Next Article in Journal
In Vivo and Ex Vivo Mitochondrial Function in COVID-19 Patients on the Intensive Care Unit
Next Article in Special Issue
Post-Synapses in the Brain: Role of Dendritic and Spine Structures
Previous Article in Journal
Obesity and Endothelial Function
Previous Article in Special Issue
Mechanisms of Synaptic Vesicle Exo- and Endocytosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Synaptic Disruption by Soluble Oligomers in Patients with Alzheimer’s and Parkinson’s Disease

by
Berenice A. Gutierrez
and
Agenor Limon
*
Mitchell Center for Neurodegenerative Diseases, Department of Neurology, School of Medicine, University of Texas Medical Branch, Galveston, TX 77555, USA
*
Author to whom correspondence should be addressed.
Biomedicines 2022, 10(7), 1743; https://doi.org/10.3390/biomedicines10071743
Submission received: 23 June 2022 / Revised: 13 July 2022 / Accepted: 15 July 2022 / Published: 19 July 2022
(This article belongs to the Special Issue News about Structure and Function of Synapses: Health and Diseases)

Abstract

:
Neurodegenerative diseases are the result of progressive dysfunction of the neuronal activity and subsequent neuronal death. Currently, the most prevalent neurodegenerative diseases are by far Alzheimer’s (AD) and Parkinson’s (PD) disease, affecting millions of people worldwide. Although amyloid plaques and neurofibrillary tangles are the neuropathological hallmarks for AD and Lewy bodies (LB) are the hallmark for PD, current evidence strongly suggests that oligomers seeding the neuropathological hallmarks are more toxic and disease-relevant in both pathologies. The presence of small soluble oligomers is the common bond between AD and PD: amyloid β oligomers (AβOs) and Tau oligomers (TauOs) in AD and α-synuclein oligomers (αSynOs) in PD. Such oligomers appear to be particularly increased during the early pathological stages, targeting synapses at vulnerable brain regions leading to synaptic plasticity disruption, synapse loss, inflammation, excitation to inhibition imbalance and cognitive impairment. Absence of TauOs at synapses in individuals with strong AD disease pathology but preserved cognition suggests that mechanisms of resilience may be dependent on the interactions between soluble oligomers and their synaptic targets. In this review, we will discuss the current knowledge about the interactions between soluble oligomers and synaptic dysfunction in patients diagnosed with AD and PD, how it affects excitatory and inhibitory synaptic transmission, and the potential mechanisms of synaptic resilience in humans.

1. Introduction

Alzheimer’s disease (AD) and Parkinson’s disease (PD) are among the most prevalent neurodegenerative diseases that shares the misfolding of proteins and synaptic dysfunction as part of their neuropathology. The World Health Organization estimates that 55 million people worldwide live with dementia, of which two-thirds are due to AD [1]. PD is the second most common age-related neurodegenerative disorder after AD with an estimated of up to 10 million people worldwide [2]. Due to the large prevalence of these disorders, the economical and psychological costs on society, caregivers and individuals affected is extremely high. Although large advances in understanding the potential causes of the clinical symptoms at different levels of analysis has been made, there is still the need of effective disease-modifying treatments that can help people with a diagnosis and those at high risk. In this review we will discuss recent studies about the role of oligomers of misfolded proteins on the vulnerability of the distinct types of synapses and the potential mechanisms of synaptic resilience in humans.
AD is a slowly progressive neurological disease, clinically characterized by a decline in memory, language, and other cognitive skills [3,4]. The diagnosis as recommended by the National Institute on Aging-Alzheimer’s disease Association (NIA-AA) research framework requires the presence of extracellular amyloid plaques and intracellular neurofibrillary tangles under a microscopic examination of several brain regions [5]; thus, the definitive diagnosis of AD is made after death. Alive people with amnestic dementia are usually diagnosed with AD clinical syndrome [6].
The pathological hallmarks of AD are the final expression of a molecular chain of events, including misfolding of proteins that form large polymers of aggregates. Amyloid-beta (Aβ) deposits are the consequence of abnormal cleavage of the Aβ precursor protein (APP), a cell surface receptor thought to regulate neuronal migration during development [7]. APP cleavage can generate amyloidogenic (pathogenic) and non-amyloidogenic (non-pathogenic) Aβ forms. The β-site APP cleaving enzyme I (BACE1) cleaves APP close to its N-terminus between residues M671 and D672 and is responsible for generating the amyloidogenic Aβ form [8]. Abnormal aggregation of Aβ monomers (Aβ40 and Aβ42 with 40 and 42 amino acids) leads to the formation of Aβ oligomers (AβOs) and the establishment of extracellular amyloid plaques that promote neuronal synaptic and cell loss [7]. Additionally, but not necessarily in parallel, tau proteins (ranging from 352 to 421 amino acids in AD), which are part of the cytoskeleton and critical for intracellular transport mechanisms [9,10,11], become hyperphosphorylated causing their detachment from the microtubules and leading to cytoskeletal destabilization. Tau monomers are normally found unfolded and do not form filaments. However, abnormal aggregation into straight and paired helical filaments may occur; this process is driven by a shift from random coil to a β-sheet structure of regions within the second and third repeat domains [12]. When tau monomers detach from the microtubules, they form oligomers (TauO), which are the foundation of the intracellular neurofibrillary tangles that disrupt the synaptic transmission and lead to neuronal death [7]. Interestingly, there is no clear association between the number of Aβ and tau deposits with the severity or duration of the disease; moreover, there are adults with normal cognition and remarkable amounts of those proteins, particularly Aβ [13,14,15].
The ‘core’ of the neuronal lesions in PD is the progressive degeneration of dopamine neurons in the central nervous system (CNS), which accounts for most of the prominent symptoms (slowness of movement, rest tremor, and rigidity) [16,17,18]. The neurodegeneration of PD is driven by accumulation of α-synuclein (αSyn) and the presence of Lewy bodies and dystrophic neurites, in the central and peripheral nervous systems [19,20]. α-syn is an unfolded protein of 140 amino acids that lacks a distinct secondary or tertiary structure, and whose conformational changes generate the insoluble neurotoxic species of PD [8]. The first association between Aβ and AD was suggested by Glenner and Wong [21], while the one between αSyn and PD was done by Goldberg and Lansbury [22]. There is strong evidence of co-occurring pathologies across these neurodegenerative diseases. For example, αSyn accumulation is frequently observed in AD, whereas amyloidosis and tauopathy are also observed in PD, indicating protein-protein interactions and cross-seeding between Aβ, tau and αSyn proteins which promote aggregation and accelerate cognitive impairment [23]. Importantly, the generation and release of all these oligomers to the extracellular space is highly interdependent on the electrical activity of neurons [24,25] and has strong effect on synaptic plasticity.

2. Neurodegeneration Driven by Small Soluble Oligomers

Oligomers are small, soluble protein aggregates which possess unique structural and functional properties. They are intermediary between soluble monomeric proteins and insoluble mature fibrils [26]. For several years, therapeutic research in AD and PD was centered in targeting insoluble fibrillar aggregates of Aβ, tau and αSyn, but recent studies have shown that soluble oligomers are the most toxic species that induce neuronal damage and dysfunction in neurodegenerative disorders [27,28,29,30,31,32,33,34,35,36], suggesting that anti-oligomeric therapeutic strategies would be a better approach to antagonize cognitive deficit symptoms.
Neurodegeneration driven by AβOs has been experimentally proven using human postmortem brain tissue from subjects clinically diagnosed with AD. AβOs extracted from these subjects have shown to alter long-term potentiation, enhance long term depression, and reduce the dendritic spine density of pyramidal neurons in the hippocampus of control mice [37]. The reduction of spine density is consequence of the loss of spine cytoskeletal proteins, a phenomenon that implicates impairment of memory-related receptors such as NMDA receptors [37,38] (See below for effects of oligomers on synaptic receptors). AβOs also contribute to loss of synaptic markers such as synaptic vesicle-associated membrane protein 2 and post-synaptic density protein 95 [39] indicating reduction of synapses and loss of communication between neurons. The synaptic deterioration manifests with memory and learning impairment as observed in control rats after injection of AβOs from human AD brains [37]. Further, levels of AβOs in fractionated brain homogenates from patients with AD correlate with the severity of cognitive impairment (assessed by Blessed Information-Memory-Concentration and the Mini-Mental State Examination scores) [39]. Nevertheless, AβOs are found in subjects without cognitive impairment; they increase physiologically in old age [40]. The fact that both demented and non-demented patients have increased levels of oligomers does not necessarily mean they have the same oligomeric structure. AβOs organize into dimers, trimers, tetramers, and higher order structures. Three oligomeric structures amyloid-β trimers, Aβ*56 and amyloid-β dimers—have been identified in AD mouse models and humans [40,41]. In a cross-sectional analysis of the Religious Orders Study, two important observations were done: first, although all three AβOs increase in old age, they are always significantly higher in demented compared to cognitively intact patients; and second, the oligomeric form Aβ*56, but not Aβ dimers nor Aβ trimers, correlate positively with soluble pathological tau proteins and negatively with postsynaptic proteins in intact subjects [40]. These results propose different planes of participation of the three oligomeric structures in physiological and pathological processes [40]. For instance, Aβ*56 (56-kDa oligomer) appears to be a major oligomer forming complexes with NMDA receptors and mediating Ca2+ influx and activation of Ca2+-dependent calmodulin kinase IIα, which in turn induces disarrangement and phosphorylation of tau [40,42]. Interestingly, control subjects or Non-Demented with AD Neuropathology (NDAN)—subjects with histopathological, but not clinical AD—lack binding of AβOs to NMDA receptors or other synaptic structures [37,43], strongly suggesting that binding of AβOs to synapses is an important trigger of synaptic degeneration.
The relation between oligomer formation and disease state remains controversial. Most studies support the pathogenic role of oligomers in neurodegeneration as mentioned above. For example, elevated levels of plasma AβOs have shown a strong correlation with the cognitive performance in patients with AD (assessed by Mini-Mental State Examination, Cognitive Abilities Screening Instrument, Alzheimer’s Disease Assessment Scale–cognitive portion, and common objects memory test) [44]. Other studies have shown correlation with severity [45,46]. Additionally, in transgenic mice (PS1V97L), inhibition of AβOs showed to improve the memory function [47]. However, a few other groups have found contradictory results. AβOs were not able to induce memory deficit in mutant mice with loss of APP, suggesting that either APP is a key component of cognitive decline or that Aβ aggregates affect cognition by a yet unknown mechanism [48]. Another study showed cognitive improvement in mice with amyloid pathology after lowering the APP/Aβ production, while the amount of AβOs remained unchanged [49]. Altogether, evidence points out that oligomers play a role in the pathophysiology of neurodegenerative disorders; however, at what time of the AD continuum they become clinically significant, and whether they are the unique and most important target still needs to be determined.
TauOs are present in neurons and astrocytes at early stages of neurodegeneration, they are increased even before the formation of neurofibrillary tangles and clinical manifestations of AD [50,51,52]. As with AβOs, human brain derived TauOs injected in mice impair synaptic plasticity in the hippocampus, and clinically manifest with anterograde memory storage dysfunction [53]. TauOs are not only increased in senile and AD brains, but also, they have been detected in high levels in animal models and brains of individuals with PD, suggesting that TauOs are neurotoxic mediators in synucleinopathies [54,55]. TauOs have been associated with stress granules and molecular markers such as eIF3η, TIA1, PABP and HNRNPA2B1, inducing a translational stress response [56,57]. Moreover, TauOs form a complex with m6A and HNRNPA2B1, which is increased up to 5-fold in subjects with AD, meaning that this complex contributes to neurodegeneration [57]. Transactive response-DNA binding protein 43 (TDP-43), has been found in stress granules of patients with frontotemporal dementia and amyotrophic lateral sclerosis, and now it was shown that TDP-43 forms oligomeric assemblies that associate with AβOs and TauOs, suggesting co-partnerships in the pathogeneses of neurodegenerative diseases [58]. Further, TauOs interact with the Musashi family of RNA-binding proteins in AD brains to form nuclear aggregates and induce reduction of LaminB1, leading to nuclear instability and thus, neuronal dysfunction [59].
Neurodegeneration driven by αSynOs depends on their interaction with cell membranes. Similar to AβOs and TauOs, αSynOs can perforate the membrane and hence alter the membrane conductance; therefore, formation of ion-permeating pores seems to be a general mechanism to destabilize the cell membrane shared by some oligomeric forms of misfolded proteins. Moreover, perfusion of αSynOs onto hippocampal neurons induce an increase of intracellular calcium level, which supports the idea of strong membrane interactions [60,61]. The elevation of calcium levels is in accordance with the calcium homeostasis dysregulation observed in PD subjects [62,63,64,65]. Importantly, αSynOs contribute significantly to dopaminergic loss and neuronal cell death, which is not observed to be caused by αSyn fibrils, indicating once again that oligomers, but not insoluble fibrils, are the most toxic species in PD and other neurodegenerative disorders [61,66].

3. Synaptic Dysfunction Leading to Cognitive Impairment

3.1. Synapse Loss

It has been well established that synaptic dysfunction occurs in AD and PD. Less synapses are found in postmortem brains of patients with AD and PD in brain regions underlying clinical manifestations of both diseases [67,68]. However, only recently it became possible to evaluate synaptic alterations in alive people by using synaptic positron emission tomography (PET) imaging. Specifically, the PET tracer [11C]UCB-J for the synaptic vesicle glycoprotein (SV2A), expressed in all synapses and located in synaptic vesicles at presynaptic terminals, was recently used to detect synaptic alterations in vivo of patients with early AD and PD. In AD, PET imaging of SV2A showed prominent reduction synapses in the hippocampus, followed by the entorhinal cortex, parahippocampal cortex, amygdala, lateral temporal cortex, PCC/precuneus, and lateral parietal cortex, but not in the prefrontal cortex, lateral occipital cortex, medial occipital cortex, or pericentral cortex [69]. The synaptic density reductions were maintained after partial volume correction of the PET images, meaning that the effect is not entirely attributed to loss of gray matter tissue. Importantly, there was a correlation between the reduction of SV2A and cognitive impairment. PET studies correlate with accumulated literature that has consistently shown evidence of synaptic loss across brain regions in AD and other neurodegenerative disorders [70]. In PD, PET imaging showed lower SV2A in the substantia nigra, followed by red nucleus and locus coeruleus as well as other clinically relevant areas [71]. Interestingly, neurocognitive assessment in PD did not correlate with SV2A levels likely because cognitive impairment is milder than that observed in AD and thus the range for correlation is narrower.

3.2. Inflammatory Response Effects on Synapses

Although AD and PD were not originally considered inflammatory disorders, neuroinflammation is a critical component in the pathogenesis and progression of cognitive impairment. Neuroinflammation involves activation of microglia and astrocytes, and the subsequent release of cytokine radicals which lead to synaptic loss and damage [72,73]. Particularly, microglia are pivotal in the control of synapse activity by establishing direct contact with neurons, meaning that an inflammatory process at this level has a negative impact on synaptic surveillance and thus, cognitive function. However, whether neuroinflammation is caused by soluble oligomers, the most toxic components in the pathology of AD and PD, is not clear yet. Here we present current studies addressing the association between soluble oligomers and synaptic dysfunction due to neuroinflammation.
Distinct Aβ conformations seem to trigger different magnitudes of microglial activation. As mentioned before, oligomeric (rather than fibrillary) forms of Aβ, are the most neurotoxic aggregates in AD [37,38,74,75,76]. Thus, it has been investigated in vitro and in vivo whether AβOs are also stronger promoters of glial activation. An in vitro glial cell culture exposed to AβOs and fibrillar-Aβ, demonstrated not only that the pro-inflammatory response of the oligomeric form of Aβ was stronger than its fibrillary counterpart, but also that the response was an M1-like phenotype [77]. Complementarily, a murine study, where brain inflammation was induced by different Aβ42 conformers, showed that the lightest AβOs can activate microglial cells and promote a violent inflammatory response, whereas heavier oligomeric and fibrillary Aβ conformations induced less glial activation and poorer inflammatory responses [78]. Another in vivo model demonstrated AβOs promoted stronger neurotoxicity and inflammatory response mediated by NF-κB, when compared to fibrillar-Aβ [79]. All these studies reinforce the idea that AβOs are the most potent activators of microglial cells, and following studies display how this inflammatory response leads to synaptic disruption and sequential neuronal dysfunction. The inflammatory response followed by synaptic disruption and neuronal loss can be clinically translated as memory, language, and visual perception decline, among other forms of cognitive impairment [80]. In animal models AβOs induce inflammatory signaling leading to this cognitive decline manifestations [81,82,83]. For instance, in an acute experimental model in C57BL/6 mice, memory impairment and inflammation were observed after an intracerebroventricular injection of AβOs, suggesting that oligomers interfere with synaptic transmission necessary to establish new memories; again, the fibrillar-Aβ did not produce this effect [32]. The molecular link between cognitive deficit and neuroinflammation lies in the release of cytokines by microglial cells. In one study, purified AβOs from human AD brain tissue were injected in wild type mice to induce microglial inflammation. The inflammatory response of this model was demonstrated when several cytokines at mRNA and protein levels were identified, including Ccl3, CCl4, and Tnf [84]. Other mechanism underlying AβO-induced microglial activation is explained by TLR-4, which likely induces aberrant TNF-α signaling [85,86]. In support of this deleterious role of the inflammatory response, the cognitive decline, induced by the intracerebral injection of AβOs, was reversed by the administration of anti-inflammatory drugs, doxycycline, and TLR-4 antagonists [87]. In another study of intracerebroventricular injection of AβOs in wild type mice, the complement factors C1q, which initiates the classic complement pathway, and C3, were found elevated at the synapse level, which would explain the synapse loss through microglial activation [88]. In addition to inflammation induced by microglia, astrocytosis is another early phenomenon in AD development, but whether AβOs induce astrocytosis remains to be determined.
TauOs are the most neurotoxic tau species involved in the development of cognitive impairment [89,90]. They induce neuroinflammation in AD and frontotemporal lobar dementia through interaction with astrocytes and microglia [91]. A model for the toxic relationship between TauOs and inflammation has been proposed, where TauOs through astrocytes and microglia trigger the release of cytokines, RAGE receptors and their ligand HMGB1. Activation of RAGE signals NF-κB and p38-MAPK pathways, which in turn promote hyperphosphorylation of Tau and subsequent aggregation of more oligomers, and thus, neuronal damage or death and a vicious cycle of chronic neuroinflammation [51,91]. Levels of IL-1β and TNF-α increase significantly when microglia cells are exposed to TauOs [92].
Inflammatory reaction induced by αSyn can take place in the microglia and astrocytes or have a direct effect on neurons. Upon activation of microglia by αSyn, the microglia release neurotoxic factors, including pro-inflammatory cytokines which may lead to neuronal dysfunction [93,94]. As in Aβ and Tau, αSyn self-aggregates to generate αSynOs, protofibrils and fibrils. It is also well established that αSynOs are the main responsible form for cognitive decline in Lewy body dementia and PD. Whether αSynOs contribute more than fibrillar α-syn to neuroinflammation in PD, in the same way as AβOs and TauOs in AD has been a recent area of interest. One study showed, injection of αSynOs in the brain ventricles of wild type mice caused memory impairment through a TLR-2 dependent mechanism, which is closely associated with activation of glial cells in the hippocampus; and contrary to AβOs, TLR-4 was not involved in memory impairment [95,96]. On an age-depending study in mice, αSynOs induced an inflammatory response in microglia cells through release of TNF-α in adult, but not young mice, mimicking the inflammatory response in PD pathology [97]. Besides TLR-2 [95,98], αSynOs has also been implicated in the TLR-1 pathway, in either case, the pro-inflammatory microglial phenotype leads to translocation of NF-κB and increased production of TNF-α and IL-1β through MyD88 dependent mechanisms [99]. The monomeric and fibrillary forms of αSyns are not able to produce an inflammatory response of microglial cells [95,96,98] nor impair memory of tested mice [95]. The cognitive deficit induced by αSynOs is antagonized with pretreatment of anti-inflammatory drugs [95]. In astrocytes, αSynOs induce neuronal cell death in a TLR-4 dependent mechanism by triggering the production of TNF-α and other cytokines by astrocytes [96]. Neurons alone are also directly susceptible to αSynOs induced TLR4-independent toxicity [96]. Additionally, αSynOs can induce the production of reactive oxygen species (ROS) on the presence of free metal ions, resulting in neuronal death [100]. At the peripheral level, neuroinflammation induced by αSynOs also enhances and aggravates cognitive deficits in mice [101].
Gamma-aminobutyric acid (GABA) plays an important role in the communication between neurons and microglia. Microglia and other CNS cells through altered GABA receptors may lead to impaired signaling and thus, lose communication with neurons [102]. GABAergic signaling in microglia decreases the activity of inflammatory mediators NF-κB and p38 MAP kinase and suppress the release of TNF-α and IL-6 outside synapses [103]. This suggests that GABA receptors may be a potential target to reduce the deleterious effects of oligomers on synaptic dysfunction. In support, enhancement of GABAA receptors through drugs such as carbamazepine, phenytoin, and valproic acid have been found to stabilize intracellular Ca2+ levels and thus ameliorate the neurotoxic effects of AβOs [104]. GABAergic signaling across neurons and glial cells that modulate homeostatic plasticity is an area that has been understudied and requires more multidisciplinary efforts from the research community.

3.3. Receptors Involved in Synaptic Dysfunction

Accumulated evidence indicates that AβOs directly activates AMPA receptors [105]. AMPA receptors are complex proteins made by the combination of four principal subunits (GluA1-GluA4) [105,106], and co-assembled auxiliary proteins [107,108], that modulate the gating, permeability, and pharmacology of the channel [108,109,110,111,112]. GluA2-lacking AMPA receptors are permeable to Ca2+ and its excessive activation leads to Ca2+ overload, excitotoxicity, and neurodegeneration [113,114,115,116,117]. Recent evidence from Reinders et al., demonstrated that AβOs cause synaptic failure only in neurons expressing GluA3 subunits [118], and mice with severe AD neuropathology but deficient in GluA3 were cognitively resilient [118], strongly indicating that synaptic vulnerability to AβOs may depend on the stoichiometry of synaptic receptors. This is consistent with human postmortem studies where lower gene expression levels for GluA3 correlated with better cognitive performance in prodromal AD [119]. Similarly, it is increasingly acknowledged that AβOs directly activate heterologously-expressed receptors composed by GRIN1/GluN2A and GRIN1/GluN2B subunits [105,120], which are the most abundant NMDA receptors in mammals’ cortical synapses; however, only the activation of receptors containing GluN2B subunits (GluN2B-NMDA receptors) leads to acute activity-dependent postsynaptic failure [121], Ca2+ dysregulation [122], synaptic depression [123,124], and neurotoxicity in in vitro systems [125,126]. Most likely due to the high Ca2+ permeability of GluN2B-NMDA receptors [127] and their downstream signaling [128]. The clinical significance of GluN2B is reinforced by a multisite postmortem study showing that lower cortical gene expression of GluN2B correlates with better cognitive performance in people diagnosed with prodromal AD [119]. These parallel lines of evidence strongly suggest that levels of expression of GluN2B-NMDA receptors are correlated with synaptic and neuronal vulnerability. Although little is known about the mechanism by which αSyn produce synaptic dysfunction, Trudler at al., recently showed that αSynOs induce Ca2+-dependent release of glutamate from astrocytes leading to a chronic increase of glutamate that activates extrasynaptic NMDA receptors and inducing synaptic loss [129]. αSynOs also bind to NMDA receptors, increasing the synaptic transmission and resulting in membrane damage and LTP impairment [130]. Also, by targeting GluN2A NMDA receptors, αSynOs can induce visual spatial memory impairment [131].
In contrast to AβOs or αSynOs, TauOs have not been demonstrated to directly interact with synaptic receptors. However, tau participates in Aβ mediated toxicity by interacting with Fyn kinase via its amino-terminal projection domain facilitating the NMDA receptors-mediated synaptotoxicity [132]. The role of tau in Aβ toxicity via Fyn-kinase modulation is further supported by studies reporting that absence of tau in dendritic spines prevented the toxic effects of AβOs mediated by GluN2B-NMDA receptors [133], and whereas a reduction in tau levels prevented the cognitive impairment in AD transgenic mice overexpressing Aβ, overexpression of Fyn can enhance their cognitive impairment [134]. Taken together, these results strongly suggest that AβOs and αSynOs may initiate their toxic effects by binding to specific subtypes of AMPA and NMDA receptors and/or the proteins they form complexes with; and tau is an important element for downstream signaling of the neurotoxicity. It follows that, differential expression of those particular targets may provide synaptic protection and underlie cognitive resilience in humans.

3.4. Impaired Excitatory/Inhibitory Ratio

Hyperexcitability of cortical and hippocampal circuits and 87-fold increase in seizures incidence in the AD population is well documented [135], particularly in early-onset familial AD [136]. Convulsive seizures occur in approximately 7–21% of sporadic AD patients [137,138], 31% of patients with PS2 mutations [139] and 56% of patients with APP duplications [140]. These data do not account for hidden hyperexcitability status that occurs early in AD pathogenesis [141]. Since oligomers act mostly on excitatory synapses leading to dysfunction first and synaptic loss later, a large question in the field is how, overall reduction of excitatory inputs leads to hyperexcitability in the AD brain. Although the causes of network hyperactivity are still under investigation by many labs; early studies in animal models suggest that impaired inhibition is a potential mechanism for network hyperactivity [142,143]. Impairment of interneuron activity with changes in their intrinsic properties have been reported in the mice models of amyloidosis [144]. Interneuron deficits reduces neurogenesis and neuronal maturation in the hilus of the hippocampus [145] and leads to age and tau-dependent learning and memory deficits [146]. Potentiation of GABA receptors by pentobarbital restores some of the deficits observed by GABAergic impairment [145]. Our initial studies transplanting human receptors and recording their electrical activity observed a dramatic reduction of GABAA receptors in AD [147]. This severe reduction of gene expression was later confirmed by other groups using high throughput microarray technology [148] and provided evidence that in addition to excitatory synaptic loss, inhibitory synapses were also affected in AD.
Because the abundance, activity, and strength of excitatory and inhibitory synapses in the neocortex are highly correlated in their amplitude and time domains [149,150,151,152]. The “global E/I ratio” defined as equal average amounts of postsynaptic AMPA and GABAA receptor activities is essential for maintaining stability of cortical neurons [150,153] and is tightly regulated within a narrow range by inhibitory plasticity following any excitatory change produced by ongoing sensory experiences and activity-dependent plasticity [154]. Taking in consideration the synaptic alterations found in AD, the immediate question is how the global E/I ratio is affected in AD? Interestingly, functional alterations in the default mode network (DMN) correlate with cognitive impairment in AD and it is early affected in AD [149]. The DMN spans several brain regions including the parietal cortex, importantly, baseline DMN activity is increased in AD and fails to deactivate during cognitive tasks, suggesting that the DMN is abnormally and continuously hyperactive in AD [149,155]. By using fluorescence deconvolution tomography (FDT), where automated systems are used to quantify 30,000 immunolabeled elements within the size constraints of synapses from 3D reconstructions of image z-stacks [156,157,158,159], it was found a larger E/I ratio (il-PSD-95 for excitatory/il-gephyrin for inhibitory synapses), in AD compared to non-demented controls. Therefore, our results indicate an overlooked link between increased regional synaptic E/I ratio, cortical hyperexcitability, and a potential enhancement of activity-dependent amyloidosis [160,161]. The pro-excitatory E/I ratio in AD was confirmed by directly recording the activity of AMPA and GABAA receptors microtransplanted from AD parietal cortex [162]. Importantly the E/I ratio was preserved in NDAN people indicating that preservation of the E/I correlates with the preservation of cognition even in presence of neuropathologic change [163].
Taken together, these data suggest that hyperexcitability in AD patients result from impaired GABAergic inhibition which leads to cortical excess of excitatory synaptic inputs. Notably, patients expressing the homomeric APOE4 allele are at the highest genetic risk to develop late-onset AD, and GABAergic interneurons and synapses are also deficient in patients with APOE4 AD. Bumetanide is a diuretic acting on chloride transporters modulating the E/I ratio via pathways enriched in GABAergic signaling. Importantly, bumetanide was found to be protective against AD as evaluated in mouse models, and more importantly through an electronic health record, which showed that the prevalence of AD in patients older than 65 years old was significantly reduced in those taking bumetanide [164].
In PD, there is also evidence of E/I imbalance, however the effects are more brain region specific. Loss of substantia nigra pars compacta (SNpc) dopaminergic neurons lead to hyperactivity of the globus pallidum and excessive inhibition onto the motor thalamus, this ultimately leads to the abnormal movement manifestation of PD [165]. While it is still not clear how aSynOs relate to the loss of SNpc, it is clear that a complex series of pathways interact to produce early dysfunction of SNpc [166]. Recent studies in human organoids with mutations linked to PD show that the E/I is altered in PD with lower inhibition and reduced levels of neurosteroid allopregrananolone even before the presence of neuropathology is observed [167]. This indicates that E/I alterations precede pathology, and its correction may have disease-modifying effects.

4. Conclusions

Oligomeric forms of Aβ, tau and aSyn are the most toxic species affecting synapses leading to synaptic dysfunction and altered neuronal communication in brain regions vulnerable to the neuropathology. The effects of oligomers precede the presence of deposits and seem to be associated to early changes in excitatory and inhibitory synapses. Therefore, oligomers seem to produce a “double hit” on synapses (Figure 1). First, they lead to calcium dys-homeostasis by binding directly to excitatory receptors and leading to a first wave of hyperexcitability, then producing GABAergic dysfunction by a mechanism that is still not understood, which leads to a second chronic wave of hyperexcitablity that ultimately leads to neuronal loss and hypoactivity. Understanding the regional and temporal relationships between oligomers, synaptic targets and E/I balance is a critical need in the field.

Author Contributions

Conceptualization, A.L.; investigation, B.A.G. and A.L.; writing—original draft preparation, B.A.G. and A.L.; writing—review and editing, B.A.G. and A.L; visualization, B.A.G. and A.L.; supervision, A.L.; funding acquisition, A.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by NIA, grants number AG070255 and AG073133 to A.L.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Julia Neuhart (University of Minnesota, Morris) for her technical assistance with the references.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. World Health Organization. Available online: https://www.who.int/news-room/fact-sheets/detail/dementia (accessed on 21 June 2022).
  2. Parkinson’s Disease Statistics. Available online: https://parkinsonsnewstoday.com/parkinsons-disease-statistics/ (accessed on 5 June 2022).
  3. Soria Lopez, J.A.; González, H.M.; Léger, G.C. Alzheimer’s Disease. Handb. Clin. Neurol. 2019, 167, 231–255. [Google Scholar] [CrossRef] [PubMed]
  4. Alzheimer’s Association. 2016 Alzheimer’s Disease Facts and Figures. Alzheimer’s Dement. 2016, 12, 459–509. [Google Scholar] [CrossRef] [PubMed]
  5. DeTure, M.A.; Dickson, D.W. The Neuropathological Diagnosis of Alzheimer’s Disease. Mol. Neurodegener. 2019, 14, 32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Jack, C.R.; Bennett, D.A.; Blennow, K.; Carrillo, M.C.; Dunn, B.; Haeberlein, S.B.; Holtzman, D.M.; Jagust, W.; Jessen, F.; Karlawish, J.; et al. NIA-AA Research Framework: Toward a Biological Definition of Alzheimer’s Disease. Alzheimer’s Dement. 2018, 14, 535–562. [Google Scholar] [CrossRef] [PubMed]
  7. Gralle, M.; Ferreira, S.T. Structure and Functions of the Human Amyloid Precursor Protein: The Whole Is More than the Sum of Its Parts. Prog. Neurobiol. 2007, 82, 11–32. [Google Scholar] [CrossRef]
  8. Ashraf, G.; Greig, N.; Khan, T.; Hassan, I.; Tabrez, S.; Shakil, S.; Sheikh, I.; Zaidi, S.; Akram, M.; Jabir, N.; et al. Protein Misfolding and Aggregation in Alzheimer’s Disease and Type 2 Diabetes Mellitus. CNS Neurol. Disord. Drug Targets 2014, 13, 1280–1293. [Google Scholar] [CrossRef]
  9. Scholz, T.; Mandelkow, E. Transport and Diffusion of Tau Protein in Neurons. Cell. Mol. Life Sci. 2014, 71, 3139–3150. [Google Scholar] [CrossRef]
  10. Chaudhary, A.R.; Berger, F.; Berger, C.L.; Hendricks, A.G. Tau Directs Intracellular Trafficking by Regulating the Forces Exerted by Kinesin and Dynein Teams. Traffic 2018, 19, 111–121. [Google Scholar] [CrossRef] [Green Version]
  11. Afreen, S.; Riherd Methner, D.N.; Ferreira, A. Tau45-230 Association with the Cytoskeleton and Membrane-Bound Organelles: Functional Implications in Neurodegeneration. Neuroscience 2017, 362, 104–117. [Google Scholar] [CrossRef]
  12. Rudenko, L.K.; Wallrabe, H.; Periasamy, A.; Siller, K.H.; Svindrych, Z.; Seward, M.E.; Best, M.N.; Bloom, G.S. Intraneuronal Tau Misfolding Induced by Extracellular Amyloid-β Oligomers. J. Alzheimer’s Dis. 2019, 71, 1125–1138. [Google Scholar] [CrossRef]
  13. Lee, H.-G.; Perry, G.; Moreira, P.I.; Garrett, M.R.; Liu, Q.; Zhu, X.; Takeda, A.; Nunomura, A.; Smith, M.A. Tau Phosphorylation in Alzheimer’s Disease: Pathogen or Protector? Trends Mol. Med. 2005, 11, 164–169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Yu, H.; Wu, J. Amyloid-β: A Double Agent in Alzheimer’s Disease? Biomed. Pharmacother. 2021, 139, 111575. [Google Scholar] [CrossRef] [PubMed]
  15. Nelson, P.T.; Alafuzoff, I.; Bigio, E.H.; Bouras, C.; Braak, H.; Cairns, N.J.; Castellani, R.J.; Crain, B.J.; Davies, P.; del Tredici, K.; et al. Correlation of Alzheimer Disease Neuropathologic Changes with Cognitive Status: A Review of the Literature. J. Neuropathol. Exp. Neurol. 2012, 71, 362–381. [Google Scholar] [CrossRef]
  16. Postuma, R.B.; Berg, D.; Stern, M.; Poewe, W.; Olanow, C.W.; Oertel, W.; Obeso, J.; Marek, K.; Litvan, I.; Lang, A.E.; et al. MDS Clinical Diagnostic Criteria for Parkinson’s Disease. Mov. Disord. 2015, 30, 1591–1601. [Google Scholar] [CrossRef]
  17. Raza, C.; Anjum, R.; Shakeel, N.U.A. Parkinson’s Disease: Mechanisms, Translational Models and Management Strategies. Life Sci. 2019, 226, 77–90. [Google Scholar] [CrossRef] [PubMed]
  18. Antony, P.M.A.; Diederich, N.J.; Krüger, R.; Balling, R. The Hallmarks of Parkinson’s Disease. FEBS J. 2013, 280, 5981–5993. [Google Scholar] [CrossRef] [Green Version]
  19. Braak, H.; de Vos, R.A.I.; Bohl, J.; del Tredici, K. Gastric α-Synuclein Immunoreactive Inclusions in Meissner’s and Auerbach’s Plexuses in Cases Staged for Parkinson’s Disease-Related Brain Pathology. Neurosci. Lett. 2006, 396, 67–72. [Google Scholar] [CrossRef]
  20. Orimo, S.; Uchihara, T.; Nakamura, A.; Mori, F.; Kakita, A.; Wakabayashi, K.; Takahashi, H. Axonal α-Synuclein Aggregates Herald Centripetal Degeneration of Cardiac Sympathetic Nerve in Parkinson’s Disease. Brain 2008, 131, 642–650. [Google Scholar] [CrossRef] [Green Version]
  21. Glenner, G.G.; Wong, C.W. Alzheimer’s Disease: Initial Report of the Purification and Characterization of a Novel Cerebrovascular Amyloid Protein. Biochem. Biophys. Res. Commun. 1984, 120, 885–890. [Google Scholar] [CrossRef]
  22. Goldberg, M.S.; Lansbury, P.T. Is There a Cause-and-Effect Relationship between α-Synuclein Fibrillization and Parkinson’s Disease? Nat. Cell Biol. 2000, 2, E115–E119. [Google Scholar] [CrossRef]
  23. Nguyen, P.H.; Ramamoorthy, A.; Sahoo, B.R.; Zheng, J.; Faller, P.; Straub, J.E.; Dominguez, L.; Shea, J.-E.; Dokholyan, N.V.; de Simone, A.; et al. Amyloid Oligomers: A Joint Experimental/Computational Perspective on Alzheimer’s Disease, Parkinson’s Disease, Type II Diabetes, and Amyotrophic Lateral Sclerosis. Chem. Rev. 2021, 121, 2545–2647. [Google Scholar] [CrossRef] [PubMed]
  24. Cirrito, J.R.; Yamada, K.A.; Finn, M.B.; Sloviter, R.S.; Bales, K.R.; May, P.C.; Schoepp, D.D.; Paul, S.M.; Mennerick, S.; Holtzman, D.M. Synaptic Activity Regulates Interstitial Fluid Amyloid-β Levels in Vivo. Neuron 2005, 48, 913–922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Pooler, A.M.; Phillips, E.C.; Lau, D.H.W.; Noble, W.; Hanger, D.P. Physiological Release of Endogenous Tau Is Stimulated by Neuronal Activity. EMBO Rep. 2013, 14, 389–394. [Google Scholar] [CrossRef] [PubMed]
  26. Verma, M.; Vats, A.; Taneja, V. Toxic Species in Amyloid Disorders: Oligomers or Mature Fibrils. Ann. Indian Acad. Neurol. 2015, 18, 138–145. [Google Scholar] [CrossRef] [PubMed]
  27. Forloni, G.; Artuso, V.; la Vitola, P.; Balducci, C. Oligomeropathies and Pathogenesis of Alzheimer and Parkinson’s Diseases. Mov. Disord. 2016, 31, 771–781. [Google Scholar] [CrossRef] [PubMed]
  28. Ono, K. The Oligomer Hypothesis in α-Synucleinopathy. Neurochem. Res. 2017, 42, 3362–3371. [Google Scholar] [CrossRef] [PubMed]
  29. Kuo, Y.M.; Emmerling, M.R.; Vigo-Pelfrey, C.; Kasunic, T.C.; Kirkpatrick, J.B.; Murdoch, G.H.; Ball, M.J.; Roher, A.E. Water-Soluble Aβ (N-40, N-42) Oligomers in Normal and Alzheimer Disease Brains. J. Biol. Chem. 1996, 271, 4077–4081. [Google Scholar] [CrossRef] [Green Version]
  30. Mucke, L.; Selkoe, D.J. Neurotoxicity of Amyloid β-Protein: Synaptic and Network Dysfunction. Cold Spring Harb. Perspect. Med. 2012, 2, a006338. [Google Scholar] [CrossRef] [Green Version]
  31. Walsh, D.M.; Klyubin, I.; Fadeeva, J.V.; Cullen, W.K.; Anwyl, R.; Wolfe, M.S.; Rowan, M.J.; Selkoe, D.J. Naturally Secreted Oligomers of Amyloid β Protein Potently Inhibit Hippocampal Long-Term Potentiation in Vivo. Nature 2002, 416, 535–539. [Google Scholar] [CrossRef]
  32. Balducci, C.; Beeg, M.; Stravalaci, M.; Bastone, A.; Sclip, A.; Biasini, E.; Tapella, L.; Colombo, L.; Manzoni, C.; Borsello, T.; et al. Synthetic Amyloid-β Oligomers Impair Long-Term Memory Independently of Cellular Prion Protein. Proc. Natl. Acad. Sci. USA 2010, 107, 2295–2300. [Google Scholar] [CrossRef] [Green Version]
  33. Haass, C.; Selkoe, D.J. Soluble Protein Oligomers in Neurodegeneration: Lessons from the Alzheimer’s Amyloid β-Peptide. Nat. Rev. Mol. Cell Biol. 2007, 8, 101–112. [Google Scholar] [CrossRef] [PubMed]
  34. Wilcox, K.C.; Lacor, P.N.; Pitt, J.; Klein, W.L. Aβ Oligomer-Induced Synapse Degeneration in Alzheimer’s Disease. Cell. Mol. Neurobiol. 2011, 31, 939–948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Ward, S.M.; Himmelstein, D.S.; Lancia, J.K.; Binder, L.I. Tau Oligomers and Tau Toxicity in Neurodegenerative Disease. Biochem. Soc. Trans. 2012, 40, 667–671. [Google Scholar] [CrossRef] [Green Version]
  36. Kayed, R.; Lasagna-Reeves, C.A. Molecular Mechanisms of Amyloid Oligomers Toxicity. J. Alzheimer’s Dis. 2013, 33 (Suppl. 1), S67–S78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Shankar, G.M.; Li, S.; Mehta, T.H.; Garcia-Munoz, A.; Shepardson, N.E.; Smith, I.; Brett, F.M.; Farrell, M.A.; Rowan, M.J.; Lemere, C.A.; et al. Amyloid-β Protein Dimers Isolated Directly from Alzheimer’s Brains Impair Synaptic Plasticity and Memory. Nat. Med. 2008, 14, 837–842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Lacor, P.N.; Buniel, M.C.; Furlow, P.W.; Clemente, A.S.; Velasco, P.T.; Wood, M.; Viola, K.L.; Klein, W.L. Aβ Oligomer-Induced Aberrations in Synapse Composition, Shape, and Density Provide a Molecular Basis for Loss of Connectivity in Alzheimer’s Disease. J. Neurosci. 2007, 27, 796–807. [Google Scholar] [CrossRef] [PubMed]
  39. Pham, E.; Crews, L.; Ubhi, K.; Hansen, L.; Adame, A.; Cartier, A.; Salmon, D.; Galasko, D.; Michael, S.; Savas, J.N.; et al. Progressive Accumulation of Amyloid-β Oligomers in Alzheimer’s Disease and in Amyloid Precursor Protein Transgenic Mice Is Accompanied by Selective Alterations in Synaptic Scaffold Proteins. FEBS J. 2010, 277, 3051–3067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Lesné, S.E.; Sherman, M.A.; Grant, M.; Kuskowski, M.; Schneider, J.A.; Bennett, D.A.; Ashe, K.H. Brain Amyloid-β Oligomers in Ageing and Alzheimer’s Disease. Brain 2013, 136, 1383–1398. [Google Scholar] [CrossRef] [Green Version]
  41. Lesné, S.; Koh, M.T.; Kotilinek, L.; Kayed, R.; Glabe, C.G.; Yang, A.; Gallagher, M.; Ashe, K.H. A Specific Amyloid-β Protein Assembly in the Brain Impairs Memory. Nature 2006, 440, 352–357. [Google Scholar] [CrossRef]
  42. Amar, F.; Sherman, M.A.; Rush, T.; Larson, M.; Boyle, G.; Chang, L.; Götz, J.; Buisson, A.; Lesné, S.E. The Amyloid-β Oligomer Aβ*56 Induces Specific Alterations in Neuronal Signaling That Lead to Tau Phosphorylation and Aggregation. Sci. Signal. 2017, 10, eaal2021. [Google Scholar] [CrossRef] [Green Version]
  43. Bjorklund, N.L.; Reese, L.C.; Sadagoparamanujam, V.-M.; Ghirardi, V.; Woltjer, R.L.; Taglialatela, G. Absence of Amyloid β Oligomers at the Postsynapse and Regulated Synaptic Zn2+ in Cognitively Intact Aged Individuals with Alzheimer’s Disease Neuropathology. Mol. Neurodegener. 2012, 7, 23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Meng, X.; Li, T.; Wang, X.; Lv, X.; Sun, Z.; Zhang, J.; Su, F.; Kang, S.; Kim, S.; An, S.S.A.; et al. Association between Increased Levels of Amyloid-β Oligomers in Plasma and Episodic Memory Loss in Alzheimer’s Disease. Alzheimer’s Res. Ther. 2019, 11, 89. [Google Scholar] [CrossRef] [PubMed]
  45. McLean, C.A.; Cherny, R.A.; Fraser, F.W.; Fuller, S.J.; Smith, M.J.; Beyreuther, K.; Bush, A.I.; Masters, C.L. Soluble Pool of Aβ Amyloid as a Determinant of Severity of Neurodegeneration in Alzheimer’s Disease. Ann. Neurol. 1999, 46, 860–866. [Google Scholar] [CrossRef]
  46. Mc Donald, J.M.; Savva, G.M.; Brayne, C.; Welzel, A.T.; Forster, G.; Shankar, G.M.; Selkoe, D.J.; Ince, P.G.; Walsh, D.M. The Presence of Sodium Dodecyl Sulphate-Stable Aβ Dimers Is Strongly Associated with Alzheimer-Type Dementia. Brain 2010, 133, 1328–1341. [Google Scholar] [CrossRef] [Green Version]
  47. Hou, T.-T.; Yang, H.-Y.; Wang, W.; Wu, Q.-Q.; Tian, Y.-R.; Jia, J.-P. Sulforaphane Inhibits the Generation of Amyloid-β Oligomer and Promotes Spatial Learning and Memory in Alzheimer’s Disease (PS1V97L) Transgenic Mice. J. Alzheimer’s Dis. 2018, 62, 1803–1813. [Google Scholar] [CrossRef] [PubMed]
  48. Kim, J.; Chakrabarty, P.; Hanna, A.; March, A.; Dickson, D.W.; Borchelt, D.R.; Golde, T.; Janus, C. Normal Cognition in Transgenic BRI2-Aβ Mice. Mol. Neurodegener. 2013, 8, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Melnikova, T.; Fromholt, S.; Kim, H.; Lee, D.; Xu, G.; Price, A.; Moore, B.D.; Golde, T.E.; Felsenstein, K.M.; Savonenko, A.; et al. Reversible Pathologic and Cognitive Phenotypes in an Inducible Model of Alzheimer-Amyloidosis. J. Neurosci. 2013, 33, 3765–3779. [Google Scholar] [CrossRef] [Green Version]
  50. Maeda, S.; Sahara, N.; Saito, Y.; Murayama, S.; Ikai, A.; Takashima, A. Increased Levels of Granular Tau Oligomers: An Early Sign of Brain Aging and Alzheimer’s Disease. Neurosci. Res. 2006, 54, 197–201. [Google Scholar] [CrossRef]
  51. Gaikwad, S.; Puangmalai, N.; Bittar, A.; Montalbano, M.; Garcia, S.; McAllen, S.; Bhatt, N.; Sonawane, M.; Sengupta, U.; Kayed, R. Tau Oligomer Induced HMGB1 Release Contributes to Cellular Senescence and Neuropathology Linked to Alzheimer’s Disease and Frontotemporal Dementia. Cell Rep. 2021, 36, 109419. [Google Scholar] [CrossRef]
  52. Patterson, K.R.; Remmers, C.; Fu, Y.; Brooker, S.; Kanaan, N.M.; Vana, L.; Ward, S.; Reyes, J.F.; Philibert, K.; Glucksman, M.J.; et al. Characterization of Prefibrillar Tau Oligomers in Vitro and in Alzheimer Disease. J. Biol. Chem. 2011, 286, 23063–23076. [Google Scholar] [CrossRef] [Green Version]
  53. Lasagna-Reeves, C.A.; Castillo-Carranza, D.L.; Sengupta, U.; Guerrero-Munoz, M.J.; Kiritoshi, T.; Neugebauer, V.; Jackson, G.R.; Kayed, R. Alzheimer Brain-Derived Tau Oligomers Propagate Pathology from Endogenous Tau. Sci. Rep. 2012, 2, 700. [Google Scholar] [CrossRef]
  54. Paleologou, K.E.; Kragh, C.L.; Mann, D.M.A.; Salem, S.A.; Al-Shami, R.; Allsop, D.; Hassan, A.H.; Jensen, P.H.; El-Agnaf, O.M.A. Detection of Elevated Levels of Soluble α-Synuclein Oligomers in Post-Mortem Brain Extracts from Patients with Dementia with Lewy Bodies. Brain 2009, 132, 1093–1101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Gerson, J.E.; Farmer, K.M.; Henson, N.; Castillo-Carranza, D.L.; Carretero Murillo, M.; Sengupta, U.; Barrett, A.; Kayed, R. Tau Oligomers Mediate α-Synuclein Toxicity and Can Be Targeted by Immunotherapy. Mol. Neurodegener. 2018, 13, 13. [Google Scholar] [CrossRef] [PubMed]
  56. Jiang, L.; Ash, P.E.A.; Maziuk, B.F.; Ballance, H.I.; Boudeau, S.; Abdullatif, A.A.; Orlando, M.; Petrucelli, L.; Ikezu, T.; Wolozin, B. TIA1 Regulates the Generation and Response to Toxic Tau Oligomers. Acta Neuropathol. 2019, 137, 259–277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Jiang, L.; Lin, W.; Zhang, C.; Ash, P.E.A.; Verma, M.; Kwan, J.; van Vliet, E.; Yang, Z.; Cruz, A.L.; Boudeau, S.; et al. Interaction of Tau with HNRNPA2B1 and N6-Methyladenosine RNA Mediates the Progression of Tauopathy. Mol. Cell. 2021, 81, 4209–4227.e12. [Google Scholar] [CrossRef] [PubMed]
  58. Montalbano, M.; McAllen, S.; Cascio, F.L.; Sengupta, U.; Garcia, S.; Bhatt, N.; Ellsworth, A.; Heidelman, E.A.; Johnson, O.D.; Doskocil, S.; et al. TDP-43 and Tau Oligomers in Alzheimer’s Disease, Amyotrophic Lateral Sclerosis, and Frontotemporal Dementia. Neurobiol. Dis. 2020, 146, 105130. [Google Scholar] [CrossRef] [PubMed]
  59. Montalbano, M.; McAllen, S.; Sengupta, U.; Puangmalai, N.; Bhatt, N.; Ellsworth, A.; Kayed, R. Tau Oligomers Mediate Aggregation of RNA-Binding Proteins Musashi1 and Musashi2 Inducing Lamin Alteration. Aging Cell 2019, 18, e13035. [Google Scholar] [CrossRef] [Green Version]
  60. Surguchev, A.; Surguchov, A. Effect of α-Synuclein on Membrane Permeability and Synaptic Transmission: A Clue to Neurodegeneration? J. Neurochem. 2015, 132, 619–621. [Google Scholar] [CrossRef] [Green Version]
  61. Winner, B.; Jappelli, R.; Maji, S.K.; Desplats, P.A.; Boyer, L.; Aigner, S.; Hetzer, C.; Loher, T.; Vilar, M.; Campioni, S.; et al. In Vivo Demonstration That α-Synuclein Oligomers Are Toxic. Proc. Natl. Acad. Sci. USA 2011, 108, 4194–4199. [Google Scholar] [CrossRef] [Green Version]
  62. Surmeier, D.J.; Schumacker, P.T.; Guzman, J.D.; Ilijic, E.; Yang, B.; Zampese, E. Calcium and Parkinson’s Disease. Biochem. Biophys. Res. Commun. 2017, 483, 1013–1019. [Google Scholar] [CrossRef] [Green Version]
  63. Subramaniam, S.R.; Chesselet, M.-F. Mitochondrial Dysfunction and Oxidative Stress in Parkinson’s Disease. Prog. Neurobiol. 2013, 106–107, 17–32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Zampese, E.; Surmeier, D.J. Calcium, Bioenergetics, and Parkinson’s Disease. Cells 2020, 9, 2045. [Google Scholar] [CrossRef] [PubMed]
  65. Poewe, W.; Seppi, K.; Tanner, C.M.; Halliday, G.M.; Brundin, P.; Volkmann, J.; Schrag, A.-E.; Lang, A.E. Parkinson Disease. Nat. Rev. Dis. Primers 2017, 3, 17013. [Google Scholar] [CrossRef] [PubMed]
  66. Bigi, A.; Ermini, E.; Chen, S.W.; Cascella, R.; Cecchi, C. Exploring the Release of Toxic Oligomers from α-Synuclein Fibrils with Antibodies and STED Microscopy. Life 2021, 11, 431. [Google Scholar] [CrossRef]
  67. Terry, R.D.; Masliah, E.; Salmon, D.P.; Butters, N.; DeTeresa, R.; Hill, R.; Hansen, L.A.; Katzman, R. Physical Basis of Cognitive Alterations in Alzheimer’s Disease: Synapse Loss Is the Major Correlate of Cognitive Impairment. Ann. Neurol. 1991, 30, 572–580. [Google Scholar] [CrossRef]
  68. Schulz-Schaeffer, W.J. The Synaptic Pathology of α-Synuclein Aggregation in Dementia with Lewy Bodies, Parkinson’s Disease and Parkinson’s Disease Dementia. Acta Neuropathol. 2010, 120, 131–143. [Google Scholar] [CrossRef] [Green Version]
  69. Mecca, A.P.; Chen, M.K.; O’Dell, R.S.; Naganawa, M.; Toyonaga, T.; Godek, T.A.; Harris, J.E.; Bartlett, H.H.; Zhao, W.; Nabulsi, N.B.; et al. In Vivo Measurement of Widespread Synaptic Loss in Alzheimer’s Disease with SV2A PET. Alzheimer’s Dement. 2020, 16, 974–982. [Google Scholar] [CrossRef]
  70. Wilde, M.C.; Overk, C.R.; Sijben, J.W.; Masliah, E. Meta-analysis of Synaptic Pathology in Alzheimer’s Disease Reveals Selective Molecular Vesicular Machinery Vulnerability. Alzheimer’s Dement. 2016, 12, 633–644. [Google Scholar] [CrossRef] [Green Version]
  71. Matuskey, D.; Tinaz, S.; Wilcox, K.C.; Naganawa, M.; Toyonaga, T.; Dias, M.; Henry, S.; Pittman, B.; Ropchan, J.; Nabulsi, N.; et al. Synaptic Changes in Parkinson Disease Assessed with in Vivo Imaging. Ann. Neurol. 2020, 87, 329–338. [Google Scholar] [CrossRef] [Green Version]
  72. Forloni, G.; Balducci, C. Alzheimer’s Disease, Oligomers, and Inflammation. J. Alzheimer’s Dis. 2018, 62, 1261–1276. [Google Scholar] [CrossRef] [Green Version]
  73. Heneka, M.T.; Carson, M.J.; El Khoury, J.; Landreth, G.E.; Brosseron, F.; Feinstein, D.L.; Jacobs, A.H.; Wyss-Coray, T.; Vitorica, J.; Ransohoff, R.M.; et al. Neuroinflammation in Alzheimer’s Disease. Lancet Neurol. 2015, 14, 388–405. [Google Scholar] [CrossRef] [Green Version]
  74. Dahlgren, K.N.; Manelli, A.M.; Stine, W.B.; Baker, L.K.; Krafft, G.A.; LaDu, M.J. Oligomeric and Fibrillar Species of Amyloid-β Peptides Differentially Affect Neuronal Viability. J. Biol. Chem. 2002, 277, 32046–32053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Ahmed, M.; Davis, J.; Aucoin, D.; Sato, T.; Ahuja, S.; Aimoto, S.; Elliott, J.I.; van Nostrand, W.E.; Smith, S.O. Structural Conversion of Neurotoxic Amyloid-β1–42 Oligomers to Fibrils. Nat. Struct. Mol. Biol. 2010, 17, 561–567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Walsh, D.M.; Selkoe, D.J. Aβ Oligomers—A Decade of Discovery. J. Neurochem. 2007, 101, 1172–1184. [Google Scholar] [CrossRef] [PubMed]
  77. Michelucci, A.; Heurtaux, T.; Grandbarbe, L.; Morga, E.; Heuschling, P. Characterization of the Microglial Phenotype under Specific Pro-Inflammatory and Anti-Inflammatory Conditions: Effects of Oligomeric and Fibrillar Amyloid-β. J. Neuroimmunol. 2009, 210, 3–12. [Google Scholar] [CrossRef]
  78. Heurtaux, T.; Michelucci, A.; Losciuto, S.; Gallotti, C.; Felten, P.; Dorban, G.; Grandbarbe, L.; Morga, E.; Heuschling, P. Microglial Activation Depends on β-Amyloid Conformation: Role of the Formylpeptide Receptor 2. J. Neurochem. 2010, 114, 576–586. [Google Scholar] [CrossRef]
  79. He, Y.; Zheng, M.-M.; Ma, Y.; Han, X.-J.; Ma, X.-Q.; Qu, C.-Q.; Du, Y.-F. Soluble Oligomers and Fibrillar Species of Amyloid β-Peptide Differentially Affect Cognitive Functions and Hippocampal Inflammatory Response. Biochem. Biophys. Res. Commun. 2012, 429, 125–130. [Google Scholar] [CrossRef]
  80. Rao, J.S.; Kellom, M.; Kim, H.-W.; Rapoport, S.I.; Reese, E.A. Neuroinflammation and Synaptic Loss. Neurochem. Res. 2012, 37, 903–910. [Google Scholar] [CrossRef]
  81. Lourenco, M.V.; Clarke, J.R.; Frozza, R.L.; Bomfim, T.R.; Forny-Germano, L.; Batista, A.F.; Sathler, L.B.; Brito-Moreira, J.; Amaral, O.B.; Silva, C.A.; et al. TNF-α Mediates PKR-Dependent Memory Impairment and Brain IRS-1 Inhibition Induced by Alzheimer’s β-Amyloid Oligomers in Mice and Monkeys. Cell Metab. 2013, 18, 831–843. [Google Scholar] [CrossRef] [Green Version]
  82. Forny-Germano, L.; e Silva, N.L.M.; Batista, A.F.; Brito-Moreira, J.; Gralle, M.; Boehnke, S.E.; Coe, B.C.; Lablans, A.; Marques, S.A.; Martinez, A.M.B.; et al. Alzheimer’s Disease-like Pathology Induced by Amyloid-β Oligomers in Nonhuman Primates. J. Neurosci. 2014, 34, 13629–13643. [Google Scholar] [CrossRef]
  83. Ledo, J.H.; Azevedo, E.P.; Clarke, J.R.; Ribeiro, F.C.; Figueiredo, C.P.; Foguel, D.; de Felice, F.G.; Ferreira, S.T. Amyloid-β Oligomers Link Depressive-like Behavior and Cognitive Deficits in Mice. Mol. Psychiatry 2013, 18, 1053–1054. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Xu, H.; Gelyana, E.; Rajsombath, M.; Yang, T.; Li, S.; Selkoe, D. Environmental Enrichment Potently Prevents Microglia-Mediated Neuroinflammation by Human Amyloid β-Protein Oligomers. J. Neurosci. 2016, 36, 9041–9056. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Ledo, J.H.; Azevedo, E.P.; Beckman, D.; Ribeiro, F.C.; Santos, L.E.; Razolli, D.S.; Kincheski, G.C.; Melo, H.M.; Bellio, M.; Teixeira, A.L.; et al. Cross Talk Between Brain Innate Immunity and Serotonin Signaling Underlies Depressive-Like Behavior Induced by Alzheimer’s Amyloid-β Oligomers in Mice. J. Neurosci. 2016, 36, 12106–12116. [Google Scholar] [CrossRef] [PubMed]
  86. Balducci, C.; Frasca, A.; Zotti, M.; la Vitola, P.; Mhillaj, E.; Grigoli, E.; Iacobellis, M.; Grandi, F.; Messa, M.; Colombo, L.; et al. Toll-like Receptor 4-Dependent Glial Cell Activation Mediates the Impairment in Memory Establishment Induced by β-Amyloid Oligomers in an Acute Mouse Model of Alzheimer’s Disease. Brain Behav. Immun. 2017, 60, 188–197. [Google Scholar] [CrossRef]
  87. Balducci, C.; Forloni, G. Doxycycline for Alzheimer’s Disease: Fighting β-Amyloid Oligomers and Neuroinflammation. Front. Pharmacol. 2019, 10, 738. [Google Scholar] [CrossRef] [Green Version]
  88. Hong, S.; Beja-Glasser, V.F.; Nfonoyim, B.M.; Frouin, A.; Li, S.; Ramakrishnan, S.; Merry, K.M.; Shi, Q.; Rosenthal, A.; Barres, B.A.; et al. Complement and Microglia Mediate Early Synapse Loss in Alzheimer Mouse Models. Science 2016, 352, 712–716. [Google Scholar] [CrossRef] [Green Version]
  89. Lasagna-Reeves, C.A.; Castillo-Carranza, D.L.; Sengupta, U.; Sarmiento, J.; Troncoso, J.; Jackson, G.R.; Kayed, R. Identification of Oligomers at Early Stages of Tau Aggregation in Alzheimer’s Disease. FASEB J. 2012, 26, 1946–1959. [Google Scholar] [CrossRef] [Green Version]
  90. Shafiei, S.S.; Guerrero-Muñoz, M.J.; Castillo-Carranza, D.L. Tau Oligomers: Cytotoxicity, Propagation, and Mitochondrial Damage. Front. Aging Neurosci. 2017, 9, 83. [Google Scholar] [CrossRef] [Green Version]
  91. Nilson, A.N.; English, K.C.; Gerson, J.E.; Barton Whittle, T.; Nicolas Crain, C.; Xue, J.; Sengupta, U.; Castillo-Carranza, D.L.; Zhang, W.; Gupta, P.; et al. Tau Oligomers Associate with Inflammation in the Brain and Retina of Tauopathy Mice and in Neurodegenerative Diseases. J. Alzheimer’s Dis. 2017, 55, 1083–1099. [Google Scholar] [CrossRef] [Green Version]
  92. Sun, X.-Y.; Li, L.-J.; Dong, Q.-X.; Zhu, J.; Huang, Y.-R.; Hou, S.-J.; Yu, X.-L.; Liu, R.-T. Rutin Prevents Tau Pathology and Neuroinflammation in a Mouse Model of Alzheimer’s Disease. J. Neuroinflamm. 2021, 18, 131. [Google Scholar] [CrossRef]
  93. Alvarez-Erviti, L.; Couch, Y.; Richardson, J.; Cooper, J.M.; Wood, M.J.A. α-Synuclein Release by Neurons Activates the Inflammatory Response in a Microglial Cell Line. Neurosci. Res. 2011, 69, 337–342. [Google Scholar] [CrossRef] [PubMed]
  94. Klegeris, A.; Pelech, S.; Giasson, B.I.; Maguire, J.; Zhang, H.; McGeer, E.G.; McGeer, P.L. α-Synuclein Activates Stress Signaling Protein Kinases in THP-1 Cells and Microglia. Neurobiol. Aging 2008, 29, 739–752. [Google Scholar] [CrossRef]
  95. La Vitola, P.; Balducci, C.; Cerovic, M.; Santamaria, G.; Brandi, E.; Grandi, F.; Caldinelli, L.; Colombo, L.; Morgese, M.G.; Trabace, L.; et al. α-Synuclein Oligomers Impair Memory through Glial Cell Activation and via Toll-like Receptor 2. Brain Behav. Immun. 2018, 69, 591–602. [Google Scholar] [CrossRef] [PubMed]
  96. Hughes, C.D.; Choi, M.L.; Ryten, M.; Hopkins, L.; Drews, A.; Botía, J.A.; Iljina, M.; Rodrigues, M.; Gagliano, S.A.; Gandhi, S.; et al. Picomolar Concentrations of Oligomeric α-Synuclein Sensitizes TLR4 to Play an Initiating Role in Parkinson’s Disease Pathogenesis. Acta Neuropathol. 2019, 137, 103–120. [Google Scholar] [CrossRef] [Green Version]
  97. Bliederhaeuser, C.; Grozdanov, V.; Speidel, A.; Zondler, L.; Ruf, W.P.; Bayer, H.; Kiechle, M.; Feiler, M.S.; Freischmidt, A.; Brenner, D.; et al. Age-Dependent Defects of α-Synuclein Oligomer Uptake in Microglia and Monocytes. Acta Neuropathol. 2016, 131, 379–391. [Google Scholar] [CrossRef] [PubMed]
  98. Kim, C.; Ho, D.-H.; Suk, J.-E.; You, S.; Michael, S.; Kang, J.; Joong Lee, S.; Masliah, E.; Hwang, D.; Lee, H.-J.; et al. Neuron-Released Oligomeric α-Synuclein Is an Endogenous Agonist of TLR2 for Paracrine Activation of Microglia. Nat. Commun. 2013, 4, 1562. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Daniele, S.G.; Béraud, D.; Davenport, C.; Cheng, K.; Yin, H.; Maguire-Zeiss, K.A. Activation of MyD88-Dependent TLR1/2 Signaling by Misfolded α-Synuclein, a Protein Linked to Neurodegenerative Disorders. Sci. Signal. 2015, 8, ra45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Deas, E.; Cremades, N.; Angelova, P.R.; Ludtmann, M.H.R.; Yao, Z.; Chen, S.; Horrocks, M.H.; Banushi, B.; Little, D.; Devine, M.J.; et al. α-Synuclein Oligomers Interact with Metal Ions to Induce Oxidative Stress and Neuronal Death in Parkinson’s Disease. Antioxid. Redox Signal. 2016, 24, 376–391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. La Vitola, P.; Balducci, C.; Baroni, M.; Artioli, L.; Santamaria, G.; Castiglioni, M.; Cerovic, M.; Colombo, L.; Caldinelli, L.; Pollegioni, L.; et al. Peripheral Inflammation Exacerbates α-Synuclein Toxicity and Neuropathology in Parkinson’s Models. Neuropathol. Appl. Neurobiol. 2021, 47, 43–60. [Google Scholar] [CrossRef]
  102. Czapski, G.A.; Strosznajder, J.B. Glutamate and GABA in Microglia-Neuron Cross-Talk in Alzheimer’s Disease. Int. J. Mol. Sci. 2021, 22, 11677. [Google Scholar] [CrossRef]
  103. Michalettos, G.; Ruscher, K. Crosstalk Between GABAergic Neurotransmission and Inflammatory Cascades in the Post-Ischemic Brain: Relevance for Stroke Recovery. Front. Cell. Neurosci. 2022, 16, 807911. [Google Scholar] [CrossRef]
  104. Calvo-Flores Guzmán, B.; Vinnakota, C.; Govindpani, K.; Waldvogel, H.J.; Faull, R.L.M.; Kwakowsky, A. The GABAergic System as a Therapeutic Target for Alzheimer’s Disease. J. Neurochem. 2018, 146, 649–669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Alberdi, E.; Sánchez-Gómez, M.V.; Cavaliere, F.; Pérez-Samartín, A.; Zugaza, J.L.; Trullas, R.; Domercq, M.; Matute, C. Amyloid β Oligomers Induce Ca2+ Dysregulation and Neuronal Death through Activation of Ionotropic Glutamate Receptors. Cell Calcium 2010, 47, 264–272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Traynelis, S.F.; Wollmuth, L.P.; McBain, C.J.; Menniti, F.S.; Vance, K.M.; Ogden, K.K.; Hansen, K.B.; Yuan, H.; Myers, S.J.; Dingledine, R. Glutamate Receptor Ion Channels: Structure, Regulation, and Function. Pharmacol. Rev. 2010, 62, 405–496. [Google Scholar] [CrossRef] [Green Version]
  107. Schwenk, J.; Harmel, N.; Zolles, G.; Bildl, W.; Kulik, A.; Heimrich, B.; Chisaka, O.; Jonas, P.; Schulte, U.; Fakler, B.; et al. Functional Proteomics Identify Cornichon Proteins as Auxiliary Subunits of AMPA Receptors. Science 2009, 323, 1313–1319. [Google Scholar] [CrossRef] [Green Version]
  108. De Boer, H.; Blok, G.J.; Voerman, H.J.; van der Veen, E.A. Is Growth Hormone Supplementation in Growth Hormone Deficiency in Adults Indicated? Ned. Tijdschr. Geneeskd. 1990, 134, 2428–2431. [Google Scholar] [PubMed]
  109. Herring, B.E.; Shi, Y.; Suh, Y.H.; Zheng, C.-Y.; Blankenship, S.M.; Roche, K.W.; Nicoll, R.A. Cornichon Proteins Determine the Subunit Composition of Synaptic AMPA Receptors. Neuron 2013, 77, 1083–1096. [Google Scholar] [CrossRef] [Green Version]
  110. Kato, A.S.; Gill, M.B.; Yu, H.; Nisenbaum, E.S.; Bredt, D.S. TARPs Differentially Decorate AMPA Receptors to Specify Neuropharmacology. Trends Neurosci. 2010, 33, 241–248. [Google Scholar] [CrossRef]
  111. Cho, C.-H.; St-Gelais, F.; Zhang, W.; Tomita, S.; Howe, J.R. Two Families of TARP Isoforms That Have Distinct Effects on the Kinetic Properties of AMPA Receptors and Synaptic Currents. Neuron 2007, 55, 890–904. [Google Scholar] [CrossRef] [Green Version]
  112. Milstein, A.D.; Zhou, W.; Karimzadegan, S.; Bredt, D.S.; Nicoll, R.A. TARP Subtypes Differentially and Dose-Dependently Control Synaptic AMPA Receptor Gating. Neuron 2007, 55, 905–918. [Google Scholar] [CrossRef] [Green Version]
  113. Noh, K.-M.; Yokota, H.; Mashiko, T.; Castillo, P.E.; Zukin, R.S.; Bennett, M.V.L. Blockade of Calcium-Permeable AMPA Receptors Protects Hippocampal Neurons against Global Ischemia-Induced Death. Proc. Natl. Acad. Sci. USA 2005, 102, 12230–12235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Liu, S.; Lau, L.; Wei, J.; Zhu, D.; Zou, S.; Sun, H.-S.; Fu, Y.; Liu, F.; Lu, Y. Expression of Ca2+-Permeable AMPA Receptor Channels Primes Cell Death in Transient Forebrain Ischemia. Neuron 2004, 43, 43–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Spaethling, J.M.; Klein, D.M.; Singh, P.; Meaney, D.F. Calcium-Permeable AMPA Receptors Appear in Cortical Neurons after Traumatic Mechanical Injury and Contribute to Neuronal Fate. J. Neurotrauma 2008, 25, 1207–1216. [Google Scholar] [CrossRef]
  116. Corona, J.C.; Tapia, R. Ca2+-Permeable AMPA Receptors and Intracellular Ca2+ Determine Motoneuron Vulnerability in Rat Spinal Cord in Vivo. Neuropharmacology 2007, 52, 1219–1228. [Google Scholar] [CrossRef]
  117. Vieira, M.; Fernandes, J.; Burgeiro, A.; Thomas, G.M.; Huganir, R.L.; Duarte, C.B.; Carvalho, A.L.; Santos, A.E. Excitotoxicity through Ca2+-Permeable AMPA Receptors Requires Ca2+-Dependent JNK Activation. Neurobiol. Dis. 2010, 40, 645–655. [Google Scholar] [CrossRef] [Green Version]
  118. Reinders, N.R.; Pao, Y.; Renner, M.C.; da Silva-Matos, C.M.; Lodder, T.R.; Malinow, R.; Kessels, H.W. Amyloid-β Effects on Synapses and Memory Require AMPA Receptor Subunit GluA3. Proc. Natl. Acad. Sci. USA 2016, 113, E6526–E6534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Berchtold, N.C.; Sabbagh, M.N.; Beach, T.G.; Kim, R.C.; Cribbs, D.H.; Cotman, C.W. Brain Gene Expression Patterns Differentiate Mild Cognitive Impairment from Normal Aged and Alzheimer’s Disease. Neurobiol. Aging 2014, 35, 1961–1972. [Google Scholar] [CrossRef] [Green Version]
  120. Texidó, L.; Martín-Satué, M.; Alberdi, E.; Solsona, C.; Matute, C. Amyloid β Peptide Oligomers Directly Activate NMDA Receptors. Cell Calcium 2011, 49, 184–190. [Google Scholar] [CrossRef]
  121. Sinnen, B.L.; Bowen, A.B.; Gibson, E.S.; Kennedy, M.J. Local and Use-Dependent Effects of β-Amyloid Oligomers on NMDA Receptor Function Revealed by Optical Quantal Analysis. J. Neurosci. 2016, 36, 11532–11543. [Google Scholar] [CrossRef]
  122. Ferreira, I.L.; Bajouco, L.M.; Mota, S.I.; Auberson, Y.P.; Oliveira, C.R.; Rego, A.C. Amyloid β Peptide 1–42 Disturbs Intracellular Calcium Homeostasis through Activation of GluN2B-Containing N-Methyl-d-Aspartate Receptors in Cortical Cultures. Cell Calcium 2012, 51, 95–106. [Google Scholar] [CrossRef]
  123. Kessels, H.W.; Nabavi, S.; Malinow, R. Metabotropic NMDA Receptor Function Is Required for β-Amyloid-Induced Synaptic Depression. Proc. Natl. Acad. Sci. USA 2013, 110, 4033–4038. [Google Scholar] [CrossRef] [Green Version]
  124. Snyder, E.M.; Nong, Y.; Almeida, C.G.; Paul, S.; Moran, T.; Choi, E.Y.; Nairn, A.C.; Salter, M.W.; Lombroso, P.J.; Gouras, G.K.; et al. Regulation of NMDA Receptor Trafficking by Amyloid-β. Nat. Neurosci. 2005, 8, 1051–1058. [Google Scholar] [CrossRef]
  125. Ferreira, I.L.; Ferreiro, E.; Schmidt, J.; Cardoso, J.M.; Pereira, C.M.F.; Carvalho, A.L.; Oliveira, C.R.; Rego, A.C. Aβ and NMDAR Activation Cause Mitochondrial Dysfunction Involving ER Calcium Release. Neurobiol. Aging 2015, 36, 680–692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Costa, R.O.; Lacor, P.N.; Ferreira, I.L.; Resende, R.; Auberson, Y.P.; Klein, W.L.; Oliveira, C.R.; Rego, A.C.; Pereira, C.M.F. Endoplasmic Reticulum Stress Occurs Downstream of GluN2B Subunit of N-Methyl-d-Aspartate Receptor in Mature Hippocampal Cultures Treated with Amyloid-β Oligomers. Aging Cell 2012, 11, 823–833. [Google Scholar] [CrossRef] [PubMed]
  127. Evans, R.C.; Morera-Herreras, T.; Cui, Y.; Du, K.; Sheehan, T.; Kotaleski, J.H.; Venance, L.; Blackwell, K.T. The Effects of NMDA Subunit Composition on Calcium Influx and Spike Timing-Dependent Plasticity in Striatal Medium Spiny Neurons. PLoS Comput. Biol. 2012, 8, e1002493. [Google Scholar] [CrossRef] [PubMed]
  128. Skeberdis, V.A.; Chevaleyre, V.; Lau, C.G.; Goldberg, J.H.; Pettit, D.L.; Suadicani, S.O.; Lin, Y.; Bennett, M.V.L.; Yuste, R.; Castillo, P.E.; et al. Protein Kinase A Regulates Calcium Permeability of NMDA Receptors. Nat. Neurosci. 2006, 9, 501–510. [Google Scholar] [CrossRef] [PubMed]
  129. Trudler, D.; Sanz-Blasco, S.; Eisele, Y.S.; Ghatak, S.; Bodhinathan, K.; Akhtar, M.W.; Lynch, W.P.; Piña-Crespo, J.C.; Talantova, M.; Kelly, J.W.; et al. α-Synuclein Oligomers Induce Glutamate Release from Astrocytes and Excessive Extrasynaptic NMDAR Activity in Neurons, Thus Contributing to Synapse Loss. J. Neurosci. 2021, 41, 2264–2273. [Google Scholar] [CrossRef]
  130. Diógenes, M.J.; Dias, R.B.; Rombo, D.M.; Vicente Miranda, H.; Maiolino, F.; Guerreiro, P.; Näsström, T.; Franquelim, H.G.; Oliveira, L.M.A.; Castanho, M.A.R.B.; et al. Extracellular α-Synuclein Oligomers Modulate Synaptic Transmission and Impair LTP via NMDA-Receptor Activation. J. Neurosci. 2012, 32, 11750–11762. [Google Scholar] [CrossRef] [Green Version]
  131. Durante, V.; de Iure, A.; Loffredo, V.; Vaikath, N.; de Risi, M.; Paciotti, S.; Quiroga-Varela, A.; Chiasserini, D.; Mellone, M.; Mazzocchetti, P.; et al. α-Synuclein Targets GluN2A NMDA Receptor Subunit Causing Striatal Synaptic Dysfunction and Visuospatial Memory Alteration. Brain 2019, 142, 1365–1385. [Google Scholar] [CrossRef] [Green Version]
  132. Roberson, E.D.; Scearce-Levie, K.; Palop, J.J.; Yan, F.; Cheng, I.H.; Wu, T.; Gerstein, H.; Yu, G.-Q.; Mucke, L. Reducing Endogenous Tau Ameliorates Amyloid β-Induced Deficits in an Alzheimer’s Disease Mouse Model. Science 2007, 316, 750–754. [Google Scholar] [CrossRef] [Green Version]
  133. Ittner, L.M.; Ke, Y.D.; Delerue, F.; Bi, M.; Gladbach, A.; van Eersel, J.; Wölfing, H.; Chieng, B.C.; Christie, M.J.; Napier, I.A.; et al. Dendritic Function of Tau Mediates Amyloid-β Toxicity in Alzheimer’s Disease Mouse Models. Cell 2010, 142, 387–397. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Kaufman, A.C.; Salazar, S.V.; Haas, L.T.; Yang, J.; Kostylev, M.A.; Jeng, A.T.; Robinson, S.A.; Gunther, E.C.; van Dyck, C.H.; Nygaard, H.B.; et al. Fyn Inhibition Rescues Established Memory and Synapse Loss in Alzheimer Mice. Ann. Neurol. 2015, 77, 953–971. [Google Scholar] [CrossRef] [Green Version]
  135. Amatniek, J.C.; Hauser, W.A.; DelCastillo-Castaneda, C.; Jacobs, D.M.; Marder, K.; Bell, K.; Albert, M.; Brandt, J.; Stern, Y. Incidence and Predictors of Seizures in Patients with Alzheimer’s Disease. Epilepsia 2006, 47, 867–872. [Google Scholar] [CrossRef] [PubMed]
  136. Palop, J.J.; Mucke, L. Epilepsy and Cognitive Impairments in Alzheimer Disease. Arch. Neurol. 2009, 66, 435–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Hauser, W.A.; Morris, M.L.; Heston, L.L.; Anderson, V.E. Seizures and Myoclonus in Patients with Alzheimer’s Disease. Neurology 1986, 36, 1226–1230. [Google Scholar] [CrossRef] [PubMed]
  138. Mendez, M.; Lim, G. Seizures in Elderly Patients with Dementia: Epidemiology and Management. Drugs Aging 2003, 20, 791–803. [Google Scholar] [CrossRef]
  139. Jayadev, S.; Leverenz, J.B.; Steinbart, E.; Stahl, J.; Klunk, W.; Yu, C.-E.; Bird, T.D. Alzheimer’s Disease Phenotypes and Genotypes Associated with Mutations in Presenilin 2. Brain 2010, 133, 1143–1154. [Google Scholar] [CrossRef] [Green Version]
  140. Cabrejo, L.; Guyant-Maréchal, L.; Laquerrière, A.; Vercelletto, M.; de la Fournière, F.; Thomas-Antérion, C.; Verny, C.; Letournel, F.; Pasquier, F.; Vital, A.; et al. Phenotype Associated with APP Duplication in Five Families. Brain 2006, 129, 2966–2976. [Google Scholar] [CrossRef] [Green Version]
  141. Sperling, R.A.; Laviolette, P.S.; O’Keefe, K.; O’Brien, J.; Rentz, D.M.; Pihlajamaki, M.; Marshall, G.; Hyman, B.T.; Selkoe, D.J.; Hedden, T.; et al. Amyloid Deposition Is Associated with Impaired Default Network Function in Older Persons without Dementia. Neuron 2009, 63, 178–188. [Google Scholar] [CrossRef] [Green Version]
  142. Palop, J.J.; Chin, J.; Roberson, E.D.; Wang, J.; Thwin, M.T.; Bien-Ly, N.; Yoo, J.; Ho, K.O.; Yu, G.-Q.; Kreitzer, A.; 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. [Google Scholar] [CrossRef] [Green Version]
  143. Verret, L.; Mann, E.O.; Hang, G.B.; Barth, A.M.I.; Cobos, I.; Ho, K.; Devidze, N.; Masliah, E.; Kreitzer, A.C.; Mody, I.; et al. Inhibitory Interneuron Deficit Links Altered Network Activity and Cognitive Dysfunction in Alzheimer Model. Cell 2012, 149, 708–721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Sos, K.E.; Mayer, M.I.; Takács, V.T.; Major, A.; Bardóczi, Z.; Beres, B.M.; Szeles, T.; Saito, T.; Saido, T.C.; Mody, I.; et al. Amyloid β Induces Interneuron-Specific Changes in the Hippocampus of APPNL-F Mice. PLoS ONE 2020, 15, e0233700. [Google Scholar] [CrossRef] [PubMed]
  145. Li, G.; Bien-Ly, N.; Andrews-Zwilling, Y.; Xu, Q.; Bernardo, A.; Ring, K.; Halabisky, B.; Deng, C.; Mahley, R.W.; Huang, Y. GABAergic Interneuron Dysfunction Impairs Hippocampal Neurogenesis in Adult Apolipoprotein E4 Knockin Mice. Cell Stem Cell 2009, 5, 634–645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Andrews-Zwilling, Y.; Bien-Ly, N.; Xu, Q.; Li, G.; Bernardo, A.; Yoon, S.Y.; Zwilling, D.; Yan, T.X.; Chen, L.; Huang, Y. Apolipoprotein E4 Causes Age- and Tau-Dependent Impairment of GABAergic Interneurons, Leading to Learning and Memory Deficits in Mice. J. Neurosci. 2010, 30, 13707–13717. [Google Scholar] [CrossRef] [PubMed]
  147. Limon, A.; Reyes-Ruiz, J.M.; Miledi, R. Loss of Functional GABAA Receptors in the Alzheimer Diseased Brain. Proc. Natl. Acad. Sci. USA 2012, 109, 10071–10076. [Google Scholar] [CrossRef] [Green Version]
  148. Berchtold, N.C.; Coleman, P.D.; Cribbs, D.H.; Rogers, J.; Gillen, D.L.; Cotman, C.W. Synaptic Genes Are Extensively Downregulated across Multiple Brain Regions in Normal Human Aging and Alzheimer’s Disease. Neurobiol. Aging 2013, 34, 1653–1661. [Google Scholar] [CrossRef] [Green Version]
  149. Zott, B.; Busche, M.A.; Sperling, R.A.; Konnerth, A. What Happens with the Circuit in Alzheimer’s Disease in Mice and Humans? Annu. Rev. Neurosci. 2018, 41, 277–297. [Google Scholar] [CrossRef]
  150. Zhou, S.; Yu, Y. Synaptic E-I Balance Underlies Efficient Neural Coding. Front. Neurosci. 2018, 12, 46. [Google Scholar] [CrossRef] [Green Version]
  151. Okun, M.; Lampl, I. Instantaneous Correlation of Excitation and Inhibition during Ongoing and Sensory-Evoked Activities. Nat. Neurosci. 2008, 11, 535–537. [Google Scholar] [CrossRef]
  152. Xue, M.; Atallah, B.V.; Scanziani, M. Equalizing Excitation-Inhibition Ratios across Visual Cortical Neurons. Nature 2014, 511, 596–600. [Google Scholar] [CrossRef]
  153. Barral, J.; Reyes, A.D. Synaptic Scaling Rule Preserves Excitatory-Inhibitory Balance and Salient Neuronal Network Dynamics. Nat. Neurosci. 2016, 19, 1690–1696. [Google Scholar] [CrossRef] [PubMed]
  154. Vogels, T.P.; Sprekeler, H.; Zenke, F.; Clopath, C.; Gerstner, W. Inhibitory Plasticity Balances Excitation and Inhibition in Sensory Pathways and Memory Networks. Science 2011, 334, 1569–1573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Lustig, C.; Snyder, A.Z.; Bhakta, M.; O’Brien, K.C.; McAvoy, M.; Raichle, M.E.; Morris, J.C.; Buckner, R.L. Functional Deactivations: Change with Age and Dementia of the Alzheimer Type. Proc. Natl. Acad. Sci. USA 2003, 100, 14504–14509. [Google Scholar] [CrossRef] [Green Version]
  156. Chen, L.Y.; Rex, C.S.; Babayan, A.H.; Kramár, E.A.; Lynch, G.; Gall, C.M.; Lauterborn, J.C. Physiological Activation of Synaptic Rac>PAK (p-21 Activated Kinase) Signaling Is Defective in a Mouse Model of Fragile X Syndrome. J. Neurosci. 2010, 30, 10977–10984. [Google Scholar] [CrossRef]
  157. Chen, L.Y.; Rex, C.S.; Sanaiha, Y.; Lynch, G.; Gall, C.M. Learning Induces Neurotrophin Signaling at Hippocampal Synapses. Proc. Natl. Acad. Sci. USA 2010, 107, 7030–7035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Rex, C.S.; Chen, L.Y.; Sharma, A.; Liu, J.; Babayan, A.H.; Gall, C.M.; Lynch, G. Different Rho GTPase-Dependent Signaling Pathways Initiate Sequential Steps in the Consolidation of Long-Term Potentiation. J. Cell Biol. 2009, 186, 85–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Lauterborn, J.C.; Rex, C.S.; Kramár, E.; Chen, L.Y.; Pandyarajan, V.; Lynch, G.; Gall, C.M. Brain-Derived Neurotrophic Factor Rescues Synaptic Plasticity in a Mouse Model of Fragile X Syndrome. J. Neurosci. 2007, 27, 10685–10694. [Google Scholar] [CrossRef] [Green Version]
  160. Cirrito, J.R.; Kang, J.-E.; Lee, J.; Stewart, F.R.; Verges, D.K.; Silverio, L.M.; Bu, G.; Mennerick, S.; Holtzman, D.M. Endocytosis Is Required for Synaptic Activity-Dependent Release of Amyloid-β in Vivo. Neuron 2008, 58, 42–51. [Google Scholar] [CrossRef] [Green Version]
  161. Hettinger, J.C.; Lee, H.; Bu, G.; Holtzman, D.M.; Cirrito, J.R. AMPA-Ergic Regulation of Amyloid-β Levels in an Alzheimer’s Disease Mouse Model. Mol. Neurodegener. 2018, 13, 22. [Google Scholar] [CrossRef]
  162. Lauterborn, J.C.; Scaduto, P.; Cox, C.D.; Schulmann, A.; Lynch, G.; Gall, C.M.; Keene, C.D.; Limon, A. Increased Excitatory to Inhibitory Synaptic Ratio in Parietal Cortex Samples from Individuals with Alzheimer’s Disease. Nat. Commun. 2021, 12, 1–5. [Google Scholar] [CrossRef]
  163. Singh, A.; Allen, D.; Fracassi, A.; Tumurbaatar, B.; Natarajan, C.; Scaduto, P.; Woltjer, R.; Kayed, R.; Limon, A.; Krishnan, B.; et al. Functional Integrity of Synapses in the Central Nervous System of Cognitively Intact Individuals with High Alzheimer’s Disease Neuropathology Is Associated with Absence of Synaptic Tau Oligomers. J. Alzheimer’s Dis. 2020, 78, 1661–1678. [Google Scholar] [CrossRef] [PubMed]
  164. Taubes, A.; Nova, P.; Zalocusky, K.A.; Kosti, I.; Bicak, M.; Zilberter, M.Y.; Hao, Y.; Yoon, S.Y.; Oskotsky, T.; Pineda, S.; et al. Experimental and Real-World Evidence Supporting the Computational Repurposing of Bumetanide for APOE4-Related Alzheimer’s Disease. Nat. Aging 2021, 1, 932–947. [Google Scholar] [CrossRef]
  165. Llinás, R.R.; Ribary, U.; Jeanmonod, D.; Kronberg, E.; Mitra, P.P. Thalamocortical Dysrhythmia: A Neurological and Neuropsychiatric Syndrome Characterized by Magnetoencephalography. Proc. Natl. Acad. Sci. USA 1999, 96, 15222–15227. [Google Scholar] [CrossRef] [Green Version]
  166. Minakaki, G.; Krainc, D.; Burbulla, L.F. The Convergence of α-Synuclein, Mitochondrial, and Lysosomal Pathways in Vulnerability of Midbrain Dopaminergic Neurons in Parkinson’s Disease. Front. Cell Dev. Biol. 2020, 8, 580634. [Google Scholar] [CrossRef] [PubMed]
  167. Foliaki, S.T.; Schwarz, B.; Groveman, B.R.; Walters, R.O.; Ferreira, N.C.; Orrù, C.D.; Smith, A.; Wood, A.; Schmit, O.M.; Freitag, P.; et al. Neuronal Excitatory-to-Inhibitory Balance Is Altered in Cerebral Organoid Models of Genetic Neurological Diseases. Mol. Brain 2021, 14, 156. [Google Scholar] [CrossRef]
Figure 1. Overview of major effects of toxic oligomers in synapses. Left. Neuroinflammatory and Reactive Oxygen Species (ROS) participate in the production and the effects of toxic oligomers on synapses. Right, Major synaptic effects on synapses. It is still not clear what is the chronological order of events, but each one influence the others and some of them are happening simultaneously at brain regions vulnerable to AD pathology.
Figure 1. Overview of major effects of toxic oligomers in synapses. Left. Neuroinflammatory and Reactive Oxygen Species (ROS) participate in the production and the effects of toxic oligomers on synapses. Right, Major synaptic effects on synapses. It is still not clear what is the chronological order of events, but each one influence the others and some of them are happening simultaneously at brain regions vulnerable to AD pathology.
Biomedicines 10 01743 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gutierrez, B.A.; Limon, A. Synaptic Disruption by Soluble Oligomers in Patients with Alzheimer’s and Parkinson’s Disease. Biomedicines 2022, 10, 1743. https://doi.org/10.3390/biomedicines10071743

AMA Style

Gutierrez BA, Limon A. Synaptic Disruption by Soluble Oligomers in Patients with Alzheimer’s and Parkinson’s Disease. Biomedicines. 2022; 10(7):1743. https://doi.org/10.3390/biomedicines10071743

Chicago/Turabian Style

Gutierrez, Berenice A., and Agenor Limon. 2022. "Synaptic Disruption by Soluble Oligomers in Patients with Alzheimer’s and Parkinson’s Disease" Biomedicines 10, no. 7: 1743. https://doi.org/10.3390/biomedicines10071743

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop