012, 2012, 1–11. [Google Scholar] [CrossRef] [Green Version]
  78. Zhang, L.; Keung, W.; Samokhvalov, V.; Wang, W.; Lopaschuk, G.D. Role of Fatty Acid Uptake and Fatty Acid β-Oxidation in Mediating Insulin Resistance in Heart and Skeletal Muscle. Biochim. Et Biophys. Acta (BBA)-Mol. Cell Biol. Lipids 2010, 1801, 1–22. [Google Scholar] [CrossRef]
  79. Aon, M.A.; Bhatt, N.; Cortassa, S.C. Mitochondrial and Cellular Mechanisms for Managing Lipid Excess. Front. Physiol. 2014, 5. [Google Scholar] [CrossRef] [Green Version]
  80. Barber, S.C.; Mead, R.J.; Shaw, P.J. Oxidative Stress in ALS: A Mechanism of Neurodegeneration and a Therapeutic Target. Biochim. Biophys. Acta (BBA)-Mol. Basis Dis. 2006, 1762, 1051–1067. [Google Scholar] [CrossRef] [Green Version]
  81. Mahoney, D.J.; Kaczor, J.J.; Bourgeois, J.; Yasuda, N.; Tarnopolsky, M.A. Oxidative Stress and Antioxidant Enzyme Upregulation in SOD1-G93A Mouse Skeletal Muscle. Muscle Nerve. 2006, 33, 809–816. [Google Scholar] [CrossRef] [PubMed]
  82. Smith, R.G.; Henry, Y.K.; Mattson, M.P.; Appel, S.H. Presence of 4-Hydroxynonenal in Cerebrospinal Fluid of Patients with Sporadic Amyotrophic Lateral Sclerosis. Ann. Neurol. 1998, 44, 696–699. [Google Scholar] [CrossRef] [PubMed]
  83. Mitsumoto, H.; Santella, R.M.; Liu, X.; Bogdanov, M.; Zipprich, J.; Wu, H.-C.; Mahata, J.; Kilty, M.; Bednarz, K.; Bell, D.; et al. Oxidative Stress Biomarkers in Sporadic ALS. Amyotroph. Lateral Scler. 2008, 9, 177–183. [Google Scholar] [CrossRef] [PubMed]
  84. Shaw, P.J.; Ince, P.G.; Falkous, G.; Mantle, D. Oxidative Damage to Protein in Sporadic Motor Neuron Disease Spinal Cord. Ann. Neurol. 1995, 38, 691–695. [Google Scholar] [CrossRef] [PubMed]
  85. Shibata, N.; Nagai, R.; Uchida, K.; Horiuchi, S.; Yamada, S.; Hirano, A.; Kawaguchi, M.; Yamamoto, T.; Sasaki, S.; Kobayashi, M. Morphological Evidence for Lipid Peroxidation and Protein Glycoxidation in Spinal Cords from Sporadic Amyotrophic Lateral Sclerosis Patients. Brain Res. 2001, 917, 97–104. [Google Scholar] [CrossRef]
  86. Chang, Y.; Kong, Q.; Shan, X.; Tian, G.; Ilieva, H.; Cleveland, D.W.; Rothstein, J.D.; Borchelt, D.R.; Wong, P.C.; Lin, C.G. Messenger RNA Oxidation Occurs Early in Disease Pathogenesis and Promotes Motor Neuron Degeneration in ALS. PLoS ONE 2008, 3, e2849. [Google Scholar] [CrossRef] [PubMed]
  87. Halter, B.; Gonzalez de Aguilar, J.-L.; Rene, F.; Petri, S.; Fricker, B.; Echaniz-Laguna, A.; Dupuis, L.; Larmet, Y.; Loeffler, J.-P. Oxidative Stress in Skeletal Muscle Stimulates Early Expression of Rad in a Mouse Model of Amyotrophic Lateral Sclerosis. Free Radic. Biol. Med. 2010, 48, 915–923. [Google Scholar] [CrossRef]
  88. Leclerc, N.; Ribera, F.; Zoll, J.; Warter, J.-M.; Poindron, P.; Lampert, E.; Borg, J. Selective Changes in Mitochondria Respiratory Properties in Oxidative or Glycolytic Muscle Fibers Isolated from G93AhumanSOD1 Transgenic Mice. Neuromuscul. Disord. 2001, 11, 722–727. [Google Scholar] [CrossRef]
  89. Xiao, Y.; Karam, C.; Yi, J.; Zhang, L.; Li, X.; Yoon, D.; Wang, H.; Dhakal, K.; Ramlow, P.; Yu, T.; et al. ROS-Related Mitochondrial Dysfunction in Skeletal Muscle of an ALS Mouse Model during the Disease Progression. Pharmacol. Res. 2018, 138, 25–36. [Google Scholar] [CrossRef]
  90. Echaniz-Laguna, A.; Zoll, J.; Ribera, F.; Tranchant, C.; Warter, J.-M.; Lonsdorfer, J.; Lampert, E. Mitochondrial Respiratory Chain Function in Skeletal Muscle of ALS Patients. Ann. Neurol. 2002, 52, 623–627. [Google Scholar] [CrossRef]
  91. Krasnianski, A.; Deschauer, M.; Neudecker, S.; Gellerich, F.N.; Müller, T.; Schoser, B.G.; Krasnianski, M.; Zierz, S. Mitochondrial Changes in Skeletal Muscle in Amyotrophic Lateral Sclerosis and Other Neurogenic Atrophies. Brain 2005, 128, 1870–1876. [Google Scholar] [CrossRef]
  92. Ryan, T.E.; Erickson, M.L.; Verma, A.; Chavez, J.; Rivner, M.H.; Mccully, K.K. Skeletal Muscle Oxidative Capacity in Amyotrophic Lateral Sclerosis: Muscle Metabolism in ALS. Muscle Nerve. 2014, 50, 767–774. [Google Scholar] [CrossRef] [PubMed]
  93. Kasarskis, E.J.; Berryman, S.; Vanderleest, J.G.; Schneider, A.R.; McClain, C.J. Nutritional Status of Patients with Amyotrophic Lateral Sclerosis: Relation to the Proximity of Death. Am. J. Clin. Nutr. 1996, 63, 130–137. [Google Scholar] [CrossRef] [Green Version]
  94. Reyes, E.T.; Perurena, O.H.; Festoff, B.W.; Jorgensen, R.; Moore, W.V. Insulin Resistance in Amyotrophic Lateral Sclerosis. J. Neurol. Sci. 1984, 63, 317–324. [Google Scholar] [CrossRef]
  95. Pradat, P.-F.; Bruneteau, G.; Gordon, P.H.; Dupuis, L.; Bonnefont-Rousselot, D.; Simon, D.; Salachas, F.; Corcia, P.; Frochot, V.; Lacorte, J.-M.; et al. Impaired Glucose Tolerance in Patients with Amyotrophic Lateral Sclerosis. Amyotroph. Lateral Scler. 2010, 11, 166–171. [Google Scholar] [CrossRef]
  96. Desport, J.C.; Preux, P.M.; Truong, T.C.; Vallat, J.M.; Sautereau, D.; Couratier, P. Nutritional Status Is a Prognostic Factor for Survival in ALS Patients. Neurology 1999, 53, 1059. [Google Scholar] [CrossRef]
  97. Marin, B.; Desport, J.C.; Kajeu, P.; Jesus, P.; Nicolaud, B.; Nicol, M.; Preux, P.M.; Couratier, P. Alteration of Nutritional Status at Diagnosis Is a Prognostic Factor for Survival of Amyotrophic Lateral Sclerosis Patients. J. Neurol. Neurosurg. Psychiatry 2011, 82, 628–634. [Google Scholar] [CrossRef] [Green Version]
  98. Nakken, O.; Meyer, H.E.; Stigum, H.; Holmøy, T. High BMI Is Associated with Low ALS Risk: A Population-Based Study. Neurology 2019, 93, e424–e432. [Google Scholar] [CrossRef]
  99. Pape, J.A.; Grose, J.H. The Effects of Diet and Sex in Amyotrophic Lateral Sclerosis. Rev. Neurol. 2020, 176, 301–315. [Google Scholar] [CrossRef]
  100. Jawaid, A.; Murthy, S.B.; Wilson, A.M.; Qureshi, S.U.; Amro, M.J.; Wheaton, M.; Simpson, E.; Harati, Y.; Strutt, A.M.; York, M.K.; et al. A Decrease in Body Mass Index Is Associated with Faster Progression of Motor Symptoms and Shorter Survival in ALS. Amyotroph. Lateral Scler. 2010, 11, 542–548. [Google Scholar] [CrossRef]
  101. Paganoni, S.; Deng, J.; Jaffa, M.; Cudkowicz, M.E.; Wills, A.-M. Body Mass Index, Not Dyslipidemia, Is an Independent Predictor of Survival in Amyotrophic Lateral Sclerosis. Muscle Nerve. 2011, 44, 20–24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Dardiotis, E.; Siokas, V.; Sokratous, M.; Tsouris, Z.; Aloizou, A.-M.; Florou, D.; Dastamani, M.; Mentis, A.-F.A.; Brotis, A.G. Body Mass Index and Survival from Amyotrophic Lateral Sclerosis: A Meta-Analysis. Neurol. Clin. Pract. 2018, 8, 437–444. [Google Scholar] [CrossRef] [PubMed]
  103. O’Reilly, É.J.; Wang, H.; Weisskopf, M.G.; Fitzgerald, K.C.; Falcone, G.; McCullough, M.L.; Thun, M.; Park, Y.; Kolonel, L.N.; Ascherio, A. Premorbid Body Mass Index and Risk of Amyotrophic Lateral Sclerosis. Amyotroph. Lateral Scler. Front. Degener. 2013, 14, 205–211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Desport, J.C.; Preux, P.M.; Magy, L.; Boirie, Y.; Vallat, J.M.; Beaufrère, B.; Couratier, P. Factors Correlated with Hypermetabolism in Patients with Amyotrophic Lateral Sclerosis. Am. J. Clin. Nutr. 2001, 74, 328–334. [Google Scholar] [CrossRef] [Green Version]
  105. Desport, J.-C.; Torny, F.; Lacoste, M.; Preux, P.-M.; Couratier, P. Hypermetabolism in ALS: Correlations with Clinical and Paraclinical Parameters. Neurodegener. Dis. 2005, 2, 202–207. [Google Scholar] [CrossRef]
  106. Fayemendy, P.; Marin, B.; Labrunie, A.; Boirie, Y.; Walrand, S.; Achamrah, N.; Coëffier, M.; Preux, P.-M.; Lautrette, G.; Desport, J.-C.; et al. Hypermetabolism Is a Reality in Amyotrophic Lateral Sclerosis Compared to Healthy Subjects. J. Neurol. Sci. 2021, 420, 117257. [Google Scholar] [CrossRef]
  107. Jésus, P.; Fayemendy, P.; Nicol, M.; Lautrette, G.; Sourisseau, H.; Preux, P.-M.; Desport, J.-C.; Marin, B.; Couratier, P. Hypermetabolism Is a Deleterious Prognostic Factor in Patients with Amyotrophic Lateral Sclerosis. Eur. J. Neurol. 2018, 25, 97–104. [Google Scholar] [CrossRef]
  108. Bouteloup, C.; Desport, J.-C.; Clavelou, P.; Guy, N.; Derumeaux-Burel, H.; Ferrier, A.; Couratier, P. Hypermetabolism in ALS Patients: An Early and Persistent Phenomenon. J. Neurol. 2009, 256, 1236–1242. [Google Scholar] [CrossRef]
  109. Steyn, F.J.; Ioannides, Z.A.; van Eijk, R.P.A.; Heggie, S.; Thorpe, K.A.; Ceslis, A.; Heshmat, S.; Henders, A.K.; Wray, N.R.; van den Berg, L.H.; et al. Hypermetabolism in ALS Is Associated with Greater Functional Decline and Shorter Survival. J. Neurol. Neurosurg. Psychiatry 2018, 89, 1016–1023. [Google Scholar] [CrossRef] [Green Version]
  110. Funalot, B.; Desport, J.-C.; Sturtz, F.; Camu, W.; Couratier, P. High Metabolic Level in Patients with Familial Amyotrophic Lateral Sclerosis. Amyotroph. Lateral Scler. 2009, 10, 113–117. [Google Scholar] [CrossRef]
  111. Peter, R.S.; Rosenbohm, A.; Dupuis, L.; Brehme, T.; Kassubek, J.; Rothenbacher, D.; Nagel, G.; Ludolph, A.C. Life Course Body Mass Index and Risk and Prognosis of Amyotrophic Lateral Sclerosis: Results from the ALS Registry Swabia. Eur. J. Epidemiol. 2017, 32, 901–908. [Google Scholar] [CrossRef]
  112. Moglia, C.; Calvo, A.; Grassano, M.; Canosa, A.; Manera, U.; D’Ovidio, F.; Bombaci, A.; Bersano, E.; Mazzini, L.; Mora, G.; et al. Early Weight Loss in Amyotrophic Lateral Sclerosis: Outcome Relevance and Clinical Correlates in a Population-Based Cohort. J. Neurol. Neurosurg. Psychiatry 2019, 90, 666–673. [Google Scholar] [CrossRef]
  113. Dupuis, L.; Corcia, P.; Fergani, A. Dyslipidemia Is a Protective Factor in Amyotrophic Lateral Sclerosis. Neurology 2008, 7, 1004–1009. [Google Scholar] [CrossRef]
  114. Ngo, S.T.; Steyn, F.J.; Huang, L.; Mantovani, S.; Pfluger, C.M.M.; Woodruff, T.M.; O’Sullivan, J.D.; Henderson, R.D.; McCombe, P.A. Altered Expression of Metabolic Proteins and Adipokines in Patients with Amyotrophic Lateral Sclerosis. J. Neurol. Sci. 2015, 357, 22–27. [Google Scholar] [CrossRef] [Green Version]
  115. Dupuis, L.; Oudart, H.; Rene, F.; de Aguilar, J.-L.G.; Loeffler, J.-P. Evidence for Defective Energy Homeostasis in Amyotrophic Lateral Sclerosis: Benefit of a High-Energy Diet in a Transgenic Mouse Model. Proc. Natl. Acad. Sci. USA 2004, 101, 11159–11164. [Google Scholar] [CrossRef] [Green Version]
  116. Palamiuc, L.; Schlagowski, A.; Ngo, S.T.; Vernay, A.; Dirrig-Grosch, S.; Henriques, A.; Boutillier, A.; Zoll, J.; Echaniz-Laguna, A.; Loeffler, J.; et al. A Metabolic Switch toward Lipid Use in Glycolytic Muscle Is an Early Pathologic Event in a Mouse Model of Amyotrophic Lateral Sclerosis. Embo. Mol. Med. 2015, 7, 526–546. [Google Scholar] [CrossRef]
  117. Joardar, A.; Manzo, E.; Zarnescu, D.C. Metabolic Dysregulation in Amyotrophic Lateral Sclerosis: Challenges and Opportunities. Curr. Genet. Med. Rep. 2017, 5, 108–114. [Google Scholar] [CrossRef]
  118. Jeoung, N.H.; Wu, P.; Joshi, M.A.; Jaskiewicz, J.; Bock, C.B.; Depaoli-Roach, A.A.; Harris, R.A. Role of Pyruvate Dehydrogenase Kinase Isoenzyme 4 (PDHK4) in Glucose Homoeostasis during Starvation. Biochem. J. 2006, 397, 417–425. [Google Scholar] [CrossRef] [Green Version]
  119. Zhang, S.; Hulver, M.W.; McMillan, R.P.; Cline, M.A.; Gilbert, E.R. The Pivotal Role of Pyruvate Dehydrogenase Kinases in Metabolic Flexibility. Nutr. Metab. (Lond.) 2014, 11, 10. [Google Scholar] [CrossRef] [Green Version]
  120. Desseille, C.; Deforges, S.; Biondi, O.; Houdebine, L.; D’amico, D.; Lamazière, A.; Caradeuc, C.; Bertho, G.; Bruneteau, G.; Weill, L.; et al. Specific Physical Exercise Improves Energetic Metabolism in the Skeletal Muscle of Amyotrophic-Lateral- Sclerosis Mice. Front. Mol. Neurosci. 2017, 10, 332. [Google Scholar] [CrossRef]
  121. Telerman-Toppet, N.; Coërs, C. Motor Innervation and Fiber Type Pattern in Amyotrophic Lateral Sclerosis and in Charcot-Marie-Tooth Disease: Motor Units in ALS and CMT. Muscle Nerve. 1978, 1, 133–139. [Google Scholar] [CrossRef] [PubMed]
  122. Hegedus, J.; Putman, C.T.; Gordon, T. Time Course of Preferential Motor Unit Loss in the SOD1G93A Mouse Model of Amyotrophic Lateral Sclerosis. Neurobiol. Dis. 2007, 28, 154–164. [Google Scholar] [CrossRef] [PubMed]
  123. Hegedus, J.; Putman, C.T.; Tyreman, N.; Gordon, T. Preferential Motor Unit Loss in the SOD1 G93A Transgenic Mouse Model of Amyotrophic Lateral Sclerosis: Motor Units in Mouse ALS. J. Physiol. 2008, 586, 3337–3351. [Google Scholar] [CrossRef] [PubMed]
  124. Atkin, J.D.; Scott, R.L.; West, J.M.; Lopes, E.; Quah, A.K.J.; Cheema, S.S. Properties of Slow- and Fast-Twitch Muscle Fibres in a Mouse Model of Amyotrophic Lateral Sclerosis. Neuromuscul. Disord. 2005, 15, 377–388. [Google Scholar] [CrossRef] [PubMed]
  125. Deforges, S.; Branchu, J.; Biondi, O.; Grondard, C.; Pariset, C.; Lécolle, S.; Lopes, P.; Vidal, P.-P.; Chanoine, C.; Charbonnier, F. Motoneuron Survival Is Promoted by Specific Exercise in a Mouse Model of Amyotrophic Lateral Sclerosis: Motoneuron Activation and Neuroprotection in ALS Mice. J. Physiol. 2009, 587, 3561–3572. [Google Scholar] [CrossRef] [PubMed]
  126. Flis, D.J.; Dzik, K.; Kaczor, J.J.; Cieminski, K.; Halon-Golabek, M.; Antosiewicz, J.; Wieckowski, M.R.; Ziolkowski, W. Swim Training Modulates Mouse Skeletal Muscle Energy Metabolism and Ameliorates Reduction in Grip Strength in a Mouse Model of Amyotrophic Lateral Sclerosis. IJMS 2019, 20, 233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Randle, P.J.; Garland, P.B.; Hales, C.N.; Newsholme, E.A. The Glucose Fatty-Acid Cycle. Its Role in Insulin Sensitivity and the Metabolic Disturbances of Diabetes Mellitus. Lancet 1963, 1, 785–789. [Google Scholar] [CrossRef]
  128. Bassel-Duby, R.; Olson, E.N. Signaling Pathways in Skeletal Muscle Remodeling. Annu. Rev. Biochem. 2006, 75, 19–37. [Google Scholar] [CrossRef] [Green Version]
  129. Talanian, J.L.; Holloway, G.P.; Snook, L.A.; Heigenhauser, G.J.F.; Bonen, A.; Spriet, L.L. Exercise Training Increases Sarcolemmal and Mitochondrial Fatty Acid Transport Proteins in Human Skeletal Muscle. Endocrinol. Metab. 2010, 299, 9. [Google Scholar] [CrossRef] [Green Version]
  130. Pette, D.; Staron, R.S. Transitions of Muscle Fiber Phenotypic Profiles. Histochem Cell Biol. 2001, 115, 359–372. [Google Scholar] [CrossRef]
  131. Lacorte, E.; Ferrigno, L.; Leoncini, E.; Corbo, M.; Boccia, S.; Vanacore, N. Physical Activity, and Physical Activity Related to Sports, Leisure and Occupational Activity as Risk Factors for ALS: A Systematic Review. Neurosci. Biobehav. Rev. 2016, 66, 61–79. [Google Scholar] [CrossRef]
  132. Bozzoni, V. Amyotrophic Lateral Sclerosis and Environmental Factors. Funct. Neurol. 2016, 31. [Google Scholar] [CrossRef]
  133. Chio, A. Severely Increased Risk of Amyotrophic Lateral Sclerosis among Italian Professional Football Players. Brain 2005, 128, 472–476. [Google Scholar] [CrossRef] [Green Version]
  134. Chio, A.; Mora, G.; Calvo, A.; Mazzini, L.; Bottacchi, E.; Mutani, R. On behalf of the PARALS Epidemiology of ALS in Italy: A 10-Year Prospective Population-Based Study. Neurology 2009, 72, 725–731. [Google Scholar] [CrossRef]
  135. Beghi, E. Are Professional Soccer Players at Higher Risk for ALS? Amyotroph. Lateral Scler. Front. Degener. 2013, 14, 501–506. [Google Scholar] [CrossRef]
  136. Piazza, O.; Sirén, A.-L.; Ehrenreich, H. Soccer, Neurotrauma and Amyotrophic Lateral Sclerosis: Is There a Connection? Curr. Med. Res. Opin. 2004, 20, 505–508. [Google Scholar] [CrossRef]
  137. Al-Chalabi, A.; Hardiman, O. The Epidemiology of ALS: A Conspiracy of Genes, Environment and Time. Nat. Rev. Neurol. 2013, 9, 617–628. [Google Scholar] [CrossRef]
  138. Scarmeas, N.; Shih, T.; Stern, Y.; Ottman, R.; Rowland, L.P. Premorbid Weight, Body Mass, and Varsity Athletics in ALS. Neurology 2002, 59, 773–775. [Google Scholar] [CrossRef] [Green Version]
  139. Huisman, M.H.B.; Seelen, M.; de Jong, S.W.; Dorresteijn, K.R.I.S.; van Doormaal, P.T.C.; van der Kooi, A.J.; de Visser, M.; Schelhaas, H.J.; van den Berg, L.H.; Veldink, J.H. Lifetime Physical Activity and the Risk of Amyotrophic Lateral Sclerosis. J. Neurol. Neurosurg. Psychiatry 2013, 84, 976–981. [Google Scholar] [CrossRef] [Green Version]
  140. Longstreth, W.T.; McGuire, V.; Koepsell, T.D.; Wang, Y.; van Belle, G. Risk of Amyotrophic Lateral Sclerosis and History of Physical Activity: A Population-Based Case-Control Study. Arch Neurol 1998, 55, 201. [Google Scholar] [CrossRef] [Green Version]
  141. Valenti, M.; Pontieri, F.E.; Conti, F.; Altobelli, E.; Manzoni, T.; Frati, L. Amyotrophic Lateral Sclerosis and Sports: A Case-Control Study. Eur. J. Neurol. 2005, 12, 223–225. [Google Scholar] [CrossRef] [PubMed]
  142. Turner, M.R.; Goldacre, R.; Ramagopalan, S.; Talbot, K.; Goldacre, M.J. Autoimmune Disease Preceding Amyotrophic Lateral Sclerosis: An Epidemiologic Study. Neurology 2013, 81, 1222–1225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Dal Bello-Haas, V.; Florence, J.M. Therapeutic Exercise for People with Amyotrophic Lateral Sclerosis or Motor Neuron Disease. Cochrane Database Syst. Rev. 2013. [Google Scholar] [CrossRef] [PubMed]
  144. Mahoney, D.J.; Rodriguez, C.; Devries, M.; Yasuda, N.; Tarnopolsky, M.A. Effects of High-Intensity Endurance Exercise Training in the G93A Mouse Model of Amyotrophic Lateral Sclerosis. Muscle Nerve. 2004, 29, 656–662. [Google Scholar] [CrossRef]
  145. Carreras, I.; Yuruker, S.; Aytan, N.; Hossain, L.; Choi, J.-K.; Jenkins, B.G.; Kowall, N.W.; Dedeoglu, A. Moderate Exercise Delays the Motor Performance Decline in a Transgenic Model of ALS. Brain Res. 2010, 1313, 192–201. [Google Scholar] [CrossRef] [Green Version]
  146. Kaspar, B.K.; Frost, L.M.; Christian, L.; Umapathi, P.; Gage, F.H. Synergy of Insulin-like Growth Factor-1 and Exercise in Amyotrophic Lateral Sclerosis. Ann. Neurol. 2005, 57, 649–655. [Google Scholar] [CrossRef]
  147. Grondard, C.; Biondi, O.; Pariset, C.; Lopes, P.; Deforges, S.; Lécolle, S.; Gaspera, B.D.; Gallien, C.-L.; Chanoine, C.; Charbonnier, F. Exercise-Induced Modulation of Calcineurin Activity Parallels the Time Course of Myofibre Transitions. J. Cell. Physiol. 2008, 214, 126–135. [Google Scholar] [CrossRef]
  148. Smittkamp, S.E.; Morris, J.K.; Bomhoff, G.L.; Chertoff, M.E.; Geiger, P.C.; Stanford, J.A. SOD1-G93A Mice Exhibit Muscle-Fiber-Type-Specific Decreases in Glucose Uptake in the Absence of Whole-Body Changes in Metabolism. Neurodegener. Dis. 2013. [Google Scholar] [CrossRef]
  149. Zubiri, I.; Lombardi, V.; Bremang, M.; Mitra, V.; Nardo, G.; Adiutori, R.; Lu, C.-H.; Leoni, E.; Yip, P.; Yildiz, O.; et al. Tissue-Enhanced Plasma Proteomic Analysis for Disease Stratification in Amyotrophic Lateral Sclerosis. Mol. Neurodegener. 2018, 13, 60. [Google Scholar] [CrossRef]
  150. Ren, J.M.; Marshall, B.A.; Mueckler, M.M.; McCaleb, M.; Amatruda, J.M.; Shulman, G.I. Overexpression of Glut4 Protein in Muscle Increases Basal and Insulin-Stimulated Whole Body Glucose Disposal in Conscious Mice. J. Clin. Investig. 1995, 95, 429–432. [Google Scholar] [CrossRef] [Green Version]
  151. Leturque, A.; Loizeau, M.; Vaulont, S.; Salminen, M.; Girard, J. Improvement of Insulin Action in Diabetic Transgenic Mice Selectively Overexpressing GLUT4 in Skeletal Muscle. Diabetes 1996, 45, 23–27. [Google Scholar] [CrossRef]
  152. Tsao, T.-S.; Burcelin, R.; Katz, E.B.; Huang, L.; Charron, M.J. Enhanced Insulin Action Due to Targeted GLUT4 Overexpression Exclusively in Muscle. Diabetes 1996, 45, 28–36. [Google Scholar] [CrossRef]
  153. Dupuis, L.; Pradat, P.-F.; Ludolph, A.C.; Loeffler, J.-P. Energy Metabolism in Amyotrophic Lateral Sclerosis. Lancet. Neurol. 2011, 10, 75–82. [Google Scholar] [CrossRef]
  154. Saccà, F.; Quarantelli, M.; Rinaldi, C.; Tucci, T.; Piro, R.; Perrotta, G.; Carotenuto, B.; Marsili, A.; Palma, V.; Michele, G.; et al. A Randomized Controlled Clinical Trial of Growth Hormone in Amyotrophic Lateral Sclerosis: Clinical, Neuroimaging, and Hormonal Results. J. Neurol. 2012, 259, 132–138. [Google Scholar] [CrossRef]
  155. Massao Hirabara, S.; de Oliveira Carvalho, C.R.; Mendonça, J.R.; Piltcher Haber, E.; Fernandes, L.C.; Curi, R. Palmitate Acutely Raises Glycogen Synthesis in Rat Soleus Muscle by a Mechanism That Requires Its Metabolization (Randle Cycle). FEBS Lett. 2003, 541, 109–114. [Google Scholar] [CrossRef] [Green Version]
  156. Puigserver, P.; Spiegelman, B.M. Peroxisome Proliferator-Activated Receptor-γ Coactivator 1α (PGC-1α): Transcriptional Coactivator and Metabolic Regulator. Endocr. Rev. 2003, 24, 78–90. [Google Scholar] [CrossRef] [Green Version]
  157. Thau, N.; Knippenberg, S.; Korner, S.; Rath, K.J.; Dengler, R.; Petri, S. Decreased MRNA Expression of PGC-1> and PGC-1>YRegulated Factors in the SOD1G93A ALS Mouse Model and in Human Sporadic ALS. J. Neuropathol. Exp. Neurol. 2012, 71, 11. [Google Scholar] [CrossRef] [Green Version]
  158. Cho, Y.; Hazen, B.C.; Russell, A.P.; Kralli, A. Peroxisome Proliferator-Activated Receptor γ Coactivator 1 (PGC-1)- and Estrogen-Related Receptor (ERR)-Induced Regulator in Muscle 1 (PERM1) Is a Tissue-Specific Regulator of Oxidative Capacity in Skeletal Muscle Cells. J. Biol. Chem. 2013, 288, 25207–25218. [Google Scholar] [CrossRef] [Green Version]
  159. Cresci, S.; Wright, L.D.; Spratt, J.A.; Briggs, F.N.; Kelly, D.P. Activation of a Novel Metabolic Gene Regulatory Pathway by Chronic Stimulation of Skeletal Muscle. Am. J. Physiol. Cell Physiol. 1996, 270, C1413–C1420. [Google Scholar] [CrossRef]
  160. Gulick, T.; Cresci, S.; Caira, T.; Moore, D.D.; Kelly, D.P. The Peroxisome Proliferator-Activated Receptor Regulates Mitochondrial Fatty Acid Oxidative Enzyme Gene Expression. Proc. Natl. Acad. Sci. USA 1994, 91, 11012–11016. [Google Scholar] [CrossRef] [Green Version]
  161. Finck, B.N.; Bernal-Mizrachi, C.; Han, D.H.; Coleman, T.; Sambandam, N.; LaRiviere, L.L.; Holloszy, J.O.; Semenkovich, C.F.; Kelly, D.P. A Potential Link between Muscle Peroxisome Proliferator- Activated Receptor-α Signaling and Obesity-Related Diabetes. Cell Metab. 2005, 1, 133–144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Da Cruz, S.; Parone, P.A.; Lopes, V.S.; Lillo, C.; McAlonis-Downes, M.; Lee, S.K.; Vetto, A.P.; Petrosyan, S.; Marsala, M.; Murphy, A.N.; et al. Elevated PGC-1α Activity Sustains Mitochondrial Biogenesis and Muscle Function without Extending Survival in a Mouse Model of Inherited ALS. Cell Metab. 2012, 15, 778–786. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Lin, J.; Wu, H.; Tarr, P.T.; Zhang, C.-Y.; Wu, Z.; Boss, O.; Michael, L.F.; Puigserver, P.; Isotani, E.; Olson, E.N.; et al. Transcriptional Co-Activator PGC-1α Drives the Formation of Slow-Twitch Muscle Fibres. Nature 2002, 418, 797–801. [Google Scholar] [CrossRef] [PubMed]
  164. Luquet, S.; Lopez-Soriano, J.; Holst, D.; Fredenrich, A.; Melki, J.; Rassoulzadegan, M.; Grimaldi, P.A. Peroxisome Proliferator-activated Receptor δ Controls Muscle Development and Oxydative Capability. FASEB J. 2003, 17, 2299–2301. [Google Scholar] [CrossRef]
  165. Wang, Y.-X.; Zhang, C.-L.; Yu, R.T.; Cho, H.K.; Nelson, M.C.; Bayuga-Ocampo, C.R.; Ham, J.; Kang, H.; Evans, R.M. Regulation of Muscle Fiber Type and Running Endurance by PPARδ. PLoS Biol. 2004, 2, e294. [Google Scholar] [CrossRef] [Green Version]
  166. McVeigh, J.J.; Lopaschuk, G.D. Dichloroacetate Stimulation of Glucose Oxidation Improves Recovery of Ischemic Rat Hearts. Am. J. Physiol.-Heart Circ. Physiol. 1990, 259, H1079–H1085. [Google Scholar] [CrossRef]
  167. Whitehouse, S.; Randle, P.J. Activation of Pyruvate Dehydrogenase in Perfused Rat Heart by Dichloroacetate (Short Communication). Biochem. J. 1973, 134, 651–653. [Google Scholar] [CrossRef] [Green Version]
  168. Miquel, E.; Cassina, A.; Martínez-Palma, L.; Bolatto, C.; Trías, E.; Gandelman, M.; Radi, R.; Barbeito, L.; Cassina, P. Modulation of Astrocytic Mitochondrial Function by Dichloroacetate Improves Survival and Motor Performance in Inherited Amyotrophic Lateral Sclerosis. PLoS ONE 2012, 7, e34776. [Google Scholar] [CrossRef]
  169. DeAngelo, A.B.; George, M.H.; House, D.E. Hepatocarcinogenicity In The Male B6c3f1 Mouse Following A Lifetime Exposure To Dichloroacetic Acid In The Drinking Water: Dose-Response Determination And Modes Of Action. J. Toxicol. Environ. Health 1999, 58, 485–507. [Google Scholar] [CrossRef]
  170. Lionetti, V.; Stanley, W.C.; Recchia, F.A. Modulating Fatty Acid Oxidation in Heart Failure. Cardiovasc. Res. 2011, 90, 202–209. [Google Scholar] [CrossRef]
  171. McCormack, J.G.; Stanley, W.C.; Wolff, A.A. Ranolazine: A Novel Metabolic Modulator for the Treatment of Angina. Gen. Pharmacol. Vasc. Syst. 1998, 30, 639–645. [Google Scholar] [CrossRef]
  172. Tafreshi, M.J.; Fisher, E. Ranolazine: A New Approach to Management of Patients with Angina. Ann. Pharm. 2006, 40, 689–693. [Google Scholar] [CrossRef]
  173. Hill, J.A.; Schofield, R.S. The Use of Ranolazine in Cardiovascular Disease. Expert Opin. Investig. Drugs 2002, 11, 117–123. [Google Scholar] [CrossRef]
  174. Clarke, B. Ranolazine Increases Active Pyruvate Dehydrogenase in Perfused Normoxic Rat Hearts: Evidence for an Indirect Mechanism. J. Mol. Cell. Cardiol. 1996, 28, 341–350. [Google Scholar] [CrossRef]
  175. Stanley, W.C. Partial Fatty Acid Oxidation Inhibitors for Stable Angina. Expert Opin. Investig. Drugs 2002, 11, 615–629. [Google Scholar] [CrossRef]
  176. Mourouzis, I.; Mantzouratou, P.; Galanopoulos, G.; Kostakou, E.; Dhalla, A.K.; Belardinelli, L.; Pantos, C. The Beneficial Effects of Ranolazine on Cardiac Function After Myocardial Infarction Are Greater in Diabetic Than in Nondiabetic Rats. J. Cardiovasc Pharm. 2014, 19, 457–469. [Google Scholar] [CrossRef]
  177. Chaitman, B.R.; Pepine, C.J.; Parker, J.O.; Skopal, J.; Chumakova, G.; Kuch, J.; Wang, W.; Skettino, S.L.; Wolff, A.A. Combination Assessment of Ranolazine In Stable Angina (CARISA) Investigators Effects of Ranolazine with Atenolol, Amlodipine, or Diltiazem on Exercise Tolerance and Angina Frequency in Patients with Severe Chronic Angina: A Randomized Controlled Trial. JAMA 2004, 291, 309–316. [Google Scholar] [CrossRef] [Green Version]
  178. Novak, K.R.; Norman, J.; Mitchell, J.R.; Pinter, M.J.; Rich, M.M. Sodium Channel Slow Inactivation as a Therapeutic Target for Myotonia Congenita. Ann. Neurol. 2015, 77, 320–332. [Google Scholar] [CrossRef] [Green Version]
  179. Fu, Z.; Zhao, L.; Chai, W.; Dong, Z.; Cao, W.; Liu, Z. Ranolazine Recruits Muscle Microvasculature and Enhances Insulin Action in Rats: Ranolazine, Microvasculature and Insulin Action. J. Physiol. 2013, 591, 5235–5249. [Google Scholar] [CrossRef]
  180. Eckel, R.H.; Henry, R.R.; Yue, P.; Dhalla, A.; Wong, P.; Jochelson, P.; Belardinelli, L.; Skyler, J.S. Effect of Ranolazine Monotherapy on Glycemic Control in Subjects With Type 2 Diabetes. Diabetes Care 2015, 38, 1189–1196. [Google Scholar] [CrossRef] [Green Version]
  181. Caminiti, G.; Fossati, C.; Battaglia, D.; Massaro, R.; Rosano, G.; Volterrani, M. Ranolazine Improves Insulin Resistance in Non-Diabetic Patients with Coronary Heart Disease. A Pilot Study. Int. J. Cardiol. 2016, 219, 127–129. [Google Scholar] [CrossRef] [PubMed]
  182. Wolff, A.A.; Rotmensch, H.H.; Stanley, W.C.; Ferrari, R. Metabolic Approaches to the Treatment of Ischemic Heart Disease: The Clinicians’ Perspective. Heart Fail. Rev. 2020, 7, 187–203. [Google Scholar] [CrossRef] [PubMed]
  183. Lopaschuk, G.D.; Ussher, J.R.; Folmes, C.D.L.; Jaswal, J.S.; Stanley, W.C. Myocardial Fatty Acid Metabolism in Health and Disease. Physiol. Rev. 2010, 90, 207–258. [Google Scholar] [CrossRef] [PubMed]
  184. Kantor, P.F.; Lucien, A.; Kozak, R.; Lopaschuk, G.D. The Antianginal Drug Trimetazidine Shifts Cardiac Energy Metabolism From Fatty Acid Oxidation to Glucose Oxidation by Inhibiting Mitochondrial Long-Chain 3-Ketoacyl Coenzyme A Thiolase. Circ. Res. 2000, 86, 580–588. [Google Scholar] [CrossRef] [Green Version]
  185. Fang, Y.-H.; Piao, L.; Hong, Z.; Toth, P.T.; Marsboom, G.; Bache-Wiig, P.; Rehman, J.; Archer, S.L. Therapeutic Inhibition of Fatty Acid Oxidation in Right Ventricular Hypertrophy: Exploiting Randle’s Cycle. J. Mol. Med. 2012, 90, 31–43. [Google Scholar] [CrossRef] [Green Version]
  186. Ferraro, E.; Pin, F.; Gorini, S.; Pontecorvo, L.; Ferri, A.; Mollace, V.; Costelli, P.; Rosano, G. Improvement of Skeletal Muscle Performance in Ageing by the Metabolic Modulator Trimetazidine: Metabolism Remodeling in Skeletal Muscle. J. CachexiaSarcopenia Muscle 2016, 7, 449–457. [Google Scholar] [CrossRef]
  187. Gatta, L.; Vitiello, L.; Gorini, S.; Chiandotto, S.; Costelli, P.; Giammarioli, A.M.; Malorni, W.; Rosano, G.; Ferraro, E. Modulating the Metabolism by Trimetazidine Enhances Myoblast Differentiation and Promotes Myogenesis in Cachectic Tumor-Bearing C26 Mice. Oncotarget 2017, 8, 113938–113956. [Google Scholar] [CrossRef] [Green Version]
  188. Karahan, G.; Kaya, H.; Erdogan, M.A.; Yigitturk, G.; Gokyayla, E.; Erbas, O. Effects of Trimetazidine on Nerve Regeneration in a Rat Sciatic Nerve Injury Model. Bratisl. Lek. Listy 2019, 120, 777–782. [Google Scholar] [CrossRef]
Figure 1. Pathogenic mechanisms thought to contribute to the development of ALS. Current data support several hypotheses that may explain the onset of ALS. Taken individually, these toxic events can mimic some of the hallmarks of ALS. This representation shows the complexity and multifactorial character of this disease.
Figure 1. Pathogenic mechanisms thought to contribute to the development of ALS. Current data support several hypotheses that may explain the onset of ALS. Taken individually, these toxic events can mimic some of the hallmarks of ALS. This representation shows the complexity and multifactorial character of this disease.
Cells 10 01449 g001
Figure 2. Altered energy metabolism in ALS skeletal muscle. In skeletal muscle, glucose enters into the cell through the GLUT4 transporter. It is then either stored as glycogen (not shown) or phosphorylated to give rise to pyruvate, the final product of glycolysis. Once formed, pyruvate enters into the mitochondria through a series of specific carriers (outer membrane: VDAC, inner membrane: MPC). In the mitochondria, pyruvate is oxidized into acetyl-CoA by PDH before entering the Krebs cycle to produce energy. Fatty acids enter the cell via FAT/CD36 transporters and get into the mitochondrial matrix thanks to CPT1 where β-oxidation will begin and produce acetyl-CoA which produces energy once in the Krebs cycle. In ALS, the oxidative pathway is greatly enhanced while the glycolytic pathway is reduced by abnormal induction of PDK4 in skeletal muscle which inactivates PDH by phosphorylation. PDK4 is believed to be at the root of this energy imbalance. Ac-CoA: acetyl-coenzyme A; CPT1: carnitine palmitoyltransferase 1; FAT/CD36: fatty acid translocase/cluster of differentiation 36; GLUT4: Glucose transporter type 4; MPC: mitochondria pyruvate carrier, PDH: pyruvate deshydrogenase; PDK4: pyruvate deshydrogenase kinase 4; VDAC: voltage-gated anionic channel.
Figure 2. Altered energy metabolism in ALS skeletal muscle. In skeletal muscle, glucose enters into the cell through the GLUT4 transporter. It is then either stored as glycogen (not shown) or phosphorylated to give rise to pyruvate, the final product of glycolysis. Once formed, pyruvate enters into the mitochondria through a series of specific carriers (outer membrane: VDAC, inner membrane: MPC). In the mitochondria, pyruvate is oxidized into acetyl-CoA by PDH before entering the Krebs cycle to produce energy. Fatty acids enter the cell via FAT/CD36 transporters and get into the mitochondrial matrix thanks to CPT1 where β-oxidation will begin and produce acetyl-CoA which produces energy once in the Krebs cycle. In ALS, the oxidative pathway is greatly enhanced while the glycolytic pathway is reduced by abnormal induction of PDK4 in skeletal muscle which inactivates PDH by phosphorylation. PDK4 is believed to be at the root of this energy imbalance. Ac-CoA: acetyl-coenzyme A; CPT1: carnitine palmitoyltransferase 1; FAT/CD36: fatty acid translocase/cluster of differentiation 36; GLUT4: Glucose transporter type 4; MPC: mitochondria pyruvate carrier, PDH: pyruvate deshydrogenase; PDK4: pyruvate deshydrogenase kinase 4; VDAC: voltage-gated anionic channel.
Cells 10 01449 g002
Figure 3. Metabolic reprogramming of skeletal muscle by pharmacological approaches.
Figure 3. Metabolic reprogramming of skeletal muscle by pharmacological approaches.
Cells 10 01449 g003
Table 1. Summary of the main NMJ alterations found in ALS patients and in ALS mouse models.
Table 1. Summary of the main NMJ alterations found in ALS patients and in ALS mouse models.
Change Relative to Onset of SymptomsResultsRef.
NMJ alterationsPatientsBefore
  • Decrease of motor unit number measured by MUNE in 2 asymptomatic SOD1 mutation carriers out of 19 (no statistics)
[30]
After
  • 33.8% of NMJ denervated in ALS (n = 10) vs. 9.8% in controls (n = 5) p < 0.05
[28]
SOD1G93A
mice
Before
  • P47: Denervation of 40% of NMJ in medial gastrocnemius (n = 890) (no statistics)
  • P80: decrease by 60% of intact ventral root axons (n = 4/genotype, n < 0.01 vs. control)
[26]
  • P58: Denervation of 78% of type IIb and 30% of type IIa fibers in medial gastrocnemius
(2 mice/genotype; 20–30 muscle fibers analyzed; no statistics)
[31]
  • P48: preferential denervation of type IIb NMJ followed by IIa and I in gastrocnemius (n = 3/genotype; 450 NMJ analyzed; no statistics)
[32]
  • P30: 40% denervation of type IIb NMJ in tibialis anterior (n = 11/genotype; p ≤ 0.05)
[33]
  • P30-36: Decrease by 15.5% of motor unit number measured by Baysian MUNE in gastrocnemius (n = 10/genotype; p = 0.018)
[34]
MUNE: motor unit number estimation.
Table 2. Summary of the main mitochondrial changes found in skeletal muscle of ALS patients and ALS mouse models.
Table 2. Summary of the main mitochondrial changes found in skeletal muscle of ALS patients and ALS mouse models.
Change Relative to Onset of SymptomsResultsRef.
Mitochondrial alterations in skeletal musclePatientsAfter
  • Complex I activity reduced by 40% (n = 26 ALS, n = 28 controls; p < 0.01)
  • Abnormal morphology (partially swollen, para-crystalline inclusions, vacuoles)
[51]
  • Mitochondrial aggregates in the sub-sarcolemma zone in 49 out of 49ALS cases
  • Ultrastructural abnormalities (giant mitochondria, para-crystalline inclusions) in 5 out of 49 ALS cases
[52]
  • 46% of patients with cytochrome c oxidase deficiency
  • Respiratory chain complex activity decreased by at least 30% in patients with severe COX deficiency (8 out of 50)
[54]
  • Increase of the maximal oxidative phosphorylation capacity of muscular mitochondria (Vmax) by 1.8 fold (n = 7 ALS, n = 7 controls; p < 0.05)
  • Progressive decrease of complex IV activity as disease progresses (n = 7 ALS, n = 7 controls; p < 0.05)
[55]
  • Complex I activity reduced by 47.5% (n = 14 ALS, n = 28 controls; p < 0.01)
[58]
  • Complex I and IV activity reduced by 37.1% and 43.6% respectively (n = 17 ALS, n = 21 controls; p < 0.01)
[59]
SOD1G93A mice Before
  • P37: Localized loss of inner membrane potential near the NMJ
  • Localized increase of calcium release by altered mitochondria after osmotic choc (n = 6/genotype; p < 0.0001)
[57]
  • P55: Complex I activity reduced by 20% in the tibialis anterior (n = 5 independent experiments, p < 0.0001);
  • P55: Decrease of oxygen consumption rate (OCR) by 30% (four independent experiments, with each sample tested in quadruplicate, p < 0.0001)
[60]
Table 3. Summary of the main alterations of metabolism and contractile properties of skeletal muscle in ALS patients and ALS mouse models.
Table 3. Summary of the main alterations of metabolism and contractile properties of skeletal muscle in ALS patients and ALS mouse models.
Change Relative to Onset of SymptomsResultsRef.
Alteration of skeletal muscle metabolism and contractile propertiesPatientsAfter
  • 3-fold increase of PDK4 mRNA level compared to controls (ALS: n = 11, control n = 7; p = 0.035)
[116]
  • 4-fold increase of PDK4 mRNA level and 20% decrease GAPDH mRNA level in ALS vs control cases (anconeus or deltoid muscle; ALS: n = 10, control: n = 6; p < 0.05)
[120]
  • Switch in muscle fiber type from glycolytic to oxidative in muscle biopsies of ALS patients (ALS: n = 9, 200-300 fibers/biopsy, p < 001)
[121]
SOD1G93A miceBefore
  • P55: Pdk4 mRNA level is increased 2.5-fold in tibialis anterior (p < 0.05)
  • P70: decrease of type IIb fiber mRNA level by 0.5-fold (n ≥ 4; p < 0.05) and Glut4 mRNA level by 0.4-fold (n ≥ 4; p < 0.05), and increase of CPT1 by 4.8-fold (n = 4, p < 0.01) in TA
→Bioenergetics defects
[60]
  • P70: glucose tolerance is significantly decreased (n = 6/genotype, p < 0.05)
[120]
  • P40: 22% decrease of motor unit number in the glycolytic gastrocnemius muscle compared to control mice (p < 0.05)
  • P80: 45% decrease of tetanic force in the glycolytic muscle (extensor digitorum longus) and 48% decrease of motor units number compared to control mice (p < 0.05)
→ Sequential denervation of glycolytic muscles with disease progression
[122]
  • P60: the tetanic contractile force developed by the glycolytic muscle TA is reduced by 80% (p < 0.01) and the number of motor units declines by 60% (p <0.01) compared to control mice
  • P60: the number of innervated type IIb fibers is reduced by 40% in TA (p < 0.01)
  • P60: decrease of cross sectioning area of type IIbfibers (p < 0.001)
→ Selective vulnerability of fast-twitch type IIb muscle fibers
→ Preferential denervation of fast motor neurons
[123]
  • P60: decrease of fast-twitch muscle fibers diameter by 12.5% compared to control mice (gastrocnemius muscle, number of fibers analyzed: ALS n = 63, control n = 65; p < 0.01)
[124]
After
  • Swimming straining started at P70 delays disease onset by 2 weeks (p < 0.001) and extends survival by 3 weeks (p < 0.01) (n = 8/genotype)
  • P115: the fast-to-slow myofiber transition in the fast-twitch plantaris and TA are significantly limited by the swimming program
[125]
  • P115: Glut4 and Gapdh mRNA level are reduced by around 75% in the TA and soleus muscles (p < 0.05); Swimming training increases Glut4 and Gapdh mRNA levels to control levels in TA but not in soleus; Pdk4 mRNA level is increased by 2-fold in TA (n = 5/genotype, p < 0.05)
→ Impairment of glycolytic pathway
→ Physical activity improves metabolism
[120]
  • P105: Swimming training started at P70 maintains grip strength in ALS mice (n = 8/genotype, p < 0.05 vs ALS sedentary mice
  • P105: Citrate synthase activity is reduced by 30% in ALS sedentary mice compared to control (p = 0.0007) and swimming training prevents this decrease (n = 8/condition).
  • P105: Malate dehydrogenase activity is increased ed by 25% (n = 8/condition, p < 0.0001)
→ Altered glucose metabolism
→ Swimming exercise modulates skeletal muscle energy metabolism
[126]
Sod1G86R miceBefore
  • P65: ALS mice have improved performance during endurance exercise
  • P65: glucose handling is altered. In TA, glycogen stores are increased, PFK activity is decreased by 23% (ALS mice n = 7, control: n = 6; p = 0;016), pyruvate level is 1.7 fold increased (n = 5/genotype; p = 0;019)
    Pdk4 mRNA level is 2.2-fold increased in TA compared to control mice (ALS mice n = 8, control: n = 7; p = 0;014) while unchanged in soleus.
  • Relative mRNA levels of genes involved in lipid handling pathway (Lpl; Cd36; Acsf2; Cpt1; PparB/∂) are increased in TA
→Metabolic switch: glycolytic pathway is strongly inhibited, and β-oxidation is enhanced
[116]
After
  • P95: grip strength is decreased in ALS mice (n = 8/genotype, p = 0.03) and dichloraoacetate treatment prevents grip strength loss (ALS mice: n = 8; control: n = 9, p = 0.0003)
DCA corrects the metabolic switch in TA
[116]
Acsf2: acyl-CoA synthetase family member 2; Cd36/Fat: fatty acid translocase; Cpt1: carnitine palmitoyltransferase 1; Gapdh: glyceraldehyde-3-phosphate dehydrogenase; Glut4: glucose transporter 4; Lpl: lipoprotein lipase; Pdk4: pyruvate dehydrogenase kinase 4; Pfk: phosphofructokinase; Pparβ/δ: Peroxisome proliferator-activated receptor β/δ; TA: tibialis anterior.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Quessada, C.; Bouscary, A.; René, F.; Valle, C.; Ferri, A.; Ngo, S.T.; Loeffler, J.-P. Skeletal Muscle Metabolism: Origin or Prognostic Factor for Amyotrophic Lateral Sclerosis (ALS) Development? Cells 2021, 10, 1449. https://doi.org/10.3390/cells10061449

AMA Style

Quessada C, Bouscary A, René F, Valle C, Ferri A, Ngo ST, Loeffler J-P. Skeletal Muscle Metabolism: Origin or Prognostic Factor for Amyotrophic Lateral Sclerosis (ALS) Development? Cells. 2021; 10(6):1449. https://doi.org/10.3390/cells10061449

Chicago/Turabian Style

Quessada, Cyril, Alexandra Bouscary, Frédérique René, Cristiana Valle, Alberto Ferri, Shyuan T. Ngo, and Jean-Philippe Loeffler. 2021. "Skeletal Muscle Metabolism: Origin or Prognostic Factor for Amyotrophic Lateral Sclerosis (ALS) Development?" Cells 10, no. 6: 1449. https://doi.org/10.3390/cells10061449

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