Next Article in Journal
Nerve Biopsy in Peripheral Neuropathies: Not All Water Is under the Bridge
Next Article in Special Issue
Mutations and Protein Interaction Landscape Reveal Key Cellular Events Perturbed in Upper Motor Neurons with HSP and PLS
Previous Article in Journal
Citation Classics in Consumer Neuroscience, Neuromarketing and Neuroaesthetics: Identification and Conceptual Analysis
Previous Article in Special Issue
Hereditary Spastic Paraplegia: From Genes, Cells and Networks to Novel Pathways for Drug Discovery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Neurophysiological Mechanisms Underlying Cortical Hyper-Excitability in Amyotrophic Lateral Sclerosis: A Review

Faculty of Medicine, School of Biomedical Sciences, The University of Queensland, Brisbane, QLD 4072, Australia
*
Author to whom correspondence should be addressed.
Brain Sci. 2021, 11(5), 549; https://doi.org/10.3390/brainsci11050549
Submission received: 2 April 2021 / Revised: 23 April 2021 / Accepted: 26 April 2021 / Published: 27 April 2021
(This article belongs to the Special Issue Factors Responsible for CSMN Vulnerability)

Abstract

:
Amyotrophic lateral sclerosis (ALS) is a progressive neuromotor disease characterized by the loss of upper and lower motor neurons (MNs), resulting in muscle paralysis and death. Early cortical hyper-excitability is a common pathological process observed clinically and in animal disease models. Although the mechanisms that underlie cortical hyper-excitability are not completely understood, the molecular and cellular mechanisms that cause enhanced neuronal intrinsic excitability and changes in excitatory and inhibitory synaptic activity are starting to emerge. Here, we review the evidence for an anterograde glutamatergic excitotoxic process, leading to cortical hyper-excitability via intrinsic cellular and synaptic mechanisms and for the role of interneurons in establishing disinhibition in clinical and experimental settings. Understanding the mechanisms that lead to these complex pathological processes will likely produce key insights towards developing novel therapeutic strategies to rescue upper MNs, thus alleviating the impact of this fatal disease.

1. Introduction

First described by Jean-Martin Charcot in 1869, amyotrophic lateral sclerosis (ALS), one of the most common neuromotor diseases, is characterized by an inexorable loss of upper and lower motor neurons (MNs). The upper motor neurons of the primary motor cortex are large pyramidal neurons in layer V (LVPNs), whose axonal projections form the corticospinal tracts, which descend to directly or indirectly excite the lower motor neurons (LMNs) of brainstem and the spinal cord [1,2]. Although lower MNs are the final output pathway that integrate motor output and regulate muscle activity, the presence of corticospinal monosynaptic or polysynaptic axonal projections that originate in the primary motor cortex, dorsal prefrontal cortex and somatosensory cortex [3,4] and relay cortical output to the LMNs allow fine regulation of neuromotor execution in humans [5]. There is substantial evidence, both clinically and in animal models, suggesting significant cortical disruption early in ALS. However, what makes the upper MNs vulnerable to degeneration and how this perturbs the neuromotor circuitry is unclear.
The clinical outcome in ALS is largely based on lower MN degeneration, presenting mostly as the onset of weakness in limb (80%) or bulbar (20%) muscles [6]. A relentless progression occurs, with muscle weakness spreading to multiple muscles, paralysis and death from respiratory failure within two to three years of diagnosis in 50% of ALS patients [7]. Cognitive abnormalities are increasingly recognized in ALS patients, with fronto-temporal dementia in about 5 to 15% ALS cases, worsening the prognosis and disease progression [8,9]. Classically, a clear family history is present in only 10% of all ALS cases and has been increasingly (but not always) associated with a genetic cause, while the remaining 90% have no family history and are classified as sporadic ALS (sALS), although a genetic cause can be identified in about 10% of these patients [2,10,11]. This makes ALS a disease of complex pathology and the continued discovery of new genetic causes highlights this.
In an attempt to identify the origin of neurodegeneration and the correlation between upper MN and lower MN dysfunction, the “dying forward” [5] hypothesis was conceptualized. One proposed pathogenic mechanism that culminates in the death of upper and lower MNs is glutamate-induced excitotoxicity which results from either excessive presynaptic glutamate release or defective glutamate reuptake. This depolarizes the post-synaptic neuron and increases calcium influx, thus leading to hyper-excitability [12,13]. Hyper-excitability is a commonly observed feature of different cell types at several locations, including LVPNs [14,15] brainstem motor neurons [16], spinal motor neurons [17,18] and skeletal muscles [19] that may contribute to pathology in ALS. Cortical hyper-excitability seems to be one of the most common features preceding the degeneration of upper MNs [14,15,20] and LMNs [21,22]. This review concentrates on the neurophysiology of hyper-excitability in upper MNs and the probable mechanisms involved.

2. Excitability and Hyper-Excitability in ALS: General Concepts

Excitability of a neuron is a measure of its electrophysiological properties. This relates to the neuron’s ability to produce an output or action potential (AP) by depolarizing its membrane potential to a threshold level in response to an input. This intrinsic ability of a neuron is referred to as its “intrinsic excitability” and is defined by factors including the types and number of excitatory receptors and voltage gated ion channels (e.g., Na+, K+) present, which shapes the neuronal output [12]. Similarly, neuronal activity typically refers to the frequency of spontaneously elicited APs or post-synaptic currents, and this depends on the neuron’s intrinsic excitability, its synaptic input strength and excitation/inhibition balance, and the coordination of excitatory and inhibitory synaptic input to generate an AP.
The concept of functional hyper-excitability or exaggerated neuronal output in ALS involving cortical dysfunction has been described by many [23,24] and is considered a hallmark feature which may contribute to an excitotoxic environment via altered intrinsic neuronal properties or synaptic mechanisms [14,15,25,26]. Clinically, hyper-excitability in upper MNs is directly associated with corticospinal tract abnormality and is characterized by the presence of enhanced muscle tone and increased reflexes, including brisk reflexes and spasticity [27,28], symptoms that are considered clinical hallmarks of cortical involvement. Moreover, cortical dysfunction and hyper-excitability are now considered as early diagnostic signs of ALS [29].

2.1. Hyper-Excitability in Upper MNs: Influence on Cortical Dysfunction

Eisen and colleagues first emphasized the direct cortical involvement in ALS with their dying forward hypothesis, whereby the primary cortical dysfunction in upper MNs triggers the subsequent pathologic degeneration and death of lower MNs [5,29]. A key part of this hypothesis was the evidence that lower MNs in Onuf’s, oculomotor and abducens nuclei, which are spared or mildly affected late in ALS, lack direct monosynaptic neuronal projections in humans, indicating a role for direct cortical inputs in relaying cortical dysfunction [5]. Additionally, cortical hyper-excitability is observed prior to the symptomatic stage clinically [21,22,30] and in experimental animal models [14,15,20,31], clearly underlining a role for anterograde cortical dysfunction in ALS. Importantly, recent advances in clinical, molecular, neuroimaging and genetic studies further emphasize the significant pathological influence of hyper-excitability-induced upper MN dysfunction in ALS [29,32]. Moreover, cortical hyper-excitability-associated excitotoxicity is a result of enhanced intrinsic properties and synaptic dysfunction [33].
A counter to the original dying forward hypothesis is that direct monosynaptic contacts between upper and lower MNs are thought to direct fine voluntary motor control and are found in humans and some primates [34], although both mono- and polysynaptic connections co-exist in the same species [35,36]. However, despite a lack of direct corticospinal inputs onto lower MNs, rodent models of ALS show prominent upper MN degeneration preceding lower MN loss [4,37,38], while selective knockdown of ALS gene mutations in rodent cortex delays disease progress and lower MN loss and extends survival [39,40]. As corticospinal influence is relayed to lower MNs by local oligosynaptic excitatory interneuronal networks in non-primate species [41,42], it seems likely that this circuitry can still mediate a dying forward influence. Hence, despite variation in the morphological and functional properties of these neurons between species, lower MNs are directly or indirectly modulated by upper MNs via corticospinal descending projections from the motor cortex [41,42,43,44] and this influence from upper MNs are deleterious for lower MN survival in ALS [39,40]. A recent clinical study in ALS onset patients showing absence of hyper-excitability in spinal MNs [45] could possibly reflect a compensatory mechanism in the surviving MNs, given that not all lower MNs are hyper-excitable and their excitability status correlates with their survival [46,47,48]. Upper MN degeneration precedes lower MN loss [4,37,38], clearly implying the possibility of a dying forward influence in ALS.

2.1.1. Intrinsic Excitability-Mediated Cortical Dysfunction

The intrinsic excitability of a neuron reflects its electrical properties, comprising the membrane density and distribution of ionic conductances and receptors present on its particular dendritic morphology. The maintenance of appropriate functional output is strongly influenced by the number and properties of synaptic inputs [49]. Increased depolarizing currents are reported to result in an enhanced firing rate [49,50], and hence, it is highly likely that increased intrinsic excitability and excitotoxic susceptibility in ALS are linked [20].
One prime factor affecting the neuronal excitability is the persistent inward current (PIC) generated by a inactivation-resistant population of Na+ channels, referred to as the persistent sodium current (INaP) [51]. In ALS, the intrinsic excitability of cortical MNs are positively skewed by the enhanced current density of INaP, increasing the firing frequency triggered by elevated Na+ current and depleted K+ conductance [20] and producing persistent depolarizations due to rapid activation and slow decay kinetics [52,53]. Several other studies reinforce this finding, including increased AP firing observed in layer V pyramidal neurons [15] and in motor cortex slices [54]. More recently, the NaV1.6 channel localized in the initial segment of the axon have been shown to be overexpressed in cortical MNs in ALS [55], to the point where AP firing initiates [56], clearly confirming sustained depolarization-induced hyper-excitability in dying cortical MNs [20]. A depolarizing stimulus will activate any PIC present to generate AP firing in MNs [50], and hence, elevated PIC observed in ALS can contribute to hyper-excitability-induced excitotoxicity [20] by directly enhancing intracellular Ca2+ influx and MN firing frequency [57,58]. Furthermore, the selective suppression of the INaP and firing frequency in LMNs in animal models of ALS [16,59,60], and in clinical studies of cortical hyper-excitability [61] by riluzole, a glutamatergic antagonist and inhibitor of Na+ channel activity, further reinforces the correlation between increased PIC and cortical dysfunction [20,54,61].
Similarly, another contributor to MN hyper-excitability in ALS is reduced K+ conductance, documented in clinical studies showing peripheral motor axonal excitability [62,63] and in patient-derived MNs showing reduced amplitude of delayed rectifier K+ currents [64]. However, whether the corticomotor neurons exhibit reduced K+ conductance remains to be explored. Activating Kv7.2/3 potassium channels (also known as the M current) by retigabine, an activator of Kv7.2/3 potassium channels drives K+ efflux, counterbalancing the Na+ influx and thus dampening MN hyper-excitability as observed in ALS patient-derived MNs [64] and LVPNs of mouse models [65]. This evidence supports the idea that intrinsic neuronal properties resulting in hyper-excitability may be a combination of increased Na+ and decreased K+ conductances, providing additional avenues for therapeutic exploration.

2.1.2. Synaptic Excitability-Mediated Cortical Dysfunction

Cortical hyper-excitability in ALS is also associated with both functional and morphological synaptic changes [14,15,31,66]. The synaptic regulation of neuronal activity is determined via the number, distribution and timing of excitatory and inhibitory neurotransmission. A major part of synaptic dysfunction in ALS is increased glutamatergic excitatory neurotransmission, giving rise to excessive glutamate, a key driver of excitotoxicity [67,68]. During excitatory neurotransmission, glutamate is secreted from the pre-synaptic neuron and acts to rapidly activate AMPA and NMDA receptors [68]. Extracellular glutamate levels are regulated within the CNS via interactive functions, namely uptake and synthesis through glutamate transporters and the levels of excitatory and inhibitory neurotransmission [68]. Taking into account the negative effects of glutamate reported on neurons [68], the enhanced glutamate observed in ALS patients provides a clear correlation between increased glutamate levels, hyper-excitability and excitotoxicity. Enhanced glutamate mediated excitotoxicity in ALS has been shown to be a combination of excessive pre-synaptic release [67,69] and inhibited uptake from the synaptic cleft due to ineffective astroglial glutamate transporters (EAAT2) observed in the spinal cord [70] and motor cortex of affected ALS patients [71,72] and in rodent models of ALS [73,74,75], resulting in glutamate retention and persistent enhanced post-synaptic NMDA and AMPA receptor activation [67,68], causing an increased intracellular influx of Ca2+ and Na+ ions resulting in a synaptic hyper-active state [69]. Additionally, the Ca2+ buffering capacity of MNs in ALS is reported to be impaired, with defective Ca2+ ATPase and Ca2+/Na+ exchangers exacerbating the increased intracellular Ca2+ burden [76] leading to Ca2+ overabundance and sustained depolarization [58,76,77]. Glutamatergic hyper-excitation is observed as the enhanced spontaneous excitatory post-synaptic current (sEPSC) frequency of LVPNs [14,15]. This cascade of events in cortical hyper-excitability and dysfunction evolves into neuronal degeneration, as mitochondrial defects cause oxidative stress-generating reactive oxygen species, eventually leading to the death of the MN [13].

2.1.3. Clinical Evidence of Functional and Structural Cortical Hyper-Excitability

Clinical observation of functional and structural cortical excitatory measurements in ALS patients serve as a marker of altered cortical output function, i.e., the cortical state of excitability. Enhanced cortical excitability prior to symptom onset due to significant lower MN degeneration has been observed in ALS patients by utilizing advanced clinical neurophysiological and neuroimaging techniques [29], suggesting that cortical dysfunction may be a significant early diagnostic marker in ALS [29,78]. Trans-cranial magnetic stimulation, an effective tool for evaluating functional cortical excitability, demonstrated a decrease in short-interval intracortical inhibition (a measure of cortical inhibitory function), increased cortico-conduction time and enhanced intracortical facilitation in Betz cells (layer V output neurons) in ALS patients [21,22,78,79,80]. Short interval intercortical inhibition (SICI) is an activation of a subthreshold inhibitory circuit via subthreshold stimulus, which represents an equilibrium between the robust inhibitory effect mediated by GABAergic interneurons in conjunction with weaker cortical glutamatergic facilitatory effects from pyramidal neurons. This process increases the threshold for the generation of an evoked response; hence, the observation of low threshold elicitation of a motor-evoked potential (MEP) indicates cortical hyper-excitability [22,24,81]. Given that synaptic transmission is a balance between inhibitory and excitatory transmission, the presence of reduced SICI, reflecting inhibitory interneuronal dysfunction [82,83], alongside increased MEP amplitude, indicating enhanced excitatory transmission mediated via glutamate [84], clearly signals that cortical hyperexcitability is most likely a combination of cortical disinhibition and enhanced excitation. Furthermore, functional MRI studies provide additional evidence for early cortical structural dysfunction, demonstrating cortical thinning of the primary motor cortex and temporal cortex [85,86]. Likewise, diffusion tensor imaging showed abnormalities in the corticospinal tracts projecting towards the lower MNs in ALS patients [87,88,89,90], suggesting the loss of cortical MNs and degeneration of their corticospinal projections. Additionally, the partial reduction of cortical hyper-excitability in ALS patients by riluzole administration further reinforces the pathogenic action of cortical hyper-excitability in ALS [61,91].

3. Morphological Changes Correlated with Cortical Hyper-Excitability

Corticospinal neurons, namely LVPNs in the deep motor cortical layer, are the source of corticospinal projections to spinal cord controlling voluntary movement execution. Besides direct corticospinal projections, LVPNs also make direct corticofugal connections to the thalamus, cortico-cortical connections to layer 4 and layer 2/3 motor cortex, and corticobulbar connections to motor nuclei in the brainstem [92]. Evidence of early upper MN degeneration has been demonstrated in various rodent ALS models [4,37,93,94] and more subtle changes in LVPN morphology have been observed as an early loss of dendritic spines and shrinkage of apical dendrites [14,66,95]. Additionally, in ALS patients, Betz cells of the motor cortex (the largest LVPNs) also display apical dendrite degeneration with intensive vacuolation, disintegration of their dendritic architecture and reduced synapse number, indicative of functional changes in cortical neurons [96].

4. The Role of Inhibitory Cortical Interneurons in Cortical Hyper-Excitability

The maintenance of normal neuronal activity in the cortical neural circuit is an interaction between the excitatory pyramidal neurons and abundant interneurons that mainly cause inhibition via γ-aminobutyric acid (GABA) neurotransmission. GABAergic interneurons in the motor cortex can be activated via a sub-threshold stimulus [97]. In ALS patients, dysfunction of cortical inhibitory interneuronal activity is observed as reduced or absent SICI mediated by cortical GABAergic interneurons [22,78,81], alongside increased cortico-neuronal excitability; this clearly implicates an impaired inhibition/excitation balance induction of cortical hyper-excitability. Additionally, reduced levels of GABA in the motor cortex [98], decreased mRNA expression of the alpha subunit of GABAA receptors [99] and functional neuroimaging studies revealing decreased binding of flumazenil, a GABAA receptor ligand, in sporadic ALS patients [100] provide direct evidence that decreased inhibition may contribute to cortical hyper-excitability. Furthermore, histological studies have shown reduced parvalbumin interneuron numbers in post-mortem motor cortices of ALS patients [101] and animal models of ALS [102,103], and a decreased number of neuropeptide Y-positive interneurons at symptom onset coupled with progressive reduction of calretinin-positive interneurons from symptom onset up to the end stage in the motor cortex of an ALS mouse model [104]. A more recent electrophysiological study revealed decreased intracortical inhibition in layer V interneurons of motor cortex and a rescue of this disinhibition-induced hyper-excitability by increasing interneuronal activity [102]. Additionally, cortical parvalbumin interneurons [105] and somatostatin interneurons [106] have also been shown to display increased intrinsic and synaptic excitability. Given that cortical interneurons can display either hyper-excitability [105,106] or hypo-excitability [102], it remains to be determined whether cortical hyper-excitability in a rodent model occurs in addition to, or as a consequence of, the hyper-excitability of cortical excitatory and inhibitory neuronal populations. Taken together, these findings clearly imply a role for interneuron dysfunction causing an imbalance in cortical excitation/inhibition and thus hyper-excitability in ALS.

5. Conclusions

Hyper-excitability is a prevalent feature observed clinically in ALS patients and in animal models for ALS. What causes cortical hyper-excitability and how hyper-excitability drives upper MN degeneration and death is just beginning to be unraveled. Evidence from electrophysiological and histological studies in humans and in animal models provides insight into an early cortical involvement in the disease process that precedes lower MN degeneration. Cortical hyper-excitability is clearly a result of increased excitation and impaired inhibition, and the evidence so far points to a number of molecular and cellular mechanisms that contribute to cortical hyper-excitability and that may drive the degeneration of upper MNs. It will be critically important to further characterize these mechanisms and understand their consequences, in order to develop rational therapeutic strategies that normalize hyper-excitation and alleviate the disease.

Author Contributions

Writing—original draft preparation, J.P.; writing—review and editing, M.C.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Australian National Health and Medical Research Council (grant APP1065884), Motor Neuron Research Institute of Australia and FightMND.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cleveland, D.W.; Rothstein, J.D. From Charcot to Lou Gehrig: Deciphering selective motor neuron death in ALS. Nat. Rev. Neurosci. 2001, 2, 806–819. [Google Scholar] [CrossRef] [PubMed]
  2. Turner, M.R.; Hardiman, O.; Benatar, M.; Brooks, B.R.; Chio, A.; de Carvalho, M.; Ince, P.G.; Lin, C.; Miller, R.G.; Mitsumoto, H.; et al. Controversies and priorities in amyotrophic lateral sclerosis. Lancet Neurol. 2013, 12, 310–322. [Google Scholar] [CrossRef] [Green Version]
  3. Seo, J.P.; Jang, S.H. Different characteristics of the corticospinal tract according to the cerebral origin: DTI study. Am. J. Neuroradiol. 2013, 34, 1359. [Google Scholar] [CrossRef] [Green Version]
  4. Marques, C.; Burg, T.; Scekic-Zahirovic, J.; Fischer, M.; Rouaux, C. Upper and lower motor neuron degenerations are somatotopically related and temporally ordered in the Sod1 mouse model of amyotrophic lateral sclerosis. Brain Sci. 2021, 11, 369. [Google Scholar] [CrossRef] [PubMed]
  5. Eisen, A.; Kim, S.; Pant, B. Amyotrophic lateral sclerosis (ALS): A phylogenetic disease of the corticomotoneuron? Muscle Nerve 1992, 15, 219–224. [Google Scholar] [CrossRef]
  6. Mitchell, J.D.; Borasio, G.D. Amyotrophic lateral sclerosis. Lancet 2007, 369, 2031–2041. [Google Scholar] [CrossRef]
  7. Pupillo, E.; Messina, P.; Logroscino, G.; Beghi, E. Long-term survival in amyotrophic lateral sclerosis: A population-based study. Ann. Neurol. 2014, 75, 287–297. [Google Scholar] [CrossRef]
  8. Al-Chalabi, A.; Hardiman, O.; Kiernan, M.C.; Chio, A.; Rix-Brooks, B.; van den Berg, L.H. Amyotrophic lateral sclerosis: Moving towards a new classification system. Lancet Neurol. 2016, 15, 1182–1194. [Google Scholar] [CrossRef]
  9. Phukan, J.; Elamin, M.; Bede, P.; Jordan, N.; Gallagher, L.; Byrne, S.; Lynch, C.; Pender, N.; Hardiman, O. The syndrome of cognitive impairment in amyotrophic lateral sclerosis: A population-based study. J. Neurol. Neurosurg. Psychiatry 2012, 83, 102–108. [Google Scholar] [CrossRef] [PubMed]
  10. Chen, S.; Sayana, P.; Zhang, X.; Le, W. Genetics of amyotrophic lateral sclerosis: An update. Mol. Neurodegener. 2013, 8, 28. [Google Scholar] [CrossRef] [Green Version]
  11. Pasinelli, P.; Brown, R.H. Molecular biology of amyotrophic lateral sclerosis: Insights from genetics. Nat. Rev. Neurosci. 2006, 7, 710–723. [Google Scholar] [CrossRef]
  12. Gunes, Z.I.; Kan, V.W.Y.; Ye, X.; Liebscher, S. Exciting Complexity: The Role of Motor Circuit Elements in ALS Pathophysiology. Front. Neurosci. 2020, 14, 573. [Google Scholar] [CrossRef]
  13. Pradhan, J.; Noakes, P.G.; Bellingham, M.C. The role of altered BDNF/TrkB signaling in amyotrophic lateral sclerosis. Front. Cell. Neurosci. 2019, 13. [Google Scholar] [CrossRef]
  14. Fogarty, M.J.; Noakes, P.G.; Bellingham, M.C. Motor cortex layer V pyramidal neurons exhibit dendritic regression, spine loss, and increased synaptic excitation in the presymptomatic hSOD1(G93A) mouse model of amyotrophic lateral sclerosis. J. Neurosci. Off. J. Soc. Neurosci. 2015, 35, 643–647. [Google Scholar] [CrossRef] [Green Version]
  15. Saba, L.; Viscomi, M.T.; Caioli, S.; Pignataro, A.; Bisicchia, E.; Pieri, M.; Molinari, M.; Ammassari-Teule, M.; Zona, C. Altered functionality, morphology, and vesicular glutamate transporter expression of cortical motor neurons from a presymptomatic mouse model of amyotrophic lateral sclerosis. Cereb. Cortex 2016, 26, 1512–1528. [Google Scholar] [CrossRef] [Green Version]
  16. Van Zundert, B.; Peuscher, M.H.; Hynynen, M.; Chen, A.; Neve, R.L.; Brown, R.H., Jr.; Constantine-Paton, M.; Bellingham, M.C. Neonatal neuronal circuitry shows hyperexcitable disturbance in a mouse model of the adult-onset neurodegenerative disease amyotrophic lateral sclerosis. J. Neurosci. Off. J. Soc. Neurosci. 2008, 28, 10864–10874. [Google Scholar] [CrossRef] [PubMed]
  17. Kuo, J.J.; Schonewille, M.; Siddique, T.; Schults, A.N.; Fu, R.; Bar, P.R.; Anelli, R.; Heckman, C.J.; Kroese, A.B. Hyperexcitability of cultured spinal motoneurons from presymptomatic ALS mice. J. Neurophysiol. 2004, 91, 571–575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Kuo, J.J.; Siddique, T.; Fu, R.; Heckman, C.J. Increased persistent Na(+) current and its effect on excitability in motoneurones cultured from mutant SOD1 mice. J. Physiol. 2005, 563, 843–854. [Google Scholar] [CrossRef] [PubMed]
  19. Camerino, G.M.; Fonzino, A.; Conte, E.; De Bellis, M.; Mele, A.; Liantonio, A.; Tricarico, D.; Tarantino, N.; Dobrowolny, G.; Musarò, A.; et al. Elucidating the Contribution of Skeletal Muscle Ion Channels to Amyotrophic Lateral Sclerosis in search of new therapeutic options. Sci. Rep. 2019, 9, 3185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Pieri, M.; Carunchio, I.; Curcio, L.; Mercuri, N.B.; Zona, C. Increased persistent sodium current determines cortical hyperexcitability in a genetic model of amyotrophic lateral sclerosis. Exp. Neurol. 2009, 215, 368–379. [Google Scholar] [CrossRef] [PubMed]
  21. Menon, P.; Geevasinga, N.; van den Bos, M. Cortical hyperexcitability and disease spread in amyotrophic lateral sclerosis. Eur. J. Neurol. 2017, 24, 816–824. [Google Scholar] [CrossRef]
  22. Menon, P.; Kiernan, M.C.; Vucic, S. Cortical hyperexcitability precedes lower motor neuron dysfunction in ALS. Clin. Neurophysiol. 2015, 126, 803–809. [Google Scholar] [CrossRef]
  23. Eisen, A.; Pant, B.; Stewart, H. Cortical excitability in amyotrophic lateral sclerosis: A clue to pathogenesis. Can. J. Neurol. Sci. J. Can. Des. Sci. Neurol. 1993, 20, 11–16. [Google Scholar] [CrossRef] [Green Version]
  24. Vucic, S.; Kiernan, M.C. Novel threshold tracking techniques suggest that cortical hyperexcitability is an early feature of motor neuron disease. Brain J. Neurol. 2006, 129, 2436–2446. [Google Scholar] [CrossRef]
  25. Bae, J.S.; Simon, N.G.; Menon, P.; Vucic, S.; Kiernan, M.C. The puzzling case of hyperexcitability in amyotrophic lateral sclerosis. J. Clin. Neurol. (Seoul Korea) 2013, 9, 65–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Fogarty, M.J. Driven to decay: Excitability and synaptic abnormalities in amyotrophic lateral sclerosis. Brain Res. Bull. 2018, 140, 318–333. [Google Scholar] [CrossRef] [PubMed]
  27. Gordon, P.H. Amyotrophic Lateral Sclerosis: An update for 2013 Clinical Features, Pathophysiology, Management and Therapeutic Trials. Aging Dis. 2013, 4, 295–310. [Google Scholar] [CrossRef] [PubMed]
  28. Wijesekera, L.C.; Leigh, P.N. Amyotrophic lateral sclerosis. Orphanet J. Rare Dis. 2009, 4, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Geevasinga, N.; Menon, P.; Ozdinler, P.H.; Kiernan, M.C.; Vucic, S. Pathophysiological and diagnostic implications of cortical dysfunction in ALS. Nat. Rev. Neurol. 2016, 12, 651–661. [Google Scholar] [CrossRef] [PubMed]
  30. Vucic, S.; Nicholson, G.A.; Kiernan, M.C. Cortical hyperexcitability may precede the onset of familial amyotrophic lateral sclerosis. Brain J. Neurol. 2008, 131, 1540–1550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Fogarty, M.J.; Klenowski, P.M.; Lee, J.D.; Drieberg-Thompson, J.R.; Bartlett, S.E.; Ngo, S.T.; Hilliard, M.A.; Bellingham, M.C.; Noakes, P.G. Cortical synaptic and dendritic spine abnormalities in a presymptomatic TDP-43 model of amyotrophic lateral sclerosis. Sci. Rep. 2016, 6, 37968. [Google Scholar] [CrossRef] [Green Version]
  32. Van den Bos, M.A.J.; Geevasinga, N.; Higashihara, M.; Menon, P.; Vucic, S. Pathophysiology and Diagnosis of ALS: Insights from Advances in Neurophysiological Techniques. Int. J. Mol. Sci. 2019, 20, 2818. [Google Scholar] [CrossRef] [Green Version]
  33. Rothstein, J.D.; Tsai, G.; Kuncl, R.W.; Clawson, L.; Cornblath, D.R.; Drachman, D.B.; Pestronk, A.; Stauch, B.L.; Coyle, J.T. Abnormal excitatory amino acid metabolism in amyotrophic lateral sclerosis. Ann. Neurol. 1990, 28, 18–25. [Google Scholar] [CrossRef] [PubMed]
  34. Lemon, R.N.; Griffiths, J. Comparing the function of the corticospinal system in different species: Organizational differences for motor specialization? Muscle Nerve 2005, 32, 261–279. [Google Scholar] [CrossRef]
  35. Welniarz, Q.; Dusart, I.; Roze, E. The corticospinal tract: Evolution, development, and human disorders. Dev. Neurobiol. 2017, 77, 810–829. [Google Scholar] [CrossRef] [PubMed]
  36. Lemon, R.N. Descending pathways in motor control. Annu. Rev. Neurosci. 2008, 31, 195–218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Ozdinler, P.H.; Benn, S.; Yamamoto, T.H.; Guzel, M.; Brown, R.H., Jr.; Macklis, J.D. Corticospinal motor neurons and related subcerebral projection neurons undergo early and specific neurodegeneration in hSOD1G(9)(3)A transgenic ALS mice. J. Neurosci. Off. J. Soc. Neurosci. 2011, 31, 4166–4177. [Google Scholar] [CrossRef] [Green Version]
  38. Zang, D.W.; Cheema, S.S. Degeneration of corticospinal and bulbospinal systems in the superoxide dismutase 1(G93A G1H) transgenic mouse model of familial amyotrophic lateral sclerosis. Neurosci. Lett. 2002, 332, 99–102. [Google Scholar] [CrossRef]
  39. Burg, T.; Bichara, C.; Scekic-Zahirovic, J.; Fischer, M.; Stuart-Lopez, G.; Brunet, A.; Lefebvre, F.; Cordero-Erausquin, M.; Rouaux, C. Absence of Subcerebral Projection Neurons Is Beneficial in a Mouse Model of Amyotrophic Lateral Sclerosis. Ann. Neurol. 2020, 88, 688–702. [Google Scholar] [CrossRef]
  40. Thomsen, G.M.; Gowing, G.; Latter, J.; Chen, M.; Vit, J.P.; Staggenborg, K.; Avalos, P.; Alkaslasi, M.; Ferraiuolo, L.; Likhite, S.; et al. Delayed disease onset and extended survival in the SOD1G93A rat model of amyotrophic lateral sclerosis after suppression of mutant SOD1 in the motor cortex. J. Neurosci. 2014, 34, 15587–15600. [Google Scholar] [CrossRef] [Green Version]
  41. Alstermark, B.; Ogawa, J. In vivo recordings of bulbospinal excitation in adult mouse forelimb motoneurons. J. Neurophysiol. 2004, 92, 1958–1962. [Google Scholar] [CrossRef] [Green Version]
  42. Alstermark, B.; Ogawa, J.; Isa, T. Lack of monosynaptic corticomotoneuronal EPSPs in rats: Disynaptic EPSPs mediated via reticulospinal neurons and polysynaptic EPSPs via segmental interneurons. J. Neurophysiol. 2004, 91, 1832–1839. [Google Scholar] [CrossRef]
  43. Nakajima, K.; Maier, M.A.; Kirkwood, P.A.; Lemon, R.N. Striking differences in transmission of corticospinal excitation to upper limb motoneurons in two primate species. J. Neurophysiol. 2000, 84, 698–709. [Google Scholar] [CrossRef] [Green Version]
  44. Pierrot-Deseilligny, E. Propriospinal transmission of part of the corticospinal excitation in humans. Muscle Nerve 2002, 26, 155–172. [Google Scholar] [CrossRef]
  45. Marchand-Pauvert, V.; Peyre, I.; Lackmy-Vallee, A.; Querin, G.; Bede, P.; Lacomblez, L.; Debs, R.; Pradat, P.F. Absence of hyperexcitability of spinal motoneurons in patients with amyotrophic lateral sclerosis. J. Physiol. 2019, 597, 5445–5467. [Google Scholar] [CrossRef] [PubMed]
  46. Martinez-Silva, M.L.; Imhoff-Manuel, R.D.; Sharma, A.; Heckman, C.J.; Shneider, N.A.; Roselli, F.; Zytnicki, D.; Manuel, M. Hypoexcitability precedes denervation in the large fast-contracting motor units in two unrelated mouse models of ALS. eLife 2018, 7. [Google Scholar] [CrossRef]
  47. Delestrée, N.; Manuel, M.; Iglesias, C.; Elbasiouny, S.M.; Heckman, C.J.; Zytnicki, D. Adult spinal motoneurones are not hyperexcitable in a mouse model of inherited amyotrophic lateral sclerosis. J. Physiol. 2014, 592, 1687–1703. [Google Scholar] [CrossRef]
  48. Leroy, F.; Lamotte d’Incamps, B.; Imhoff-Manuel, R.D.; Zytnicki, D. Early intrinsic hyperexcitability does not contribute to motoneuron degeneration in amyotrophic lateral sclerosis. eLife 2014, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Powers, R.K.; Binder, M.D. Input-output functions of mammalian motoneurons. Rev. Physiol. Biochem. Pharmacol. 2001, 143, 137–263. [Google Scholar]
  50. Granit, R.; Kernell, D.; Shortess, G.K. Quantitative aspects of repetitive firing of mammalian motoneurones, caused by injected currents. J. Physiol. 1963, 168, 911–931. [Google Scholar] [CrossRef] [Green Version]
  51. Lee, R.H.; Heckman, C.J. Essential role of a fast persistent inward current in action potential initiation and control of rhythmic firing. J. Neurophysiol. 2001, 85, 472–475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Carter, B.C.; Giessel, A.J.; Sabatini, B.L.; Bean, B.P. Transient sodium current at subthreshold voltages: Activation by EPSP waveforms. Neuron 2012, 75, 1081–1093. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Heckman, C.J.; Johnson, M.; Mottram, C.; Schuster, J. Persistent inward currents in spinal motoneurons and their influence on human motoneuron firing patterns. Neurosci. Rev. J. Bringing Neurobiol. Neurol. Psychiatry 2008, 14, 264–275. [Google Scholar] [CrossRef] [Green Version]
  54. Carunchio, I.; Curcio, L.; Pieri, M.; Pica, F.; Caioli, S.; Viscomi, M.T.; Molinari, M.; Canu, N.; Bernardi, G.; Zona, C. Increased levels of p70S6 phosphorylation in the G93A mouse model of Amyotrophic Lateral Sclerosis and in valine-exposed cortical neurons in culture. Exp. Neurol. 2010, 226, 218–230. [Google Scholar] [CrossRef]
  55. Saba, L.; Viscomi, M.T.; Martini, A.; Caioli, S.; Mercuri, N.B.; Guatteo, E.; Zona, C. Modified age-dependent expression of NaV1.6 in an ALS model correlates with motor cortex excitability alterations. Neurobiol. Dis. 2019, 130, 104532. [Google Scholar] [CrossRef]
  56. Kole, M.H.; Ilschner, S.U.; Kampa, B.M.; Williams, S.R.; Ruben, P.C.; Stuart, G.J. Action potential generation requires a high sodium channel density in the axon initial segment. Nat. Neurosci. 2008, 11, 178–186. [Google Scholar] [CrossRef]
  57. ElBasiouny, S.M.; Schuster, J.E.; Heckman, C.J. Persistent inward currents in spinal motoneurons: Important for normal function but potentially harmful after spinal cord injury and in amyotrophic lateral sclerosis. Clin. Neurophysiol. Off. J. Int. Fed. Clin. Neurophysiol. 2010, 121, 1669–1679. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Le Masson, G.; Przedborski, S.; Abbott, L.F. A computational model of motor neuron degeneration. Neuron 2014, 83, 975–988. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Bellingham, M.C. A review of the neural mechanisms of action and clinical efficiency of riluzole in treating amyotrophic lateral sclerosis: What have we learned in the last decade? CNS Neurosci. Ther. 2011, 17, 4–31. [Google Scholar] [CrossRef] [PubMed]
  60. Bellingham, M.C. Pre- and postsynaptic mechanisms underlying inhibition of hypoglossal motor neuron excitability by riluzole. J. Neurophysiol. 2013, 110, 1047–1061. [Google Scholar] [CrossRef]
  61. Geevasinga, N.; Menon, P.; Ng, K.; Van Den Bos, M.; Byth, K.; Kiernan, M.C.; Vucic, S. Riluzole exerts transient modulating effects on cortical and axonal hyperexcitability in ALS. Amyotroph. Lateral Scler. Front. Degener. 2016, 17, 580–588. [Google Scholar] [CrossRef]
  62. Kanai, K.; Kuwabara, S.; Misawa, S.; Tamura, N.; Ogawara, K.; Nakata, M.; Sawai, S.; Hattori, T.; Bostock, H. Altered axonal excitability properties in amyotrophic lateral sclerosis: Impaired potassium channel function related to disease stage. Brain 2006, 129, 953–962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Vucic, S.; Kiernan, M.C. Axonal excitability properties in amyotrophic lateral sclerosis. Clin. Neurophysiol. Off. J. Int. Fed. Clin. Neurophysiol. 2006, 117, 1458–1466. [Google Scholar] [CrossRef]
  64. Wainger, B.J.; Kiskinis, E.; Mellin, C.; Wiskow, O.; Han, S.S.; Sandoe, J.; Perez, N.P.; Williams, L.A.; Lee, S.; Boulting, G.; et al. Intrinsic membrane hyperexcitability of amyotrophic lateral sclerosis patient-derived motor neurons. Cell Rep. 2014, 7, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Buskila, Y.; Kekesi, O.; Bellot-Saez, A.; Seah, W.; Berg, T.; Trpceski, M.; Yerbury, J.J.; Ooi, L. Dynamic interplay between H-current and M-current controls motoneuron hyperexcitability in amyotrophic lateral sclerosis. Cell Death Dis. 2019, 10, 310. [Google Scholar] [CrossRef] [PubMed]
  66. Fogarty, M.J.; Mu, E.W.; Noakes, P.G.; Lavidis, N.A.; Bellingham, M.C. Marked changes in dendritic structure and spine density precede significant neuronal death in vulnerable cortical pyramidal neuron populations in the SOD1(G93A) mouse model of amyotrophic lateral sclerosis. Acta Neuropathol. Commun. 2016, 4, 77. [Google Scholar] [CrossRef] [Green Version]
  67. King, A.E.; Woodhouse, A.; Kirkcaldie, M.T.; Vickers, J.C. Excitotoxicity in ALS: Overstimulation, or overreaction? Exp. Neurol. 2016, 275 Pt 1, 162–171. [Google Scholar] [CrossRef]
  68. Lewerenz, J.; Maher, P. Chronic glutamate toxicity in neurodegenerative diseases-what is the evidence? Front. Neurosci. 2015, 9, 469. [Google Scholar] [CrossRef]
  69. Van Den Bosch, L.; Van Damme, P.; Bogaert, E.; Robberecht, W. The role of excitotoxicity in the pathogenesis of amyotrophic lateral sclerosis. Biochim. Biophys. Acta 2006, 1762, 1068–1082. [Google Scholar] [CrossRef] [Green Version]
  70. Rothstein, J.D.; Van Kammen, M.; Levey, A.I.; Martin, L.J.; Kuncl, R.W. Selective loss of glial glutamate transporter GLT-1 in amyotrophic lateral sclerosis. Ann. Neurol. 1995, 38, 73–84. [Google Scholar] [CrossRef]
  71. Bristol, L.A.; Rothstein, J.D. Glutamate transporter gene expression in amyotrophic lateral sclerosis motor cortex. Ann. Neurol. 1996, 39, 676–679. [Google Scholar] [CrossRef]
  72. Lin, C.L.; Bristol, L.A.; Jin, L.; Dykes-Hoberg, M.; Crawford, T.; Clawson, L.; Rothstein, J.D. Aberrant RNA processing in a neurodegenerative disease: The cause for absent EAAT2, a glutamate transporter, in amyotrophic lateral sclerosis. Neuron 1998, 20, 589–602. [Google Scholar] [CrossRef] [Green Version]
  73. Howland, D.S.; Liu, J.; She, Y.; Goad, B.; Maragakis, N.J.; Kim, B.; Erickson, J.; Kulik, J.; DeVito, L.; Psaltis, G.; et al. Focal loss of the glutamate transporter EAAT2 in a transgenic rat model of SOD1 mutant-mediated amyotrophic lateral sclerosis (ALS). Proc. Natl. Acad. Sci. USA 2002, 99, 1604–1609. [Google Scholar] [CrossRef] [Green Version]
  74. Tanaka, K.; Watase, K.; Manabe, T.; Yamada, K.; Watanabe, M.; Takahashi, K.; Iwama, H.; Nishikawa, T.; Ichihara, N.; Kikuchi, T.; et al. Epilepsy and exacerbation of brain injury in mice lacking the glutamate transporter GLT-1. Science 1997, 276, 1699–1702. [Google Scholar] [CrossRef] [PubMed]
  75. Warita, H.; Manabe, Y.; Murakami, T.; Shiote, M.; Shiro, Y.; Hayashi, T.; Nagano, I.; Shoji, M.; Abe, K. Tardive decrease of astrocytic glutamate transporter protein in transgenic mice with ALS-linked mutant SOD1. Neurol. Res. 2002, 24, 577–581. [Google Scholar] [CrossRef] [PubMed]
  76. Sirabella, R.; Valsecchi, V.; Anzilotti, S.; Cuomo, O.; Vinciguerra, A.; Cepparulo, P.; Brancaccio, P.; Guida, N.; Blondeau, N.; Canzoniero, L.M.T.; et al. Ionic Homeostasis Maintenance in ALS: Focus on New Therapeutic Targets. Front. Neurosci. 2018, 12, 510. [Google Scholar] [CrossRef] [PubMed]
  77. LeMasson, G.; Marder, E.; Abbott, L.F. Activity-dependent regulation of conductances in model neurons. Science 1993, 259, 1915–1917. [Google Scholar] [CrossRef] [Green Version]
  78. Geevasinga, N.; Menon, P.; Yiannikas, C.; Kiernan, M.C.; Vucic, S. Diagnostic utility of cortical excitability studies in amyotrophic lateral sclerosis. Eur. J. Neurol. Off. J. Eur. Fed. Neurol. Soc. 2014, 21, 1451–1457. [Google Scholar] [CrossRef]
  79. Di Lazzaro, V.; Ranieri, F.; Profice, P.; Pilato, F.; Mazzone, P.; Capone, F.; Insola, A.; Oliviero, A. Transcranial direct current stimulation effects on the excitability of corticospinal axons of the human cerebral cortex. Brain Stimul. 2013, 6, 641–643. [Google Scholar] [CrossRef]
  80. Vucic, S.; Ziemann, U.; Eisen, A.; Hallett, M.; Kiernan, M.C. Transcranial magnetic stimulation and amyotrophic lateral sclerosis: Pathophysiological insights. J. Neurol. Neurosurg. Psychiatry 2013, 84, 1161–1170. [Google Scholar] [CrossRef]
  81. Menon, P.; Higashihara, M.; van den Bos, M.; Geevasinga, N.; Kiernan, M.C.; Vucic, S. Cortical hyperexcitability evolves with disease progression in ALS. Ann. Clin. Transl. Neurol. 2020, 7, 733–741. [Google Scholar] [CrossRef] [PubMed]
  82. Rusu, C.V.; Murakami, M.; Ziemann, U.; Triesch, J. A model of TMS-induced I-waves in motor cortex. Brain Stimul. 2014, 7, 401–414. [Google Scholar] [CrossRef]
  83. Turner, M.R.; Kiernan, M.C. Does interneuronal dysfunction contribute to neurodegeneration in amyotrophic lateral sclerosis? Amyotroph. Lateral Scler. Off. Publ. World Fed. Neurol. Res. Group Mot. Neuron Dis. 2012, 13, 245–250. [Google Scholar] [CrossRef] [PubMed]
  84. Ziemann, U.; Reis, J.; Schwenkreis, P.; Rosanova, M.; Strafella, A.; Badawy, R.; Müller-Dahlhaus, F. TMS and drugs revisited 2014. Clin. Neurophysiol. Off. J. Int. Fed. Clin. Neurophysiol. 2015, 126, 1847–1868. [Google Scholar] [CrossRef]
  85. Verstraete, E.; Veldink, J.H.; Hendrikse, J.; Schelhaas, H.J.; van den Heuvel, M.P.; van den Berg, L.H. Structural MRI reveals cortical thinning in amyotrophic lateral sclerosis. J. Neurol. Neurosurg. Psychiatry 2012, 83, 383–388. [Google Scholar] [CrossRef] [PubMed]
  86. Walhout, R.; Westeneng, H.J.; Verstraete, E.; Hendrikse, J.; Veldink, J.H.; van den Heuvel, M.P.; van den Berg, L.H. Cortical thickness in ALS: Towards a marker for upper motor neuron involvement. J. Neurol. Neurosurg. Psychiatry 2015, 86, 288–294. [Google Scholar] [CrossRef]
  87. Blain, C.R.; Brunton, S.; Williams, V.C.; Leemans, A.; Turner, M.R.; Andersen, P.M.; Catani, M.; Stanton, B.R.; Ganesalingham, J.; Jones, D.K.; et al. Differential corticospinal tract degeneration in homozygous ‘D90A’ SOD-1 ALS and sporadic ALS. J. Neurol. Neurosurg. Psychiatry 2011, 82, 843–849. [Google Scholar] [CrossRef] [PubMed]
  88. Lillo, P.; Mioshi, E.; Burrell, J.R.; Kiernan, M.C.; Hodges, J.R.; Hornberger, M. Grey and white matter changes across the amyotrophic lateral sclerosis-frontotemporal dementia continuum. PLoS ONE 2012, 7, e43993. [Google Scholar] [CrossRef]
  89. Menke, R.A.; Abraham, I.; Thiel, C.S.; Filippini, N.; Knight, S.; Talbot, K.; Turner, M.R. Fractional anisotropy in the posterior limb of the internal capsule and prognosis in amyotrophic lateral sclerosis. Arch. Neurol. 2012, 69, 1493–1499. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Senda, J.; Kato, S.; Kaga, T.; Ito, M.; Atsuta, N.; Nakamura, T.; Watanabe, H.; Tanaka, F.; Naganawa, S.; Sobue, G. Progressive and widespread brain damage in ALS: MRI voxel-based morphometry and diffusion tensor imaging study. Amyotroph. Lateral Scler. Off. Publ. World Fed. Neurol. Res. Group Mot. Neuron Dis. 2011, 12, 59–69. [Google Scholar] [CrossRef]
  91. Vucic, S.; Lin, C.S.; Cheah, B.C.; Murray, J.; Menon, P.; Krishnan, A.V.; Kiernan, M.C. Riluzole exerts central and peripheral modulating effects in amyotrophic lateral sclerosis. Brain 2013, 136, 1361–1370. [Google Scholar] [CrossRef] [Green Version]
  92. Oswald, M.J.; Tantirigama, M.L.; Sonntag, I.; Hughes, S.M.; Empson, R.M. Diversity of layer 5 projection neurons in the mouse motor cortex. Front. Cell. Neurosci. 2013, 7, 174. [Google Scholar] [CrossRef] [Green Version]
  93. Gautam, M.; Jara, J.H.; Kocak, N.; Rylaarsdam, L.E.; Kim, K.D.; Bigio, E.H.; Hande Özdinler, P. Mitochondria, ER, and nuclear membrane defects reveal early mechanisms for upper motor neuron vulnerability with respect to TDP-43 pathology. Acta Neuropathol. 2019, 137, 47–69. [Google Scholar] [CrossRef] [PubMed]
  94. Gautam, M.; Jara, J.H.; Sekerkova, G.; Yasvoina, M.V.; Martina, M.; Ozdinler, P.H. Absence of alsin function leads to corticospinal motor neuron vulnerability via novel disease mechanisms. Hum. Mol. Genet. 2016, 25, 1074–1087. [Google Scholar] [CrossRef] [Green Version]
  95. Jara, J.H.; Genç, B.; Cox, G.A.; Bohn, M.C.; Roos, R.P.; Macklis, J.D.; Ulupınar, E.; Özdinler, P.H. Corticospinal Motor Neurons Are Susceptible to Increased ER Stress and Display Profound Degeneration in the Absence of UCHL1 Function. Cereb. Cortex 2015, 25, 4259–4272. [Google Scholar] [CrossRef] [Green Version]
  96. Genç, B.; Jara, J.H.; Lagrimas, A.K.; Pytel, P.; Roos, R.P.; Mesulam, M.M.; Geula, C.; Bigio, E.H.; Özdinler, P.H. Apical dendrite degeneration, a novel cellular pathology for Betz cells in ALS. Sci. Rep. 2017, 7, 41765. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Wagle-Shukla, A.; Ni, Z.; Gunraj, C.A.; Bahl, N.; Chen, R. Effects of short interval intracortical inhibition and intracortical facilitation on short interval intracortical facilitation in human primary motor cortex. J. Physiol. 2009, 587, 5665–5678. [Google Scholar] [CrossRef] [PubMed]
  98. Foerster, B.R.; Callaghan, B.C.; Petrou, M.; Edden, R.A.; Chenevert, T.L.; Feldman, E.L. Decreased motor cortex γ-aminobutyric acid in amyotrophic lateral sclerosis. Neurology 2012, 78, 1596–1600. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Petri, S.; Krampfl, K.; Hashemi, F.; Grothe, C.; Hori, A.; Dengler, R.; Bufler, J. Distribution of GABAA receptor mRNA in the motor cortex of ALS patients. J. Neuropathol. Exp. Neurol. 2003, 62, 1041–1051. [Google Scholar] [CrossRef] [Green Version]
  100. Lloyd, C.M.; Richardson, M.P.; Brooks, D.J.; Al-Chalabi, A.; Leigh, P.N. Extramotor involvement in ALS: PET studies with the GABA(A) ligand [(11)C]flumazenil. Brain 2000, 123 Pt 11, 2289–2296. [Google Scholar] [CrossRef] [Green Version]
  101. Nihei, K.; McKee, A.C.; Kowall, N.W. Patterns of neuronal degeneration in the motor cortex of amyotrophic lateral sclerosis patients. Acta Neuropathol. 1993, 86, 55–64. [Google Scholar] [CrossRef] [PubMed]
  102. Khademullah, C.S.; Aqrabawi, A.J.; Place, K.M.; Dargaei, Z.; Liang, X.; Pressey, J.C.; Bedard, S.; Yang, J.W.; Garand, D.; Keramidis, I.; et al. Cortical interneuron-mediated inhibition delays the onset of amyotrophic lateral sclerosis. Brain J. Neurol. 2020, 143, 800–810. [Google Scholar] [CrossRef] [PubMed]
  103. Nieto-Gonzalez, J.L.; Moser, J.; Lauritzen, M.; Schmitt-John, T.; Jensen, K. Reduced GABAergic inhibition explains cortical hyperexcitability in the wobbler mouse model of ALS. Cereb. Cortex 2011, 21, 625–635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Clark, R.M.; Blizzard, C.A.; Young, K.M.; King, A.E.; Dickson, T.C. Calretinin and Neuropeptide Y interneurons are differentially altered in the motor cortex of the SOD1(G93A) mouse model of ALS. Sci. Rep. 2017, 7, 44461. [Google Scholar] [CrossRef] [Green Version]
  105. Kim, J.; Hughes, E.G.; Shetty, A.S.; Arlotta, P.; Goff, L.A.; Bergles, D.E.; Brown, S.P. Changes in the Excitability of Neocortical Neurons in a Mouse Model of Amyotrophic Lateral Sclerosis Are Not Specific to Corticospinal Neurons and Are Modulated by Advancing Disease. J. Neurosci. Off. J. Soc. Neurosci. 2017, 37, 9037–9053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Zhang, W.; Zhang, L.; Liang, B.; Schroeder, D.; Zhang, Z.W.; Cox, G.A.; Li, Y.; Lin, D.T. Hyperactive somatostatin interneurons contribute to excitotoxicity in neurodegenerative disorders. Nat. Neurosci. 2016, 19, 557–559. [Google Scholar] [CrossRef] [Green Version]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pradhan, J.; Bellingham, M.C. Neurophysiological Mechanisms Underlying Cortical Hyper-Excitability in Amyotrophic Lateral Sclerosis: A Review. Brain Sci. 2021, 11, 549. https://doi.org/10.3390/brainsci11050549

AMA Style

Pradhan J, Bellingham MC. Neurophysiological Mechanisms Underlying Cortical Hyper-Excitability in Amyotrophic Lateral Sclerosis: A Review. Brain Sciences. 2021; 11(5):549. https://doi.org/10.3390/brainsci11050549

Chicago/Turabian Style

Pradhan, Jonu, and Mark C. Bellingham. 2021. "Neurophysiological Mechanisms Underlying Cortical Hyper-Excitability in Amyotrophic Lateral Sclerosis: A Review" Brain Sciences 11, no. 5: 549. https://doi.org/10.3390/brainsci11050549

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