Editorial Open Access
Copyright ©2010 Baishideng Publishing Group Co., Limited. All rights reserved.
World J Biol Chem. May 26, 2010; 1(5): 62-68
Published online May 26, 2010. doi: 10.4331/wjbc.v1.i5.62
State of the art and the dark side of amyotrophic lateral sclerosis
Antonio Musarò, Department of Histology and Medical Embryology, Institute Pasteur Cenci-Bolognetti, IIM, Sapienza University of Rome, Rome 00161, Italy
Author contributions: Musarò A wrote the paper.
Supported by Seventh Framework Programme-Myoage, Telethon, Fondazione Roma, AFM, MIUR and ASI
Correspondence to: Antonio Musarò, PhD, Department of Histology and Medical Embryology, Institute Pasteur Cenci-Bolognetti, IIM, Sapienza University of Rome, Via A. Scarpa, 14, Rome 00161, Italy. antonio.musaro@uniroma1.it
Telephone: +39-6-49766956 Fax: +39-6-4462854
Received: May 4, 2010
Revised: May 22, 2010
Accepted: May 24, 2010
Published online: May 26, 2010

Abstract

Amyotrophic lateral sclerosis (ALS) is a disorder that involves the degeneration of motor neurons, muscle atrophy, and paralysis. In a few familiar forms of ALS, mutations in the superoxide dismutase-1 (SOD1) gene have been held responsible for the degeneration of motor neurons. Nevertheless, after the discovery of the SOD1 mutations no consensus has emerged as to which cells, tissues and pathways are primarily implicated in the pathogenic events that lead to ALS. Ubiquitous overexpression of mutant SOD1 in transgenic animals recapitulates the pathological features of ALS. However, the toxicity of mutant SOD1 is not necessarily limited to the central nervous system. Views about ALS pathogenesis are now enriched by the recent discovery of mutations in a pair of DNA/RNA-binding proteins called TDP-43 and FUS/TLS as causes of familial and sporadic forms of ALS. Although the steps that lead to the pathological state are well defined, several fundamental issues are still controversial: are the motor neurons the first direct targets of ALS; and what is the contribution of non-neuronal cells, if any, to the pathogenesis of ALS? The state of the art of ALS pathogenesis and the open questions are discussed in this review.

Key Words: Amyotrophic lateral sclerosis, Neurodegenerative disease, Muscle wasting, Oxidative stress, Excitotoxicity, Protein aggregation, Mitochondrial dysfunction, Insulin-like growth factor 1



INTRODUCTION

Amyotrophic lateral sclerosis (ALS) is a fatal neurodegenerative disorder that mainly affects pyramidal neurons in the motor cortex and lower motor neurons that originate in the brainstem and spinal cord. The disease was first described in 1869 by the French neurobiologist and physician Charcot et al[1] who linked the symptoms of ALS to a group of nerves specifically affected by the disease; the motor neurons that originate in the spinal cord. The name of the disease reflects the different tissue compartments that are severely affected. In particular, “amyotrophic” refers to the atrophy of muscle fibers and loss of muscle mass; “lateral” refers to the nerve tracks that run down both sides of the spinal cord, where many neurons affected by ALS are found; and “sclerosis” refers to the scar tissue that remains following degeneration of the nerves.

The most typical feature of this progressive lethal disease is the degeneration of cortical, bulbar and spinal motor neurons, except for the neurons that control the bladder, and the oculomotor neurons[2]. This leads to muscle weakness, fasciculations, muscle atrophy, speech and swallowing disabilities, progressive paralysis, and death caused by respiratory failure.

ALS is epidemiologically classified into two forms: sporadic (90%-95%) and familial (5%-10%)[3]. Among the familial cases, approximately 20% are caused by dominantly inherited mutations in the Cu/Zn superoxide dismutase-1 (SOD1) protein[4]. Although the major part of these mutations are missense, deletions and insertions have been observed at the level of the C-terminal region of the protein[5]. The best function of the metalloenzyme SOD1 is to convert superoxide, a toxic by-product of mitochondrial oxidative phosphorylation, to water or hydrogen peroxide. Initially, it has been suggested that mutation in the SOD1 gene that lead to a decrease in the protein enzymatic activity (loss of function hypothesis). However, subsequent studies have clarified that mutant SOD1 possesses a neurotoxic property (gain of function hypothesis) that is responsible for the pathogenic mechanism of the disease[3,6-8]. Indeed, the finding that overexpression of mutant SOD1 in transgenic mice recapitulates several clinical feature of ALS disease, even in the presence of endogenous mouse SOD1, has led to the conclusion that the disease results from a toxic gain of function[9,10]. In addition to SOD1 mutations, other gene defects have been reported to cause ALS[3]. In particular, mutations in more than 50 different human genes are implicated in the pathogenesis of motor neuron cell death[11].

Among these, mutation of synaptobrevin/vesicle-associated membrane protein-associated protein B (VAPB) gene causes familial forms of motor neuron diseases, including a rare, slowly progressing form of ALS (ALS8), typical severe ALS with rapid progression, as well as a late-onset form of spinal muscular atrophy[12-14]. Several functions have been ascribed to VAPB proteins, including membrane trafficking, cytoskeleton association, and membrane docking interactions for cytoplasmic factors.

More recently, Fasana et al[15] have demonstrated that ALS-linked VAPB mutation causes dramatic endoplasmic reticulum (ER) restructuring that might underlie its pathogenicity in motor neurons. In addition, Langou et al[16] have provided evidence that ER stress and impaired homeostatic regulation of calcium are implicated in the death process. It has also been demonstrated that VAPBP56S transgenic mice develop cytoplasmic accumulations of TDP-43, a DNA/RNA-binding protein, which suggests a link between abnormal VAPBP56S function and TDP-43 mis-localization[17].

The research field of ALS has been improved by the recent discovery of two new mutations in the DNA/RNA-binding proteins TDP-43 and FUS/TLS, which have triggered new interest in ALS pathogenesis[18-21]. However, even in these cases, where a well-defined mutation has been linked to the disease, a clear correlation between the genetic defect and the pathophysiology of the disease has not yet been disclosed. Notably, alteration in TDP-43 sub-localization is also associated with inclusion body myositis[22], which suggests that this nucleic-acid-binding protein plays a pathologic role in skeletal muscle.

Although most efforts have been aimed at defining the potential genes and pathways associated with motor neuron degeneration and to understand ALS pathogenesis, no consensus has yet emerged as to the primary toxicity of SOD1 mutations. The primary causes of ALS are therefore still unknown and no effective and decisive treatments are available.

PATHOGENIC MECHANISMS OF ALS

Several pathogenic mechanisms have been proposed to account for ALS, including glutamate-induced excitotoxicity, oxidative stress, protein aggregation, and mitochondrial dysfunction (Figure 1).

Figure 1
Figure 1 Schematic representation of potential mechanisms associated with the pathogenesis of amyotrophic lateral sclerosis (ALS). Mutations in superoxide dismutase-1 (SOD1) and/or other genes and environmental factors are responsible for the activation of the pathogenic mechanisms that lead to ALS. Several pathogenic mechanisms have been proposed. (1) Glutamate-induced excitotoxicity: overstimulation of neurons by glutamate causes the accumulation of calcium ions in cellular compartments, which leads to activation of apoptotic pathways; (2) Oxidative stress that is caused by an imbalance between the production of reactive oxygen species and antioxidant defenses; (3) Protein aggregation: oxidative stress and mitochondrial alteration could be responsible for protein aggregation; mutant SOD1 protein or other ALS-related genes tend to be misfolded or form aggregates that, in turn, trigger a toxic cascade that leads to neuronal degeneration; (4) Mitochondrial dysfunction, which leads to oxidative stress, decreased activity of respiratory complexes, decreased ATP levels and cytochrome c release; and (5) Deficit in neurotrophic growth factors, including insulin-like growth factor 1, and activation of proteolytic systems.
Glutamate-induced excitotoxicity

Several lines of evidence implicate glutamatergic toxicity as a contributory factor in the neuronal injury in ALS[23]. Excessive stimulation of glutamate receptors causes excitotoxicity. Under physiological conditions, excitotoxicity is prevented by rapid binding and clearance of synapse-released glutamate by high-affinity, Na(+)-dependent glutamate transporters. The excitatory amino acid transporters (EAATs) play a crucial role in the regulation of extracellular glutamate concentration, which stimulates the reuptake of glutamate and therefore maintains a physiological concentration, of about 1-3 μmol/L, in the synaptic cleft[24]. Five different subtypes of EAAT (1-5) have been described and they are differently distributed throughout the brain[24]. Notably, EAAT2 is the major glutamate transporter and is widely distributed in the human central nervous system. The selective localization of EAAT2 in glial cells has implications for its excitotoxic mechanisms.

How does glutamate induces excitotoxicity? It has been demonstrated that glutamate transport is diminished in sporadic and familial ALS patients, due to a selective loss of the astroglial glutamate transporter, EAAT2, in the motor cortex and spinal cord[25]. Alterations in this protein might interfere with the normal clearance, which allows glutamate to remain in the environment and continue to activate the receptors. Once activated, the receptors cause calcium influx that cells are not able to buffer because of an insufficient number of calcium-binding proteins. Altered concentration of calcium activates apoptotic pathways, which leads to motor neuron death and degeneration.

In addition, alteration in the antioxidant enzyme SOD1 induces oxidative damage to the intracellular C-terminal part of EAAT2, which results in decreased glutamate transport[26]. Although glutamate-induced excitotoxicity represents a pathogenic event of ALS, it remains to be determined whether these changes represent a primary defect that is responsible for motor neuron degeneration, or are the result of ALS.

Oxidative stress and ALS

The observation that approximately 20% of familial ALS is caused by mutations in the SOD1 gene has placed oxidative stress as one of the potential pathogenic events that are associated with the disease. Several lines of evidence suggest a primary role of oxidative stress in the pathogenesis of ALS, both in neurons and in muscle where antioxidant enzyme activity is altered[27]. To date more than 150 mutations have been described for the SOD1 gene, most of which cause dominantly inherited disease[4]. Mutations in SOD1 that impair its functions can lead to increased oxidative damage, which promotes activation of apoptotic pathways. Mutant SOD1 causes mitochondrial alteration as membrane depolarization, decreased activity of respiratory complexes and cytochrome c release[28-30]. Of note, the role of oxidative stress in tissue homeostasis is complex. It is clear that transiently increased levels of oxidative stress might reflect a potentially health-promoting process, whereas uncontrolled accumulation of oxidative stress might have pathological implications[31]. Additional work is therefore necessary to understand and define precisely whether the manipulation of the redox balance represents a useful approach in the design of therapeutic strategies for neuromuscular diseases.

Protein aggregation and ALS

One of the characteristic features of ALS is the occurrence of extra- or intracellular fibrillar aggregates that contain misfolded proteins with β-sheet conformation[32,33]. Biochemical studies have demonstrated that mutant SOD1 protein tends to be misfolded or forms aggregates that in turn trigger a toxic cascade that leads to neuronal degeneration. More recently, it has been demonstrated that TDP-43 spontaneously forms aggregates that bear a remarkable ultrastructural similarity to TDP-43 deposits in degenerating neurons of ALS patients[34]. The C-terminal domain of TDP-43 is crucial for spontaneous aggregation. Several ALS-linked TDP-43 mutations within this domain (Q331K, M337V, Q343R, N345K, R361S, and N390D) increase the number of TDP-43 aggregates and promote toxicity in vivo[34]. Under most circumstances, cells activate the protein chaperones and the ubiquitin-proteasome system to maintain protein quality control. Normally, cells are capable of handling mutant proteins sufficiently to prevent them from exerting toxic effects and/or being sequestered into inclusions[35]. However, under circumstances of increased physiological or environmental stress, the ubiquitin-proteasome system can become overloaded and impaired. This results in engulfment of the cells, which become defective in the disposal of altered macromolecules and more prone to damage.

Mitochondrial dysfunction and ALS

Several lines of evidence have shown alteration in mitochondria associated with ALS[36]. Mitochondria represent a primary site of intracellular production of reactive oxygen species, and hence a major source of oxidative stress that, in turn, impairs the normal function of mitochondria[37]. Despite numerous reports demonstrating mitochondrial abnormalities associated with ALS, the role of mitochondrial dysfunction in disease onset and progression remains unknown. It has been suggested that mutant SOD1 causes dysfunction and structural damage of mitochondria in human patients and mouse models of ALS[38].

A proposed mechanism suggests that mutant SOD1 is imported into mitochondria[28,29,39], which causes direct damage of the organelle and activation of cell death[30]. The mechanisms that regulate SOD1 mitochondrial import involve the redox state of the cell, the intracellular distribution of the copper chaperone for SOD1, and the folding of SOD1[40]. Indeed, overexpression of copper chaperone for SOD1 increases mitochondrial localization of mutant SOD1, which in turn causes early loss of mitochondrial function and disease progression[41].

Mitochondrial dysfunction, which appears early in the course of ALS pathology, does not seem to be restricted to motor neurons, and it is present in other tissues, particularly skeletal muscle[42].

ALS: IS IT JUST A MOTOR NEURON DISEASE?

Although most efforts have been aimed at defining the potential genes and mechanisms that are associated with motor neuron degeneration and to understand ALS pathogenesis, no consensus has yet emerged about which cells, tissues and pathways are directly affected by mutant SOD1. Several lines of evidence have suggested that the neurodegenerative action of mutant SOD1 genes operate through a dominant paracrine activity that emanates from non-neuronal tissues.

The obvious loss of motor neurons in the spinal cord initially focused attention on how mutant SOD1 might act within motor neurons to provoke neuronal degeneration and death. However, the mutant gene products are expressed widely, which raises the possibility that the toxicity might result from the action of mutant SOD1 protein in non-neuronal cells. This notion is supported by recent experimental evidence.

Notably, restriction of SOD1 mutant expression selectively to postnatal motor neurons fails to produce detectable signs of pathology or motor neuron disease[43]. More recently, Jaarsma et al[44] have demonstrated that transgenic mice in which mutant SOD1 was largely restricted to neurons, under the transcriptional control of Thy1.2 promoter, developed disease only at an old age. However, the disease progressed slowly without reaching the same degree of paralysis compared to the classical animal model of ALS in which the same mutant SOD1 gene is ubiquitously expressed[9]. This suggests that other cell types are involved in ALS-associated neurodegeneration. In fact, analysis of chimeras that are generated between wild-type and SOD1 mutant mouse embryonic cells has revealed that wild-type non-neuronal cells in adult chimeric animals extend the survival of SOD1 mutant motor neurons. This suggests that the neurodegenerative action of mutant SOD1 operates through a dominant paracrine activity that emanates from non-neuronal cells[45].

In particular, it has been demonstrated that, diminishing the SOD1 mutant levels in microglia has little effect on the early disease phase, but markedly slows later disease progression[46]. These results suggest that mutant SOD1 in motor neurons affect disease onset, whereas mutant SOD1 in microglia contributes to the propagation of disease at a late stage.

Notably, astrocyte activity normally increases at later stages of ALS disease and is concomitant with motor neuron degeneration, which suggests that astrocytes act as deadly neighbors that exacerbate motor neuron damage. Indeed, co-cultured motor neurons are less likely to survive when they are on astrocytes that express mutant SOD1, or exposed to astrocyte-conditioned medium, than on astrocytes that express normal SOD1[47].

Although mutant SOD1 is also expressed by muscle, it is not clear whether its presence in skeletal muscle directly contributes to any pathological sign of ALS. This issue has been recently investigated by work in our laboratory, which has demonstrated that muscle selective expression of SOD1 mutation causes pathological alterations and induces pre-symptomatic sign of ALS[48]. Our data might explain previous findings that have shown how the ubiquitous expression of SOD1G93A in transgenic mice causes first muscle atrophy, which is later followed by alteration of the neuromuscular junction (NMJ), retrograde axonal degeneration, and lastly, motor neuron death[49]. This retrograde and progressive (muscle-to-NMJ-to-nerve) sequential pattern of degeneration suggests the possibility that certain muscle abnormalities indeed precede motor neuron death rather than result from it.

These studies formally prove that: (1) muscle atrophy is not necessarily determined by motor neuron degeneration or activity, but it is causally linked to the toxic gain of function of SOD1G93A expression; and (2) skeletal muscle is a direct contributor of ALS pathogenesis. Our study, however, seems apparently in contrast with that reported by Miller et al[50] who found that mutant SOD1 does not cause toxicity by its action within the muscle, which suggests that muscle is not a primary target for non-cell-autonomous toxicity in familial ALS. In these experiments, the authors found that the partial reduction of mutant SOD1 within muscle, using either a lentivirus that encodes a siRNA directed against mutant SOD1 or muscle selective SOD1 mutant gene excision, does not affect disease[50]. The apparent discrepancy between the two studies can be explained considering that: (1) partial suppression of mutant SOD1 accumulation within muscle is not sufficient to delay the progression of the disease in SOD1G93A mice[50], which suggests that residual expression of SOD1G93A mutant gene is able to maintain a pathological muscle phenotype; and (2) low expression levels of MLC/SOD1G93A are sufficient to cause muscle atrophy and alteration in functional performance of the soleus muscle[48].

Moreover other studies support the evidence that skeletal muscle is a primary target of mutant SOD1 toxicity in mice. Wong and Martin have reported that skeletal-muscle-restricted expression of human mutant SOD1 gene causes motor neuron degeneration in old transgenic mice[51]. Dupuis et al[52] have reported that muscle selective alterations in mitochondrial function might initiate NMJ destruction, which is followed by distal axonopathy, astrocytosis in the spinal cord, and mild motor neuron loss[52]. Moreover, Zhou et al[42] have reported that alterations in the potential of mitochondrial inner membrane of fiber segments near NMJs occur in young SOD1G93A mice prior to disease onset.

All the above suggests that skeletal muscle is an important candidate to consider as a primary target of the toxicity that results from mutations in the SOD1 gene, and that the effective connection between muscle and nerve is crucial to the capacity of both partners to survive and function adequately throughout life. Of note, skeletal muscle is also a source of signals that influence neuron survival, axonal growth and maintenance of synaptic connections[53]. Indeed, development in the absence of skeletal muscle results in the sequential ablation of motor neurons from the spinal cord to the brain[54]. Thus, muscle clearly plays an important role in providing guidance and cues to the developing motor neurons, and in providing trophic support to maintain motor neuron and axon function[54]. In absence of this trophic support, muscle can have a negative impact on the nervous system, and therefore, can contribute to the alteration in the functional connection between muscle and nerve.

Among growth factors, insulin-like growth factor 1 (IGF-1) has been implicated in anabolism of muscle and nerve tissues, which induces muscle hypertrophy and promotes neuronal survival[55,56]. The therapeutic potential of IGF-1 in ALS was underscored by injection of SOD1G93A mouse muscle with an adeno-associated virus that carried an IGF-1 gene, and by a transgenic approach in which the local form of IGF-1 was selectively expressed in skeletal muscle[57,58]. The two approaches have demonstrated that muscle IGF-1 expression counteracts the symptoms of ALS and reduces components of catabolism, which activates satellite cells and markers of muscle regeneration[59,60]. However, the use of IGF-1 in ALS patients has previously produced conflicting results and a new study has reported no benefit in either survival or functional scales[61]. The reason for these discrepant results could reside in the different delivery approaches used in human and mouse models and/or in the different isoforms of IGF-1 used in the these studies.

MICRORNA AND ALS

Recent studies have also uncovered profound and unexpected roles for a family of small regulatory RNAs, known as microRNAs (miRNAs) (or miR), in the control of cell proliferation, differentiation and development, and in the pathogenesis of several diseases.

miRNAs are endogenous, approximately 22 nucleotides long, and inhibit translation or promote mRNA degradation by annealing to complementary sequences in the 3′ untranslated regions of specific target mRNAs[62]. miRNA expression profiles are highly dynamic during embryonic development and in adulthood. Mis-expression of miRNAs can perturb embryogenesis, organogenesis, tissue homeostasis and the cell cycle[63].

Several miRNAs are expressed in the nervous system where they play a pivotal role in neuronal differentiation, synaptogenesis and plasticity[64], and deregulation of miRNAs can have profound effects on neuronal physiology and pathology[65]. Although it is known that deregulation of specific miRNA-dependent regulatory circuitries correlates with the initiation and progression of several neurological disorders, the underlying mechanisms of these phenomena are still not fully understood. Both TDP-43 and FUS proteins, mutations of which can be involved in ALS pathogenesis, have been shown to interact specifically with the Drosha protein[66], thus suggesting that they are involved in the regulation of miRNA expression by modulating the activity of this processing enzyme.

In addition, several miRNAs, such as miR 1, miR 133, miR 214, miR 181 and miR 206, are specifically expressed in skeletal muscles[67]. It has been reported recently that miR-206 delays ALS progression and promotes regeneration of neuromuscular synapses in mouse models[68]. In particular, it has been demonstrated that loss of miR-206 does not affect disease onset, but does accelerate disease progression and atrophy of skeletal muscle, which leads to kyphosis, paralysis and death[68].

The mechanism by which miR-206 promotes a partially successful compensatory response to denervation involves the histone deacetylase 4 (HDAC4), which normally inhibits nerve re-innervation by blocking the expression of fibroblast growth factor binding protein 1 (FGFBP1)[68]. Thus, miR-206 blocking the activity of HDAC4 guarantees the activation of FGFBP1 expression, which promotes the maintenance of NMJ integrity and plasticity. This evidence further reinforces the hypothesis that muscle-derived factors can promote a functional nerve-muscle interaction, which delays motor neuron degeneration.

CONCLUSION

Although there have been significant advances in understanding the biology of ALS, no consensus has emerged as to which cells, tissues and pathways are directly affected by mutant SOD1. Several pathways have been implicated in disease pathogenesis, including glutamate-mediated excitotoxicity, mitochondrial dysfunction, neuro-inflammation, apoptosis, oxidative stress, protein aggregation, and aberrant axonal transport (Figure 1). The results reviewed in this manuscript support the redefinition of ALS as a multi-systemic disease in which alterations in structural, physiological and metabolic parameters in different cell types (muscle, motor neurons, and glia) act synergistically to exacerbate the disease. Multi-interventional approaches, including novel methods to intercept the damage and to deliver molecules to vulnerable cells, have recently been shown to be effective[69]. Thus, new avenues for promising therapeutic approaches can be derived from multidrug treatments and/or the delivery of growth factors by viral vectors, in combination with exercise and/or dietary regimes[69]. From a clinical point of view, the most powerful future approach would be to target motor neurons, non-motor neuronal cells and skeletal muscle.

Footnotes

Peer reviewers: Yiider Tseng, PhD, Associate Professor, Department of Chemical Engineering, University of Florida, Room 223, Museum Road, Chemical Engineering Building, Gainesville, FL 32611-6005, United States; Wolfgang Obermann, PhD, Department of Physiology, Ruhr-University Bochum, Universitätsstrasse 150, 44801 Bochum, Germany

S- Editor Cheng JX L- Editor Kerr C E- Editor Zheng XM

References
1.  Charcot JM, Joffroy A. Two cases of progressive muscle atrophy with damage of grey matter and of anterior-lateral bundles of the spinal cord. Arch Physiol Norm Path. 1869;2 354-367, 745-760.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Wijesekera LC, Leigh PN. Amyotrophic lateral sclerosis. Orphanet J Rare Dis. 2009;4:3.  [PubMed]  [DOI]  [Cited in This Article: ]
3.  Pasinelli P, Brown RH. Molecular biology of amyotrophic lateral sclerosis: insights from genetics. Nat Rev Neurosci. 2006;7:710-723.  [PubMed]  [DOI]  [Cited in This Article: ]
4.  Rosen DR, Siddique T, Patterson D, Figlewicz DA, Sapp P, Hentati A, Donaldson D, Goto J, O'Regan JP, Deng HX. Mutations in Cu/Zn superoxide dismutase gene are associated with familial amyotrophic lateral sclerosis. Nature. 1993;362:59-62.  [PubMed]  [DOI]  [Cited in This Article: ]
5.  Shaw BF, Valentine JS. How do ALS-associated mutations in superoxide dismutase 1 promote aggregation of the protein? Trends Biochem Sci. 2007;32:78-85.  [PubMed]  [DOI]  [Cited in This Article: ]
6.  Bruijn LI, Miller TM, Cleveland DW. Unraveling the mechanisms involved in motor neuron degeneration in ALS. Annu Rev Neurosci. 2004;27:723-749.  [PubMed]  [DOI]  [Cited in This Article: ]
7.  Boillée S, Vande Velde C, Cleveland DW. ALS: a disease of motor neurons and their nonneuronal neighbors. Neuron. 2006;52:39-59.  [PubMed]  [DOI]  [Cited in This Article: ]
8.  Gonzalez de Aguilar JL, Echaniz-Laguna A, Fergani A, René F, Meininger V, Loeffler JP, Dupuis L. Amyotrophic lateral sclerosis: all roads lead to Rome. J Neurochem. 2007;101:1153-1160.  [PubMed]  [DOI]  [Cited in This Article: ]
9.  Gurney ME, Pu H, Chiu AY, Dal Canto MC, Polchow CY, Alexander DD, Caliendo J, Hentati A, Kwon YW, Deng HX. Motor neuron degeneration in mice that express a human Cu,Zn superoxide dismutase mutation. Science. 1994;264:1772-1775.  [PubMed]  [DOI]  [Cited in This Article: ]
10.  Ripps ME, Huntley GW, Hof PR, Morrison JH, Gordon JW. Transgenic mice expressing an altered murine superoxide dismutase gene provide an animal model of amyotrophic lateral sclerosis. Proc Natl Acad Sci USA. 1995;92:689-693.  [PubMed]  [DOI]  [Cited in This Article: ]
11.  Lefebvre S, Bürglen L, Reboullet S, Clermont O, Burlet P, Viollet L, Benichou B, Cruaud C, Millasseau P, Zeviani M. Identification and characterization of a spinal muscular atrophy-determining gene. Cell. 1995;80:155-165.  [PubMed]  [DOI]  [Cited in This Article: ]
12.  Nishimura AL, Mitne-Neto M, Silva HC, Richieri-Costa A, Middleton S, Cascio D, Kok F, Oliveira JR, Gillingwater T, Webb J. A mutation in the vesicle-trafficking protein VAPB causes late-onset spinal muscular atrophy and amyotrophic lateral sclerosis. Am J Hum Genet. 2004;75:822-831.  [PubMed]  [DOI]  [Cited in This Article: ]
13.  Marques VD, Barreira AA, Davis MB, Abou-Sleiman PM, Silva WA Jr, Zago MA, Sobreira C, Fazan V, Marques W Jr. Expanding the phenotypes of the Pro56Ser VAPB mutation: proximal SMA with dysautonomia. Muscle Nerve. 2006;34:731-739.  [PubMed]  [DOI]  [Cited in This Article: ]
14.  Gkogkas C, Middleton S, Kremer AM, Wardrope C, Hannah M, Gillingwater TH, Skehel P. VAPB interacts with and modulates the activity of ATF6. Hum Mol Genet. 2008;17:1517-1526.  [PubMed]  [DOI]  [Cited in This Article: ]
15.  Fasana E, Fossati M, Ruggiano A, Brambillasca S, Hoogenraad CC, Navone F, Francolini M, Borgese N. A VAPB mutant linked to amyotrophic lateral sclerosis generates a novel form of organized smooth endoplasmic reticulum. FASEB J. 2010;24:1419-1430.  [PubMed]  [DOI]  [Cited in This Article: ]
16.  Langou K, Moumen A, Pellegrino C, Aebischer J, Medina I, Aebischer P, Raoul C. AAV-mediated expression of wildtype and ALS-linked mutant VAPB selectively triggers death of motoneurons through a Ca-dependent ER-associated pathway. J Neurochem. 2010;Epub ahead of print.  [PubMed]  [DOI]  [Cited in This Article: ]
17.  Tudor EL, Galtrey CM, Perkinton MS, Lau KF, De Vos KJ, Mitchell JC, Ackerley S, Hortobágyi T, Vámos E, Leigh PN. Amyotrophic lateral sclerosis mutant vesicle-associated membrane protein-associated protein-B transgenic mice develop TAR-DNA-binding protein-43 pathology. Neuroscience. 2010;167:774-785.  [PubMed]  [DOI]  [Cited in This Article: ]
18.  Kwiatkowski TJ Jr, Bosco DA, Leclerc AL, Tamrazian E, Vanderburg CR, Russ C, Davis A, Gilchrist J, Kasarskis EJ, Munsat T. Mutations in the FUS/TLS gene on chromosome 16 cause familial amyotrophic lateral sclerosis. Science. 2009;323:1205-1208.  [PubMed]  [DOI]  [Cited in This Article: ]
19.  Vance C, Rogelj B, Hortobágyi T, De Vos KJ, Nishimura AL, Sreedharan J, Hu X, Smith B, Ruddy D, Wright P. Mutations in FUS, an RNA processing protein, cause familial amyotrophic lateral sclerosis type 6. Science. 2009;323:1208-1211.  [PubMed]  [DOI]  [Cited in This Article: ]
20.  Wegorzewska I, Bell S, Cairns NJ, Miller TM, Baloh RH. TDP-43 mutant transgenic mice develop features of ALS and frontotemporal lobar degeneration. Proc Natl Acad Sci USA. 2009;106:18809-18814.  [PubMed]  [DOI]  [Cited in This Article: ]
21.  Lagier-Tourenne C, Cleveland DW. Rethinking ALS: the FUS about TDP-43. Cell. 2009;136:1001-1004.  [PubMed]  [DOI]  [Cited in This Article: ]
22.  Salajegheh M, Pinkus JL, Taylor JP, Amato AA, Nazareno R, Baloh RH, Greenberg SA. Sarcoplasmic redistribution of nuclear TDP-43 in inclusion body myositis. Muscle Nerve. 2009;40:19-31.  [PubMed]  [DOI]  [Cited in This Article: ]
23.  Foran E, Trotti D. Glutamate transporters and the excitotoxic path to motor neuron degeneration in amyotrophic lateral sclerosis. Antioxid Redox Signal. 2009;11:1587-1602.  [PubMed]  [DOI]  [Cited in This Article: ]
24.  Heath PR, Shaw PJ. Update on the glutamatergic neurotransmitter system and the role of excitotoxicity in amyotrophic lateral sclerosis. Muscle Nerve. 2002;26:438-458.  [PubMed]  [DOI]  [Cited in This Article: ]
25.  Van Damme P, Dewil M, Robberecht W, Van Den Bosch L. Excitotoxicity and amyotrophic lateral sclerosis. Neurodegener Dis. 2005;2:147-159.  [PubMed]  [DOI]  [Cited in This Article: ]
26.  Trotti D, Rolfs A, Danbolt NC, Brown RH Jr, Hediger MA. SOD1 mutants linked to amyotrophic lateral sclerosis selectively inactivate a glial glutamate transporter. Nat Neurosci. 1999;2:848.  [PubMed]  [DOI]  [Cited in This Article: ]
27.  Barber SC, Mead RJ, Shaw PJ. Oxidative stress in ALS: a mechanism of neurodegeneration and a therapeutic target. Biochim Biophys Acta. 2006;1762:1051-1067.  [PubMed]  [DOI]  [Cited in This Article: ]
28.  Mattiazzi M, D'Aurelio M, Gajewski CD, Martushova K, Kiaei M, Beal MF, Manfredi G. Mutated human SOD1 causes dysfunction of oxidative phosphorylation in mitochondria of transgenic mice. J Biol Chem. 2002;277:29626-29633.  [PubMed]  [DOI]  [Cited in This Article: ]
29.  Higgins CM, 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.  [PubMed]  [DOI]  [Cited in This Article: ]
30.  Takeuchi H, Kobayashi Y, Ishigaki S, Doyu M, Sobue G. Mitochondrial localization of mutant superoxide dismutase 1 triggers caspase-dependent cell death in a cellular model of familial amyotrophic lateral sclerosis. J Biol Chem. 2002;277:50966-50972.  [PubMed]  [DOI]  [Cited in This Article: ]
31.  Musarò A, Fulle S, Fanò G. Oxidative stress and muscle homeostasis. Curr Opin Clin Nutr Metab Care. 2010;13:236-242.  [PubMed]  [DOI]  [Cited in This Article: ]
32.  Kabashi E, Valdmanis PN, Dion P, Rouleau GA. Oxidized/misfolded superoxide dismutase-1: the cause of all amyotrophic lateral sclerosis? Ann Neurol. 2007;62:553-559.  [PubMed]  [DOI]  [Cited in This Article: ]
33.  Vande Velde C, Miller TM, Cashman NR, Cleveland DW. Selective association of misfolded ALS-linked mutant SOD1 with the cytoplasmic face of mitochondria. Proc Natl Acad Sci USA. 2008;105:4022-4027.  [PubMed]  [DOI]  [Cited in This Article: ]
34.  Johnson BS, Snead D, Lee JJ, McCaffery JM, Shorter J, Gitler AD. TDP-43 is intrinsically aggregation-prone, and amyotrophic lateral sclerosis-linked mutations accelerate aggregation and increase toxicity. J Biol Chem. 2009;284:20329-20339.  [PubMed]  [DOI]  [Cited in This Article: ]
35.  Kabashi E, Durham HD. Failure of protein quality control in amyotrophic lateral sclerosis. Biochim Biophys Acta. 2006;1762:1038-1050.  [PubMed]  [DOI]  [Cited in This Article: ]
36.  Dupuis L, Gonzalez de Aguilar JL, Oudart H, de Tapia M, Barbeito L, Loeffler JP. Mitochondria in amyotrophic lateral sclerosis: a trigger and a target. Neurodegener Dis. 2004;1:245-254.  [PubMed]  [DOI]  [Cited in This Article: ]
37.  Lenaz G, Bovina C, D'Aurelio M, Fato R, Formiggini G, Genova ML, Giuliano G, Merlo Pich M, Paolucci U, Parenti Castelli G. Role of mitochondria in oxidative stress and aging. Ann N Y Acad Sci. 2002;959:199-213.  [PubMed]  [DOI]  [Cited in This Article: ]
38.  Higgins CM, Jung C, Xu Z. ALS-associated mutant SOD1G93A causes mitochondrial vacuolation by expansion of the intermembrane space and by involvement of SOD1 aggregation and peroxisomes. BMC Neurosci. 2003;4:16.  [PubMed]  [DOI]  [Cited in This Article: ]
39.  Okado-Matsumoto A, Fridovich I. Subcellular distribution of superoxide dismutases (SOD) in rat liver: Cu,Zn-SOD in mitochondria. J Biol Chem. 2001;276:38388-38393.  [PubMed]  [DOI]  [Cited in This Article: ]
40.  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.  [PubMed]  [DOI]  [Cited in This Article: ]
41.  Son M, Puttaparthi K, Kawamata H, Rajendran B, Boyer PJ, Manfredi G, Elliott JL. Overexpression of CCS in G93A-SOD1 mice leads to accelerated neurological deficits with severe mitochondrial pathology. Proc Natl Acad Sci USA. 2007;104:6072-6077.  [PubMed]  [DOI]  [Cited in This Article: ]
42.  Zhou J, Yi J, Fu R, Liu E, Siddique T, Ríos E, Deng HX. Hyperactive intracellular calcium signaling associated with localized mitochondrial defects in skeletal muscle of an animal model of amyotrophic lateral sclerosis. J Biol Chem. 2010;285:705-712.  [PubMed]  [DOI]  [Cited in This Article: ]
43.  Lino MM, Schneider C, Caroni P. Accumulation of SOD1 mutants in postnatal motoneurons does not cause motoneuron pathology or motoneuron disease. J Neurosci. 2002;22:4825-4832.  [PubMed]  [DOI]  [Cited in This Article: ]
44.  Jaarsma D, Teuling E, Haasdijk ED, De Zeeuw CI, Hoogenraad CC. Neuron-specific expression of mutant superoxide dismutase is sufficient to induce amyotrophic lateral sclerosis in transgenic mice. J Neurosci. 2008;28:2075-2088.  [PubMed]  [DOI]  [Cited in This Article: ]
45.  Clement AM, Nguyen MD, Roberts EA, Garcia ML, Boillée S, Rule M, McMahon AP, Doucette W, Siwek D, Ferrante RJ. Wild-type nonneuronal cells extend survival of SOD1 mutant motor neurons in ALS mice. Science. 2003;302:113-117.  [PubMed]  [DOI]  [Cited in This Article: ]
46.  Boillée S, Yamanaka K, Lobsiger CS, Copeland NG, Jenkins NA, Kassiotis G, Kollias G, Cleveland DW. Onset and progression in inherited ALS determined by motor neurons and microglia. Science. 2006;312:1389-1392.  [PubMed]  [DOI]  [Cited in This Article: ]
47.  Julien JP. ALS: astrocytes move in as deadly neighbors. Nat Neurosci. 2007;10:535-537.  [PubMed]  [DOI]  [Cited in This Article: ]
48.  Dobrowolny G, Aucello M, Rizzuto E, Beccafico S, Mammucari C, Boncompagni S, Belia S, Wannenes F, Nicoletti C, Del Prete Z. Skeletal muscle is a primary target of SOD1G93A-mediated toxicity. Cell Metab. 2008;8:425-436.  [PubMed]  [DOI]  [Cited in This Article: ]
49.  Dupuis L, Loeffler JP. Neuromuscular junction destruction during amyotrophic lateral sclerosis: insights from transgenic models. Curr Opin Pharmacol. 2009;9:341-346.  [PubMed]  [DOI]  [Cited in This Article: ]
50.  Miller TM, Kim SH, Yamanaka K, Hester M, Umapathi P, Arnson H, Rizo L, Mendell JR, Gage FH, Cleveland DW. Gene transfer demonstrates that muscle is not a primary target for non-cell-autonomous toxicity in familial amyotrophic lateral sclerosis. Proc Natl Acad Sci USA. 2006;103:19546-19551.  [PubMed]  [DOI]  [Cited in This Article: ]
51.  Wong M, Martin LJ. Skeletal muscle-restricted expression of human SOD1 causes motor neuron degeneration in transgenic mice. Hum Mol Genet. 2010;19:2284-2302.  [PubMed]  [DOI]  [Cited in This Article: ]
52.  Dupuis L, Gonzalez de Aguilar JL, Echaniz-Laguna A, Eschbach J, Rene F, Oudart H, Halter B, Huze C, Schaeffer L, Bouillaud F. Muscle mitochondrial uncoupling dismantles neuromuscular junction and triggers distal degeneration of motor neurons. PLoS One. 2009;4:e5390.  [PubMed]  [DOI]  [Cited in This Article: ]
53.  Funakoshi H, Belluardo N, Arenas E, Yamamoto Y, Casabona A, Persson H, Ibáñez CF. Muscle-derived neurotrophin-4 as an activity-dependent trophic signal for adult motor neurons. Science. 1995;268:1495-1499.  [PubMed]  [DOI]  [Cited in This Article: ]
54.  Kablar B, Rudnicki MA. Development in the absence of skeletal muscle results in the sequential ablation of motor neurons from the spinal cord to the brain. Dev Biol. 1999;208:93-109.  [PubMed]  [DOI]  [Cited in This Article: ]
55.  Suzuki M, Svendsen CN. Combining growth factor and stem cell therapy for amyotrophic lateral sclerosis. Trends Neurosci. 2008;31:192-198.  [PubMed]  [DOI]  [Cited in This Article: ]
56.  Scicchitano BM, Rizzuto E, Musarò A. Counteracting muscle wasting in aging and neuromuscular diseases: the critical role of IGF-1. Aging (Albany NY). 2009;1:451-457.  [PubMed]  [DOI]  [Cited in This Article: ]
57.  Kaspar BK, Lladó J, Sherkat N, Rothstein JD, Gage FH. Retrograde viral delivery of IGF-1 prolongs survival in a mouse ALS model. Science. 2003;301:839-842.  [PubMed]  [DOI]  [Cited in This Article: ]
58.  Musarò A, McCullagh K, Paul A, Houghton L, Dobrowolny G, Molinaro M, Barton ER, Sweeney HL, Rosenthal N. Localized Igf-1 transgene expression sustains hypertrophy and regeneration in senescent skeletal muscle. Nat Genet. 2001;27:195-200.  [PubMed]  [DOI]  [Cited in This Article: ]
59.  Dobrowolny G, Giacinti C, Pelosi L, Nicoletti C, Winn N, Barberi L, Molinaro M, Rosenthal N, Musarò A. Muscle expression of a local Igf-1 isoform protects motor neurons in an ALS mouse model. J Cell Biol. 2005;168:193-199.  [PubMed]  [DOI]  [Cited in This Article: ]
60.  Dobrowolny G, Aucello M, Molinaro M, Musarò A. Local expression of mIgf-1 modulates ubiquitin, caspase and CDK5 expression in skeletal muscle of an ALS mouse model. Neurol Res. 2008;30:131-6.  [PubMed]  [DOI]  [Cited in This Article: ]
61.  Sorenson EJ, Windbank AJ, Mandrekar JN, Bamlet WR, Appel SH, Armon C, Barkhaus PE, Bosch P, Boylan K, David WS. Subcutaneous IGF-1 is not beneficial in 2-year ALS trial. Neurology. 2008;71:1770-1775.  [PubMed]  [DOI]  [Cited in This Article: ]
62.  Bartel DP. MicroRNAs: genomics, biogenesis, mechanism, and function. Cell. 2004;116:281-297.  [PubMed]  [DOI]  [Cited in This Article: ]
63.  Asli NS, Pitulescu ME, Kessel M. MicroRNAs in organogenesis and disease. Curr Mol Med. 2008;8:698-710.  [PubMed]  [DOI]  [Cited in This Article: ]
64.  Vo NK, Cambronne XA, Goodman RH. MicroRNA pathways in neural development and plasticity. Curr Opin Neurobiol. 2010;Epub ahead of print.  [PubMed]  [DOI]  [Cited in This Article: ]
65.  Shafi G, Aliya N, Munshi A. MicroRNA signatures in neurological disorders. Can J Neurol Sci. 2010;37:177-185.  [PubMed]  [DOI]  [Cited in This Article: ]
66.  Gregory RI, Yan KP, Amuthan G, Chendrimada T, Doratotaj B, Cooch N, Shiekhattar R. The Microprocessor complex mediates the genesis of microRNAs. Nature. 2004;432:235-240.  [PubMed]  [DOI]  [Cited in This Article: ]
67.  van Rooij E, Liu N, Olson EN. MicroRNAs flex their muscles. Trends Genet. 2008;24:159-166.  [PubMed]  [DOI]  [Cited in This Article: ]
68.  Williams AH, Valdez G, Moresi V, Qi X, McAnally J, Elliott JL, Bassel-Duby R, Sanes JR, Olson EN. MicroRNA-206 delays ALS progression and promotes regeneration of neuromuscular synapses in mice. Science. 2009;326:1549-1554.  [PubMed]  [DOI]  [Cited in This Article: ]
69.  Carrì MT, Grignaschi G, Bendotti C. Targets in ALS: designing multidrug therapies. Trends Pharmacol Sci. 2006;27:267-273.  [PubMed]  [DOI]  [Cited in This Article: ]