Next Article in Journal
Simplifying Genotyping of Mutants from Genome Editing with a Parallel qPCR-Based iGenotype Index
Next Article in Special Issue
Cooperation of Various Cytoskeletal Components Orchestrates Intercellular Spread of Mitochondria between B-Lymphoma Cells through Tunnelling Nanotubes
Previous Article in Journal
Dual Targeting of Glioblastoma Cells with Bispecific Killer Cell Engagers Directed to EGFR and ErbB2 (HER2) Facilitates Effective Elimination by NKG2D-CAR-Engineered NK Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mitochondria: A Promising Convergent Target for the Treatment of Amyotrophic Lateral Sclerosis

by
Teresa Cunha-Oliveira
1,*,
Liliana Montezinho
2,
Rui F. Simões
1,
Marcelo Carvalho
1,
Elisabete Ferreiro
1 and
Filomena S. G. Silva
1,3,*
1
CNC—Center for Neuroscience and Cell Biology, CIBB—Centre for Innovative Biomedicine and Biotechnology, University of Coimbra, 3004-504 Coimbra, Portugal
2
Center for Investigation Vasco da Gama (CIVG), Escola Universitária Vasco da Gama, 3020-210 Coimbra, Portugal
3
Mitotag Lda, Biocant Park, 3060-197 Cantanhede, Portugal
*
Authors to whom correspondence should be addressed.
Cells 2024, 13(3), 248; https://doi.org/10.3390/cells13030248
Submission received: 7 December 2023 / Revised: 18 January 2024 / Accepted: 24 January 2024 / Published: 29 January 2024
(This article belongs to the Collection Feature Papers in Mitochondria)

Abstract

:
Amyotrophic lateral sclerosis (ALS) is a devastating neurodegenerative disease characterized by the progressive loss of motor neurons, for which current treatment options are limited. Recent studies have shed light on the role of mitochondria in ALS pathogenesis, making them an attractive therapeutic intervention target. This review contains a very comprehensive critical description of the involvement of mitochondria and mitochondria-mediated mechanisms in ALS. The review covers several key areas related to mitochondria in ALS, including impaired mitochondrial function, mitochondrial bioenergetics, reactive oxygen species, metabolic processes and energy metabolism, mitochondrial dynamics, turnover, autophagy and mitophagy, impaired mitochondrial transport, and apoptosis. This review also highlights preclinical and clinical studies that have investigated various mitochondria-targeted therapies for ALS treatment. These include strategies to improve mitochondrial function, such as the use of dichloroacetate, ketogenic and high-fat diets, acetyl-carnitine, and mitochondria-targeted antioxidants. Additionally, antiapoptotic agents, like the mPTP-targeting agents minocycline and rasagiline, are discussed. The paper aims to contribute to the identification of effective mitochondria-targeted therapies for ALS treatment by synthesizing the current understanding of the role of mitochondria in ALS pathogenesis and reviewing potential convergent therapeutic interventions. The complex interplay between mitochondria and the pathogenic mechanisms of ALS holds promise for the development of novel treatment strategies to combat this devastating disease.

Graphical Abstract

1. Introduction

Amyotrophic lateral sclerosis (ALS) disease, also known as Lou Gehrig’s disease or Charcot disease, is a progressive neurodegenerative disorder that often manifests quietly, initially concealing its full spectrum of symptoms. The disease advances swiftly, not affecting cognition and self-awareness, yielding a definitive diagnosis that compels patients and their loved ones to confront a daunting reality—a verdict that entails grappling with an immensely debilitating condition, necessitating increasing levels of both human and mechanical support. Regrettably, ALS lacks a cure and frequently culminates in rapid fatality.
Although advances in the understanding of ALS have been made, an effective treatment remains to be discovered. The incidence of ALS ranges from 0.6 to 3.8 per 100,000 person-years, with a prevalence of 4.1 to 8.4 per 100,000 individuals. This predominantly affects individuals between the ages of 51 and 66 [1]. The survival period from the onset of symptoms to demise spans from 24 to 50 months, although a minority—10% of ALS patients—experience a slower variant with survival exceeding a decade [1]. Roughly 90–95% of patients develop sporadic ALS (sALS), characterized by an enigmatic etiology and the absence of familial precedent [2]. The remaining 5–10% manifest familial ALS (fALS) [2], an inherited form, frequently linked to an earlier age of onset [3,4].
Presently, approximately 50 genes have been identified as having a role in disease modification [5], albeit only a fraction of these genes demonstrate a significative prevalence in ALS, particularly within the context of fALS [6]. Among this collection of genes, a subset has shown involvement in both sALS and fALS. Notably, the genes C9ORF72 (encoding chromosome 9 open reading frame 72 (C9orf72) protein) [7,8], SOD1 (encoding Cu/Zn superoxide dismutase 1 [SOD1] protein) [9,10,11,12], TARDP (encoding for TAR DNA-binding protein 43 [TDP-43]) [13], and FUS (encoding fused in sarcoma/translocated in sarcoma [FUS/TLS] protein) have been implicated [14,15]. Moreover, it has been proposed that ALS pathogenesis arises from the collaboration of multiple genes, each acting independently, thereby underscoring a specific genetic susceptibility among ALS patients [5]. External factors also seem to play a pertinent role in triggering neurodegenerative processes linked to ALS, particularly in individuals with a susceptible genetic predisposition [16,17,18]. Indeed, various risk factors have been identified as potential catalysts for ALS onset, encompassing age, smoking, body mass index (BMI), physical fitness level, and exposure to environmental hazards like chemicals, pesticides, metals, and electromagnetic fields, among others [19]. Nevertheless, uncertainties persist regarding the precise contribution of these factors, as the root cause of the disease has not been identified.
To date, a cure for ALS remains elusive and existing therapies primarily address symptom management, yielding minimal impact on survival time. Currently, seven drugs are approved by the US Food and Drugs Administration (FDA): Qalsody (tofersen, # N215887)), RELYVRIO (AMX0035, #N216660), Radicava™ (edaravone; #N209176), Rilutek (riluzole, now generic; #N020599), Tiglutik™ (thickened riluzole; #N209080), Exservan (riluzole oral film; #N212640), and Nuedexta® (#N021879). The recent approval of tofersen, AMX0035, and edaravone have benefited from the application of the American “Accelerating Access to Critical Therapies for ALS Act” (Public Law 117–79). In Europe, to date, only riluzole has been approved by the European Medicines Agency (EMA) (under the reference EMEA/H/C/000109).
ALS is marked by the progressive degeneration of both upper motor neurons (UMN) in the cerebral cortex and lower motor neurons (LMN) in the brain stem and spinal cord. This degeneration precipitates muscle weakness, which gradually evolves into muscle atrophy and paralysis, ultimately culminating in respiratory failure and death [4,20]. Increasing evidence suggests that fALS and sALS may share common pathological mechanisms due to their analogous clinical presentations [2,21]. A multitude of molecular pathways potentially underlie disease development, encompassing disrupted RNA metabolism, glutamate excitotoxicity, protein misfolding and aggregation, endoplasmic reticulum stress, impaired protein trafficking, compromised axonal transport, oxidative stress, inflammation, and mitochondrial dysfunction [2,3,22,23]. Much like numerous other neurodegenerative diseases, as mentioned earlier, mitochondrial dysfunctions assume a significant role in ALS progression. Indeed, various mitochondrial abnormalities have been described not only within the central nervous system (CNS) of ALS patients but also in peripheral tissues, such as skeletal muscle [24,25], liver [26], and lymphocytes [23]. These perturbations encompass mitochondrial bioenergetics dysfunction, inefficient calcium buffering, initiation of mitochondrial-related apoptosis, and aberrations in mitochondrial transport and morphology, among other aspects [27,28]. However, despite the compelling evidence of multifaceted mitochondrial dysfunction in ALS, the precise extent of its involvement, and whether it acts as a triggering factor or is a consequence of other intracellular toxic mechanisms, remains enigmatic.
This review seeks to synthesize recent insights into the role of mitochondria and mitochondria-mediated mechanisms in driving cellular damage. Additionally, it aims to offer a deeper understanding of the interplay between different cell types implicated in ALS. Ultimately, only through a more thorough comprehension of these intricate processes can effective mitochondria-targeted therapies be identified for potential pharmacological interventions in ALS treatment.

2. Mitochondrial Dysfunctions in ALS

Mitochondria are highly dynamic organelles, vital for normal metabolism and cellular survival, that participate in a diverse array of cellular functions, including energy metabolism, calcium homeostasis, redox regulation, and cell-death mechanisms [2]. Much like in other neurodegenerative diseases, disruptions to mitochondrial structure, dynamics, bioenergetics, and calcium buffering have been extensively documented in ALS patients, as well as in in vitro and in vivo models of ALS [2,29]. Varied morphological changes have been observed in the early stages of ALS disease within motor neurons (MNs) from transgenic mouse models with SOD1 [30,31,32], FUS [33], and TDP-43 [34,35] mutations; cortical neurons from the C9orf72 mouse model [36]; and muscle samples from ALS patients [24]. These findings underscore the involvement of mitochondria right from the onset of ALS, manifesting not only in the CNS but also in muscle tissue. The indispensable role of mitochondria in driving ALS progression is further underscored by the identification of specific mitochondrial components interacting with ALS-associated proteins. Notable examples include B-cell lymphoma 2 (Bcl-2) [37], voltage-dependent anion channel 1 (VDAC1) [38], and the mitochondrial form of lysyl-tRNA synthetase [38,39], which interact with mutant SOD1; mRNAs encoding mitochondrial respiratory chain components, which interact with TDP-43 [40] and FUS [41] proteins; and VDAC3, a translocase of the inner mitochondrial membrane 50 (TIMM50) [42] and mitochondrial ribosomal proteins [43], which interact with C9orf72. However, the precise extent of the mitochondrial role in driving disease progression remains an enigma.

2.1. Alterations in Mitochondrial Respiration and ATP Production

The most well-known function of mitochondria is their generation of approximately 90% of cellular energy in the form of adenosine triphosphate (ATP), an essential currency for virtually every facet of cellular activity. This energy synthesis stems from the intricate process of oxidative phosphorylation (OXPHOS) taking place in the electron transport chain (ETC), situated within the inner mitochondrial membrane (IMM) [44] (Figure 1).
Discrepant findings have emerged concerning mitochondrial respiration and ATP production across distinct ALS models. Examination of postmortem spinal cord tissues from sALS patients unveiled mitochondrial dysfunction characterized by diminished activity of complexes I, II, III, and IV in tandem with decreased mitochondrial DNA (mtDNA) content and citrate synthase activity [45]. Conversely, no discernable alteration in ETC activity surfaced within crude mitochondrial preparations obtained from the motor and parietal cortices and cerebellum of sALS patients [46]. This dichotomy might signify a more pronounced ETC functional impairment in the spinal cord compared to brain tissues. An alternate interpretation of these contradictory outcomes could be rooted in the limited number of samples analyzed in both studies as well as the inherent heterogeneity found among sALS patients.
Remarkably, beyond the CNS, altered ETC activities have also been documented in sALS patients. Instances include decreased activities of complexes I and IV in skeletal muscle [24,47] and decreased complex I activity coupled with lower ATP levels in lymphocytes [48]. Furthermore, when looking to patient fibroblasts, a noteworthy elevation in mitochondrial membrane potential (ΔΨm) has been observed [49,50]. This potentially alludes to a compensatory mechanism aimed at ameliorating inefficient ATP synthesis. However, comprehensive investigations are imperative to validate these alterations.
Disruption in mitochondrial respiration has been extensively documented in mouse models carrying mutated SOD (mutSOD) [46,51,52,53,54,55] as well as in ALS patients harboring mutSOD1 [46,56]. SOD1, a Cu-Zn metalloprotein, plays a pivotal role in converting superoxide anion radicals (O2) into molecular oxygen and hydrogen peroxide (H2O2); it is primarily localized within the cytosol, but also found in the nucleus, peroxisomes, and mitochondria [11]. Although there is evidence of SOD1 accumulation within mitochondria [51,57,58,59,60,61], the precise function of this enzyme within mitochondria remains unclear.
Most studies suggest that both wild-type and mutSOD1 expressed in transgenic mice, particularly SOD1G93A, predominantly accumulate in the intermembrane space (IMS) of spinal cord mitochondria [51,57,58,59,61]. However, some reports have also indicated localization on the outer mitochondrial membrane (OMM) [37,62], association with components of the IMM [51], or accumulation in the mitochondrial matrix [63]. Yet, the accumulation of mutSOD1 within the mitochondrial matrix has not been definitively established.
Regarding the impact of mutSOD1 on mitochondrial respiration, studies have revealed the decreased activities of complexes I+III, II+III, and IV in the spinal cord and its ventral horn at the onset of the disease in SOD1G93A mice [51]. Furthermore, a decrease in the activities of complexes I, II, and IV in the spinal cord preceding disease onset was found in the same mouse model [52]. Consistently, decreased mitochondrial respiration rates have been observed in the spinal cord and brain at the pre-symptomatic stage, persisting throughout the disease in SOD1G93A mice. Notably, a decline in complex IV activity in the brain of these mice was associated with cytochrome c-IMM dissociation [53], suggesting an increase in reactive oxygen species (ROS) production leading to enhanced cardiolipin oxidation and subsequent cytochrome c dissociation from the IMM. Decreased complex II and IV activities were also documented in SOD1G93A or SOD1G37R in NSC-34 motor-neuron-like cell lines [54], and significant impairment of the complex I-linked OXPHOS, with a lower ΔΨm, was also identified in SOD1G93A NSC-34 cells [55].
Contrasting findings were reported in patients carrying mutSOD1A4V, who exhibited increased complex I activity in the frontal cortex [56] and in the motor and parietal cortices [46]. Moreover, elevated activities of complexes II–III were noted in the motor cortex and cerebellum of patients with mutSOD1A4V as well as in the motor and parietal cortices and cerebellum of a single patient with mutSOD1Ill3T, with similar findings in the motor cortex and cerebellum of ALS patients without identified SOD1 mutations to date [46]. In this specific context, the heightened activities of the ETC could signify a compensatory response to oxidative damage to the IMM, possibly resulting from OXPHOS uncoupling. In the forebrain of SOD1G93A mice, an elevation in complex I activity was observed during the pre-symptomatic stage when compared to levels in transgenic wild-type counterparts [46]. Discrepancies in these findings may arise from several factors, including differences in experimental conditions, variations among distinct SOD1 mutations, diverse SOD1G93A mouse strains, discrepancies between animal and human physiology, and the disease stage itself. These disparities highlight the intricate nature of alterations in ETC dynamics, which can emerge as either a causal factor or a consequence of oxidative stress mechanisms. In some instances, these alterations may be counterbalanced by the activation of compensatory mechanisms. It is crucial to consider that the SOD1 transgenic animal model features a high copy number of human mutSOD1, rendering it a non-physiological model that may lack many of the phenotypic alterations observed in ALS patients [64,65]. The prevalence of studies documenting alterations in mitochondrial respiration in SOD1 transgenic animal models in contrast to the absence of impairment in mitochondrial respiration observed in studies involving SOD1A4V or SOD1G93A human MNs [66] suggests that mitochondrial dysfunction observed in the SOD1 transgenic animal model may be overestimated and should be carefully interpreted.
Another gene mutation associated with ALS is TDP-43, which has demonstrated various cellular respiratory system alterations in both patients and cellular models [67,68]. TDP-43, a DNA- and RNA-binding protein, plays a critical role in RNA processing [69]. Studies have shown that TDP-43 can lead to decreased activity of complex I and to a reduction in ΔΨm. This was associated with increased expression of mitochondrial uncoupling protein 2 (UCP2) in NSC-34 cell lines transfected with TDP-43 [67]. In neurons from ALS patients, mutant TDP-43 was found to accumulate within mitochondria [40] and to preferentially bind to mitochondrial mRNAs encoding the ND3 and ND6 subunits of complex I. This binding resulted in the disassembly of complex I and decreased ΔΨm in HEK293 cells overexpressing mutant TDP-43 variants (G298S, A315T, or A382T) or in fibroblasts from ALS patients carrying G298S or A382T mutations [40]. These findings suggest that TDP-43 toxicity directly impacts mitochondrial bioenergetics. However, fibroblasts from ALS/FTD patients carrying TDP-43A382T did not present alterations in mitochondrial respiration when compared to controls, although a decrease in ΔΨm was observed in these cell lines [70]. FUS, another multifunctional DNA/RNA-binding protein associated with ALS [71], has also been associated with alterations in mitochondrial respiration [41]. Similar to TDP-43, both wild-type FUS and ALS mutant derivatives were found to bind to mRNAs encoding subunits of mitochondrial respiratory chain complexes (RCC) [41]. This binding disrupted mitochondrial networks and led to impaired mitochondrial respiration, both in HEK293T cells overexpressing wild-type FUS and in those expressing mutant FUSR521C. Similar effects were observed in fibroblasts derived from ALS patients [41].
As for C9orf72, the most prevalent mutation in ALS, characterized by the expansion of the GGGGCC (G4C2) hexanucleotide repeat in the first intron of the C9ORF72 gene [72], observations have pointed to reduced activities of mitochondrial complexes I and V within the frontal cortex of 6-month-old C9orf72-ALS/FTD mice [36]. Similarly, a significant reduction in ΔΨm has been noted in human induced pluripotent stem cells (iPSC)-derived MNs reprogrammed from the fibroblasts of C9orf72 patients [73]. Conversely, C9orf72 fibroblasts from ALS/FTD patients exhibited increased ΔΨm along with elevated oxygen-consumption rates, heightened ATP levels, greater complex II activity, and enhanced production of ROS compared to controls [70].
Collectively, these findings underscore the presence of mitochondrial respiration impairment in both the CNS and muscles of ALS patients and animal models. Despite variations in results among different studies, it is apparent that these alterations collectively represent processes of mitochondrial dysfunction that contribute to the progression of ALS.

2.2. Role of Oxidative Stress Mechanisms

Mitochondria serve both as a source and a target of ROS [74]. Their most studied function is the production of cellular energy in the form of ATP, a process involving the controlled transport of electrons from substrates to molecular oxygen. As previously mentioned, this intricate electron transfer occurs through a series of enzymatic complexes composing the ETC (Figure 1). When electrons inadvertently leak from mitochondrial ETC complexes directly to molecular oxygen, they generate O2 [74], which carries out pivotal roles in cellular signaling [75].
Under normal physiological conditions, cells meticulously maintain a defined redox balance by counteracting ROS with the assistance of the cellular antioxidant defense system [76]. Oxidative stress occurs when this balance is disrupted, either due to excessive ROS production, a compromise in antioxidant defenses, or both [77,78]. In times of cellular stress, elevated ROS levels perturb the redox homeostasis, causing damage to essential biomolecules including proteins, DNA, and lipids. This disruption severely impairs organelle function and eventually leads to cell death [79].
A wealth of evidence indicates that both heightened ROS levels and inadequate antioxidant defenses play substantial roles in ALS [41]. Indeed, post-mortem brain tissue [80,81,82], cerebrospinal fluid (CSF) [83,84,85,86], plasma [87], and urine [88] from ALS patients exhibited increased oxidative damage to proteins, lipids, and DNA. In spinal cord tissues from sALS patients, oxidative damage manifested as elevated levels of protein carbonyls [56,81,82,89], 8-hydroxy-2′-deoxyguanosine (8-OHdG) [81,90], 4-hydroxynonenal (4-HNE) protein conjugates [90,91], nitrotyrosine products [92,93,94], and malondialdehyde-modified proteins [81]. Furthermore, sALS patients’ erythrocytes revealed increased lipid peroxidation concurrent with a decrease in glucose-6-phosphate dehydrogenase activity [95] and in antioxidant defenses, including catalase, glutathione reductase, and glutathione (GSH). These effects appear to intensify with the progression of the disease [95].
Given that SOD1 is a Cu-Zn metalloprotein responsible for converting O2.− into molecular oxygen and H2O2, [11] it is highly plausible that ALS-associated mutations in SOD1 may disrupt the delicate balance of ROS and lead to oxidative damage [96]. Indeed, lymphoblasts from ALS patients with or without identified SOD1 mutations and healthy controls showed distinct redox signatures [97]. Although SOD1 is primarily considered a cytosolic enzyme, it has also been found within the IMS, where it may play a crucial role in safeguarding mitochondria against O2.−, thereby complementing the function of SOD2, which eliminates O2.− released into the mitochondrial matrix [9]. However, the precise mechanism by which SOD1 enters mitochondria has not yet been fully elucidated. In yeast, only an immature form of SOD1, lacking both Cu and Zn and existing in a reduced disulfide state, was found to be able to cross the OMM [98]. It is presumed that the SOD1 apoenzyme gains access to the IMS through the TOM complex [99].
Like other IMS proteins, the import and retention of SOD1 within the IMS are tightly linked to its folding and maturation processes, which include the formation of disulfide bonds. The cysteine residues within SOD1 play a pivotal role in localizing it to the IMS [98,99]. Moreover, it has been proposed that the IMS import of SOD1 involves a copper chaperone (CCS), while its mitochondrial distribution might be regulated by the Mia40/Erv1 disulfide relay system [99,100]. Despite evidence of SOD1 accumulation in mitochondria, the exact mechanisms through which mutSOD1 induces oxidative stress and mitochondrial damage remain to be fully understood, and it is not clear whether these effects are a cause or a consequence of the disease. Considering that only an immature form of SOD1 can traverse the OMM [98], it is possible that mutSOD1 increases the fraction of intramitochondrial SOD1 due to inefficient cytosolic maturation, potentially explaining the heightened mutSOD1 accumulation under pathological conditions. Additionally, the increased mitochondrial accumulation of both wild-type and mutSOD1 in vitro following experimental copper depletion suggests that heightened mitochondrial mutSOD1 accumulation might also be involved in certain pathological mechanisms associated with this mutation [101]. Despite the uncertainties surrounding the deleterious effects of this mutation, it has been proposed that mutSOD1 exerts its pathological impact through a toxic gain of function mechanism rather than by altering SOD1′s enzymatic activity [102].
Various hypotheses have been put forth to elucidate the toxic effects of mutSOD1. One such hypothesis suggests that mutSOD1 induces oxidative damage through an abnormal association with zinc [103] or through exposure to toxic copper at the active site, leading to an elevation in O2.− levels [104]. Additionally, it has been proposed that mutSOD1 can exert its toxicity through the following mechanisms: (1) peroxidase activity–mutSOD1 may function as a peroxidase, demonstrating reverse activity, by utilizing H2O2 as a substrate [58,105]; (2) reactivity with peroxynitrite–mutSOD1 might react with peroxynitrite, resulting in tyrosine nitration [77,106,107]; and (3) formation of soluble aggregates–misfolded SOD1 can form soluble aggregates, leading to decreased stability of SOD1 monomer/dimers [108]. Results observed in MNs derived from iPSCs showed that SOD1 aggregates are progressively and stably formed and transported to neighboring cells [109]. Indeed, there is evidence of increased susceptibility to oxidative stress and mitochondrial damage attributed to misfolded SOD1. This has been observed in mitochondria isolated from the spinal cords of SOD1G93A rats and SOD1G37R mice [110]. Furthermore, in the NSC-34 mouse motoneuronal cell line, mutSOD1 was found to accumulate in an oxidized and aggregated state. This accumulation appears to be associated with mutSOD1 interaction with mitochondria and the oxidation of cysteine residues. Thus, mutSOD1 may consequently contribute to respiratory chain impairment and an imbalance in the ratio of reduced glutathione (GSH) to oxidized glutathione (GSSG), shifting towards a more oxidizing environment [111]. Supporting this, decreased GSH levels and elevated GSSG levels were observed in the lumbar tissues of the spinal cords of mutant SODG93A mice. Additionally, genetic induction of diminished antioxidant defenses led to notable MN degeneration in hSOD1WT mice, a hemizygous mice model with moderate over-expression of wild-type hSOD1 that does not typically exhibit paralysis or a shortened lifespan [112]. A similar approach was applied to investigate the effects of decreased antioxidant defenses in SOD1G93A mice, which overexpress ALS-linked mutSOD1 [113]. In this case, a 70% decrease in GSH levels further exacerbated the detrimental impact of mutSOD1 on life span. This was associated with increased oxidative stress, aggravated mitochondrial dysfunction, and an elevated association of mutSOD1 with mitochondria [113]. Further evidence of mutSOD1′s negative effects on redox balance has been observed in SH-SY5Y human neuroblastoma cells overexpressing mutSOD1 [114]. Here, it was noted that mutSOD1 induced the activation of Src homology 2 domain containing (Shc) transforming protein 1 (p66Shc). p66Shc is an alternatively-spliced isoform of a growth factor adapter that is phosphorylated upon oxidative stress, resulting in its translocation to mitochondria, where it binds to cytochrome c and, thereby, functions as an oxidoreductase and amplifies ROS production [114].
Apart from mitochondria, another significant source of ROS within the cell stems from the activity of adenine dinucleotide phosphate (NADPH) oxidases (NOX). These NOX enzymes form a family of enzyme complexes that facilitate the transfer of electrons from NADPH to molecular oxygen, ultimately generating O2.− and H2O2 [115]. A pivotal feed-forward loop of ROS production has been identified, driven by the crosstalk between mitochondria and NOX [115]. Essentially, ROS produced by NOX activity can assail mitochondria, while ROS generated within mitochondria can stimulate NOX activity, culminating in amplified ROS production. Members of the NOX family have emerged as potential sources of ROS in ALS. Specifically, the deletion of Nox2 and, to a lesser extent, Nox1 has been shown to decelerate disease progression and enhance the survival of SOD1G93A mice, underscoring the role of NOX activity in the oxidative imbalance observed in ALS [116,117]. This is substantiated by evidence revealing elevated NOX activity, particularly NADPH-dependent O2.− production, in SOD1G93A mice. This increase appears to result from the physical interaction of mutSOD1 with Rho-like small GTPases, specifically Rac1, which serves as a regulator of NOX activity [118]. Under non-physiological conditions, such as those encountered in an oxidative environment, the dissociation of mutSOD1 from Rac-GTP is impeded, causing the enzyme to remain in an active state and, consequently, intensifying Nox2-derived O2.− production [118]. NADPH oxidase inhibition, particularly with apocynin and its derivative diapocynin, was reported to significantly extend the survival of SOD1G93A ALS mice [118]. However, this was challenged by a follow-up study conducted at two different institutions [119], which failed to replicate the remarkable survival extension previously reported [118]. These results cast some doubt on the role of NADPH oxidase in the SOD1 mouse model of ALS. Moreover, the lack of consistent and significant protective effects in various trials raises important questions about the translational potential of NADPH oxidase inhibition as a therapeutic strategy for ALS.
The significance of disrupted redox homeostasis for ALS was further underscored by a study involving MNs from both sALS and fALS patients as well as rodent mutSOD1 ALS models (H46R/G93A rats and G1H/G1L-G93A mice) [120]. As the disease progressed, and concurrent with the decline in spinal MNs, a diminishing number of neurons was found to retain peroxiredoxin-II (PrxII) and glutathione peroxidase-l (GPxl), which are key regulators of the redox system [120]. However, during disease progression, a small population of MNs was also observed to exhibit overexpression of PrxII/GPxl, indicating an up-regulation of the redox system in these cells. This suggests that some neurons retain the ability to respond to redox stress, at least until the advanced stages of the disease [120].
Another vitally important antioxidant pathway that responds to oxidative stress is the nuclear factor erythroid-2-related factor 2 (Nrf2) pathway. In balanced conditions, Nrf2 is sequestered in the cytosol by Keap1 (Kelch ECH-associating protein 1), a cytoplasmic regulator that acts as a sensor for ROS. When ROS levels increase, cysteine residues on Keap1 undergo oxidative modification, releasing Nrf2. This liberated Nrf2 then translocates to the nucleus, where it upregulates the expression of antioxidant genes. These include GPX, catalase, and enzymes involved in glutathione metabolism, such as glutathione S-transferase, glutathione cysteine ligase modifier subunit, and glutathione cysteine ligase catalytic subunit (GCLC) [121,122,123,124,125]. Notably, SOD1 is also among the genes upregulated by Nrf2, making this pathway particularly intriguing for ALS research, especially in cases associated with SOD1 gene mutations.
Despite the precise role of Nrf2 in ALS pathogenesis remaining unclear, accumulating evidence suggests that Nrf2 plays a pivotal role, and its activity may serve as a novel therapeutic target [22]. In both NSC34 cells expressing mutSOD1 and MNs isolated from familial SOD1-associated ALS patients, Nrf2 levels have been shown to be lower than those found in control conditions [126,127]. This implies a disturbance in the Nrf2 pathway and a defective response to oxidative stress. It is important to note that Nrf2 levels may fluctuate throughout the course of the disease. Interestingly, the Nrf2 pathway is naturally activated in muscle tissue during the onset of pathology in the SODG93A mouse model [128]. Dysregulation of the Nrf2 pathway has also been observed in sALS patients who do not express mutSOD1, suggesting that Nrf2 may play a role in ALS independently of mutSOD1 [128]. Further research is needed to fully comprehend the mechanistic underpinnings of Nrf2 pathway activation and impairment in ALS.
In the context of the redox system, catalase, an enzyme responsible for converting H2O2 into water and oxygen, which is normally absent from mitochondria, was found to enhance mitochondrial antioxidant defenses and mitochondrial function in primary cultures of astrocytes carrying SOD1G93A when genetically modified to target mitochondria [129]. However, overexpression of catalase in SOD1G93A mice had no discernible effect on their lifespan, indicating that intervention in this pathway alone is insufficient to halt disease progression [129].
Oxidative damage in ALS has also been attributed to other mutated ALS-linked genes. For instance, mutations in TDP-43, a protein belonging to the heterogeneous nuclear ribonucleoproteins (hnRNP) family, have been found to disrupt the Nrf2 pathway [130,131,132]. This disruption likely occurs through the interaction of TDP-43 with another hnRNP family member, hnRNP K, which influences its expression and the RNA-binding properties of the protein [131], subsequently affecting its antioxidant gene transcripts’ binding and function. Consequently, while Nrf2 signaling remains activated in response to stress, its function is compromised due to the abnormal binding of hnRNP K to antioxidant gene transcripts, leading to uncontrolled oxidative damage and toxicity [131]. Multiple lines of evidence support the association between TDP-43 and the dysregulated Nrf2 pathway.
In fibroblasts from patients with the TDP-43M337V mutation and in astrocyte cultures from TDP-43Q331K mice, GSH levels (downstream of Nrf2 activation) were found to be decreased, indicating an impaired oxidative stress response. Moreover, NSC-34 cells overexpressing TDP-43M337V showed a decline in antioxidant defenses, leading to increased intracellular lipid peroxidation, lower cell viability, nuclear accumulation of Nrf2, and decreased protein expression of NAD(P)H quinone dehydrogenase 1 (NQO1, downstream of Nrf2 signaling) [130]. Similar results were obtained in another study using cells expressing TDP43 with two mutations (M337V/Q331K) [132].
Regarding C9orf72, few studies directly implicate oxidative damage as a causative factor in the disease. Nevertheless, evidence suggests a connection between oxidative stress and ALS in the context of C9orf72 mutations. Oxidative stress and DNA damage were found to be increased in iPSC-derived C9orf72 MNs [43]. Additionally, another study demonstrated that dysfunction in iPSC-derived astrocytes from ALS patients with mutated C9orf72 was also associated with elevated oxidative stress. Interestingly, the medium from these cell cultures induced oxidative stress in wild-type MNs [133].
While substantial evidence strongly underscores the contribution of excessive ROS in ALS, which are primarily generated within mitochondria, as well as compromised antioxidant defenses [22,134], the precise role of oxidative damage in the development and progression of the disease remains a subject of inquiry. Specifically, it remains to be elucidated whether oxidative damage serves as a primary instigator of the disease or arises as a secondary consequence stemming from the initial toxic mechanisms that trigger the onset of ALS.

2.3. Metabolic Dysregulation

The investigation of energy homeostasis in ALS research has revealed significant dysregulation in both ALS patients [135] and in models of the disease involving SOD1 and TDP-43 [136,137,138,139,140]. ALS patients exhibited an imbalance between food intake and energy expenditure [141], and research has shown increased energy demands in the muscles of SOD1 (G86R and G93A) mice compared to controls [136]. Interestingly, by using mice expressing human SOD1 exclusively in skeletal muscle, Martin and Wong showed that the presence of hSOD1 only in skeletal muscle resulted not only in limb weakness and muscle wasting but also in abnormalities in adipose tissue, suggesting a relationship between muscle and adipose tissue in the context of ALS disease [142].
Malnutrition is a recognized problem among ALS patients, with its prevalence ranging from 15% to 55%, and it is negatively correlated with disease severity [143,144,145,146]. There is substantial evidence suggesting that malnutrition serves as a prognostic factor for poorer survival rates among ALS patients [144,145,146,147,148]. The progression of malnutrition typically leads to weight loss [147,149], attributable to dysphagia resulting from bulbar weakness or heightened metabolic activity [148,150,151,152].
Interestingly, studies have indicated that individuals with a lower BMI are at an increased risk of developing ALS [153], while, conversely, overweight individuals face a lower risk of up to 40% [154]. Consequently, weight loss emerges as a crucial prognostic factor for patients [155] and can serve as a valuable clinical indicator of the necessity for nutritional support.
Several studies have reported instances of hypermetabolism, characterized by an increase in measured metabolic rate compared to calculated values, in a significant proportion of ALS patients [144,148,149,150,151,156,157,158,159]. Many of these studies have applied equations for predicting resting energy expenditure [160], although these equations still require validation for ALS [161] and often do not account for the muscle atrophy frequently observed in ALS patients [159]. The prevalence of hypermetabolism in sALS patients has been documented to range from 25% to 68% [144,148,149,150,151,158,159], while it was reportedly found in 100% of fALS individuals [156]. Notably, approximately 80% of ALS patients in one study were found to maintain a consistent metabolic phenotype over time, and this phenotype showed no correlation with disease characteristics [148]. However, it was observed that hypermetabolism negatively correlated with age but positively correlated with BMI and fat-free mass [148]. In contrast, some studies have reported an increase in hypermetabolism as patients neared the end of life [149], although this finding was not replicated in other research [148,150,151]. Nevertheless, a trend towards worse survival among hypermetabolic ALS patients was subsequently noted [158]. Further insights come from reports of increased prevalence of hypermetabolism in ALS patients compared to age- and sex-matched control individuals [159]. In this study, hypermetabolism was associated with a significant reduction in fat-free mass in ALS patients [159]. Notably, this study was the first to consider the muscle atrophy observed in these patients, reporting a prevalence of 41% hypermetabolism in ALS patients compared to 12% in healthy, matched subjects [159]. Importantly, hypermetabolic ALS individuals in this study experienced a significantly faster rate of functional decline and worse survival than normometabolic ALS patients [159], underscoring the potential utility of metabolic indices in predicting ALS outcomes. However, a systematic review [162] analyzing 29 selected articles from various databases and aiming to correlate changes in metabolic parameters with ALS progression and survival concluded that the use of different metabolic parameters and methodologies across studies made it challenging to establish correlations [162]. The parameters examined varied widely, including BMI (65.5%), hypertension (20.7%), cardiovascular risk (6.89%), obesity (10.3%), diabetes (13.8%), and glycaemia (10.3%) [162]. Despite the assumption that metabolism influences ALS, the study could not definitively establish an association between metabolism and disease development [162]. The authors emphasized the need for standardized methodologies and the analysis of metabolic parameters across different studies to yield more conclusive results [162].
The dysregulation of metabolic pathways associated with stress and oxidative responses may contribute significantly to the progression of ALS [163]. Despite extensive research, the root causes of metabolic dysregulation and its effects on neuronal metabolism in ALS patients remain incompletely understood. Nevertheless, mitochondria have emerged as central players in this process [148]. Numerous studies employing proteomic approaches have reported substantial alterations in proteins linked to glycolysis, mitochondrial function, and stress responses, particularly in SOD1G93A mice when compared to control animals [163,164,165,166,167,168]. A recent study employed targeted proteomics to quantify proteins involved in metabolic pathways at the pre-onset, onset, and end stages of the disease in the spinal cord of SOD1G93A mice [163]. Their findings revealed significant changes in metabolic protein profiles, even at early stages preceding disease onset. Notably, an increased abundance of phosphofructokinase and lactate dehydrogenase—crucial enzymes regulating glycolysis and tricarboxylic acid (TCA) cycle rates, respectively—was observed [163]. These results align with a prior study that reported heightened glycolytic flux in a neuronal cell line expressing SOD1G93A [169]. Moreover, increased glucose uptake was detected in the spinal cords of transgenic SOD1G93A mice during the pre-symptomatic stage, which progressively declined as the disease advanced [170]. Importantly, an elevated rate of glycolysis can impact other metabolic pathways, including the pentose phosphate pathway, potentially decreasing cellular antioxidant capacity and triggering cellular apoptosis [171]. At the later stages of the disease, reductions in the levels of various glycolytic proteins (aldolase, glyceraldehyde phosphatase, and phosphoglycerate kinase), which are important proteins involved in the malate–aspartate shuttle (cytoplasmic and mitochondrial aspartate aminotransferases, and mitochondrial malate dehydrogenase), pyruvate dehydrogenase, and TCA cycle enzymes (malate and isocitrate dehydrogenase-related proteins) were reported [163]. Conversely, certain proteins related to fatty acid metabolism and the ETC exhibited increased abundance in the spinal cords of SOD1G93A mice [163]. These alterations in protein content are consistent with a shift from glycolysis to oxidative metabolism, which was also been reported in the muscles of SOD1 mice [172,173], which have higher energetic needs [136]. This shift in metabolic strategy is probably driven by the heightened energy demands of large MNs, which are particularly susceptible to changes in metabolism [163]. Interestingly, the glycolytic ATP production rate showed potential diagnostic and therapeutic interest in lymphoblasts from ALS patients [174].
When glucose demand outstrips the available supply, neurons begin to utilize ketone bodies produced by the liver and, notably, by astrocytes [175,176,177,178,179]. The production of ketone bodies by astrocytes, which enhances oxidative metabolism, is reflected in the upregulation of proteins related to β-oxidation, fatty acid metabolism [137,163], and the levels of 3-hydroxy-3-methylglutaryl-CoA lyase, a key enzyme in the ketogenic process [179]. Dysregulated fatty acid metabolism has been reported in both ALS animal models and ALS patients, with evidence suggesting that dyslipidemias play a pivotal role in disease progression [141].
Positive protein inclusions of TDP-43 are an almost ubiquitous feature in ALS cases, establishing them as a fundamental pathological hallmark of the disease [180,181]. Through metabolic profiling, significant changes in glycolysis and pentose phosphate pathways were found in Drosophila models of TDP-43 proteinopathy [140]. The TDP-43 mutants exhibited increased levels of pyruvate, suggesting an upsurge in glycolytic activity and a substantial upregulation of phosphofructokinase, the enzyme governing the rate-limiting step in glycolysis [140]. These findings align with previous research showing elevated pyruvate levels in the plasma of ALS patients [182,183]. The Drosophila TDP-43 mutants also presented an intensified glycolytic input into the pentose phosphate pathway and heightened levels of glucose-6-phosphate dehydrogenase, potentially providing a mechanism to counteract oxidative stress by enhancing NADPH production [140].
Another study adopted a metabolic profiling approach using astrocytes reprogrammed from fibroblasts obtained from C9orf72, sALS, and age-matched control subjects to explore C9orf72-dependent metabolic dysfunctions [184]. This study unveiled distinct metabolic profiles and the loss of metabolic flexibility in C9orf72 and sALS cells in comparison to controls [184]. Specifically, deficiencies were noted in the metabolism of adenosine, fructose, and glycogen along with disruptions in the membrane transport of mitochondrial substrates, contributing to the loss of metabolic flexibility in C9orf72 and sALS cells. These findings underscore the importance of metabolic flexibility in scenarios characterized by bioenergetic stress [184].
Metabolomic investigations have identified distinct metabolite profiles in samples from both ALS patients and mouse models when compared to controls [182,183,185,186,187,188,189,190,191,192]. These studies have revealed alterations in metabolites linked to glucose metabolism [185,189,190,191], with correlations established between these changes and ALS disease progression [186,190,191]. Some preliminary studies have even explored the potential of these metabolic profiles to predict ALS onset [182]. Additionally, elevated levels of ketone bodies have been observed in the cerebrospinal fluid of ALS patients, suggesting potential alterations in lipid beta oxidation [185,186]. Notably, distinct metabolite profiles have been observed among different subtypes of ALS patients, underscoring the heterogeneity of the disease [187,188].

2.4. Mitochondrial Dynamics and Biogenesis

Mitochondrial dynamics and biogenesis are tightly regulated by a complex network of signaling pathways within cells [193]. Mitochondria exist as part of a highly dynamic network that relies on a delicate balance between fission and fusion events to maintain their functional integrity. These processes control mitochondrial size, number, and shape, facilitating the mixing of mitochondrial contents, including proteins, lipids, and DNA [194,195]. Fission plays a crucial role in eliminating dysfunctional mitochondria, while fusion enables the intermingling of contents from both functional and defective mitochondria, thereby preserving overall mitochondrial integrity [196].
The key regulator of mitochondrial fission is dynamin-related protein 1 (Drp1), which becomes phosphorylated and relocates from the cytosol to the OMM when a mitochondrion is targeted for fission [197,198]. Drp1 is recruited by various receptor proteins, including mitochondrial fission factor (Mff) [199], mitochondrial fission 1 protein (Fis1) [200], and mitochondrial dynamics proteins of 49 and 51 kDa (MiD49/51) [201]. Following recruitment, Drp1 oligomerizes and forms a ring-like structure through self-assembly and GTP hydrolysis, and this structure severs the mitochondrial membrane, leading to mitochondrial fragmentation [202,203].
Conversely, the process of mitochondrial fusion involves both the IMM and the OMM. Major players in OMM fusing are mitofusins 1 and 2 (Mfn1 and Mfn2), while optic atrophy protein 1 (OPA1) governs the IMM fusion [204,205]. Mfn1 and Mfn2, situated in the OMM, facilitate the fusion of mitochondrial membranes. OPA1, located in the IMM, interacts with both Mfn isoforms, forming intermembrane protein complexes that orchestrate the fusion of the OMM to the IMM [206,207]. Given the critical dependency and sensitivity of neurons to alterations in mitochondrial dynamics, several studies have underscored the link between the disruption of mitochondrial dynamics regulators, such as Drp1, Mfn2, and OPA1, and the emergence of dysfunctional mitochondria, often leading to neurodegenerative processes [208,209,210,211].
The disruption of mitochondrial dynamics has been extensively documented in both animal models [31,34,35,212,213] and ALS patients [214]; however, the precise mechanisms underlying these alterations are still a subject of debate. While specific changes may vary among different ALS models, most studies consistently reveal an increase in mitochondrial fission relative to fusion [30,31,215]. In the case of SOD1G93A mice, researchers noted an initial increase in the mitochondrial fusion proteins Mfn1 and OPA1 during the early stages of the disease, followed by a significant decrease in later stages [31]. This biphasic pattern could suggest an initial protective response involving mitochondrial fusion followed by a gradual shift towards increased fission. Lines of evidence have revealed that mitochondrial dynamics are also under cellular redox control. It was also already discussed here that mutSOD1 may contribute to an imbalance in the GSH:GSSG ratio towards a more oxidizing environment [111]. Using HeLa cells, Shutt and collaborators have shown that GSSG strongly induces mitochondrial fusion by inducing the formation of disulfide-mediated Mfn oligomers in a process requiring GTP hydrolysis, whereas neutralizing oxidative stress with GSH or ROS scavengers impaired the fusion process [216]. On the other hand, a study conducted using endothelial cells showed that the depletion of the mitochondrial-located protein disulfide isomerase A1, an oxidoreductase with thiol reductase activity for Drp1, induced a ROS-mediated protein post-translational modification in which a reactive cysteine thiol is transformed into cysteine sulfenic acid. This Drp1 post-translational modification, termed protein sulfenyltion, is a key initial readout of redox signaling that ultimately leads to mitochondrial fragmentation, further ROS increases, and mitochondrial respiratory impairment [217]. This way, one can hypothesize that this or similar ROS-mediated sulfenylation processes can occur in ALS and further increase mitochondrial fission and its detrimental effects. Additionally, mitochondrial fission-related proteins, such as pSer637-Drp1 and Fis1, were found to increase during the early stages of the disease, potentially targeting cells for apoptosis and mitophagy [31]. Similar alterations in mitochondrial dynamics were observed in skeletal muscle biopsies from ALS patients [214] and in the MNs of rat spinal cords [212].
Although less extensively studied, models of ALS associated with TDP-43 [34,35,213], FUS [218,219], and C9orf72 [70] mutations have also exhibited evidence of mitochondrial fragmentation. For instance, in TDP-43 models, there was an increase in Fis1 protein levels and phosphorylated Drp1 along with decreased Mfn1 protein expression in the brain lysates of transgenic (TDP43PrP) mice expressing full-length human TDP-43 (hTDP-43) driven by the mouse prion promoter [213]. Moreover, OPA1 protein expression was found to be increased in brain lysates of 3-month-old TDP-43A315TKi animals, decreasing with age, particularly at 15 months of age. This suggests that young TDP-43A315TKi ALS mice experience enhanced mitochondrial fusion that shifts toward greater mitochondrial fragmentation as they age [34]. Interestingly, overexpression of mutant TDP-43 (Q331K and M337V) in cultured neurons led to a shift in the mitochondrial fission–fusion balance towards fission, causing mitochondrial fragmentation in the dendrites and cell bodies of MNs rather than in axons. Conversely, TDP-43 knockdown shifted the balance towards fusion, resulting in mitochondrial elongation. These findings indicate a specific role of TDP-43 in regulating mitochondrial dynamics [35]. Mitochondrial fragmentation has also been reported in MNs expressing either R521G or R521H FUS [218] and in HT-22 mouse neuronal-like cells expressing the P525L-mutant FUS [219] compared to neurons expressing wild-type FUS. Similarly, FUS transgenic flies exhibited mitochondrial fragmentation compared to wild-type flies [219]. Additionally, a reduction in the length of mitochondria was observed in fibroblasts from patients expressing ALS mutant C9orf72, suggesting a disruption in mitochondrial dynamic balance driven by the most prevalent ALS mutation [70].
Mitochondrial biogenesis plays a pivotal role in maintaining mitochondrial homeostasis throughout the cellular life cycle, ensuring that the physiological demands of eukaryotic cells are met by synthesizing new mitochondrial components to replace damaged mitochondria [29,220]. This tightly regulated process involves the coordinated efforts of both mtDNA and the nuclear genome to create new mitochondria [221]. One of the principal regulators of mitochondrial biogenesis is peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC-1α), which stimulates mitochondrial biogenesis by activating various transcription factors, including nuclear respiratory factor 1 (NRF1) and nuclear respiratory factor 2 (NRF2, also known as GA-binding protein-GABP and not to be confused with Nrf-2) [29,220]. These transcription factors, in turn, activate mitochondrial transcription factor A (Tfam), which is responsible for the transcription of nuclear-encoded mitochondrial proteins and the maintenance and replication of mtDNA [29].
Numerous studies have provided evidence of altered mitochondrial biogenesis markers in both ALS mouse models and patients [222,223,224]. In the mouse model carrying SOD1G93A, research has revealed diminished levels of PGC-1α mRNA in the spinal cord [222,223,224] and skeletal muscle [214,224]. Notably, increased levels of PGC-1α mRNA were reported in the skeletal muscle of the same mouse model [222]. Furthermore, protein expression levels of PGC-1α were found to be reduced in the skeletal muscle of the SOD1G93A model [214,224] and in the skeletal muscle of ALS patients [214]. Collectively, these findings suggest that both PGC-1α mRNA and protein expression may play a role in the pathogenesis of the disease, particularly in the context of SOD1 mutations. In this same model, Tfam mRNA levels were shown to decrease in the spinal cord and skeletal muscle [224]. Additionally, protein levels of NRF1 were found to be diminished in the skeletal muscle of the SOD1G93A mouse model [214,224] and in the skeletal muscle of ALS patients [214]. Taken together, these observations of varying protein and mRNA levels of mitochondrial markers in both animal models and patients, particularly in the spinal cord and skeletal muscle of individuals with SOD1 mutations, indicate a potential disruption in the mitochondrial biogenesis process affecting the cellular capacity to maintain a healthy network of functional mitochondria, which further underscoring mitochondrial relevance in ALS pathogenesis. Interestingly, Tfam protein levels in the lymphoblasts of ALS patients are of potential diagnostic and therapeutic interest [174].
Several studies have reported variations in mtDNA levels in both ALS mouse models and patients, shedding light on the complexity of mitochondrial dynamics [225,226]. For instance, the SOD1 knockout mouse model (SOD−/−) exhibited heightened mtDNA levels in skeletal muscle [225]. In patients—notably those with SOD1 and C9orf72 mutations—peripheral blood mtDNA levels were also found to be elevated [226]. Additionally, fibroblasts from C9orf72 patients displayed increased mtDNA levels and mitochondrial mass [70]. These observed increases in mtDNA content likely represent a compensatory mechanism in response to the abnormal mitochondrial function observed in these individuals [70,226]. Conversely, fibroblasts from patients harboring TARBDP mutations exhibited reduced mitochondrial mass, without alteration in mtDNA levels [70]. This suggests that TARBDP mutant fibroblasts might not possess the capacity to activate mitochondrial proliferation mechanisms to counterbalance dysfunctional mitochondria.
In summary, various studies indicate that mtDNA levels are subject to alterations, spanning from mouse models to human patients, hinting at their potential as a marker for the pathological processes underlying ALS. Furthermore, these mtDNA alterations seem to differ when comparing different mutations [70,226], especially between mutSOD1 and C9orf72 patients versus TARBDP patients. As a result, these findings prompt speculation that distinct mutations may induce specific modifications in mitochondrial biogenesis markers, although further research is needed to elucidate the precise mechanisms involved.

2.5. Mitochondrial Trafficking in ALS

Mitochondrial trafficking in neurons is a highly orchestrated process that varies depending on the direction of the movement, the polarity, and the arrangement of neuronal microtubules (MTs) [227]. MTs are polymeric structures composed of α- and β- tubulin heterodimers with a polar head-to-tail structure. Within MTs, α-tubulin monomers are oriented toward the slower-growing end (the minus end), while β-tubulin faces the faster-growing end (the plus end). In neurons, the minus end faces the cell body and the plus end extends toward the cellular extremities [228,229]. Mitochondria can move in both directions along these MTs. Retrograde transport, from the extremities to the cell body, is powered by dynein motors, whereas the anterograde transport, from the cell body to the extremities, is orchestrated by the kinesin superfamily [230,231]. These movements are energy-dependent, relying on ATP hydrolysis [232].
In addition to the transporting complexes, various protein adaptors play critical roles in the trafficking of neuronal mitochondria. These adaptors include the mitochondrial rho (MIRO)/Milton complex in Drosophila [233,234], trafficking kinesin protein (TRAK) 1 and 2 in mammals [235,236], syntabulin [237], and RAN-binding protein 2 (RANBP2) [238], among others. Beyond MTs, actin filaments and their associated protein motor adaptors also contribute significantly to mitochondrial movement. Myosins 2 and 19, for instance, play crucial roles in the short-distance movement of mitochondria along these filaments [239,240]. During mitochondrial transit, there are instances when they must halt to facilitate proper axonal and synaptic functions, which necessitates dissociation from the motor proteins and docking [241]. Syntaphilin appears to be a key protein involved in this process, immobilizing mitochondria by anchoring them to the MTs (Figure 2) [242,243].
While the precise biological and molecular origins of the phenomenon remain a topic of significant debate, the defective axonal trafficking of mitochondria stands out as one of the earliest neuropathological features observed in ALS models [244]. Various lines of evidence have consistently demonstrated that alterations in mitochondrial trafficking and morphology occur within the sciatic nerve of ALS SOD1 mutant mice. Notably, these abnormalities manifest in the early stages of ALS, even before any clinical symptoms become apparent. This malfunction typically commences with disruptions in retrograde mitochondrial transport, which precede defects in anterograde transport and ultimately culminate in mitochondrial fragmentation, as also observed in TDP43 ALS mutant mice [245]. These retrograde transport impairments have been associated with the inadequate removal of damaged mitochondria via mitophagy [246,247]. In both ALS models, anomalies in anterograde transport appeared at later stages of the disease and were linked to abnormal mitochondrial clustering (Figure 2) [245]. The net increase in the retrograde trafficking of mitochondria may signify an enhanced clearance of damaged mitochondria via mitophagy [248]. Nevertheless, it is important to note that in rat cortical neurons transfected with mutSOD1, a different scenario unfolds; these neurons exhibited decreased anterograde transport of mitochondria, without concurrent retrograde transport issues. This disturbance resulted in increased mitochondrial content in cell bodies compared to the cell axons, potentially triggering axonal stress by limiting mitochondrial Ca2+ buffering and diminishing ATP synthesis derived from mitochondria (Figure 2) [249,250].
Exploring the molecular mechanisms underlying the reduction of anterograde transport in cortical neurons from the E18 embryos of ALS SOD1 mutant rats, researchers have identified a decreased expression of Miro protein in these neurons. This is a hallmark of PINK1/Parkin-dependent mitophagy [248], where PINK1 phosphorylates Miro, targeting it for Parkin-dependent degradation [251] and ultimately leading to the cessation of mitochondrial transport and the buildup of damaged mitochondria. Interestingly, it has been found that either the restoration of Miro1 expression or the knockdown of Pink1 can restore mitochondrial motility [248]. Additionally, various causes of alterations in mitochondrial have been proposed, including pathogenic mutations affecting the axonal transport machinery itself, such as kinesin-1, dynein, Profilin 1, Dynactin, and α-tubulin [244,252,253,254]. Furthermore, indirect processes appear to contribute significantly to this phenomenon, involving pathogenic kinase signaling, microtubule destabilization, and protein aggregation [244].
It is essential to emphasize that different mutant models yield distinct effects, underscoring the significance of recognizing that diverse alterations may arise when utilizing different biological models and cell types.

2.6. Mitochondria and Endoplasmic Reticulum Crosstalk

Mitochondria play a significant role in managing Ca2+ levels, influencing energy metabolism and cellular activities. This is especially relevant for neurons given that mitochondria localize in neuronal cell bodies and in dendrites, axons, and synaptic terminals. When there is a high concentration of Ca2+ in microdomains near mitochondria, the organelle efficiently accumulates Ca2+, primarily through the Ca2+ uniporter (MCU). This electrogenic carrier is situated at the IMM and operates based on the negative transmembrane potential of mitochondria. This accumulation is facilitated by microdomains adjacent to endoplasmic reticulum (ER) subcompartments that physically interact with mitochondria, referred to as endoplasmic reticulum–mitochondria-associated membranes (MAMs) (reviewed in [255]). MAMs have also been demonstrated to play a crucial role in regulating autophagy and mitochondrial dynamics through the modulation of Drp1 or Mfn2 [256,257]. These complex structures are composed of numerous proteins from both mitochondria and the ER, with additional components yet to be identified [258]. Intriguingly, mass spectrometric analysis has identified more than 1000 proteins in MAM fragments [259]. As MAMs function as a signaling hub, their effectiveness heavily relies on proteins that perform diverse tasks. These proteins are categorized based on their primary functions. For instance, in the domain of Ca2+ transport, inositol 1,4,5-triphosphate receptor (IP3R), VDAC1, 2b isoform of Sarco/Endoplasmic Reticulum Calcium ATPase (SERCA), Sigma-1 receptor (Sigmar1), and glucose-regulated protein 75 (Grp75) [260,261,262,263] have been extensively studied. Regarding autophagy, autophagy-related 14 (ATG14) and autophagy-related 5 (ATG5) play pivotal roles [264]. Tethering proteins responsible for the physical contact between mitochondria and the ER, such as ER protein vesicle-associated membrane protein-associated protein B (VAPB) and protein tyrosine phosphatase interacting protein 51 (PTPIP51) [265], are also noteworthy.
Numerous studies indicate that ALS-associated mutations are linked to impaired enzymatic activities within MAM domains, specifically impacting calcium transfer between the ER and mitochondria (reviewed in [266]). Notably, proteins like TDP-43 and FUS induce the dissociation of VAPB and PTPIP51 [265,267], thereby modifying calcium transfer dynamics between these organelles [265]. Moreover, mutations in the VAPB gene have been demonstrated to induce a familial variant of ALS. This could be attributed to a decrease in the effectiveness of Ca2+ elimination, potentially resulting from inadequate energy to power the Ca2+ ATPases, leading to the accumulation of surplus Ca2+ within mutated neurons and triggering changes in synaptic vesicle release and the development of synapses [268]. These findings are further supported by a study showing that disruption of VAPB–PTPIP51 tethering and the related delivery of Ca2+ from ER stores to mitochondria occurs in neurons derived from iPS cells obtained from patients carrying pathogenic c9orf72 expansions associated with ALS/FTD [269]. Recently, Pilotto and colleagues discovered that ER stress in human C9ORF72-ALS/FTD iPSCs and C9orf72 mouse neurons initiates a compensatory mechanism [270]. This involves the upregulation of GRP75 expression at MAMs, effectively counteracting early mitochondrial Ca2+ imbalance. Over time, PolyGA inclusions sequester GRP75, resulting in its diminished presence at the MAMs and leading to functional loss and subsequent mitochondrial dysfunction [270].
Several molecules have been investigated as potential therapeutic agents addressing dysfunction at MAMs in ALS (reviewed in [271]). For instance, salubrinal, an ER stress inhibitor, displayed promising results in SOD1G93A mice, alleviating disease progression by enhancing survival rates and promoting neuromuscular junction reinnervation [272]. Additionally, PRE-084, a sigma-1 receptor agonist, demonstrated notable effects, reducing SOD1G93A-induced cell death and extending the survival time of mice [273]. Another example is Sephin1, an ER stress inhibitor, which exhibited significant efficacy in preventing motor neuron degeneration in both in vitro and in vivo ALS models [274,275]. These studies highlight diverse strategies and potential therapeutic avenues for addressing MAM dysfunction in ALS.

2.7. Autophagy and Mitophagy

Autophagy is a pivotal catabolic process in cellular recycling and the maintenance of intracellular homeostasis that involves the breakdown of unnecessary or damaged cellular components through lysosomal degradation [276]. More specifically, a form of autophagy known as macroautophagy encompasses a series of intricate signaling events that initiate with the sequestration of proteins and damaged organelles into autophagosomes, followed by their digestion via lysosomal hydrolysis [277]. Dysregulation of autophagic flux has been closely associated with various pathological processes [278,279], including neurodegenerative diseases, such as ALS [280].
In ALS, autophagy is known to occur in the MNs of the spinal cord in both postmortem human ALS patients [281,282] and animal models of the disease [283,284,285,286]; however, its precise functional role in disease progression remains elusive. Mutations in genes encoding proteins for crucial selective autophagy receptors, such as optineurin (OPTN) [217], p62/Sequestosome-1, and Tank-binding protein (TBK1) [287], have been implicated in ALS, contributing to the impairment of autophagic flux. For instance, ALS patients with C9orf72 expansion exhibited p62 inclusions [281,282], and pathogenic mutations in SOD1 were found to lead to polyubiquitinated aggregates of misfolded mutSOD1 that interact with p62, thereby enhancing the association between mutSOD1 and microtubule-associated light chain 3-II (LC3-II), a primary autophagic marker embedded into autophagosomes [288,289].
Moreover, mutations in the OPTN gene have been found in ALS patients [217] and colocalize with SOD1 inclusions [290]. Overexpression of missense mutated OPTN resulted in increased LC3-II levels, potentially impairing autophagosome transport or fusion events [291]. ALS cases have been associated with genes involved in vesicular trafficking, including VCP, CHMP2B, VAPB, ALS2, and DCTN1 (genes encoding valosin-containing protein, charged multivesicular body protein 2B, VAMP associated protein B, alsin Rho guanine nucleotide exchange factor, and dynactin subunit 1 protein, respectively), which may influence autophagy flux either directly or indirectly [292].
Several studies underscore the critical role of autophagy in ALS pathogenesis. mutSOD1 mouse models have exhibited hyperactivated autophagy, characterized by AMPK activation and mTOR downregulation [293], along with increased conversion of LC3-I into LC3-II [294], indicative of heightened autophagic vacuole formation [295,296]. These models have also shown elevated levels of autophagic regulators, such as TFEB, involved in autophagy initiation and lysosomal biogenesis [297], and Beclin1, an autophagic initiator [298], resulting in increased LC3-II levels [299]. Additionally, mutSOD1′s toxic effect extends to excessive interaction with dynein, a protein responsible for axonal retrograde transport. This overload disrupts autophagosome retrograde transport, contributing to the failure of late-stage fusion steps in autophagy [300].
In the context of TDP-43 and FUS, both proteins, much like SOD1, tend to undergo misfolding and aggregation, a process that coincides with autophagic impairment [301,302]. Notably, previous studies have demonstrated that the pharmacological enhancement of autophagy using rapamycin decreased cell death in neurons carrying FUS mutations [303]. Additionally, it promoted the clearance of TDP-43 aggregates in Drosophila and mouse models, resulting in a reduction of ALS symptoms [304,305].
Concerning C9orf72, human neurons derived from iPSC obtained from individuals with C9orf72 repeat expansions exhibited elevated p62 levels and heightened sensitivity to autophagy inhibitors like chloroquine and 3-methyladenine compared to control neurons [306]. These findings suggest a compromised autophagy function in these neurons. In addition, the loss-of-function of C9orf72 can disrupt the mTOR signaling pathway, inhibiting its activity [307,308]. Furthermore, there is a correlation between C9orf72 loss and an increase in lysosomal biogenesis along with an upsurge in swollen lysosomes [308]. Supporting these observations, C9orf72 KO mice displayed elevated levels of lysosomal markers such as LC3I, LAMP1, and PSAP [309], indicating that C9orf72 likely plays a regulatory role in the autophagy/lysosome pathway.
Although extensive research has explored the autophagy process in ALS, the selective degradation of mitochondria through autophagy, known as mitophagy, has only recently gained attention. Mitophagy is of great significance as it serves to eliminate defective and irreparable mitochondria, preventing their accumulation and potential harm to the cell [310]. Consequently, mitophagy plays a central role in mitochondrial quality control.
In cases of mitochondrial dysfunction, the protein PINK1, which typically relocates from the OMM to the IMM for cleavage under normal circumstances, accumulates on the OMM. This accumulation acts as a signal for the recruitment of the E3 ubiquitin ligase Parkin [311]. PINK1 phosphorylates Parkin and ubiquitin, activating Parkin’s E3 ligase activity [311]. Subsequent pathways have been proposed to describe the connection to the autophagy process, although some uncertainties persist. For instance, parkin-independent mitophagy pathways have been identified, notably through the recruitment of the autophagy receptor OPTN by PINK1 [312]. Nevertheless, during mitophagy, defective mitochondria must either be enveloped by an autophagosome or fused with a lysosome [311]. Of note, mutations associated with ALS in genes such as OPTN and TBK1 have been shown to disrupt mitophagy. This disruption may result in the accumulation of dysfunctional mitochondria, potentially contributing to the neuronal degeneration observed in ALS [313,314].
Recently, a novel concept has emerged regarding the removal of dysfunctional mitochondria within cells, termed “mitoautophagy” [315]. In a groundbreaking study, it was discovered that a substantial number of flawed mitochondria in corticospinal motor neurons (CMNs), particularly in young prpTDP-43A315T and PFN1G118V mice, underwent a self-destructive process without relying on lysosomes or autophagosomes as typically observed in mitophagy [315]. Interestingly, distinct mechanisms of mitochondrial removal were identified in CMNs from SOD1G93A and prpTDP-43A315T mice, with the former being autophagosomal dependent. This suggests that mitoautophagy might be specific to TDP-43-associated pathology [315]. This difference was attributed to the absence of TDP-43 pathology in SOD1G93A mice [315]. As ALS patients with SOD1 mutations also lack TDP-43 pathology [315,316], this raised the possibility that various pathways for mitochondrial removal are activated in MNs, each potentially playing a distinct role in ALS development. Consequently, further research is warranted to comprehensively unravel the intricate relationship between mitochondria and ALS pathological processes.

2.8. Apoptotic Mechanisms

The apoptosis in MNs plays a pivotal role in the neurodegeneration observed in ALS. Various apoptotic markers have been identified in postmortem spinal cord tissues of ALS patients [317,318] and ALS mouse models [319]. The activation of several caspases has been closely associated with ALS neurodegeneration, encompassing upstream caspases 1 and 9 as well as downstream executioner caspases 3 and 7 [320,321,322]. Additionally, the release of cytochrome c from mitochondria to the cytosol has been observed in the spinal cords of SOD1G93A transgenic mice, underscoring the clear involvement of the mitochondrial apoptotic pathway in mutSOD1-mediated toxicity [322].
In concert with apoptotic pathways, ROS and calcium overload have emerged as common pathological features in the neurodegenerative process of ALS, implicating mitochondrial dysfunctions and cell-death mechanisms. In terms of calcium uptake, it was noted that mutSOD1 impaired mitochondrial calcium-loading capacity in the brain and spinal cord of G93A mutant transgenic mice early in the disease course, long before motor weakness sets in, and massive neuronal death occurred [323].
With respect to calcium dysregulation, several studies have explored the role of the mitochondrial permeability transition pore (mPTP) in the degeneration of MNs in ALS. The mPTP is a megachannel located in the IMM known to function in both physiological and pathological conditions [324], including various neurodegenerative disorders [325]. While under normal physiological conditions, the mPTP serves as a rapid mitochondrial efflux mechanism for managing ROS and calcium overload, under pathological conditions, prolonged mPTP opening can lead to rapid mitochondrial swelling driven by the osmotic pressure of matrix solutes. This ultimately results in the rupture of the OMM and the collapse of ΔΨm, culminating in apoptotic cell death when sufficient ATP is available or necrosis when ATP levels are insufficient [326,327].
Despite the ongoing quest to fully understand the molecular identity of the pore-forming unit, various models have been proposed. These range from models that involve regulatory proteins, such as cyclophilin D (Cyp-D), adenine nucleotide translocator (ANT), VDAC, hexokinase-II (HK-II), mitochondrial creatine kinase (mtCK), and translocator protein of 18 kDa (TSPO) [328,329], to more recent hypotheses that also encompass ATP synthase [329,330].
Specifically concerning ALS, high expression levels of mPTP regulators, including VDAC, ANT, and Cyp-D, were found in mouse MNs, as well as nitration of Cyp-D and ANT in spinal cord fractions of symptomatic G93Ahigh-mSOD1 mice [331]. Additionally, the genetic deletion of Cyp-D delayed disease onset and extended the survival of G93A-mSOD1 mice expressing high and low levels of the mutant protein. These findings suggest that mPTP is involved in the causal mechanisms of MN degeneration in mouse ALS, partly due to oxidative and nitrative damage to pore regulatory proteins [331]. Moreover, a study comparing knockout Cyp-D mice overexpressing SOD1 mutations (G93A, G85R, and G37R) with littermates overexpressing Cyp-D found that the absence of Cyp-D improved mitochondrial ATP synthesis, reduced the levels of misfolded SOD1 aggregates in the spinal cord, and suppressed MN death [332]. Intriguingly, an increased expression of Cyp-D was also observed in association with mitochondrial ROS production and enhanced mitoflash activity in the muscle of SOD1G93A transgenic mice at the onset of symptoms [333]. Furthermore, the application of cyclosporin A (CsA), an inhibitor of Cyp-D, attenuated SOD1G93A-induced mitoflash activity, suggesting the involvement of mPTP opening in muscle dysfunction [333].
Considering that the denervation of skeletal muscle deprives it of any electric stimulation—resulting in the loss of physiological calcium uptake by mitochondria, increasing mitochondrial ROS production and triggering mPTP opening—it is plausible to propose that mPTP opening is a key element in the denervation process observed in ALS [334]. Further research is required to elucidate this aspect comprehensively.

2.9. Mitochondrial Dysfunction as Cause and/or Consequence in ALS

As described above, mitochondrial dysfunction in ALS is a multifaceted phenomenon, and determining whether it is a cause or consequence of the disease remains a challenge. While the literature suggests that mitochondrial impairment is observed early in the disease process (e.g., [335,336,337]), pinpointing the exact initiating event is complicated by the heterogeneity of ALS cases [187,338,339].
The fact that numerous studies reported mitochondrial abnormalities in ALS patients and animal models at early, pre-symptomatic stages suggests that mitochondrial dysfunction is an early event in the disease cascade, possibly contributing to the initiation or acceleration of ALS pathogenesis. On the other hand, mitochondrial dysfunction could also be a consequence of the disease’s evolution. Motor neuron degeneration, protein aggregation, and other ALS-related mechanisms may induce stress on mitochondria, leading to their dysfunction as a downstream effect [340]. Importantly, mitochondrial dysfunction and ALS pathogenesis may engage in a reciprocal relationship, forming a vicious cycle; initial mitochondrial impairment could exacerbate disease progression, while ongoing neurodegeneration may further compromise mitochondrial function. The presence of specific genetic mutations associated with both mitochondrial function and ALS adds another layer of complexity. Interactions between genetic susceptibility and environmental factors may influence the dynamics of mitochondrial involvement in ALS. The elusive nature of causality underscores the urgency of continued investigations to unravel the intricate relationship between mitochondria and ALS, paving the way for targeted and effective therapeutic interventions.

3. Preclinical and Clinical Endeavors Targeting Mitochondria

Understanding and harnessing ways to enhance mitochondrial function has emerged as a promising avenue for therapeutic development in ALS and other neurodegenerative diseases. This section delves into innovative approaches, compounds, and therapies that target mitochondria to restore their vitality, promote cellular energy production, and ultimately address the underlying causes of diseases associated with mitochondrial dysfunctions, such as ALS. From metabolic enhancers to antioxidants and mitochondria-targeted agents, these approaches offer hope for more effective treatments and improved quality of life for those affected by ALS and other mitochondrial-related conditions (Figure 3). The results of relevant clinical trials (Table 1) will also be discussed in this section.

3.1. Approaches for Direct Enhancement of Mitochondrial Function

Metabolic enhancers represent a compelling class of therapeutic interventions in the context of ALS and other neurodegenerative disorders. These approaches aim to boost cellular energy production and stimulate mitochondrial function. As metabolic deficits are a hallmark feature in ALS, emerging research suggests that enhancing metabolic pathways within mitochondria may hold the key to slowing or mitigating disease progression.

3.1.1. Dichloroacetate

Dichloroacetate (DCA), investigated in multiple neurodegenerative diseases [341], represents an alternative approach for indirectly enhancing mitochondrial function. As a pyruvate dehydrogenase kinase (PDK) inhibitor, DCA stimulates the conversion of pyruvate to acetyl coenzyme A (AcCoA), supplying additional energy substrates to the TCA cycle (Figure 3A).
Administered to mutant SOD1G93A or SOD1G86R mice at 500 mg/kg/day during the pre-symptomatic stage, DCA improved survival, delayed disease onset, reduced spinal motor neuron loss, and enhanced mitochondrial function [173,342]. For example, DCA treatment in SOD1G86R mice not only improved glycolytic capacity but also upregulated the expression of genes associated with mitochondrial biogenesis (e.g., Pgc-1α and Mfn2), believed to impede disease progression in this animal model [173]. Moreover, DCA (5 mM) improved the mitochondrial function of abnormal glial cells isolated from the spinal cords of adult paralytic SOD1G93A rats, enhancing respiratory capacity and decreasing toxicity to MNs [343].
Additionally, administering DCA (100 mg/kg for 10 days) to symptomatic SOD1G93A rats reduced MN degeneration, gliosis, and the number of GFAP/S100β double-labeled hypertrophic glial cells in the spinal cord. These findings indicate DCA’s potential therapeutic strategy for ALS by modulating glial metabolism and mitigating MN degeneration [343].
However, no clinical trials have assessed DCA effectiveness in ALS patients, necessitating further research.

3.1.2. Ketogenic and High-Fat Diets

Various dietary interventions aim to delay ALS progression and counteract the hypermetabolism associated with the condition. Two notable approaches include the Ketogenic Diet (KD) [344,345] and the High-Fat Diet (HFD) [136,346,347,348] (Figure 3A).
A KD, characterized by high fat intake, increases circulating ketones, while restricting carbohydrates and proteins [349,350]. The primary ketones acetoacetate and D-β-3-hydroxybutyrate, produced in the liver, serve as an energy source when glucose availability is limited [351]. Studies report that a KD, comprising 60% fat, 20% carbohydrates, and 20% protein, improves motor function and safeguards MNs in SOD1G93A mice [344,345].
This diet exerts its effects partially through alterations in mitochondrial function, promoting ATP synthesis, and restoring the activity of complex I in the ETC, which is often impaired in ALS [345]. Furthermore, it was demonstrated that caprylic triglyceride, a precursor to ketone bodies, enhanced motor function and protected MNs in SOD1G93A mice by boosting oxidative metabolism, thereby increasing mitochondrial basal and maximum oxygen consumption [344]. Converted rapidly to caprylic acid, it easily traverses membranes, becoming β-oxidized to AcCoA in the mitochondria. Consequently, it supplies ketone bodies to the TCA cycle, serving as a rapid energy source when cellular glucose levels are low. Notably, neither the KD nor caprylic triglyceride significantly altered the survival of SOD1-G93A transgenic mice [344,345].
Regarding the HFD, studies have indicated its potential to slow disease progression in ALS mouse models. One study employing a diet composed of 47% fat, 38% carbohydrates, and 15% protein reduced disease progression in a SOD1G93A mouse model [348]. Similarly, another study utilizing a HFD comprising 21% fat and 0.15% cholesterol extended the mean survival of SOD1G86R mice [136]. To investigate the impact of high-caloric diets on ALS patients, a double-blinded trial known as the LIPCAL-ALS study (NCT02306590) enrolled 201 patients; participants were assigned to receive either a high-caloric fatty diet (HCFD, 405 kcal/day, 100% fat) or a placebo in conjunction with riluzole (100 mg/day). However, the results did not provide conclusive evidence of a life-prolonging effect of the diet on the overall ALS patient population. Nevertheless, post-hoc analysis revealed a significant survival benefit for a subgroup of fast-progressing patients [352]. Given the potential influence of caloric content on the intervention’s efficacy, a clinical Phase I LIPCALII study (NCT04172792) is set to explore whether an ultra-high caloric diet (UHCD), featuring twice the caloric content compared to LIPCAL-ALS, can be well tolerated by ALS patients over four weeks and if it can exceed the beneficial effects of LIPCAL-ALS. In another study [353] assessing the effects of a high-caloric nutrition protocol on ALS patients with percutaneous gastrostomy, 40 patients were randomly assigned to either a routine diet (control group) or high-caloric nutrition combined with the routine diet (Ensure group) for six months. This study demonstrated a significant increase in cumulative survival rates in the Ensure group, suggesting the potential of this approach to enhance nutritional status and longevity.
In summary, the mechanisms through which HFD and KD modify ALS disease progression in mouse models are unclear. However, it is suggested that the high fat content of these diets might elevate phospholipids and cholesterol, which are crucial components for axonal membrane assembly and regeneration [354]. In alignment with this hypothesis, it was demonstrated that abnormally elevated cholesterol levels were associated with increased survival in ALS patients, by more than 12 months, suggesting hyperlipidemia as a prognostic factor [355].

3.1.3. Acetyl-Carnitine

Acetyl-carnitine (ALC) is a crucial cellular source of acetyl groups (Figure 3A), particularly in high-energy demanding situations, playing a pivotal role in transporting long-chain fatty acids across mitochondrial membranes and limiting β-oxidation rates [356].
Studies administering 50 mg/kg/day of ALC orally before disease onset significantly delayed symptoms, slowed motor function deterioration, and extended lifespan in mutant SOD1G93A mice; subcutaneous ALC injection to symptomatic mutant SOD1G93A mice improved survival [357].
In a randomized double-blind, placebo-controlled Phase II trial involving 82 patients receiving 3 g/day of ACL or a placebo with riluzole (100 mg/day), ALC demonstrated mild enhancements in ALSFRS score and respiratory capacity. Notably, it doubled the median survival range compared to the placebo group, indicating effectiveness, tolerability, and safety in ALS treatment [358]. A recent study involving 32 ALS patients revealed ALC’s improvement in the redox state, persisting six months post-treatment, offering potential disease biomarkers and drug effects indicators in clinical practice and trials [359].

3.2. Antioxidants

Mitochondrial dysfunction, either causal or consequential to oxidative stress, contributes to ALS pathogenesis. While oxidative damage and mitochondrial dysfunction are recognized contributors to ALS pathogenesis, it is disheartening that all antioxidant treatments tested in patients have thus far proved ineffective. This highlights the urgent need for advancements in antioxidant therapies for ALS treatment. In the following section, we delve into various preclinical and clinical trials, both completed and ongoing, that explore antioxidants capable of influencing mitochondrial homeostasis (Figure 3B).

3.2.1. N-acetyl-L-cysteine (NAC)

N-acetyl-L-cysteine (NAC), a membrane-permeable antioxidant, replenishes cellular cysteine and glutathione pools, mitigating free radical damage [360,361] (Figure 3B). In a preclinical study, NAC (1 mM, 24 h) attenuated mitochondrial ROS production, restored MTT reduction rates to control levels, and elevated ATP levels in human neuroblastoma SH-SY5Y cell lines harboring the SOD1G93A mutation [362]. In SOD1G93A transgenic mice, a daily dose of 2.0 mg/kg significantly extended survival and improved motor performance [360]. However, a double-blind, placebo-controlled clinical trial involving 110 ALS patients using subcutaneous NAC infusion (50 mg/kg daily) did not show a substantial increase in 12-month survival or disease progression slowdown [363]. In G93A mice, intranasal NAC administration with the nanocarrier PEG-PCL-Tat significantly increased spinal cord accumulation, extending median survival by 11.5 days. These findings highlight the potential of this approach as a promising Drug Delivery System for ALS therapeutics [364]. However, the potential benefits of NAC in ALS remain uncertain, necessitating further clinical trials in humans. The ability of NAC to traverse the Blood–Brain Barrier (BBB) has been a subject of controversy and is likely influenced by dosage and administration routes [365]. Consequently, multiple NAC derivatives have been synthesized to overcome this limitation. Important examples include N-acetylcysteine ethyl ester (NACET), which is proposed to improve pharmacokinetics but undergoes rapid transformation into NAC and cysteine resulting low plasma levels [366]; N-acetylcysteine butyl ester (NACBE), which is highly lipophilic and was shown to have superior effects after oxidative insult exposure compared to NAC [367]; and N-acetylcysteine amide (NACA), which was developed to enhance lipophilicity, membrane permeability, and the capability to traverse the BBB [368]. Previous studies have supported the protective properties of NACA, suggesting its potential clinical utility [369].

3.2.2. Edaravone

Edaravone, the active component of Radicut®, is a potent free radical scavenger widely employed in the treatment of cerebral ischemia in Japan [370,371]. Its neuroprotective role arose from its ability to eliminate lipid peroxides and hydroxyl radicals [372,373] (Figure 3B). While the precise mechanisms are not fully understood, it has been proposed that in addition to its radical-scavenging properties, edaravone might inhibit the mPTP, contributing to its neuroprotective effects [374]. It benefits various CNS cell types, including neurons, microglia [375], astrocytes [376], and oligodendrocytes [377], partly attributable to its anti-inflammatory properties [378]. Preclinical studies indicate improved motor function, slowed disease progression, and mitigated motor neuron degeneration in transgenic SOD1 rodent models of ALS treated with edaravone doses ranging from 1.5 to 15 mg/kg [379,380].
In an open-label Phase II study with 20 ALS patients, intravenous administration of edaravone (30 or 60 mg daily) was safe, well-tolerated, and slowed disease progression, as measured by the ALS-FRS scale, during the six-month treatment period compared to the six months before edaravone administration [381]. A subsequent double-blind, placebo-controlled study with 102 ALS patients showed a smaller reduction in ALSFRS-R scores in the edaravone group over a 24-week treatment period [371].
In a recent Phase III study (NCT01492686), a 24-week, double-blind, parallel-group study of edaravone showed less decline in ALSFRS-R scores at 6 months and less deterioration in quality of life in patients receiving edaravone compared to those receiving standard care [382]. Currently, edaravone is approved for ALS treatment in Japan and South Korea and it was also approved by the FDA in May 2017 [383]. However, the precise mechanism of action of edaravone in ALS treatment remains to be fully elucidated.
A Phase I trial of an oral formulation of edaravone (TW001) developed by the Treeway company demonstrated safety, tolerability, and adequate exposure levels [384]. Furthermore, in a Phase III trial (NCT04165824), oral edaravone showed a favorable safety profile in ALS patients after 48 weeks of treatment [385]. A recent meta-analysis [386] suggests potential clinical benefits of edaravone in ALS treatment, with no significant increase in adverse events or deaths in compiled randomized clinical trials. However, more high-quality research is needed for further confirmation due to the small sample sizes in the included studies.

3.2.3. Melatonin

Melatonin (N-acetyl-5-methoxytryptamine), a neurohormone secreted by the pineal gland, possesses ROS-scavenging activity and amphiphilic properties, permeating both lipophilic and hydrophilic cellular environments [387] (Figure 3B). Its potential as an experimental drug has been explored in various neurodegenerative diseases characterized by excessive ROS production owing to its robust antioxidant properties [388].
In addition to acting as a potent free radical scavenger, melatonin augments cellular antioxidant defenses by stimulating vital antioxidant enzymes (SOD, glutathione peroxidase, and glutathione reductase) and elevating GSH levels [389]. Furthermore, melatonin plays a role in preserving mitochondrial homeostasis, decreasing free radical generation, and protecting mitochondrial ATP synthesis by stimulating the activities of complexes I and IV [390].
In transgenic SOD1G93A mice, the oral administration of melatonin (57–88 mg/kg/day) at the pre-symptomatic stage delayed disease progression and extended survival [87]. The same study demonstrated that the rectal administration of 300 mg/day of melatonin in 31 sALS patients was well tolerated over 2 years, reducing circulating serum protein carbonyl levels. However, it did not show upregulation of genes encoding antioxidant enzymes [87]. The decreased oxidative damage in ALS patients under melatonin treatment, coupled with its established safety in humans, emphasizes the need for further clinical trials to elucidate its neuroprotective effects in ALS.
More recently, it was reported that administering melatonin (30 mg/kg) to pre-symptomatic SOD1G93A-transgenic mice significantly delayed disease onset, neurological deterioration, and mortality [391]. These effects involved inhibition of the caspase-1/cytochrome c/caspase-3 pathways and the loss of melatonin receptor 1A. Conversely, melatonin administration (0.5, 2.5, and 50 mg/kg, i.p.) to pre-symptomatic SOD1G93A-transgenic mice reduced their survival [392]. This study revealed increased MN loss, elevated 4-hydroxynonenal (4-HNE) levels, and upregulated expression of human SOD1 in mice treated with melatonin compared to untreated animals. These findings suggest that melatonin might exacerbate the phenotype in the SOD1G93A mouse ALS model due to the upregulation of toxic human SOD1, which may override its antioxidant and anti-apoptotic effects [392]. This raises the possibility that the SOD1G93A ALS mouse model may not be ideal for evaluating the neuroprotective properties of melatonin or other compounds with complex antioxidative properties. Further research is warranted to unravel the mechanisms of melatonin’s action and determine whether its antioxidant and anti-apoptotic effects can translate into clinical benefits.

3.2.4. Mitochondria-Targeted Antioxidants

One promising avenue in ALS research involves mitochondria-targeted antioxidants, such as 10-(60-ubiquinonyl) decyltriphenylphosphonium, known as MitoQ (Figure 3B). This compound features a triphenylphosphonium (TPP) functional group linked to the antioxidant ubiquinone [393] that penetrates biological membranes and selectively accumulates within mitochondria, driven by the ΔΨm [394]. Positioned within mitochondria, MitoQ effectively shields these critical organelles from oxidative damage [393,395].
Inside mitochondria, complex II reduces the ubiquinone moiety of MitoQ to its active ubiquinol form, bolstering the defense against oxidative damage [395]. Various oxidants can convert ubiquinol back to ubiquinone, efficiently reversed by the respiratory chain [396] and ensuring continual recycling. MitoQ efficiently mitigates oxidative damage in chronic hepatitis C virus patients, decreasing liver damage [397]. Additionally, it has shown promise in neurodegenerative diseases like Parkinson’s [398,399] and Alzheimer’s [400,401], by decreasing oxidative damage. However, a Phase II clinical trial for Parkinson’s disease (NCT00329056) yielded disappointing results [402].
Despite the setback in Parkinson’s disease trials, MitoQ has shown potential in ALS; it ameliorated nitroxidative stress and mitochondrial dysfunction in astrocytes expressing SOD1G93A, decreasing toxicity to MNs in co-cultures [403]. In SOD1G93A mice, MitoQ (500 µM) administration improved the ALS phenotype by slowing mitochondrial function decline in the spinal cord and quadriceps muscle, increasing the lifespan of the animals [404]. This treatment reduced nitroxidative markers and pathological signs in the spinal cord along with the recovery of neuromuscular junctions and a marked increase in hindlimb strength [404]. MitoQ rapidly crosses the BBB [400] and is well tolerated in animals and humans [397], with nausea being the most common side effect [405].
These findings underscore the potential of mitochondria-directed antioxidants as a strategy to delay ALS symptoms, warranting further development.
Another mitochondria-targeted antioxidant in ALS research is the mitochondria-targeted carboxy-proxyl (Mito-CP). Like MitoQ, this features a TPP cation covalently coupled to carboxy-proxyl nitroxide, accumulating within mitochondria [406]. Low doses of Mito-CP (1–10 nM) effectively prevented MN death expressing SOD1G93A induced by proapoptotic stimuli that trigger ROS formation [407]. It also prevented mitochondrial dysfunction, decreasing O2.− production in SOD1G93A astrocytes and promoting MN survival [403].
While these results are promising, further investigations are necessary to explore other mitochondriotropic compounds, focusing on reducing toxicity and enhancing therapeutic efficacy.
In addition, it is crucial to prioritize investment in the development of ALS models that faithfully represent the diverse subtypes of the disease. This approach aims to address the challenge of translating positive effects observed in preclinical trials with antioxidants into meaningful efficacy during clinical trials. These models should serve as robust platforms for conducting preclinical trials before advancing to clinical trials [22]. Human induced pluripotent stem cells have opened avenues to explore therapeutic development relevant to human diseases [408,409]. An example is the generation of MNs from a patient-derived iPSC line carrying the SOD1-A4V mutation that demonstrated significant disease phenotypes, including proteinopathy, structural attrition, axonopathy, synaptic pathology, and functional defects. This model holds the potential to emerge as a robust preclinical platform for evaluating the therapeutic efficacy of diverse molecules in addressing this disease [66].

3.3. Antiapoptotic Agents

3.3.1. mPTP-Targeting Agents

As discussed in Section 2.8, emerging evidence points towards the involvement of mPTP in ALS pathogenesis. Initial studies used CsA, which, when administered intracerebroventricularly (25 µg every other day) at the symptomatic stage in SOD1G93A mice, extended their survival [410]. Similarly, intracerebroventricular administration of CsA (20 µg/mouse/week) in pre-symptomatic SOD1G93A mice delayed the onset of hindlimb weakness, prolonged the time from onset to paralysis, and extended life span [411]. However, it is important to note that CyA struggles to cross the BBB [411,412] and its benefits could be confounded by its immunosuppressant effects due to calcineurin inhibition. Therefore, it is advisable to explore other mPTP inhibitors that lack calcineurin effects.
One such strategy involves mPTP stabilization using cholest-4-en-3-one oxime, known as olesoxime (TRO19622), which binds to VDAC and TSPO [413,414] (Figure 3C). Subcutaneous administration of TRO19622 (3 or 30 mg/kg) in pre-symptomatic SOD1G93A mice improved motor function and prolonged survival [415]. Additionally, administration of TRO19622 (600 mg/kg of food pellets) in pre-symptomatic SOD1G93A mice delayed muscle denervation, decreased astrogliosis, prevented microglia activation, and protected MNs in the lumbar spinal cord, suggesting its potential as a neuroprotective agent to delay ALS neurodegeneration [416]. However, a Phase II–III clinical trial failed to demonstrate efficacy in 512 ALS patients receiving 330 mg TRO19622 daily compared to a matching placebo group receiving 50 mg riluzole twice a day for 18 months. None of the assessed parameters, including slow vital capacity, manual muscle testing, and rates of deterioration of ALSFRS-R scores, revealed a significant clinical benefit of TRO19622 treatment compared with the placebo, except for a minimal increase in the ALSFRS-R global score over 9 months [417]. Several factors may explain the absence of clinical efficacy, including differences in the timing of administration relative to disease onset, the limitations of animal models in predicting clinical outcomes [418], and the need for further investigation into the survival-promoting effects of TRO19622 on human MNs derived from ALS patient induced pluripotent stem cells [419].
Another compound of interest is GNX-4728, a cinnamic anilide derivative that inhibits mPTP opening and has shown promise in the SODG93A mouse model (Figure 3C). Systemic treatment with GNX-4728 (15 mg/kg) in C57BL/6 mice significantly increased mitochondrial calcium retention capacity (CRC) in the heart and brain. Importantly, the increase in CRC in the brain suggests its potential to cross the BBB [420]. Furthermore, systemic administration of GNX-4728 (300 µg every other day, i.p.) in pre-symptomatic SOD1G37R mice delayed disease onset, increased lifespan, protected against motor neuron and mitochondrial degeneration, attenuated spinal cord inflammation, and preserved neuromuscular junction (NMJ) innervation in the diaphragm in ALS mice [420].
These findings highlight the potential of mPTP-targeting agents as a therapeutic approach in ALS, with the need for further research to optimize their clinical translation.

3.3.2. Rasagiline

Rasagiline, a monoamine oxidase B inhibitor primarily used to treat Parkinson’s disease [421,422], has shown neuroprotective properties beyond its MAO inhibitory activity [423] (Figure 3C). In vitro studies suggest its neuroprotective role, partly mediated by anti-apoptotic properties [424,425,426]. Rasagiline may protect mitochondria by preventing mPTP opening, inhibiting caspase 3 activation [425], or by upregulating the anti-apoptotic Bcl-2 and Bcl-xL genes and downregulating the pro-apoptotic Bad and Bax genes [426].
In ALS, oral rasagiline (0.5–2 mg/kg/day), administered alone or with riluzole (30 mg/kg/day) at the pre-symptomatic stage, improved motor performance and extended survival in SOD1G93A mice [427]. A Phase II clinical trial involving 23 ALS patients receiving rasagiline (2 mg/day) showed increased ΔΨm and oxygen radical antioxidant capacity (ORAC) in lymphocytes and a decrease in the apoptotic marker Bcl-2/Bax ratio [428]. While this trial did not find significant improvements in the rate of ALSFRS-R score decline, it did identify differences in symptom duration among patients administered who received rasagiline compared to placebo controls (NCT01232738) [428]. This raises the question of whether rasagiline-induced mitochondrial changes could slow motor function decline and extend ALS patient survival.
However, another Phase II clinical trial involving 80 ALS patients receiving rasagiline (2 mg/day for 12 months) did not demonstrate improvements in mitochondrial and molecular biomarkers (ΔΨm, ORAC, and Bcl-2/Bax ratio) or differences in the average 12-month ALSFRS-R slope between the rasagiline and placebo groups. Variability among patients and differences in sample processing timing at various centers involved in this clinical trial (NCT01786603) may explain the discrepant results regarding rasagiline’s impact on mitochondrial markers [429].
Post-hoc analysis of a Phase II clinical trial involving 252 patients (NCT01879241) suggested that 1 mg of rasagiline, in addition to riluzole (100 mg/day), could slow disease progression in patients with normal to fast disease progression. This effect was observed in function (ALSFRS-R decline at 6, 12, and 18 months) and survival (at 6 and 12 months) [430]. This study emphasizes the importance of stratifying disease progression rates in clinical trials, as the response to drugs and their disease-modifying effects may vary according to progression rate [430]. However, as the trial did not initially specify an analysis by progression rate, cautious interpretation is necessary. Future clinical trials are needed to establish rasagiline’s therapeutic potential in ALS, especially for normal- to fast-progressing forms.
In conclusion, rasagiline’s effects in ALS research reveal its complex potential as a therapeutic agent, with varying outcomes in clinical trials. Further research and careful patient stratification are essential to unlock its full therapeutic benefits in ALS treatment.

4. Conclusions

Throughout this review, we have delved into the complex interplay between mitochondrial dysfunction and the pathogenesis of ALS, a multifaceted disease characterized by the progressive degeneration of MNs and with elusive therapeutic options.
Mitochondria have emerged as a compelling convergent target for ALS treatment. Mitochondrial dysfunction, encompassing bioenergetic deficits, oxidative stress, impaired calcium handling, and disrupted dynamics, plays a pivotal role in the cascade of events that ultimately leads to motor neuron demise in ALS. The evidence gleaned from preclinical models and clinical trials underscores the significance of mitochondrial abnormalities in this devastating disease. As the intricate web of mechanisms underpinning ALS has unraveled, mitochondria have emerged as a nexus where various pathological processes converge. The potential of mitochondria as a therapeutic target in ALS is multifaceted and encompasses:
  • Restoring Energy Balance: Mitochondrial dysfunction in ALS often leads to a decline in cellular energy production, which is particularly detrimental to highly energy-demanding MNs. Strategies aimed at rejuvenating mitochondrial function and ensuring an adequate supply of ATP hold promise in sustaining motor neuron health and function.
  • Attenuating Oxidative Stress: Mitochondria are both sources and targets of oxidative stress in ALS. Therapies aimed at reducing mitochondrial-generated ROS and bolstering endogenous antioxidant defenses offer potential avenues for alleviating oxidative damage.
  • Regulating Calcium Homeostasis: Impaired calcium homeostasis in ALS contributes to mitochondrial dysfunction and subsequent neuronal death. Interventions that restore the calcium balance within neurons and mitochondria may mitigate excitotoxicity and improve neuronal survival.
  • Enhancing Mitochondrial Dynamics: Maintaining a healthy mitochondrial network through processes such as fission, fusion, and mitophagy is crucial for cellular health. Targeting these dynamic processes to ensure the timely removal of damaged mitochondria and the efficient distribution of healthy ones may have therapeutic implications.
  • Unveiling Genetic Insights: Genetic mutations associated with ALS, such as those in the SOD1, C9ORF72, and FUS genes, can directly impact mitochondrial function. Understanding the precise mechanisms by which these mutations affect mitochondria could pave the way for gene-specific therapies.
As we reflect on the current state of ALS research and treatment, it is evident that mitochondrial-targeted strategies hold significant potential. Several promising compounds, such as NAC, CoQ10, MitoQ, and other mitochondria-targeting antioxidants, have shown efficacy in preclinical models by mitigating mitochondrial dysfunction and delaying disease progression. These candidates provide a glimmer of hope for ALS patients and their families.

Funding

This work was financed by the European Regional Development Fund (ERDF) through the COMPETE 2020—Operational Programme for Competitiveness and Internationalization and Portuguese national funds via FCT–Fundação para a Ciência e a Tecnologia under projects PTDC/MED-FAR/29391/2017, POCI-01–0145-FEDER-029391, PTDC/BTM-SAL/29297/2017, POCI-01–0145-FEDER-029297, PTDC/BTM-ORG/0055/2021 (https://doi.org/10.54499/PTDC/BTM-ORG/0055/2021), https://doi.org/10.54499/DL57/2016/CP1448/CT0016 [TCO], CEECIND/00322/2017, and 2022.00011.CEECIND [EF], UIDP/04539/2020, UIDB/04539/2020, and UIDB/00081/2020.

Conflicts of Interest

The authors of this review paper disclose the following potential conflicts of interest: Teresa Cunha-Oliveira has received financial support in the form of a grant (PTDC/BTM-SAL/29297/2017, POCI-01–0145-FEDER-029297, https://app.dimensions.ai/details/grant/grant.9567304, accessed on 23 January 2024) and salary (DL57/2016/CP1448/CT0016) from Fundação para a Ciência e a Tecnologia (Portugal). Elisabete Ferreiro has received financial support in the form of a grant (PTDC/BTM-ORG/0055/2021, https://doi.org/10.54499/PTDC/BTM-ORG/0055/2021, accessed on 23 January 2024) and salary (CEECIND/00322/2017 and 2022.00011.CEECIND) from Fundação para a Ciência e a Tecnologia (Portugal). Filomena S.G. Silva has received financial support in the form of a grant and salary (PTDC/MED-FAR/29391/2017, POCI-01–0145-FEDER-029391, https://app.dimensions.ai/details/grant/grant.9567351, accessed on 23 January 2024). Additionally, F. Silva is currently employed by Mitotag Lda to work on the development of mitochondria-targeted antioxidants, which are discussed in this review. These funding sources and affiliations do not influence the objectivity and integrity of the content presented in this review. The authors are committed to providing a comprehensive and unbiased overview of the topic, and any potential conflicts of interest have been disclosed for transparency.

References

  1. Longinetti, E.; Fang, F. Epidemiology of Amyotrophic Lateral Sclerosis: An Update of Recent Literature. Curr. Opin. Neurol. 2019, 32, 771–776. [Google Scholar] [CrossRef]
  2. 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]
  3. Valdmanis, P.N.; Rouleau, G.A. Genetics of Familial Amyotrophic Lateral Sclerosis. Neurology 2008, 70, 144–152. [Google Scholar] [CrossRef]
  4. Zarei, S.; Carr, K.; Reiley, L.; Diaz, K.; Guerra, O.; Altamirano, P.; Pagani, W.; Lodin, D.; Orozco, G.; Chinea, A. A Comprehensive Review of Amyotrophic Lateral Sclerosis. Surg. Neurol. Int. 2015, 6, 171. [Google Scholar] [CrossRef]
  5. Boylan, K. Familial Amyotrophic Lateral Sclerosis. Neurol. Clin. 2015, 33, 807–830. [Google Scholar] [CrossRef]
  6. Kodavati, M.; Wang, H.; Hegde, M.L. Altered Mitochondrial Dynamics in Motor Neuron Disease: An Emerging Perspective. Cells 2020, 9, 1065. [Google Scholar] [CrossRef]
  7. DeJesus-Hernandez, M.; Mackenzie, I.R.; Boeve, B.F.; Boxer, A.L.; Baker, M.; Rutherford, N.J.; Nicholson, A.M.; Finch, N.A.; Flynn, H.; Adamson, J.; et al. Expanded GGGGCC Hexanucleotide Repeat in Noncoding Region of C9ORF72 Causes Chromosome 9p-Linked FTD and ALS. Neuron 2011, 72, 245–256. [Google Scholar] [CrossRef]
  8. Renton, A.E.; Majounie, E.; Waite, A.; Simon-Sanchez, 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]
  9. Deng, H.X.; Hentati, A.; Tainer, J.A.; Iqbal, Z.; Cayabyab, A.; Hung, W.Y.; Getzoff, E.D.; Hu, P.; Herzfeldt, B.; Roos, R.P.; et al. Amyotrophic Lateral Sclerosis and Structural Defects in Cu,Zn Superoxide Dismutase. Science 1993, 261, 1047–1051. [Google Scholar] [CrossRef]
  10. Kaur, S.J.; McKeown, S.R.; Rashid, S. Mutant SOD1 Mediated Pathogenesis of Amyotrophic Lateral Sclerosis. Gene 2016, 577, 109–118. [Google Scholar] [CrossRef]
  11. Rosen, D.R.; Siddique, T.; Patterson, D.; Figlewicz, D.A.; Sapp, P.; Hentati, A.; Donaldson, D.; Goto, J.; O’Regan, J.P.; Deng, H.-X.; et al. Mutations in Cu/Zn Superoxide Dismutase Gene Are Associated with Familial Amyotrophic Lateral Sclerosis. Nature 1993, 362, 59–62. [Google Scholar] [CrossRef]
  12. Synofzik, M.; Ronchi, D.; Keskin, I.; Basak, A.N.; Wilhelm, C.; Gobbi, C.; Birve, A.; Biskup, S.; Zecca, C.; Fernandez-Santiago, R.; et al. Mutant Superoxide Dismutase-1 Indistinguishable from Wild-Type Causes ALS. Hum. Mol. Genet. 2012, 21, 3568–3574. [Google Scholar] [CrossRef] [PubMed]
  13. Sreedharan, J.; Blair, I.P.; Tripathi, V.B.; Hu, X.; Vance, C.; Rogelj, B.; Ackerley, S.; Durnall, J.C.; Williams, K.L.; Buratti, E.; et al. TDP-43 Mutations in Familial and Sporadic Amyotrophic Lateral Sclerosis. Science 2008, 319, 1668–1672. [Google Scholar] [CrossRef] [PubMed]
  14. Kwiatkowski, T.J., Jr.; 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] [PubMed]
  15. 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]
  16. 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]
  17. 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]
  18. Yu, B.; Pamphlett, R. Environmental Insults: Critical Triggers for Amyotrophic Lateral Sclerosis. Transl. Neurodegener. 2017, 6, 15. [Google Scholar] [CrossRef]
  19. Ingre, C.; Roos, P.M.; Piehl, F.; Kamel, F.; Fang, F. Risk Factors for Amyotrophic Lateral Sclerosis. Clin. Epidemiol. 2015, 7, 181–193. [Google Scholar] [CrossRef]
  20. Tafuri, F.; Ronchi, D.; Magri, F.; Comi, G.P.; Corti, S. SOD1 Misplacing and Mitochondrial Dysfunction in Amyotrophic Lateral Sclerosis Pathogenesis. Front. Cell. Neurosci. 2015, 9, 336. [Google Scholar] [CrossRef]
  21. Bacman, S.R.; Bradley, W.G.; Moraes, C.T. Mitochondrial Involvement in Amyotrophic Lateral Sclerosis: Trigger or Target? Mol. Neurobiol. 2006, 33, 113–131. [Google Scholar] [CrossRef]
  22. Cunha-Oliveira, T.; Montezinho, L.; Mendes, C.; Firuzi, O.; Saso, L.; Oliveira, P.J.; Silva, F.S.G. Oxidative Stress in Amyotrophic Lateral Sclerosis: Pathophysiology and Opportunities for Pharmacological Intervention. Oxidative Med. Cell. Longev. 2020, 2020, 5021694. [Google Scholar] [CrossRef] [PubMed]
  23. Curti, D.; Malaspina, A.; Facchetti, G.; Camana, C.; Mazzini, L.; Tosca, P.; Zerbi, F.; Ceroni, M. Amyotrophic Lateral Sclerosis: Oxidative Energy Metabolism and Calcium Homeostasis in Peripheral Blood Lymphocytes. Neurology 1996, 47, 1060–1064. [Google Scholar] [CrossRef] [PubMed]
  24. 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]
  25. Vielhaber, S.; Winkler, K.; Kirches, E.; Kunz, D.; Buchner, 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]
  26. Nakano, Y.; Hirayama, K.; Terao, K. Hepatic Ultrastructural Changes and Liver Dysfunction in Amyotrophic Lateral Sclerosis. Arch. Neurol. 1987, 44, 103–106. [Google Scholar] [CrossRef]
  27. Palomo, G.M.; Manfredi, G. Exploring New Pathways of Neurodegeneration in ALS: The Role of Mitochondria Quality Control. Brain Res. 2015, 1607, 36–46. [Google Scholar] [CrossRef] [PubMed]
  28. Cozzolino, M.; Carri, M.T. Mitochondrial Dysfunction in ALS. Prog. Neurobiol. 2012, 97, 54–66. [Google Scholar] [CrossRef]
  29. Golpich, M.; Amini, E.; Mohamed, Z.; Azman Ali, R.; Mohamed Ibrahim, N.; Ahmadiani, A. Mitochondrial Dysfunction and Biogenesis in Neurodegenerative Diseases: Pathogenesis and Treatment. CNS Neurosci. Ther. 2017, 23, 5–22. [Google Scholar] [CrossRef]
  30. Vande Velde, C.; McDonald, K.K.; Boukhedimi, Y.; McAlonis-Downes, M.; Lobsiger, C.S.; Bel Hadj, S.; Zandona, A.; Julien, J.P.; Shah, S.B.; Cleveland, D.W. Misfolded SOD1 Associated with Motor Neuron Mitochondria Alters Mitochondrial Shape and Distribution Prior to Clinical Onset. PLoS ONE 2011, 6, e22031. [Google Scholar] [CrossRef] [PubMed]
  31. Liu, W.; Yamashita, T.; Tian, F.; Morimoto, N.; Ikeda, Y.; Deguchi, K.; Abe, K. Mitochondrial Fusion and Fission Proteins Expression Dynamically Change in a Murine Model of Amyotrophic Lateral Sclerosis. Curr. Neurovasc. Res. 2013, 10, 222–230. [Google Scholar] [CrossRef] [PubMed]
  32. Kong, J.; Xu, Z. Massive Mitochondrial Degeneration in Motor Neurons Triggers the Onset of Amyotrophic Lateral Sclerosis in Mice Expressing a Mutant SOD1. J. Neurosci. 1998, 18, 3241–3250. [Google Scholar] [CrossRef]
  33. Deng, J.; Yang, M.; Chen, Y.; Chen, X.; Liu, J.; Sun, S.; Cheng, H.; Li, Y.; Bigio, E.H.; Mesulam, M.; et al. FUS Interacts with HSP60 to Promote Mitochondrial Damage. PLoS Genet. 2015, 11, e1005357. [Google Scholar] [CrossRef]
  34. Stribl, C.; Samara, A.; Trumbach, D.; Peis, R.; Neumann, M.; Fuchs, H.; Gailus-Durner, V.; Hrabe de Angelis, M.; Rathkolb, B.; Wolf, E.; et al. Mitochondrial Dysfunction and Decrease in Body Weight of a Transgenic Knock-in Mouse Model for TDP-43. J. Biol. Chem. 2014, 289, 10769–10784. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, W.; Li, L.; Lin, W.L.; Dickson, D.W.; Petrucelli, L.; Zhang, T.; Wang, X. The ALS Disease-Associated Mutant TDP-43 Impairs Mitochondrial Dynamics and Function in Motor Neurons. Hum. Mol. Genet. 2013, 22, 4706–4719. [Google Scholar] [CrossRef]
  36. Choi, S.Y.; Lopez-Gonzalez, R.; Krishnan, G.; Phillips, H.L.; Li, A.N.; Seeley, W.W.; Yao, W.D.; Almeida, S.; Gao, F.B. C9ORF72-ALS/FTD-Associated Poly(GR) Binds Atp5a1 and Compromises Mitochondrial Function in Vivo. Nat. Neurosci. 2019, 22, 851–862. [Google Scholar] [CrossRef] [PubMed]
  37. Pasinelli, P.; Belford, M.E.; Lennon, N.; Bacskai, B.J.; Hyman, B.T.; Trotti, D.; Brown, R.H., Jr. Amyotrophic Lateral Sclerosis-Associated SOD1 Mutant Proteins Bind and Aggregate with Bcl-2 in Spinal Cord Mitochondria. Neuron 2004, 43, 19–30. [Google Scholar] [CrossRef]
  38. Israelson, A.; Arbel, N.; Da Cruz, S.; Ilieva, H.; Yamanaka, K.; Shoshan-Barmatz, V.; Cleveland, D.W. Misfolded Mutant SOD1 Directly Inhibits VDAC1 Conductance in a Mouse Model of Inherited ALS. Neuron 2010, 67, 575–587. [Google Scholar] [CrossRef]
  39. Kawamata, H.; Magrane, J.; Kunst, C.; King, M.P.; Manfredi, G. Lysyl-tRNA Synthetase is a Target for Mutant SOD1 Toxicity in Mitochondria. J. Biol. Chem. 2008, 283, 28321–28328. [Google Scholar] [CrossRef]
  40. Wang, W.; Wang, L.; Lu, J.; Siedlak, S.L.; Fujioka, H.; Liang, J.; Jiang, S.; Ma, X.; Jiang, Z.; da Rocha, E.L.; et al. The Inhibition of TDP-43 Mitochondrial Localization Blocks Its Neuronal Toxicity. Nat. Med. 2016, 22, 869–878. [Google Scholar] [CrossRef] [PubMed]
  41. Tsai, Y.L.; Coady, T.H.; Lu, L.; Zheng, D.; Alland, I.; Tian, B.; Shneider, N.A.; Manley, J.L. ALS/FTD-Associated Protein FUS Induces Mitochondrial Dysfunction by Preferentially Sequestering Respiratory Chain Complex mRNAs. Genes Dev. 2020, 34, 785–805. [Google Scholar] [CrossRef]
  42. Blokhuis, A.M.; Koppers, M.; Groen, E.J.N.; van den Heuvel, D.M.A.; Dini Modigliani, S.; Anink, J.J.; Fumoto, K.; van Diggelen, F.; Snelting, A.; Sodaar, P.; et al. Comparative Interactomics Analysis of Different ALS-Associated Proteins Identifies Converging Molecular Pathways. Acta Neuropathol. 2016, 132, 175–196. [Google Scholar] [CrossRef] [PubMed]
  43. Lopez-Gonzalez, R.; Lu, Y.; Gendron, T.F.; Karydas, A.; Tran, H.; Yang, D.; Petrucelli, L.; Miller, B.L.; Almeida, S.; Gao, F.B. Poly(GR) in C9ORF72-Related ALS/FTD Compromises Mitochondrial Function and Increases Oxidative Stress and DNA Damage in iPSC-Derived Motor Neurons. Neuron 2016, 92, 383–391. [Google Scholar] [CrossRef] [PubMed]
  44. Kausar, S.; Wang, F.; Cui, H. The Role of Mitochondria in Reactive Oxygen Species Generation and Its Implications for Neurodegenerative Diseases. Cells 2018, 7, 274. [Google Scholar] [CrossRef]
  45. Wiedemann, F.R.; Manfredi, G.; Mawrin, C.; Beal, M.F.; Schon, E.A. Mitochondrial DNA and respiratory chain function in spinal cords of ALS patients. J. Neurochem. 2002, 80, 616–625. [Google Scholar] [CrossRef]
  46. Browne, S.E.; Bowling, A.C.; Baik, M.J.; Gurney, M.; Brown, R.H., Jr.; Beal, M.F. Metabolic dysfunction in familial, but not sporadic, amyotrophic lateral sclerosis. J. Neurochem. 1998, 71, 281–287. [Google Scholar] [CrossRef]
  47. Vielhaber, S.; Kunz, D.; Winkler, K.; Wiedemann, F.R.; Kirches, E.; Feistner, H.; Heinze, H.J.; Elger, C.E.; Schubert, W.; Kunz, W.S. Mitochondrial DNA Abnormalities in Skeletal Muscle of Patients with Sporadic Amyotrophic Lateral Sclerosis. Brain 2000, 123 Pt 7, 1339–1348. [Google Scholar] [CrossRef]
  48. Ghiasi, P.; Hosseinkhani, S.; Noori, A.; Nafissi, S.; Khajeh, K. Mitochondrial complex I Deficiency and ATP/ADP Ratio in Lymphocytes of Amyotrophic Lateral Sclerosis Patients. Neurol. Res. 2012, 34, 297–303. [Google Scholar] [CrossRef]
  49. Kirk, K.; Gennings, C.; Hupf, J.C.; Tadesse, S.; D’Aurelio, M.; Kawamata, H.; Valsecchi, F.; Mitsumoto, H.; Groups, A.P.C.S.; Manfredi, G. Bioenergetic Markers in Skin Fibroblasts of Sporadic Amyotrophic Lateral Sclerosis and Progressive Lateral Sclerosis Patients. Ann. Neurol. 2014, 76, 620–624. [Google Scholar] [CrossRef]
  50. Konrad, C.; Kawamata, H.; Bredvik, K.G.; Arreguin, A.J.; Cajamarca, S.A.; Hupf, J.C.; Ravits, J.M.; Miller, T.M.; Maragakis, N.J.; Hales, C.M.; et al. Fibroblast Bioenergetics to Classify Amyotrophic Lateral Sclerosis Patients. Mol. Neurodegener. 2017, 12, 76. [Google Scholar] [CrossRef] [PubMed]
  51. Mattiazzi, M.; D’Aurelio, M.; Gajewski, C.D.; Martushova, K.; Kiaei, M.; Beal, M.F.; Manfredi, G. Mutated Human SOD1 Causes Dysfunction of Oxidative Phosphorylation in Mitochondria of Transgenic Mice. J. Biol. Chem. 2002, 277, 29626–29633. [Google Scholar] [CrossRef] [PubMed]
  52. Jung, C.; Higgins, C.M.; Xu, Z. Mitochondrial Electron Transport Chain Complex Dysfunction in a Transgenic Mouse Model for Amyotrophic Lateral Sclerosis. J. Neurochem. 2002, 83, 535–545. [Google Scholar] [CrossRef] [PubMed]
  53. Kirkinezos, I.G.; Bacman, S.R.; Hernandez, D.; Oca-Cossio, J.; Arias, L.J.; Perez-Pinzon, M.A.; Bradley, W.G.; Moraes, C.T. Cytochrome C Association with the Inner Mitochondrial Membrane is Impaired in the CNS of G93A-SOD1 Mice. J. Neurosci. 2005, 25, 164–172. [Google Scholar] [CrossRef]
  54. Menzies, F.M.; Cookson, M.R.; Taylor, R.W.; Turnbull, D.M.; Chrzanowska-Lightowlers, Z.M.; Dong, L.; Figlewicz, D.A.; Shaw, P.J. Mitochondrial Dysfunction in a Cell Culture Model of Familial Amyotrophic Lateral Sclerosis. Brain 2002, 125, 1522–1533. [Google Scholar] [CrossRef] [PubMed]
  55. Calabria, E.; Scambi, I.; Bonafede, R.; Schiaffino, L.; Peroni, D.; Potrich, V.; Capelli, C.; Schena, F.; Mariotti, R. ASCs-Exosomes Recover Coupling Efficiency and Mitochondrial Membrane Potential in an in Vitro Model of ALS. Front. Neurosci. 2019, 13, 1070. [Google Scholar] [CrossRef] [PubMed]
  56. Bowling, A.C.; Schulz, J.B.; Brown, R.H., Jr.; 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]
  57. Jaarsma, D.; Rognoni, F.; van Duijn, W.; Verspaget, H.W.; Haasdijk, E.D.; Holstege, J.C. CuZn Superoxide Dismutase (SOD1) Accumulates in Vacuolated Mitochondria in Transgenic Mice Expressing Amyotrophic Lateral Sclerosis-Linked SOD1 Mutations. Acta Neuropathol. 2001, 102, 293–305. [Google Scholar] [CrossRef]
  58. Higgins, C.M.J.; Jung, C.; Ding, H.; Xu, Z. Mutant Cu, Zn superoxide dismutase that causes motoneuron degeneration is present in mitochondria in the CNS. J. Neurosci. 2002, 22, RC215. [Google Scholar] [CrossRef]
  59. Liu, J.; Lillo, C.; Jonsson, P.A.; Vande Velde, C.; Ward, C.M.; Miller, T.M.; Subramaniam, J.R.; Rothstein, J.D.; Marklund, S.; Andersen, P.M.; et al. Toxicity of Familial ALS-Linked SOD1 Mutants from Selective Recruitment to Spinal Mitochondria. Neuron 2004, 43, 5–17. [Google Scholar] [CrossRef]
  60. Bergemalm, D.; Jonsson, P.A.; Graffmo, K.S.; Andersen, P.M.; Brannstrom, T.; Rehnmark, A.; Marklund, S.L. Overloading of Stable and Exclusion of Unstable Human Superoxide Dismutase-1 Variants in Mitochondria of Murine Amyotrophic Lateral Sclerosis Models. J. Neurosci. 2006, 26, 4147–4154. [Google Scholar] [CrossRef]
  61. Deng, H.X.; Shi, Y.; Furukawa, Y.; Zhai, H.; Fu, R.; Liu, E.; Gorrie, G.H.; Khan, M.S.; Hung, W.Y.; Bigio, E.H.; et al. Conversion to the Amyotrophic Lateral Sclerosis Phenotype is Associated with Intermolecular Linked Insoluble Aggregates of SOD1 in Mitochondria. Proc. Natl. Acad. Sci. USA 2006, 103, 7142–7147. [Google Scholar] [CrossRef]
  62. Vande Velde, C.; Miller, T.M.; Cashman, N.R.; Cleveland, D.W. Selective Association of Misfolded ALS-Linked Mutant SOD1 with the Cytoplasmic Face of Mitochondria. Proc. Natl. Acad. Sci. USA 2008, 105, 4022–4027. [Google Scholar] [CrossRef]
  63. Vijayvergiya, C.; Beal, M.F.; Buck, J.; Manfredi, G. Mutant Superoxide Dismutase 1 Forms Aggregates in the Brain Mitochondrial Matrix of Amyotrophic Lateral Sclerosis Mice. J. Neurosci. 2005, 25, 2463–2470. [Google Scholar] [CrossRef]
  64. Philips, T.; Rothstein, J.D. Rodent Models of Amyotrophic Lateral Sclerosis. Curr. Protoc. Pharmacol. 2015, 69, 5–67. [Google Scholar] [CrossRef]
  65. Ludolph, A.C.; Bendotti, C.; Blaugrund, E.; Chio, A.; Greensmith, L.; Loeffler, J.P.; Mead, R.; Niessen, H.G.; Petri, S.; Pradat, P.F.; et al. Guidelines for Preclinical Animal Research in ALS/MND: A Consensus Meeting. Amyotroph. Lateral Scler. 2010, 11, 38–45. [Google Scholar] [CrossRef]
  66. Kim, B.W.; Ryu, J.; Jeong, Y.E.; Kim, J.; Martin, L.J. Human Motor Neurons with SOD1-G93A Mutation Generated from CRISPR/Cas9 Gene-Edited iPSCs Develop Pathological Features of Amyotrophic Lateral Sclerosis. Front. Cell. Neurosci. 2020, 14, 604171. [Google Scholar] [CrossRef]
  67. Lu, J.; Duan, W.; Guo, Y.; Jiang, H.; Li, Z.; Huang, J.; Hong, K.; Li, C. Mitochondrial Dysfunction in Human TDP-43 Transfected NSC34 Cell Lines and the Protective Effect of Dimethoxy Curcumin. Brain Res. Bull. 2012, 89, 185–190. [Google Scholar] [CrossRef]
  68. Wang, M.D.; Little, J.; Gomes, J.; Cashman, N.R.; Krewski, D. Identification of Risk Factors Associated with Onset and Progression of Amyotrophic Lateral Sclerosis Using Systematic Review and Meta-Analysis. Neurotoxicology 2016, 61, 101–130. [Google Scholar] [CrossRef]
  69. Warraich, S.T.; Yang, S.; Nicholson, G.A.; Blair, I.P. TDP-43: A DNA and RNA Binding Protein with Roles in Neurodegenerative Diseases. Int. J. Biochem. Cell Biol. 2010, 42, 1606–1609. [Google Scholar] [CrossRef]
  70. Onesto, E.; Colombrita, C.; Gumina, V.; Borghi, M.O.; Dusi, S.; Doretti, A.; Fagiolari, G.; Invernizzi, F.; Moggio, M.; Tiranti, V.; et al. Gene-Specific Mitochondria Dysfunctions in Human TARDBP and C9ORF72 Fibroblasts. Acta Neuropathol. Commun. 2016, 4, 47. [Google Scholar] [CrossRef]
  71. Yamaguchi, A.; Takanashi, K. FUS Interacts with Nuclear Matrix-Associated Protein SAFB1 as Well as Matrin3 to Regulate Splicing and Ligand-Mediated Transcription. Sci. Rep. 2016, 6, 35195. [Google Scholar] [CrossRef]
  72. Lai, J.D.; Ichida, J.K. C9ORF72 Protein Function and Immune Dysregulation in Amyotrophic Lateral Sclerosis. Neurosci. Lett. 2019, 713, 134523. [Google Scholar] [CrossRef] [PubMed]
  73. Dafinca, R.; Scaber, J.; Ababneh, N.; Lalic, T.; Weir, G.; Christian, H.; Vowles, J.; Douglas, A.G.; Fletcher-Jones, A.; Browne, C.; et al. C9orf72 Hexanucleotide Expansions Are Associated with Altered Endoplasmic Reticulum Calcium Homeostasis and Stress Granule Formation in Induced Pluripotent Stem Cell-Derived Neurons from Patients with Amyotrophic Lateral Sclerosis and Frontotemporal Dementia. Stem Cells 2016, 34, 2063–2078. [Google Scholar] [CrossRef] [PubMed]
  74. Starkov, A.A. The Role of Mitochondria in Reactive Oxygen Species Metabolism and Signaling. Ann. N. Y. Acad. Sci. 2008, 1147, 37–52. [Google Scholar] [CrossRef] [PubMed]
  75. Sies, H.; Jones, D.P. Reactive Oxygen Species (ROS) as Pleiotropic Physiological Signalling Agents. Nat. Rev. Mol. Cell Biol. 2020, 21, 363–383. [Google Scholar] [CrossRef]
  76. Esra, B.; MuratUmit, S.; Cansin, S.; Serpil, E.; Omer, K. Oxidative Stress and Antioxidant Defense. World Allergy Organ. J. 2012, 5, 9–19. [Google Scholar]
  77. Barber, S.C.; Mead, R.J.; Shaw, P.J. Oxidative Stress in ALS: A Mechanism of Neurodegeneration and a Therapeutic Target. Biochim. Biophys. Acta 2006, 1762, 1051–1067. [Google Scholar] [CrossRef]
  78. Bodega, G.; Alique, M.; Puebla, L.; Carracedo, J.; Ramírez, R.M. Microvesicles: ROS scavengers and ROS producers. J. Extracell. Vesicles 2019, 8, 1626654. [Google Scholar] [CrossRef]
  79. Schieber, M.; Chandel, N.S. ROS Function in Redox Signaling and Oxidative Stress. Curr. Biol. 2014, 24, R453–R462. [Google Scholar] [CrossRef]
  80. Calingasan, N.Y.; Chen, J.; Kiaei, M.; Beal, M.F. Beta-Amyloid 42 Accumulation in the Lumbar Spinal Cord Motor Neurons of Amyotrophic Lateral Sclerosis Patients. Neurobiol. Dis. 2005, 19, 340–347. [Google Scholar] [CrossRef]
  81. Ferrante, R.J.; Browne, S.E.; Shinobu, L.A.; Bowling, A.C.; Baik, M.J.; MacGarvey, U.; Kowall, N.W.; Brown, R.H., Jr.; Beal, M.F. Evidence of Increased Oxidative Damage in Both Sporadic and Familial Amyotrophic Lateral Sclerosis. J. Neurochem. 1997, 69, 2064–2074. [Google Scholar] [CrossRef]
  82. 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]
  83. Simpson, E.P.; Henry, Y.K.; Henkel, J.S.; Smith, R.G.; Appel, S.H. Increased Lipid Peroxidation in Sera of ALS Patients: A Potential Biomarker of Disease Burden. Neurology 2004, 62, 1758–1765. [Google Scholar] [CrossRef]
  84. 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]
  85. Ihara, Y.; Nobukuni, K.; Takata, H.; Hayabara, T. Oxidative Stress and Metal Content in Blood and Cerebrospinal Fluid of Amyotrophic Lateral Sclerosis Patients with and without a Cu, Zn-Superoxide Dismutase Mutation. Neurol. Res. 2005, 27, 105–108. [Google Scholar] [CrossRef]
  86. Tohgi, H.; Abe, T.; Yamazaki, K.; Murata, T.; Ishizaki, E.; Isobe, C. Remarkable Increase in Cerebrospinal Fluid 3-Nitrotyrosine in Patients with Sporadic Amyotrophic Lateral Sclerosis. Ann. Neurol. 1999, 46, 129–131. [Google Scholar] [CrossRef]
  87. Weishaupt, J.H.; Bartels, C.; Polking, E.; Dietrich, J.; Rohde, G.; Poeggeler, B.; Mertens, N.; Sperling, S.; Bohn, M.; Huther, G.; et al. Reduced Oxidative Damage in ALS by High-Dose Enteral Melatonin Treatment. J. Pineal Res. 2006, 41, 313–323. [Google Scholar] [CrossRef]
  88. 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]
  89. Niebrój-Dobosz, I.; Dziewulska, D.; Kwieciński, H. Oxidative Damage to Proteins in the Spinal Cord in Amyotrophic Lateral Sclerosis (ALS). Folia Neuropathol. 2004, 42, 151–156. [Google Scholar]
  90. 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]
  91. Pedersen, W.A.; Fu, W.; Keller, J.N.; Markesbery, W.R.; Appel, S.; Smith, R.G.; Kasarskis, E.; Mattson, M.P. Protein Modification by the Lipid Peroxidation Product 4-Hydroxynonenal in the Spinal Cords of Amyotrophic Lateral Sclerosis Patients. Ann. Neurol. 1998, 44, 819–824. [Google Scholar] [CrossRef]
  92. Beal, M.F.; Ferrante, R.J.; Browne, S.E.; Matthews, R.T.; Kowall, N.W.; Brown, R.H., Jr. Increased 3-Nitrotyrosine in Both Sporadic and Familial Amyotrophic Lateral Sclerosis. Ann. Neurol. 1997, 42, 644–654. [Google Scholar] [CrossRef] [PubMed]
  93. Abe, K.; Pan, L.H.; Watanabe, M.; Konno, H.; Kato, T.; Itoyama, Y. Upregulation of Protein-Tyrosine Nitration in the Anterior Horn Cells of Amyotrophic Lateral Sclerosis. Neurol. Res. 1997, 19, 124–128. [Google Scholar] [CrossRef] [PubMed]
  94. Abe, K.; Pan, L.H.; Watanabe, M.; Kato, T.; Itoyama, Y. Induction of Nitrotyrosine-Like Immunoreactivity in the Lower Motor Neuron of Amyotrophic Lateral Sclerosis. Neurosci. Lett. 1995, 199, 152–154. [Google Scholar] [CrossRef] [PubMed]
  95. Babu, G.N.; Kumar, A.; Chandra, R.; Puri, S.K.; Singh, R.L.; Kalita, J.; Misra, U.K. Oxidant-Antioxidant Imbalance in the Erythrocytes of Sporadic Amyotrophic Lateral Sclerosis Patients Correlates with the Progression of Disease. Neurochem. Int. 2008, 52, 1284–1289. [Google Scholar] [CrossRef] [PubMed]
  96. Bozzo, F.; Mirra, A.; Carri, M.T. Oxidative Stress and Mitochondrial Damage in the Pathogenesis of ALS: New Perspectives. Neurosci. Lett. 2017, 636, 3–8. [Google Scholar] [CrossRef]
  97. Cunha-Oliveira, T.; Silva, D.F.; Segura, L.; Baldeiras, I.; Marques, R.; Rosenstock, T.; Oliveira, P.J.; Silva, F.S.G. Redox Profiles of Amyotrophic Lateral Sclerosis Lymphoblasts with or without Known SOD1 Mutations. Eur. J. Clin. Investig. 2022, 52, e13798. [Google Scholar] [CrossRef] [PubMed]
  98. Field, L.S.; Furukawa, Y.; O’Halloran, T.V.; Culotta, V.C. Factors Controlling the Uptake of Yeast Copper/Zinc Superoxide Dismutase into Mitochondria. J. Biol. Chem. 2003, 278, 28052–28059. [Google Scholar] [CrossRef]
  99. Kawamata, H.; Manfredi, G. Different Regulation of Wild-Type and Mutant Cu,Zn Superoxide Dismutase Localization in Mammalian Mitochondria. Hum. Mol. Genet. 2008, 17, 3303–3317. [Google Scholar] [CrossRef]
  100. Reddehase, S.; Grumbt, B.; Neupert, W.; Hell, K. The Disulfide Relay System of Mitochondria is Required for the Biogenesis of Mitochondrial CCS1 and SOD1. J. Mol. Biol. 2009, 385, 331–338. [Google Scholar] [CrossRef]
  101. Arciello, M.; Capo, C.R.; D’Annibale, S.; Cozzolino, M.; Ferri, A.; Carri, M.T.; Rossi, L. Copper Depletion Increases the Mitochondrial-Associated SOD1 in Neuronal Cells. Biometals 2011, 24, 269–278. [Google Scholar] [CrossRef]
  102. Menzies, F.M.; Ince, P.G.; Shaw, P.J. Mitochondrial Involvement in Amyotrophic Lateral Sclerosis. Neurochem. Int. 2002, 40, 543–551. [Google Scholar] [CrossRef]
  103. Beckman, J.S.; Estevez, A.G.; Crow, J.P.; Barbeito, L. Superoxide Dismutase and the Death of Motoneurons in ALS. Trends Neurosci. 2001, 24, S15–S20. [Google Scholar] [CrossRef]
  104. Estévez, A.G.; Crow, J.P.; Sampson, J.B.; Reiter, C.; Zhuang, Y.; Richardson, G.J.; Tarpey, M.M.; Barbeito, L.; Beckman, J.S. Induction of Nitric Oxide- Dependent Apoptosis in Motor Neurons by Zinc-Deficient Superoxide Dismutase. Science 1999, 286, 2498–2500. [Google Scholar] [CrossRef]
  105. Yim, M.B.; Kang, J.H.; Yim, H.S.; Kwak, H.S.; Chock, P.B.; Stadtman, E.R. A Gain-of-Function of an Amyotrophic Lateral Sclerosis-Associated Cu,Zn-Superoxide Dismutase Mutant: An Enhancement of Free Radical Formation Due to a Decrease in Km for Hydrogen Peroxide. Proc. Natl. Acad. Sci. USA 1996, 93, 5709–5714. [Google Scholar] [CrossRef] [PubMed]
  106. Liu, D. The roles of Free Radicals in Amyotrophic Lateral Sclerosis. J. Mol. Neurosci. 1996, 7, 159–167. [Google Scholar] [CrossRef]
  107. Beckman, J.S.; Carson, M.; Smith, C.D.; Koppenol, W.H. ALS, SOD and Peroxynitrite. Nature 1993, 364, 584. [Google Scholar] [CrossRef]
  108. Valentine, J.S.; Doucette, P.A.; Zittin Potter, S. Copper-Zinc Superoxide Dismutase and Amyotrophic Lateral Sclerosis. Annu. Rev. Biochem. 2005, 74, 563–593. [Google Scholar] [CrossRef] [PubMed]
  109. Benkler, C.; O’Neil, A.L.; Slepian, S.; Qian, F.; Weinreb, P.H.; Rubin, L.L. Aggregated SOD1 Causes Selective Death of Cultured Human Motor Neurons. Sci. Rep. 2018, 8, 16393. [Google Scholar] [CrossRef] [PubMed]
  110. Pickles, S.; Destroismaisons, L.; Peyrard, S.L.; Cadot, S.; Rouleau, G.A.; Brown, R.H., Jr.; Julien, J.P.; Arbour, N.; Vande Velde, C. Mitochondrial Damage Revealed by Immunoselection for ALS-Linked Misfolded SOD1. Hum. Mol. Genet. 2013, 22, 3947–3959. [Google Scholar] [CrossRef]
  111. Ferri, A.; Cozzolino, M.; Crosio, C.; Nencini, M.; Casciati, A.; Gralla, E.B.; Rotilio, G.; Valentine, J.S.; Carri, M.T. Familial ALS-Superoxide Dismutases Associate with Mitochondria and Shift Their Redox Potentials. Proc. Natl. Acad. Sci. USA 2006, 103, 13860–13865. [Google Scholar] [CrossRef] [PubMed]
  112. Killoy, K.M.; Harlan, B.A.; Pehar, M.; Helke, K.L.; Johnson, J.A.; Vargas, M.R. Decreased Glutathione Levels Cause Overt Motor Neuron Degeneration in hSOD1(WT) Over-Expressing Mice. Exp. Neurol. 2018, 302, 129–135. [Google Scholar] [CrossRef] [PubMed]
  113. Vargas, M.R.; Johnson, D.A.; Johnson, J.A. Decreased Glutathione Accelerates Neurological Deficit and Mitochondrial Pathology in Familial ALS-Linked hSOD1(G93A) Mice Model. Neurobiol. Dis. 2011, 43, 543–551. [Google Scholar] [CrossRef]
  114. Pesaresi, M.G.; Amori, I.; Giorgi, C.; Ferri, A.; Fiorenzo, P.; Gabanella, F.; Salvatore, A.M.; Giorgio, M.; Pelicci, P.G.; Pinton, P.; et al. Mitochondrial Redox Signalling by p66Shc Mediates ALS-Like Disease through Rac1 Inactivation. Hum. Mol. Genet. 2011, 20, 4196–4208. [Google Scholar] [CrossRef]
  115. Dikalov, S. Cross Talk between Mitochondria and NADPH Oxidases. Free. Radic. Biol. Med. 2011, 51, 1289–1301. [Google Scholar] [CrossRef]
  116. Wu, D.C.; Re, D.B.; Nagai, M.; Ischiropoulos, H.; Przedborski, S. The Inflammatory NADPH Oxidase Enzyme Modulates Motor Neuron Degeneration in Amyotrophic Lateral Sclerosis Mice. Proc. Natl. Acad. Sci. USA 2006, 103, 12132–12137. [Google Scholar] [CrossRef]
  117. Marden, J.J.; Harraz, M.M.; Williams, A.J.; Nelson, K.; Luo, M.; Paulson, H.; Engelhardt, J.F. Redox Modifier Genes in Amyotrophic Lateral Sclerosis in Mice. J. Clin. Investig. 2007, 117, 2913–2919. [Google Scholar] [CrossRef]
  118. Harraz, M.M.; Marden, J.J.; Zhou, W.; Zhang, Y.; Williams, A.; Sharov, V.S.; Nelson, K.; Luo, M.; Paulson, H.; Schöneich, C.; et al. SOD1 Mutations Disrupt Redox-Sensitive Rac Regulation of NADPH Oxidase in a Familial ALS Model. J. Clin. Investig. 2008, 118, 659–670. [Google Scholar] [CrossRef]
  119. Trumbull, K.A.; McAllister, D.; Gandelman, M.M.; Fung, W.Y.; Lew, T.; Brennan, L.; Lopez, N.; Morre, J.; Kalyanaraman, B.; Beckman, J.S. Diapocynin and Apocynin Administration Fails to Significantly Extend Survival in G93A SOD1 ALS Mice. Neurobiol. Dis. 2012, 45, 137–144. [Google Scholar] [CrossRef] [PubMed]
  120. Kato, S.; Kato, M.; Abe, Y.; Matsumura, T.; Nishino, T.; Aoki, M.; Itoyama, Y.; Asayama, K.; Awaya, A.; Hirano, A.; et al. Redox System Expression in the Motor Neurons in Amyotrophic Lateral Sclerosis (ALS): Immunohistochemical Studies on Sporadic ALS, Superoxide Dismutase 1 (SOD1)-Mutated Familial ALS, and SOD1-Mutated ALS Animal Models. Acta Neuropathol. 2005, 110, 101–112. [Google Scholar] [CrossRef] [PubMed]
  121. Thimmulappa, R.K.; Mai, K.H.; Srisuma, S.; Kensler, T.W.; Yamamoto, M.; Biswal, S. Identification of Nrf2-Regulated Genes Induced by the Chemopreventive Agent Sulforaphane by Oligonucleotide Microarray. Cancer Res. 2002, 62, 5196–5203. [Google Scholar]
  122. Lee, J.M.; Calkins, M.J.; Chan, K.; Kan, Y.W.; Johnson, J.A. Identification of the NF-E2-Related Factor-2-Dependent Genes Conferring Protection against Oxidative Stress in Primary Cortical Astrocytes Using Oligonucleotide Microarray Analysis. J. Biol. Chem. 2003, 278, 12029–12038. [Google Scholar] [CrossRef] [PubMed]
  123. Petri, S.; Korner, S.; Kiaei, M. Nrf2/ARE Signaling Pathway: Key Mediator in Oxidative Stress and Potential Therapeutic Target in ALS. Neurol. Res. Int. 2012, 2012, 878030. [Google Scholar] [CrossRef] [PubMed]
  124. Nioi, P.; McMahon, M.; Itoh, K.; Yamamoto, M.; Hayes, J.D. Identification of a Novel Nrf2-Regulated Antioxidant Response Element (ARE) in the Mouse NAD(P)H:Quinone Oxidoreductase 1 Gene: Reassessment of the ARE Consensus Sequence. Biochem. J. 2003, 374, 337–348. [Google Scholar] [CrossRef]
  125. Kwak, M.K.; Wakabayashi, N.; Greenlaw, J.L.; Yamamoto, M.; Kensler, T.W. Antioxidants enhance Mammalian Proteasome Expression through the Keap1-Nrf2 Signaling Pathway. Mol. Cell. Biol. 2003, 23, 8786–8794. [Google Scholar] [CrossRef] [PubMed]
  126. Kirby, J.; Halligan, E.; Baptista, M.J.; Allen, S.; Heath, P.R.; Holden, H.; Barber, S.C.; Loynes, C.A.; Wood-Allum, C.A.; Lunec, J.; et al. Mutant SOD1 Alters the Motor Neuronal Transcriptome: Implications for Familial ALS. Brain 2005, 128, 1686–1706. [Google Scholar] [CrossRef] [PubMed]
  127. Sarlette, A.; Krampfl, K.; Grothe, C.; Neuhoff, N.; Dengler, R.; Petri, S. Nuclear Erythroid 2-Related Factor 2-Antioxidative Response Element Signaling Pathway in Motor Cortex and Spinal Cord in Amyotrophic Lateral Sclerosis. J. Neuropathol. Exp. Neurol. 2008, 67, 1055–1062. [Google Scholar] [CrossRef] [PubMed]
  128. Kraft, A.D.; Resch, J.M.; Johnson, D.A.; Johnson, J.A. Activation of the Nrf2-ARE Pathway in Muscle and Spinal Cord during ALS-Like Pathology in Mice Expressing Mutant SOD1. Exp. Neurol. 2007, 207, 107–117. [Google Scholar] [CrossRef]
  129. Pehar, M.; Beeson, G.; Beeson, C.C.; Johnson, J.A.; Vargas, M.R. Mitochondria-Targeted Catalase Reverts the Neurotoxicity of hSOD1G(9)(3)A Astrocytes without Extending the Survival of ALS-Linked Mutant hSOD1 Mice. PLoS ONE 2014, 9, e103438. [Google Scholar] [CrossRef]
  130. Tian, Y.P.; Che, F.Y.; Su, Q.P.; Lu, Y.C.; You, C.P.; Huang, L.M.; Wang, S.G.; Wang, L.; Yu, J.X. Effects of Mutant TDP-43 on the Nrf2/ARE Pathway and Protein Expression of MafK and JDP2 in NSC-34 Cells. Genet. Mol. Res. 2017, 16, 1–13. [Google Scholar] [CrossRef]
  131. Moujalled, D.; Grubman, A.; Acevedo, K.; Yang, S.; Ke, Y.D.; Moujalled, D.M.; Duncan, C.; Caragounis, A.; Perera, N.D.; Turner, B.J.; et al. TDP-43 Mutations Causing Amyotrophic Lateral Sclerosis Are Associated with Altered Expression of RNA-Binding Protein hnRNP K and Affect the Nrf2 Antioxidant Pathway. Hum. Mol. Genet. 2017, 26, 1732–1746. [Google Scholar] [CrossRef] [PubMed]
  132. Duan, W.; Li, X.; Shi, J.; Guo, Y.; Li, Z.; Li, C. Mutant TAR DNA-Binding Protein-43 Induces Oxidative Injury in Motor Neuron-Like Cell. Neuroscience 2010, 169, 1621–1629. [Google Scholar] [CrossRef] [PubMed]
  133. Birger, A.; Ben-Dor, I.; Ottolenghi, M.; Turetsky, T.; Gil, Y.; Sweetat, S.; Perez, L.; Belzer, V.; Casden, N.; Steiner, D.; et al. Human iPSC-Derived Astrocytes from ALS Patients with Mutated C9ORF72 Show Increased Oxidative Stress and Neurotoxicity. EBioMedicine 2019, 50, 274–289. [Google Scholar] [CrossRef] [PubMed]
  134. Bono, S.; Feligioni, M.; Corbo, M. Impaired Antioxidant KEAP1-NRF2 System in Amyotrophic Lateral Sclerosis: NRF2 Activation as a Potential Therapeutic Strategy. Mol. Neurodegener. 2021, 16, 71. [Google Scholar] [CrossRef] [PubMed]
  135. 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] [PubMed]
  136. Dupuis, L.; Oudart, H.; Rene, F.; Gonzalez de Aguilar, J.L.; 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]
  137. Fergani, A.; Oudart, H.; Gonzalez De Aguilar, J.L.; Fricker, B.; Rene, F.; Hocquette, J.F.; Meininger, V.; Dupuis, L.; Loeffler, J.P. Increased Peripheral Lipid Clearance in an Animal Model of Amyotrophic Lateral Sclerosis. J. Lipid Res. 2007, 48, 1571–1580. [Google Scholar] [CrossRef]
  138. Chiang, P.M.; Ling, J.; Jeong, Y.H.; Price, D.L.; Aja, S.M.; Wong, P.C. Deletion of Tdp-43 Down-Regulates Tbc1d1, a Gene Linked to Obesity, and Alters Body Fat Metabolism. Proc. Natl. Acad. Sci. USA 2010, 107, 16320–16324. [Google Scholar] [CrossRef]
  139. Coughlan, K.S.; Halang, L.; Woods, I.; Prehn, J.H. A High-Fat Jelly Diet Restores Bioenergetic Balance and Extends Lifespan in the Presence of Motor Dysfunction and Lumbar Spinal Cord Motor Neuron Loss in TDP-43A315T Mutant C57BL6/J Mice. Dis. Models Mech. 2016, 9, 1029–1037. [Google Scholar] [CrossRef]
  140. Manzo, E.; Lorenzini, I.; Barrameda, D.; O’Conner, A.G.; Barrows, J.M.; Starr, A.; Kovalik, T.; Rabichow, B.E.; Lehmkuhl, E.M.; Shreiner, D.D.; et al. Glycolysis Upregulation Is Neuroprotective as a Compensatory Mechanism in ALS. eLife 2019, 8, e45114. [Google Scholar] [CrossRef]
  141. Schmitt, F.; Hussain, G.; Dupuis, L.; Loeffler, J.P.; Henriques, A. A Plural Role for Lipids in Motor Neuron Diseases: Energy, Signaling and Structure. Front. Cell. Neurosci. 2014, 8, 25. [Google Scholar] [CrossRef] [PubMed]
  142. 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, 592851. [Google Scholar] [CrossRef]
  143. Park, Y.; Park, J.; Kim, Y.; Baek, H.; Kim, S.H. Association between Nutritional Status and Disease Severity Using the Amyotrophic Lateral Sclerosis (ALS) Functional Rating Scale in ALS Patients. Nutrition 2015, 31, 1362–1367. [Google Scholar] [CrossRef]
  144. Desport, J.C.; Preux, P.M.; Truong, C.T.; Courat, L.; Vallat, J.M.; Couratier, P. Nutritional Assessment and Survival in ALS Patients. Amyotroph. Lateral Scler. Other Mot. Neuron Disord. 2000, 1, 91–96. [Google Scholar] [CrossRef]
  145. 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, 1059–1063. [Google Scholar] [CrossRef] [PubMed]
  146. 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]
  147. Kasarskis, E.J.; Neville, H.E. Management of ALS: Nutritional care. Neurology 1996, 47, S118–S120. [Google Scholar] [CrossRef]
  148. 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]
  149. 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]
  150. Desport, J.C.; Preux, P.M.; Magy, L.; Boirie, Y.; Vallat, J.M.; Beaufrere, B.; Couratier, P. Factors Correlated with Hypermetabolism in Patients with Amyotrophic Lateral Sclerosis. Am. J. Clin. Nutr. 2001, 74, 328–334. [Google Scholar] [CrossRef]
  151. 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]
  152. Genton, L.; Viatte, V.; Janssens, J.P.; Heritier, A.C.; Pichard, C. Nutritional State, Energy Intakes and Energy Expenditure of Amyotrophic Lateral Sclerosis (ALS) Patients. Clin. Nutr. 2011, 30, 553–559. [Google Scholar] [CrossRef]
  153. Gallo, V.; Wark, P.A.; Jenab, M.; Pearce, N.; Brayne, C.; Vermeulen, R.; Andersen, P.M.; Hallmans, G.; Kyrozis, A.; Vanacore, N.; et al. Prediagnostic Body Fat and Risk of Death from Amyotrophic Lateral Sclerosis: The EPIC Cohort. Neurology 2013, 80, 829–838. [Google Scholar] [CrossRef]
  154. O’Reilly, E.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]
  155. 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]
  156. 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] [PubMed]
  157. Vaisman, N.; Lusaus, M.; Nefussy, B.; Niv, E.; Comaneshter, D.; Hallack, R.; Drory, V.E. Do Patients with Amyotrophic Lateral Sclerosis (ALS) Have Increased Energy Needs? J. Neurol. Sci. 2009, 279, 26–29. [Google Scholar] [CrossRef] [PubMed]
  158. Jesus, 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] [PubMed]
  159. 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] [PubMed]
  160. Harris, J.A.; Benedict, F.G. A Biometric Study of Human Basal Metabolism. Proc. Natl. Acad. Sci. USA 1918, 4, 370–373. [Google Scholar] [CrossRef] [PubMed]
  161. Sherman, M.S.; Pillai, A.; Jackson, A.; Heiman-Patterson, T. Standard Equations Are Not Accurate in Assessing Resting Energy Expenditure in Patients with Amyotrophic Lateral Sclerosis. JPEN J. Parenter. Enter. Nutr. 2004, 28, 442–446. [Google Scholar] [CrossRef]
  162. Brito, M.D.; da Silva, G.F.G.; Tilieri, E.M.; Araujo, B.G.; Calio, M.L.; Rosenstock, T.R. Metabolic Alteration and Amyotrophic Lateral Sclerosis Outcome: A Systematic Review. Front. Neurol. 2019, 10, 1205. [Google Scholar] [CrossRef] [PubMed]
  163. Pharaoh, G.; Sataranatarajan, K.; Street, K.; Hill, S.; Gregston, J.; Ahn, B.; Kinter, C.; Kinter, M.; Van Remmen, H. Metabolic and Stress Response Changes Precede Disease Onset in the Spinal Cord of Mutant SOD1 ALS Mice. Front. Neurosci. 2019, 13, 487. [Google Scholar] [CrossRef] [PubMed]
  164. Perluigi, M.; Fai Poon, H.; Hensley, K.; Pierce, W.M.; Klein, J.B.; Calabrese, V.; De Marco, C.; Butterfield, D.A. Proteomic Analysis of 4-Hydroxy-2-Nonenal-Modified Proteins in G93A-SOD1 Transgenic Mice—A Model of Familial Amyotrophic Lateral Sclerosis. Free. Radic. Biol. Med. 2005, 38, 960–968. [Google Scholar] [CrossRef] [PubMed]
  165. Poon, H.F.; Hensley, K.; Thongboonkerd, V.; Merchant, M.L.; Lynn, B.C.; Pierce, W.M.; Klein, J.B.; Calabrese, V.; Butterfield, D.A. Redox Proteomics Analysis of Oxidatively Modified Proteins in G93A-SOD1 Transgenic Mice—A Model of Familial Amyotrophic Lateral Sclerosis. Free Radic. Biol. Med. 2005, 39, 453–462. [Google Scholar] [CrossRef] [PubMed]
  166. Lukas, T.J.; Luo, W.W.; Mao, H.; Cole, N.; Siddique, T. Informatics-Assisted Protein Profiling in a Transgenic Mouse Model of Amyotrophic Lateral Sclerosis. Mol. Cell. Proteom. 2006, 5, 1233–1244. [Google Scholar] [CrossRef] [PubMed]
  167. Massignan, T.; Casoni, F.; Basso, M.; Stefanazzi, P.; Biasini, E.; Tortarolo, M.; Salmona, M.; Gianazza, E.; Bendotti, C.; Bonetto, V. Proteomic Analysis of Spinal Cord of Presymptomatic Amyotrophic Lateral Sclerosis G93A SOD1 Mouse. Biochem. Biophys. Res. Commun. 2007, 353, 719–725. [Google Scholar] [CrossRef]
  168. Li, Q.; Vande Velde, C.; Israelson, A.; Xie, J.; Bailey, A.O.; Dong, M.Q.; Chun, S.J.; Roy, T.; Winer, L.; Yates, J.R.; et al. ALS-Linked Mutant Superoxide Dismutase 1 (SOD1) Alters Mitochondrial Protein Composition and Decreases Protein Import. Proc. Natl. Acad. Sci. USA 2010, 107, 21146–21151. [Google Scholar] [CrossRef]
  169. Valbuena, G.N.; Rizzardini, M.; Cimini, S.; Siskos, A.P.; Bendotti, C.; Cantoni, L.; Keun, H.C. Metabolomic Analysis Reveals Increased Aerobic Glycolysis and Amino Acid Deficit in a Cellular Model of Amyotrophic Lateral Sclerosis. Mol. Neurobiol. 2016, 53, 2222–2240. [Google Scholar] [CrossRef]
  170. Miyazaki, K.; Masamoto, K.; Morimoto, N.; Kurata, T.; Mimoto, T.; Obata, T.; Kanno, I.; Abe, K. Early and Progressive Impairment of Spinal Blood Flow-Glucose Metabolism Coupling in Motor Neuron Degeneration of ALS Model Mice. J. Cereb. Blood Flow Metab. 2012, 32, 456–467. [Google Scholar] [CrossRef]
  171. Belanger, M.; Allaman, I.; Magistretti, P.J. Brain Energy Metabolism: Focus on Astrocyte-Neuron Metabolic Cooperation. Cell Metab. 2011, 14, 724–738. [Google Scholar] [CrossRef]
  172. Sharp, P.S.; Dick, J.R.; Greensmith, L. The Effect of Peripheral Nerve Injury on Disease Progression in the SOD1(G93A) Mouse Model of Amyotrophic Lateral Sclerosis. Neuroscience 2005, 130, 897–910. [Google Scholar] [CrossRef]
  173. Palamiuc, L.; Schlagowski, A.; Ngo, S.T.; Vernay, A.; Dirrig-Grosch, S.; Henriques, A.; Boutillier, A.L.; Zoll, J.; Echaniz-Laguna, A.; Loeffler, J.P.; 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] [PubMed]
  174. Cunha-Oliveira, T.; Carvalho, M.; Sardao, V.; Ferreiro, E.; Mena, D.; Pereira, F.B.; Borges, F.; Oliveira, P.J.; Silva, F.S.G. Integrative Profiling of Amyotrophic Lateral Sclerosis Lymphoblasts Identifies Unique Metabolic and Mitochondrial Disease Fingerprints. Mol. Neurobiol. 2022, 59, 6373–6396. [Google Scholar] [CrossRef] [PubMed]
  175. Guzman, M.; Blazquez, C. Ketone Body Synthesis in the Brain: Possible Neuroprotective Effects. Prostaglandins Leukot. Essent. Fat. Acids 2004, 70, 287–292. [Google Scholar] [CrossRef] [PubMed]
  176. LaManna, J.C.; Salem, N.; Puchowicz, M.; Erokwu, B.; Koppaka, S.; Flask, C.; Lee, Z. Ketones Suppress Brain Glucose Consumption. Adv. Exp. Med. Biol. 2009, 645, 301–306. [Google Scholar] [CrossRef]
  177. Yi, C.X.; Habegger, K.M.; Chowen, J.A.; Stern, J.; Tschop, M.H. A Role for Astrocytes in the Central Control of Metabolism. Neuroendocrinology 2011, 93, 143–149. [Google Scholar] [CrossRef]
  178. Le Foll, C.; Levin, B.E. Fatty Acid-Induced Astrocyte Ketone Production and the Control of Food Intake. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2016, 310, R1186–R1192. [Google Scholar] [CrossRef] [PubMed]
  179. Iglesias, J.; Morales, L.; Barreto, G.E. Metabolic and Inflammatory Adaptation of Reactive Astrocytes: Role of PPARs. Mol. Neurobiol. 2017, 54, 2518–2538. [Google Scholar] [CrossRef]
  180. Neumann, M.; Sampathu, D.M.; Kwong, L.K.; Truax, A.C.; Micsenyi, M.C.; Chou, T.T.; Bruce, J.; Schuck, T.; Grossman, M.; Clark, C.M.; et al. Ubiquitinated TDP-43 in Frontotemporal Lobar Degeneration and Amyotrophic Lateral Sclerosis. Science 2006, 314, 130–133. [Google Scholar] [CrossRef]
  181. 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]
  182. Lawton, K.A.; Brown, M.V.; Alexander, D.; Li, Z.; Wulff, J.E.; Lawson, R.; Jaffa, M.; Milburn, M.V.; Ryals, J.A.; Bowser, R.; et al. Plasma Metabolomic Biomarker Panel to Distinguish Patients with Amyotrophic Lateral Sclerosis from Disease Mimics. Amyotroph. Lateral Scler. Front. Degener. 2014, 15, 362–370. [Google Scholar] [CrossRef]
  183. Lawton, K.A.; Cudkowicz, M.E.; Brown, M.V.; Alexander, D.; Caffrey, R.; Wulff, J.E.; Bowser, R.; Lawson, R.; Jaffa, M.; Milburn, M.V.; et al. Biochemical Alterations Associated with ALS. Amyotroph. Lateral Scler. 2012, 13, 110–118. [Google Scholar] [CrossRef]
  184. Allen, S.P.; Hall, B.; Woof, R.; Francis, L.; Gatto, N.; Shaw, A.C.; Myszczynska, M.; Hemingway, J.; Coldicott, I.; Willcock, A.; et al. C9orf72 Expansion within Astrocytes Reduces Metabolic Flexibility in Amyotrophic Lateral Sclerosis. Brain 2019, 142, 3771–3790. [Google Scholar] [CrossRef] [PubMed]
  185. Blasco, H.; Corcia, P.; Moreau, C.; Veau, S.; Fournier, C.; Vourc’h, P.; Emond, P.; Gordon, P.; Pradat, P.F.; Praline, J.; et al. 1H-NMR-Based Metabolomic Profiling of CSF in Early Amyotrophic Lateral Sclerosis. PLoS ONE 2010, 5, e13223. [Google Scholar] [CrossRef]
  186. Kumar, A.; Bala, L.; Kalita, J.; Misra, U.K.; Singh, R.L.; Khetrapal, C.L.; Babu, G.N. Metabolomic Analysis of Serum by (1) H NMR Spectroscopy in Amyotrophic Lateral Sclerosis. Clin. Chim. Acta 2010, 411, 563–567. [Google Scholar] [CrossRef] [PubMed]
  187. Wuolikainen, A.; Andersen, P.M.; Moritz, T.; Marklund, S.L.; Antti, H. ALS Patients with Mutations in the SOD1 Gene Have an Unique Metabolomic Profile in the Cerebrospinal Fluid Compared with ALS Patients without Mutations. Mol. Genet. Metab. 2012, 105, 472–478. [Google Scholar] [CrossRef] [PubMed]
  188. Wuolikainen, A.; Moritz, T.; Marklund, S.L.; Antti, H.; Andersen, P.M. Disease-Related Changes in the Cerebrospinal Fluid Metabolome in Amyotrophic Lateral Sclerosis Detected by GC/TOFMS. PLoS ONE 2011, 6, e17947. [Google Scholar] [CrossRef] [PubMed]
  189. Gray, E.; Larkin, J.R.; Claridge, T.D.; Talbot, K.; Sibson, N.R.; Turner, M.R. The Longitudinal Cerebrospinal Fluid Metabolomic Profile of Amyotrophic Lateral Sclerosis. Amyotroph. Lateral Scler. Front. Degener. 2015, 16, 456–463. [Google Scholar] [CrossRef] [PubMed]
  190. Blasco, H.; Nadal-Desbarats, L.; Pradat, P.F.; Gordon, P.H.; Madji Hounoum, B.; Patin, F.; Veyrat-Durebex, C.; Mavel, S.; Beltran, S.; Emond, P.; et al. Biomarkers in Amyotrophic Lateral Sclerosis: Combining Metabolomic and Clinical Parameters to Define Disease Progression. Eur. J. Neurol. 2016, 23, 346–353. [Google Scholar] [CrossRef] [PubMed]
  191. Blasco, H.; Patin, F.; Descat, A.; Garcon, G.; Corcia, P.; Gele, P.; Lenglet, T.; Bede, P.; Meininger, V.; Devos, D.; et al. A Pharmaco-Metabolomics Approach in a Clinical Trial of ALS: Identification of Predictive Markers of Progression. PLoS ONE 2018, 13, e0198116. [Google Scholar] [CrossRef]
  192. Patin, F.; Baranek, T.; Vourc’h, P.; Nadal-Desbarats, L.; Goossens, J.F.; Marouillat, S.; Dessein, A.F.; Descat, A.; Hounoum, B.M.; Bruno, C.; et al. Combined Metabolomics and Transcriptomics Approaches to Assess the IL-6 Blockade as a Therapeutic of ALS: Deleterious Alteration of Lipid Metabolism. Neurotherapeutics 2016, 13, 905–917. [Google Scholar] [CrossRef]
  193. Palikaras, K.; Lionaki, E.; Tavernarakis, N. Balancing Mitochondrial Biogenesis and Mitophagy to Maintain Energy Metabolism Homeostasis. Cell Death Differ. 2015, 22, 1399–1401. [Google Scholar] [CrossRef]
  194. Nakada, K.; Inoue, K.; Ono, T.; Isobe, K.; Ogura, A.; Goto, Y.I.; Nonaka, I.; Hayashi, J.I. Inter-Mitochondrial Complementation: Mitochondria-Specific System Preventing Mice from Expression of Disease Phenotypes by Mutant mtDNA. Nat. Med. 2001, 7, 934–940. [Google Scholar] [CrossRef]
  195. Chen, H.; Chomyn, A.; Chan, D.C. Disruption of Fusion Results in Mitochondrial Heterogeneity and Dysfunction. J. Biol. Chem. 2005, 280, 26185–26192. [Google Scholar] [CrossRef]
  196. Khalil, B.; Lievens, J.C. Mitochondrial Quality Control in Amyotrophic Lateral Sclerosis: Towards a Common Pathway? Neural Regen. Res. 2017, 12, 1052–1061. [Google Scholar] [CrossRef]
  197. Smirnova, E.; Griparic, L.; Shurland, D.L.; van der Bliek, A.M. Dynamin-Related Protein Drp1 is Required for Mitochondrial Division in Mammalian Cells. Mol. Biol. Cell 2001, 12, 2245–2256. [Google Scholar] [CrossRef]
  198. Tamura, Y.; Itoh, K.; Sesaki, H. SnapShot: Mitochondrial Dynamics. Cell 2011, 145, 1158.e1. [Google Scholar] [CrossRef] [PubMed]
  199. Otera, H.; Wang, C.; Cleland, M.M.; Setoguchi, K.; Yokota, S.; Youle, R.J.; Mihara, K. Mff is an Essential Factor for Mitochondrial Recruitment of Drp1 During Mitochondrial Fission in Mammalian Cells. J. Cell Biol. 2010, 191, 1141–1158. [Google Scholar] [CrossRef] [PubMed]
  200. James, D.I.; Parone, P.A.; Mattenberger, Y.; Martinou, J.C. Hfis1, a Novel Component of the Mammalian Mitochondrial Fission Machinery. J. Biol. Chem. 2003, 278, 36373–36379. [Google Scholar] [CrossRef] [PubMed]
  201. Palmer, C.S.; Osellame, L.D.; Laine, D.; Koutsopoulos, O.S.; Frazier, A.E.; Ryan, M.T. MiD49 and MiD51, New Components of the Mitochondrial Fission Machinery. EMBO Rep. 2011, 12, 565–573. [Google Scholar] [CrossRef]
  202. Mears, J.A.; Lackner, L.L.; Fang, S.; Ingerman, E.; Nunnari, J.; Hinshaw, J.E. Conformational Changes in Dnm1 Support a Contractile Mechanism for Mitochondrial Fission. Nat. Struct. Mol. Biol. 2011, 18, 20–26. [Google Scholar] [CrossRef] [PubMed]
  203. Fukushima, N.H.; Brisch, E.; Keegan, B.R.; Bleazard, W.; Shaw, J.M. The GTPase Effector Domain Sequence of the Dnm1p GTPase Regulates Self-Assembly and Controls a Rate-Limiting Step in Mitochondrial Fission. Mol. Biol. Cell 2001, 12, 2756–2766. [Google Scholar] [CrossRef] [PubMed]
  204. Santel, A.; Fuller, M.T. Control of Mitochondrial Morphology by a Human Mitofusin. J. Cell Sci. 2001, 114 Pt 5, 867–874. [Google Scholar] [CrossRef] [PubMed]
  205. Chen, H.; Chan, D.C. Physiological Functions of Mitochondrial Fusion. Ann. N. Y. Acad. Sci. 2010, 1201, 21–25. [Google Scholar] [CrossRef] [PubMed]
  206. Song, Z.; Ghochani, M.; McCaffery, J.M.; Frey, T.G.; Chan, D.C. Mitofusins and OPA1 Mediate Sequential Steps in Mitochondrial Membrane Fusion. Mol. Biol. Cell 2009, 20, 3525–3532. [Google Scholar] [CrossRef]
  207. Cipolat, S.; Martins de Brito, O.; Dal Zilio, B.; Scorrano, L. OPA1 Requires Mitofusin 1 to Promote Mitochondrial Fusion. Proc Natl. Acad. Sci. USA 2004, 101, 15927–15932. [Google Scholar] [CrossRef] [PubMed]
  208. Chen, H.; McCaffery, J.M.; Chan, D.C. Mitochondrial Fusion Protects against Neurodegeneration in the Cerebellum. Cell 2007, 130, 548–562. [Google Scholar] [CrossRef]
  209. Ishihara, N.; Nomura, M.; Jofuku, A.; Kato, H.; Suzuki, S.O.; Masuda, K.; Otera, H.; Nakanishi, Y.; Nonaka, I.; Goto, Y.; et al. Mitochondrial Fission Factor Drp1 is Essential for Embryonic Development and Synapse Formation in Mice. Nat. Cell Biol. 2009, 11, 958–966. [Google Scholar] [CrossRef]
  210. Wakabayashi, J.; Zhang, Z.; Wakabayashi, N.; Tamura, Y.; Fukaya, M.; Kensler, T.W.; Iijima, M.; Sesaki, H. The Dynamin-Related GTPase Drp1 is Required for Embryonic and Brain Development in Mice. J. Cell Biol. 2009, 186, 805–816. [Google Scholar] [CrossRef]
  211. Delettre, C.; Lenaers, G.; Griffoin, J.M.; Gigarel, N.; Lorenzo, C.; Belenguer, P.; Pelloquin, L.; Grosgeorge, J.; Turc-Carel, C.; Perret, E.; et al. Nuclear Gene OPA1, Encoding a Mitochondrial Dynamin-Related Protein, is Mutated in Dominant Optic Atrophy. Nat. Genet. 2000, 26, 207–210. [Google Scholar] [CrossRef]
  212. Song, W.; Song, Y.; Kincaid, B.; Bossy, B.; Bossy-Wetzel, E. Mutant SOD1G93A Triggers Mitochondrial Fragmentation in Spinal Cord Motor Neurons: Neuroprotection by SIRT3 and PGC-1alpha. Neurobiol. Dis. 2013, 51, 72–81. [Google Scholar] [CrossRef]
  213. Xu, Y.F.; Gendron, T.F.; Zhang, Y.J.; Lin, W.L.; D’Alton, S.; Sheng, H.; Casey, M.C.; Tong, J.; Knight, J.; Yu, X.; et al. Wild-Type Human TDP-43 Expression Causes TDP-43 Phosphorylation, Mitochondrial Aggregation, Motor Deficits, and Early Mortality in Transgenic Mice. J. Neurosci. 2010, 30, 10851–10859. [Google Scholar] [CrossRef] [PubMed]
  214. Russell, A.P.; Wada, S.; Vergani, L.; Hock, M.B.; Lamon, S.; Leger, B.; Ushida, T.; Cartoni, R.; Wadley, G.D.; Hespel, P.; et al. Disruption of Skeletal Muscle Mitochondrial Network Genes and miRNAs in Amyotrophic Lateral Sclerosis. Neurobiol. Dis. 2013, 49, 107–117. [Google Scholar] [CrossRef] [PubMed]
  215. Vinsant, S.; Mansfield, C.; Jimenez-Moreno, R.; Del Gaizo Moore, V.; 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]
  216. Shutt, T.; Geoffrion, M.; Milne, R.; McBride, H.M. The Intracellular Redox State Is a Core Determinant of Mitochondrial Fusion. EMBO Rep. 2012, 13, 909–915. [Google Scholar] [CrossRef]
  217. Kim, Y.M.; Youn, S.W.; Sudhahar, V.; Das, A.; Chandhri, R.; Cuervo Grajal, H.; Kweon, J.; Leanhart, S.; He, L.; Toth, P.T.; et al. Redox Regulation of Mitochondrial Fission Protein Drp1 by Protein Disulfide Isomerase Limits Endothelial Senescence. Cell Rep. 2018, 23, 3565–3578. [Google Scholar] [CrossRef] [PubMed]
  218. Tradewell, M.L.; Yu, Z.; Tibshirani, M.; Boulanger, M.C.; Durham, H.D.; Richard, S. Arginine Methylation by PRMT1 Regulates Nuclear-Cytoplasmic Localization and Toxicity of FUS/TLS Harbouring ALS-Linked Mutations. Hum. Mol. Genet. 2012, 21, 136–149. [Google Scholar] [CrossRef] [PubMed]
  219. Deng, Q.; Holler, C.J.; Taylor, G.; Hudson, K.F.; Watkins, W.; Gearing, M.; Ito, D.; Murray, M.E.; Dickson, D.W.; Seyfried, N.T.; et al. FUS Is Phosphorylated by DNA-PK and Accumulates in the Cytoplasm after DNA Damage. J. Neurosci. 2014, 34, 7802–7813. [Google Scholar] [CrossRef]
  220. Li, P.A.; Hou, X.; Hao, S. Mitochondrial Biogenesis in Neurodegeneration. J. Neurosci. Res. 2017, 95, 2025–2029. [Google Scholar] [CrossRef]
  221. Panchal, K.; Tiwari, A.K. Mitochondrial Dynamics, a Key Executioner in Neurodegenerative Diseases. Mitochondrion 2019, 47, 151–173. [Google Scholar] [CrossRef]
  222. Bayer, H.; Lang, K.; Buck, E.; Higelin, J.; Barteczko, L.; Pasquarelli, N.; Sprissler, J.; Lucas, T.; Holzmann, K.; Demestre, M.; et al. ALS-Causing Mutations Differentially Affect PGC-1alpha Expression and Function in the Brain vs. Peripheral Tissues. Neurobiol. Dis. 2017, 97 Pt A, 36–45. [Google Scholar] [CrossRef]
  223. Palomo, G.M.; Granatiero, V.; Kawamata, H.; Konrad, C.; Kim, M.; Arreguin, A.J.; Zhao, D.; Milner, T.A.; Manfredi, G. Parkin Is a Disease Modifier in the Mutant SOD1 Mouse Model of ALS. EMBO Mol. Med. 2018, 10, e8888. [Google Scholar] [CrossRef]
  224. Thau, N.; Knippenberg, S.; Korner, S.; Rath, K.J.; Dengler, R.; Petri, S. Decreased mRNA Expression of PGC-1alpha and PGC-1alpha-Regulated Factors in the SOD1G93A ALS Mouse Model and in Human Sporadic ALS. J. Neuropathol. Exp. Neurol. 2012, 71, 1064–1074. [Google Scholar] [CrossRef] [PubMed]
  225. Masser, D.R.; Clark, N.W.; Van Remmen, H.; Freeman, W.M. Loss of the Antioxidant Enzyme CuZnSOD (Sod1) Mimics an Age-Related Increase in Absolute Mitochondrial DNA Copy Number in the Skeletal Muscle. Age 2016, 38, 323–333. [Google Scholar] [CrossRef] [PubMed]
  226. Stoccoro, A.; Mosca, L.; Carnicelli, V.; Cavallari, U.; Lunetta, C.; Marocchi, A.; Migliore, L.; Coppede, F. Mitochondrial DNA Copy Number and D-Loop Region Methylation in Carriers of Amyotrophic Lateral Sclerosis Gene Mutations. Epigenomics 2018, 10, 1431–1443. [Google Scholar] [CrossRef] [PubMed]
  227. Sheng, Z.H.; Cai, Q. Mitochondrial Transport in Neurons: Impact on Synaptic Homeostasis and Neurodegeneration. Nat. Rev. Neurosci. 2012, 13, 77–93. [Google Scholar] [CrossRef]
  228. Melkov, A.; Abdu, U. Regulation of Long-Distance Transport of Mitochondria along Microtubules. Cell. Mol. Life Sci. 2017, 75, 163–176. [Google Scholar] [CrossRef]
  229. Conde, C.; Caceres, A. Microtubule Assembly, Organization and Dynamics in Axons and Dendrites. Nat. Rev. Neurosci. 2009, 10, 319–332. [Google Scholar] [CrossRef] [PubMed]
  230. Tanaka, Y.; Kanai, Y.; Okada, Y.; Nonaka, S.; Takeda, S.; Harada, A.; Hirokawa, N. Targeted Disruption of Mouse Conventional Kinesin Heavy Chain, kif5B, Results in Abnormal Perinuclear Clustering of Mitochondria. Cell 1998, 93, 1147–1158. [Google Scholar] [CrossRef]
  231. Pilling, A.D.; Horiuchi, D.; Lively, C.M.; Saxton, W.M. Kinesin-1 and Dynein Are the Primary Motors for Fast Transport of Mitochondria in Drosophila Motor Axons. Mol. Biol. Cell 2006, 17, 2057–2068. [Google Scholar] [CrossRef] [PubMed]
  232. Hirokawa, N.; Niwa, S.; Tanaka, Y. Molecular Motors in Neurons: Transport Mechanisms and Roles in Brain Function, Development, and Disease. Neuron 2010, 68, 610–638. [Google Scholar] [CrossRef] [PubMed]
  233. Guo, X.; Macleod, G.T.; Wellington, A.; Hu, F.; Panchumarthi, S.; Schoenfield, M.; Marin, L.; Charlton, M.P.; Atwood, H.L.; Zinsmaier, K.E. The GTPase dMiro is Required for Axonal Transport of Mitochondria to Drosophila Synapses. Neuron 2005, 47, 379–393. [Google Scholar] [CrossRef] [PubMed]
  234. Stowers, R.S.; Megeath, L.J.; Gorska-Andrzejak, J.; Meinertzhagen, I.A.; Schwarz, T.L. Axonal Transport of Mitochondria to Synapses Depends on Milton, a Novel Drosophila Protein. Neuron 2002, 36, 1063–1077. [Google Scholar] [CrossRef] [PubMed]
  235. MacAskill, A.F.; Brickley, K.; Stephenson, F.A.; Kittler, J.T. GTPase Dependent Recruitment of Grif-1 by Miro1 Regulates Mitochondrial Trafficking in Hippocampal Neurons. Mol. Cell. Neurosci. 2009, 40, 301–312. [Google Scholar] [CrossRef] [PubMed]
  236. Brickley, K.; Stephenson, F.A. Trafficking Kinesin Protein (TRAK)-Mediated Transport of Mitochondria in Axons of Hippocampal Neurons. J. Biol. Chem. 2011, 286, 18079–18092. [Google Scholar] [CrossRef]
  237. Cai, Q.; Gerwin, C.; Sheng, Z.H. Syntabulin-Mediated Anterograde Transport of Mitochondria along Neuronal Processes. J. Cell Biol. 2005, 170, 959–969. [Google Scholar] [CrossRef]
  238. Cho, K.I.; Cai, Y.; Yi, H.; Yeh, A.; Aslanukov, A.; Ferreira, P.A. Association of the Kinesin-Binding Domain of RanBP2 to KIF5B and KIF5C Determines Mitochondria Localization and Function. Traffic 2007, 8, 1722–1735. [Google Scholar] [CrossRef]
  239. Altmann, K.; Frank, M.; Neumann, D.; Jakobs, S.; Westermann, B. The Class V Myosin Motor Protein, Myo2, Plays a Major Role in Mitochondrial Motility in Saccharomyces Cerevisiae. J. Cell Biol. 2008, 181, 119–130. [Google Scholar] [CrossRef]
  240. Boldogh, I.R.; Yang, H.C.; Nowakowski, W.D.; Karmon, S.L.; Hays, L.G.; Yates, J.R., 3rd; Pon, L.A. Arp2/3 Complex and Actin Dynamics Are Required for Actin-Based Mitochondrial Motility in Yeast. Proc. Natl. Acad. Sci. USA 2001, 98, 3162–3167. [Google Scholar] [CrossRef]
  241. Sheng, Z.H. Mitochondrial Trafficking and Anchoring in Neurons: New Insight and Implications. J. Cell Biol. 2014, 204, 1087–1098. [Google Scholar] [CrossRef]
  242. Kang, J.S.; Tian, J.H.; Pan, P.Y.; Zald, P.; Li, C.; Deng, C.; Sheng, Z.H. Docking of Axonal Mitochondria by Syntaphilin Controls Their Mobility and Affects Short-Term Facilitation. Cell 2008, 132, 137–148. [Google Scholar] [CrossRef]
  243. Chen, Y.M.; Gerwin, C.; Sheng, Z.H. Dynein Light Chain LC8 Regulates Syntaphilin-Mediated Mitochondrial Docking in Axons. J. Neurosci. 2009, 29, 9429–9438. [Google Scholar] [CrossRef]
  244. De Vos, K.J.; Hafezparast, M. Neurobiology of Axonal Transport Defects in Motor Neuron Diseases: Opportunities for Translational Research? Neurobiol. Dis. 2017, 105, 283–299. [Google Scholar] [CrossRef]
  245. Magrane, J.; Cortez, C.; Gan, W.B.; Manfredi, G. Abnormal Mitochondrial Transport and Morphology Are Common Pathological Denominators in SOD1 and TDP43 ALS Mouse Models. Hum. Mol. Genet. 2014, 23, 1413–1424. [Google Scholar] [CrossRef]
  246. Okamoto, K.; Hirai, S.; Shoji, M.; Senoh, Y.; Yamazaki, T. Axonal Swellings in the Corticospinal Tracts in Amyotrophic Lateral Sclerosis. Acta Neuropathol. 1990, 80, 222–226. [Google Scholar] [CrossRef]
  247. Sasaki, S.; Iwata, M. Ultrastructural Study of Synapses in the Anterior Horn Neurons of Patients with Amyotrophic Lateral Sclerosis. Neurosci. Lett. 1996, 204, 53–56. [Google Scholar] [CrossRef] [PubMed]
  248. Moller, A.; Bauer, C.S.; Cohen, R.N.; Webster, C.P.; De Vos, K.J. Amyotrophic Lateral Sclerosis-Associated Mutant SOD1 Inhibits Anterograde Axonal Transport of Mitochondria by Reducing Miro1 Levels. Hum. Mol. Genet. 2017, 26, 4668–4679. [Google Scholar] [CrossRef] [PubMed]
  249. De Vos, K.J.; Chapman, A.L.; Tennant, M.E.; Manser, C.; Tudor, E.L.; Lau, K.F.; Brownlees, J.; Ackerley, S.; Shaw, P.J.; McLoughlin, D.M.; et al. Familial Amyotrophic Lateral Sclerosis-Linked SOD1 Mutants Perturb Fast Axonal Transport to Reduce Axonal Mitochondria Content. Hum. Mol. Genet. 2007, 16, 2720–2728. [Google Scholar] [CrossRef] [PubMed]
  250. 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, 252. [Google Scholar] [CrossRef] [PubMed]
  251. Wang, X.; Winter, D.; Ashrafi, G.; Schlehe, J.; Wong, Y.L.; Selkoe, D.; Rice, S.; Steen, J.; LaVoie, M.J.; Schwarz, T.L. PINK1 and Parkin Target Miro for Phosphorylation and Degradation to Arrest Mitochondrial Motility. Cell 2011, 147, 893–906. [Google Scholar] [CrossRef] [PubMed]
  252. Smith, B.N.; Ticozzi, N.; Fallini, C.; Gkazi, A.S.; Topp, S.; Kenna, K.P.; Scotter, E.L.; Kost, J.; Keagle, P.; Miller, J.W.; et al. Exome-Wide Rare Variant Analysis Identifies TUBA4A Mutations Associated with Familial ALS. Neuron 2014, 84, 324–331. [Google Scholar] [CrossRef] [PubMed]
  253. Puls, I.; Jonnakuty, C.; LaMonte, B.H.; Holzbaur, E.L.; Tokito, M.; Mann, E.; Floeter, M.K.; Bidus, K.; Drayna, D.; Oh, S.J.; et al. Mutant Dynactin in Motor Neuron Disease. Nat. Genet. 2003, 33, 455–456. [Google Scholar] [CrossRef]
  254. Wu, C.H.; Fallini, C.; Ticozzi, N.; Keagle, P.J.; Sapp, P.C.; Piotrowska, K.; Lowe, P.; Koppers, M.; McKenna-Yasek, D.; Baron, D.M.; et al. Mutations in the Profilin 1 Gene Cause Familial Amyotrophic Lateral Sclerosis. Nature 2012, 488, 499–503. [Google Scholar] [CrossRef] [PubMed]
  255. Naia, L.; Ferreira, I.L.; Ferreiro, E.; Rego, A.C. Mitochondrial Ca(2+) Handling in Huntington’s and Alzheimer’s Diseases—Role of ER-Mitochondria Crosstalk. Biochem. Biophys. Res. Commun. 2017, 483, 1069–1077. [Google Scholar] [CrossRef]
  256. Ji, C.; Zhang, Z.; Li, Z.; She, X.; Wang, X.; Li, B.; Xu, X.; Song, D.; Zhang, D. Mitochondria-Associated Endoplasmic Reticulum Membranes: Inextricably Linked with Autophagy Process. Oxidative Med. Cell. Longev. 2022, 2022, 7086807. [Google Scholar] [CrossRef]
  257. Sack, R.B.; Kline, R.L.; Spira, W.M. Oral Immunization of Rabbits with Enterotoxigenic Escherichia coli Protects against Intraintestinal Challenge. Infect. Immun. 1988, 56, 387–394. [Google Scholar] [CrossRef]
  258. He, Q.; Qu, M.; Shen, T.; Su, J.; Xu, Y.; Xu, C.; Barkat, M.Q.; Cai, J.; Zhu, H.; Zeng, L.H.; et al. Control of Mitochondria-Associated Endoplasmic Reticulum Membranes by Protein S-Palmitoylation: Novel Therapeutic Targets for Neurodegenerative Diseases. Ageing Res. Rev. 2023, 87, 101920. [Google Scholar] [CrossRef]
  259. Sala-Vila, A.; Navarro-Lerida, I.; Sanchez-Alvarez, M.; Bosch, M.; Calvo, C.; Lopez, J.A.; Calvo, E.; Ferguson, C.; Giacomello, M.; Serafini, A.; et al. Interplay between Hepatic Mitochondria-Associated Membranes, Lipid Metabolism and Caveolin-1 in Mice. Sci. Rep. 2016, 6, 27351. [Google Scholar] [CrossRef]
  260. Tubbs, E.; Theurey, P.; Vial, G.; Bendridi, N.; Bravard, A.; Chauvin, M.A.; Ji-Cao, J.; Zoulim, F.; Bartosch, B.; Ovize, M.; et al. Mitochondria-Associated Endoplasmic Reticulum Membrane (MAM) Integrity Is Required for Insulin Signaling and Is Implicated in Hepatic Insulin Resistance. Diabetes 2014, 63, 3279–3294. [Google Scholar] [CrossRef]
  261. D’Eletto, M.; Rossin, F.; Occhigrossi, L.; Farrace, M.G.; Faccenda, D.; Desai, R.; Marchi, S.; Refolo, G.; Falasca, L.; Antonioli, M.; et al. Transglutaminase Type 2 Regulates ER-Mitochondria Contact Sites by Interacting with GRP75. Cell Rep. 2018, 25, 3573–3581.e4. [Google Scholar] [CrossRef]
  262. Couly, S.; Yasui, Y.; Su, T.P. SIGMAR1 Confers Innate Resilience against Neurodegeneration. Int. J. Mol. Sci. 2023, 24, 7767. [Google Scholar] [CrossRef]
  263. Gutierrez, T.; Simmen, T. Endoplasmic Reticulum Chaperones Tweak the Mitochondrial Calcium Rheostat to Control Metabolism and Cell Death. Cell Calcium 2018, 70, 64–75. [Google Scholar] [CrossRef] [PubMed]
  264. Hamasaki, M.; Furuta, N.; Matsuda, A.; Nezu, A.; Yamamoto, A.; Fujita, N.; Oomori, H.; Noda, T.; Haraguchi, T.; Hiraoka, Y.; et al. Autophagosomes form at ER-Mitochondria Contact Sites. Nature 2013, 495, 389–393. [Google Scholar] [CrossRef] [PubMed]
  265. Stoica, R.; De Vos, K.J.; Paillusson, S.; Mueller, S.; Sancho, R.M.; Lau, K.F.; Vizcay-Barrena, G.; Lin, W.L.; Xu, Y.F.; Lewis, J.; et al. ER-Mitochondria Associations Are Regulated by the VAPB-PTPIP51 Interaction and Are Disrupted by ALS/FTD-Associated TDP-43. Nat. Commun. 2014, 5, 3996. [Google Scholar] [CrossRef]
  266. Bernal, A.F.; Mota, N.; Pamplona, R.; Area-Gomez, E.; Portero-Otin, M. Hakuna MAM-Tata: Investigating the Role of Mitochondrial-Associated Membranes in ALS. Biochim. Biophys. Acta Mol. Basis Dis. 2023, 1869, 166716. [Google Scholar] [CrossRef]
  267. Stoica, R.; Paillusson, S.; Gomez-Suaga, P.; Mitchell, J.C.; Lau, D.H.; Gray, E.H.; Sancho, R.M.; Vizcay-Barrena, G.; De Vos, K.J.; Shaw, C.E.; et al. ALS/FTD-Associated FUS Activates GSK-3beta to Disrupt the VAPB-PTPIP51 Interaction and ER-Mitochondria Associations. EMBO Rep. 2016, 17, 1326–1342. [Google Scholar] [CrossRef]
  268. Karagas, N.E.; Gupta, R.; Rastegari, E.; Tan, K.L.; Leung, H.H.; Bellen, H.J.; Venkatachalam, K.; Wong, C.O. Loss of Activity-Induced Mitochondrial ATP Production Underlies the Synaptic Defects in a Drosophila Model of ALS. J. Neurosci. 2022, 42, 8019–8037. [Google Scholar] [CrossRef] [PubMed]
  269. Gomez-Suaga, P.; Morotz, G.M.; Markovinovic, A.; Martin-Guerrero, S.M.; Preza, E.; Arias, N.; Mayl, K.; Aabdien, A.; Gesheva, V.; Nishimura, A.; et al. Disruption of ER-Mitochondria Tethering and Signalling in C9orf72-Associated Amyotrophic Lateral Sclerosis and Frontotemporal Dementia. Aging Cell 2022, 21, e13549. [Google Scholar] [CrossRef]
  270. Pilotto, F.; Schmitz, A.; Maharjan, N.; Diab, R.; Odriozola, A.; Tripathi, P.; Yamoah, A.; Scheidegger, O.; Oestmann, A.; Dennys, C.N.; et al. PolyGA Targets the ER Stress-Adaptive Response by Impairing GRP75 Function at the MAM in C9ORF72-ALS/FTD. Acta Neuropathol. 2022, 144, 939–966. [Google Scholar] [CrossRef]
  271. Jeon, Y.M.; Kwon, Y.; Lee, S.; Kim, H.J. Potential Roles of the Endoplasmic Reticulum Stress Pathway in Amyotrophic Lateral Sclerosis. Front. Aging Neurosci. 2023, 15, 1047897. [Google Scholar] [CrossRef]
  272. Saxena, S.; Cabuy, E.; Caroni, P. A Role for Motoneuron Subtype-Selective ER Stress in Disease Manifestations of FALS Mice. Nat. Neurosci. 2009, 12, 627–636. [Google Scholar] [CrossRef]
  273. Ono, Y.; Tanaka, H.; Takata, M.; Nagahara, Y.; Noda, Y.; Tsuruma, K.; Shimazawa, M.; Hozumi, I.; Hara, H. SA4503, a Sigma-1 Receptor Agonist, Suppresses Motor Neuron Damage in In Vitro and In Vivo Amyotrophic Lateral Sclerosis Models. Neurosci. Lett. 2014, 559, 174–178. [Google Scholar] [CrossRef] [PubMed]
  274. Das, I.; Krzyzosiak, A.; Schneider, K.; Wrabetz, L.; D’Antonio, M.; Barry, N.; Sigurdardottir, A.; Bertolotti, A. Preventing Proteostasis Diseases by Selective Inhibition of a Phosphatase Regulatory Subunit. Science 2015, 348, 239–242. [Google Scholar] [CrossRef]
  275. Chen, Y.; Podojil, J.R.; Kunjamma, R.B.; Jones, J.; Weiner, M.; Lin, W.; Miller, S.D.; Popko, B. Sephin1, Which Prolongs the Integrated Stress Response, Is a Promising Therapeutic for Multiple Sclerosis. Brain 2019, 142, 344–361. [Google Scholar] [CrossRef]
  276. Fujikake, N.; Shin, M.; Shimizu, S. Association between Autophagy and Neurodegenerative Diseases. Front. Neurosci. 2018, 12, 255. [Google Scholar] [CrossRef]
  277. Mizushima, N. Autophagy: Process and Function. Genes Dev. 2007, 21, 2861–2873. [Google Scholar] [CrossRef] [PubMed]
  278. Kroemer, G.; Levine, B. Autophagic Cell Death: The Story of a Misnomer. Nat. Rev. Mol. Cell Biol. 2008, 9, 1004–1010. [Google Scholar] [CrossRef] [PubMed]
  279. Song, C.Y.; Guo, J.F.; Liu, Y.; Tang, B.S. Autophagy and Its Comprehensive Impact on ALS. Int. J. Neurosci. 2012, 122, 695–703. [Google Scholar] [CrossRef]
  280. Park, H.; Kang, J.H.; Lee, S. Autophagy in Neurodegenerative Diseases: A Hunter for Aggregates. Int. J. Mol. Sci. 2020, 21, 3369. [Google Scholar] [CrossRef]
  281. King, A.; Al-Sarraj, S.; Troakes, C.; Smith, B.N.; Maekawa, S.; Iovino, M.; Spillantini, M.G.; Shaw, C.E. Mixed tau, TDP-43 and p62 Pathology in FTLD Associated with a C9ORF72 Repeat Expansion and p.Ala239Thr MAPT (tau) Variant. Acta Neuropathol. 2013, 125, 303–310. [Google Scholar] [CrossRef]
  282. Al-Sarraj, S.; King, A.; Troakes, C.; Smith, B.; Maekawa, S.; Bodi, I.; Rogelj, B.; Al-Chalabi, A.; Hortobagyi, T.; Shaw, C.E. p62 Positive, TDP-43 Negative, Neuronal Cytoplasmic and Intranuclear Inclusions in the Cerebellum and Hippocampus Define the Pathology of C9orf72-Linked FTLD and MND/ALS. Acta Neuropathol. 2011, 122, 691–702. [Google Scholar] [CrossRef]
  283. Duan, W.; Guo, M.; Yi, L.; Zhang, J.; Bi, Y.; Liu, Y.; Li, Y.; Li, Z.; Ma, Y.; Zhang, G.; et al. Deletion of Tbk1 Disrupts Autophagy and Reproduces Behavioral and Locomotor Symptoms of FTD-ALS in Mice. Aging 2019, 11, 2457–2476. [Google Scholar] [CrossRef]
  284. Chen, T.; Huang, B.; Shi, X.; Gao, L.; Huang, C. Mutant UBQLN2(P497H) in Motor Neurons Leads to ALS-Like Phenotypes and Defective Autophagy in Rats. Acta Neuropathol. Commun. 2018, 6, 122. [Google Scholar] [CrossRef] [PubMed]
  285. Zhang, J.J.; Zhou, Q.M.; Chen, S.; Le, W.D. Repurposing Carbamazepine for the Treatment of Amyotrophic Lateral Sclerosis in SOD1-G93A Mouse Model. CNS Neurosci. Ther. 2018, 24, 1163–1174. [Google Scholar] [CrossRef] [PubMed]
  286. Vicencio, E.; Beltran, S.; Labrador, L.; Manque, P.; Nassif, M.; Woehlbier, U. Implications of Selective Autophagy Dysfunction for ALS Pathology. Cells 2020, 9, 381. [Google Scholar] [CrossRef]
  287. Freischmidt, A.; Wieland, T.; Richter, B.; Ruf, W.; Schaeffer, V.; Muller, K.; Marroquin, N.; Nordin, F.; Hubers, A.; Weydt, P.; et al. Haploinsufficiency of TBK1 Causes Familial ALS and Fronto-Temporal Dementia. Nat. Neurosci. 2015, 18, 631–636. [Google Scholar] [CrossRef]
  288. Rea, S.L.; Majcher, V.; Searle, M.S.; Layfield, R. SQSTM1 Mutations—Bridging Paget Disease of Bone and ALS/FTLD. Exp. Cell Res. 2014, 325, 27–37. [Google Scholar] [CrossRef]
  289. Gal, J.; Strom, A.L.; Kwinter, D.M.; Kilty, R.; Zhang, J.; Shi, P.; Fu, W.; Wooten, M.W.; Zhu, H. Sequestosome 1/p62 Links Familial ALS Mutant SOD1 to LC3 via an Ubiquitin-Independent Mechanism. J. Neurochem. 2009, 111, 1062–1073. [Google Scholar] [CrossRef]
  290. Korac, J.; Schaeffer, V.; Kovacevic, I.; Clement, A.M.; Jungblut, B.; Behl, C.; Terzic, J.; Dikic, I. Ubiquitin-Independent Function of Optineurin in Autophagic Clearance of Protein Aggregates. J. Cell Sci. 2013, 126, 580–592. [Google Scholar] [CrossRef] [PubMed]
  291. Shen, W.C.; Li, H.Y.; Chen, G.C.; Chern, Y.; Tu, P.H. Mutations in the Ubiquitin-Binding Domain of OPTN/Optineurin Interfere with Autophagy-Mediated Degradation of Misfolded Proteins by a Dominant-Negative Mechanism. Autophagy 2015, 11, 685–700. [Google Scholar] [CrossRef]
  292. Taylor, J.P.; Brown, R.H., Jr.; Cleveland, D.W. Decoding ALS: From Genes to Mechanism. Nature 2016, 539, 197–206. [Google Scholar] [CrossRef]
  293. Kim, J.; Kundu, M.; Viollet, B.; Guan, K.L. AMPK and mTOR Regulate Autophagy through Direct Phosphorylation of Ulk1. Nat. Cell Biol. 2011, 13, 132–141. [Google Scholar] [CrossRef]
  294. Wei, Y. Autophagic Induction of Amyotrophic Lateral Sclerosis-Linked Cu/Zn Superoxide Dismutase 1 G93A Mutant in NSC34 Cells. Neural Regen. Res. 2014, 9, 16–24. [Google Scholar] [CrossRef]
  295. Morimoto, N.; Nagai, M.; Ohta, Y.; Miyazaki, K.; Kurata, T.; Morimoto, M.; Murakami, T.; Takehisa, Y.; Ikeda, Y.; Kamiya, T.; et al. Increased Autophagy in Transgenic Mice with a G93A Mutant SOD1 Gene. Brain Res. 2007, 1167, 112–117. [Google Scholar] [CrossRef]
  296. An, T.; Shi, P.; Duan, W.; Zhang, S.; Yuan, P.; Li, Z.; Wu, D.; Xu, Z.; Li, C.; Guo, Y. Oxidative Stress and Autophagic Alteration in Brainstem of SOD1-G93A Mouse Model of ALS. Mol. Neurobiol. 2014, 49, 1435–1448. [Google Scholar] [CrossRef] [PubMed]
  297. Settembre, C.; Ballabio, A. TFEB Regulates Autophagy: An Integrated Coordination of Cellular Degradation and Recycling Processes. Autophagy 2011, 7, 1379–1381. [Google Scholar] [CrossRef]
  298. Kang, R.; Zeh, H.J.; Lotze, M.T.; Tang, D. The Beclin 1 Network Regulates Autophagy and Apoptosis. Cell Death Differ. 2011, 18, 571–580. [Google Scholar] [CrossRef] [PubMed]
  299. Chen, Y.; Liu, H.; Guan, Y.; Wang, Q.; Zhou, F.; Jie, L.; Ju, J.; Pu, L.; Du, H.; Wang, X. The Altered Autophagy Mediated by TFEB in Animal and Cell Models of Amyotrophic Lateral Sclerosis. Am. J. Transl. Res. 2015, 7, 1574–1587. [Google Scholar] [PubMed]
  300. Zhang, F.; Strom, A.L.; Fukada, K.; Lee, S.; Hayward, L.J.; Zhu, H. Interaction between Familial Amyotrophic Lateral Sclerosis (ALS)-Linked SOD1 Mutants and the Dynein Complex. J. Biol. Chem. 2007, 282, 16691–16699. [Google Scholar] [CrossRef]
  301. Johnson, B.S.; Snead, D.; Lee, J.J.; McCaffery, J.M.; Shorter, J.; Gitler, A.D. TDP-43 Is Intrinsically Aggregation-Prone, and Amyotrophic Lateral Sclerosis-Linked Mutations Accelerate Aggregation and Increase Toxicity. J. Biol. Chem. 2009, 284, 20329–20339. [Google Scholar] [CrossRef] [PubMed]
  302. Sun, Z.; Diaz, Z.; Fang, X.; Hart, M.P.; Chesi, A.; Shorter, J.; Gitler, A.D. Molecular Determinants and Genetic Modifiers of Aggregation and Toxicity for the ALS Disease Protein FUS/TLS. PLoS Biol. 2011, 9, e1000614. [Google Scholar] [CrossRef] [PubMed]
  303. Ryu, H.H.; Jun, M.H.; Min, K.J.; Jang, D.J.; Lee, Y.S.; Kim, H.K.; Lee, J.A. Autophagy Regulates Amyotrophic Lateral Sclerosis-Linked Fused in Sarcoma-Positive Stress Granules in Neurons. Neurobiol. Aging 2014, 35, 2822–2831. [Google Scholar] [CrossRef] [PubMed]
  304. Cheng, C.W.; Lin, M.J.; Shen, C.K. Rapamycin Alleviates Pathogenesis of a New Drosophila Model of ALS-TDP. J. Neurogenet. 2015, 29, 59–68. [Google Scholar] [CrossRef]
  305. Wang, I.F.; Guo, B.S.; Liu, Y.C.; Wu, C.C.; Yang, C.H.; Tsai, K.J.; Shen, C.K. Autophagy Activators Rescue and Alleviate Pathogenesis of a Mouse Model with Proteinopathies of the TAR DNA-Binding Protein 43. Proc. Natl. Acad. Sci. USA 2012, 109, 15024–15029. [Google Scholar] [CrossRef] [PubMed]
  306. Almeida, S.; Gascon, E.; Tran, H.; Chou, H.J.; Gendron, T.F.; Degroot, S.; Tapper, A.R.; Sellier, C.; Charlet-Berguerand, N.; Karydas, A.; et al. Modeling Key Pathological Features of Frontotemporal Dementia with C9ORF72 Repeat Expansion in iPSC-Derived Human Neurons. Acta Neuropathol. 2013, 126, 385–399. [Google Scholar] [CrossRef] [PubMed]
  307. Ugolino, J.; Ji, Y.J.; Conchina, K.; Chu, J.; Nirujogi, R.S.; Pandey, A.; Brady, N.R.; Hamacher-Brady, A.; Wang, J. Loss of C9orf72 Enhances Autophagic Activity via Deregulated mTOR and TFEB Signaling. PLoS Genet. 2016, 12, e1006443. [Google Scholar] [CrossRef]
  308. Amick, J.; Roczniak-Ferguson, A.; Ferguson, S.M. C9orf72 Binds SMCR8, Localizes to Lysosomes, and Regulates mTORC1 Signaling. Mol. Biol. Cell 2016, 27, 3040–3051. [Google Scholar] [CrossRef]
  309. Sullivan, P.M.; Zhou, X.; Robins, A.M.; Paushter, D.H.; Kim, D.; Smolka, M.B.; Hu, F. The ALS/FTLD Associated Protein C9orf72 Associates with SMCR8 and WDR41 to Regulate the Autophagy-Lysosome Pathway. Acta Neuropathol. Commun. 2016, 4, 51. [Google Scholar] [CrossRef]
  310. Wang, Y.; Liu, N.; Lu, B. Mechanisms and Roles of Mitophagy in Neurodegenerative Diseases. CNS Neurosci. Ther. 2019, 25, 859–875. [Google Scholar] [CrossRef]
  311. Edens, B.M.; Miller, N.; Ma, Y.C. Impaired Autophagy and Defective Mitochondrial Function: Converging Paths on the Road to Motor Neuron Degeneration. Front. Cell. Neurosci. 2016, 10, 44. [Google Scholar] [CrossRef]
  312. Lazarou, M.; Sliter, D.A.; Kane, L.A.; Sarraf, S.A.; Wang, C.; Burman, J.L.; Sideris, D.P.; Fogel, A.I.; Youle, R.J. The Ubiquitin Kinase PINK1 Recruits Autophagy Receptors to Induce Mitophagy. Nature 2015, 524, 309–314. [Google Scholar] [CrossRef] [PubMed]
  313. Moore, A.S.; Holzbaur, E.L. Dynamic Recruitment and Activation of ALS-Associated TBK1 with Its Target Optineurin Are Required for Efficient Mitophagy. Proc. Natl. Acad. Sci. USA 2016, 113, E3349–E3358. [Google Scholar] [CrossRef] [PubMed]
  314. Wong, Y.C.; Holzbaur, E.L. Optineurin Is an Autophagy Receptor for Damaged Mitochondria in Parkin-Mediated Mitophagy That Is Disrupted by an ALS-Linked Mutation. Proc. Natl. Acad. Sci. USA 2014, 111, E4439–E4448. [Google Scholar] [CrossRef]
  315. Gautam, M.; Xie, E.F.; Kocak, N.; Ozdinler, P.H. Mitoautophagy: A Unique Self-Destructive Path Mitochondria of Upper Motor Neurons with TDP-43 Pathology Take, Very Early in ALS. Front. Cell. Neurosci. 2019, 13, 489. [Google Scholar] [CrossRef] [PubMed]
  316. Mackenzie, I.R.; Bigio, E.H.; Ince, P.G.; Geser, F.; Neumann, M.; Cairns, N.J.; Kwong, L.K.; Forman, M.S.; Ravits, J.; Stewart, H.; et al. Pathological TDP-43 Distinguishes Sporadic Amyotrophic Lateral Sclerosis from Amyotrophic Lateral Sclerosis with SOD1 Mutations. Ann. Neurol. 2007, 61, 427–434. [Google Scholar] [CrossRef] [PubMed]
  317. Martin, L.J. Neuronal Death in Amyotrophic Lateral Sclerosis Is Apoptosis: Possible Contribution of a Programmed Cell Death Mechanism. J. Neuropathol. Exp. Neurol. 1999, 58, 459–471. [Google Scholar] [CrossRef] [PubMed]
  318. Ilzecka, J.; Stelmasiak, Z.; Dobosz, B. Interleukin-1beta Converting Enzyme/Caspase-1 (ICE/Caspase-1) and Soluble APO-1/Fas/CD 95 Receptor in Amyotrophic Lateral Sclerosis Patients. Acta Neurol. Scand. 2001, 103, 255–258. [Google Scholar] [CrossRef]
  319. Li, M.; Ona, V.O.; Guegan, C.; Chen, M.; Jackson-Lewis, V.; Andrews, L.J.; Olszewski, A.J.; Stieg, P.E.; Lee, J.P.; Przedborski, S.; et al. Functional Role of Caspase-1 and Caspase-3 in an ALS Transgenic Mouse Model. Science 2000, 288, 335–339. [Google Scholar] [CrossRef]
  320. Pasinelli, P.; Houseweart, M.K.; Brown, R.H., Jr.; Cleveland, D.W. Caspase-1 and -3 Are Sequentially Activated in Motor Neuron Death in Cu,Zn Superoxide Dismutase-Mediated Familial Amyotrophic Lateral Sclerosis. Proc. Natl. Acad. Sci. USA 2000, 97, 13901–13906. [Google Scholar] [CrossRef]
  321. Vukosavic, S.; Stefanis, L.; Jackson-Lewis, V.; Guegan, C.; Romero, N.; Chen, C.; Dubois-Dauphin, M.; Przedborski, S. Delaying Caspase Activation by Bcl-2: A Clue to Disease Retardation in a Transgenic Mouse Model of Amyotrophic Lateral Sclerosis. J. Neurosci. 2000, 20, 9119–9125. [Google Scholar] [CrossRef]
  322. Guégan, C.; Vila, M.; Rosoklija, G.; Hays, A.P.; Przedborski, S. Recruitment of the Mitochondrial-Dependent Apoptotic Pathway in Amyotrophic Lateral Sclerosis. J. Neurosci. 2001, 21, 6569–6576. [Google Scholar] [CrossRef]
  323. Damiano, M.; Starkov, A.A.; Petri, S.; Kipiani, K.; Kiaei, M.; Mattiazzi, M.; Flint Beal, M.; Manfredi, G. Neural Mitochondrial Ca2+ Capacity Impairment Precedes the Onset of Motor Symptoms in G93A Cu/Zn-Superoxide Dismutase Mutant Mice. J. Neurochem. 2006, 96, 1349–1361. [Google Scholar] [CrossRef]
  324. Silva, F.S.; Costa, C.F.; Marques, R.J.; Oliveira, P.J.; Pereira, G.C. Pharmacological Targeting of the Mitochondrial Permeability Transition Pore for Cardioprotection. In Mitochondrial Biology and Experimental Therapeutics; Oliveira, P.J., Ed.; Springer: Berlin/Heidelberg, Germany, 2018; pp. 423–490. [Google Scholar] [CrossRef]
  325. Kalani, K.; Yan, S.F.; Yan, S.S. Mitochondrial Permeability Transition Pore: A Potential Drug Target for Neurodegeneration. Drug Discov. Today 2018, 23, 1983–1989. [Google Scholar] [CrossRef]
  326. Halestrap, A.P. What Is the Mitochondrial Permeability Transition Pore? J. Mol. Cell. Cardiol. 2009, 46, 821–831. [Google Scholar] [CrossRef] [PubMed]
  327. Kinnally, K.W.; Peixoto, P.M.; Ryu, S.Y.; Dejean, L.M. Is mPTP the Gatekeeper for Necrosis, Apoptosis, or Both? Biochim. Biophys. Acta 2011, 1813, 616–622. [Google Scholar] [CrossRef] [PubMed]
  328. Hurst, S.; Hoek, J.; Sheu, S.S. Mitochondrial Ca(2+) and Regulation of the Permeability Transition Pore. J. Bioenerg. Biomembr. 2017, 49, 27–47. [Google Scholar] [CrossRef] [PubMed]
  329. Bernardi, P.; Rasola, A.; Forte, M.; Lippe, G. The Mitochondrial Permeability Transition Pore: Channel Formation by F-ATP Synthase, Integration in Signal Transduction, and Role in Pathophysiology. Physiol. Rev. 2015, 95, 1111–1155. [Google Scholar] [CrossRef]
  330. Bonora, M.; Morganti, C.; Morciano, G.; Pedriali, G.; Lebiedzinska-Arciszewska, M.; Aquila, G.; Giorgi, C.; Rizzo, P.; Campo, G.; Ferrari, R.; et al. Mitochondrial Permeability Transition Involves Dissociation of F1FO ATP Synthase Dimers and C-Ring Conformation. EMBO Rep. 2017, 18, 1077–1089. [Google Scholar] [CrossRef] [PubMed]
  331. Martin, L.J.; Gertz, B.; Pan, Y.; Price, A.C.; Molkentin, J.D.; Chang, Q. The Mitochondrial Permeability Transition Pore in Motor Neurons: Involvement in the Pathobiology of ALS Mice. Exp. Neurol. 2009, 218, 333–346. [Google Scholar] [CrossRef] [PubMed]
  332. Parone, P.A.; Da Cruz, S.; Han, J.S.; McAlonis-Downes, M.; Vetto, A.P.; Lee, S.K.; Tseng, E.; Cleveland, D.W. Enhancing Mitochondrial Calcium Buffering Capacity Reduces Aggregation of Misfolded SOD1 and Motor Neuron Cell Death without Extending Survival in Mouse Models of Inherited Amyotrophic Lateral Sclerosis. J. Neurosci. 2013, 33, 4657–4671. [Google Scholar] [CrossRef]
  333. 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]
  334. Karam, C.; Yi, J.; Xiao, Y.; Dhakal, K.; Zhang, L.; Li, X.; Manno, C.; Xu, J.; Li, K.; Cheng, H.; et al. Absence of Physiological Ca(2+) Transients Is an Initial Trigger for Mitochondrial Dysfunction in Skeletal Muscle Following Denervation. Skelet. Muscle 2017, 7, 6. [Google Scholar] [CrossRef]
  335. Gautam, M.; Gunay, A.; Chandel, N.S.; Ozdinler, P.H. Mitochondrial Dysregulation Occurs early in ALS Motor Cortex with TDP-43 Pathology and Suggests Maintaining NAD(+) Balance as a Therapeutic Strategy. Sci. Rep. 2022, 12, 4287. [Google Scholar] [CrossRef] [PubMed]
  336. So, E.; Mitchell, J.C.; Memmi, C.; Chennell, G.; Vizcay-Barrena, G.; Allison, L.; Shaw, C.E.; Vance, C. Mitochondrial Abnormalities and Disruption of the Neuromuscular Junction Precede the Clinical Phenotype and Motor Neuron Loss in hFUSWT Transgenic Mice. Hum. Mol. Genet. 2018, 27, 463–474. [Google Scholar] [CrossRef]
  337. Belosludtseva, N.V.; Matveeva, L.A.; Belosludtsev, K.N. Mitochondrial Dyshomeostasis as an Early Hallmark and a Therapeutic Target in Amyotrophic Lateral Sclerosis. Int. J. Mol. Sci. 2023, 24, 16833. [Google Scholar] [CrossRef] [PubMed]
  338. Goyal, N.A.; Berry, J.D.; Windebank, A.; Staff, N.P.; Maragakis, N.J.; van den Berg, L.H.; Genge, A.; Miller, R.; Baloh, R.H.; Kern, R.; et al. Addressing Heterogeneity in Amyotrophic Lateral Sclerosis Clinical Trials. Muscle Nerve 2020, 62, 156–166. [Google Scholar] [CrossRef]
  339. Beghi, E.; Mennini, T.; Bendotti, C.; Bigini, P.; Logroscino, G.; Chio, A.; Hardiman, O.; Mitchell, D.; Swingler, R.; Traynor, B.J.; et al. The Heterogeneity of Amyotrophic Lateral Sclerosis: A Possible Explanation of Treatment Failure. Curr. Med. Chem. 2007, 14, 3185–3200. [Google Scholar] [CrossRef] [PubMed]
  340. Obrador, E.; Salvador-Palmer, R.; Lopez-Blanch, R.; Jihad-Jebbar, A.; Valles, S.L.; Estrela, J.M. The Link between Oxidative Stress, Redox Status, Bioenergetics and Mitochondria in the Pathophysiology of ALS. Int. J. Mol. Sci. 2021, 22, 6352. [Google Scholar] [CrossRef]
  341. Jha, M.K.; Jeon, S.; Suk, K. Pyruvate Dehydrogenase Kinases in the Nervous System: Their Principal Functions in Neuronal-glial Metabolic Interaction and Neuro-metabolic Disorders. Curr. Neuropharmacol. 2012, 10, 393–403. [Google Scholar] [CrossRef]
  342. Miquel, E.; Cassina, A.; Martinez-Palma, L.; Bolatto, C.; Trias, 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]
  343. Martinez-Palma, L.; Miquel, E.; Lagos-Rodriguez, V.; Barbeito, L.; Cassina, A.; Cassina, P. Mitochondrial Modulation by Dichloroacetate Reduces Toxicity of Aberrant Glial Cells and Gliosis in the SOD1G93A Rat Model of Amyotrophic Lateral Sclerosis. Neurotherapeutics 2019, 16, 203–215. [Google Scholar] [CrossRef] [PubMed]
  344. Zhao, W.; Varghese, M.; Vempati, P.; Dzhun, A.; Cheng, A.; Wang, J.; Lange, D.; Bilski, A.; Faravelli, I.; Pasinetti, G.M. Caprylic Triglyceride as a Novel Therapeutic Approach to Effectively Improve the Performance and Attenuate the Symptoms Due to the Motor Neuron Loss in ALS Disease. PLoS ONE 2012, 7, e49191. [Google Scholar] [CrossRef] [PubMed]
  345. Zhao, Z.; Lange, D.J.; Voustianiouk, A.; MacGrogan, D.; Ho, L.; Suh, J.; Humala, N.; Thiyagarajan, M.; Wang, J.; Pasinetti, G.M. A Ketogenic Diet as a Potential Novel Therapeutic Intervention in Amyotrophic Lateral Sclerosis. BMC Neurosci. 2006, 7, 29. [Google Scholar] [CrossRef] [PubMed]
  346. Pedersen, W.A.; Mattson, M.P. No Benefit of Dietary Restriction on Disease Onset or Progression in Amyotrophic Lateral Sclerosis Cu/Zn-Superoxide Dismutase Mutant Mice. Brain Res. 1999, 833, 117–120. [Google Scholar] [CrossRef] [PubMed]
  347. Hamadeh, M.J.; Rodriguez, M.C.; Kaczor, J.J.; Tarnopolsky, M.A. Caloric Restriction Transiently Improves Motor Performance but Hastens Clinical Onset of Disease in the Cu/Zn-Superoxide Dismutase Mutant G93A Mouse. Muscle Nerve 2005, 31, 214–220. [Google Scholar] [CrossRef] [PubMed]
  348. Mattson, M.P.; Cutler, R.G.; Camandola, S. Energy Intake and Amyotrophic Lateral Sclerosis. Neuromol. Med. 2007, 9, 17–20. [Google Scholar] [CrossRef] [PubMed]
  349. Tefera, T.W.; Borges, K. Metabolic Dysfunctions in Amyotrophic Lateral Sclerosis Pathogenesis and Potential Metabolic Treatments. Front. Neurosci. 2016, 10, 611. [Google Scholar] [CrossRef] [PubMed]
  350. Branco, A.F.; Ferreira, A.; Simoes, R.F.; Magalhaes-Novais, S.; Zehowski, C.; Cope, E.; Silva, A.M.; Pereira, D.; Sardao, V.A.; Cunha-Oliveira, T. Ketogenic Diets: From Cancer to Mitochondrial Diseases and Beyond. Eur. J. Clin. Investig. 2016, 46, 285–298. [Google Scholar] [CrossRef]
  351. Laffel, L. Ketone Bodies: A Review of Physiology, Pathophysiology and Application of Monitoring to Diabetes. Diabetes Metab. Res. Rev. 1999, 15, 412–426. [Google Scholar] [CrossRef]
  352. Ludolph, A.C.; Dorst, J.; Dreyhaupt, J.; Weishaupt, J.H.; Kassubek, J.; Weiland, U.; Meyer, T.; Petri, S.; Hermann, A.; Emmer, A.; et al. Effect of High-Caloric Nutrition on Survival in Amyotrophic Lateral Sclerosis. Ann. Neurol. 2020, 87, 206–216. [Google Scholar] [CrossRef]
  353. Wang, S.; Yuan, T.; Yang, H.; Zhou, X.; Cao, J. Effect of Complete High-Caloric Nutrition on the Nutritional Status and Survival Rate of Amyotrophic Lateral Sclerosis Patients after Gastrostomy. Am. J. Transl. Res. 2022, 14, 7842–7851. [Google Scholar] [PubMed]
  354. Vance, J.E.; Campenot, R.B.; Vance, D.E. The Synthesis and Transport of Lipids for Axonal Growth and Nerve Regeneration. Biochim. Biophys. Acta 2000, 1486, 84–96. [Google Scholar] [CrossRef] [PubMed]
  355. Dupuis, L.; Corcia, P.; Fergani, A.; Gonzalez De Aguilar, J.L.; Bonnefont-Rousselot, D.; Bittar, R.; Seilhean, D.; Hauw, J.J.; Lacomblez, L.; Loeffler, J.P.; et al. Dyslipidemia Is a Protective Factor in Amyotrophic Lateral Sclerosis. Neurology 2008, 70, 1004–1009. [Google Scholar] [CrossRef] [PubMed]
  356. Onofrj, M.; Ciccocioppo, F.; Varanese, S.; di Muzio, A.; Calvani, M.; Chiechio, S.; Osio, M.; Thomas, A. Acetyl-L-Carnitine: From a Biological Curiosity to a Drug for the Peripheral Nervous System and Beyond. Expert Rev. Neurother. 2013, 13, 925–936. [Google Scholar] [CrossRef] [PubMed]
  357. Kira, Y.; Nishikawa, M.; Ochi, A.; Sato, E.; Inoue, M. L-Carnitine Suppresses the Onset of Neuromuscular Degeneration and Increases the Life Span of Mice with Familial Amyotrophic Lateral Sclerosis. Brain Res. 2006, 1070, 206–214. [Google Scholar] [CrossRef] [PubMed]
  358. Beghi, E.; Pupillo, E.; Bonito, V.; Buzzi, P.; Caponnetto, C.; Chio, A.; Corbo, M.; Giannini, F.; Inghilleri, M.; Bella, V.L.; et al. Randomized Double-Blind Placebo-Controlled Trial Of Acetyl-L-Carnitine for ALS. Amyotroph. Lateral Scler. Front. Degener. 2013, 14, 397–405. [Google Scholar] [CrossRef] [PubMed]
  359. Grossini, E.; De Marchi, F.; Venkatesan, S.; Mele, A.; Ferrante, D.; Mazzini, L. Effects of Acetyl-L-Carnitine on Oxidative Stress in Amyotrophic Lateral Sclerosis Patients: Evaluation on Plasma Markers and Members of the Neurovascular Unit. Antioxidants 2023, 12, 1887. [Google Scholar] [CrossRef] [PubMed]
  360. Andreassen, O.A.; Dedeoglu, A.; Klivenyi, P.; Beal, M.F.; Bush, A.I. N-Acetyl-L-Cysteine Improves Survival and Preserves Motor Performance in an Animal Model of Familial Amyotrophic Lateral Sclerosis. Neuroreport 2000, 11, 2491–2493. [Google Scholar] [CrossRef]
  361. Burgunder, J.M.; Varriale, A.; Lauterburg, B.H. Effect of N-Acetylcysteine on Plasma Cysteine and Glutathione Following Paracetamol Administration. Eur. J. Clin. Pharmacol. 1989, 36, 127–131. [Google Scholar] [CrossRef]
  362. Beretta, S.; Sala, G.; Mattavelli, L.; Ceresa, C.; Casciati, A.; Ferri, A.; Carri, M.T.; Ferrarese, C. Mitochondrial Dysfunction Due to Mutant Copper/Zinc Superoxide Dismutase Associated with Amyotrophic Lateral Sclerosis Is Reversed by N-Acetylcysteine. Neurobiol. Dis. 2003, 13, 213–221. [Google Scholar] [CrossRef] [PubMed]
  363. Louwerse, E.S.; Weverling, G.J.; Bossuyt, P.M.; Meyjes, F.E.; de Jong, J.M. Randomized, Double-Blind, Controlled Trial of Acetylcysteine in Amyotrophic Lateral Sclerosis. Arch. Neurol. 1995, 52, 559–564. [Google Scholar] [CrossRef]
  364. Kurano, T.; Kanazawa, T.; Iioka, S.; Kondo, H.; Kosuge, Y.; Suzuki, T. Intranasal Administration of N-Acetyl-L-Cysteine Combined with Cell-Penetrating Peptide-Modified Polymer Nanomicelles as a Potential Therapeutic Approach for Amyotrophic Lateral Sclerosis. Pharmaceutics 2022, 14, 2590. [Google Scholar] [CrossRef] [PubMed]
  365. Bavarsad Shahripour, R.; Harrigan, M.R.; Alexandrov, A.V. N-Acetylcysteine (NAC) in Neurological Disorders: Mechanisms of Action and Therapeutic Opportunities. Brain Behav. 2014, 4, 108–122. [Google Scholar] [CrossRef]
  366. Giustarini, D.; Milzani, A.; Dalle-Donne, I.; Tsikas, D.; Rossi, R. N-Acetylcysteine Ethyl Ester (NACET): A Novel Lipophilic Cell-Permeable Cysteine Derivative with an Unusual Pharmacokinetic Feature and Remarkable Antioxidant Potential. Biochem. Pharmacol. 2012, 84, 1522–1533. [Google Scholar] [CrossRef] [PubMed]
  367. Kularatne, R.N.; Bulumulla, C.; Catchpole, T.; Takacs, A.; Christie, A.; Stefan, M.C.; Csaky, K.G. Protection of Human Retinal Pigment Epithelial Cells from Oxidative Damage Using Cysteine Prodrugs. Free Radic. Biol. Med. 2020, 152, 386–394. [Google Scholar] [CrossRef]
  368. Sunitha, K.; Hemshekhar, M.; Thushara, R.M.; Santhosh, M.S.; Yariswamy, M.; Kemparaju, K.; Girish, K.S. N-Acetylcysteine Amide: A Derivative to Fulfill the Promises of N-Acetylcysteine. Free Radic. Res. 2013, 47, 357–367. [Google Scholar] [CrossRef]
  369. Atlas, D. Emerging Therapeutic Opportunities of Novel Thiol-Amides, NAC-Amide (AD4/NACA) and Thioredoxin Mimetics (TXM-Peptides) for Neurodegenerative-Related Disorders. Free Radic. Biol. Med. 2021, 176, 120–141. [Google Scholar] [CrossRef]
  370. Shichinohe, H.; Kuroda, S.; Yasuda, H.; Ishikawa, T.; Iwai, M.; Horiuchi, M.; Iwasaki, Y. Neuroprotective Effects of the Free Radical Scavenger Edaravone (MCI-186) in Mice Permanent Focal Brain Ischemia. Brain Res. 2004, 1029, 200–206. [Google Scholar] [CrossRef]
  371. Abe, K.; Itoyama, Y.; Sobue, G.; Tsuji, S.; Aoki, M.; Doyu, M.; Hamada, C.; Kondo, K.; Yoneoka, T.; Akimoto, M.; et al. Confirmatory Double-Blind, Parallel-Group, Placebo-Controlled Study of Efficacy and Safety of Edaravone (MCI-186) in Amyotrophic Lateral Sclerosis Patients. Amyotroph. Lateral Scler. Front. Degener. 2014, 15, 610–617. [Google Scholar] [CrossRef]
  372. Mizuno, A.; Umemura, K.; Nakashima, M. Inhibitory Effect of MCI-186, a Free Radical Scavenger, on Cerebral Ischemia Following Rat Middle Cerebral Artery Occlusion. Gen. Pharmacol. 1998, 30, 575–578. [Google Scholar] [CrossRef]
  373. Yamamoto, T.; Yuki, S.; Watanabe, T.; Mitsuka, M.; Saito, K.I.; Kogure, K. Delayed Neuronal Death Prevented by Inhibition of Increased Hydroxyl Radical Formation in a Transient Cerebral Ischemia. Brain Res. 1997, 762, 240–242. [Google Scholar] [CrossRef]
  374. Takayasu, Y.; Nakaki, J.; Kawasaki, T.; Koda, K.; Ago, Y.; Baba, A.; Matsuda, T. Edaravone, a Radical Scavenger, Inhibits Mitochondrial Permeability Transition Pore in Rat Brain. J. Pharmacol. Sci. 2007, 103, 434–437. [Google Scholar] [CrossRef]
  375. Banno, M.; Mizuno, T.; Kato, H.; Zhang, G.; Kawanokuchi, J.; Wang, J.; Kuno, R.; Jin, S.; Takeuchi, H.; Suzumura, A. The Radical Scavenger Edaravone Prevents Oxidative Neurotoxicity Induced by Peroxynitrite and Activated Microglia. Neuropharmacology 2005, 48, 283–290. [Google Scholar] [CrossRef]
  376. Wang, Q.; Zhang, X.; Chen, S.; Zhang, X.; Zhang, S.; Youdium, M.; Le, W. Prevention of Motor Neuron Degeneration by Novel Iron Chelators in SOD1(G93A) Transgenic Mice of Amyotrophic Lateral Sclerosis. Neurodegener. Dis. 2011, 8, 310–321. [Google Scholar] [CrossRef] [PubMed]
  377. Miyamoto, N.; Maki, T.; Pham, L.D.; Hayakawa, K.; Seo, J.H.; Mandeville, E.T.; Mandeville, J.B.; Kim, K.W.; Lo, E.H.; Arai, K. Oxidative Stress Interferes with White Matter Renewal after Prolonged Cerebral Hypoperfusion in Mice. Stroke 2013, 44, 3516–3521. [Google Scholar] [CrossRef] [PubMed]
  378. Zhang, M.; Teng, C.H.; Wu, F.F.; Ge, L.Y.; Xiao, J.; Zhang, H.Y.; Chen, D.Q. Edaravone Attenuates Traumatic Brain Injury through Anti-Inflammatory and Anti-Oxidative Modulation. Exp. Ther. Med. 2019, 18, 467–474. [Google Scholar] [CrossRef] [PubMed]
  379. Aoki, M.; Warita, H.; Mizuno, H.; Suzuki, N.; Yuki, S.; Itoyama, Y. Feasibility Study for Functional Test Battery of SOD Transgenic Rat (H46R) and Evaluation of Edaravone, a Free Radical Scavenger. Brain Res. 2011, 1382, 321–325. [Google Scholar] [CrossRef] [PubMed]
  380. Ito, H.; Wate, R.; Zhang, J.; Ohnishi, S.; Kaneko, S.; Ito, H.; Nakano, S.; Kusaka, H. Treatment with Edaravone, Initiated at Symptom Onset, Slows Motor Decline and Decreases SOD1 Deposition in ALS Mice. Exp. Neurol. 2008, 213, 448–455. [Google Scholar] [CrossRef] [PubMed]
  381. Yoshino, H.; Kimura, A. Investigation of the Therapeutic Effects of Edaravone, a Free Radical Scavenger, on Amyotrophic Lateral Sclerosis (Phase II Study). Amyotroph. Lateral Scler. 2006, 7, 241–245. [Google Scholar] [CrossRef] [PubMed]
  382. Tanaka, M.; Akimoto, M.; Palumbo, J.; Sakata, T. A Double-Blind, Parallel-Group, Placebo-Controlled, 24-Week, Exploratory Study of Edaravone (MCI-186) for the Treatment of Advanced Amyotrophic Lateral Sclerosis (ALS) (P3.191). Neurology 2016, 86. [Google Scholar] [CrossRef]
  383. Hardiman, O.; van den Berg, L.H. Edaravone: A New Treatment for ALS on the Horizon? Lancet Neurol. 2017, 16, 490–491. [Google Scholar] [CrossRef]
  384. CPHI-Online. Treeway Announces Positive Data from Two Separate Phase I Tw001 Clinical Trials. Available online: https://www.cphi-online.com/treeway-announces-positive-data-from-two-separate-news038315.html (accessed on 23 January 2024).
  385. Genge, A.; Pattee, G.L.; Sobue, G.; Aoki, M.; Yoshino, H.; Couratier, P.; Lunetta, C.; Petri, S.; Selness, D.; Bidani, S.; et al. Oral Edaravone Demonstrated a Favorable Safety Profile in Patients with Amyotrophic Lateral Sclerosis after 48 Weeks of Treatment. Muscle Nerve 2023, 67, 124–129. [Google Scholar] [CrossRef]
  386. Gao, M.; Zhu, L.; Chang, J.; Cao, T.; Song, L.; Wen, C.; Chen, Y.; Zhuo, Y.; Chen, F. Safety and Efficacy of Edaravone in Patients with Amyotrophic Lateral Sclerosis: A Systematic Review and Meta-Analysis. Clin. Drug Investig. 2023, 43, 1–11. [Google Scholar] [CrossRef]
  387. Tan, D.X.; Reiter, R.J.; Manchester, L.C.; Yan, M.T.; El-Sawi, M.; Sainz, R.M.; Mayo, J.C.; Kohen, R.; Allegra, M.; Hardeland, R. Chemical and Physical Properties and Potential Mechanisms: Melatonin as a Broad Spectrum Antioxidant and Free Radical Scavenger. Curr. Top. Med. Chem. 2002, 2, 181–197. [Google Scholar] [CrossRef]
  388. Ganie, S.A.; Dar, T.A.; Bhat, A.H.; Dar, K.B.; Anees, S.; Zargar, M.A.; Masood, A. Melatonin: A Potential Anti-Oxidant Therapeutic Agent for Mitochondrial Dysfunctions and Related Disorders. Rejuvenation Res. 2016, 19, 21–40. [Google Scholar] [CrossRef] [PubMed]
  389. Pandi-Perumal, S.R.; BaHammam, A.S.; Brown, G.M.; Spence, D.W.; Bharti, V.K.; Kaur, C.; Hardeland, R.; Cardinali, D.P. Melatonin Antioxidative Defense: Therapeutical Implications for Aging and Neurodegenerative Processes. Neurotox. Res. 2013, 23, 267–300. [Google Scholar] [CrossRef] [PubMed]
  390. Leon, J.; Acuna-Castroviejo, D.; Escames, G.; Tan, D.X.; Reiter, R.J. Melatonin Mitigates Mitochondrial Malfunction. J. Pineal Res. 2005, 38, 1–9. [Google Scholar] [CrossRef] [PubMed]
  391. Zhang, Y.; Cook, A.; Kim, J.; Baranov, S.V.; Jiang, J.; Smith, K.; Cormier, K.; Bennett, E.; Browser, R.P.; Day, A.L.; et al. Melatonin Inhibits the Caspase-1/Cytochrome C/Caspase-3 Cell Death Pathway, Inhibits MT1 Receptor Loss and Delays Disease Progression in a Mouse Model of Amyotrophic Lateral Sclerosis. Neurobiol. Dis. 2013, 55, 26–35. [Google Scholar] [CrossRef] [PubMed]
  392. Dardiotis, E.; Panayiotou, E.; Feldman, M.L.; Hadjisavvas, A.; Malas, S.; Vonta, I.; Hadjigeorgiou, G.; Kyriakou, K.; Kyriakides, T. Intraperitoneal Melatonin Is Not Neuroprotective in the G93ASOD1 Transgenic Mouse Model of Familial ALS and May Exacerbate Neurodegeneration. Neurosci. Lett. 2013, 548, 170–175. [Google Scholar] [CrossRef]
  393. Murphy, M.P.; Smith, R.A. Targeting Antioxidants to Mitochondria by Conjugation to Lipophilic Cations. Annu. Rev. Pharmacol. Toxicol. 2007, 47, 629–656. [Google Scholar] [CrossRef] [PubMed]
  394. Finichiu, P.G.; James, A.M.; Larsen, L.; Smith, R.A.; Murphy, M.P. Mitochondrial Accumulation of a Lipophilic Cation Conjugated to an Ionisable Group Depends on Membrane Potential, pH Gradient and pK(a): Implications for the Design of Mitochondrial Probes and Therapies. J. Bioenerg. Biomembr. 2013, 45, 165–173. [Google Scholar] [CrossRef] [PubMed]
  395. Kelso, G.F.; Porteous, C.M.; Coulter, C.V.; Hughes, G.; Porteous, W.K.; Ledgerwood, E.C.; Smith, R.A.; Murphy, M.P. Selective Targeting of a Redox-Active Ubiquinone to Mitochondria within Cells: Antioxidant and Antiapoptotic Properties. J. Biol. Chem. 2001, 276, 4588–4596. [Google Scholar] [CrossRef] [PubMed]
  396. Cocheme, H.M.; Kelso, G.F.; James, A.M.; Ross, M.F.; Trnka, J.; Mahendiran, T.; Asin-Cayuela, J.; Blaikie, F.H.; Manas, A.R.; Porteous, C.M.; et al. Mitochondrial Targeting of Quinones: Therapeutic Implications. Mitochondrion 2007, 7, S94–S102. [Google Scholar] [CrossRef]
  397. Smith, R.A.; Murphy, M.P. Animal and Human Studies with the Mitochondria-Targeted Antioxidant MitoQ. Ann. N. Y. Acad. Sci. 2010, 1201, 96–103. [Google Scholar] [CrossRef]
  398. Solesio, M.E.; Prime, T.A.; Logan, A.; Murphy, M.P.; Del Mar Arroyo-Jimenez, M.; Jordan, J.; Galindo, M.F. The Mitochondria-Targeted Anti-Oxidant MitoQ Reduces Aspects of Mitochondrial Fission in the 6-OHDA Cell Model of Parkinson’s Disease. Biochim. Biophys. Acta 2013, 1832, 174–182. [Google Scholar] [CrossRef]
  399. Ghosh, A.; Chandran, K.; Kalivendi, S.V.; Joseph, J.; Antholine, W.E.; Hillard, C.J.; Kanthasamy, A.; Kanthasamy, A.; Kalyanaraman, B. Neuroprotection by a Mitochondria-Targeted Drug in a Parkinson’s Disease Model. Free Radic. Biol. Med. 2010, 49, 1674–1684. [Google Scholar] [CrossRef]
  400. McManus, M.J.; Murphy, M.P.; Franklin, J.L. The Mitochondria-Targeted Antioxidant MitoQ Prevents Loss of Spatial Memory Retention and Early Neuropathology in a Transgenic Mouse Model of Alzheimer’s Disease. J. Neurosci. 2011, 31, 15703–15715. [Google Scholar] [CrossRef]
  401. Manczak, M.; Mao, P.; Calkins, M.J.; Cornea, A.; Reddy, A.P.; Murphy, M.P.; Szeto, H.H.; Park, B.; Reddy, P.H. Mitochondria-Targeted Antioxidants Protect against Amyloid-Beta Toxicity in Alzheimer’s Disease Neurons. J. Alzheimer’s Dis. 2010, 20 (Suppl. S2), S609–S631. [Google Scholar] [CrossRef]
  402. Snow, B.J.; Rolfe, F.L.; Lockhart, M.M.; Frampton, C.M.; O’Sullivan, J.D.; Fung, V.; Smith, R.A.; Murphy, M.P.; Taylor, K.M.; Protect Study, G. A Double-Blind, Placebo-Controlled Study to Assess the Mitochondria-Targeted Antioxidant MitoQ as a Disease-Modifying Therapy in Parkinson’s Disease. Mov. Disord. 2010, 25, 1670–1674. [Google Scholar] [CrossRef]
  403. Cassina, P.; Cassina, A.; Pehar, M.; Castellanos, R.; Gandelman, M.; de Leon, A.; Robinson, K.M.; Mason, R.P.; Beckman, J.S.; Barbeito, L.; et al. Mitochondrial Dysfunction in SOD1G93A-Bearing Astrocytes Promotes Motor Neuron Degeneration: Prevention by Mitochondrial-Targeted Antioxidants. J. Neurosci. 2008, 28, 4115–4122. [Google Scholar] [CrossRef]
  404. Miquel, E.; Cassina, A.; Martinez-Palma, L.; Souza, J.M.; Bolatto, C.; Rodriguez-Bottero, S.; Logan, A.; Smith, R.A.; Murphy, M.P.; Barbeito, L.; et al. Neuroprotective Effects of the Mitochondria-Targeted Antioxidant MitoQ in a Model of Inherited Amyotrophic Lateral Sclerosis. Free Radic. Biol. Med. 2014, 70, 204–213. [Google Scholar] [CrossRef] [PubMed]
  405. Gane, E.J.; Weilert, F.; Orr, D.W.; Keogh, G.F.; Gibson, M.; Lockhart, M.M.; Frampton, C.M.; Taylor, K.M.; Smith, R.A.; Murphy, M.P. The Mitochondria-Targeted Anti-Oxidant Mitoquinone Decreases Liver Damage in a Phase II Study of Hepatitis C Patients. Liver Int. 2010, 30, 1019–1026. [Google Scholar] [CrossRef] [PubMed]
  406. Zielonka, J.; Joseph, J.; Sikora, A.; Hardy, M.; Ouari, O.; Vasquez-Vivar, J.; Cheng, G.; Lopez, M.; Kalyanaraman, B. Mitochondria-Targeted Triphenylphosphonium-Based Compounds: Syntheses, Mechanisms of Action, and Therapeutic and Diagnostic Applications. Chem. Rev. 2017, 117, 10043–10120. [Google Scholar] [CrossRef] [PubMed]
  407. Pehar, M.; Vargas, M.R.; Robinson, K.M.; Cassina, P.; Diaz-Amarilla, P.J.; Hagen, T.M.; Radi, R.; Barbeito, L.; Beckman, J.S. Mitochondrial Superoxide Production and Nuclear Factor Erythroid 2-Related Factor 2 Activation in p75 Neurotrophin Receptor-Induced Motor Neuron Apoptosis. J. Neurosci. 2007, 27, 7777–7785. [Google Scholar] [CrossRef] [PubMed]
  408. Xu, X.L.; Yi, F.; Pan, H.Z.; Duan, S.L.; Ding, Z.C.; Yuan, G.H.; Qu, J.; Zhang, H.C.; Liu, G.H. Progress and Prospects in Stem Cell Therapy. Acta Pharmacol. Sin. 2013, 34, 741–746. [Google Scholar] [CrossRef] [PubMed]
  409. Ohnuki, M.; Takahashi, K. Present and Future Challenges of Induced Pluripotent Stem Cells. Philos. Trans. R. Soc. B Biol. Sci. 2015, 370, 20140367. [Google Scholar] [CrossRef] [PubMed]
  410. Keep, M.; Elmer, E.; Fong, K.S.; Csiszar, K. Intrathecal Cyclosporin Prolongs Survival of Late-Stage ALS Mice. Brain Res. 2001, 894, 327–331. [Google Scholar] [CrossRef] [PubMed]
  411. Karlsson, J.; Fong, K.S.; Hansson, M.J.; Elmer, E.; Csiszar, K.; Keep, M.F. Life Span Extension and Reduced Neuronal Death after Weekly Intraventricular Cyclosporin Injections in the G93A Transgenic Mouse Model of Amyotrophic Lateral Sclerosis. J. Neurosurg. 2004, 101, 128–137. [Google Scholar] [CrossRef]
  412. Kirkinezos, I.G.; Hernandez, D.; Bradley, W.G.; Moraes, C.T. An ALS Mouse Model with a Permeable Blood-Brain Barrier Benefits from Systemic Cyclosporine A Treatment. J. Neurochem. 2004, 88, 821–826. [Google Scholar] [CrossRef]
  413. Martin, L.J. The Mitochondrial Permeability Transition Pore: A Molecular Target for Amyotrophic Lateral Sclerosis Therapy. Biochim. Biophys. Acta 2010, 1802, 186–197. [Google Scholar] [CrossRef] [PubMed]
  414. Martin, L.J. Olesoxime, a Cholesterol-Like Neuroprotectant for the Potential Treatment of Amyotrophic Lateral Sclerosis. IDrugs 2010, 13, 568–580. [Google Scholar]
  415. Bordet, T.; Buisson, B.; Michaud, M.; Drouot, C.; Galea, P.; Delaage, P.; Akentieva, N.P.; Evers, A.S.; Covey, D.F.; Ostuni, M.A.; et al. Identification and Characterization of Cholest-4-En-3-One, Oxime (TRO19622), a Novel Drug Candidate for Amyotrophic Lateral Sclerosis. J. Pharmacol. Exp. Ther. 2007, 322, 709–720. [Google Scholar] [CrossRef] [PubMed]
  416. Sunyach, C.; Michaud, M.; Arnoux, T.; Bernard-Marissal, N.; Aebischer, J.; Latyszenok, V.; Gouarne, C.; Raoul, C.; Pruss, R.M.; Bordet, T.; et al. Olesoxime Delays Muscle Denervation, Astrogliosis, Microglial Activation and Motoneuron Death in an ALS Mouse Model. Neuropharmacology 2012, 62, 2346–2352. [Google Scholar] [CrossRef] [PubMed]
  417. Lenglet, T.; Lacomblez, L.; Abitbol, J.L.; Ludolph, A.; Mora, J.S.; Robberecht, W.; Shaw, P.J.; Pruss, R.M.; Cuvier, V.; Meininger, V.; et al. A Phase II–III Trial of Olesoxime in Subjects with Amyotrophic Lateral Sclerosis. Eur. J. Neurol. 2014, 21, 529–536. [Google Scholar] [CrossRef]
  418. Scott, S.; Kranz, J.E.; Cole, J.; Lincecum, J.M.; Thompson, K.; Kelly, N.; Bostrom, A.; Theodoss, J.; Al-Nakhala, B.M.; Vieira, F.G.; et al. Design, Power, and Interpretation of Studies in the Standard Murine Model of ALS. Amyotroph. Lateral Scler. 2008, 9, 4–15. [Google Scholar] [CrossRef]
  419. Yang, Y.M.; Gupta, S.K.; Kim, K.J.; Powers, B.E.; Cerqueira, A.; Wainger, B.J.; Ngo, H.D.; Rosowski, K.A.; Schein, P.A.; Ackeifi, C.A.; et al. A Small Molecule Screen in Stem-Cell-Derived Motor Neurons Identifies a Kinase Inhibitor as a Candidate Therapeutic for ALS. Cell Stem Cell 2013, 12, 713–726. [Google Scholar] [CrossRef]
  420. Martin, L.J.; Fancelli, D.; Wong, M.; Niedzwiecki, M.; Ballarini, M.; Plyte, S.; Chang, Q. GNX-4728, a Novel Small Molecule Drug Inhibitor of Mitochondrial Permeability Transition, Is Therapeutic in a Mouse Model of Amyotrophic Lateral Sclerosis. Front. Cell. Neurosci. 2014, 8, 433. [Google Scholar] [CrossRef]
  421. Parkinson Study, G. A Controlled Trial of Rasagiline in Early Parkinson Disease: The Tempo Study. Arch. Neurol. 2002, 59, 1937–1943. [Google Scholar] [CrossRef]
  422. Olanow, C.W.; Rascol, O.; Hauser, R.; Feigin, P.D.; Jankovic, J.; Lang, A.; Langston, W.; Melamed, E.; Poewe, W.; Stocchi, F.; et al. A Double-Blind, Delayed-Start Trial of Rasagiline in Parkinson’s Disease. N. Engl. J. Med. 2009, 361, 1268–1278. [Google Scholar] [CrossRef]
  423. Youdim, M.B.; Weinstock, M. Molecular Basis of Neuroprotective Activities of Rasagiline and the Anti-Alzheimer Drug TV3326 [(N-propargyl-(3R)aminoindan-5-YL)-ethyl methyl carbamate]. Cell. Mol. Neurobiol. 2001, 21, 555–573. [Google Scholar] [CrossRef] [PubMed]
  424. Maruyama, W.; Akao, Y.; Youdim, M.B.; Davis, B.A.; Naoi, M. Transfection-Enforced Bcl-2 Overexpression and an Anti-Parkinson Drug, Rasagiline, Prevent Nuclear Accumulation of Glyceraldehyde-3-Phosphate Dehydrogenase Induced by an Endogenous Dopaminergic Neurotoxin, N-methyl(R)salsolinol. J. Neurochem. 2001, 78, 727–735. [Google Scholar] [CrossRef] [PubMed]
  425. Maruyama, W.; Youdim, M.B.; Naoi, M. Antiapoptotic Properties of Rasagiline, N-Propargylamine-1(R)-Aminoindan, and Its Optical (S)-isomer, TV1022. Ann. N. Y. Acad. Sci. 2001, 939, 320–329. [Google Scholar] [CrossRef] [PubMed]
  426. Youdim, M.B.; Amit, T.; Falach-Yogev, M.; Bar Am, O.; Maruyama, W.; Naoi, M. The Essentiality of Bcl-2, PKC and Proteasome-Ubiquitin Complex Activations in the Neuroprotective-Antiapoptotic Action of the Anti-Parkinson Drug, Rasagiline. Biochem. Pharmacol. 2003, 66, 1635–1641. [Google Scholar] [CrossRef] [PubMed]
  427. Waibel, S.; Reuter, A.; Malessa, S.; Blaugrund, E.; Ludolph, A.C. Rasagiline Alone and in Combination with Riluzole Prolongs Survival in an ALS Mouse Model. J. Neurol. 2004, 251, 1080–1084. [Google Scholar] [CrossRef]
  428. Macchi, Z.; Wang, Y.; Moore, D.; Katz, J.; Saperstein, D.; Walk, D.; Simpson, E.; Genge, A.; Bertorini, T.; Fernandes, J.A.; et al. A Multi-Center Screening Trial of Rasagiline in Patients with Amyotrophic Lateral Sclerosis: Possible Mitochondrial Biomarker Target Engagement. Amyotroph. Lateral Scler. Front. Degener. 2015, 16, 345–352. [Google Scholar] [CrossRef]
  429. Statland, J.M.; Moore, D.; Wang, Y.; Walsh, M.; Mozaffar, T.; Elman, L.; Nations, S.P.; Mitsumoto, H.; Fernandes, J.A.; Saperstein, D.; et al. Rasagiline for Amyotrophic Lateral Sclerosis: A Randomized, Controlled Trial. Muscle Nerve 2019, 59, 201–207. [Google Scholar] [CrossRef]
  430. Ludolph, A.C.; Schuster, J.; Dorst, J.; Dupuis, L.; Dreyhaupt, J.; Weishaupt, J.H.; Kassubek, J.; Weiland, U.; Petri, S.; Meyer, T.; et al. Safety and Efficacy of Rasagiline as an Add-on Therapy to Riluzole in Patients with Amyotrophic Lateral Sclerosis: A Randomised, Double-Blind, Parallel-Group, Placebo-Controlled, Phase 2 Trial. Lancet Neurol. 2018, 17, 681–688. [Google Scholar] [CrossRef]
Figure 1. Overview of mitochondrial functions. Mitochondrial bioenergetics is driven by the oxidation of different substrates and is stimulated by calcium. Flux of electrons through the electron transport chain creates a transmembrane proton gradient of about 160 mV in the resting state, which fuels Adenosine Triphosphate (ATP) synthesis in the mitochondrial matrix. Leakage of electrons in some bioenergetic reactions generates reactive oxygen species (ROS) that are involved in important cellular signaling processes. Abbreviations: ADP: Adenosine Diphosphate; CI: Complex I; CII: Complex II; CIII: Complex III; CIV: Complex IV; Cyt c: Cytochrome c; ETF: Electron Transfer Flavoprotein; FAD: Flavin Adenine Dinucleotide; IMM: Inner Mitochondrial Membrane; MCU: Mitochondrial Calcium Uniporter; MPC: Mitochondrial Pyruvate Carrier; ΔΨm: Mitochondrial Transmembrane Electrochemical Potential; NAD: β-Nicotinamide adenine dinucleotide; NADH: β-Nicotinamide adenine dinucleotide 2′-phosphate reduced form.
Figure 1. Overview of mitochondrial functions. Mitochondrial bioenergetics is driven by the oxidation of different substrates and is stimulated by calcium. Flux of electrons through the electron transport chain creates a transmembrane proton gradient of about 160 mV in the resting state, which fuels Adenosine Triphosphate (ATP) synthesis in the mitochondrial matrix. Leakage of electrons in some bioenergetic reactions generates reactive oxygen species (ROS) that are involved in important cellular signaling processes. Abbreviations: ADP: Adenosine Diphosphate; CI: Complex I; CII: Complex II; CIII: Complex III; CIV: Complex IV; Cyt c: Cytochrome c; ETF: Electron Transfer Flavoprotein; FAD: Flavin Adenine Dinucleotide; IMM: Inner Mitochondrial Membrane; MCU: Mitochondrial Calcium Uniporter; MPC: Mitochondrial Pyruvate Carrier; ΔΨm: Mitochondrial Transmembrane Electrochemical Potential; NAD: β-Nicotinamide adenine dinucleotide; NADH: β-Nicotinamide adenine dinucleotide 2′-phosphate reduced form.
Cells 13 00248 g001
Figure 2. Dysregulated mitochondrial trafficking in ALS Models. Anomalies in mitochondrial trafficking are a prevalent phenomenon across various ALS models. Notably, both anterograde and retrograde transport processes are compromised in this context. Intriguingly, emerging research suggests a temporal sequence in which retrograde transport disruption precedes anterograde transport impairment. Dysfunction in retrograde transport is intimately associated with mitophagy deficits, resulting in the inefficient removal of dysfunctional mitochondria in TDP 43 ALS mutant mice. Conversely, cortical neurons from the E18 embryos of ALS SOD1 mutant rats exhibit alterations in anterograde transport linked to a reduction in Miro expression, ultimately leading to mitochondrial transport stagnation. These perturbations in both transport pathways culminate in abnormal mitochondrial clustering, diminished Ca2+ buffering capacity, and compromised ATP production. In extreme cases, these abnormalities can trigger axonal stress, ultimately contributing to cell death and neurodegeneration.
Figure 2. Dysregulated mitochondrial trafficking in ALS Models. Anomalies in mitochondrial trafficking are a prevalent phenomenon across various ALS models. Notably, both anterograde and retrograde transport processes are compromised in this context. Intriguingly, emerging research suggests a temporal sequence in which retrograde transport disruption precedes anterograde transport impairment. Dysfunction in retrograde transport is intimately associated with mitophagy deficits, resulting in the inefficient removal of dysfunctional mitochondria in TDP 43 ALS mutant mice. Conversely, cortical neurons from the E18 embryos of ALS SOD1 mutant rats exhibit alterations in anterograde transport linked to a reduction in Miro expression, ultimately leading to mitochondrial transport stagnation. These perturbations in both transport pathways culminate in abnormal mitochondrial clustering, diminished Ca2+ buffering capacity, and compromised ATP production. In extreme cases, these abnormalities can trigger axonal stress, ultimately contributing to cell death and neurodegeneration.
Cells 13 00248 g002
Figure 3. Mitochondria-targeted therapies for ALS featuring (A) metabolic enhancers, (B) antioxidants, and (C) anti-apoptotic agents. Abbreviations: AcCoA: acetyl coenzyme A; CK: Creatine kinase; ETC: electron transport chain; FAO: fatty acid beta-oxidation; mPTP: mitochondrial permeability transition pore; PDH, pyruvate dehydrogenase; PDK, pyruvate dehydrogenase kinase; PDP, pyruvate dehydrogenase phosphatase; ROS: reactive oxygen species; RNS: reactive nitrogen species; TCA: tricarboxylic acid cycle.
Figure 3. Mitochondria-targeted therapies for ALS featuring (A) metabolic enhancers, (B) antioxidants, and (C) anti-apoptotic agents. Abbreviations: AcCoA: acetyl coenzyme A; CK: Creatine kinase; ETC: electron transport chain; FAO: fatty acid beta-oxidation; mPTP: mitochondrial permeability transition pore; PDH, pyruvate dehydrogenase; PDK, pyruvate dehydrogenase kinase; PDP, pyruvate dehydrogenase phosphatase; ROS: reactive oxygen species; RNS: reactive nitrogen species; TCA: tricarboxylic acid cycle.
Cells 13 00248 g003
Table 1. Clinical trials discussed in this review.
Table 1. Clinical trials discussed in this review.
Clinical Trial IDStudy NameConditionPhase/TypeSponsorStarting DateEnd DateNumber of Patients
NCT01257581Phase 2 Selection Trial of High Dosage Creatine and Two Dosages of Tamoxifen in Amyotrophic Lateral Sclerosis (ALS)Amyotrophic Lateral Sclerosis (ALS)Phase 2Nazem Atassi, Massachusetts General Hospital2011-032013-0260
NCT02306590Efficacy, Safety and Tolerability of High Lipid and Calorie Supplementation in Amyotrophic Lateral SclerosisAmyotrophic Lateral Sclerosis (ALS)Prospective, multicenter, randomized, stratified, parallel-group, double-blind trialAlbert Christian Ludolph, Prof., University of Ulm2015-022018-09207
NCT04172792Safety and Tolerability of Fat-rich vs. Carbohydrate-rich High-caloric Food Supplements in Patients with Amyotrophic Lateral Sclerosis (ALS)Amyotrophic Lateral Sclerosis (ALS)Phase 1Albert Christian Ludolph, Prof., University of Ulm2019-112021-0464
NCT00243932Clinical Trial of High Dose CoQ10 in ALSAmyotrophic Lateral Sclerosis (ALS)Phase 2Columbia University2005-042008-03185
NCT01492686Efficacy and Safety Study of MCI-186 for Treatment of Patients with Amyotrophic Lateral Sclerosis (ALS) 2Amyotrophic Lateral Sclerosis (ALS)Phase 3Mitsubishi Tanabe Pharma Corporation2011-122014-10137
NCT04165824Safety Study of Oral Edaravone Administered in Subjects with ALSAmyotrophic Lateral Sclerosis (ALS)Phase 3Mitsubishi Tanabe Pharma America Inc.2019-112021-10185
NCT01281189A Randomized, Double-Blind, Placebo-Controlled, Multi-Center Study of the Safety and Efficacy of Dexpramipexole in Subjects with Amyotrophic Lateral SclerosisAmyotrophic Lateral Sclerosis (ALS)Phase 3Knopp Biosciences2011-032012-11942
NCT00329056A Double-Blind, Prospective, Randomized Comparison of 2 Doses of MitoQ and Placebo for the Treatment of Patients with Parkinson’s DiseaseParkinson’s Disease (PD)Double-Blind, Prospective, Randomized ComparisonAntipodean Pharmaceuticals, Inc.2006-052007-11128
NCT00047723Minocycline to Treat Amyotrophic Lateral SclerosisAmyotrophic Lateral Sclerosis (ALS)Phase 3National Institute of Neurological Disorders and Stroke (NINDS)2003-012007-01400
NCT01232738A Multi-Center Controlled Screening Trial of Safety and Efficacy of Rasagiline in Subjects with Amyotrophic Lateral Sclerosis (ALS)Amyotrophic Lateral Sclerosis (ALS)Phase 2Yunxia Wang, MD, University of Kansas Medical Center2011-122013-0536
NCT01786603Rasagiline in Subjects with Amyotrophic Lateral Sclerosis (ALS)Amyotrophic Lateral Sclerosis (ALS)Phase 2Richard Barohn, MD, University of Kansas Medical Center2013-112016-0780
NCT01879241Efficacy, Safety, and Tolerability Study of 1 mg Rasagiline in Patients with Amyotrophic Lateral Sclerosis (ALS) Receiving Standard Therapy (Riluzole)-An AMG Trial with a Market Authorized SubstanceAmyotrophic Lateral Sclerosis (ALS)Phase 2Albert Christian Ludolph, Prof., University of Ulm2013-062016-08252
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cunha-Oliveira, T.; Montezinho, L.; Simões, R.F.; Carvalho, M.; Ferreiro, E.; Silva, F.S.G. Mitochondria: A Promising Convergent Target for the Treatment of Amyotrophic Lateral Sclerosis. Cells 2024, 13, 248. https://doi.org/10.3390/cells13030248

AMA Style

Cunha-Oliveira T, Montezinho L, Simões RF, Carvalho M, Ferreiro E, Silva FSG. Mitochondria: A Promising Convergent Target for the Treatment of Amyotrophic Lateral Sclerosis. Cells. 2024; 13(3):248. https://doi.org/10.3390/cells13030248

Chicago/Turabian Style

Cunha-Oliveira, Teresa, Liliana Montezinho, Rui F. Simões, Marcelo Carvalho, Elisabete Ferreiro, and Filomena S. G. Silva. 2024. "Mitochondria: A Promising Convergent Target for the Treatment of Amyotrophic Lateral Sclerosis" Cells 13, no. 3: 248. https://doi.org/10.3390/cells13030248

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