Skip to main content
Translational Psychiatry logoLink to Translational Psychiatry
. 2022 Jun 22;12:257. doi: 10.1038/s41398-022-02024-7

Neuronal hyperexcitability in Alzheimer’s disease: what are the drivers behind this aberrant phenotype?

Helena Targa Dias Anastacio 1,2, Natalie Matosin 1,2, Lezanne Ooi 1,2,
PMCID: PMC9217953  PMID: 35732622

Abstract

Alzheimer’s disease (AD) is a progressive neurodegenerative disorder leading to loss of cognitive abilities and ultimately, death. With no cure available, limited treatments mostly focus on symptom management. Identifying early changes in the disease course may provide new therapeutic targets to halt or reverse disease progression. Clinical studies have shown that cortical and hippocampal hyperactivity are a feature shared by patients in the early stages of disease, progressing to hypoactivity during later stages of neurodegeneration. The exact mechanisms causing neuronal excitability changes are not fully characterized; however, animal and cell models have provided insights into some of the factors involved in this phenotype. In this review, we summarize the evidence for neuronal excitability changes over the course of AD onset and progression and the molecular mechanisms underpinning these differences. Specifically, we discuss contributors to aberrant neuronal excitability, including abnormal levels of intracellular Ca2+ and glutamate, pathological amyloid β (Aβ) and tau, genetic risk factors, including APOE, and impaired inhibitory interneuron and glial function. In light of recent research indicating hyperexcitability could be a predictive marker of cognitive dysfunction, we further argue that the hyperexcitability phenotype could be leveraged to improve the diagnosis and treatment of AD, and present potential targets for future AD treatment development.

Subject terms: Molecular neuroscience, Diseases

Alzheimer’s disease and neuronal hyperexcitability

It is estimated that approximately 50 million people worldwide suffer from dementia [1]. Alzheimer’s disease (AD) is the most common cause of dementia, accounting for 60–80% of all cases [2]. AD causes a progressive decline in cognitive function and eventually leads to death. Amyloid plaque deposition and neurofibrillary tangles (NFT) in the brain are the main pathological hallmarks of this disease. Amyloid plaques are comprised of insoluble Aβ peptides that accumulate in the extracellular space [3], while NFTs are intraneuronal aggregates containing hyperphosphorylated and misfolded tau [4]. AD can be caused by mutations in genes involved in the processing of Aβ, including APP, PSEN1, and PSEN2 [5]. However, familial inherited genetic cases of the disease (familial AD, FAD) represent only around 2% of total cases [6]. The majority of cases are sporadic AD or late-onset Alzheimer’s disease (LOAD), where the disease is driven by a combination of genetic and other risk factors. One of the major genetic risk factors for LOAD is the APOE4 allele. There are three different alleles for the APOE gene: ε2, ε3, and ε4. The allele ε3 is the most common in the general population, and while ε4 increases the risk for AD, ε2 is protective against the disease [7]. Despite decades of research into the genetic basis and pathological mechanisms of AD, there is no treatment available to halt or reverse its progression, leading to an urgency for the development of new treatment options. Neuronal hyperexcitability can be detected, in a non-invasive manner, before the onset of dementia [8, 9]. Therefore, potentially, this phenotype could be harnessed to identify individuals at risk and start treatments before degeneration progresses too far.

Hyperexcitability can be defined as the increased likelihood that a neuron will be activated by a certain stimulus. In patients, neuronal hyperexcitability has been observed via enhanced brain activity when individuals perform a memory-encoding task [810]. In preclinical animal and cell models of AD, neurons show a higher frequency of action potential firing, as well as a lower threshold for firing [1115]. Importantly, activity recordings at the level of single neurons, neuronal networks or even entire brain regions consistently show hyperexcitability in the early stages of AD (summarized in Table 1). This evidence comes from laboratory models of AD, as well as living patients, showcasing the potential of neuronal excitability changes as a biomarker for early detection of AD. Beyond that, the molecular mechanisms driving neuronal hyperexcitability are also potential targets for therapeutic intervention, which holds promise for halting or reversing AD. Recent data from mouse studies show that reversing abnormal neuronal excitability in brain regions affected early in the process of Aβ deposition not only reduces Aβ load in the same brain region but also prevents the spread of Aβ pathology to other areas of the brain [16]. Since Aβ accumulation is thought to contribute to memory deficits and neurodegeneration, these results provide initial evidence that reducing neuronal excitability could provide a potential therapeutic avenue for AD. However, the exact mechanisms leading to hyperexcitability are not yet fully understood and several questions need to be answered before we can effectively advance with treatment development. In this review, we summarize the current evidence available on neuronal excitability changes in AD, highlight the gaps in the field and identify pathways forward for future research.

Table 1.

Evidence of neuronal hyperexcitability in patients, animal and cell models of Alzheimer’s disease.

Study type/ Neuronal subtype Model/Group of study Functional phenotype Methodology Reference
Human studies
Clinical study 16 non-demented APOE4 carriers (two APOE4/4 and 14 APOE3/4 subjects) Increased activity of parahippocampal, left hippocampal, parietal, temporal, and prefrontal regions of APOE4 carriers fMRI [18]
14 non-demented APOE3/3 individuals
Clinical study Ten controls Increased hippocampal activity in MCI individuals compared to controls fMRI [9]
Nine mild MCI patients Decreased hippocampal and entorhinal activity in AD patients compared to controls
Ten AD patients
Clinical study 90 controls Increased activity of the hippocampus, frontal, and temporal lobes of asymptomatic offspring of AD patients fMRI [17]
95 asymptomatic offspring of AD patients
Clinical study 15 controls Increased hippocampal activity in less impaired MCI subjects compared to controls fMRI [8]
15 less impaired MCI patients Decreased hippocampal activity in more impaired MCI and mild AD subjects compared to control
12 more impaired MCI patients
Ten mild AD patients
Clinical study 19 controls Increased activity in the posterior hippocampal, parahippocampal and fusiform regions of MCI patients vs controls fMRI [10]
14 subjects with MCI Brain activity of AD patients was not significantly different from control
11 patients with mild AD
Animal models
Frontal, central, parietal, and occipital cortices of freely moving control and hAPP-J20 mice 3–7 months non-transgenic mice (n = not specified)  hAPP-J20 mice had frequent generalized cortical epileptiform discharges, which were absent in the control EEG [26]
3–7 months mice expressing hAPP
Swedish and Indiana mutations (n = 6 mice)
Neurons of L2/3 frontal cortex of live control and APP23xPS45 mice 6–10 months control mice (n = 10 mice; 564 neurons) Greater number of hyperactive (>4 transients/min) neurons in APPswe/PS1G384 mice Calcium imaging [11]
6–10 months APPswe/PS1G384A mice (n = 20 mice; 564 neurons)
L2/3 pyramidal neurons of cortical slices 3–4.5 months non-transgenic mice (n = 10 mice; 6 neurons) Current injection induced action potential firing in APdE9 mice at subthreshold stimulus compared to controls Whole-cell patch-clamp [15]
Frontal cortex of freely moving control and APdE9 mice 3–4.5 months mice harboring APPswe and PSEN1dE9 mutations 25–65% of APdE9 mice had seizures, while none of the control animals exhibited this phenotype EEG
(n = 20 mice; nine neurons)
Pyramidal neurons of lateral amygdala slices of APOE3 and APOE4 mice 1 and 7 months human APOE3/E3 knock-in mice (n = 11 mice; 23 neurons) Reduced frequency of spontaneous excitatory postsynaptic currents in APOE4/E4 mice Whole-cell patch-clamp [32]
1 and 7 months human APOE4/E4 knock-in mice (n = 12 mice; 28 neurons)
CA1 pyramidal neurons of live control and APP23xPS45 mice 1–2 months control mice (n = 6 mice; 693 neurons) Greater number of hyperactive (>20 transients/min) neurons in APPswe/PS1G384A mice Calcium imaging [28]
1–2 months APPswe/PS1G384A mice (n = 7 mice; 818 neurons)
6–7 months control mice (n = 5 mice; 312 neurons)
6–7 months APPswe/PS1G384A mice (n = 5 mice; 349 neurons)
Neurons of CA1 hippocampal slices of APP/PS1 mice 10–14 months control mice (n = 16 mice; 35–75 neurons) Higher frequency of spontaneous action potential in APPswe/PS1M146V animals Whole-cell patch-clamp [12]
10–14 months APPswe/PS1M146V mice (n = 16 mice; 44–69 neurons)
Cortex and hippocampus of live control and Tg2576 mice 5wo WT mice (n = 17) Synchronized transient spike-like events were detected in Tg2576 mice but absent in WT mice EEG [29]
5wo APPswe mice (n = 9)
Neurons of L2/3 frontal cortex of live control and APP23xPS45 mice 10–14 months control mice (n = 9 mice; 226 neurons) Greater number of hyperactive (>4 transients/min) neurons in APPswe/PS1G384A Calcium imaging [27]
10–14 months APPswe/PS1G384A (n = 10 mice; 260 neurons)
2D and 3D cell culture models
iPSC-derived neurons iPSCs from one sporadic AD patient Sporadic AD neurons had spontaneous Ca2+ responses and control neurons remained inactive in the absence of stimulus Calcium imaging [13]
(n > 30 from three independent experiments) iPSCs from one healthy individual
iPSC-derived neurons APOE4/4 iPSCs from one sporadic AD patient Higher frequency of miniature excitatory postsynaptic current in APOE4/4 neurons Whole-cell patch-clamp [37]
(n = 7–9 from three independent cultures) APOE3/3 isogenic control iPSCs
3D coculture of iPSC-derived neurons and astrocytes (n = >3 independent experiments) Lentiviral-transduced human neural progenitor cells expressing both Increased spontaneous Ca2+ transients in FAD neurons Calcium imaging [38]
Swedish and London APP
mutations
Control human neural progenitor cells
2D and 3D cultures of cortical neurons iPSCs from one healthy individual Higher frequency of spontaneous action potential in 2D AD neurons Whole-cell patch-clamp [14]
(n = 13) iPSCs from healthy individuals edited by CRISPR/Cas9 to express Increase in spontaneous action potential firing rate in 3D AD neurons MEA
APPswe or PS1M146V mutation
iPSCs from another healthy individual
iPSCs from healthy individuals edited by TALEN to express PS1dE9 mutation

AD Alzheimer’s disease, EEG electroencephalogram, FAD familial Alzheimer’s disease, fMRI functional magnetic resonance imaging, iPSC induced pluripotent stem cells, MCI mild cognitive impairment, MEA microelectrode array, WT wild-type.

Evidence of neuronal network hyperactivity from human studies

A plethora of clinical studies shows hyperactivity of the cortical and hippocampal brain areas in patients with both sporadic and familial forms of AD [810, 17, 18] This hyperactive phenotype is observable at the earlier stages of the disease before there is pronounced cell loss and hypoactivity [17, 19]. This is important because it could allow for early identification and intervention.

Neuronal hyperactivity has been observed in patients with mild cognitive impairment (MCI), a state between the expected cognitive changes during aging and dementia. Individuals suffering from MCI have a greater risk of developing AD, but MCI does not always progress to dementia [20]. A number of clinical studies have used functional magnetic resonance imaging (fMRI) to measure changes in blood oxygenation as a proxy for brain activity. While performing associative memory-encoding tasks, patients with MCI displayed increased brain activity in the hippocampal, parahippocampal, and fusiform regions. The AD group, on the other hand, showed hippocampal and entorhinal hypoactivation [8, 9], although one study found no difference between AD and control [10]. Together these findings suggest neuronal hyperactivity precedes hypoactivity in AD. An interesting study by Basset et al. [17] investigated the neuronal activity of non-symptomatic children of autopsy-confirmed AD patients with at least one more first-degree relative clinically diagnosed with AD. The fMRI scans acquired during an associative learning task showed greater activity in the frontal and temporal lobes, including the hippocampus, compared to age-matched controls without any history of first-degree relatives with suspected dementia. Their results demonstrated that individuals at genetic risk of developing LOAD showed increased brain activity, starting as early as a decade before the onset age of their parents’ disease.

It is important to note that excessive neuronal network activation can also manifest as epilepsy. A meta-analysis of 30 studies concluded that AD patients have an increased risk of suffering from epilepsy and seizures [21]. The risk for epileptic seizures is particularly higher in patients with younger-onset AD and is greater during earlier stages of the disease [22, 23]. Together, these results suggest brain hyperactivity occurs early in the development of AD. Since imaging studies examining living patients are unable to provide us with the resolution necessary to observe pathology at the cellular and molecular levels, we turn to animal and cell models of the disease for this purpose.

Animal models examining hyperexcitability

A variety of mouse models to study AD has been generated by genetically modifying animals to express AD-related mutations, most commonly in the APP and PSEN1 genes. Less commonly, but still of great importance, are mouse models expressing the human APOE4 allele, the main genetic risk factor for LOAD [24]. Mouse models of AD develop several of the main hallmarks of the disease, such as amyloid plaques and/or tau pathology and cognitive impairment [25], providing a model system that can be manipulated to identify pathological mechanisms and test potential treatments.

Aberrant neuronal excitability has been reported in the frontal cortex and hippocampus of several animal models of AD [11, 12, 2628]. For example, by measuring cortical and hippocampal electrical activity with electroencephalogram (EEG) recordings, Palop et al. [26] detected spontaneous non-convulsive seizure activity in hAPPFAD mice, which were absent in controls. Busche et al. [11] used two-photon Ca2+ imaging to analyse brain activity in real-time in vivo by measuring intracellular Ca2+ transients of individual neurons of layer 2/3 frontal cortex. In double transgenic 1.5 to 2-month-old APP23xPS45 mice, no significant differences in Ca2+ transients were found between AD and control mice. At this age, no memory deficits were detected, and amyloid plaques were absent, as they start to form at the age of 3 months. However, at an older age (6 to 10-month-old), the number of silent neurons was increased in plaque-bearing APP23xPS45 mice, and the proportion of hyperactive cells was 15 times higher than in wild-type (WT) animals. At this age, spatial and working memory was impaired in AD mice. Furthermore, at older ages, 10 to 14-month-old APP23xPS45 mice had more hyperactive neurons than WT [27]. Compared to 6 to 8-month-old mice, both WT and APP23xPS45 (10 to 14-month-old) mice had an increase in hyperactive neurons, but only APP23xPS45 mice had fewer silent neurons [27]. This suggests hyperactivity occurs in both healthy aging and AD but is more pronounced in the AD phenotype/model. An important observation from these studies is that the animal model used did not show the shift from hyper to hypoexcitability in the cortex that is reported in patient studies [8, 9]. Hence, the use of this particular animal model to study hyperexcitability in AD may be performed with caution and the results should be corroborated by other models.

Curiously, in the cortex, the hyperactive neurons were localized near the amyloid plaques (<60 µm), whereas silent and normal cells were randomly distributed [11]. This same pattern was observed in the hippocampus of APP23xPS45 mice [28], though with an earlier presentation of the hyperactivity. CA1 pyramidal neurons of the hippocampus had a greater number of hyperactive neurons compared to WT by 1 to 2 months of age. At an older age (6 to 8-month-old), AD mice had increased hyper- and hypoactive neurons. The number of hyperactive neurons, however, was significantly reduced compared to younger mice [28]. These results suggest that, in the hippocampus, hyperactivity starts before there is marked neuronal silencing and decreases with age, while hypoactivity increases over time, which is in line with clinical findings [8, 9]. These studies also indicate that functional neuronal deficits start earlier in the hippocampus and eventually progress to the cortex. It is relevant to note that most animal models used to study AD express more than one AD-related mutation, which differs from the clinic, as FAD patients harbor only single mutations. However, studies using animal models expressing a single AD-related mutation have also detected neuronal hyperexcitability [2931].

In the amygdala, analysis of single neurons by patch-clamp revealed contrasting results compared to the hippocampus and cortex. The flow of ions that depolarizes postsynaptic neurons and renders them more likely to fire an action potential is termed the excitatory postsynaptic current. Reduced spontaneous excitatory postsynaptic current frequency was found in 1 and 7-month-old mice bearing the human APOE4 gene, compared to mice with the human APOE3 gene, with no differences in inhibitory currents [32, 33]. Aged APOE4 mice (18 to 20-month-old mice) demonstrated an increased frequency of both excitatory and inhibitory transmission, with greater inhibition overall [34]. This was observed in the absence of Aβ plaque deposition or NFT formation.

Taken together, these results suggest that neuronal hyperactivity occurs in distinct brain regions of animal models of AD and manifests at different stages of the disease, depending on the area of the brain. These brain region differences might be explained by differences in neuronal vulnerability to excitability changes. Selective vulnerability is common in neurodegenerative diseases and may explain why some brain regions are affected earlier during the disease course, while others remain intact. Hence, some neuronal subtypes may be more susceptible to the different pathways underlying hyperexcitability, and therefore would be the first to show such alterations. As the disease progresses, this pathological change spreads to less vulnerable areas of the brain, while others may be unaffected. More studies are necessary to confirm the exact timeline in which hyperexcitability occurs and which brain regions are affected. Then, more efforts can be directed at identifying the precise neuronal populations that are most vulnerable and understanding what makes them vulnerable. Targeting proteins or receptors exclusively expressed in those neuronal subtypes would allow us to specifically target these neurons to reduce neuronal hyperexcitability.

Cell models of neuronal hyperexcitability

Induced pluripotent stem cells (iPSCs) are beneficial in the study of human neurological disorders that are caused by a complex range of genetic risk factors, including AD. One strength of iPSC disease modeling is that the cells maintain the genetic background of patients and can be used to identify how specific genes lead to pathology at the cellular and molecular levels. Pluripotent stem cells can be differentiated into neurons in two-dimensional cultures; however, the lack of Aβ aggregates due to regular cell media change is one of the disadvantages that has been documented by some studies in this model [35]. This issue may be addressed to a certain extent with 3D neuronal cultures/organoids, which can recapitulate amyloid-β aggregation, as well as NFT pathology [36].

A handful of studies have looked into neuronal excitability in LOAD using cell models. LOAD iPSC-derived neurons showed an increase in spontaneous Ca2+ signals, compared to controls that were silent in the absence of stimulus [13]. In another study, iPSC-derived neurons from a LOAD patient harboring the APOE4 allele displayed higher excitatory current frequencies compared to APOE3 isogenic control neurons [37]. A thorough study by Ghatak et al. found that iPSC-derived neurons bearing either PSEN1dE9, PSEN1M146V or APPswe mutations were hyperactive, compared to their isogenic controls [14]. Organoid models of AD, which displayed cortical layer formation and Aβ accumulation, also showed an enhanced extracellular action potential firing rate, detected with microelectrode arrays [14]. Another 3D cell model using neural stem cells genetically modified to express the K670N/M671L (Swedish) and V717I (London) mutations in the APP gene, found neurons from the AD group had increased spontaneous Ca2+ transients, compared to controls [38]. The proportion of hyperactive neurons increased over time in culture, with just over 15% of neurons showing hyperactivity at 3-weeks and more than 65% at 9-weeks of differentiation. Animal and human studies suggest that later stages of the disease lead to reduced neuronal activation [9, 39, 40]. Thus, it would be interesting to verify whether the number of hyperactive neurons would continue to increase or if it would decrease, in line with in vivo observations.

In summary, all 2D and 3D cell models of AD harboring different FAD mutations or derived from LOAD patients appear to show a similar hyperexcitable phenotype, corroborating findings from human and animal studies. These results suggest that hyperexcitability is detected across different systems used to model and study AD.

Potential causes of neuronal hyperexcitability

Hyperexcitability is observed in LOAD and FAD patients, as well as various laboratory models of the disease. Because AD is a multifactorial disease it is likely that several pathways contribute to hyperexcitability. In this section, we present the evidence pertaining to a range of factors that have been proposed to be involved in hyperexcitability: Ca2+ dyshomeostasis; glutamate and N-methyl-d-aspartate receptors (NMDAR); amyloid-β; tau; genetic risk factors for LOAD, glial cells, and inhibitory interneurons (Fig. 1).

Fig. 1. Mechanisms causing neuronal hyperexcitability in Alzheimer’s disease.

Fig. 1

Representation of a synapse in the healthy (left) and in the AD (right) brain. A Increased release of calcium from intracellular stores from pre and postsynaptic neurons results in higher levels of cytosolic calcium. B Enhanced glutamatergic signaling can be caused by reduced astrocytic uptake, reduced levels of glutamine synthetase (not shown in the Figure), and/or increased vGLUT expression. C Amyloid-β can form ionic pores in the plasma membrane. It also reduces the expression of Kv4 channels and increases NMDAR activation via increased d-serine and glutamate release and reduced glutamate uptake. D Protein tau can contribute to hyperexcitability by altering glutamate levels as well as the expression and function of Kv4.2 channels and NMDAR. E Compared to apoE3, apoE4 reduces the clearance and uptake of Aβ42 by astrocytes and microglia, respectively. F, G The release of pro-inflammatory cytokines, such as IL-1β and TNF-α, by glial cells can promote hyperexcitability. F During gliosis, reactive microglia fail to properly regulate neuronal excitability. G In AD, astrocytes show increased release and reduced uptake of glutamate, as well as reduced expression of potassium channel Kir4.1 (not shown in the Figure). Reactive astrocytes also increase neuronal excitability by reducing synaptic inhibition. H Reduced firing frequency and the number of inhibitory GABAergic neurons is another contributing factor to hyperexcitability in AD. AD Alzheimer’s disease, NMDAR NMDA receptor, vGLUT vesicular glutamate transporter.

Ca2+ dyshomeostasis

Mutations in the presenilin genes, PSEN1 and PSEN2, that are part of the γ-secretase complex, can cause FAD [5]. Dysregulation of intracellular calcium levels via mutations in these genes has been reported to contribute to AD pathology [41]. Apart from their role in the processing of amyloid precursor protein (APP), the presenilins also play a role in regulating Ca2+ levels in the cytoplasm by acting on different receptors in the endoplasmic reticulum (ER) membrane [4245]. Knocking out PSEN1 and PSEN2 or expressing a FAD mutation in mouse hippocampal neurons increased ER Ca2+ levels and consequently elevated Ca2+ release into the cytoplasm in the presence of a stimulus [42]. Presenilins have been suggested to function as Ca2+ leak channels in the ER, which become defective with PSEN mutations [43]. Alternative mechanisms of how presenilin disturbs Ca2+ signaling have also been proposed. For example, the sarcoendoplasmic reticulum Ca2+ ATPase (SERCA) pump maintains the low cytosolic levels of Ca2+ by pumping this cation from the cytosol into the ER. Knocking out both PSEN1 and PSEN2 genes in fibroblasts reduced the activity of SERCA and consequently, ER Ca2+ stores, and increased cytosolic Ca2+ [44]. Finally, the inositol trisphosphate receptor (InsP3R), an ER Ca2+ channel that releases Ca2+ into the cytoplasm, is also affected by presenilin-1 alterations. Mutations in PSEN1 enhanced InsP3R gating, elevating cytosolic Ca2+ levels [45]. Additionally, emptying presynaptic Ca2+ stores reduced amplitude and frequency of presynaptic Ca2+ transients and normalized neuronal network activity in both APPswe/PS1G384A and PS45 mice [27]. Since intracellular levels of Ca2+ mediate neurotransmitter release, elevated cytosolic Ca2+ can trigger excitatory neurotransmitter release and lead to exacerbated neuronal excitability.

Glutamate and NMDA receptors

Excessive glutamate is toxic for cells and has been associated with neurodegeneration [46]. As the main excitatory neurotransmitter in the central nervous system, glutamate must be maintained at appropriate levels to promote cell survival [47]. At high levels, glutamate causes excitotoxicity and neuronal death [48]. To prevent overstimulation of the receptors, after glutamate is released in the synapse, it is taken up by astrocytes through the excitatory amino acid transporters (EAAT). Reduced glutamate transport has been detected in the cortex and hippocampus of AD patients, which some studies reported to be due to a reduction in the expression of EAAT1 and EAAT2 [4951]. When taken up by astrocytes, glutamate is converted into glutamine by glutamine synthetase. Glutamine is then transferred back to the neurons where it is recycled into glutamate and transported into synaptic vesicles and becomes available to be released again. In the frontal and temporal cortices of postmortem brain tissue of AD patients, there is a reduction in the levels of glutamine synthetase, thus increasing the availability of glutamate [5254]. The release of glutamate in the synapse is dependent on the import of glutamate into synaptic vesicles by vesicular glutamate transporters (vGLUT). An iPSC-derived neuronal culture found increased expression of vGLUT1 in the AD group, which could lead to higher release probability [14]. Another study, however, found reduced levels of vGLUT in hippocampal neurons of APOE4 mice [55]. As for cerebrospinal fluid (CSF) levels of glutamate in AD patients, results so far have been controversial. While some authors have reported higher levels of glutamate [56, 57], others observed no changes [58] or even reductions in glutamate levels [59, 60]. Differences in the duration of illness, history of medication, and the inclusion or not of biomarkers as criteria for the diagnosis of AD could explain such variability in the results.

There are three types of ionotropic glutamatergic receptors and of these, NMDARs have been implicated in neuronal hyperexcitability in AD [27, 61] Application of  the activity-dependent NMDAR blocker MK-801 in AD mice reduced the frequency of neuronal firing in normal and hyperactive neurons, suggesting aberrant activation of this receptor is involved in hyperexcitability [27]. The use of the drug Memantine, which preferentially blocks extrasynaptic NMDAR, to treat moderate to severe AD in the clinic, supports the theory that synaptic NMDAR are required for neuronal survival, while extrasynaptic NMDAR mediate neurotoxicity [62]. Consistent with this hypothesis, enhancing the synaptic glutamate NMDAR subunit epsilon-1 (GluN2A) with the positive allosteric modulator, GNE-0723, reduced network hypersynchrony and epileptic discharge and ameliorated memory deficit in an AD mouse model [61]. The authors speculated this could be a consequence of activation of GluN2A expressing interneurons, thus suppressing excitatory neuronal activity. However, further investigation is required to fully understand the mechanisms behind these results. In conclusion, several alterations in glutamatergic transmission are observed in AD, thus modulation of specific NMDAR subunits poses a potential treatment strategy for hyperexcitability in AD.

Amyloid-β

Amyloid-β is one of the main components of amyloid plaques, insoluble extracellular deposits found in the brain of AD patients. The presence of amyloid plaques in postmortem brain tissue is one of the criteria for a definitive diagnosis of AD. Importantly, increased neuronal excitability is observed in neurons in the vicinity of Aβ plaques [11, 28] and a few reports have shown both soluble and fibrillar forms of Aβ can provoke changes in neuronal activity. Administration of Aβ dimers directly to neurons increased action potential firing, meanwhile, γ-secretase inhibitors, which reduced the levels of soluble Aβ, also restored neuronal activity to control levels in mice and in cell culture [14, 28]. In APPdE9 mice, administration of fibrillary but not oligomeric Aβ depolarized both dentate gyrus and layer 2/3 pyramidal neurons and enhanced their activity. The authors proposed the insertion of fibrillar Aβ in the membrane could lead to the disruption of voltage-gated ion channels responsible for maintaining neuronal membrane potential [15]. Supporting this theory, a Drosophila model expressing human Aβ42 found a role for Kv4 channels in Aβ-induced hyperexcitability. Neurons overexpressing Aβ42 were hyperexcitable and expressed less Kv4 protein, while Kv2 and Kv3 channels were unaffected. Increasing Kv4 channel levels in the animals not only restored normal neuronal excitability but also rescued learning impairment [63]. Aβ can also perturb Ca2+ homeostasis by stimulating voltage-gated calcium channels or directly by forming ion-permeable pores in the plasma membrane and increasing Ca2+ influx [64, 65]. Another possible mechanism through which Aβ could participate in neuronal hyperactivity is by modulating glutamatergic signaling. In cultured cells, Aβ42 augmented the release of the NMDAR co-agonist d-serine by microglia and glutamate from astrocytes [66, 67]. Soluble Aβ oligomers decreased the uptake of glutamate and increased extrasynaptic NMDAR activation, which is probably due to a spillover of excessive extracellular glutamate [6870]. Moreover, the application of soluble Aβ in the CA1 region of the hippocampus in WT mice increased neuronal activity by blocking glutamate reuptake [71]. Studies have also shown neuronal hyperactivity can promote Aβ accumulation in mice, therefore creating a vicious cycle that is detrimental to the cells [72, 73]. In sum, accumulation of Aβ affects numerous pathways that converge on a similar outcome, an increase in neuronal excitability.

Tau

Alzheimer’s disease can be classified as a tauopathy, a group of neurodegenerative diseases characterized by the aggregation of hyperphosphorylated tau proteins into neurofibrillary tangles in neurons [74]. Tau is a multifunctional protein, initially identified as a microtubule-binding protein. Mutations in the gene encoding tau (MAPT) can cause frontotemporal dementia, another tauopathy that is also associated with increased excitability of neuronal networks [75]. In mice and Drosophila models of epilepsy, reduction or ablation of endogenous tau expression improves cortical and hippocampal neuronal hyperexcitability and reduces the incidence and severity of seizures [7679]. In transgenic mice, overexpression of the human tau protein harboring the A152T mutation increased levels of extracellular glutamate and neuronal loss in vivo [80]. A mouse model of tauopathy, expressing human tau with a P301L mutation, displayed increased neuronal excitability in layer 3 of the cortex even before the formation of NFT, which could indicate a role of soluble hyperphosphorylated tau in driving this aberrant phenotype [81]. In the same mouse model it was also observed that while pyramidal neurons in the CA1 region of the hippocampus had increased firing, inhibitory interneurons were less active, suggesting a failure in inhibitory synaptic transmission [82]. In animal models of AD specifically, tau reduction was also protective against hyperexcitability. Transgenic hAPP mice expressing tau had heightened sensitivity to the γ-aminobutyric acid (GABA) receptor antagonist, pentylenetetrazole, and the glutamate receptor agonist, kainate, suffering from seizures at lower doses compared to control mice. Tau reduction blocked this effect and reversed memory deficits without changing Aβ levels [83]. More precisely, Roberson et al. [78] found that decreased tau levels reversed NMDAR dysfunction and the imbalance of inhibitory and excitatory currents that shifted towards excitation in hAPP-tau+/+ mice. Similar to Aβ, the release and propagation of tau is stimulated by neuronal activity, generating a toxic vicious cycle that could accelerate the progression of AD [84, 85]. Looking deeper into the pathways modulated by tau, Hall et al. [86] reported that knocking out tau in AD mice rescued dendritic hyperexcitability and Kv4.2 potassium channel depletion in the CA1 region of the hippocampus. Knocking out Kv4.2 channels in mice contributed to behavioral abnormalities and elevated epileptiform spiking. Interestingly, EEG recordings of frontal and parietal cortices of transgenic mice detected increased epileptiform spikes in hTau-A152T mice but decreased in hTau-WT mice compared to controls [87], indicating that overexpression of tau in animal models can lead to different outcomes, depending on the presence or absence of a disease-causing mutation.

Recent work indicates tau may also drive hypoexcitability. Transgenic mice expressing human tau with the P301L mutation had reduced neuronal activity compared to the control [28, 88]. Furthermore, the reduction in neuronal activity appears to be driven by soluble tau as this phenotype was observed in the absence of NFT [28, 89]. It is important to note, however, that mutations in the MAPT gene cause frontotemporal dementia but have not been linked to AD directly, therefore the interpretation of these results in the context of AD must be taken with caution. Using mouse models of AD, studies have shown that while Aβ pathology increases excitability, tau pathology drives hypoexcitability, and when both pathologies are combined, the effect of tau seems to predominate, leading to decreased neuronal firing [90, 91]. A proposed mechanism through which tau could lead to neuronal hypoexcitability involves the axon initial segment (AIS), where the initiation of the action potential occurs. Hatch et al. [88] found tau hyperphosphorylation destabilizes microtubules, shifting the AIS further away from the soma and consequently reducing neuronal firing frequency. Another interesting finding by the authors was that phosphorylation of different tau residues is commonly observed in AD (identified using the antibodies AT180 (T231), 12E8 (S262/S356), and PHF1 (S396/404), led to different outcomes. While AT180E and 12E8E hyperphosphorylation in primary hippocampal neurons caused the relocation of the AIS further from the soma, PHF1 had no effects on the location of the AIS.

Taken together, these results indicate the effects of tau can vary depending on phosphorylation, mutation, and brain region, thus deeper investigation into the role of tau in mediating neuronal excitability changes is required to understand its impact on AD.

APOE4 and genetic risk factors for LOAD

Sporadic AD or LOAD comprises more than 98% of all cases of AD [6]. Whilst LOAD is not caused by a mutation in a single gene, there is a significant genetic component, with multiple genetic variants affecting the risk of developing LOAD. The APOE ε4 allele is the major genetic risk factor for LOAD, elevating the risk of developing the disease 3 to 10-fold, compared to the general population [92]. The human APOE gene encodes three isoforms of the apolipoprotein E (apoE): APOE ε2, APOE ε3, and APOE ε4. The protein apoE is involved in the transport of cholesterol and other lipids in the periphery and in the brain [93]. Increasing evidence shows hyperexcitability in humans and animal models harboring the APOE4 allele, indicating this phenotype is not limited to the familial form of the disease. In humans, young and healthy individuals carrying the APOE4 allele had greater hippocampal activation at rest and during a memory-encoding task relative to non-APOE4 carriers. This was observed along with a lack of difference in brain perfusion or whole-brain, gray matter, white matter, cerebrospinal fluid (CSF) or hippocampal volumes [19]. Increased brain activity was also observed in the prefrontal, temporal and parietal cortices of older, cognitively-intact APOE4 carriers [18, 9496]. Mice expressing human APOE4 developed a seizure phenotype that is absent (or less pronounced) in mice expressing human APOE2 or APOE3. APOE4 mice showed increased hippocampal and cortical activity and were also more sensitive to the GABA receptor antagonist, pentylenetetrazole. Interestingly, these alterations were observed in the absence of Aβ plaque formation, with no significant changes in the levels of Aβ42, although Aβ40 was increased in APOE4 animals [97]. Apart from enhanced excitability, APOE4 mice have impaired short-term synaptic plasticity [98]. Moreover, Apoe knock-out mice displayed reduced α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA)/NMDA receptor ratios, while mice lacking Apoe expression only in the brain displayed similar NMDAR and AMPAR currents to APOE3 mice. This suggests plasma apoE plays an important role in synaptic function [99].

Studies have identified defective dendritic morphology in APOE4 mice, including reduced dendritic length, branch number, spine size, and spine density [98, 100, 101]. This could be explained by a deficiency in the ability of the apoE4 isoform to deliver lipids, such as cholesterol. Cholesterol and other lipids are important for neuronal morphology, synapse formation, and ion channel function, thus alterations in their levels can dramatically affect the excitability of neurons [33]. In culture, APOE4 human iPSC-derived neurons are more excitable than APOE3 isogenic controls, which could be due to an increase in the expression of synaptic proteins, including synaptophysin and PSD-95, and an upregulation of genes involved in neuronal differentiation [37]. The APOE genotype also has an effect on the levels of Aβ. Compared to APOE3-expressing cells, iPSC-derived astrocytes and microglia carrying the APOE4 allele have reduced clearance and slower uptake of Aβ42, respectively [37]. However, Konttinen et al. [102] found there was no significant difference in Aβ42 uptake in APOE4 iPSC microglia, compared to APOE3 isogenic control lines. In regards to the difference between APOE2 and APOE3 astrocytes, Brookhouser et al. [103] observed that the effect of the APOE genotype in Aβ uptake in iPSC-derived astrocytes varies depending on the AD-related mutation harbored by the cell. Thus, further research is required to understand neuronal excitability changes and the complex interaction between microglia, astrocytes and neurons under different APOE genotypes. This could be accomplished by coculturing these cell types with different combinations of APOE alleles.

Neuroinflammation is a common feature of AD [104], and genome-wide association studies (GWAS) have identified several genes involved in the immune response to be associated with AD [105, 106]. Some of these genes are CR1, CD33, MS4A, TREM2, CLU, ABCA7, EPHA1, and BIN1. In the context of the immune response, BIN1 knock-out mice display a higher incidence of inflammation [107]; however, this gene is also involved in endocytosis, calcium homeostasis, and apoptosis [106]. BIN1 encodes the bridging integrator 1 (BIN1), myc box-dependent-interacting protein, which is expressed throughout the body with higher levels in muscle and brain tissue [108]. Little is known about BIN1 function in the nervous system, however, it has recently been implicated in neuronal hyperexcitability. In rat hippocampal neuronal cultures, overexpression of BIN1 increased calcium influx and action potential firing frequency due to an increase in the interaction between BIN1 and L-type voltage-gated calcium channels. The interaction between these proteins is tau dependent, since reducing tau expression inhibited their interaction and consequently, reversed this hyperexcitable phenotype [109].

In summary, numerous studies support the role of APOE4 in hyperexcitability and more recent data has revealed BIN1 to be another potential contributing factor to AD neuronal excitability changes. GWAS offers the first step in elucidating the genetic risk factors for developing AD. From there, candidate genes can be manipulated in laboratory models to clarify which pathways are involved in neuronal hyperexcitability.

Glial cells and neuroinflammation

Glia are non-neuronal cells of the central nervous system, including astrocytes, microglia, and oligodendrocytes. The latter are responsible for the formation of myelin sheaths around axons [110], and even though oligodendrocyte abnormalities have been reported in AD (see refs. [111, 112] for more on oligodendrocytes in AD), in this review we will focus on the role of astrocytes and microglia. Astrocytes and microglia are activated by, and respond to, pathogens and stressors by releasing pro-inflammatory molecules to promote repair; however, chronic activation of glial cells and inflammation can promote seizures [113], which are more prevalent in AD patients than in healthy elderly people [21]. Injection of the pro-inflammatory cytokine, interleukin-1 β (IL-1β), exacerbates seizures [114], while blockade of the IL-1β receptor reduces seizure susceptibility in mice [115, 116]. Inhibition of IL-1β synthesis also shows anticonvulsive effects [117]. The pro-convulsive effect of IL-1β is reversed by treatment with an NR2B receptor antagonist, indicating the involvement of the NMDAR in this pathway [118]. The pro-inflammatory cytokine, tumor necrosis factor-α (TNF-α), regulates glutamatergic signaling in both neurons and astrocytes. Exposure of cultured hippocampal neurons to TNF-α increased the expression of AMPA receptors in the plasma membrane, which resulted in a higher frequency of miniature excitatory postsynaptic currents [119]. In the dentate gyrus of mice, TNF-α is required for glutamate release from astrocytes, causing an increase in excitatory neuronal activity via activation of NMDA receptors [120]. Reduction of neuroinflammation, including anti-TNF-α therapy, has been shown to protect against epilepsy in patients and animals [121, 122]. Therefore, this approach to reduce neuroinflammation and consequently hyperexcitability could also be relevant in the context of AD.

Microglia

Genes involved in the immune response have been linked to a higher risk of developing AD [106]. Many of these genes are highly expressed in microglia, such as ABCA7, CD33, and TREM2. TREM2 encodes triggering receptors expressed on myeloid cells 2 (TREM2), a protein that stimulates microglial phagocytosis [123]. Deletion of TREM2 enhances tau and amyloid-β pathology [124, 125], suggesting a direct role for microglia in containing the spread of abnormal protein aggregation in AD. Microglia are the brain’s resident macrophages, playing an important function in brain injury and inflammation, as well as synaptic pruning [126]. Apart from these functions, microglia can also sense and modulate neuronal activity [127]. Similar to neurons, microglia residing close to amyloid plaques show increased Ca2+ transients [128], which can occur as a response to neuronal hyperactivity [129]. In zebrafish, resting microglia sense neuronal activity and extend processes towards these neurons. This interaction, in turn, reduces both spontaneous and evoked neuronal activity [130]. A similar outcome is observed in mice, where the increased neuronal firing was followed by a greater number of microglial extensions contacting the active neurons, and pharmacological inhibition of microglia exacerbated neuronal responses to neurostimulants [131]. Microglial depletion also aggravated seizure severity in mice [132]. However, microglia-neuronal contact can also result in increased synaptic activity and network synchronization [133]. In mice, Ca2+ responses of motor cortical neurons were elevated in dendritic spines in close contact with microglia. This enhanced response was reduced after the retraction of microglial processes. Ablation of microglia reduced the synchronous firing of neurons located close to each other. These contrasting results probably highlight the function of microglia as a regulator of neuronal activity. Microglia can sense neuronal hypo- and hyperexcitability [127] and can promote either an increase or decrease in neuronal firing, which indicates these cells may be critical for maintaining homeostatic network activity in the brain. Interestingly, microglial activation with lipopolysaccharide (LPS) treatment abolished the regulation of neuronal activity, suggesting brain inflammation can have a detrimental effect in microglial modulation of neuronal circuits [133].

Astrocytes

Astrocytes exert a range of functions in the brain, including nutrient supply to neurons, maintenance of the blood–brain barrier, and regulation of the extracellular ionic environment [134]. Astrocytes regulate ion homeostasis by clearing extracellular K+ during neuronal repolarization [135]. Defective K+ buffering by astrocytes increases extracellular K+ levels and is associated with the initiation of seizures [136]. Protein expression of the astrocytic potassium channel Kir4.1 is reduced in a mouse model of AD and in the brain of AD patients [137], which could impact the resting membrane potential and neuronal excitability. In addition to K+, astrocytes also regulate extracellular glutamate levels by taking up this neurotransmitter from the synapse via EAAT1 and EAAT2 receptors, which have reduced expression in the AD brain [4951]. Elevated glutamate release by astrocytes was also found in a mouse model of AD [138]. Astrocytic dysregulation of glutamate levels contributes to neuronal hyperexcitability and therefore, represents a potential target in re-establishing physiological neuronal activity. Under pathological conditions, such as AD, astrocytes go through molecular and functional changes in a process known as astrogliosis or reactive astrogliosis [139]. The selective induction of astrogliosis in mice decreased synaptic inhibition of CA1 pyramidal neurons, as well as expression of glutamine synthetase [140]. At the cellular level, inhibitory postsynaptic currents were smaller in neurons neighboring reactive astrocytes, while excitatory postsynaptic currents were unaltered. These changes resulted in hyperexcitability of the neuronal network. The application of glutamine reversed these aberrant phenotypes, suggesting reactive astrocytes can impair network excitability by depleting glutamine available for neuronal uptake and reducing the availability of GABA for release by inhibitory neurons.

Inhibitory interneurons

Inhibitory interneurons regulate neuronal network oscillations, such as gamma waves, which are required for cognitive functions [141]. Increased gamma oscillations during memory-encoding tasks are predictive of effective memory formation and are associated with reduced epileptiform discharges [142]. A loss of inhibitory neuron function has been associated with hyperexcitability in AD. Reduced inhibitory tone, probably due to decreased responsiveness to GABAergic stimuli, lower numbers of GABAergic interneurons, and reduced GABAergic signaling, has been reported in APOE4 mice [100, 143]. Neuronal cultures with either PSEN1M146V or APPswe mutation, which were hyperactive, presented reduced frequency of inhibitory postsynaptic currents, as well as decreased staining for GABA and the vesicular GABA transporter [14]. Inhibitory neuron deficiency has been reported in the cortex and hippocampus of animal models of AD. Treating APP23xPS45 mice with the GABAA receptor antagonist, gabazine increased the firing frequency of hyperactive, silent, and normal neurons of cortical layer 2/3 to a similar level. The increase was, however, significantly smaller for hyperactive neurons than for normal cells [11]. EEG recordings of hAPPFAD mice showed increased spontaneous epileptiform discharges during low-intensity gamma oscillations, which were reduced in hAPPFAD animals, compared to controls [144]. Since gamma waves are generated by the activity of parvalbumin (PV) positive GABAergic interneurons, the authors tested if abnormalities in PV cells could lead to network hypersynchrony [144]. Patch-clamp confirmed reduced inhibitory postsynaptic current frequency in hAPPFAD mice. These animals also had reduced expression of the voltage-gated sodium channel Nav1.1 in the parietal cortex and PV cells had reduced mRNA and protein expression of Nav1.1. Overexpressing Nav1.1 in PV cells prevented inhibitory postsynaptic current abnormalities and reduced epileptiform discharges by decreasing the number of spikes during low-intensity gamma oscillations. Impaired Nav1.1 appears to contribute to cognitive decline, as increasing its expression improved spatial learning and memory in mice [144]. Hamm et al. [145] found similar results in the hippocampus of TgCRND8 mice, transgenic mice containing human APP with both K670N/M671L (Swedish) and V717F (Indiana) mutations. The animals exhibited fewer PV interneurons in the CA1, as well as depressed gamma oscillations. A decreased expression of Nav1.6 and a tendency towards a reduction for Nav1.1 was also detected. These aberrations occurred concomitantly with impaired memory performance. Another study has further corroborated the role of Nav1.1 in impaired synaptic inhibition. Nav1.1 expressing interneuron transplant in the hippocampus and cortex of hAPPFAD mice reduced epileptic spike frequency and enhanced gamma oscillatory activity. Transplantation of Nav1.1 interneurons also reversed learning deficits [146]. To sum up, these studies show a clear contribution of Nav1.1 and impaired synaptic inhibition to elevated neuronal excitability in AD. Targeting this channel in future studies has the potential for developing new treatments for AD. Even though transplantation studies have shown beneficial effects, it could still take years for this approach to be translated into the clinic. A potentially easier solution would be enhancing Nav1.1 activity with selective drugs, a strategy that has been shown effective in reducing seizure activity in animal models of Dravet syndrome epilepsy [147, 148]. The effect of these compounds in reducing hyperexcitability in AD has yet to be investigated.

Is hyperexcitability neuroprotective or neurodegenerative?

Synaptic deficits, neurodegeneration, and brain atrophy are common features of AD [149]. Because of this, it has been suggested that neuronal hyperexcitability could be an attempt to compensate for the neuronal loss or synaptic deficits by recruiting greater neuronal resources, to maintain the same level of performance. However, neuronal hyperactivation has been documented in individuals with MCI who still present poorer memory performance compared to controls [10]. Also, previous studies have found enhanced brain activity in individuals at risk for developing AD, even with unchanged hippocampal volume [19]. This hypothesis is further refuted by evidence that antiepileptic drugs decrease hippocampal activation and improve memory performance in individuals with MCI, as well as in animal models of AD [150, 151]. Thus, increased brain activity is more likely a driver of AD pathogenesis rather than a compensatory effect. In fact, augmented brain activity can promote AD pathology. Increased neuronal activity not only causes increases in Aβ levels [72, 73] but also stimulates the release of tau in vivo and in vitro, leading to the spread of tau pathology, which contributes to cognitive deficits [84, 85]. Besides, reducing neuronal excitability decreases Aβ deposition and synaptic loss and prevents the spread of Aβ plaques in mice [16, 152].

A longitudinal study recording hippocampal activity in older individuals without dementia found people with the most pronounced hyperactivation at the first assessment suffered the greatest hypoactivation two years later and the fastest cognitive decline [39]. AD patients with subclinical epileptiform activity had faster cognitive decline than those who did not present such abnormalities [153]. Therefore, network hyperexcitability may be predictive of imminent hypoactivity and cognitive impairment [39]. A similar pattern of neuronal excitability dysregulation is observed in another neurodegenerative disease, amyotrophic lateral sclerosis (ALS; reviewed in ref. [154]). Motor neuron hyperexcitability is a common pathology observed in both sporadic and familial cases of ALS. Hyperexcitability occurs in the early stages of ALS, even prior to motor symptom onset, and then progresses to hypoactivity [154]. It has been shown that the neurons most vulnerable to degeneration in ALS are also the ones that receive greater excitatory input, compared to the resistant neuronal subtypes, and eliminating this input protects against degeneration [155]. Whether the same process occurs in AD has yet to be identified.

Together, current evidence suggests that hyperactivation occurs in the early stages of AD, even before there is significant neuronal loss, contradicting the hypothesis that overexcitation is a compensatory effect. On the contrary, neuronal hyperactivity appears to be detrimental by promoting AD pathology and possibly, cognitive decline. Thus, excessive neuronal excitability could be a potential target for therapeutic intervention.

Conclusions and future directions

Emerging evidence shows that abnormally elevated neuronal activity is a common functional feature of AD that is associated with greater cognitive decline. There are probably a few reasons for this: increased neuronal activity stimulates the spread of amyloid and tau pathology [72, 73, 84, 85], and in the long term it can lead to excitotoxicity and cell death [48]. A more sophisticated explanation may be that neuronal networks require fine regulation to function appropriately, as changes in the firing properties of a single cortical neuron are sufficient to elicit changes in behavior [156, 157]. Neuronal hyperactivity can also render neurons unable to encode spatial information or discriminate visual stimuli [158, 159]. Considering AD is a multifactorial disease, it is likely that several factors, rather than a single cause, act in conjunction to disrupt neuronal excitability regulation. In this review, we discussed some of these factors: excessive levels of cytoplasmic Ca2+ and glutamate; proteins commonly associated with AD, such as Aβ, apoE4, and tau; and the dysregulation of inhibitory interneuron and glial cell function.

Many questions remain to be answered about hyperexcitability in AD. Clinical fMRI studies demonstrate specific brain areas are hyperactive in patients with MCI, as well as asymptomatic offspring of AD patients and at-risk individuals, while AD patients show brain hypoactivity. However, the exact timeline for these events is unclear. When does hyperactivity start, and at which point does it begin to decrease? Would medical intervention be beneficial and if so at what point of disease progression or when would it be too late? Other studies highlight neuronal hyperexcitability as a predictor of subsequent hypoexcitability, but the mechanisms causing this shift in neuronal activity are not fully characterized. How does hyperexcitability eventually lead to reduced neuronal activity? To answer these questions we need to consider which models should be used to investigate this phenotype. Ideally, we want a model that is able to recapitulate what we see in human studies: a shift from hyper- to hypoexcitability in specific brain areas. Since all models have advantages and disadvantages, there is strength in combining information from human studies with animal and cell models of AD.

Human studies have been crucial to confirm that hyperactivity is not an artifact of laboratory models but also occurs in patients. Yet, clinical studies are observational and limited in the mechanistic information that can be tested in living patients. A vast amount of information has come from animal studies, which permit the manipulation of variables and the investigation of effects on an entire organism. This model, however, is disadvantaged by species differences. Recent data has been provided via human iPSCs, which give us the opportunity to study cells from AD patients, and to understand how specific mutations or allele variants affect cellular function. The downside is that neuronal cultures do not recapitulate the entire brain and its complex networks. Compared to organoids, 2D culture systems are an easier and faster method to study the mechanisms governing hyperexcitability. They are also well-established, meaning there is more data available for comparison. However, they lack the cytoarchitecture of the brain, and studies have shown that cells isolated from the brain exhibit significant changes in gene expression when in two-dimensional cell cultures [160, 161]. This disadvantage can be overcome with 3D cultures/organoids [162], which allow more complex interactions between cells and the interaction between cells and the extracellular matrix. They also provide a better spatial organization of the cells [105, 163]. Nonetheless, this model still presents drawbacks. The cultures are labor-intensive, more costly, and can show high variability between batches. It can also take months for the neurons within the organoids to mature [164]. Overcoming these difficulties is therefore critical for future studies to continue providing reliable data on neuronal hyperexcitability.

The use of fMRI to detect elevated brain activity may help to identify individuals at risk before symptom onset. However, because hyperexcitability is not exclusive to AD, but observed in other neurodegenerative diseases, such as ALS [154], its use as a biomarker for AD should be done in combination with other diagnostic methods currently available. As we discussed in this review, other cell types, such as inhibitory neurons and glial cells, influence neuronal excitability. Coculture systems of iPSC-derived neurons and glia combining healthy and diseased cells can give us more information on the role of each cell type [165]. In both cells and animals, gene-editing techniques, such as CRISPR-Cas9, can be used to study the role of specific cell subtypes or specific proteins and receptors. Another approach to investigate which channels are dysfunctional in the AD brain is microtransplanting native receptors from human postmortem tissue into frog oocytes [166168]. This technique enables the study of the electrophysiological properties of disease-related receptors and can be used to validate the efficiency of novel drugs in human receptors. A deeper understanding of the specific cell subtypes and channels altered in AD will permit the development of drugs with minimal off-target effects.

Ultimately, understanding the circuits that are affected in AD, as well as their composite cell types, will be important for pinpointing the changes in neuronal excitability and how they drive or interact with pathology. To achieve this, the continuation of clinical and preclinical studies is essential to gather the information that can be translated into novel diagnostic and/or treatment strategies for AD.

Acknowledgements

HTDA is supported by the International Postgraduate Tuition Award from the University of Wollongong. LO is supported by a National Health and Medical Research Council (NHMRC) of Australia Boosting Dementia Research Leadership Fellowship (APP1135720). NM was supported by the Al and Val Rosenstrauss Fellowship from the Rebecca L. Cooper Medical Research Foundation (#F2021971), an NHMRC Early Career Fellowship (APP1105445), and grants from the Brain Behavior Research Foundation (NARSAD Young Investigator Grant, #26486) and the Rebecca L. Cooper Medical Research Foundation (#PG2020645).

Author contributions

HTDA: initial draft manuscript writing, Figure and Table construction, manuscript editing, and final approval of the manuscript; NM: manuscript editing, supervision and final approval of the manuscript; LO: study conception, funding, manuscript editing, supervision and final approval of the manuscript.

Competing interests

The authors declare no competing interests.

Footnotes

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

References

  • 1.Prince M, Wimo A, Guerchet M, Gemma-Claire Ali M, Wu Y-T, Prina M, et al. World Alzheimer Report 2015–The global impact of dementia: an analysis of prevalence, incidence, cost and trends. London: Alzheimer's Disease International; 2015.
  • 2.Alzheimer’s Association. 2019 Alzheimer’s disease facts and figures. Includes a special report on Alzheimer’s detection in the primary care setting: connecting patients and physicians. Alzheimer’s Dement J Alzheimer’s Assoc. 2019;15:321–87. doi: 10.1016/j.jalz.2019.01.010. [DOI] [Google Scholar]
  • 3.Glenner GG, Wong CW. Alzheimer’s disease: Initial report of the purification and characterization of a novel cerebrovascular amyloid protein. Biochem Biophys Res Commun. 1984;120:885–90. doi: 10.1016/S0006-291X(84)80190-4. [DOI] [PubMed] [Google Scholar]
  • 4.Brion JP, Couck AM, Passareiro E, Flament-Durand J. Neurofibrillary tangles of Alzheimer’s disease: An immunohistochemical study. J Submicrosc Cytol. 1985;17:89–96. [PubMed] [Google Scholar]
  • 5.Blacker D, Tanzi RE. The genetics of Alzheimer disease: current status and future prospects. Arch Neurol. 1998;55:294–6. doi: 10.1001/archneur.55.3.294. [DOI] [PubMed] [Google Scholar]
  • 6.Bird TD. Alzheimer disease overview. GeneReviews; 1998.
  • 7.Saunders AM, Strittmatter WJ, Schmechel D, St. George-Hyslop PH, Pericak-Vance MA, Joo SH, et al. Association of apolipoprotein E allele ϵ4 with late-onset familial and sporadic alzheimer’s disease. Neurology. 1993;43:1467–72. doi: 10.1212/WNL.43.8.1467. [DOI] [PubMed] [Google Scholar]
  • 8.Celone KA, Calhoun VD, Dickerson BC, Atri A, Chua EF, Miller SL, et al. Alterations in memory networks in mild cognitive impairment and Alzheimer’s disease: An independent component analysis. J Neurosci. 2006;26:10222–31. doi: 10.1523/JNEUROSCI.2250-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Dickerson BC, Salat DH, Greve DN, Chua EF, Rand-Giovannetti E, Rentz DM, et al. Increased hippocampal activation in mild cognitive impairment compared to normal aging and AD. Neurology. 2005;65:404–11. doi: 10.1212/01.wnl.0000171450.97464.49. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Hämäläinen A, Pihlajamäki M, Tanila H, Hänninen T, Niskanen E, Tervo S, et al. Increased fMRI responses during encoding in mild cognitive impairment. Neurobiol Aging. 2007;28:1889–903. doi: 10.1016/j.neurobiolaging.2006.08.008. [DOI] [PubMed] [Google Scholar]
  • 11.Busche MA, Eichhoff G, Adelsberger H, Abramowski D, Wiederhold KH, Haass C, et al. Clusters of hyperactive neurons near amyloid plaques in a mouse model of Alzheimer’s disease. Science. 2008;321:1686–9. doi: 10.1126/science.1162844. [DOI] [PubMed] [Google Scholar]
  • 12.Šišková Z, Justus D, Kaneko H, Friedrichs D, Henneberg N, Beutel T, et al. Dendritic structural degeneration is functionally linked to cellular hyperexcitability in a mouse model of alzheimer’s disease. Neuron. 2014;84:1023–33. doi: 10.1016/j.neuron.2014.10.024. [DOI] [PubMed] [Google Scholar]
  • 13.Balez R, Steiner N, Engel M, Muñoz SS, Lum JS, Wu Y, et al. Neuroprotective effects of apigenin against inflammation, neuronal excitability and apoptosis in an induced pluripotent stem cell model of Alzheimer’s disease. Sci Rep. 2016;6:1–16.. doi: 10.1038/srep31450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Ghatak S, Dolatabadi N, Trudler D, Zhang X, Wu Y, Mohata M, et al. Mechanisms of hyperexcitability in alzheimer’s disease hiPSC-derived neurons and cerebral organoids vs. Isogenic control. Elife. 2019;8:1–22.. doi: 10.7554/eLife.50333. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Minkeviciene R, Rheims S, Dobszay MB, Zilberter M, Hartikainen J, Fülöp L, et al. Amyloid β-induced neuronal hyperexcitability triggers progressive epilepsy. J Neurosci. 2009;29:3453–62. doi: 10.1523/JNEUROSCI.5215-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Gail Canter R, Huang WC, Choi H, Wang J, Ashley Watson L, Yao CG, et al. 3D mapping reveals network-specific amyloid progression and subcortical susceptibility in mice. Commun Biol. 2019;2:1–12.. doi: 10.1038/s42003-019-0599-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Bassett SS, Yousem DM, Cristinzio C, Kusevic I, Yassa MA, Caffo BS, et al. Familial risk for Alzheimer’s disease alters fMRI activation patterns. Brain. 2006;129:1229–39. doi: 10.1093/brain/awl089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Bookheimer SY, Strojwas MH, Cohen MS, Saunders AM, Pericak-Vance MA, Mazziotta JC, et al. Patterns of brain activation in people at risk for Alzheimer’s disease. N. Engl J Med. 2000;343:450–6. doi: 10.1056/NEJM200008173430701. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Filippini N, MacIntosh BJ, Hough MG, Goodwin GM, Frisoni GB, Smith SM, et al. Distinct patterns of brain activity in young carriers of the APOE-ε4 allele. Proc Natl Acad Sci USA. 2009;106:7209–14. doi: 10.1073/pnas.0811879106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Petersen RC. Mild cognitive impairment. Contin Lifelong Learn Neurol. 2016;22:404–18. doi: 10.1212/CON.0000000000000313. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Miranda DDC, Brucki SMD. Epilepsia em pacientes com doença de alzheimer: Uma revisão sistemática. Dement e Neuropsychol. 2014;8:66–71. doi: 10.1590/S1980-57642014DN81000010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Amatniek JC, Hauser WA, DelCastillo-Castaneda C, Jacobs DM, Marder K, Bell K, et al. Incidence and predictors of seizures in patients with Alzheimer’s disease. Epilepsia. 2006;47:867–72. doi: 10.1111/j.1528-1167.2006.00554.x. [DOI] [PubMed] [Google Scholar]
  • 23.Mendez MF, Catanzaro P, Doss RC, Arguello R, Frey WH. Seizures in Alzheimer’s disease: clinicopathologic study. J Geriatr Psychiatry Neurol. 1994;7:230–3. doi: 10.1177/089198879400700407. [DOI] [PubMed] [Google Scholar]
  • 24.Hall AM, Roberson ED. Mouse models of Alzheimer’s disease. Brain Res Bull. 2012;88:3–12. doi: 10.1016/j.brainresbull.2011.11.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Drummond E, Wisniewski T. Alzheimer’s disease: experimental models and reality. Acta Neuropathologica. 2017;133:155–75. doi: 10.1007/s00401-016-1662-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Palop JJ, Chin J, Roberson ED, Wang J, Thwin MT, Bien-Ly N, et al. Aberrant excitatory neuronal activity and compensatory remodeling of inhibitory hippocampal circuits in mouse models of Alzheimer’s disease. Neuron. 2007;55:697–711. doi: 10.1016/j.neuron.2007.07.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Lerdkrai C, Asavapanumas N, Brawek B, Kovalchuk Y, Mojtahedi N, Del Moral MO, et al. Intracellular Ca2+ stores control in vivo neuronal hyperactivity in a mouse model of Alzheimer’s disease. Proc Natl Acad Sci USA. 2018;115:E1279–88.. doi: 10.1073/pnas.1714409115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Busche MA, Chen X, Henning HA, Reichwald J, Staufenbiel M, Sakmann B, et al. Critical role of soluble amyloid- for early hippocampal hyperactivity in a mouse model of Alzheimer’s disease. Proc Natl Acad Sci USA. 2012;109:8740–5. doi: 10.1073/pnas.1206171109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Kam K, Duffy ÁM, Moretto J, LaFrancois JJ, Scharfman HE. Interictal spikes during sleep are an early defect in the Tg2576 mouse model of β-amyloid neuropathology. Sci Rep. 2016;6:1–16.. doi: 10.1038/srep20119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Bezzina C, Verret L, Juan C, Remaud J, Halley H, Rampon C, et al. Early onset of hypersynchronous network activity and expression of a marker of chronic seizures in the Tg2576 mouse model of Alzheimer’s disease. PLoS ONE. 2015;10:e0119910. [DOI] [PMC free article] [PubMed]
  • 31.Ziyatdinova S, Rönnbäck A, Gurevicius K, Miszczuk D, Graff C, Winblad B, et al. Increased epileptiform EEG activity and decreased seizure threshold in arctic APP transgenic mouse model of Alzheimer’s disease. Curr Alzheimer Res. 2016;13:817–30. doi: 10.2174/1567205013666160129095508. [DOI] [PubMed] [Google Scholar]
  • 32.Klein RC, Mace BE, Moore SD, Sullivan PM. Progressive loss of synaptic integrity in human apolipoprotein E4 targeted replacement mice and attenuation by apolipoprotein E2. Neuroscience. 2010;171:1265–72. doi: 10.1016/j.neuroscience.2010.10.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Wang C, Wilson WA, Moore SD, Mace BE, Maeda N, Schmechel DE, et al. Human apoE4-targeted replacement mice display synaptic deficits in the absence of neuropathology. Neurobiol Dis. 2005;18:390–8. doi: 10.1016/j.nbd.2004.10.013. [DOI] [PubMed] [Google Scholar]
  • 34.Klein RC, Acheson SK, Mace BE, Sullivan PM, Moore SD. Altered neurotransmission in the lateral amygdala in aged human apoE4 targeted replacement mice. Neurobiol Aging. 2014;35:2046–52. doi: 10.1016/j.neurobiolaging.2014.02.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Young JE, Goldstein LSB. Alzheimer’s disease in a dish: promises and challenges of human stem cell models. Hum Mol Genet. 2012;21:R82. doi: 10.1093/hmg/dds319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Choi SH, Kim YH, Hebisch M, Sliwinski C, Lee S, D’Avanzo C, et al. A three-dimensional human neural cell culture model of Alzheimer’s disease. Nature. 2014;515:274–8. doi: 10.1038/nature13800. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Lin YT, Seo J, Gao F, Feldman HM, Wen HL, Penney J, et al. APOE4 causes widespread molecular and cellular alterations associated with Alzheimer’s disease phenotypes in human iPSC-derived brain cell types. Neuron. 2018;98:1141–54.e7. doi: 10.1016/j.neuron.2018.05.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Park J, Wetzel I, Marriott I, Dréau D, D’Avanzo C, Kim DY, et al. A 3D human triculture system modeling neurodegeneration and neuroinflammation in Alzheimer’s disease. Nat Neurosci. 2018;21:941–51. doi: 10.1038/s41593-018-0175-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.O’Brien JL, O’Keefe BKM, Laviolette PS, Deluca AN, Blacker BD, Dickerson SBC, et al. Longitudinal fMRI in elderly reveals loss of hippocampal activation with clinical decline. Neurology. 2010;74:1969–76. doi: 10.1212/WNL.0b013e3181e3966e. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Busche MA, Konnerth A. Neuronal hyperactivity-A key defect in Alzheimer’s disease? Bioessays. 2015;37:624–32.. doi: 10.1002/bies.201500004. [DOI] [PubMed] [Google Scholar]
  • 41.Bezprozvanny I, Mattson MP. Neuronal calcium mishandling and the pathogenesis of Alzheimer’s disease. Trends Neurosci. 2008;31:454–63. doi: 10.1016/j.tins.2008.06.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Zhang H, Sun S, Herreman A, De Strooper B, Bezprozvanny I. Neurobiology of disease role of presenilins in neuronal calcium homeostasis. J Neurosci. 2010;30:8566–80. doi: 10.1523/JNEUROSCI.1554-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Tu H, Nelson O, Bezprozvanny A, Wang Z, Lee SF, Hao YH, et al. Presenilins form ER Ca2+ leak channels, a function disrupted by familial Alzheimer’s disease-linked mutations. Cell. 2006;126:981–93. doi: 10.1016/j.cell.2006.06.059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Green KN, Demuro A, Akbari Y, Hitt BD, Smith IF, Parker I, et al. SERCA pump activity is physiologically regulated by presenilin and regulates amyloid β production. J Cell Biol. 2008;181:1107–16. doi: 10.1083/jcb.200706171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Cheung K-H, Shineman D, Müller M, Cárdenas C, Mei L, Yang J, et al. Mechanism of Ca2+ disruption in Alzheimer’s disease by presenilin regulation of InsP3 receptor channel gating. Neuron. 2008;58:871–83. doi: 10.1016/j.neuron.2008.04.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Choi DW. Glutamate neurotoxicity and diseases of the nervous system. Neuron. 1988;1:623–34. doi: 10.1016/0896-6273(88)90162-6. [DOI] [PubMed] [Google Scholar]
  • 47.Jansson LC, Åkerman KE. The role of glutamate and its receptors in the proliferation, migration, differentiation and survival of neural progenitor cells. J Neural Transm 2014 1218. 2014;121:819–36. doi: 10.1007/s00702-014-1174-6. [DOI] [PubMed] [Google Scholar]
  • 48.Olney JW. Brain lesions, obesity, and other disturbances in mice treated with monosodium glutamate. Science. 1969;164:719–21. doi: 10.1126/science.164.3880.719. [DOI] [PubMed] [Google Scholar]
  • 49.Jacob CP, Koutsilieri E, Bartl J, Neuen-Jacob E, Arzberger T, Zander N, et al. Alterations in expression of glutamatergic transporters and receptors in sporadic Alzheimer’s disease. J Alzheimer’s Dis. 2007;11:97–116. doi: 10.3233/JAD-2007-11113. [DOI] [PubMed] [Google Scholar]
  • 50.Scott HA, Gebhardt FM, Mitrovic AD, Vandenberg RJ, Dodd PR. Glutamate transporter variants reduce glutamate uptake in Alzheimer’s disease. Neurobiol Aging. 2011;32:553.e1–553.e11. doi: 10.1016/j.neurobiolaging.2010.03.008. [DOI] [PubMed] [Google Scholar]
  • 51.Li S, Mallory M, Alford M, Tanaka S, Masliah E. Glutamate transporter alterations in Alzheimer disease are possibly associated with abnormal APP expression. J Neuropathol Exp Neurol. 1997;56:901–11. doi: 10.1097/00005072-199708000-00008. [DOI] [PubMed] [Google Scholar]
  • 52.Walton HS, Dodd PR. Glutamate-glutamine cycling in Alzheimer’s disease. Neurochem Int. 2007;50:1052–66. doi: 10.1016/j.neuint.2006.10.007. [DOI] [PubMed] [Google Scholar]
  • 53.Smith CD, Carney JM, Starke-Reed PE, Oliver CN, Stadtman ER, Floyd RA, et al. Excess brain protein oxidation and enzyme dysfunction in normal aging and in Alzheimer disease. Proc Natl Acad Sci USA. 1991;88:10540–3. doi: 10.1073/pnas.88.23.10540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Le Prince G, Delaere P, Fages C, Lefrançois T, Touret M, Salanon M, et al. Glutamine synthetase (GS) expression is reduced in senile dementia of the Alzheimer type. Neurochem Res. 1995;20:859–62. doi: 10.1007/BF00969698. [DOI] [PubMed] [Google Scholar]
  • 55.Liraz O, Boehm-Cagan A, Michaelson DM. ApoE4 induces Aβ42, tau, and neuronal pathology in the hippocampus of young targeted replacement apoE4 mice. Mol Neurodegener. 2013;8:16. doi: 10.1186/1750-1326-8-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Madeira C, Vargas-Lopes C, Brandão CO, Reis T, Laks J, Panizzutti R, et al. Elevated glutamate and glutamine levels in the cerebrospinal fluid of patients with probable Alzheimer’s disease and depression. Front Psychiatry. 2018;9:561. [DOI] [PMC free article] [PubMed]
  • 57.Pomara N, Singh R, Deptula D, Chou JCY, Schwartz MB, LeWitt PA. Glutamate and other CSF amino acids in Alzheimer’s disease. Am J Psychiatry. 1992;149:251–4. doi: 10.1176/ajp.149.2.251. [DOI] [PubMed] [Google Scholar]
  • 58.Smith CCT, Bowen DM, Francis PT, Snowden JS, Neary D. Putative amino acid transmitters in lumbar cerebrospinal fluid of patients with histologically verified Alzheimer’s dementia. J Neurol Neurosurg Psychiatry. 1985;48:469–71. doi: 10.1136/jnnp.48.5.469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Basun H, Forssell LG, Almkvist O, Cowburn RF, Eklöf R, Winblad B, et al. Amino acid concentrations in cerebrospinal fluid and plasma in Alzheimer’s disease and healthy control subjects. J Neural Transm Park Dis Dement Sect. 1990;2:295–304. doi: 10.1007/BF02252924. [DOI] [PubMed] [Google Scholar]
  • 60.Kuiper MA, Teerlink T, Visser JJ, Bergmans PLM, Scheltens P, Wolters EC. L-glutamate, L-arginine and L-citrulline levels in cerebrospinal fluid of Parkinson’s disease, multiple system atrophy, and Alzheimer’s disease patients. J Neural Transm. 2000;107:183–9. doi: 10.1007/s007020050016. [DOI] [PubMed] [Google Scholar]
  • 61.Hanson JE, Ma K, Elstrott J, Weber M, Saillet S, Khan AS, et al. GluN2A NMDA receptor enhancement improves brain oscillations, synchrony, and cognitive functions in Dravet syndrome and Alzheimer’s disease models. Cell Rep. 2020;30:381–96.e4. doi: 10.1016/j.celrep.2019.12.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Parsons MP, Raymond LA. Extrasynaptic NMDA receptor involvement in central nervous system disorders. Neuron. 2014;82:279–93. doi: 10.1016/j.neuron.2014.03.030. [DOI] [PubMed] [Google Scholar]
  • 63.Ping Y, Hahm E-T, Waro G, Song Q, Vo-Ba D-A, Licursi A, et al. Linking aβ42-induced hyperexcitability to neurodegeneration, learning and motor deficits, and a shorter lifespan in an Alzheimer’s model. PLoS Genet. 2015;11:1–25.. doi: 10.1371/journal.pgen.1005025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Ho R, Ortiz D, Shea TB. Amyloid-β promotes calcium influx and neurodegeneration via stimulation of L voltage-sensitive calcium channels rather than NMDA channels in cultured neurons. J Alzheimer’s Dis. 2001;3:479–83. doi: 10.3233/JAD-2001-3507. [DOI] [PubMed] [Google Scholar]
  • 65.Arispe N, Pollard HB, Rojas E. Giant multilevel cation channels formed by Alzheimer disease amyloid β- protein [AβP-(1-40)] in bilayer membranes. Proc Natl Acad Sci USA. 1993;90:10573–7. [DOI] [PMC free article] [PubMed]
  • 66.Talantova M, Sanz-Blasco S, Zhang X, Xia P, Akhtar MW, Okamoto SI, et al. Aβ induces astrocytic glutamate release, extrasynaptic NMDA receptor activation, and synaptic loss. Proc Natl Acad Sci USA. 2013;110:E2518–27. doi: 10.1073/pnas.1306832110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Wu SZ, Bodles AM, Porter MM, Griffin WST, Basile AS, Barger SW. Induction of serine racemase expression and D-serine release from microglia by amyloid β-peptide. J Neuroinflammation. 2004;1:2. doi: 10.1186/1742-2094-1-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Li S, Hong S, Shepardson NE, Walsh DM, Shankar GM, Selkoe D. Soluble oligomers of amyloid β protein facilitate hippocampal long-term depression by disrupting neuronal glutamate uptake. Neuron. 2009;62:788–801. doi: 10.1016/j.neuron.2009.05.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Arias C, Arrieta I, Tapia R. β‐Amyloid peptide fragment 25–35 potentiates the calcium‐dependent release of excitatory amino acids from depolarized hippocampal slices. J Neurosci Res. 1995;41:561–6. doi: 10.1002/jnr.490410416. [DOI] [PubMed] [Google Scholar]
  • 70.Parpura-Gill A, Beitz D, Uemura E. The inhibitory effects of β-amyloid on glutamate and glucose uptakes by cultured astrocytes. Brain Res. 1997;754:65–71. doi: 10.1016/S0006-8993(97)00043-7. [DOI] [PubMed] [Google Scholar]
  • 71.Zott B, Simon MM, Hong W, Unger F, Chen-Engerer HJ, Frosch MP, et al. A vicious cycle of β amyloid−dependent neuronal hyperactivation. Science. 2019;365:559–65. doi: 10.1126/science.aay0198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Cirrito JR, Yamada KA, Finn MB, Sloviter RS, Bales KR, May PC, et al. Synaptic activity regulates interstitial fluid amyloid-β levels in vivo. Neuron. 2005;48:913–22. doi: 10.1016/j.neuron.2005.10.028. [DOI] [PubMed] [Google Scholar]
  • 73.Yamamoto K, Tanei ZI, Hashimoto T, Wakabayashi T, Okuno H, Naka Y, et al. Chronic optogenetic activation augments Aβ pathology in a mouse model of Alzheimer disease. Cell Rep. 2015;11:859–65. doi: 10.1016/j.celrep.2015.04.017. [DOI] [PubMed] [Google Scholar]
  • 74.Götz J, Halliday G, Nisbet RM. Molecular pathogenesis of the tauopathies. Annu Rev Pathol. 2019;14:239–61. doi: 10.1146/annurev-pathmechdis-012418-012936. [DOI] [PubMed] [Google Scholar]
  • 75.Beagle AJ, Darwish SM, Ranasinghe KG, La AL, Karageorgiou E, Vossel KA. Relative incidence of seizures and myoclonus in Alzheimer’s disease, dementia with lewy bodies, and frontotemporal dementia. J Alzheimer’s Dis. 2017;60:211–23. doi: 10.3233/JAD-170031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.DeVos SL, Goncharoff DK, Chen G, Kebodeaux CS, Yamada K, Stewart FR, et al. Antisense reduction of tau in adult mice protects against seizures. J Neurosci. 2013;33:12887–97. doi: 10.1523/JNEUROSCI.2107-13.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Holth JK, Bomben VC, Graham Reed J, Inoue T, Younkin L, Younkin SG, et al. Tau loss attenuates neuronal network hyperexcitability in mouse and drosophila genetic models of epilepsy. J Neurosci. 2013;33:1651–9. doi: 10.1523/JNEUROSCI.3191-12.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Roberson ED, Halabisky B, Yoo JW, Yao J, Chin J, Yan F, et al. Amyloid-β/fyn-induced synaptic, network, and cognitive impairments depend on tau levels in multiple mouse models of alzheimer’s disease. J Neurosci. 2011;31:700–11. doi: 10.1523/JNEUROSCI.4152-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Ittner LM, Ke YD, Delerue F, Bi M, Gladbach A, van Eersel J, et al. Dendritic function of tau mediates amyloid-β toxicity in alzheimer’s disease mouse models. Cell. 2010;142:387–97. doi: 10.1016/j.cell.2010.06.036. [DOI] [PubMed] [Google Scholar]
  • 80.Decker JM, Krüger L, Sydow A, Dennissen FJ, Siskova Z, Mandelkow E, et al. The Tau/A152T mutation, a risk factor for frontotemporal-spectrum disorders, leads to NR2B receptor-mediated excitotoxicity. EMBO Rep. 2016;17:552–69. doi: 10.15252/embr.201541439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Crimins JL, Rocher AB, Luebke JI. Electrophysiological changes precede morphological changes to frontal cortical pyramidal neurons in the rTg4510 mouse model of progressive tauopathy. Acta Neuropathol. 2012;124:777–95. doi: 10.1007/s00401-012-1038-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Witton J, Staniaszek LE, Bartsch U, Randall AD, Jones MW, Brown JT. Disrupted hippocampal sharp-wave ripple-associated spike dynamics in a transgenic mouse model of dementia. J Physiol. 2016;594:4615–30. doi: 10.1113/jphysiol.2014.282889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Roberson ED, Scearce-Levie K, Palop JJ, Yan F, Cheng IH, Wu T, et al. Reducing endogenous tau ameliorates amyloid β-induced deficits in an Alzheimer’s disease mouse model. Science. 2007;316:750–4. doi: 10.1126/science.1141736. [DOI] [PubMed] [Google Scholar]
  • 84.Pooler AM, Phillips EC, Lau DHW, Noble W, Hanger DP. Physiological release of endogenous tau is stimulated by neuronal activity. EMBO Rep. 2013;14:389–94. doi: 10.1038/embor.2013.15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Wu JW, Hussaini SA, Bastille IM, Rodriguez GA, Mrejeru A, Rilett K, et al. Neuronal activity enhances tau propagation and tau pathology in vivo. Nat Neurosci. 2016;19:1085–92. doi: 10.1038/nn.4328. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Hall AM, Throesch BT, Buckingham SC, Markwardt SJ, Peng Y, Wang Q, et al. Tau-dependent Kv4.2 depletion and dendritic hyperexcitability in a mouse model of Alzheimer’s disease. J Neurosci. 2015;35:6221–30. doi: 10.1523/JNEUROSCI.2552-14.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Maeda S, Djukic B, Taneja P, Yu G-Q, Lo I, Davis A, et al. Expression of A152T human tau causes age-dependent neuronal dysfunction and loss in transgenic mice. EMBO Rep. 2016;17:530–51.. doi: 10.15252/embr.201541438. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Hatch RJ, Wei Y, Xia D, Götz J. Hyperphosphorylated tau causes reduced hippocampal CA1 excitability by relocating the axon initial segment. Acta Neuropathol. 2017;133:717–30. doi: 10.1007/s00401-017-1674-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Marinković P, Blumenstock S, Goltstein PM, Korzhova V, Peters F, Knebl A, et al. In vivo imaging reveals reduced activity of neuronal circuits in a mouse tauopathy model. Brain. 2019;142:1051–62. doi: 10.1093/brain/awz035. [DOI] [PubMed] [Google Scholar]
  • 90.Angulo SL, Orman R, Neymotin SA, Liu L, Buitrago L, Cepeda-Prado E, et al. Tau and amyloid-related pathologies in the entorhinal cortex have divergent effects in the hippocampal circuit. Neurobiol Dis. 2017;108:261–76. doi: 10.1016/j.nbd.2017.08.015. [DOI] [PubMed] [Google Scholar]
  • 91.Busche MA, Wegmann S, Dujardin S, Commins C, Schiantarelli J, Klickstein N, et al. Tau impairs neural circuits, dominating amyloid-β effects, in Alzheimer models in vivo. Nat Neurosci. 2019;22:57–64. doi: 10.1038/s41593-018-0289-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Corder EH, Saunders AM, Strittmatter WJ, Schmechel DE, Gaskell PC, Small GW, et al. Gene dose of apolipoprotein E type 4 allele and the risk of Alzheimer’s disease in late onset families. Science. 1993;261:921–3. doi: 10.1126/science.8346443. [DOI] [PubMed] [Google Scholar]
  • 93.Mahley RW. Apolipoprotein E: cholesterol transport protein with expanding role in cell biology. Science. 1988;240:622–30. doi: 10.1126/science.3283935. [DOI] [PubMed] [Google Scholar]
  • 94.Burggren AC, Small GW, Sabb FW, Bookheimer SY. Specificity of brain activation patterns in people at genetic risk for Alzheimer disease. Am J Geriatr Psychiatry. 2002;10:44–51. doi: 10.1097/00019442-200201000-00006. [DOI] [PubMed] [Google Scholar]
  • 95.Fleisher AS, Houston WS, Eyler LT, Frye S, Jenkins C, Thal LJ, et al. Identification of Alzheimer disease risk by functional magnetic resonance imaging. Arch Neurol. 2005;62:1881–8. doi: 10.1001/archneur.62.12.1881. [DOI] [PubMed] [Google Scholar]
  • 96.Wishart HA, Saykin AJ, Rabin LA, Santulli RB, Flashman LA, Guerin SJ, et al. Increased brain activation during working memory in cognitively intact adults with the APOE ε4 allele. Am J Psychiatry. 2006;163:1603–10. doi: 10.1176/ajp.2006.163.9.1603. [DOI] [PubMed] [Google Scholar]
  • 97.Hunter JM, Cirrito JR, Restivo JL, Kinley RD, Sullivan PM, Holtzman DM, et al. Emergence of a seizure phenotype in aged apolipoprotein epsilon 4 targeted replacement mice. Brain Res. 2012;1467:120–32. doi: 10.1016/j.brainres.2012.05.048. [DOI] [PubMed] [Google Scholar]
  • 98.Sun G-Z, He Y-C, Ma XK, Li S-T, Chen D-J, Gao M, et al. Hippocampal synaptic and neural network deficits in young mice carrying the human APOE4 gene. CNS Neurosci Ther. 2017;23:748–58. doi: 10.1111/cns.12720. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Lane-Donovan C, Wong WM, Durakoglugil MS, Wasser CR, Jiang S, Xian X, et al. Genetic restoration of plasma apoe improves cognition and partially restores synaptic defects in ApoE-deficient mice. J Neurosci. 2016;36:10141–50. doi: 10.1523/JNEUROSCI.1054-16.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Li G, Bien-Ly N, Andrews-Zwilling Y, Xu Q, Bernardo A, Ring K, et al. GABAergic interneuron dysfunction impairs hippocampal neurogenesis in adult apolipoprotein E4 knockin mice. Cell Stem Cell. 2009;5:634–45. doi: 10.1016/j.stem.2009.10.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Dumanis SB, Tesoriero JA, Babus LW, Nguyen MT, Trotter JH, Ladu MJ, et al. ApoE4 decreases spine density and dendritic complexity in cortical neurons in vivo. J Neurosci. 2009;29:15317–22. doi: 10.1523/JNEUROSCI.4026-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Konttinen H, Cabral-da-Silva MeC, Ohtonen S, Wojciechowski S, Shakirzyanova A, Caligola S, et al. PSEN1ΔE9, APPswe, and APOE4 confer disparate phenotypes in human iPSC-derived microglia. Stem Cell Rep. 2019;13:669–83. doi: 10.1016/j.stemcr.2019.08.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Brookhouser N, Raman S, Frisch C, Srinivasan G, Brafman DA. APOE2 mitigates disease-related phenotypes in an isogenic hiPSC-based model of Alzheimer’s disease. Mol Psychiatry. 2021;26:5715–32. [DOI] [PMC free article] [PubMed]
  • 104.Rogers J, Webster S, Lue LF, Brachova L, Civin WH, Emmerling M, et al. Inflammation and Alzheimer’s disease pathogenesis. Neurobiol Aging. 1996;17:681–6. doi: 10.1016/0197-4580(96)00115-7. [DOI] [PubMed] [Google Scholar]
  • 105.Ooi L, Dottori M, Cook AL, Engel M, Gautam V, Grubman A, et al. If human brain organoids are the answer to understanding dementia, what are the questions? Neuroscientist. 2020;26:438–54. doi: 10.1177/1073858420912404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Karch CM, Goate AM. Alzheimer’s disease risk genes and mechanisms of disease pathogenesis. Biol Psychiatry. 2015;77:43–51. doi: 10.1016/j.biopsych.2014.05.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Chang MY, Boulden J, Katz JB, Wang L, Meyer TJ, Soler AP, et al. Bin1 ablation increases susceptibility to cancer during aging, particularly lung cancer. Cancer Res. 2007;67:7605–12. doi: 10.1158/0008-5472.CAN-07-1100. [DOI] [PubMed] [Google Scholar]
  • 108.Butler MH, David C, Ochoa GC, Freyberg Z, Daniell L, Grabs D, et al. Amphiphysin II (SH3p9; BIN1), a member of the amphiphysin/Rvs family, is concentrated in the cortical cytomatrix of axon initial segments and nodes of ranvier in brain and around T tubules in skeletal muscle. J Cell Biol. 1997;137:1355–67. doi: 10.1083/jcb.137.6.1355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Voskobiynyk Y, Roth JR, Cochran JN, Rush T, Carullo NVN, Mesina JS, et al. Alzheimer’s disease risk gene BIN1 induces Tau-dependent network hyperexcitability. Elife. 2020;9:1–25.. doi: 10.7554/eLife.57354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Baumann N, Pham-Dinh D. Biology of oligodendrocyte and myelin in the mammalian central nervous system. Physiol Rev. 2001;81:871–927. doi: 10.1152/physrev.2001.81.2.871. [DOI] [PubMed] [Google Scholar]
  • 111.Butt AM, De La Rocha IC, Rivera A. Oligodendroglial cells in alzheimer’s disease. In: Verkhratsky A, Ho MS, Zorec R, Parpura V, editors. Advances in experimental medicine and biology. Singapore: Springer; 2019;325–33. [DOI] [PubMed]
  • 112.Nasrabady SE, Rizvi B, Goldman JE, Brickman AM. White matter changes in Alzheimer’s disease: a focus on myelin and oligodendrocytes. Acta Neuropathol Commun. 2018;6:22. doi: 10.1186/s40478-018-0515-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Devinsky O, Vezzani A, Najjar S, De Lanerolle NC, Rogawski MA. Glia and epilepsy: excitability and inflammation. Trends Neurosci. 2013;36:174–84. doi: 10.1016/j.tins.2012.11.008. [DOI] [PubMed] [Google Scholar]
  • 114.Dubé C, Vezzani A, Behrens M, Bartfai T, Baram TZ. Interleukin-1β contributes to the generation of experimental febrile seizures. Ann Neurol. 2005;57:152. doi: 10.1002/ana.20358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Bertani I, Iori V, Trusel M, Maroso M, Foray C, Mantovani S, et al. Inhibition of IL-1β signaling normalizes NMDA-dependent neurotransmission and reduces seizure susceptibility in a mouse model of Creutzfeldt-Jakob disease. J Neurosci. 2017;37:10278–89. doi: 10.1523/JNEUROSCI.1301-17.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Rodgers KM, Hutchinson MR, Northcutt A, Maier SF, Watkins LR, Barth DS. The cortical innate immune response increases local neuronal excitability leading to seizures. Brain. 2009;132:2478. doi: 10.1093/brain/awp177. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Ravizza T, Lucas SM, Balosso S, Bernardino L, Ku G, Noé F, et al. Inactivation of caspase-1 in rodent brain: a novel anticonvulsive strategy. Epilepsia. 2006;47:1160–8. doi: 10.1111/j.1528-1167.2006.00590.x. [DOI] [PubMed] [Google Scholar]
  • 118.Balosso S, Maroso M, Sanchez-Alavez M, Ravizza T, Frasca A, Bartfai T, et al. A novel non-transcriptional pathway mediates the proconvulsive effects of interleukin-1β. Brain. 2008;131:3256. doi: 10.1093/brain/awn271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Beattie EC, Stellwagen D, Morishita W, Bresnahan JC, Byeong KH, Von Zastrow M, et al. Control of synaptic strength by glial TNFα. Science. 2002;295:2282–5. doi: 10.1126/science.1067859. [DOI] [PubMed] [Google Scholar]
  • 120.Santello M, Bezzi P, Volterra A. TNFα controls glutamatergic gliotransmission in the hippocampal dentate gyrus. Neuron. 2011;69:988–1001. doi: 10.1016/j.neuron.2011.02.003. [DOI] [PubMed] [Google Scholar]
  • 121.Frigerio F, Pasqualini G, Craparotta I, Marchini S, Van Vliet EA, Foerch P, et al. N-3 docosapentaenoic acid-derived protectin D1 promotes resolution of neuroinflammation and arrests epileptogenesis. Brain. 2018;141:3130–43. doi: 10.1093/brain/awy247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Lagarde S, Villeneuve N, Trébuchon A, Kaphan E, Lepine A, McGonigal A, et al. Anti-tumor necrosis factor alpha therapy (adalimumab) in Rasmussen’s encephalitis: an open pilot study. Epilepsia. 2016;57:956–66. doi: 10.1111/epi.13387. [DOI] [PubMed] [Google Scholar]
  • 123.Takahashi K, Rochford CDP, Neumann H. Clearance of apoptotic neurons without inflammation by microglial triggering receptor expressed on myeloid cells-2. J Exp Med. 2005;201:647–57. doi: 10.1084/jem.20041611. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Leyns CEG, Gratuze M, Narasimhan S, Jain N, Koscal LJ, Jiang H, et al. TREM2 function impedes tau seeding in neuritic plaques. Nat Neurosci. 2019;22:1217. doi: 10.1038/s41593-019-0433-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Yuan P, Condello C, Keene CD, Wang Y, Bird TD, Paul SM, et al. TREM2 haplodeficiency in mice and humans impairs the microglia barrier function leading to decreased amyloid compaction and severe axonal dystrophy. Neuron. 2016;90:724. doi: 10.1016/j.neuron.2016.05.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Lenz KM, Nelson LH. Microglia and beyond: Innate immune cells as regulators of brain development and behavioral function. Front Immunol. 2018;9:698. doi: 10.3389/fimmu.2018.00698. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Umpierre AD, Wu LJ. How microglia sense and regulate neuronal activity. Glia. 2021;69:1637–53. doi: 10.1002/glia.23961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Brawek B, Schwendele B, Riester K, Kohsaka S, Lerdkrai C, Liang Y, et al. Impairment of in vivo calcium signaling in amyloid plaque-associated microglia. Acta Neuropathol. 2014;127:495–505. doi: 10.1007/s00401-013-1242-2. [DOI] [PubMed] [Google Scholar]
  • 129.Umpierre AD, Bystrom LL, Ying Y, Liu YU, Worrell G, Wu LJ. Microglial calcium signaling is attuned to neuronal activity in awake mice. Elife. 2020;9:1–24.. doi: 10.7554/eLife.56502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Li Y, Du XF, Liu CS, Wen ZL, Du JL. Reciprocal regulation between resting microglial dynamics and neuronal activity in vivo. Dev Cell. 2012;23:1189–202. doi: 10.1016/j.devcel.2012.10.027. [DOI] [PubMed] [Google Scholar]
  • 131.Badimon A, Strasburger HJ, Ayata P, Chen X, Nair A, Ikegami A, et al. Negative feedback control of neuronal activity by microglia. Nature. 2020;586:417–23. doi: 10.1038/s41586-020-2777-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Wu W, Li Y, Wei Y, Bosco DB, Xie M, Zhao MG, et al. Microglial depletion aggravates the severity of acute and chronic seizures in mice. Brain Behav Immun. 2020;89:245–55. doi: 10.1016/j.bbi.2020.06.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Akiyoshi R, Wake H, Kato D, Horiuchi H, Ono R, Ikegami A, et al. Microglia enhance synapse activity to promote local network synchronization. eNeuro. 2018;5:88–106.. doi: 10.1523/ENEURO.0088-18.2018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Wang DD, Bordey A. The astrocyte odyssey. Prog Neurobiol. 2008;86:342–67. doi: 10.1016/j.pneurobio.2008.09.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Kofuji P, Newman EA. Potassium buffering in the central nervous system. Neuroscience. 2004;129:1043–54. doi: 10.1016/j.neuroscience.2004.06.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Wang F, Qi X, Zhang J, Huang J. Astrocytic modulation of potassium under seizures. Neural Regeneration Res. 2020;15:980–7. doi: 10.4103/1673-5374.270295. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Wilcock DM, Vitek MP, Colton CA. Vascular amyloid alters astrocytic water and potassium channels in mouse models and humans with Alzheimer’s disease. Neuroscience. 2009;159:1055. doi: 10.1016/j.neuroscience.2009.01.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Pirttimaki TM, Codadu NK, Awni A, Pratik P, Nagel DA, Hill EJ, et al. α7 nicotinic receptor-mediated astrocytic gliotransmitter release: Aβ effects in a preclinical Alzheimer’s mouse model. PLoS One. 2013;8:e81828. doi: 10.1371/journal.pone.0081828. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Buffo A, Rolando C, Ceruti S. Astrocytes in the damaged brain: Molecular and cellular insights into their reactive response and healing potential. Biochem Pharm. 2010;79:77–89. doi: 10.1016/j.bcp.2009.09.014. [DOI] [PubMed] [Google Scholar]
  • 140.Ortinski PI, Dong J, Mungenast A, Yue C, Takano H, Watson DJ, et al. Selective induction of astrocytic gliosis generates deficits in neuronal inhibition. Nat Neurosci. 2010;13:584–91. doi: 10.1038/nn.2535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Kepecs A, Fishell G. Interneuron cell types are fit to function. Nature. 2014;505:318–26. doi: 10.1038/nature12983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Matsumoto JY, Stead M, Kucewicz MT, Matsumoto AJ, Peters PA, Brinkmann BH, et al. Network oscillations modulate interictal epileptiform spike rate during human memory. Brain. 2013;136:2444–56. doi: 10.1093/brain/awt159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Nuriel T, Angulo SL, Khan U, Ashok A, Chen Q, Figueroa HY, et al. Neuronal hyperactivity due to loss of inhibitory tone in APOE4 mice lacking Alzheimer’s disease-like pathology. Nat Commun. 2017;8:1–14.. doi: 10.1038/s41467-017-01444-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Verret L, Mann EO, Hang GB, Barth AMI, Cobos I, Ho K, et al. Inhibitory interneuron deficit links altered network activity and cognitive dysfunction in alzheimer model. Cell. 2012;149:708–21. doi: 10.1016/j.cell.2012.02.046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Hamm V, Héraud C, Bott JB, Herbeaux K, Strittmatter C, Mathis C, et al. Differential contribution of APP metabolites to early cognitive deficits in a TgCRND8 mouse model of Alzheimer’s disease. Sci Adv. 2017;3:e1601068. doi: 10.1126/sciadv.1601068. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Martinez-Losa M, Tracy TE, Ma K, Verret L, Clemente-Perez A, Khan AS, et al. Nav1.1-overexpressing interneuron transplants restore brain rhythms and cognition in a mouse model of Alzheimer’s disease. Neuron. 2018;98:75–89. doi: 10.1016/j.neuron.2018.02.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Chow CY, Chin YKY, Ma L, Undheim EAB, Herzig V, King GF. A selective NaV1.1 activator with potential for treatment of Dravet syndrome epilepsy. Biochem Pharmacol. 2020;181:113991. [DOI] [PubMed]
  • 148.Richards KL, Milligan CJ, Richardson RJ, Jancovski N, Grunnet M, Jacobson LH, et al. Selective NaV1.1 activation rescues Dravet syndrome mice from seizures and premature death. Proc Natl Acad Sci USA. 2018;115:E8077–85.. doi: 10.1073/pnas.1804764115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Deture MA, Dickson DW. The neuropathological diagnosis of Alzheimer’s disease. Mol Neurodegeneration. 2019;14:1–18.. doi: 10.1186/s13024-019-0333-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Bakker A, Krauss GL, Albert MS, Speck CL, Jones LR, Stark CE, et al. Reduction of hippocampal hyperactivity improves cognition in amnestic mild cognitive impairment. Neuron. 2012;74:467–74. doi: 10.1016/j.neuron.2012.03.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Sanchez PE, Zhu L, Verret L, Vossel KA, Orr AG, Cirrito JR, et al. Levetiracetam suppresses neuronal network dysfunction and reverses synaptic and cognitive deficits in an Alzheimer’s disease model. Proc Natl Acad Sci USA. 2012;109:E2895–903. [DOI] [PMC free article] [PubMed]
  • 152.Yuan P, Grutzendler J. Attenuation of β-amyloid deposition and neurotoxicity by chemogenetic modulation of neural activity. J Neurosci. 2016;36:632–41. doi: 10.1523/JNEUROSCI.2531-15.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Vossel KA, Ranasinghe KG, Beagle AJ, Mizuiri D, Honma SM, Dowling AF, et al. Incidence and impact of subclinical epileptiform activity in Alzheimer’s disease. Ann Neurol. 2016;80:858–70. doi: 10.1002/ana.24794. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Do-Ha D, Buskila Y, Ooi L. Impairments in motor neurons, interneurons and astrocytes contribute to hyperexcitability in ALS: underlying mechanisms and paths to therapy. Mol Neurobiol. 2018;55:1410–8. doi: 10.1007/s12035-017-0392-y. [DOI] [PubMed] [Google Scholar]
  • 155.Lalancette-Hebert M, Sharma A, Lyashchenko AK, Shneider NA. Gamma motor neurons survive and exacerbate alpha motor neuron degeneration in ALS. Proc Natl Acad Sci USA. 2016;113:E8316–25.. doi: 10.1073/pnas.1605210113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Houweling AR, Brecht M. Behavioural report of single neuron stimulation in somatosensory cortex. Nature. 2008;451:65–8. doi: 10.1038/nature06447. [DOI] [PubMed] [Google Scholar]
  • 157.Brecht M, Schneider M, Sakmann B, Margie TW. Whisker movements evoked by stimulation of single pyramidal cells in rat motor cortex. Nature. 2004;427:704–10. doi: 10.1038/nature02266. [DOI] [PubMed] [Google Scholar]
  • 158.Grienberger C, Rochefort NL, Adelsberger H, Henning HA, Hill DN, Reichwald J, et al. Staged decline of neuronal function in vivo in an animal model of Alzheimer’s disease. Nat Commun. 2012;3:1–10. doi: 10.1038/ncomms1783. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Wilson IA, Ikonen S, Gallagher M, Eichenbaum H, Tanila H. Age-associated alterations of hippocampal place cells are subregion specific. J Neurosci. 2005;25:6877. doi: 10.1523/JNEUROSCI.1744-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Gosselin D, Skola D, Coufal NG, Holtman IR, Schlachetzki JCM, Sajti E, et al. An environment-dependent transcriptional network specifies human microglia identity. Science. 2017;356:1248–59.. doi: 10.1126/science.aal3222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Bohlen CJ, Bennett FC, Tucker AF, Collins HY, Mulinyawe SB, Barres BA. Diverse requirements for microglial survival, specification, and function revealed by defined-medium cultures. Neuron. 2017;94:759. doi: 10.1016/j.neuron.2017.04.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Astashkina AI, Mann BK, Prestwich GD, Grainger DW. Comparing predictive drug nephrotoxicity biomarkers in kidney 3-D primary organoid culture and immortalized cell lines. Biomaterials. 2012;33:4712–21. doi: 10.1016/j.biomaterials.2012.03.001. [DOI] [PubMed] [Google Scholar]
  • 163.Centeno EGZ, Cimarosti H, Bithell A. 2D versus 3D human induced pluripotent stem cell-derived cultures for neurodegenerative disease modelling. Mol Neurodegener. 2018;13:27. doi: 10.1186/s13024-018-0258-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Quadrato G, Nguyen T, Macosko EZ, Sherwood JL, Yang SM, Berger DR, et al. Cell diversity and network dynamics in photosensitive human brain organoids. Nature. 2017;545:48–53. doi: 10.1038/nature22047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Oksanen M, Petersen AJ, Naumenko N, Puttonen K, Lehtonen Š, Gubert Olivé M, et al. PSEN1 mutant iPSC-derived model reveals severe astrocyte pathology in Alzheimer’s disease. Stem Cell Rep. 2017;9:1885–97. doi: 10.1016/j.stemcr.2017.10.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Bernareggi A, Dueñas Z, Reyes-Ruiz JM, Ruzzier F, Miledi R. Properties of glutamate receptors of Alzheimer’s disease brain transplanted to frog oocytes. Proc Natl Acad Sci USA. 2007;104:2956. doi: 10.1073/pnas.0611513104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Limon A, Reyes-Ruiz JM, Miledi R. Loss of functional GABA A receptors in the Alzheimer diseased brain. Proc Natl Acad Sci USA. 2012;109:10071–6. doi: 10.1073/pnas.1204606109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Miledi R, Dueñas Z, Martinez-Torres A, Kawas CH, Eusebi F. From the cover: microtransplantation of functional receptors and channels from the Alzheimer’s brain to frog oocytes. Proc Natl Acad Sci USA. 2004;101:1760. doi: 10.1073/pnas.0308224100. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Translational Psychiatry are provided here courtesy of Nature Publishing Group

RESOURCES