Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Apr 20.
Published in final edited form as: Neurol Disord Epilepsy J. 2019 Jul 30;3(2):124.

Ion channel dysfunction and altered motoneuron excitability in ALS

Eric LoRusso 1, James J Hickman 1, Xiufang Guo 1,*
PMCID: PMC7170321  NIHMSID: NIHMS1579516  PMID: 32313901

Abstract

Dysregulated excitability is a hallmark of Amyotrophic Lateral Sclerosis (ALS) pathology both in ALS research models and in clinical settings. This primarily results from the dysfunction of Na+, K+, and Ca2+ ion channels responsible for maintaining neuronal thresholds and executing signal transduction or synaptic transmission. The exact dysfunction that each of these ion channel currents display in ALS pathology can vary between different ALS models, mainly induced pluripotent stem cell (iPSC) derived human motoneurons and ALS mouse models. Moreover, results can vary further across ALS mutations and between different developmental periods of these disease models. This review attempts to gather observations regarding ion channel dysfunction contributing to both hyperexcitable and hypoexcitable phenotypes in ALS motoneurons both in vivo and in vitro, so as to assess their potential as therapeutic targets.

Keywords: ALS, Ion Channels, Electrophysiology, Human

Introduction

Amyotrophic Lateral Sclerosis (ALS) is a fatal neurodegenerative disease characterized by the progressive loss of both upper and lower motoneurons leading to voluntary muscle weakness, muscle atrophy, and ultimately respiratory failure [1]. Approximately 5 – 10% of ALS cases are familial, following a Mendelian pattern of inheritance, with 80% of familial patients exhibiting known disease-causing mutations [2]. The vast majority of remaining cases are apparently sporadic in nature, with only 5 – 10% expressing these mutations [3]. Over 20 gene mutations have been linked to ALS pathology, seen in both familial and sporadic cases [4]. The most common gene mutation is found in C9orf72 (coding for chromosome 9 open reading frame 2), making up 10–15% of total ALS cases. Mutations in the gene coding for SOD1 (Super-oxide dismutase 1), a Zn/Cu binding protein responsible for destroying free superoxide radicals, is the most well studied and responsible for 2% of ALS cases. Other commonly studied mutations are found in the genes coding for TDP-43 (TAR DNA-binding protein 43), responsible for 0.9% of ALS cases; and FUS (Fused in Sarcoma), responsible for 0.7% of ALS cases [2, 5], however, many other less common mutations have also been linked to ALS. Exactly how these mutations lead to progressive motoneuron cell death is currently unknown, and the topic is of great interest to researchers.

Cortical hyperexcitability is frequently observed in ALS patients through the utilization of threshold-tracking transcranial magnetic stimulation (TMS) combined with motor electrode potential recording [68]. Many studies also have reported axonal dysfunction in ALS patients that is characterized by increased excitability, abnormalities in threshold potential, and a prolonged strength-duration time constant (TSD), a parameter which measures the rate of the decline of threshold currents as the stimulus duration is increased, all pointing to hyperexcitability [9, 10]. These abnormalities have been specifically attributed to irregular ionic conductances, more specifically the increased Na+ and decreased axonal K+ conductances, a profile that may underlie neurodegeneration and contribute to symptoms, such as fasciculations and muscle cramps found in ALS patients [1113]. In vitro approaches to studying this phenomenon utilizes SOD1 mouse models, as well as iPSC-derived motoneurons. Hyperexcitability is often seen in these ALS models, typically attributed to altered voltage-gated Na+ and K+ channel activity leading to an increased propensity for neuronal firing, and possibly cell death due to increased oxidative stress, intracellular Ca2+ build up, followed by mitochondrial dysfunction [1316]. Dysregulated Ca2+ channels may contribute to altered excitability in a similar fashion. More recently, hypoexcitable phenotypes have also been implicated in ALS pathology [17]. Regardless, altered excitability evolves with disease progression and correlates with survival [18, 19], and can hereby serve as a potential therapeutic biomarker in ALS. However, since modulation of excitability in ALS has resulted in significant symptomatic improvement, ion channels and ion homeostasis represents a significant group of therapeutic targets for ALS drug development [20]. Due to the multifactorial nature of ALS pathology, it is likely that this process involves many pathological inputs from surrounding astrocytes, microglia and interneurons, however this review is focused on the intrinsic properties of ALS spinal motoneurons that contribute to the altered excitability.

Ca2+ channels

Molecular basis of AMPA receptor dysfunction in ALS

Dysregulated calcium homeostasis in motoneurons affected by ALS is a key player in ALS pathology and has been proposed to be a convergent point of major ALS dysfunctional pathways [21]. Increased intracellular calcium levels in affected motoneuron pools can upregulate many signal transduction pathways and promote pathological mechanisms including excitotoxicity, oxidative stress, ER stress, mitochondrial dysfunction, protein misfolding and aggregation and ultimately cell death [21, 22]. The exaggerated Ca2+ influx from the extracellular environment by Ca2+ -permeable channels is a likely cause of increased intracellular Ca2+ concentrations [23, 24]. AMPA receptors (AMPAR) are ionotropic glutamate receptors composed of multiple subunits primarily responsible for the fast excitatory transmission in the central nervous system (CNS). It is well known that glutamate mediated excitotoxicity is present in ALS pathology, and that it is at least in part mediated by increased intracellular calcium concentrations, some are caused by an exaggerated influx through Ca2+-permeable AMPA-receptors [25, 26]. The GluA2 subunit is the dominant functional determinant for AMPA receptor Ca2+ permeability, and its expression is normally increased after early post-natal development, leading to a switch from Ca2+ -permeable to Ca2+ -impermeable AMPA receptors [27]. In some ALS models, this switch appears to be reversed or inhibited at some point in disease onset or progression. Studies have shown that Ca2+-permeable AMPA receptors lacking the regulatory GluA2 subunit are present and cause the exaggerated influx of Ca2+ into the cytoplasm in both SOD1 ALS and other motoneuron disease models [2830]. However, other subunits of the AMPA receptors are implicated as well. Some have been found in SOD1 mouse models where an increased expression of the GluA3 subunit was coupled with a decrease in the GluA2 subunit [31, 32], and that an antisense peptide targeting GluA3 delayed disease onset and progression [32]. This event occurred even though GluA3 isn’t the regulatory subunit for AMPAR Ca2+ permeability as is GluA2. Similar findings indicated an increase in Ca2+-permeable AMPA receptors in human C9orf72 iPSC-derived motoneurons, coupled with an upregulation of GluA1 subunits not commonly present in previous mouse model studies [24]. In the same study an increased GluA1 expression was also seen in C9orf72 post-mortem spinal cord tissue and not in cortical neurons, indicating a motoneuron specific mechanism, as well as validating pathological similarities between the C9orf72 iPSC derived motoneurons and patient primary motoneurons [24]. However, decreased expression of PICK1, an adaptor protein that regulates trafficking of the GluA2 subunit, was also observed, implying that dysregulated mechanisms involving GluA2 may still underlie the mutant AMPAR phenotype [24]. These results indicate differential mechanisms in which AMAPR components may increase Ca2+-permeable AMPA receptors beyond just the GluA2 regulatory subunit [24]. However, although multiple AMPAR subunits have been implicated in the receptor’s disease phenotype, there is little consensus on their role other than a primary role of the regulatory GluA2 subunit.

Aside from alterations in subunit composition, insufficient transcriptional editing of the Q/R site in the GluA2 domain of the AMPAR is also thought to contribute to its Ca2+ -permeable phenotype [33]. Q/R editing is a result of edition of GluR2 pre-mRNA in which a gene-encoded glutamine (Q) codon in the channel-forming intramembrane segment is changed to an arginine (R) codon (or A-I conversion in pre-mRNA). This edition is necessary to switch GluA2 from a Ca2+-permeable state to impermeable state during early development stages [34, 35]. Q/R editing in GluA2 is necessary for normal neuron function, where its impairment produces Ca2+ -permeable AMPARs resulting in a higher rate of excitotoxicity [36]. This process occurs by the action of the enzyme Adenosine Deaminase Acting on RNA2 (ADAR2), which converts an adenosine at the Q/R site to an inosine (A-I conversion) [36]. The direct correlation between ADAR2 and motor diseases have been shown in knockout mice models with ADAR2 conditionally targeted in motoneurons, resulting in an increase in Ca2+ -permeable AMPA activity due to less Q/R editing, followed by a decline in motor function and slow motoneuron death [36]. The same study also found that neurons in the nuclei of oculomotor nerves were resistant to the mutation despite showing decreased editing, which is consistent with ALS disease progression [36]. Although this suggests a correlation between ADAR2 activity and motoneuron disease, reduced RNA editing of the Q/R site is not found in the majority of familial ALS models or patients [37, 38]. However, it is a strong pathological marker of sporadic ALS cases making it a target for sporadic ALS therapy, and giving insight into possible disease mechanisms [24, 37]. In fact, decreased R/Q editing has been demonstrated in the majority of sporadic ALS cases regardless of clinical manifestation and is absent in other motoneuron disease cases, implying its disease-specific role in ALS [37, 39]. In addition, reduced ADAR2 activity in sporadic patients is almost exclusively coupled with decreased nuclear TDP-43 and increased abnormal cytoplasmic TDP-43 fragments (a common observation in both familial and sporadic ALS), suggesting a molecular link between these two ALS-specific molecular defects [37, 40].

Molecular basis of voltage-gated calcium channels in ALS

In addition to glutamate receptors, Ca2+ flux into motoneurons can occur through voltage gated calcium channels (VGCCs). VGCC activation results from a depolarizing current, and is responsible for depolarizing the neuron, regulating excitability, release of neurotransmitters, and influx of calcium as a secondary messenger modulating gene expression and cell behavior. Because AMPAR receptor targeting has not translated well into therapies for ALS, another source of intracellular Ca2+ overload may result from defects in VGCCs [41], as VGCC activity has been shown to modulate motoneuron excitability [23, 41, 42]. N-, P/Q-, L-, and R-type Ca2+ channels are seen in wildtype motoneurons, although the proportion of the inward current responsible for each type is varied [23, 43]. Research regarding VGCC defects in ALS pathology is not common, however a study by Chang and Martin [23] was done comparing VGCC activity between cultured spinal motoneurons from healthy and G93A-SOD1 mice. The study found that N-type high voltage-activated (HVA) Ca2+ currents were increased in the SOD1 mouse model, possibly due to transcriptional upregulation of VGCC subunits verified by RT-PCR, and increased localization of Ca1B (the subtype of α1 subunit in Calcium channels for N-type VGCCs) on the plasma membrane. The mechanism responsible for this was speculated to be a change in the redox state of the receptor due to mutant SOD1 induced oxidative stress. Interestingly, this observation was limited to large alpha-motoneurons. The high metabolic load of large alpha-motoneurons is consistent with the previously proposed theory that oxidative stress attributes to VGCC over-activity. Persistent Ca2+ currents (PCCa) are important in generating plateau potentials, amplifying synaptic inputs and modulating motoneuron excitability [23]. Additionally, abnormally increased PCCa amplitude, regulated by the L-type Ca2+ channel, is also observed in SOD1 mouse motoneurons [23]. High PCCa may increase motoneuron excitability in response to an excitatory post-synaptic potential, contributing to the hyperexcitability phenotype. However, VGGCs may not play a pathological role in all ALS subtypes as seen by a study in which the Ca2+ channel blocker lomerizine was found to be a successful neuroprotective agent for SOD1 but not for TDP-43 ALS mutants [42].

To date there has been little success specifically targeting either AMPAR or VGCC in clinical settings. However, the drug Riluzole, the most successful ALS therapeutic drug, acts in part as an AMPAR antagonist by possibly blocking the exaggerated influx of Ca2+ from the above pathological mechanisms.

Na+ channels

Altered Na+ currents in SOD1 mouse models

The regulation of Na+ currents is also often dysfunctional in ALS models, leading to altered neuronal excitability in both in vivo and in vitro cases [4447]. The hyperexcitability of these cells reflects both intrinsic motoneuron hyperexcitability and increased/altered excitatory inputs derived from interneurons and/or modulating astrocytes [48, 49]. Hyperexcitable phenotypes are predominantly found in embryonic or early postnatal SOD1 ALS models [45, 47, 50, 51]. For this reason, most excitability studies in mouse models utilized this developmental stage. This may indicate that hyperexcitability affects vulnerable neuron populations before disease onset. Embryonic ALS mouse studies have revealed an early hyperexcitability phenotype, attributed to an increase in persistent inward Na+ current amplitudes (PICNa) [45]. This current, in contrast to the fast inactivating Na+ current, is noninactivating or persistent. It is responsible for regulating membrane potential in the subthreshold range, and dendritic depolarizations, facilitating repetitive firing and enhancing synaptic transmission [52]. Increased PICNa amplitude may allow for increased repetitive firing or synaptic potentials leading to hyperexcitability. An increase in repetitive firing can be seen in multiple ALS mouse model studies as well as in iPSC-derived neurons [53, 54]. In other studies, the recovery from fast-inactivation after action potential propagation rather than increased Na+ current amplitude is accelerated in ALS models, which increases depolarization frequency [55]. Altered Na+ currents coupled with hyperexcitability has also been seen in neonatal SOD1 mouse models [47, 49, 56]. Increased PICNa currents coupled with increased spontaneous transmission are seen in this developmental period, also altering the rate of motoneuron firing to injected current relationship (f-I relationship) [47, 49, 54]. However, one study asserted that motoneurons affected in ALS do not exhibit intrinsic hyperexcitability at this stage [56]. Clinically, prolongation of Tsp has been consistently identified in sporadic ALS and linked to neurodegeneration [9], as well as in familial forms of ALS linked to mutations in SOD1 and C9orf72 [57]. It is proposed that increased TSD (strength duration time constant, a measure for axonal excitability) that presents in clinical ALS pathology is due to an exaggerated PICNa current that facilitates depolarization towards threshold potentials, although reduced K+ current was also identified in some cases [9, 57, 58].

Different Na+ channel isoforms are expressed in motoneurons at different developmental stages. Embryonic motoneurons express predominantly the isoforms Nav1.2 and Nav1.3, followed by a decrease in Nav1.3 expression and persistent Nav1.1, Nav1.2, and Nav1.6 expression in the adult CNS [59, 60]. More recently, it has been shown that one of the most common mutations occurring in ALS patients, SOD1A4V, increased peak Na+ amplitudes from Nav1.3 and induced a hyperpolarizing shift in the voltage dependence of the Na+ channel in embryonic ALS mouse motoneurons [44]. Since Nav1.3 is not commonly present in adult mice, it is hard to speculate on its role in disease onset and progression. In adult humans Nav1.3 is also rare, however axonal transection in a dorsal root ganglia model caused an increase in its expression [61]. This implicates axotomy could be a mechanism that de-differentiates neurons to an “embryonic state” that leads to the re-expression of Nav1.3. It would be intriguing to find out whether this could be the case in traumatic brain injuries, which is speculated to be a risk factor for ALS [62, 63]. In contrast to embryonic models, the majority of electrophysiology studies in adult ALS mouse models both in vivo and in vitro exhibit no hyperexcitability but a slight hypoexcitable phenotype [50, 51]. However, ventral-root recordings of adult SOD1 mice displayed a lessened depression of compound action potentials (APs) evoked by dorsal root stimulation compared to control, indicating an increased tendency to fire [46]. Although intrinsic hyperexcitability is not observed in adult SOD1 mouse models, an increase in beta3 protein (a voltage-gated Na+ channel subunit) is seen in presympotmatic adult mice and an increase in levels of both transcript and total protein is seen in symptomatic adult mice [64]. This indicates that mutant SOD1 may facilitate the transport, folding, and assembly of the beta3 protein in presympotmatic mice, before an increase in transcript is observed. This beta3 protein subunit was found to directly affect persistent Na+ currents, altering the neurons excitability near firing threshold [65, 66]. The dysregulation of beta3 protein and the electrophysiological implications correlate with observations in embryonic and neonatal models, but it does not correlate with observations in adult mice, suggesting the presence of some other mechanism(s). These results suggest that increased Na+ current and intrinsic hyperexcitability may play a more critical role in pre-symptomatic stages rather than during disease onset and progression.

Altered Na+ currents using iPSC-derived motoneurons from ALS patients

The use of a more recent model, iPSC-derived neurons from human ALS patients, it has been reported that there exists the same discrepancy in the hyperexcitability vs hypoexcitability phenotypes, and their role in disease progression. Studies utilizing TDP-43 and C9orf72 mutant iPSC-motoneurons reported hypoexcitability phenotypes [53], whereas those using SOD1 mutant iPSC-motoneurons reported clear hyperexcitability [67]. One recent study by Devlin et al. (2015) found a way to link both phenotypes [68]. The intrinsic properties of TDP-43 and C9orf72 iPSC-motoneurons were studied, as in previous reports, however monitoring over a longer time period resulted in an overall switch from a hyperexcitable phenotype to a hypoexcitable phenotype. This observation was coupled with a decrease in fast-inactivating Na+ currents. Further studies in SOD1 and FUS motoneurons revealed strong hypoexcitability correlated with attenuated sodium channel expression levels and reduced Na+ current peak amplitudes [53]. Due to iPSC-derived motoneurons being a relatively recent advancement, data comparing multiple ALS mutants are limited. Further studies need to be carried out comparing multiple mutants in parallel, to elucidate mutant-specific molecular mechanisms affecting excitability. It is worth noting that in these studies, it is difficult to determine whether late stage hypoexcitability is due to disease pathological progression or long-term culture conditions, even with reports of hypoexcitability while maintaining cell viability [68].

Regardless of the phenotype, alteration in Na+ channel function regulates excitability in ALS pathology and has been employed as a therapeutic target. Riluzole, the only successful drug available for ALS treatment, blocks TTX-sensitive Na+ channels. Although this drug’s therapeutic effect most likely proceeds by a number of different mechanisms, ranging from ion channel activity to other cellular processes, Riluzole treatment in vitro decreased PICNa with a concurrent decrease in excitability to levels of control [54]. This indicates that the therapeutic action of Riluzole, in part, may be due to regulating Na+ channel activity and decreasing intrinsic excitability. Na+ channel blockage by other compounds such as Flecainide (currently in clinical trials) has been seen to stabilize peripheral nerve excitability [69]. However, this mechanism of action does not explain the growing amount of data implicating hypoexcitability in ALS. However, the molecular foundation behind Na+ channel alteration in ALS is not very well understood. Oxidative stress resulting from mutant SOD1 may directly affect Na+ channel function via oxidation [45], and more specifically, oxidative stress can activate kinases such as protein kinase C (PKC) [45]. PKC is known to phosphorylate the alpha unit of certain Na+ channels, and PKC levels are elevated in the SOD1 mouse model [44]. Other indirect mechanisms may include increased or decreased axonal trafficking of ion channels. Investigation of the molecular link between ALS gene mutations and ion channel dysfunction may provide better insight on strategies targeting these channels for ALS treatment.

K+ channels

K+ channel dysfunction in ALS

Voltage-gated K+ channels are responsible for opposing the inward current created by voltage-activated Na+ channels during AP propagation. Due to their direct action on membrane excitability, K+ channels are also implicated in pathological excitability changes in ALS motoneurons. In fact, a decrease in protein expression of delayed rectifier K+ channels is specifically seen in the ventral roots of post-mortem sporadic ALS patient spinal cords [70]. Hypermethylation and downregulation of K+ channels is also found in ALS postmortem spinal cords [71]. This evidence in human ALS cases further solidifies the idea that molecular dysfunction of K+ channels occurs in disease pathology and should somewhat be reflected in ALS models. One study has observed a decrease in protein expression of the voltage-activated ion channels Kv1.6 and Nav1.6 in rat motoneurons exposed to the cerebral spinal fluid from sporadic ALS patients [72]. However, although most embryonic and post-natal ALS mouse studies do exhibit some amount of hyperexcitability, evidence supporting specific K+ current dysfunction in SOD1 mouse model in vitro studies is sparse, with most studies focusing on the previously mentioned Voltage-activated Na+ currents.

K+ channel dysfunction in iPSC-derived motoneurons from ALS patients

However, in iPSC-derived motoneurons, altered K+ channel currents are directly correlated with cell excitability in which both hyper- and hypoexcitability is observed [67, 68]. In one such study, decreased delayed rectifier K+ channel currents were found to be solely responsible for the hyperexcitable phenotype independent of Na+ channel activity [67]. In the same study, Retigabine, an anti-convulsant used to treat seizures, pharmacologically activated the voltage-activated K+ channel Kv7, lowering its firing rate and decreasing hyperexcitability to control levels in C9orf72, SOD1, Fus, and TDP-43 iPSC-derived mutant motoneurons [67]. In contrast to the previous study, a shift from early stage hyperexcitability to late stage hypoexcitability reflected a decrease in persistent voltage-gated K+ currents in C9orf72 and TDP-43 iPSC-derived motoneurons [68]. However, this study did not explore the ion currents responsible for this early hyperexcitability, and the observation was also coupled with a decrease in Na+ currents indicating that a general loss of ion channel function may result in hypoexcitability, rather than a specific loss of either the K+ or Na+ current. Others have seen a specific elevated K+ current responsible for reduced Na/K ratios in hypoexcitable FUS mutants, where both elevated K+ current and attenuated Na+ current were responsible for the same observation in hypoexcitable SOD1 mutants [53]. Differential K+ channel alterations across these different studies likely reflect mutation specific mechanisms, as well as different developmental stages likely responsible for this hyper-/hypoexcitability discrepancy. Although evidence suggesting the role of altered K+ currents in excitability changes in ALS is lacking, compared to Ca2+ and Na+ channels, axonal excitability studies in ALS patients indicated direct abnormalities in Na+ and K+ channels correlating to an observed hyperexcitable phenotype [9].

Summary and Future Directions

Research to date suggests that ion channel dysfunction in motoneurons likely contributes to altered intrinsic excitability in ALS pathology. Most electrophysiology studies in early ALS mouse models implicate dysregulated Na+ currents as the most impactful factor in a hyperexcitable phenotype [44, 45, 47, 4951, 56], whereas in iPSC-derived motoneuron models, both Na+ and K+ channel dysfunction are implicated in altered electrophysiology [53, 67, 68]. Ca2+ channels may also contribute to altered excitability by raising the membrane potential closer to threshold. However, exaggerated influx of extracellular calcium most likely contributes to altered excitability due to mitochondrial dysfunction, oxidative damage, and altered cellular transport rather than directly affecting membrane potential properties [21]. Although the debate on whether hyper- or hypoexcitability contributes to cell death and denervation is far from settled, the SOD1 mouse model seems to primarily have pre-symptomatic hyperexcitability [44, 45, 47, 49, 50] followed by late stage normal or hypoexcitability [50, 51]. This is consistent with clinical manifestations in which patients express increased peripheral nerve excitability months before disease onset [9]. Studies in the iPSC-derived motoneurons are much more variable, with many observing either hyper- [53, 68] or hypoexcitability [67, 68]. This may also be due to a pathological switch between the two phenotypes during disease progression [68]. Clinically, both hyper- and hypoexcitability are observed in ALS patients, while axonal hyperexcitability seems to be more prominent [9, 73]. Overall, significant progress has been made concerning the mechanism of hyperexcitability in ALS and its potential as a therapeutic target. However, debates will continue regarding the roles and mechanisms of hyper- or hypoexcitability in ALS pathology, such as which one is more responsible for denervation, or whether one is a compensatory mechanism or pathological subsequent of the other, or both are simply pathological phenotypes representing different disease stages or different ALS subtypes.

Although there is still question about the role of altered excitability in disease onset and progression, it is an attractive target for ALS therapy development. The potential for pharmacological action targeting ion channels is evidenced by the antagonistic action of Riluzole on AMPA receptors and voltage-gated Na+ channels, as well as other compounds including the Na+ channel blocker Flecainide [69] and the K+ channel agonist Retigabine [67]. Edavarone, a new antioxidant drug approved for ALS, has also displayed therapeutic potential by reducing oxidative damage caused by increased intracellular calcium [74]. Novel therapeutics targeting these ion channels may be used in conjunction with other therapeutic methods to limit excitotoxicity in motoneurons, while employing regenerative techniques such as neural grafts and stem cell therapies to enhance recovery. Also, with the advent of iPSC technology, human-derived motoneuron models expressing different ALS mutations are being widely employed for the study of ALS. Studying pathological differences in channel activity between different mutants, as shown in the above review, may lead to more efficient patient-specific therapies.

Acknowledgements:

This research was funded by NIH grant R01-NS050452 and a Department of the Army grant number W81XWH1410162.

Footnotes

Conflict of Interest: XG and EL confirm they have no competing financial interests and there has been no financial support for this research that could have influenced its outcome. However, JJH has a potential competing financial interest, in that a company has been formed to market services for types of cells like this in body-on-a-chip devices.

References:

  • 1.Zarei S, Carr K, Reiley L, Diaz K, Guerra O, Altamirano PF, et al. , A comprehensive review of amyotrophic lateral sclerosis. Surgical Neurology International, 2015. 6: p. 171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Al-Chalabi A, van den Berg LH, and Veldink J, Gene discovery in amyotrophic lateral sclerosis: implications for clinical management. Nature Reviews Neurology, 2016. 13: p. 96. [DOI] [PubMed] [Google Scholar]
  • 3.Debray S, Race V, Crabbé V, Herdewyn S, Matthijs G, Goris A, et al. , Frequency of C9orf72 repeat expansions in amyotrophic lateral sclerosis: a Belgian cohort study. Neurobiology of Aging, 2013. 34(12): p. 2890.e7–2890.e12. [DOI] [PubMed] [Google Scholar]
  • 4.Al-Chalabi A, Jones A, Troakes C, King A, Al-Sarraj S, and van den Berg LH, The genetics and neuropathology of amyotrophic lateral sclerosis. Acta Neuropathologica, 2012. 124(3): p. 339–352. [DOI] [PubMed] [Google Scholar]
  • 5.Renton AE, Chiò A, and Traynor BJ, State of play in amyotrophic lateral sclerosis genetics. Nature Neuroscience, 2013. 17: p. 17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Menon P, Geevasinga N, van den Bos M, Yiannikas C, Kiernan MC, and Vucic S, Cortical hyperexcitability and disease spread in amyotrophic lateral sclerosis. Eur J Neurol, 2017. 24(6): p. 816–824. [DOI] [PubMed] [Google Scholar]
  • 7.Vucic S, Ziemann U, Eisen A, Hallett M, and Kiernan MC, Transcranial magnetic stimulation and amyotrophic lateral sclerosis: pathophysiological insights. J Neurol Neurosurg Psychiatry, 2013. 84(10): p. 1161–70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Mills KR and Nithi KA, Corticomotor threshold is reduced in early sporadic amyotrophic lateral sclerosis. Muscle Nerve, 1997. 20(9): p. 1137–41. [DOI] [PubMed] [Google Scholar]
  • 9.Park SB, Kiernan MC, and Vucic S, Axonal Excitability in Amyotrophic Lateral Sclerosis. Neurotherapeutics, 2017. 14(1): p. 78–90. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Vucic S and Kiernan MC, Axonal excitability properties in amyotrophic lateral sclerosis. Clinical Neurophysiology, 2006. 117(7): p. 1458–1466. [DOI] [PubMed] [Google Scholar]
  • 11.Layzer RB, The origin of muscle fasciculations and cramps. Muscle & Nerve, 1994. 17(11): p. 1243–1249. [DOI] [PubMed] [Google Scholar]
  • 12.Kuwabara S, Shibuya K, and Misawa S, Fasciculations, axonal hyperecitability, and motoneuronal death in amyotrophic lateral sclerosis. Clinical Neurophysiology, 2014. 125(5): p. 872–873. [DOI] [PubMed] [Google Scholar]
  • 13.Iwai Y, Shibuya K, Misawa S, Sekiguchi Y, Watanabe K, Amino H, et al. , Axonal Dysfunction Precedes Motor Neuronal Death in Amyotrophic Lateral Sclerosis. PLoS ONE, 2016. 11(7): p. e0158596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Martin LJ, Mitochondrial pathobiology in ALS. Journal of bioenergetics and biomembranes, 2011. 43(6): p. 569–579. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Kanai K, Kuwabara S, Misawa S, Tamura N, Ogawara K, Nakata M, et al. , Altered axonal excitability properties in amyotrophic lateral sclerosis: impaired potassium channel function related to disease stage. Brain, 2006. 129(4): p. 953–962. [DOI] [PubMed] [Google Scholar]
  • 16.Pieri M, Carunchio I, Curcio L, Mercuri NB, and Zona C, Increased persistent sodium current determines cortical hyperexcitability in a genetic model of amyotrophic lateral sclerosis. Experimental Neurology, 2009. 215(2): p. 368–379. [DOI] [PubMed] [Google Scholar]
  • 17.Martínez-Silva M.d.L., Imhoff-Manuel RD, Sharma A, Heckman CJ, Shneider NA, Roselli F, et al. , Hypoexcitability precedes denervation in the large fast-contracting motor units in two unrelated mouse models of ALS. eLife, 2018. 7: p. e30955. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Shimizu T, Fujimaki Y, Nakatani-Enomoto S, Matsubara S, Watabe K, Rossini PM, et al. , Complex fasciculation potentials and survival in amyotrophic lateral sclerosis. Clinical Neurophysiology, 2014. 125(5): p. 1059–1064. [DOI] [PubMed] [Google Scholar]
  • 19.Cheah BC, Lin CSY, Park SB, Vucic S, Krishnan AV, and Kiernan MC, Progressive axonal dysfunction and clinical impairment in amyotrophic lateral sclerosis. Clinical Neurophysiology, 2012. 123(12): p. 2460–2467. [DOI] [PubMed] [Google Scholar]
  • 20.Sirabella R, Valsecchi V, Anzilotti S, Cuomo O, Vinciguerra A, Cepparulo P, et al. , Ionic Homeostasis Maintenance in ALS: Focus on New Therapeutic Targets. Frontiers in Neuroscience, 2018. 12: p. 510. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Leal SS and Gomes CM, Calcium dysregulation links ALS defective proteins and motor neuron selective vulnerability. Frontiers in Cellular Neuroscience, 2015. 9(225). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Jaiswal MK, Calcium, mitochondria, and the pathogenesis of ALS: the good, the bad, and the ugly. Frontiers in Cellular Neuroscience, 2013. 7: p. 199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Chang Q and Martin LJ, Voltage-gated calcium channels are abnormal in cultured spinal motoneurons in the G93A-SOD1 transgenic mouse model of ALS. Neurobiology of Disease, 2016. 93: p. 78–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Selvaraj BT, Livesey MR, Zhao C, Gregory JM, James OT, Cleary EM, et al. , C9ORF72 repeat expansion causes vulnerability of motor neurons to Ca2+-permeable AMPA receptor-mediated excitotoxicity. Nature Communications, 2018. 9(1): p. 347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Van Den Bosch L, Van Damme P, Bogaert E, and Robberecht W, The role of excitotoxicity in the pathogenesis of amyotrophic lateral sclerosis. Biochimica et Biophysica Acta (BBA) - Molecular Basis of Disease, 2006. 1762(11): p. 1068–1082. [DOI] [PubMed] [Google Scholar]
  • 26.Platt SR, The role of glutamate in central nervous system health and disease – A review. The Veterinary Journal, 2007. 173(2): p. 278–286. [DOI] [PubMed] [Google Scholar]
  • 27.Traynelis SF, Wollmuth LP, McBain CJ, Menniti FS, Vance KM, Ogden KK, et al. , Glutamate Receptor Ion Channels: Structure, Regulation, and Function. Pharmacological Reviews, 2010. 62(3): p. 405–496. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Corona JC and Tapia R, Ca2+-permeable AMPA receptors and intracellular Ca2+ determine motoneuron vulnerability in rat spinal cord in vivo. Neuropharmacology, 2007. 52(5): p. 1219–1228. [DOI] [PubMed] [Google Scholar]
  • 29.Norante RP, Massimino ML, Lorenzon P, De Mario A, Peggion C, Vicario M, et al. , Generation and validation of novel adeno-associated viral vectors for the analysis of Ca2+ homeostasis in motor neurons. Scientific Reports, 2017. 7(1): p. 6521. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Zhao P, Ignacio S, Beattie E, and Abood ME, Altered pre-symptomatic AMPA and cannabinoid receptor trafficking in motor neurons of ALS model mice: implications for excitotoxicity. The European journal of neuroscience, 2008. 27(3): p. 572–579. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Tortarolo M, Grignaschi G, Calvaresi N, Zennaro E, Spaltro G, Colovic M, et al. , Glutamate AMPA receptors change in motor neurons of SOD1G93A transgenic mice and their inhibition by a noncompetitive antagonist ameliorates the progression of amytrophic lateral sclerosis-like disease. Journal of Neuroscience Research, 2006. 83(1): p. 134–146. [DOI] [PubMed] [Google Scholar]
  • 32.Rembach A, Turner BJ, Bruce S, Cheah IK, Scott RL, Lopes EC, et al. , Antisense peptide nucleic acid targeting GluR3 delays disease onset and progression in the SOD1 G93A mouse model of familial ALS. Journal of Neuroscience Research, 2004. 77(4): p. 573–582. [DOI] [PubMed] [Google Scholar]
  • 33.Kawahara Y, Ito K, Sun H, Aizawa H, Kanazawa I, and Kwak S, RNA editing and death of motor neurons. Nature, 2004. 427: p. 801. [DOI] [PubMed] [Google Scholar]
  • 34.Whitney NP, Peng H, Erdmann NB, Tian C, Monaghan DT, and Zheng JC, Calcium-permeable AMPA receptors containing Q/R-unedited GluR2 direct human neural progenitor cell differentiation to neurons. The FASEB Journal, 2008. 22(8): p. 2888–2900. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Burnashev N, Monyer H, Seeburg PH, and Sakmann B, Divalent ion permeability of AMPA receptor channels is dominated by the edited form of a single subunit. Neuron, 1992. 8(1): p. 189–198. [DOI] [PubMed] [Google Scholar]
  • 36.Hideyama T, Yamashita T, Suzuki T, Tsuji S, Higuchi M, Seeburg PH, et al. , Induced Loss of ADAR2 Engenders Slow Death of Motor Neurons from Q/R Site-Unedited GluR2. The Journal of Neuroscience, 2010. 30(36): p. 11917–11925. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Yamashita T and Kwak S, The molecular link between inefficient GluA2 Q/R site-RNA editing and TDP-43 pathology in motor neurons of sporadic amyotrophic lateral sclerosis patients. Brain Research, 2014. 1584: p. 28–38. [DOI] [PubMed] [Google Scholar]
  • 38.Kawahara Y, Sun H, Ito K, Hideyama T, Aoki M, Sobue G, et al. , Underediting of GluR2 mRNA, a neuronal death inducing molecular change in sporadic ALS, does not occur in motor neurons in ALS1 or SBMA. Neuroscience Research, 2006. 54(1): p. 11–14. [DOI] [PubMed] [Google Scholar]
  • 39.Hideyama T, Yamashita T, Aizawa H, Tsuji S, Kakita A, Takahashi H, et al. , Profound downregulation of the RNA editing enzyme ADAR2 in ALS spinal motor neurons. Neurobiology of Disease, 2012. 45(3): p. 1121–1128. [DOI] [PubMed] [Google Scholar]
  • 40.Yamashita T, Akamatsu M, and Kwak S, Altered Intracellular Milieu of ADAR2-Deficient Motor Neurons in Amyotrophic Lateral Sclerosis. Genes, 2017. 8(2): p. 60. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Musarò A, Understanding ALS: new therapeutic approaches. The FEBS Journal, 2013. 280(17): p. 4315–4322. [DOI] [PubMed] [Google Scholar]
  • 42.Tran LT, Gentil BJ, Sullivan KE, and Durham HD, The voltage-gated calcium channel blocker lomerizine is neuroprotective in motor neurons expressing mutant SOD1, but not TDP-43. Journal of Neurochemistry, 2014. 130(3): p. 455–466. [DOI] [PubMed] [Google Scholar]
  • 43.Carlin KP, Jiang Z, and Brownstone RM, Characterization of calcium currents in functionally mature mouse spinal motoneurons. European Journal of Neuroscience, 2000. 12(5): p. 1624–1634. [DOI] [PubMed] [Google Scholar]
  • 44.Kubat Öktem E, Mruk K, Chang J, Akin A, Kobertz WR, and Brown RH, Mutant SOD1 protein increases Nav1.3 channel excitability. Journal of Biological Physics, 2016. 42(3): p. 351–370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Kuo JJ, Siddique T, Fu R, and Heckman CJ, Increased persistent Na+ current and its effect on excitability in motoneurones cultured from mutant SOD1 mice. The Journal of Physiology, 2005. 563(3): p. 843–854. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Jiang MC, Adimula A, Birch D, and Heckman CJ, Hyperexcitability in synaptic and firing activities of spinal motoneurons in an adult mouse model of amyotrophic lateral sclerosis. Neuroscience, 2017. 362: p. 33–46. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Kuo JJ, Schonewille M, Siddique T, Schults ANA, Fu R, Bär PR, et al. , Hyperexcitability of Cultured Spinal Motoneurons From Presymptomatic ALS Mice. Journal of Neurophysiology, 2004. 91(1): p. 571–575. [DOI] [PubMed] [Google Scholar]
  • 48.Sareen D, O’Rourke JG, Meera P, Muhammad AKMG, Grant S, Simpkinson M, et al. , Targeting RNA Foci in iPSC-Derived Motor Neurons from ALS Patients with a C9ORF72 Repeat Expansion. Science Translational Medicine, 2013. 5(208): p. 208ra149–208ra149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.van Zundert B, Peuscher MH, Hynynen M, Chen A, Neve RL, Brown RH, et al. , Neonatal Neuronal Circuitry Shows Hyperexcitable Disturbance in a Mouse Model of the Adult-Onset Neurodegenerative Disease Amyotrophic Lateral Sclerosis. The Journal of Neuroscience, 2008. 28(43): p. 10864–10874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Delestrée N, Manuel M, Iglesias C, Elbasiouny SM, Heckman CJ, and Zytnicki D, Adult spinal motoneurones are not hyperexcitable in a mouse model of inherited amyotrophic lateral sclerosis. The Journal of Physiology, 2014. 592(7): p. 1687–1703. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Hadzipasic M, Tahvildari B, Nagy M, Bian M, Horwich AL, and McCormick DA, Selective degeneration of a physiological subtype of spinal motor neuron in mice with SOD1-linked ALS. Proceedings of the National Academy of Sciences, 2014. 111(47): p. 16883–16888. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.T K, Persistent Na-channels: origin and function. Areview Janos Salanki memory lecture. Acta Biologica Hungarica, 2008. 59(Supplement 2). [DOI] [PubMed] [Google Scholar]
  • 53.Naujock M, Stanslowsky N, Bufler S, Naumann M, Reinhardt P, Sterneckert J, et al. , 4-Aminopyridine Induced Activity Rescues Hypoexcitable Motor Neurons from Amyotrophic Lateral Sclerosis Patient-Derived Induced Pluripotent Stem Cells. STEM CELLS, 2016. 34(6): p. 1563–1575. [DOI] [PubMed] [Google Scholar]
  • 54.Pieri M, Albo F, Gaetti C, Spalloni A, Bengtson CP, Longone P, et al. , Altered excitability of motor neurons in a transgenic mouse model of familial amyotrophic lateral sclerosis. Neuroscience Letters, 2003. 351(3): p. 153–156. [DOI] [PubMed] [Google Scholar]
  • 55.Zona C, Pieri M, and Carunchio I, Voltage-Dependent Sodium Channels in Spinal Cord Motor Neurons Display Rapid Recovery From Fast Inactivation in a Mouse Model of Amyotrophic Lateral Sclerosis. Journal of Neurophysiology, 2006. 96(6): p. 3314–3322. [DOI] [PubMed] [Google Scholar]
  • 56.Leroy F, Lamotte d’Incamps B, Imhoff-Manuel RD, and Zytnicki D, Early intrinsic hyperexcitability does not contribute to motoneuron degeneration in amyotrophic lateral sclerosis. eLife, 2014. 3: p. e04046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Geevasinga N, Menon P, Howells J, Nicholson GA, Kiernan MC, and Vucic S, Axonal ion channel dysfunction in c9orf72 familial amyotrophic lateral sclerosis. JAMA Neurology, 2015. 72(1): p. 49–57. [DOI] [PubMed] [Google Scholar]
  • 58.Vucic S and Kiernan MC, Upregulation of persistent sodium conductances in familial ALS. Journal of Neurology, Neurosurgery & Psychiatry, 2010. 81(2): p. 222–227. [DOI] [PubMed] [Google Scholar]
  • 59.Alessandri-Haber N, Alcaraz G, Deleuze C, Jullien F, Manrique C, Couraud F, et al. , Molecular determinants of emerging excitability in rat embryonic motoneurons. The Journal of Physiology, 2002. 541(1): p. 25–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Goldin AL, Diversity of Mammalian Voltage-Gated Sodium Channels. Annals of the New York Academy of Sciences, 1999. 868(1): p. 38–50. [DOI] [PubMed] [Google Scholar]
  • 61.Waxman SG, Kocsis JD, and Black JA, Type III sodium channel mRNA is expressed in embryonic but not adult spinal sensory neurons, and is reexpressed following axotomy. Journal of Neurophysiology, 1994. 72(1): p. 466–470. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Chiò A, Benzi G, Dossena M, Mutani R, and Mora G, Severely increased risk of amyotrophic lateral sclerosis among Italian professional football players. Brain, 2005. 128(3): p. 472–476. [DOI] [PubMed] [Google Scholar]
  • 63.Pupillo E, Poloni M, Bianchi E, Giussani G, Logroscino G, Zoccolella S, et al. , Trauma and amyotrophic lateral sclerosis: a european population-based case-control study from the EURALS consortium. Amyotrophic Lateral Sclerosis and Frontotemporal Degeneration, 2018. 19(1–2): p. 118–125. [DOI] [PubMed] [Google Scholar]
  • 64.Nutini M, Spalloni A, Florenzano F, Westenbroek RE, Marini C, Catterall WA, et al. , Increased expression of the beta3 subunit of voltage-gated Na+ channels in the spinal cord of the SOD1G93A mouse. Molecular and Cellular Neuroscience, 2011. 47(2): p. 108118. [DOI] [PubMed] [Google Scholar]
  • 65.Crill WE, Persistent Sodium Current in Mammalian Central Neurons. Annual Review of Physiology, 1996. 58(1): p. 349–362. [DOI] [PubMed] [Google Scholar]
  • 66.Shah BS, Stevens EB, Pinnock RD, Dixon AK, and Lee K, Developmental expression of the novel voltage-gated sodium channel auxiliary subunit β3, in rat CNS. The Journal of Physiology, 2001. 534(Pt 3): p. 763–776. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Wainger Brian J., Kiskinis E, Mellin C, Wiskow O, Han Steve S.W., Sandoe J, et al. , Intrinsic Membrane Hyperexcitability of Amyotrophic Lateral Sclerosis Patient-Derived Motor Neurons. Cell Reports, 2014. 7(1): p. 1–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Devlin A-C, Burr K, Borooah S, Foster JD, Cleary EM, Geti I, et al. , Human iPSC-derived motoneurons harbouring TARDBP or C9ORF72 ALS mutations are dysfunctional despite maintaining viability. Nature Communications, 2015. 6: p. 5999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Park SB, Vucic S, Cheah BC, Lin CSY, Kirby A, Mann KP, et al. , Flecainide in Amyotrophic Lateral Sclerosis as a Neuroprotective Strategy (FANS): A Randomized Placebo-Controlled Trial. EBioMedicine, 2015. 2(12): p. 1916–1922. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Shibuya K, Misawa S, Arai K, Nakata M, Kanai K, Yoshiyama Y, et al. , Markedly reduced axonal potassium channel expression in human sporadic amyotrophic lateral sclerosis: An immunohistochemical study. Experimental Neurology, 2011. 232(2): p. 149–153. [DOI] [PubMed] [Google Scholar]
  • 71.Figueroa-Romero C, Hur J, Bender DE, Delaney CE, Cataldo MD, Smith AL, et al. , Identification of Epigenetically Altered Genes in Sporadic Amyotrophic Lateral Sclerosis. PLoS ONE, 2012. 7(12): p. e52672. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Gunasekaran R, Narayani RS, Vijayalakshmi K, Alladi PA, Shobha K, Nalini A, et al. , Exposure to cerebrospinal fluid of sporadic Amyotrophic Lateral Sclerosis patients alters Nav1.6 and Kv1.6 channel expression in rat spinal motor neurons. Brain Research, 2009. 1255: p. 170–179. [DOI] [PubMed] [Google Scholar]
  • 73.Vucic S, Krishnan AV, and Kiernan MC, Fatigue and activity dependent changes in axonal excitability in amyotrophic lateral sclerosis. Journal of Neurology, Neurosurgery & Psychiatry, 2007. 78(11): p. 1202–1208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Rothstein JD, Edaravone: A new drug approved for ALS. Cell, 2017. 171(4): p. 725. [DOI] [PubMed] [Google Scholar]

RESOURCES