Abstract
Amyotrophic lateral sclerosis is a devastating neurodegenerative disease that, at present, has no effective cure. Evidence of increased circulating glutamate and hyperexcitability of the motor cortex in patients with amyotrophic lateral sclerosis have provided an empirical support base for the ‘dying forward’ excitotoxicity hypothesis. The hypothesis postulates that increased activation of upper motor neurons spreads pathology to lower motor neurons in the spinal cord in the form of excessive glutamate release, which triggers excitotoxic processes. Many clinical trials have focused on therapies that target excitotoxicity via dampening neuronal activation, but not all are effective. As such, there is a growing tension between the rising tide of evidence for the ‘dying forward’ excitotoxicity hypothesis and the failure of therapies that target neuronal activation.
One possible solution to these contradictory outcomes is that our interpretation of the current evidence requires revision in the context of appreciating the complexity of the nervous system and the limitations of the neurobiological assays we use to study it. In this review we provide an evaluation of evidence relevant to the ‘dying forward’ excitotoxicity hypothesis and by doing so, identify key gaps in our knowledge that need to be addressed. We hope to provide a road map from hyperexcitability to excitotoxicity so that we can better develop therapies for patients suffering from amyotrophic lateral sclerosis.
We conclude that studies of upper motor neuron activity and their synaptic output will play a decisive role in the future of amyotrophic lateral sclerosis therapy.
Keywords: amyotrophic lateral sclerosis, hyperexcitability, excitotoxicity, dying forward, homeostasis
Odierna et al. evaluate preclinical and clinical evidence for brain-initiated toxicity in amyotrophic lateral sclerosis, and identify key roadblocks and gaps in understanding that must be tackled in order to develop effective therapies.
Introduction
It has become increasingly evident that a key pathophysiological event in the progression of amyotrophic lateral sclerosis is the induction of hyperexcitability of the motor cortex.1 Studies of patients with amyotrophic lateral sclerosis have converged upon the conclusion that early motor cortex hyperexcitability, as defined by measures of cortical network function (such as reduced short intracortical inhibition and increased motor evoked potential amplitude), is a pathogenic feature.2-4 Such findings have significantly contributed to our understanding of this rapidly progressing neurodegenerative disease. They have also strengthened support for the ‘dying forward’ hypothesis, which traditionally describes pathological spread as a feed-forward process that relies on glutamate excitotoxicity (see Shaw and Ince5 and King et al.6 for more detailed reviews). Much attention has been turned towards the development of therapies that aim to suppress activation of neurons by glutamate as a treatment for amyotrophic lateral sclerosis. However, apart from riluzole, most of these treatments have failed to extend patient lifespans or effectively treat their symptoms.7-12 Despite frequent glimmers of hope, the amyotrophic lateral sclerosis research field has a history of therapeutic trails targeting excitotoxicity that fail to reach efficacy, beginning with lamotrigine and dextromethorphan close to 30 years ago.13-15 There are many possible explanations for why the efficacy of new treatments that target glutamate-mediated hyperexcitability are not consistently effective, such as the long delay between presentation and diagnosis (typically 10–16 months from symptom onset),16 a lack of effective biomarkers,17 or the overwhelming heterogeneity of the disease.18,19 One conclusion that needs to be considered is that our current hypotheses about the mechanisms that take a hyperexcitable motor cortex to excitotoxicity and neuronal degeneration in amyotrophic lateral sclerosis need to be re-examined. An incomplete understanding of disease mechanisms can result in targeting incorrect pathways, or inadvertently suppressing homeostatic processes that have been induced to counteract a neuronal dysfunction. A recent example of this came from a randomized phase 2 clinical trial for perampanel, an AMPA receptor antagonist used to treat epilepsy, as a therapy for patients with sporadic amyotrophic lateral sclerosis. Daily administration of perampanel worsened patients’ functional rating scores and increased the frequency of adverse events.20 Collectively, these studies provide the impetus needed to consider broadening interpretations of the role hyperexcitability plays in the disease.
The aim of the present review is to reassess recent preclinical research to identify exactly what is missing with respect to the ‘dying forward’ excitotoxicity hypothesis so that we can better position ourselves to develop effective therapies. We aim to evaluate interpretations of clinical and preclinical pathophysiological amyotrophic lateral sclerosis studies with a specific focus on excitability, activity and neurotransmission.
With this in mind, it is important to define some core terminology. ‘Excitability’ broadly refers to input-output relationships and can be used to describe entire networks or individual neurons, depending on the assay being used. In the context of networks, it refers to the propensity to produce a measurable functional output of the network in question in response to spatially broad stimulation (e.g. the motor evoked potential following stimulation of the motor cortex). In the context of neurons, it refers to the propensity to produce action potentials in response to depolarizing currents. ‘Activity’ is a separate term from ‘excitability’ and instead refers to the aggregate output of a network or a neuron over time. This can refer to resting levels of activity that occur spontaneously or the activity profile that occurs during recruitment of networks or neurons in their relevant context (e.g. the pattern of lower motor neuron activation during performance of a motor task). It is important to recognize that excitability alone is not predictive of activity and that for the ‘dying forward’ hypothesis, it is activity that is ultimately of interest.
Defining ‘dying forward’ spread of pathology and key evidence for the excitability hypothesis
Although ‘excitotoxicity’ and ‘dying forward’ are often conflated in amyotrophic lateral sclerosis research, they are technically not mutually inclusive ideas. Excitotoxicity classically refers to abnormal acute or chronic activation of intracellular calcium signalling cascades via glutamate receptor-dependent pathways.21 It is a process that is strongly associated with demise of neurons via either apoptotic or necroptotic cell death.22,23 The dying forward hypothesis highlights the primacy of the motor cortex in the initiation and propagation of degenerative pathways that converge onto lower motor neurons, leading to neurodegeneration and the amyotrophic lateral sclerosis phenotype.24 The most widely recognized form of the ‘dying forward’ hypothesis is one that integrates the idea of excitotoxicity as the physiological instrument that propagates pathology from projection neurons in the motor cortex to lower motor neurons in the spinal cord. Empirical support for this hypothesis started with the occurrence of increased CSF glutamate levels in patients with amyotrophic lateral sclerosis and was later supported by observations of motor cortex hyperexcitability. These results provide insightful evidence that the nervous system in amyotrophic lateral sclerosis patients functions differently at a very fundamental level and as such, have shaped the way we think about pathological mechanisms.
The earliest evidence for excitotoxicity in amyotrophic lateral sclerosis came from the striking similarities between the cellular pathology induced by excitotoxins and the observed pathological features in post-mortem tissue of patients.25 Of particular interest were axonal, dendritic and somatic swellings seen throughout the spinal cord of amyotrophic lateral sclerosis patients,25 which were highly reminiscent of those seen in rodents treated with glutamate or kainic acid.26,27 Studies of amino acid levels in plasma and CSF supported this connection by finding that patients with amyotrophic lateral sclerosis had high concentrations of circulating glutamate.28-32 The most common interpretation of these findings is that the increased circulating glutamate is a by-product of underlying neuronal hyperactivity, and that excess glutamate likely drives the excitotoxic processes. Critical evaluation of raised glutamate levels in patients, however, indicates that our understanding of the role of glutamate in amyotrophic lateral sclerosis remains incomplete. The initial findings of raised glutamate levels were not found in all patient subgroups; some studies claimed that patients had high plasma or CSF glutamate concentrations,28-32 whereas others found that glutamate levels were comparable to controls.33,34 Later studies provided some ad hoc clarification by identifying that high circulating glutamate in patients with amyotrophic lateral sclerosis is specifically associated with spinal onset.35-37 Importantly, many of the original studies reported a striking depletion of glutamate levels in brain and spinal cord tissue of patients31,38-42 but this was never meaningfully incorporated into later hypotheses of the disease mechanisms that focused on excitotoxicity. The reports of glutamate depletion, specifically in the ventral horn of the spinal cord, indicate the existence of a yet unrecognized degree of complexity of amyotrophic lateral sclerosis that challenges hypotheses that rely on excitotoxicity.
The ‘dying forward’ hypothesis is supported by early hyperexcitability that is observed in cortex of patients with amyotrophic lateral sclerosis.3,4,43-45 Cortical network excitability of patients has been explored in detail by clinical studies that have leveraged transcranial magnetic stimulation (TMS) technology alongside the accessibility of measuring corticomotor system output in the periphery. TMS uses brief pulses of electricity through a coil to generate a current that can penetrate the intact scalp and stimulate the motor cortex and all the microcircuitry within.46 By doing so, TMS elicits a response at muscles, which is measurable via EMG. Comparing the minimum required current to elicit an EMG response can produce an approximation of how excitable the motor cortex is. Clinical TMS studies indicate that the motor cortex in amyotrophic lateral sclerosis patients is hyperexcitable.2-4 This change is detectable before onset of motor symptoms in individuals with familial amyotrophic lateral sclerosis44 and before manifestation of measurable changes in peripheral nerve evoked responses.2 Cortical hyperexcitability is one of the most reliable biosignatures of amyotrophic lateral sclerosis and can differentiate the disease from its mimics.47-50 However, how cortical excitability and neurodegeneration in amyotrophic lateral sclerosis mechanistically relate to each other remains to be fully elucidated.
The difficulty in forming causal links between glutamate, cortical hyperexcitability and neurodegeneration in amyotrophic lateral sclerosis arises from the complexity of the corticomotor system and the limitations of the tools we use to study it. Clinical studies provide directly relevant and diagnostically rich information but must be tempered by patient care. Hence, animal models play a vital role in advancing our understanding of pathogenesis in amyotrophic lateral sclerosis. They permit for targeted interrogation of genetic, biochemical and physiological processes and thus provide an irreplaceable platform to test the validity of hypotheses (such as the dying forward hypothesis) at high mechanistic resolution. Despite this, different neurobiological assays are not definitive and only provide limited windows through which networks or brain cell function can be assessed. With all this in mind, we will descend in scales of complexity from networks to synapses (Fig. 1) to identify the strengths and limitations of the evidence each level can provide and hopefully build a road map to get from hyperexcitability to excitotoxicity.
Figure 1.
The complexity of the corticomotor system and the limitations of the tools we use to study it. Amyotrophic lateral sclerosis has been studied at three levels across the nervous system: networks, neurons and synapses. Network: Our understanding of network physiology has been predominantly derived from research implementing transcranial magnetic stimulation (TMS) to patients with amyotrophic lateral sclerosis. TMS can induce activation of brain regions to measure the excitability of networks (glowing brain region under the coil). Clinical TMS studies have revealed that the cortex of patients is often more excitable, such that peripheral motor responses can be more easily induced for any intensity of stimulus provided to the motor region. TMS is limited by cellular resolution and only comments on excitability. As such, the endogenous activity of motor networks, and the cells within them, in patients remains unknown. Neuron: Our understanding of neuronal physiology in amyotrophic lateral sclerosis has been predominantly derived from research implementing whole-cell patch-clamp of neurons in the corticomotor system of preclinical rodent models. Patch-clamp can measure currents that underlie neuronal function, but this is limited to the region of neurons that can be confidently controlled by the experimenter (blue region of the neuron depicts area of a neuron that is likely clamped by the pipette and accurately recorded from). Patch-clamp studies have revealed that layer 5 projection neurons of the motor cortex are often intrinsically hyperexcitable, such that their likelihood of firing action potentials is increased for any intensity of injected current. Patch-clamp is limited by its poor synaptic resolution, can only measure single neurons in isolation and only comments on excitability. As such, the endogenous activity of upper motor neurons remains unknown. Synapse: Our understanding of synaptic physiology in amyotrophic lateral sclerosis has been predominantly derived from research implementing intracellular recordings of muscle end plate potentials. End plate potential recordings measure properties of synaptic neurotransmission but are biased towards lower motor neurons and neuromuscular physiology. End plate potential recordings across multiple different model organism species have revealed that evoked neurotransmitter release is often decreased in amyotrophic lateral sclerosis. There currently are no assays to measure neurotransmission properties of central neurons of the corticomotor system. As such, basal neurotransmitter release from axons of upper motor neurons remains unknown. Network-, neuron- and synapse-level changes must always be considered in the context of homeostatic adaptions, which occur to stabilize outputs based on global and local set points.
Networks
Bridging the gap from excitability to activity: cortical network calcium imaging in mice
Studies using preclinical animal models of amyotrophic lateral sclerosis allow for interrogation of network and/or neuron physiology at high spatiotemporal resolution and broadly across brain regions. Recent studies investigating neuronal activity using in vivo calcium imaging in small animal models of amyotrophic lateral sclerosis have begun to bridge the gap between observations of network hyperexcitability and what this means for network activity. The importance of exploring network activity cannot be understated because it holds the potential to inform the development of meaningful activity-modulating therapies for patients. One study on mice harbouring expansion repeats of C9orf72 (Table 1) applied wireless photometry to indirectly measure the compound activity of corticospinal projection neurons via GCaMP6f signals from their axonal projections in white matter tracts. Despite finding a marked loss of neurons in the motor cortex, the authors reported that the output from those that remained was enhanced during high-speed locomotion, implying increased endogenous task-specific output.64 This kind of increased neuronal activity has also been observed in a study of a nuclear localization sequence-deficient fused in sarcoma mouse model (FUSΔNLS, Table 1), which assessed spontaneous GCaMP6s signals of layer 2/3 motor cortex neurons in anaesthetized mice via cranial window.65 Although neither of these studies measured whole networks or directly assessed corticospinal projection neurons, they both provide tantalizing hints that network activity might skew towards hyperactivity in amyotrophic lateral sclerosis. Contrasting these studies, assessment of neuronal activity of layer 2/3 motor cortex neurons of mice that overexpress human superoxide dismutase 1 harbouring the G93A mutation (SOD1G93A, Table 1) found that neuronal activity is indistinguishable from controls and the authors concluded that network activity in amyotrophic lateral sclerosis may be stabilized by strong homeostatic adaptations.66 Given that this work was performed at a late symptomatic stage, it raises the possibility that there is a complex progression of destabilization and stabilization of network activity over the time course of the disease. In support of this idea, a study that applied in vivo calcium imaging to awake mice found that conditional knock out of transactive response DNA-binding protein 43 kDa (TDP-43, Table 1) resulted in hyperactivity of prefrontal cortex neurons, followed by silencing and then cellular death.67 Although this assessment was not performed in the motor cortex, it demonstrates that the network dysfunction can manifest as increased or decreased activity depending on the stage of disease.
Table 1.
Genes and mutations used to model amyotrophic lateral sclerosis in mice discussed in this review article
Gene | Function | Insult used | Description |
---|---|---|---|
Superoxide dismutase 1 (SOD1) | Free radical scavenging | SOD1G93A | Mutations in SOD1 are strongly associated with development of familial amyotrophic lateral sclerosis.51 The SOD1G93A mouse overexpresses human SOD1 harbouring the disease-linked G93A amino acid substitution mutation. Expression is driven off the human SOD1 promoter.52 |
Transactive response DNA binding protein 43 kDA (TDP-43) | DNA and RNA metabolism | TDP-43A315T | TDP-43 is a major component of ubiquitinated cytoplasmic inclusions found ubiquitously in amyotrophic lateral sclerosis patients.53 Mutations in TDP-43 are also associated with development of amyotrophic lateral sclerosis.54 The TDP-43A315T mouse overexpresses human TDP-43 harbouring the disease-linked A315T amino acid substitution mutation. Expression is driven off the murine prion protein (PrP) promoter.55 |
TDP-43Q331K | The TDP-43Q331K mouse overexpresses human TDP-43 harbouring the disease-linked Q331K amino acid substitution mutation. Expression is driven off the murine PrP promoter.56 | ||
TDP-43ΔNLS | The TDP-43ΔNLS mouse model is based on the doxycycline-inducible gene expression system. It uses a promoter-specific tetracycline transactivator to direct cell type-specific overexpression of human TDP-43 with an ablated nuclear localization sequence (NLS).57 | ||
Fused in sarcoma (FUS) | DNA and RNA metabolism | FUSΔNLS | Mutations in FUS are associated with familial amyotrophic lateral sclerosis patients.58 Disease-linked mutant FUS proteins preferentially associate with cytoplasmic stress granules.59 The FUSΔNLS mouse features engineering of the endogenous mouse Fus gene to incorporate a premature STOP codon, resulting in a C-terminally truncated NLS-deficient protein that accumulates in the cytoplasm.60 |
Chromosome 9 open reading frame 72 (C9orf72) | Autophagy and vesicle trafficking | C9-500 | Intronic GGGGCC (G4C2) repeats are typically found within the C9orf72 gene between exon 1a and exon 1b.61 In healthy controls, the repeats occur <30 times62 but in some amyotrophic lateral sclerosis patients, the sequence repeats can be expanded up to hundreds of times.61 The C9-500 mouse harbours ∼500 G4C2 expansion repeats.63 |
The importance of these findings lies in the distinction between excitability and activity. Despite the many studies that have demonstrated motor cortical network hyperexcitability in amyotrophic lateral sclerosis, the activity status of networks along the corticomotor system remains a mystery. This is a significant obstacle to our development of effective therapies; if the network is not hyperactive then excitability-targeting treatments are unlikely to provide therapeutic benefits to patients. As such, additional work is needed to more completely assess how activity of the corticomotor network is affected over time in amyotrophic lateral sclerosis. In moving forward, it is important to recognize that all studies of networks in vivo suffer from the limitation that they are multi-faceted measures (Fig. 1). They are subject to many factors that complicate deciphering mechanistic information including (but likely not limited to) neuronal population composition, network connectomes, neuronal activation properties, cell intrinsic electrophysiological factors, synaptic inputs, synaptic strength, neuromodulation, brain states, glial biology and ephaptic factors. This makes it difficult to understand how the altered network excitability or activity manifests and, consequently, what kinds of therapeutics need to be designed to treat patients’ needs. As such, each part of the corticomotor system needs to be considered at the level of individual neurons so that we can understand how the insults that drive pathology of amyotrophic lateral sclerosis affects their function.
Neurons
Patch-clamp and interrogation of intrinsic upper and lower motor neuron excitability
A core change that must manifest to satisfy the ‘dying forward’ hypothesis is increased output of corticospinal projection neurons (hereon referred to as ‘upper motor neurons’). These neurons represent the main conduit through which information processed in the cortex is relayed to the spinal cord and as such, their role in propagating pathophysiology is fundamental. Neuronal excitbility is the product of intrinsic and extrinsic factors. Intrinsic properties of corticospinal projection neurons have been investigated extensively in amyotrophic lateral sclerosis mouse models using whole-cell patch-clamp, which involves invasively creating a tight seal between a pipette and a neuron’s membrane. This connection is used to inject current to gain insight into the electrophysiological properties of neurons, most notably intrinsic excitability. This method enables the excitability of neurons to be resolved at the expense of being able to measure activity over time (days to weeks) or circuit activity (Fig. 1). The intrinsic excitability of a neuron is determined by the types, density, distribution and kinetics of voltage-gated ion channels in the membrane. In brain slices of mice overexpressing the familial disease associated SOD1G93A (Table 1), intrinsic excitability of layer 5 pyramidal neurons of the motor cortex (of which upper motor neurons are a subset) has been shown to be either increased or decreased, implying a complex progression of excitability phenotypes. Intrinsic hyperexcitability of layer 5 pyramidal neurons of the motor cortex in SOD1G93A mice occurs early in postnatal developmental stages.66,68 Early intrinsic hyperexcitability of these neurons has also been observed in TDP-43 mouse models of amyotrophic lateral sclerosis, including overexpression of the disease-linked mutant TDP-43A315T and TDP-43ΔNLS (Table 1), which accumulates human TDP-43 in the cytoplasm.69,70 The observation of a common layer 5 pyramidal neuron hyperexcitability phenotype in rodent models suggests that increased neuronal excitability may represent a conserved pathological feature of amyotrophic lateral sclerosis. In support of this, human induced pluripotent stem cell-derived neurons also display hyperexcitability in vitro.71 Although these neurons do not have a strict upper motor neuron identity, they provide confidence that the hyperexcitability observed in rodents is meaningful to human disease.
The mechanism of increased upper motor neuron neuronal excitability has been attributed to a range of ion channels and ionic currents. Increased persistent sodium channel currents, in particular, have been observed in layer 5 pyramidal neurons of the motor cortex in SOD1G93A mice and this was associated with increased contribution of Nav1.6 sodium channels.68 Increased intrinsic excitability of layer 5 pyramidal neurons has separately been linked to an altered balance between hyperpolarization-activated cyclic nucleotide gated (HCN)- and Kv7 potassium channel (also known as KCNQ)-mediated currents in ex vivo brain slices of SOD1G93A mice.72 Different sets of voltage-gated potassium, sodium and calcium channels have been reported to be up- and downregulated in corticocortical and corticospinal layer 5 pyramidal neurons.66 This is perhaps not surprising given that the expression of ion channels is variably co-regulated in different subtypes of neurons to establish and maintain functional outputs.73,74 It is important to note that multiple adaptational configurations are possible to maintain functional output at the neuron, network and behavioural levels. For instance, network level output can be maintained despite variations in the roles of neurons within the network, and behavioural output can be maintained using multiple network configurations.74
The excitability of lower motor neurons has been assessed in detail in rodent models of amyotrophic lateral sclerosis. The earliest investigation into lower motor neuron excitability was performed on SOD1G93A E17.5 embryos.75 The study found that embryonic SOD1G93A lower motor neurons were hyperexcitable, as measured by higher input resistance, lower rheobase and reduced dendritic arbour. Much more work has been performed on neonatal and presymptomatic mice, many of which report seemingly contradictory findings. Some studies claim that lower motor neurons display signs of hypoexcitability,76-78 whereas others show that lower motor neurons are instead hyperexcitable.79-82 Although it is possible that these discrepancies arise due to the use of different backgrounds or models (such as the low copy number SOD1G93A or the SOD1G85R versus the high copy number SOD1G93A), a more intriguing possibility is that the differing reports arise due to a combination of two factors: (i) that the effects of mutant transgenes on lower motor neuron excitability is subtype-dependent; and (ii) that the excitability profile of lower motor neurons may change over time as the disease progresses. In support of these points, most studies that segregate lower motor neurons into subtypes based on electrophysiological or molecular markers of motor units find subtype-specific responses. Martinez-Silva et al.77 found that fast fatigable lower motor neurons, which are the most vulnerable to degeneration in amyotrophic lateral sclerosis, were more likely to display signs of hypoexcitability in SOD1G93A mice, whereas slow lower motor neurons appeared unaffected. Conversely, Leroy et al.82 reported that motor neurons with a delayed response to current injection were not affected, whereas those with an immediate response became hyperexcitable, a change that was attributable in part to lower rheobase and a reduction in dendritic arbour. Venugopal et al.83 investigated ocular motor neurons, which are spared in amyotrophic lateral sclerosis, and compared their responses to different types of trigeminal motor neurons. Their results suggested that fast fatigable trigeminal motor neurons exhibit hyperexcitability, slow fatigable trigeminal motor neurons exhibit hypoexcitability, and ocular motor neurons are unaffected in SOD1G93A mice. As for timelines of change, Quinlan et al.84 performed a longitudinal study on the evolution of excitability changes of lower motor neurons in SOD1G93A mice from P0 to 12. The authors found that although the transgenic mice did have hyperexcitable lower motor neurons, this could be attributed to an acceleration of developmental processes that drive increased excitability of these neurons.
Together, these data indicate that lower motor neurons undergo complex changes in excitability in amyotrophic lateral sclerosis. Whether these changes exacerbate or protect them from excitotoxic damage remains unclear. An interesting question that emerges from these studies is what the effects of restricted expression of amyotrophic lateral sclerosis-related transgenes only in lower motor neurons would be. Are the observed changes in excitability a physiological, adaptive response to descending pathology or do they arise from intrinsic molecular processes within the lower motor neurons following gain- or loss-of-function of proteins such as SOD1 or TDP-43? Further, how would the motor cortex change in response to focal expression of transgenes only in lower motor neurons? More work will be needed to answer these questions.
Presynaptic input onto upper motor neurons and the excitability of cortical interneurons
Many factors influence neuronal inputs including the number of excitatory and inhibitory synapses, the frequency, strength and timing of synaptic currents mediated by the activity of presynaptic neurons. Electrophysiological measurements of excitatory and inhibitory post-synaptic currents (EPSCs and IPSCs, respectively) have been explored in corticospinal projection neurons using patch-clamp techniques and have been used to draw conclusions about synaptic inputs. Postsynaptic currents display marked changes that follow a pattern along the progression of disease. Increased EPSC frequency has been observed in layer 5 pyramidal neurons of the motor cortex before symptom onset in SOD1G93A and TDP-43Q331K (Table 1) mice,85-87 while it is decreased upon symptom onset in TDP-43A315T and TDP-43ΔNLS mice.70,88 Conversely, IPSC frequency is decreased before symptom onset in SOD1G93A, TDP-43A315T and VPS54 loss-of-function mice.69,89,90 Comprehensive timeline studies of EPSC and IPSC frequency in layer 5 pyramidal neurons across disease progression in the same mouse model are currently lacking. Ignoring neurotransmission-specific factors, these results can imply that upstream input onto upper motor neurons changes early and skews towards an excitatory imbalance, mirroring findings in patients with amyotrophic lateral sclerosis.44,91 It must be noted that EPSC and IPSC measurements via patch-clamp suffer limitations with respect to how they can be interpreted. One issue relates to the physical limitations of clamping neuronal membranes and detecting currents. The amount of control over which experimenters can reliably clamp a membrane, and confidently measure currents, drops as a function of distance from the pipette seal site (known as space clamping issues, Fig. 1). Further, due to separate physical limitations imposed by cable theory, currents that occur at distal sites of axons and dendrites are not resolvable from the soma, which is the most common site where patch-clamp is performed. Another issue with interpreting the frequency of postsynaptic currents is that they are technically the product of two discrete biological processes; notably the physical number of connections formed by presynaptic neurons (neuronal wiring) and the rate of spontaneous neurotransmitter release from presynaptic active zones (synaptic). It is impossible to distinguish which of these two drive changes in the frequency of postsynaptic currents measured from a neuron using whole-cell patch-clamp without supplementary evidence, such as morphological assessment of dendritic spines. The overall result of these limitations is that measurements of EPSCs and IPSCs using patch-clamp likely skew heavily towards axo-somatic connections and cannot readily discriminate between synaptic and wiring defects. One way to overcome these shortcomings is by mapping structural proxies for synapses such as dendritic spines. This approach has been used successfully in rodent models of amyotrophic lateral sclerosis86,92 but it is not reliable for silent synapses or inhibitory connections, which are typically aspiny.93 The use of membrane- or synaptic-localized calcium indicators to study input at the level of individual synapses and all along the dendritic arbour of upper motor neurons could provide unique insight into how extrinsic excitability changes in amyotrophic lateral sclerosis.
Although the synaptic connections received by upper motor neurons is a useful metric of extrinsic contributions to excitability, it cannot be used as a perfect readout because it does not comment on neuronal activity. It is possible, for example, that increases or decreases in the number of synaptic inputs formed by interneurons onto upper motor neurons are matched by opposite and proportional changes in interneuron activity. In such cases, the extrinsic contribution to upper motor neuron excitability would remain largely unchanged. Interrogation of cortical interneuron excitability, their activity and firing patterns, is a core piece of the puzzle with respect to hyperexcitability in amyotrophic lateral sclerosis. Evidence from TMS clinical and electrophysiological preclinical studies of inhibitory interneurons provides evidence that functional inhibitory drive to upper motor neurons may be pathologically altered. Results from paired-pulse TMS experiments suggest that hyperexcitability of the motor cortex in patients with amyotrophic lateral sclerosis can be ascribed to reduced contribution of cortical inhibitory circuits along with excessive activation of high threshold excitatory pathways.91,94 This difference in functional inhibitory-excitatory weighting across the cortex is highly predictive of amyotrophic lateral sclerosis, such that it can differentiate the disease from its mimics47-50 and the added presence of increased resting motor threshold can even differentiate amyotrophic lateral sclerosis patients with cognitive impairment from those without.95 Evidence from animal models has begun to reveal which specific populations of interneurons are affected in amyotrophic lateral sclerosis. Parvalbumin-positive interneurons, which provide strong direct inhibition to layer 5 pyramidal neurons of the motor cortex, have been found to display increased or decreased excitability in SOD1G93A and TDP-43A315T mice,66,69,89 as have GABAergic SOD1G93A cortical neurons in vitro.96 Somatostatin-positive interneurons, which instead play a more upstream modulatory role via disinhibition, also undergo changes in excitability and notably display a hyperexcitable phenotype in TDP-43A315T mice.69 The contribution of these neurons to pathology in amyotrophic lateral sclerosis is highlighted by the finding that ablating them reverses the hyperexcitability phenotype of layer 5 pyramidal neurons.89 Cortical excitatory interneurons also display changes that may serve to increase the excitability of upper motor neurons. In particular, corticocortical neurons in the motor cortex of SOD1G93A mice display increased intrinsic excitability.66 The relevance of interneurons to amyotrophic lateral sclerosis is highlighted by genome-wide association studies, as well as those employing single cell/nucleus RNA sequencing, which often report that corticomotor system interneurons undergo significant (and usually cell type-specific) transcriptional changes in disease.97-99 Although it is evident that upstream changes to upper motor neurons likely play an important role in the pathology of amyotrophic lateral sclerosis, more information is needed to dissect out if these are causal in upper motor neuron pathophysiology. Also, similarly to upper motor neurons, the question of how their activity changes in amyotrophic lateral sclerosis remains unanswered.
Pre-motor interneurons and spinal cord networks
Although upper motor neurons can form direct connections with lower motor neurons, and thus directly drive glutamate-dependent synaptic transmission, their influence is ultimately dictated by networks of spinal cord interneurons. As such, sources of excitotoxic pressure onto lower motor neurons must always consider the effects of spinal cord network activity and, specifically, the contributions of premotor interneurons. Most studies examining spinal interneurons in patients or preclinical animal models report loss of interneurons as the disease progresses. In some cases, the loss of interneurons precedes loss of lower motor neurons100,101 whereas in others, the loss (or magnitude of loss) occurs afterwards102,103 or appears to be concomitant.102,104,105 This frank loss of interneurons is likely to result in broad changes to the activity of spinal cord networks and, ultimately, lower motor neurons. Consistent with this notion, presynaptic inputs onto lower motor neurons are often impacted in disease. In the SOD1G93A mouse model, lower motor neurons experience generalized loss of presynaptic boutons,100,106 indicating that they may become increasingly disconnected from descending and local control. In support of this, many studies have demonstrated that lower motor neurons progressively lose glycinergic connections,100,101,106 and that this occurs earliest in the most vulnerable fast fatigable lower motor neurons.100 Cholinergic C-boutons, which originate from premotor V0c interneurons, are also lost in SOD1G93A and TDP-43ΔNLS mice,107-110 suggesting that affliction of cholinergic input might represent a common pathological outcome of amyotrophic lateral sclerosis. This is consistent with alterations in neurotransmitter release at another cholinergic synapse—the neuromuscular junction.111-114 Lastly, studies investigating the recurrent inhibitory circuit mediated by V1 premotor interneurons, known as Renshaw cells, indicate another possible form of disconnect between interneurons and lower motor neurons. Renshaw cell input onto lower motor neurons has been shown to increase at early stages of disease, specifically onto lower motor neurons undergoing degenerative changes.102,103 Subsequently, collateral intraspinal axon projections from lower motor neurons onto Renshaw cells retract,103 which likely results in a functional decoupling of the recurrent inhibitory circuit. Taken together, these data indicate that spinal cord circuitry undergoes significant remodelling in amyotrophic lateral sclerosis and that the net effect of these changes is a loss of synaptic regulatory control over lower motor neurons.
Assessments of interneuron excitability and activity have been sparse and provide somewhat conflicting reports. Van Zundert et al.115 examined midbrain interneurons in SOD1G93A neonatal mice (P4–10) and found that they displayed increased excitability and increased frequency of EPSCs and IPSCs. However, a more recent study by Cavarsan et al.116 reported that glycinergic interneurons in the lumbar spinal cord are hypoexcitable at P6–P10. The most affected glycinergic neurons were those found in the ventral-most domain of the ventral horn (which is largely populated by Renshaw cells), whereas glycinergic neurons in the intermediate region of laminae 7 and 8 were largely unaffected. Because disparate results are found from different regions of the corticomotor system, it is likely that there are interneuron subtype-specific effects on excitability in response to pathology in amyotrophic lateral sclerosis. As such, exploring the neurophysiological profile of different spinal interneuron populations is needed to produce a more complete picture of how spinal network activity is affected in the disease state. Whether this will support or refute the excitotoxicity ‘’dying forward’ hypothesis remains to be seen.
Synapses
Basal neurotransmitter release in amyotrophic lateral sclerosis
Core to the ‘dying forward’ excitotoxicity hypothesis is the idea that hyperactivity of upper motor neurons leads to chronic release of glutamate in the spinal cord. A critical question in this hypothesis is: does hyperactivity of upper motor neurons sufficiently produce chronic glutamate release at the synapses in the spinal cord? The neurotransmitter release properties of upper motor neurons have not yet been assessed in the context of amyotrophic lateral sclerosis. This is likely due to the technical challenges associated with accurately measuring basal neurotransmission in the CNS. As such, evidence needs to be extrapolated from existing sources to consider how upper motor neuron synapses might be impacted in disease.
Multiple proteins associated with amyotrophic lateral sclerosis have been shown to localize to presynaptic axon termini,117-120 localize to postsynaptic regions118,120,121 or directly affect the transcripts of synaptic genes,119,122-124 strongly implicating the synapse as a core locus of pathophysiology in the disease.125 Assessment of what these changes mean for neurotransmission is challenging but common themes can be extracted. Some RNA sequencing studies indicate that neurotransmission may be impaired/reduced in amyotrophic lateral sclerosis. D’Erchia et al.126 reported downregulation of over 15 genes associated with synaptic vesicle biology in the spinal cord of patients with amyotrophic lateral sclerosis. These notably included SNAP25 and Syntaxin 1B, which play a critical role in neurotransmitter release. Similar results were found in a study by Liu et al.,98 which performed single cell RNA sequencing of the SOD1G93A mouse brainstem. A large number of transcripts associated with synaptic transmission, function and assembly were found to be downregulated in symptomatic mice (P100). The most notable drivers of these pathways were decreased expression of neuroligin 1 and neurexin 3, both of which localize to excitatory glutamatergic synapses where they promote synaptic maturation and neurotransmission.127-129
Given the many observations of neuronal hyperexcitability in amyotrophic lateral sclerosis, it is possible that reduced synaptic strength occurs as a homeostatic mechanism downstream in the pathological cascade to compensate for increased firing. However, there are many lines of evidence that suggest alterations in basal neurotransmission occur both independently, and in parallel to, synaptic homeostatic responses to changes in neuronal excitability. The first of these is that changes to basal neurotransmission have been observed to occur very early in preclinical amyotrophic lateral sclerosis models112,130 and that these changes are also observed in vitro131 and in development-focused model organisms.132-134 The ‘earliness’ of these changes imply a fundamentality to synaptic changes caused by disease-associated mutations. The second is that disease-associated processes are commonly observed to directly interact with or indirectly affect the function of synaptic proteins that are strong determinants of synaptic strength, including synapsin,131 SV2,135 UNC13A124 and GluR1.136 The ‘directness’ of these interactions strengthen the fundamentality of the synaptic changes caused by disease-associated mutations. Together, these findings all indicate that neurotransmission in amyotrophic lateral sclerosis may be under constant suppressive pressure in the form of biochemical and molecular pathways that impair neurotransmitter release, and this might play a key role in driving degenerative mechanisms.
In the context of the ‘dying forward’ excitotoxicity hypothesis, these studies recontextualize any observations of increased upper motor neuron or premotor interneuron activity. This is because if these excitatory neurons suffer from suppressed neurotransmitter release, any increases in their activity might only serve to normalie glutamate release over time. It must be noted that changes in glutamate handling extrasynaptically have been postulated to occur in amyotrophic lateral sclerosis due to changes in the glutamate transporter, EAAT2. Decreased expression of EAAT2 is strongly correlated with degeneration in the ventral spinal cord of deceased amyotrophic lateral sclerosis patients137 and patient-derived synaptosomes are less competent at transporting glutamate.138 EAAT2 expression is also decreased in rodent models of amyotrophic lateral sclerosis139,140 as well as in dogs with canine degenerative myelopathy, which is a naturally occurring model of motor neuron disease.141,142 Together, these suggest that reduced EAAT2 may represent a strongly conserved pathological signature of amyotrophic lateral sclerosis. How this fits into our picture of pathogenesis remains to be determined as clinical trials targeting EAAT2 have proved ineffective.11 However, changes in glutamate handling must always be considered in the context of neuronal activity over time because activity-dependent synaptic vesicle release is a major source of extracellular glutamate. One possibility that has received little appreciation is that the low glutamate concentrations in the ventral horn of amyotrophic lateral sclerosis patients’ spinal cords might suggest that the combined changes in neuronal activation and glutamate handling results in a net reduction. More work on deciphering the sources and sinks of glutamate in the motor cortex and spinal cord will be needed to resolve this question. For example, a key missing piece of evidence for the excitotoxicity ‘dying forward’ hypothesis, and our understanding of the corticomotor system in general, are the neurotransmission properties of upper motor neuron axon terminations in the spinal cord (Fig. 1). Do they release more, or less, glutamate per action potential or over time? Although technically challenging, exploring upper motor neuron synaptic output will undoubtedly provide a core piece of evidence towards our understanding of the ‘dying forward’ hypothesis, and thereby the role of glutamate-mediated excitotoxicity.
Conclusion and key messages
The past three decades of attempting to treat amyotrophic lateral sclerosis by suppressing neuronal activation across the whole nervous system has demonstrated that our theories of the disease need to be revisited. The ‘dying forward’ hypothesis remains an attractive explanation for pathogenesis in amyotrophic lateral sclerosis. In support of this, studies on ALS patients and preclinical animal models converge on the concept that the motor cortex and the upper motor neurons display hallmarks of increased excitability either in the early stages of ALS or in the presymptomatic phase of familial ALS.44 While motor cortex hyperexcitability is clearly associated with amyotrophic lateral sclerosis, the precise molecular processes underlying excitotoxicity warrants further assessment. Mechanisms responding to potential perturbations in excitability need to be differentiated from pathological drivers of amyotrophic lateral sclerosis at a neuronal network, cellular and synaptic level along the corticospinal tract (Fig. 1). Fortunately, there are no technological limitations to our ability to ask these questions with the flurry of diverse and spatiotemporally accurate tools that the past decades have provided. Hopefully, recognition of this promotes increased focus on research into upper motor neuron activity in vivo as well as assessments of synaptic biology in amyotrophic lateral sclerosis.
Acknowledgements
We would like to thank Jose Vicente Garcia Cesar for his graphic art design of Fig. 1.
Contributor Information
G Lorenzo Odierna, Tasmanian School of Medicine, University of Tasmania, Hobart, TAS 7000, Australia.
Steve Vucic, Brain and Nerve Research Center, The University of Sydney, Sydney 2050, Australia.
Marcus Dyer, Menzies Institute for Medical Research, University of Tasmania, Hobart, TAS 7000, Australia; Department of Pharmaceutical and Pharmacological Sciences, Center for Neurosciences, Vrije Universiteit Brussel (VUB), 1090 Brussels, Belgium.
Tracey Dickson, Menzies Institute for Medical Research, University of Tasmania, Hobart, TAS 7000, Australia.
Adele Woodhouse, The Wicking Dementia Centre, University of Tasmania, Hobart, TAS 7000, Australia.
Catherine Blizzard, Tasmanian School of Medicine, University of Tasmania, Hobart, TAS 7000, Australia; Menzies Institute for Medical Research, University of Tasmania, Hobart, TAS 7000, Australia.
Funding
This work was supported by the Motor Neurone Disease Research Institute of Australia (MNDRA) Daniel McLoone Grant; the National Health and Medical Research Council (Ideas Grant) and the Tasmanian Masonic Research Foundation.
Competing interests
The authors report no competing interests.
References
- 1. Gunes ZI, Kan VW, Jiang S, Logunov E, Ye X, Liebscher S. Cortical hyperexcitability in the driver’s seat in ALS. Clin Transl Neurosci. 2022;6:5. [Google Scholar]
- 2. Menon P, Kiernan MC, Vucic S. Cortical hyperexcitability precedes lower motor neuron dysfunction in ALS. Clin Neurophysiol. 2015;126:803–809. [DOI] [PubMed] [Google Scholar]
- 3. Menon P, Higashihara M, van den Bos M, Geevasinga N, Kiernan MC, Vucic S. Cortical hyperexcitability evolves with disease progression in ALS. Ann Clin Transl Neurol. 2020;7:733–741. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4. Vucic S, Pavey N, Haidar M, Turner BJ, Kiernan MC. Cortical hyperexcitability: diagnostic and pathogenic biomarker of ALS. Neurosci Lett. 2021;759:136039. [DOI] [PubMed] [Google Scholar]
- 5. Shaw PJ, Ince PG. Glutamate, excitotoxicity and amyotrophic lateral sclerosis. J Neurol. 1997;244(Suppl 2):S3–S14. [DOI] [PubMed] [Google Scholar]
- 6. King AE, Woodhouse A, Kirkcaldie MT, Vickers JC. Excitotoxicity in ALS: Overstimulation, or overreaction? Exp Neurol. 2016;275(Pt 1):162–171. [DOI] [PubMed] [Google Scholar]
- 7. Pascuzzi RM, Shefner J, Chappell AS, et al. A phase II trial of talampanel in subjects with amyotrophic lateral sclerosis. Amyotroph Lateral Scler. 2010;11:266–271. [DOI] [PubMed] [Google Scholar]
- 8. Adiao KJB, Espiritu AI, Bagnas MAC. Efficacy and safety of mexiletine in amyotrophic lateral sclerosis: A systematic review of randomized controlled trials. Neurodegener Dis Manag. 2020;10:397–407. [DOI] [PubMed] [Google Scholar]
- 9. Park SB, Vucic S, Cheah BC, et al. Flecainide in amyotrophic lateral sclerosis as a neuroprotective strategy (FANS): a randomized placebo-controlled trial. EBioMed. 2015;2:1916–1922. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Diana A, Pillai R, Bongioanni P, O'Keeffe AG, Miller RG, Moore DH. Gamma aminobutyric acid (GABA) modulators for amyotrophic lateral sclerosis/motor neuron disease. Cochrane Database Syst Rev. 2017;1:Cd006049. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11. Cudkowicz ME, Titus S, Kearney M, et al. Safety and efficacy of ceftriaxone for amyotrophic lateral sclerosis: a multi-stage, randomised, double-blind, placebo-controlled trial. Lancet Neurol. 2014;13:1083–1091. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12. Ryberg H, Askmark H, Persson LI. A double-blind randomized clinical trial in amyotrophic lateral sclerosis using lamotrigine: effects on CSF glutamate, aspartate, branched-chain amino acid levels and clinical parameters. Acta Neurol Scand. 2003;108:1–8. [DOI] [PubMed] [Google Scholar]
- 13. Eisen A, Stewart H, Schulzer M, Cameron D. Anti-glutamate therapy in amyotrophic lateral sclerosis: a trial using lamotrigine. Can J Neurol Sci. 1993;20:297–301. [PubMed] [Google Scholar]
- 14. Hollander D, Pradas J, Kaplan R, McLeod HL, Evans WE, Munsat TL. High-dose dextromethorphan in amyotrophic lateral sclerosis: phase I safety and pharmacokinetic studies. Ann Neurol. 1994;36:920–924. [DOI] [PubMed] [Google Scholar]
- 15. Gredal O, Werdelin L, Bak S, et al. A clinical trial of dextromethorphan in amyotrophic lateral sclerosis. Acta Neurol Scand. 1997;96:8–13. [DOI] [PubMed] [Google Scholar]
- 16. Richards D, Morren JA, Pioro EP. Time to diagnosis and factors affecting diagnostic delay in amyotrophic lateral sclerosis. J Neurol Sci. 2020;417:117054. [DOI] [PubMed] [Google Scholar]
- 17. Sturmey E, Malaspina A. Blood biomarkers in ALS: Challenges, applications and novel frontiers. Acta Neurol Scand. 2022;146:375–388. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18. Bendotti C, Bonetto V, Pupillo E, et al. Focus on the heterogeneity of amyotrophic lateral sclerosis. Amyotroph Lateral Scler Frontotemporal Degener. 2020;21(7–8):485–495. [DOI] [PubMed] [Google Scholar]
- 19. Devine MS, Kiernan MC, Heggie S, McCombe PA, Henderson RD. Study of motor asymmetry in ALS indicates an effect of limb dominance on onset and spread of weakness, and an important role for upper motor neurons. Amyotroph Lateral Scler Frontotemporal Degener. 2014;15(7–8):481–487. [DOI] [PubMed] [Google Scholar]
- 20. Aizawa H, Kato H, Oba K, et al. Randomized phase 2 study of perampanel for sporadic amyotrophic lateral sclerosis. J Neurol. 2022;269:885–896. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21. Arundine M, Tymianski M. Molecular mechanisms of calcium-dependent neurodegeneration in excitotoxicity. Cell Calcium. 2003;34(4–5):325–337. [DOI] [PubMed] [Google Scholar]
- 22. Olney JW. Glutaate-induced retinal degeneration in neonatal mice. Electron microscopy of the acutely evolving lesion. J Neuropathol Exp Neurol. 1969;28:455–474. [DOI] [PubMed] [Google Scholar]
- 23. Ferrer I, Martin F, Serrano T, et al. Both apoptosis and necrosis occur following intrastriatal administration of excitotoxins. Acta Neuropathol. 1995;90:504–510. [DOI] [PubMed] [Google Scholar]
- 24. Eisen A. The dying forward hypothesis of ALS: Tracing its history. Brain Sci. 2021;11:300. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Delisle MB, Carpenter S. Neurofibrillary axonal swellings and amyotrophic lateral sclerosis. J Neurol Sci. 1984;63:241–250. [DOI] [PubMed] [Google Scholar]
- 26. Olney JW, Fuller T, De Gubareff T. Acute dendrotoxic changes in the hippocampus of kainate treated rats. Brain Res. 1979;176:91–100. [DOI] [PubMed] [Google Scholar]
- 27. Olney JW, Rhee V, Gubareff TD. Neurotoxic effects of glutamate on mouse area postrema. Brain Res. 1977;120:151–157. [DOI] [PubMed] [Google Scholar]
- 28. Plaitakis A, Caroscio JT. Abnormal glutamate metabolism in amyotrophic lateral sclerosis. Ann Neurol. 1987;22:575–579. [DOI] [PubMed] [Google Scholar]
- 29. Rothstein JD, Tsai G, Kuncl RW, et al. Abnormal excitatory amino acid metabolism in amyotrophic lateral sclerosis. Ann Neurol. 1990;28:18–25. [DOI] [PubMed] [Google Scholar]
- 30. Iwasaki Y, Ikeda K, Kinoshita M. Plasma amino acid levels in patients with amyotrophic lateral sclerosis. J Neurol Sci. 1992;107:219–222. [DOI] [PubMed] [Google Scholar]
- 31. Plaitakis A, Constantakakis E. Altered metabolism of excitatory amino acids, N-acetyl-aspartate and N-acetyl-aspartyl-glutamate in amyotrophic lateral sclerosis. Brain Res Bull. 1993;30(3–4):381–386. [DOI] [PubMed] [Google Scholar]
- 32. Shaw PJ, Forrest V, Ince PG, Richardson JP, Wastell HJ. CSF and plasma amino acid levels in motor neuron disease: elevation of CSF glutamate in a subset of patients. Neurodegeneration. 1995;4:209–216. [DOI] [PubMed] [Google Scholar]
- 33. Patten BM, Harati Y, Acosta L, Jung SS, Felmus MT. Free amino acid levels in amyotrophic lateral sclerosis. Ann Neurol. 1978;3:305–309. [DOI] [PubMed] [Google Scholar]
- 34. Perry TL, Krieger C, Hansen S, Eisen A. Amyotrophic lateral sclerosis: amino acid levels in plasma and cerebrospinal fluid. Ann Neurol. 1990;28:12–17. [DOI] [PubMed] [Google Scholar]
- 35. Camu W, Billiard M, Baldy-Moulinier M. Fasting plasma and CSF amino acid levels in amyotrophic lateral sclerosis: A subtype analysis. Acta Neurol Scand. 1993;88:51–55. [DOI] [PubMed] [Google Scholar]
- 36. Spreux-Varoquaux O, Bensimon G, Lacomblez L, et al. Glutamate levels in cerebrospinal fluid in amyotrophic lateral sclerosis: a reappraisal using a new HPLC method with coulometric detection in a large cohort of patients. J Neurol Sci. 2002;193:73–78. [DOI] [PubMed] [Google Scholar]
- 37. Andreadou E, Kapaki E, Kokotis P, et al. Plasma glutamate and glycine levels in patients with amyotrophic lateral sclerosis: the effect of riluzole treatment. Clin Neurol Neurosurg. 2008;110:222–226. [DOI] [PubMed] [Google Scholar]
- 38. Perry TL, Hansen S, Jones K. Brain glutamate deficiency in amyotrophic lateral sclerosis. Neurology. 1987;37:1845–1848. [DOI] [PubMed] [Google Scholar]
- 39. Plaitakis A, Constantakakis E, Smith J. The neuroexcitotoxic amino acids glutamate and aspartate are altered in the spinal cord and brain in amyotrophic lateral sclerosis. Ann Neurol. 1988;24:446–449. [DOI] [PubMed] [Google Scholar]
- 40. Tsai GC, Stauch-Slusher B, Sim L, et al. Reductions in acidic amino acids and N-acetylaspartylglutamate in amyotrophic lateral sclerosis CNS. Brain Res. 1991;556:151–156. [DOI] [PubMed] [Google Scholar]
- 41. Fujita K, Nagata Y, Honda M. [Free amino acid contents in the spinal cord of amyotrophic lateral sclerosis]. Rinsho Shinkeigaku. 1993;33:985–987. Japanese. [PubMed] [Google Scholar]
- 42. Ono S, Imai T, Takahashi K, et al. Alteration in amino acids in motor neurons of the spinal cord in amyotrophic lateral sclerosis. J Neurol Sci. 1999;167:121–126. [DOI] [PubMed] [Google Scholar]
- 43. Geevasinga N, Menon P, Özdinler PH, Kiernan MC, Vucic S. Pathophysiological and diagnostic implications of cortical dysfunction in ALS. Nat Rev Neurol. 2016;12:651–661. [DOI] [PubMed] [Google Scholar]
- 44. Vucic S, Nicholson GA, Kiernan MC. Cortical hyperexcitability may precede the onset of familial amyotrophic lateral sclerosis. Brain. 2008;131:1540–1550. [DOI] [PubMed] [Google Scholar]
- 45. Vucic S, Stanley Chen KH, Kiernan MC, et al. Clinical diagnostic utility of transcranial magnetic stimulation in neurological disorders. Updated report of an IFCN committee. Clin Neurophysiol. 2023;150:131–175. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46. Geevasinga N, Van den Bos M, Menon P, Vucic S. Utility of transcranial magnetic simulation in studying upper motor neuron dysfunction in amyotrophic lateral sclerosis. Brain Sci. 2021; 11(7):906. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47. Vucic S, Kiernan MC. Cortical excitability testing distinguishes Kennedy's disease from amyotrophic lateral sclerosis. Clin Neurophysiol. 2008;119:1088–1096. [DOI] [PubMed] [Google Scholar]
- 48. Geevasinga N, Menon P, Sue CM, et al. Cortical excitability changes distinguish the motor neuron disease phenotypes from hereditary spastic paraplegia. Eur J Neurol. 2015; 22:826–831.e57-8. [DOI] [PubMed] [Google Scholar]
- 49. Vucic S, Cheah BC, Yiannikas C, Kiernan MC. Cortical excitability distinguishes ALS from mimic disorders. Clin Neurophysiol. 2011;122:1860–1866. [DOI] [PubMed] [Google Scholar]
- 50. Menon P, Geevasinga N, Yiannikas C, Howells J, Kiernan MC, Vucic S. Sensitivity and specificity of threshold tracking transcranial magnetic stimulation for diagnosis of amyotrophic lateral sclerosis: a prospective study. Lancet Neurol. 2015;14:478–484. [DOI] [PubMed] [Google Scholar]
- 51. Rosen DR, Siddique T, Patterson D, et al. Mutations in Cu/Zn superoxide dismutase gene are associated with familial amyotrophic lateral sclerosis. Nature. 1993;362:59–62. [DOI] [PubMed] [Google Scholar]
- 52. Gurney ME, Pu H, Chiu AY, et al. Motor neuron degeneration in mice that express a human Cu, Zn superoxide dismutase mutation. Science. 1994;264:1772–1775. [DOI] [PubMed] [Google Scholar]
- 53. Neumann M, Sampathu DM, Kwong LK, et al. Ubiquitinated TDP-43 in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Science. 2006;314:130–133. [DOI] [PubMed] [Google Scholar]
- 54. Sreedharan J, Blair IP, Tripathi VB, et al. TDP-43 mutations in familial and sporadic amyotrophic lateral sclerosis. Science. 2008;319:1668–1672. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55. Wegorzewska I, Bell S, Cairns NJ, Miller TM, Baloh RH. TDP-43 mutant transgenic mice develop features of ALS and frontotemporal lobar degeneration. Proc Natl Acad Sci U S A. 2009;106:18809–18814. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56. Arnold ES, Ling SC, Huelga SC, et al. ALS-linked TDP-43 mutations produce aberrant RNA splicing and adult-onset motor neuron disease without aggregation or loss of nuclear TDP-43. Proc Natl Acad Sci U S A. 2013;110:E736–E745. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57. Igaz LM, Kwong LK, Lee EB, et al. Dysregulation of the ALS-associated gene TDP-43 leads to neuronal death and degeneration in mice. J Clin Invest. 2011;121:726–738. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. Blair IP, Williams KL, Warraich ST, et al. FUS mutations in amyotrophic lateral sclerosis: clinical, pathological, neurophysiological and genetic analysis. J Neurol Neurosurg Psychiatry. 2010;81:639–645. [DOI] [PubMed] [Google Scholar]
- 59. Bosco DA, Lemay N, Ko HK, et al. Mutant FUS proteins that cause amyotrophic lateral sclerosis incorporate into stress granules. Hum Mol Genet. 2010;19:4160–4175. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60. Scekic-Zahirovic J, Sendscheid O, El Oussini H, et al. Toxic gain of function from mutant FUS protein is crucial to trigger cell autonomous motor neuron loss. Embo J. 2016;35:1077–1097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61. DeJesus-Hernandez M, Mackenzie IR, Boeve BF, et al. Expanded GGGGCC hexanucleotide repeat in noncoding region of C9ORF72 causes chromosome 9p-linked FTD and ALS. Neuron. 2011;72:245–256. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62. Renton AE, Majounie E, Waite A, et al. A hexanucleotide repeat expansion in C9ORF72 is the cause of chromosome 9p21-linked ALS-FTD. Neuron. 2011;72:257–268. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63. Liu Y, Pattamatta A, Zu T, et al. C9orf72 BAC mouse model with motor deficits and neurodegenerative features of ALS/FTD. Neuron. 2016;90:521–534. [DOI] [PubMed] [Google Scholar]
- 64. Amalyan S, Tamboli S, Lazarevich I, Topolnik D, Bouman LH, Topolnik L. Enhanced motor cortex output and disinhibition in asymptomatic female mice with C9orf72 genetic expansion. Cell Rep. 2022;40:111043. [DOI] [PubMed] [Google Scholar]
- 65. Scekic-Zahirovic J, Sanjuan-Ruiz I, Kan V, et al. Cytoplasmic FUS triggers early behavioral alterations linked to cortical neuronal hyperactivity and inhibitory synaptic defects. Nat Commun. 2021;12:3028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66. Kim J, Hughes EG, Shetty AS, et al. 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. 2017;37:9037–9053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67. Liang B, Thapa R, Zhang G, et al. Aberrant neural activity in prefrontal pyramidal neurons lacking TDP-43 precedes neuron loss. Prog Neurobiol. 2022;215:102297. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68. Saba L, Viscomi MT, Martini A, et al. Modified age-dependent expression of NaV1.6 in an ALS model correlates with motor cortex excitability alterations. Neurobiol Dis. 2019;130:104532. [DOI] [PubMed] [Google Scholar]
- 69. Zhang W, Zhang L, Liang B, et al. Hyperactive somatostatin interneurons contribute to excitotoxicity in neurodegenerative disorders. Nat Neurosci. 2016;19:557–559. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70. Dyer MS, Reale LA, Lewis KE, et al. Mislocalisation of TDP-43 to the cytoplasm causes cortical hyperexcitability and reduced excitatory neurotransmission in the motor cortex. J Neurochem. 2021;157:1300–1315. [DOI] [PubMed] [Google Scholar]
- 71. Devlin AC, Burr K, Borooah S, et al. Human iPSC-derived motoneurons harbouring TARDBP or C9ORF72 ALS mutations are dysfunctional despite maintaining viability. Nat Commun. 2015;6:5999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. Buskila Y, Kékesi O, Bellot-Saez A, et al. Dynamic interplay between H-current and M-current controls motoneuron hyperexcitability in amyotrophic lateral sclerosis. Cell Death Dis. Apr. 5 2019;10:310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73. Schulz DJ, Goaillard JM, Marder E. Variable channel expression in identified single and electrically coupled neurons in different animals. Nat Neurosci. 2006;9:356–362. [DOI] [PubMed] [Google Scholar]
- 74. Goaillard JM, Marder E. Ion channel degeneracy, variability, and covariation in neuron and circuit resilience. Annu Rev Neurosci. 2021;44:335–357. [DOI] [PubMed] [Google Scholar]
- 75. Martin E, Cazenave W, Cattaert D, Branchereau P. Embryonic alteration of motoneuronal morphology induces hyperexcitability in the mouse model of amyotrophic lateral sclerosis. Neurobiol Dis. Jun 2013;54:116–126. [DOI] [PubMed] [Google Scholar]
- 76. Filipchuk A, Pambo-Pambo A, Gaudel F, et al. Early hypoexcitability in a subgroup of spinal motoneurons in superoxide dismutase 1 transgenic mice, a model of amyotrophic lateral sclerosis. Neuroscience. 2021;463:337–353. [DOI] [PubMed] [Google Scholar]
- 77. Martínez-Silva ML, Imhoff-Manuel RD, Sharma A, et al. Hypoexcitability precedes denervation in the large fast-contracting motor units in two unrelated mouse models of ALS. Elife. 2018;7:e30955. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78. Delestrée N, Manuel M, Iglesias C, Elbasiouny SM, Heckman CJ, Zytnicki D. Adult spinal motoneurones are not hyperexcitable in a mouse model of inherited amyotrophic lateral sclerosis. J Physiol. 2014;592:1687–1703. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79. Meehan CF, Moldovan M, Marklund SL, Graffmo KS, Nielsen JB, Hultborn H. Intrinsic properties of lumbar motor neurones in the adult G127insTGGG superoxide dismutase-1 mutant mouse in vivo: evidence for increased persistent inward currents. Acta Physiol. 2010;200:361–376. [DOI] [PubMed] [Google Scholar]
- 80. Pambo-Pambo A, Durand J, Gueritaud JP. Early excitability changes in lumbar motoneurons of transgenic SOD1G85R and SOD1G(93A-Low) mice. J Neurophysiol. 2009;102:3627–3642. [DOI] [PubMed] [Google Scholar]
- 81. Jensen DB, Kadlecova M, Allodi I, Meehan CF. Spinal motoneurones are intrinsically more responsive in the adult G93A SOD1 mouse model of amyotrophic lateral sclerosis. J Physiol. 2020;598:4385–4403. [DOI] [PubMed] [Google Scholar]
- 82. Leroy F, Lamotte d'Incamps B, Imhoff-Manuel RD, Zytnicki D. Early intrinsic hyperexcitability does not contribute to motoneuron degeneration in amyotrophic lateral sclerosis. Elife. 2014;3:e04046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83. Venugopal S, Hsiao CF, Sonoda T, Wiedau-Pazos M, Chandler SH. Homeostatic dysregulation in membrane properties of masticatory motoneurons compared with oculomotor neurons in a mouse model for amyotrophic lateral sclerosis. J Neurosci. 2015;35:707–720. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84. Quinlan KA, Schuster JE, Fu R, Siddique T, Heckman CJ. Altered postnatal maturation of electrical properties in spinal motoneurons in a mouse model of amyotrophic lateral sclerosis. J Physiol. 2011;589(Pt 9):2245–2260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85. Fogarty MJ, Klenowski PM, Lee JD, et al. Cortical synaptic and dendritic spine abnormalities in a presymptomatic TDP-43 model of amyotrophic lateral sclerosis. Sci Rep. 2016;6:37968. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86. Fogarty MJ, Noakes PG, Bellingham MC. 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. 2015;35:643–647. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87. Saba L, Viscomi MT, Caioli S, et al. 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. [DOI] [PubMed] [Google Scholar]
- 88. Handley EE, Pitman KA, Dawkins E, et al. Synapse dysfunction of layer V pyramidal neurons precedes neurodegeneration in a mouse model of TDP-43 proteinopathies. Cereb Cortex. 2017;27:3630–3647. [DOI] [PubMed] [Google Scholar]
- 89. Khademullah CS, Aqrabawi AJ, Place KM, et al. Cortical interneuron-mediated inhibition delays the onset of amyotrophic lateral sclerosis. Brain. 2020;143:800–810. [DOI] [PubMed] [Google Scholar]
- 90. Nieto-Gonzalez JL, Moser J, Lauritzen M, Schmitt-John T, Jensen K. Reduced GABAergic inhibition explains cortical hyperexcitability in the wobbler mouse model of ALS. Cerebral Cortex. 2010;21:625–635. [DOI] [PubMed] [Google Scholar]
- 91. Van den Bos MAJ, Higashihara M, Geevasinga N, Menon P, Kiernan MC, Vucic S. Imbalance of cortical facilitatory and inhibitory circuits underlies hyperexcitability in ALS. Neurology. 2018;91:e1669–e1676. [DOI] [PubMed] [Google Scholar]
- 92. Fogarty MJ, Mu EWH, Noakes PG, Lavidis NA, Bellingham MC. Marked changes in dendritic structure and spine density precede significant neuronal death in vulnerable cortical pyramidal neuron populations in the SOD1G93A mouse model of amyotrophic lateral sclerosis. Acta Neuropathol Commun. 2016;4:77. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93. Berry KP, Nedivi E. Spine dynamics: are they all the same? Neuron. 2017;96:43–55. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94. Vucic S, Cheah BC, Kiernan MC. Defining the mechanisms that underlie cortical hyperexcitability in amyotrophic lateral sclerosis. Exp Neurol. 2009;220:177–182. [DOI] [PubMed] [Google Scholar]
- 95. Agarwal S, Highton-Williamson E, Caga J, et al. Motor cortical excitability predicts cognitive phenotypes in amyotrophic lateral sclerosis. Sci Rep. 2021;11:2172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96. Clark RM, Brizuela M, Blizzard CA, Dickson TC. Reduced excitability and increased neurite complexity of cortical interneurons in a familial mouse model of amyotrophic lateral sclerosis. Front Cell Neurosci. 2018;12:328. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97. Saez-Atienzar S, Bandres-Ciga S, Langston RG, et al. Genetic analysis of amyotrophic lateral sclerosis identifies contributing pathways and cell types. Sci Adv. 2021;7:eabd9036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98. Liu W, Venugopal S, Majid S, et al. Single-cell RNA-seq analysis of the brainstem of mutant SOD1 mice reveals perturbed cell types and pathways of amyotrophic lateral sclerosis. Neurobiol Dis. 2020;141:104877. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99. Gittings LM, Alsop EB, Antone J, et al. Cryptic exon detection and transcriptomic changes revealed in single-nuclei RNA sequencing of C9ORF72 patients spanning the ALS-FTD spectrum. Acta Neuropathol. 2023;146:433–450. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100. Allodi I, Montañana-Rosell R, Selvan R, Löw P, Kiehn O. Locomotor deficits in a mouse model of ALS are paralleled by loss of V1-interneuron connections onto fast motor neurons. Nat Commun. 2021;12:3251. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101. Chang Q, Martin LJ. Glycinergic innervation of motoneurons is deficient in amyotrophic lateral sclerosis mice: A quantitative confocal analysis. Am J Pathol. 2009;174:574–585. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102. Salamatina A, Yang JH, Brenner-Morton S, et al. Differential loss of spinal interneurons in a mouse model of ALS. Neuroscience. 2020;450:81–95. [DOI] [PubMed] [Google Scholar]
- 103. Wootz H, Fitzsimons-Kantamneni E, Larhammar M, et al. Alterations in the motor neuron-Renshaw cell circuit in the Sod1(G93A) mouse model. J Comp Neurol. 2013;521:1449–1469. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104. Morrison BM, Janssen WG, Gordon JW, Morrison JH. Time course of neuropathology in the spinal cord of G86R superoxide dismutase transgenic mice. J Comp Neurol. 1998;391:64–77. [DOI] [PubMed] [Google Scholar]
- 105. Stephens B, Guiloff RJ, Navarrete R, Newman P, Nikhar N, Lewis P. Widespread loss of neuronal populations in the spinal ventral horn in sporadic motor neuron disease. A morphometric study. J Neurol Sci. 2006;244(1–2):41–58. [DOI] [PubMed] [Google Scholar]
- 106. Sunico CR, Domínguez G, García-Verdugo JM, Osta R, Montero F, Moreno-López B. Reduction in the motoneuron inhibitory/excitatory synaptic ratio in an early-symptomatic mouse model of amyotrophic lateral sclerosis. Brain Pathol. 2011;21:1–15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107. Casas C, Herrando-Grabulosa M, Manzano R, Mancuso R, Osta R, Navarro X. Early presymptomatic cholinergic dysfunction in a murine model of amyotrophic lateral sclerosis. Brain Behav. 2013;3:145–158. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108. Nagao M, Misawa H, Kato S, Hirai S. Loss of cholinergic synapses on the spinal motor neurons of amyotrophic lateral sclerosis. J Neuropathol Exp Neurol. 1998;57:329–333. [DOI] [PubMed] [Google Scholar]
- 109. Bak AN, Djukic S, Kadlecova M, Braunstein TH, Jensen DB, Meehan CF. Cytoplasmic TDP-43 accumulation drives changes in C-bouton number and size in a mouse model of sporadic Amyotrophic Lateral Sclerosis. Mol Cell Neurosci. 2023;125:103840. [DOI] [PubMed] [Google Scholar]
- 110. Gallart-Palau X, Tarabal O, Casanovas A, et al. Neuregulin-1 is concentrated in the postsynaptic subsurface cistern of C-bouton inputs to α-motoneurons and altered during motoneuron diseases. FASEB J. 2014;28:3618–3632. [DOI] [PubMed] [Google Scholar]
- 111. Rizzuto E, Pisu S, Musarò A, Del Prete Z. Measuring neuromuscular junction functionality in the SOD1(G93A) animal model of amyotrophic lateral sclerosis. Ann Biomed Eng. 2015;43:2196–2206. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112. Chand KK, Lee KM, Lee JD, et al. Defects in synaptic transmission at the neuromuscular junction precede motor deficits in a TDP-43Q331K transgenic mouse model of amyotrophic lateral sclerosis. FASEB J. 2018;32:2676–2689. [DOI] [PubMed] [Google Scholar]
- 113. Tremblay E, Martineau É, Robitaille R. Opposite synaptic alterations at the neuromuscular junction in an ALS mouse model: when motor units matter. J Neurosci. 2017;37:8901–8918. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114. Maselli RA, Wollman RL, Leung C, et al. Neuromuscular transmission in amyotrophic lateral sclerosis. Muscle Nerve. 1993;16:1193–1203. [DOI] [PubMed] [Google Scholar]
- 115. van Zundert B, Peuscher MH, Hynynen M, et al. Neonatal neuronal circuitry shows hyperexcitable disturbance in a mouse model of the adult-onset neurodegenerative disease amyotrophic lateral sclerosis. J Neurosci. 2008;28:10864–10874. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116. Cavarsan CF, Steele PR, Genry LT, et al. Inhibitory interneurons show early dysfunction in a SOD1 mouse model of amyotrophic lateral sclerosis. J Physiol. 2023;601:647–667. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117. Schoen M, Reichel JM, Demestre M, et al. Super-resolution microscopy reveals presynaptic localization of the ALS/FTD related protein FUS in hippocampal neurons. Front Cell Neurosci. 2015;9:496. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118. Xiao S, McKeever PM, Lau A, Robertson J. Synaptic localization of C9orf72 regulates post-synaptic glutamate receptor 1 levels. Acta Neuropathol Commun. 2019;7:161. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119. Narayanan RK, Mangelsdorf M, Panwar A, Butler TJ, Noakes PG, Wallace RH. Identification of RNA bound to the TDP-43 ribonucleoprotein complex in the adult mouse brain. Amyotroph Lateral Scler Frontotemporal Degener. 2013;14:252–260. [DOI] [PubMed] [Google Scholar]
- 120. Deshpande D, Higelin J, Schoen M, et al. Synaptic FUS localization during motoneuron development and its accumulation in human ALS synapses. Front Cell Neurosci. 2019;13:256. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121. Wang IF, Wu LS, Chang HY, Shen CK. TDP-43, the signature protein of FTLD-U, is a neuronal activity-responsive factor. J Neurochem. 2008;105:797–806. [DOI] [PubMed] [Google Scholar]
- 122. Polymenidou M, Lagier-Tourenne C, Hutt KR, et al. Long pre-mRNA depletion and RNA missplicing contribute to neuronal vulnerability from loss of TDP-43. Nat Neurosci. 2011;14:459–468. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123. Qiu H, Lee S, Shang Y, et al. ALS-associated mutation FUS-R521C causes DNA damage and RNA splicing defects. J Clin Invest. 2014;124:981–999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124. Ma XR, Prudencio M, Koike Y, et al. TDP-43 represses cryptic exon inclusion in the FTD-ALS gene UNC13A. Nature. 2022;603:124–130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125. Augustine GJ, Kasai H. Bernard Katz, quantal transmitter release and the foundations of presynaptic physiology. J Physiol. 2007;578:623–625. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126. D’Erchia AM, Gallo A, Manzari C, et al. Massive transcriptome sequencing of human spinal cord tissues provides new insights into motor neuron degeneration in ALS. Sci Rep. 2017;7:10046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127. Zhang R, Jiang H, Liu Y, He G. Structure, function, and pathology of Neurexin-3. Genes Dis. 2023;10:1908–1919. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128. Song J-Y, Ichtchenko K, Südhof TC, Brose N. Neuroligin 1 is a postsynaptic cell-adhesion molecule of excitatory synapses. Proc Natl Acad Sci U S A. 1999;96:1100–1105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129. Zeidan A, Ziv NE. Neuroligin-1 loss is associated with reduced tenacity of excitatory synapses. PLoS One. 2012;7:e42314. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130. Rocha MC, Pousinha PA, Correia AM, Sebastião AM, Ribeiro JA. Early changes of neuromuscular transmission in the SOD1(G93A) mice model of ALS start long before motor symptoms onset. PLoS One. 2013;8:e73846. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131. Bauer CS, Cohen RN, Sironi F, et al. An interaction between synapsin and C9orf72 regulates excitatory synapses and is impaired in ALS/FTD. Acta Neuropathol. 2022;144:437–464. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132. Shahidullah M, Le Marchand SJ, Fei H, et al. Defects in synapse structure and function precede motor neuron degeneration in Drosophila models of FUS-related ALS. J Neurosci. 2013;33:19590–19598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133. Butti Z, Pan YE, Giacomotto J, Patten SA. Reduced C9orf72 function leads to defective synaptic vesicle release and neuromuscular dysfunction in zebrafish. Commun Biol. 2021;4:792. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134. Diaper DC, Adachi Y, Sutcliffe B, et al. Loss and gain of Drosophila TDP-43 impair synaptic efficacy and motor control leading to age-related neurodegeneration by loss-of-function phenotypes. Hum Mol Genet. 2013;22:1539–1557. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135. Jensen BK, Schuldi MH, McAvoy K, et al. Synaptic dysfunction induced by glycine-alanine dipeptides in C9orf72-ALS/FTD is rescued by SV2 replenishment. EMBO Mol Med. 2020;12:e10722. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136. Wong CE, Jin LW, Chu YP, Wei WY, Ho PC, Tsai KJ. TDP-43 proteinopathy impairs mRNP granule mediated postsynaptic translation and mRNA metabolism. Theranostics. 2021;11:330–345. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137. Sasaki S, Komori T, Iwata M. Excitatory amino acid transporter 1 and 2 immunoreactivity in the spinal cord in amyotrophic lateral sclerosis. Acta Neuropathol. 2000;100:138–144. [DOI] [PubMed] [Google Scholar]
- 138. Rothstein JD, Martin LJ, Kuncl RW. Decreased glutamate transport by the brain and spinal cord in amyotrophic lateral sclerosis. N Engl J Med. 1992;326:1464–1468. [DOI] [PubMed] [Google Scholar]
- 139. Bruijn LI, Becher MW, Lee MK, et al. ALS-linked SOD1 mutant G85R mediates damage to astrocytes and promotes rapidly progressive disease with SOD1-containing inclusions. Neuron. 1997;18:327–338. [DOI] [PubMed] [Google Scholar]
- 140. Bendotti C, Tortarolo M, Suchak SK, et al. Transgenic SOD1 G93A mice develop reduced GLT-1 in spinal cord without alterations in cerebrospinal fluid glutamate levels. J Neurochem. 2001;79:737–746. [DOI] [PubMed] [Google Scholar]
- 141. Ogawa M, Uchida K, Yamato O, Inaba M, Uddin MM, Nakayama H. Neuronal loss and decreased GLT-1 expression observed in the spinal cord of Pembroke Welsh Corgi dogs with canine degenerative myelopathy. Vet Pathol. 2014;51:591–602. [DOI] [PubMed] [Google Scholar]
- 142. Nardone R, Höller Y, Taylor AC, et al. Canine degenerative myelopathy: A model of human amyotrophic lateral sclerosis. Zoology (Jena ). 2016; 119:64–73. [DOI] [PubMed] [Google Scholar]