Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2023 May 30.
Published in final edited form as: Pharmacol Res. 2023 Apr 20;191:106743. doi: 10.1016/j.phrs.2023.106743

Discoveries and future significance of research into amyloid-beta/α7-containing nicotinic acetylcholine receptor (nAChR) interactions

Paul Whiteaker 1, Andrew A George 1,*
PMCID: PMC10228377  NIHMSID: NIHMS1899178  PMID: 37084859

Abstract

Initiated by findings that Alzheimer’s disease is associated with a profound loss of cholinergic markers in human brain, decades of studies have examined the interactions between specific subtypes of nicotinic acetylcholine receptors and amyloid-β [derived from the amyloid precursor protein (APP), which is cleaved to yield variable isoforms of amyloid-β]. We review the evolving understanding of amyloid-β’s roles in Alzheimer’s disease and pioneering studies that highlighted a role of nicotinic acetylcholine receptors in mediating important aspects of amyloid-β’s effects. This review also surveys the current state of research into amyloid-β / nicotinic acetylcholine receptor interactions. The field has reached an exciting point in which common themes are emerging from the wide range of prior research and a range of accessible, relevant model systems are available to drive further progress. We highlight exciting new areas of inquiry and persistent challenges that need to be considered while conducting this research. Studies of amyloid-β and the nicotinic acetylcholine receptor populations that it interacts with provide opportunities for innovative basic and translational scientific breakthroughs related to nicotinic receptor biology, Alzheimer’s disease, and cholinergic contributions to cognition more broadly.

Keywords: Nicotinic receptors, Alzheimer’s disease, Memory, Neuronal hyperexcitation, Cholinergic neurons, Oligomeric amyloid-beta

1. Alzheimer’s disease and Amyloid-β

Alzheimer’s disease (AD) is a devastating neurodegenerative disease that accounts for approximately 2/3 of all cases of dementia. The impact of AD is staggering: in the US alone, the cost of caring for individuals with diagnoses of dementia was estimated to exceed 300B USD in 2020. In the next three decades, as the population ages, this cost of care is expected to triple [153]. AD was originally diagnosed post mortem, using histological hallmarks. These include accumulation of fibrillar amyloid-β (fAβ) plaques and neurofibrillary tangles [10]. For many years, scientific efforts focused on understanding the complex relationship between neuropathological signs of Alzheimer’s disease (AD) and its clinical manifestation (i.e. cognitive impairment). These efforts largely focused on approaches that correlated late-stage clinical symptoms in living patients with these known AD-related biomarkers [amyloid deposition (insoluble amyloid plaques), neurofibrillary tangles (hyperphosphorylated tau), and neurodegeneration; [62]. This was the underlying basis of the “amyloid cascade hypothesis,” that aberrant processing of amyloid precursor protein led Aβ to aggregate into fAβ deposits (in particular, neuritic plaques), and that this was causally related to AD development and cognitive decline [123,126,52]. This hypothesis has developed into a useful conceptual framework that provides researchers and clinicians with testable hypotheses and diagnostic recommendations spanning different phases of the AD continuum [29]. However, as we will address next, continuing research has increasingly brought into question the role of amyloid plaques in AD pathological change.

2. Cognitive decline and oligomeric forms of Aβ

Studies over approximately the last two decades have deemphasized or shown marginal contributions of neuritic plaques and neurofibrillary tangles to dementia [15,159,84,96,98]. In addition, repeated failures of plaque-removing therapies during clinical trials have brought the classical amyloid hypothesis into further doubt or, at the least, indicated that other factors may play more important roles earlier during disease development [106,60,89,88]. While a significant amount of evidence indicates that Aβ plays a causative and initiating role in the development of AD, only a limited number of anti-Aβ agents have proven effective in clinical trials. For example, lecanumab-irmb (aka Laqembi), an immunoglobulin targeted against insoluble forms of Aβ, reduces markers of Aβ in early AD [111] but is associated with amyloid-related imaging abnormities (i.e. brain swelling or bleeding). Even the US Food and Drug Administration’s approval of Aducanumab, a monoclonal antibody treatment that reduces amyloid plaque loads [128], to treat AD has been controversial since patients receiving the drug did not see statistically-significant benefits over subjects receiving placebo and suffered significant side-effects [150]. Lastly, clinical trials using gantenerumab, a fully-human monoclonal IgG1 antibody designed to target and bind to aggregated forms of beta-amyloid, including oligomers, fibrils and plaques, exhibited an underwhelming level of Aβ removal and failed to improve the rate of cognitive and functional decline in a pair of phase 3 clinical trials [40]. These factors have all fueled the overall impression that other mechanisms than deposition of neuritic plaques may be more important [115].

As the primacy of the classical amyloid hypothesis was being challenged, several lines of evidence emphasized instead the important role of soluble, oligomeric forms of amyloid-β (oAβ), and their impact on the physiological processes that mediate cognition [21]. For example, oAβ begins to accumulate earlier during AD development than plaques [51,73,75]. The discovery of the Osaka familial AD mutation demonstrated that individuals could suffer from severe cognitive loss [130,138] while simultaneously exhibiting elevated soluble oAβ and extremely low levels of neuritic plaques [130,138,61,72]. Further studies have shown that plaque loads correlate poorly, while concentrations of soluble oAβ correlate better, with cognitive decline in both human AD patients and mouse models of AD [1,35,47,97].

The preceding findings raised questions about how circulating, soluble forms of amyloid play roles during the early stages of AD pathologic change. Specifically, how elevated levels of circulating oAβ alter neuronal and network-level function and, in turn, the development, integration, and retrieval of memory long before the expression of clinical symptoms [104]. Findings from animal models have provided new insights into causal upstream roles for oAβ in the alterations of neural circuits that are commonly affected in AD [16,105]. One key question is which form or forms of oAβ are most responsible for AD initiation and/or progression. Many in vivo and in vitro models support the hypothesis that oAβ42 (oligomers of the Aβ amino acids 1–42) play a particularly critical role in neurodegeneration and subsequent memory deficits [120,48]. Together, the factors mentioned in this section have provided substantial momentum to inquiries into soluble oAβ in general, and oAβ42, in particular. As we will address later, oAβ42 likely interferes with stable neuronal function [104], and this understanding opens important avenues for understanding the basis of neuronal and network-level instability, neurodegeneration, and memory loss in AD (Table 1).

Table 1.

Summary of studies investigating the role of Aβ42 on cellular and circuit-level function.

Publication Model and/or Aβ application Site of Action Phenotype
[41] APP/PS1 mice; Basal forebrain slices; SH-EPI cells; oligomeric Aβ42 [100 nM] α7- and a7b2-nAChR Direct activation of α7 nAChR with preferential alteration of α7β2 nAChR kinetics; septal cholinergic hyperexcitation
[50] Hippocampal CA1 pyramidal neurons; oligomeric Aβ42 [200 pM] α7*-nAChR Enhances mEPSP frequency; increased length of postsynaptic density; maintenance of hippocampal LTP
[58] APP/PS1 mice Unknown Reduction of Purkinje cell intrinsic membrane excitability; altered GABAergic and glutamatergic release from climbing fiber interneurons
[119] Hippocampal slices; exogenous Aβ42 [200 pM] α7*-nAChR Aβ regulation of presynaptic neurotransmitter release; maintenance of hippocampal LTP
[13] PSAPP mice Voltage dependent Na+ channels CA1 pyramidal cell hyperexcitability
[64] Hippocampal slices; 5XFAD mice Carbachol-sensitive mAChR decreased CA1 hippocampal neuronal excitability; enhancement of the afterhyperpolarization (AHP)
[139] NG108–15 cells α7*-nAChR Pronounced and sustained increases in Ca2+ transients from axonal varicosities of NG108–15 cells
[66] NG108–15 cells; oligomeric Aβ42 [100 pM to 100 nM] α7*-nAChR induces Ca2+ responses in NG108–15 neuritic varicosities
[158] Drosophila giant fiber system; (Gal4)–UAS driven Aβ42 Unknown Age-dependent depletion of presynaptic mitochondria and neurotransmission failure
[90] Cortical and hippocampal synaptosomes; oligomeric Aβ42 [10PM to 100 nM] α7*-nAChRs Alterations in presynaptic neurotransmission; differential neuromodulation by Aβ of synapses in hippocampus and cortex
[34] APPα7KO mouse model α7*-nAChR α7KO normalizes Aβ-induced LTP deficits; prevents learning and memory deficits observed in APP-overexpressing animals
[101] Cultured hippocampal neurons; oligomeric Aβ42 [100 nM] P/Q-type calcium channels Inhibition of P/Q-Type Calcium Currents; suppression of spontaneous synaptic activity
[118] Hippocampal slices; monomeric and oligomeric Aβ42 [200 pM] α7*-nAChR Enhancement of synaptic plasticity and enhancement of reference and contextual fear memory
[18] Basal forebrain slices; Aβ42 [100 nM] Non- α7*-nAChRs Differential effects on mEPSC frequency but not amplitude in DBB neurons
[17] Hippocampal slices α7*-nAChR Reduced neurotransmission and excitatory post-synaptic potentials (EPSPs); impairment of LTP
[33] 42 [1 pM to 100 nM] α7*-nAChR Sustained, increases in presynaptic Ca2+ via α7-containing and non- α7-containing nAChR
[39] Basal forebrain cholinergic neurons; oligomeric Aβ42 [100 nM] Non- α7*-nAChR Activates non- α7 nAChR; Ab-induced DBB neuronal depolarizations; Ab-induced inward currents from DBB neurons
[30] Acute hippocampal slices and Tg2576 mice α7*-nAChRs In vivo elevation of Aβ leads to the upregulation of α7*-nAChR protein.

3. AD and cholinergic signaling

Initial signs of cholinergic deficits in AD brain came from postmortem studies. These showed greater losses of cholinergic markers in cortical and hippocampal regions of AD patients than age-matched, cognitively-normal controls [110,11,149,28]. These findings gave rise to the “cholinergic hypothesis of AD” which “suggests that a dysfunction of acetylcholine containing neurons in the brain contributes substantially to the cognitive decline observed in those with advanced age and Alzheimer’s disease” [136]. The majority of cholinergic neurons in the mammalian brain are found in four regions. These include the brainstem pedunculo-pontine and lateral dorsal tegmental nuclei, a subset of thalamic nuclei, the striatum, and the basal forebrain cholinergic nuclei [6]. Basal forebrain cholinergic neurons (BFCNs) serve as the major sources of cholinergic projection neurons to neocortex, hippocampus, and amygdala [6], provide widespread innervation of the hippocampus and cerebral cortex, and important control of cognitive processing and memory acquisition [141]. Cholinergic projections from the medial septum and ventral diagonal band (MSDB) project to the hippocampus, parahippocampus, olfactory bulb, and midline cortical structures [6]. Based on their anterior-to-posterior and medial-to-lateral distribution, the horizontal limb of the diagonal band (HDB), the nucleus basalis (NB), and the substantia innominata (SI) project to medial frontal cortical targets, basal lateral amygdala, and the perirhinal cortex [6].

Early studies into the cholinergic hypothesis of AD provide correlative evidence that a key pathological characteristic of mild-cognitive impairment (MCI) and AD includes age-related reductions in basal forebrain volume and cholinergic transmission [59,80,79]. BFCN loss is an early feature both in humans [151] and in mouse models of AD [156,8]. Further support for an important role of cholinergic signaling in AD is provided by the fact that acetylcholinesterase inhibitors provide one of the few, albeit time-limited, therapeutic options for AD [44,45,7,9]. While BFCN signaling is clearly an important mediator of the cognitive decline associated with AD, the exact mechanisms by which the degeneration of BFCNs occurs in AD remain under active investigation. This review article focuses on mechanisms that involve interactions between oAβ and nicotinic acetylcholine receptors.

4. nAChR and AD

The initial findings of cholinergic deficits in postmortem AD brains were followed up with examinations of effects on cholinergic receptor expression. As reviewed in [102], changes in muscarinic acetylcholine receptor (mAChR) expression were not consistently seen. The story was very different for nAChR, however. Before mentioning these early findings, an overview will be provided of nAChR diversity.

nAChR subtypes exist as a diverse family of subtypes, each being a pentameric complex of specific, homologous, but genetically-distinct subunits [160,27,63]. These subtypes can, to a first approximation, be divided into three categories:

  1. Muscle-subtype receptors are expressed only on muscle tissue, and are composed of α1, β1, δ, and either γ or ε subunits. We will not address them further in this review.

  2. Non-muscle subtypes can be subdivided into two categories, heteromeric and homomeric. Heteromeric, non-muscle nAChR subtypes also contain multiple different subunits, typically at least one of a set of α subunits (in mammalian CNS; α2, α3, α4, or α6) and at least one of the β2 or β4 subunits. These α and β subunits host orthosteric agonist binding pockets at the subunit interfaces, with a “principal” face provided by the α subunit, and a “complementary” face provided by the β subunit. Further diversity can be produced by incorporation of an accessory subunit (α5 or β3) that modifies receptor functional properties but does not directly participate in agonist binding [160,27,63] For example, the most commonly expressed subtype in mammalian brain is the α4β2*-nAChR [85], where * indicates the potential presence of additional subunits [82] such as the α5 subunit [14]. In mammalian brain, heteromeric nAChR, especially the α;4β2*-nAChR subtype, are associated with high-affinity agonist binding sites as detected in radioligand binding and autoradiography experiments [85].

  3. Homomeric subtypes contain five identical subunits, and include the α7-only-nAChR subtype which is among the most common and widespread within mammalian brain [85]. In a homomeric nAChR context, both the principal and the complementary faces of agonist binding sites are provided by opposite sides of the same subunit [160,27,63]. As we will explore later most, but not all, α7*-nAChRα7* -nAChR are homomeric α7-only-nAChR. In mammalian brain, α7*-nAChR are associated with high-affinity binding of the selective antagonist α-bungarotoxin (α-Bgt) in radioligand binding and autoradiography experiments [85].

A large number of investigations used combinations of autoradiographic, immunochemical, and radioligand binding approaches to probe for changes in nAChR expression. As carefully reviewed by [23], the overall conclusions were that widespread, significant, and consistent reductions of high-affinity agonist binding site expression (i.e., of heteromeric nAChR, likely predominantly corresponding to α4β2*-nAChR) occurred in AD brain. In addition, significant losses of α-Bgt-binding sites (i.e., α7*-nAChR) also occurred in AD brain, although in a more restricted set of locations.

5. Early studies of nAChR-mediated protection against the cellular toxicity of Aβ

Inspired by the discovery that the brains of AD patients exhibit deficits of nAChR expression (see preceding section), a wide range of studies examined the ability of nAChR pharmacological agents to protect cells (often neurons) from the toxic effects of Aβ. A comprehensive overview of all of these studies is not the focus of this review, so an abridged outline will be provided here.

In an early example, rat cultured cortical neurons were shown to be protected from cell death induced by exposure to amyloid-β peptide (amino acids 25–35; Aβ25–35) by co-incubation with nicotine, in a concentration dependent manner [68]. This same study showed that nicotine’s protective effect could be replicated using a selective α7*-nAChR agonist (DMXB), and blocked by the α7*-nAChR-selective antagonist α-Bgt. The same research team also demonstrated a similar protective effect, in the same preparation, by exposure to the α4β2*-nAChR-selective agonist cytisine, that could be blocked by the α4β2-selective antagonist DHβE [69]. However, partial protection of cultured mouse against Aβ42-induced toxicity by the α7*-nAChR-selective antagonist MLA was also reported [86]. This observation questioned whether activation of nAChR, or persistent desensitization due to prolonged exposure was the cause of agonist-mediated protection [86]. Effects of compounds outside of typical competitive nAChR agonists/antagonists were also reported. Examples include protection by β-estradiol of rat PC12 cells from Aβ25–35-induced toxicity (an effect blocked by the α7*-nAChR-selective competitive MLA and the less subtype-selective non-competitive nAChR antagonist mecamylamine [134], RNA-interference knockdown of α7*-nAChR expression (which increased vulnerability of SH-SY5Y cells to Aβ25–35-induced toxicity; [122], and protective effects of acetylcholinesterase inhibitors such as tacrine and donepezil (protection at PC12 cells blocked by the poorly subtype-selective nAChR inhibitors mecamylamine and d-tubocurarine [135]) or galantamine [protection against Aβ40-induced cell death, reversed by α-Bgt, in SH-SY5Y cells [4].

Similarly, a wide range of studies examined potential mechanisms through which nAChR-mediated protection against Aβ-induced cell death might occur. One of the earliest such studies demonstrated nicotine inhibition of Aβ-evoked activation of phospholipases A2 and D [131]. Other examples include nAChR-mediated modulation of amyloid precursor protein (APP) processing and tau phosphorylation [127,55], α7*-nAChR-mediated signaling through the phosphoinositol-3-kinase signaling [70], and Janus kinase 2 pathways [129], and nicotine-evoked changes in oxidative stress responses [46].

The many and varied outcomes of the preceding publications (and of the many other related studies not mentioned here) are hard to integrate in detail, likely due in large part to the wide range of reagents, systems, and forms of Aβ used. Nonetheless, a clear overall message does emerge: manipulation of nAChR, often α7*-nAChR in particular, can impact the toxic effects of Aβ. However, few of these studies examined whether Aβ directly interacts with nAChR, the subject of our next section.

6. Early studies of direct Aβ/nAChR interactions

A pair of studies claimed to show direct binding of amyloid-β protein (either amino acids 1–40 or 1–42; Aβ40 or Aβ42, respectively) to α7*-nAChR [148,147]. These studies should be treated with caution since the claimed affinities were both exceptionally high, and apparently variable across experiments (Ki values claimed at α7*-nAChR ranged from 4.1 or 5.0 pM in [147], to as low as 8 fM in [148]. Nevertheless, two subsequent studies showed rapid and reversible antagonism of whole-cell α7*-nAChR function on rat hippocampal neurons by rat Aβ42. Potencies reported in both cases were in the nanomolar range (no specific IC50 value; [112], IC50 = 7.5 nM; [76]). Of interest, [76] also reported that inhibition was selective towards α7*-nAChR over other subtypes expressed in the preparation, and appeared to be non-competitive in nature. These observations were reinforced by [42], who demonstrated non-competitive block of α7*-nAChR with an IC50 of 90 nM, and by [121], who found that nM concentrations of either Aβ40 or Aβ42 were able to inhibit function of α7*-nAChR, transiently enhanced that of α4β2*-nAChR, and were without effect on that of α3β4*-nAChR. An interesting counterpoint is provided by [32], in which pM concentrations of Aβ42 were shown to activate α7*-nAChR expressed X. laevis oocytes, an effect diminished at higher concentrations (nM range). Contemporaneous studies did not replicate this activation of α7*-nAChR function by sub-nM Aβ42 [for example [121] and [42], although the concentration dependent effects reported by [32] closely match those reported in a recent single-channel study [74]. To complicate matters further, one study [39] showed Aβ activation of non--α7*-nAChR (putatively of the α4*-nAChR subtype) in rat basal forebrain neurons.

As may be seen, the detailed results of these early studies are mixed and, in some instances, contradictory. This may reflect technical differences across the studies in how Aβ40 and Aβ42 solutions were prepared [and therefore the forms of Aβ aggregates present [132], and of how nAChR currents were recorded [α7*-nAChR responses are particularly sensitive to the timing and concentration of ligand application [144]. Nevertheless, an overall picture emerges that α7*-nAChR are particularly sensitive to the effects of Aβ40 and Aβ42, and that high concentrations of these ligands inhibit or negatively modulate function of α7*-nAChR, possibly through a non-competitive mechanism. It has been generally considered that α7 nAChR subunits assemble into a homomeric (α7-only) subtype. However, as we will explore next, evidence for expression of heteromeric α7*-nAChR emerged contemporaneously with evidence of α7-only-nAChR expression.

7. Evidence for the native expression of heteromeric α7*-nAChR

As reviewed in [20], it was only with the cloning and expression in 1990 of what is now known as the nAChR α7 subunit [124,24] that the existence of a functional, non-muscle, α-Bgt sensitive nAChR subtype was widely accepted. Even at this early stage, immunochemical experiments [124] indicated that a minority of chick α7-nAChR contained another subunit (now named α8). This initial finding of heteromeric avian α7*-nAChR expression was quickly supported by immunohistochemical evidence of co-location in chick brain of expression of α7 with α8 or β2 subunits [12], although caution must be exercised regarding the specificity of immunocytochemical labeling of nAChR α7 subunits [56]. Further circumstantial evidence for native heteromeric α7*-nAChR expression was found in differences in functional pharmacology between α7-only- and putatively α7α8-nAChR isolated from chick retina [3], and subtle differences of agonist pharmacology between artificially-expressed α7-only-nAChR and α7*-nAChR immunoisolated from chick brain [2]. However, the α8 subunit is not expressed in mammalian species, and findings suggestive of heteromeric α7*-nAChR expression in mammals were not obtained for several more years. Whole-cell patch-clamp recordings from isolated rat intracardiac and superior cervical ganglion neurons identified α7*-nAChR with unusually slow desensitization properties and the ability to recover rapidly from α-Bgt antagonism [25,26]. In the rat CNS, dual-labeling in situ mRNA hybridization demonstrated co-expression of α7 and β2 subunits, frequently in the absence of the α4 subunit that typically combines with β2, across cholinergic cells within the basal forebrain, mesopontine tegmentum, and striatum [5]. Some GABAergic neurons associated with cholinergic nuclei were also found in this study to have similar co-expression patterns of α7 with β2 but not α4 nAChR subunits. This finding was highly reminiscent of the earlier immunohistochemistry finding of α7 and β2 nAChR subunit co-expression in chick brain [12]. More recently, we and others [137,94] used affinity purification and immunoprecipitation methods to isolate α7*-nAChR and probe them for the presence of an α7β2-nAChR subpopulation. A consistent finding between both studies is that western blotting, using antibodies whose specificity has been verified in subunit-null mutant mice, identified co-assembly of β2 subunits into α7β2-nAChR in both mouse and human forebrain. This region, of course, contains the basal forebrain cholinergic neurons (BFCNs) identified in [5]. The evidence to date indicates that α7β2-nAChR are the most prominent, perhaps only, heteromeric α7*-nAChR population expressed in mammalian brain, and that their expression is most pronounced in cholinergic nuclei (and cholinergic neurons in particular). This makes BFCNs an ideal native system in which to study α7β2-nAChR, and their effects on normal and disease pharmacology.

8. Functional properties of α7β2-nAChR

However, functional pharmacology profiles of α7β2-nAChR have been most extensively assessed using artificial expression systems that allow better control over nAChR subtype expression compared to native neurons. In an early study, α7 and β2 subunits were co-expressed in Xenopus laevis oocytes, resulting in a nAChR population with significantly slower desensitization properties than α7-only-nAChR [67]. These same authors confirmed assembly of β2 subunit into functional α7β2-nAChR using a reporter-mutation approach. They also showed co-immunoprecipitation of α7 and β2 subunits, as well as reduced efficacy of a small number of agonists at α7β2-compared to α7-only-nAChR. In a later example we used a fluorescence resonance energy transfer (FRET) approach to demonstrate directly that fluorescently-tagged human α7 and β2 subunits assemble into α7β2-nAChR when stably transfected into SH-EP1 cells [95]. Initial functional studies in this paper indicated that agonist and antagonist profiles were generally similar between α7-only- and α7β2-nAChR, although possible increased sensitivity of α7β2-nAChR to the competitive antagonist dihydr-β-erythroidine (DHβE) was noted. We later performed more extensive pharmacological profiling of α7-only- vs. α7β2-nAChR expressed in X. laevis oocytes using a linked-subunit approach to ensure full control of subunit assembly and association [94]. The resulting concatenated nAChR subtypes again showed little difference in agonist or antagonist potencies, although some differences in agonist efficacy again were noted, a finding further emphasized in α7β2-nAChR formed from unlinked subunits [161]. Broadly similar findings were reported from α7-only- vs. α7β2-nAChR expressed in X. laevis oocytes using unlinked subunits [137], although this study did note somewhat increased potency of DHβE antagonism at α7β2- over α7-only-nAChR, and reduced potency of Compound B (an α7*-nAChR selective partial agonist compound).

β subunits typically make important contributions to the pharmacological profiles of competitive agonists and antagonists [81]. For this reason, the very similar pharmacological profiles of α7-only- and α7β2-nAChR suggested that both subtypes might be activated only by agonist binding at their common α7/α7 subunit interfaces. Indeed, we tested this hypothesis using a cysteine accessibility mutant approach [95]. Altering a critical residue on the complementary face of α7 (L119C) and subsequent covalent modification by the sulfhydryl reagent methanethiosulfonate ethylammonium (MTSA) blocks access of agonists to the competitive binding site of α7-only-nAChR [109]. We replicated this finding for α7β2-nAChR (please see the earlier section “Nicotinic Receptors (nAChR) and AD” for an explanation of the terms “complementary” and “principal” faces of nAChR competitive binding sites). However, when the equivalent mutation was made in the complementary face of the β2 subunit (L121C), agonist (ACh) activation of α7β2-nAChR was unaffected. We concluded that these results showed that complementary faces of α7 subunits are critical contributors to agonist binding at both α7-only- and α7β2-nAChR, while those of β2 subunits are not [95]. This conclusion was reinforced by an elegant single-channel electrophysiology study that used an “electrical fingerprinting” approach [100]. Here, α7 subunits were tagged using mutations that significantly reduced single-channel amplitude and conductance. The effect is additive, meaning that each “electrically fingerprinted” α7 subunit that is incorporated into an α7*-nAChR complex reduces open-channel amplitude by an incremental amount, allowing α7:β2 subunit ratios to be counted. This technique demonstrated that α7β2-nAChR incorporating up to three β2 subunits can be functional, a finding compatible with the concept that two adjacent α7 subunits are required to form an agonist activation site. This requirement was confirmed using a further mutation to the complementary face of the α7 nAChR subunit that prevents agonist activation via the α7/α7 subunit-interface binding site (in this case, α7[W55T]). Incorporation of this mutation into α7β2-nAChR prevented all function, showing that loss of function at α7/α7 sites could not be rescued by agonist activation via any α7/β2 sites that were present.

Two prominent exceptions have been seen to the trend of similar pharmacology between α7-only- and α7β2-nAChR. The first exception is in the case of positive allosteric modulator (PAM) compounds. This presumably arises from the fact that PAM compounds, by definition, have sites of action outside of the conventional competitive agonist binding pocket. Differences were seen for two Type I PAM compounds [100]. Most notably, 5-hydroxyindole (5-HI) had much less ability to potentiate macroscopic current responses of α7β2-than α7-only-nAChR. This finding was matched by decreased ability of 5-HI to increase open and burst durations of single-channel responses of α7β2- compared to α7-only-nAChR, with increased incorporation of β2 subunits producing progressively lower enhancement by 5-HI. In the case of the second Type I PAM (NS-1738) the ability to prolong single-channel open and burst durations was again lessened in α7β2- compared to α7-only-nAChR, although significant enhancement of macroscopic responses by NS-1738 was seen. Interestingly, the sole Type II PAM tested (PNU-120596) had indistinguishable effects at α7-only and α7β2-nAChR (regardless of the α7:β2 subunit ratio). The second exception is in responses to soluble oligomers of amyloid-β (amino acids 1–42; oAβ42). We demonstrated that oAβ42 (100 nM; within the range of pathophysiologically-relevant concentrations described in humans) can evoke single-channel responses in both α7-only- and α7β2-nAChR [41]. Measured in the presence of the PAM PNU-120596, burst durations fell into two categories for both subtypes (short- and long-duration bursts). For α7-only-nAChR, short-duration bursts induced by ACh or oAβ42 were indistinguishable from each other (in terms of mean duration or open amplitude), and the same was true for long-duration bursts. However, the situation was more complex in the case of α7β2-nAChR. While short-duration bursts induced by ACh or oAβ42 at α7β2-nAChR were again indistinguishable from each other, long-duration bursts were approximately 5 times longer when induced by oAβ42 than ACh. Applying ACh and oAβ42 simultaneously to α7β2-nAChR resulted in outcomes that were indistinguishable from those induced by applying oAβ42 alone, suggesting a lack of competition between oAβ42 and ACh (i.e., that OAβ42 likely is producing activation through an allosteric mechanism, at least at α7β2-nAChR). A further publication determined that the effects of oAβ42 are concentration dependent: concentrations as low as 100 pM (applied alone) produce single-channel activation of α7-only-nAChR, while high-nanomolar concentrations predominantly act as negative modulators of ACh-induced function [74]. This range of effects and concentration ranges closely matches that seen in the study of macroscopic function by [32]. Application of a spectroscopic approach also suggested that OAβ42 binds outside of the ACh binding pocket [74], further reinforcing the concept that oAβ42 may act in an allosteric manner at α7*-nAChR.

The overall conclusion from this section is that functional and pharmacological properties of α7-only- and α7β2-nAChR are very similar. This is due to activation of both α7*-nAChR subtypes through conserved α7/α7 subunit interface ligand binding sites, at least in the case of conventional competitive agonists and antagonists. Despite this, some differences in agonist efficacy are seen consistently and there is some evidence of slower desensitization and increased sensitivity to DHβE antagonism of α7β2- compared to α7-only-nAChR). However, there is accumulating evidence from multiple studies that some allosteric ligands are able to distinguish convincingly between these closely-related subtypes.

9. Evidence for mammalian α7β2-nAChR function and interactions with oAβ in mammalian CNS

As detailed earlier, there has been substantial evidence over the last two decades that heterologous α7*-nAChR populations may occur in both peripheral and CNS locations. The use of artificial expression systems has provided critical information regarding the functional properties of α7β2*-nAChR in particular (see preceding section), allowing comparison to the properties recorded from native α7*-nAChR that do not conform to those typically associated with α7-only-nAChR. Earlier studies provided hints that these “atypical” α7*-nAChR might be α7β2-nAChR. For example, the Yakel laboratory identified α7*-nAChR with slow desensitization properties in rat hippocampal interneurons together with a significant correlation between expression of α7 and β2 subunits in these neurons [133]. This finding of slow-desensitizing α7*-nAChR function in hippocampal interneurons was extended in [77], which also showed high sensitivity of these responses to oAβ42 and DHβE. The same group had earlier [78] identified similar responses in mouse BFCNs, a population highlighted in [5] as being likely to host α7β2-nAChR. We found further evidence of α7β2-nAChR function contributions in specific populations of mouse BFCNs. Chronic application of oAβ42 (100 nM, 9 days) to cultured basal forebrain slices induced hyperexcitation in BFCNs of the medial septal diagonal band (MSDB) and horizontal diagonal band (HDB). This hyperexcitation phenotype was abolished both by application of α7*-nAChR-selective antagonists, and in recordings from slices from nAChR β2 subunit-null mice, indicating dependence on α7β2-nAChR expression [41]. Interestingly, BFCNs of the nucleus basalis (NB) did not show the same excitation phenotype. This may be explained by the earlier in situ hybridization study [5], which showed that α4 mRNA expression was far more prominent in NB BFCNs than the populations found in the HDB and MSDB. This may result in a higher proportion of α4β2- instead of α7β2-nAChR expression in NB BFCNs. Interestingly, AD and MCI (mild cognitive impairment) patients exhibit up-regulation of α7*-nAChR messenger RNA expression in NB BFCNs, with no differences found for other nAChR subtypes [22]. The usual caveat must be considered in this case, that changes in mRNA expression do not necessarily correspond to proportionate changes in protein expression (see for example the [23] paper cited earlier in this review).

While there is consistent evidence for α7β2-nAChR expression on hippocampal interneurons and, especially, BFCNs, it has to be noted that α7β2-nAChR may be expressed elsewhere in the brain. Certainly there is suggestive evidence that α7β2-nAChR also may be expressed in cholinergic neurons in the striatum and mesopontine tegmentum [5], but the question of whether α7β2-nAChR might be expressed elsewhere in the brain has not been addressed systematically. This likely reflects the difficulty in distinguishing α7β2-nAChR from the larger population of α7-only-nAChR expressed across the brain. However, even our current knowledge of the locations of α7β2-nAChR expression indicates that they can have profound physiological impacts. Network function within the hippocampus is critically regulated by hippocampal interneurons [38]. Similarly, BFCNs project widely across hippocampal and cortical regions and play prominent roles in modulating neuronal circuits that mediate cognitive processing [113,157,91]. Our recent finding that abolishing α7β2-nAChR function can prevent BFCN hyperexcitation induced by chronic oAβ42 exposure and rescue cognitive deficits observed in the APP/PS1 transgenic mouse model of Alzheimer’s disease [41] illustrates the physiological importance of α7β2-nAChR function and oAβ42/α7β2-nAChR interactions.

10. Impacts of soluble forms of Aβ on neuronal, synaptic, and circuit function

It is critical to acknowledge that not all effects of soluble Aβ are deleterious. Various forms of soluble Aβ (including Aβ42 and Aβ40) are observed within the central nervous system throughout an individual’s lifetime [43,92,103]. Deficits in synaptic transmission and synaptic plasticity (including those that mediate learning and memory) are observed with pathologically-relevant levels of soluble Aβ (Snyder et al., 2005; Chapman et al., 1999; [146]; D’Amelio et al., 2011 [140]. However, lower, physiologically-relevant levels of soluble Aβ have been shown to enhance specific engrams of memory, including LTP (Puzzo et al., 2008; [119]; Gulisano et al., 2019). Interestingly, enhancement of both reference and contextual fear memory, and the mechanism(s) of action of picomolar levels of Aβ42 on both synaptic plasticity and memory, involves α7-containing nicotinic acetylcholine receptors (Puzzo et al., 2008). This extraordinary sensitivity fits neatly with imaging studies of isolated rat hippocampus presynaptic nerve endings that demonstrated directly-evoked, sustained, increases in presynaptic Ca2+ via α7-containing and non-α7-containing nAChR by picomolar concentrations of Aβ42 (Dougherty et al., 2003), and with the even earlier claims of exceptionally-high affinity binding to α7-only nAChR [148,147] noted earlier in this review. The apparently contradictory effects of soluble Aβ can be explained by investigations into concentration-dependent effects of Aβ on neural function. These have clearly shown that low concentrations play a positive, modulatory role on the physiological process that govern neurotransmission and memory [119], and, paradoxically, higher concentrations of Aβ can have an inverse effect by reducing potentiation in in neural circuits that are involved in memory formation (Puzzo et al., 2008). Another important consideration is that production of soluble Aβ and neuronal/synaptic activity are not independent. For example, Aβ levels increase during the induction of long-term potentiation (LTP; Palmeri et al., 2017) and many studies have demonstrated that Aβ can modulate neurotransmission and synaptic plasticity in neural circuits implicated in the acquisition and consolidation of memory (Araki et al., 1991; Mucke et al., 1994; Ishida et al., 1997; Meziane et al., 1998; Dougherty et al., 2003; Puzzo et al., 2008). Further, Aβ is regulated within presynaptic terminals and is implicated in vesicle cycling (Cirrito et al., 2005, 2008).

Our recent investigation illuminates the importance of soluble Aβ/nAChR interactions in the context of neuronal function. As described earlier, we showed that oAβ42 exposure leads to persistent activation of α7*-nAChR, with a selective enhancement of α7β2-nAChR function in particular. We also showed hyperexcitation of BFCNs from the medial septum diagonal band (MSDB) and horizontal diagonal band (HDB), but not the nucleus basalis (NB) that was dependent on interactions of α7β2-nAChR with a pathophysiologically-relevant concentration of oAβ42 [41]. The hyperexcitation phenotype (enhanced neuronal intrinsic excitability and action potential firing rates) resulted from a reduction in action potential afterhyperpolarization (AHP) and alterations in the maximal rates of voltage change during BFCN spike depolarization and repolarization. We also demonstrated that aged male and female APP/PS1 transgenic mice, genetically null for the β2 nAChR subunit gene, showed improved spatial reference memory compared with APP/PS1 aged-matched littermates. Indeed, genetic deletion of the α7-nAChR subunit (which will eliminate expression of both homomeric α7-only and heteromeric α7β2-nAChR subtypes) causes an impairment of hippocampal synaptic plasticity and memory at 12 months of age, paralleled by an increase of amyloid precursor protein (APP) expression and Aβ levels [140]. Combined, findings from this study provide a molecular mechanism supporting a role for α7β2-nAChR in mediating the effects of oAβ42 on excitability of specific populations of cholinergic neurons, and provide a framework for understanding the role of α7β2-nAChR in oAβ42-induced cognitive decline (Fig. 1).

Fig. 1.

Fig. 1.

Oligomeric amyloid-β (oAb42) interactions and α7*-nAChR. (A) Under neuronormal conditions, both homomeric α7- and heteromeric α7β2-nAChRs are activated by the endogenous ligand acetylcholine (ACh; yellow circle). Homomeric α7-only nAChR harbor competitive agonist binding pockets at the interfaces between “principal” (+) faces of each α7 subunit and the opposing “complementary” faces (−) provided by adjacent α7 subunits. Heteromeric α7β2-nAChRs host competitive agonist binding pockets only at the interfaces between adjacent α7 subunit interfaces. (B) During periods of elevated oAb42, oAb42 binds to and activates both α7-only- and α7β2-nAChR subtypes. However, as described in [41] oAβ42 selectively enhances the single-channel open-dwell times of α7β2-containing nAChRs. The location(s) through which oAβ produce α7*-nAChR activation are not yet defined (as indicated by a question mark, although there is evidence for a non-competitive / allosteric mode of action (please see text).

The studies mentioned in this section predominantly focus on the ability of soluble Aβ to bind and functionally modulate α7*-nAChR and, in turn, alter the cellular properties of specific neuronal populations implicated in cognition. Yet, there is more to study. It is important to consider not only 1) how soluble Aβ/nAChR interactions directly alter the functional output of individual neurons (i.e. cell autonomous effects), but also 2) how oAβ42/α7*-nAChR interactions indirectly affect neuronal function by altering the functional output of local inhibitory (i. e. GABA) and/or excitatory (i.e. glutamate) neurons. Some work assessing the effects of extracellular OAβ42 administration on presynaptic and postsynaptic mechanisms underlying synaptic function and memory has been described [50]. However, these observations have largely been limited to an analysis of homogenous populations of neurons or extracellular electrophysiological field recordings and measurements of hippocampal long-term potentiation. As described in [142], investigation into the effects of Aβ42 on GABA-mediated synaptic transmission within the somatosensory cortex revealed reduced GABAergic input within layer V pyramidal neurons resulting from GABAR internalization [142]. Conversely, using an oligomeric form of Aβ1–42, Nimmrich et al. reported depressed vesicular release at GABAergic and glutamatergic synapses, due to Aβ-induced decrease in the probability of vesicle exocytosis from presynaptic terminals [101]. Given these findings, it is evident that in order to fully define the effects of oAβ42 on neural function, one must consider the effect(s) of oAβ42/α7*-nAChR interactions on local inhibitory and excitatory cell populations and how those local interactions can alter the balance of excitation and inhibition within local circuits, in turn changing the functional output of distinct brain nuclei (Fig. 2).

Fig. 2.

Fig. 2.

Local excitatory and inhibitory cellular processes that mediate basal forebrain cholinergic (BFCN) excitability. Within the medial septum, glutamatergic (light blue) and GABAergic (yellow) inputs regulate BFCN (green) excitability through α7*-nAChR-mediated synaptic transmission. Given the potential expression of α7*-nAChR (α7-only and/or (α7β2-nAChR; indicated by green and red cylinders, respectively) on glutamatergic and GABAergic terminals, the observed Aβ-induced BFCN hyperexcitability [41] may occur through enhanced glutamatergic synaptic transmission directly onto BFCNs or through disinhibition of GABAergic transmission onto BFCNs (i.e., circuit-based effects) In addition, oAβ42/α7*-nAChR interactions directly on BFCNs produce enhanced BFCN excitability. We have shown that BFCN hyperexcitation occurs through alterations in the intrinsic properties of BFCNs [likely via alterations in function of Ca2+ mediated or voltage-dependent potassium (K+) channels; blue triangles] [41].

11. Challenges and future opportunities

In this review, we have explored early findings that indicated an important role for studies of cholinergic biology, and of nAChR in particular, in AD research. These prior studies consistently indicate the importance interactions between soluble oligomeric forms of Aβ and nAChR and consistently highlight that oAβ/α7*-nAChR are especially of interest. More recent developments have shown that while most α7*-nAChR are of the α7-only (homomeric) subtype, the less-common heteromeric α7β2-nAChR population is particularly sensitive to the effects of oAβ AND is located on neuronal populations known to be particularly vulnerable early in the development of AD. These insights give us, as a field, accessible and relevant model systems to work with and thus opportunities to make further scientific breakthroughs. We will conclude by outlining some of the specific challenges and opportunities associated with ongoing studies of oAβ/nAChR interactions, informed by the studies performed to date:

Differentiation of α7-only from α7β2-nAChR:

Extensive pharmacological characterization has proved the pharmacological profiles of α7-only and α7β2-nAChR are very similar, likely due to the fact that agonism and antagonism by conventional competitive compounds appears to only occur at α7/α7 subunit interfaces that are found in both α7*-nAChR subtypes. There is some evidence that some non-competitive/allosteric ligands are able to discriminate between α7-only and α7β2-nAChR, however. Identification of ligands that are highly-selective for α7β2-over α7-only-nAChR would be an exceptionally significant achievement, allowing definitive assignment of roles to α7β2- and/or α7-only-nAChR in mediating oAβ-induced outcomes, and permitting exploration of all locations in which the rarer (but functionally important) α7β2-nAChR population is found.

Forms of oAβ:

Especially in early studies, different Aβ fragments were studied. More recently, studies have tended to emphasize the use of Aβ40 and/or oAβ42 fragments. Even if focusing on these particular peptides, however, it is critically important to know the aggregation state present in the experimental system. Careful control of time and solution conditions is absolutely critical to having known, consistent, distributions of oligomeric Aβ forms [132,99]. It is also important to note that other Aβ fragments are produced in the CNS and that these can also be biologically active [114,37,36]. Careful investigations of Aβ fragments outside of Aβ40 and/or oAβ42 are likely to be highly informative.

oAβ concentrations:

As highlighted in earlier sections of this review, studies support both α7*-nAChR activation and inhibition by oAβ. The dissimilarities in results across prior studies appear to be directly related to the model system used to study receptor function (including receptor subunit stoichiometry) and the aggregation state and concentration of Aβ42 used due to differences in preparation and handling of Aβ stocks (see preceding paragraph and [31,53]. The consideration of oAβ concentration is particularly difficult to address satisfactorily since estimates of physiologically-relevant concentrations of oAβ in normal control and AD brain vary widely from low picomolar to high micromolar [154,19]. Accordingly, both a firmer understanding of what concentrations of oAβ are pathophysiologically-relevant, and studies that examine effects across a range of oAβ concentrations are likely required to gain consistency across the field.

Preclinical model systems:

Somewhat related to the preceding is the choice of preclinical models for oAβ/AD-related studies. The use of AD animal models has given investigators valuable insight into the pathology underlying AD. While the wide range of transgenic mouse models mimic specific aspects of human AD [93,105], overexpression of specific genes [e.g. human amyloid precursor protein (APP) and presenilin 1 (PS1)] needed to reproduce AD neuropathology, including cognitive and behavioral deficits commonly associated with AD, correlate poorly to behavioral phenotypes observed in the majority of humans with AD [117,16]. Careful selection of animal models that mimic applicable aspects of aging, AD pathology, and expression of soluble oAβ concentrations appropriate to particular sets of experiments will be an important consideration for investigators interested in using these animal models toward understanding the role of cholinergic transmission within the context of AD.

Neuron / circuit model systems:

Findings to date highlight the importance of BFCNs. They are a neuronal population with proved, significant, roles in early AD. BFCNs also are the neuronal population for which the best evidence of α7β2-nAChR expression has been collected, and are readily identified [41]. The local and projection circuitry of BFCN nuclei is also well understood [6] As a result, BFCNs provide a superb native system with proven AD relevance in which to study effects of oAβ/α7β2-nAChR interactions on cell-autonomous neuronal, and local and projection circuit outcomes.

Cellular excitability:

Aβ’s ability to influence synaptic function has been well-described [116,119,125,146,53,87] and more recent findings using mouse models describe the exogenous administration of Aβ exerting its effects on synaptic plasticity through α7*-nAChR [49]. As described in [50], synaptic plasticity (driven by oAβ/α7*-nAChR interactions) modulates the cellular mechanisms that underlie transmitter release (presynaptic) and, concomitantly, treatment with picomolar concentrations of oAβ42 before a weak tetanic stimulation converts the early, induction phase of hippocampal long-term potentiation (LTP) into a longer, maintained phase of LTP [49]. These recent findings underscore the need to identify the precise cellular mechanism(s) that contribute to changes in neuronal excitability in context of oAβ/α7*-nAChR associations. At the presynaptic level, the effects of Aβ on the cellular mechanisms that regulate neurotransmitter release have been described [101]. More recent studies, including our own, have focused on Aβ-induced alterations in the intrinsic excitability of single neurons. These changes have been reported in both excitatory and inhibitory neurons [13,145,41,54,58,64,65,93]. Clearly, in order to understand how elevated levels of Aβ can influence short and/or long-term changes in neuronal plasticity, Aβ/nAChR interactions occurring at both sides of the synapse must be considered. Further, more recent findings demonstrate α7*-nAChR- calcium transients mediated by a G-protein-coupled mechanism that functions synergistically to augment calcium release in response to α7*-nAChR activation [71]. These findings may provide a novel intracellular signaling cascade in which Aβ/α7*-nAChR interactions mediate not only short-term synaptic plasticity, but also downstream changes in postsynaptic neuronal excitability. Understanding the relationship between Aβ and α7*-nAChR and its effects on ion channels that mediate specific intrinsic processes governing neuronal excitability may deepen our understanding of how Aβ-induced long-term changes in neuronal function could potentially explain the increase in epileptiform activity commonly observed in AD patients [105].

Technical difficulty of measuring α7*-nAChR function:

As extensively studied by the Papke laboratory, studies of macroscopic function of α7*-nAChR are significantly complicated by the complex relationship between agonist occupation, receptor activation, and receptor desensitization [109,152]. The exceptionally rapid desensitization (in some circumstances) of α7*-nAChR also makes correct assessment of pharmacological parameters technically challenging when considering macroscopic responses [108,107,143]. As pointed out in our recent publication [41], these considerations may be especially challenging in the case of an unconventional α7*-nAChR ligand such as oAβ and there may be significant advantages to studying α7*-nAChR function at the single-channel level, instead.

12. Significance and concluding remarks

Studies of oAβ/nAChR interactions have reached an exciting phase in which earlier findings are beginning to coalesce into a more-coherent picture of disease relevance and disease-related processes. Addressing carefully and effectively the items listed in the preceding section should result in opportunities to build an internally consistent and scientifically significant research effort on this foundation. Such progress has promise to provide further insights into cholinergic control of learning and memory, and may provide us with further (and direly needed) clinical / drug-development targets for AD. Critically, investigating the cellular and circuit-level mechanisms altered by oAβ42/nAChR interactions has the potential to provide a common understanding applicable to future scientific and clinical work in other conditions in which amyloidogenic processes are prominently involved. For example, similar mechanisms are known to be in play in aspects of Down syndrome [155,57,83] and studies in both areas may define common threads spanning multiple disorders. These shared mechanisms could then be targeted for therapeutic intervention by, e.g., manipulating oAβ42/nAChR associations directly or through intervention at newly-defined pathways altered by these interactions.

The realization that α7β2-nAChR may be especially-important partners in oAβ/nAChR interactions also provides significant opportunities. This subtype appears to have restricted distribution and physiologically-significant roles. Accordingly, it is a potentially-promising target for the kind of therapeutic interventions just mentioned. In addition, the need to develop α7β2-nAChR using ligands that interact outside of the conventional α7/α7 agonist binding pocket seems likely to yield invaluable insights regarding how best to manipulate nAChR functional outcomes across a wide range of subtypes and associated normal and disease processes.

Funding

Dr. George’s research is funded by National Institutes of Health award [R21AG067029]. Dr. Whiteaker’s research is funded by National Institutes of Health awards [R01DA042749, R01DA043567]. The National Institutes of Health did not influence the writing or submission of this article in any way; the opinions expressed are solely those of the authors.

Abbreviations:

5-HI

5-hydroxyindole

α

Bgt

α

bungarotoxin

amyloid-β

AD

Alzheimer’s disease

BFCN

basal forebrain cholinergic neuron

DHβE

dihydro-β-erythroidine

fAβ

fibrillar amyloid-β

FRET

Förster/fluorescence resonance energy transfer

HDB

horizontal limb of the diagonal band

mAChR

muscarinic acetylcholine choline receptor

MCI

mild cognitive impairment

MLA

methyllycaconitine

MSDB

medial septum and ventral diagonal band

nAChR

nicotinic acetylcholine receptor

NB

nucleus basalis

oAβ

oligomeric amyloid-β

PAM

positive allosteric modulator

Footnotes

The multifaceted activities of nervous and non-nervous neuronal nicotinic acetylcholine receptors in physiology and pathology. Eds: Dr Cecilia Gotti, Prof Francesco Clementi, Prof Michele Zoli.

CRediT authorship contribution statement

Andrew A. George and Paul Whiteaker contributed equally to the preparation, writing, editing, and presentation of the published work.

Declaration of Competing Interest

The authors declare that they have no competing interests that could bias the contents of this article.

Data availability

Data will be made available on request.

References

  • [1].Aizenstein HJ, Nebes RD, Saxton JA, Price JC, Mathis CA, Tsopelas ND, Ziolko SK, James JA, Snitz BE, Houck PR, Bi W, Cohen AD, Lopresti BJ, DeKosky ST, Halligan EM, Klunk WE, Frequent amyloid deposition without significant cognitive impairment among the elderly. Arch. Neurol 65 (2008) 1509–1517. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [2].Anand R, Peng X, Lindstrom J, Homomeric and native α7 acetylcholine receptors exhibit remarkably similar but non-identical pharmacological properties, suggesting that the native receptor is a heteromeric protein complex, FEBS Lett. 327 (1993) 241–246. [DOI] [PubMed] [Google Scholar]
  • [3].Anand R, Peng X, Ballesta JJ, Lindstrom J, Pharmacological characterization of alpha-bungarotoxin-sensitive acetylcholine receptors immunoisolated from chick retina: contrasting properties of alpha 7 and alpha 8 subunit-containing subtypes, Mol. Pharmacol 44 (1993) 1046–1050. [PubMed] [Google Scholar]
  • [4].Arias E, Alés E, Gabilan NH, Cano-Abad MF, Villarroya M, García AG, López MG, Galantamine prevents apoptosis induced by β-amyloid and thapsigargin: involvement of nicotinic acetylcholine receptors, Neuropharmacology 46 (2004) 103–114. [DOI] [PubMed] [Google Scholar]
  • [5].Azam L, Winzer-Serhan U, Leslie FM, Co-expression of α7 and β2 nicotinic acetylcholine receptor subunit mRNAs within rat brain cholinergic neurons, Neuroscience 119 (2003) 965–977. [DOI] [PubMed] [Google Scholar]
  • [6].Ballinger EC, Ananth M, Talmage DA, Role LW, Basal forebrain cholinergic circuits and signaling in cognition and cognitive decline, Neuron 91 (2016) 1199–1218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [7].Bazzari HF, Abdallah MD, El-Abhar SH, Pharmacological Interventions to Attenuate Alzheimer’s Disease Progression: The Story So Far, Curr. Alzheimer Res 16 (2019) 261–277. [DOI] [PubMed] [Google Scholar]
  • [8].Bilkei-Gorzo A, Genetic mouse models of brain ageing and Alzheimer’s disease, Pharmacol. Ther 142 (2014) 244–257. [DOI] [PubMed] [Google Scholar]
  • [9].Birks JS, Harvey RJ, Donepezil for dementia due to Alzheimer’s disease, Cochrane Database Syst. Rev 6 (2018), CD001190. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [10].Bondi MW, Edmonds EC, Salmon DP, Alzheimer’s Disease: Past, Present, and Future, J. Int Neuropsychol. Soc 23 (2017) 818–831. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [11].Bowen DM, Smith CB, White P, Davison AN, Neurotransimitter-related enzymes and indices of hypoxia in senile dementia and other abiotrophies, Brain 99 (1976) 459–496. [DOI] [PubMed] [Google Scholar]
  • [12].Britto LRG, Hamassaki-Britto DE, Ferro ES, Keyser KT, Karten HJ, Lindstrom JM, Neurons of the chick brain and retina expressing both α-bungarotoxin-sensitive and α-bungarotoxin-insensitive nicotinic acetylcholine receptors: an immunohistochemical analysis, Brain Res 590 (1992) 193–200. [DOI] [PubMed] [Google Scholar]
  • [13].Brown JT, Chin J, Leiser SC, Pangalos MN, Randall AD, Altered intrinsic neuronal excitability and reduced Na+ currents in a mouse model of Alzheimer’s disease, Neurobiol. Aging 32 (2109) (2011) e2101–e2114. [DOI] [PubMed] [Google Scholar]
  • [14].Brown RWB, Collins AC, Lindstrom JM, Whiteaker P, Nicotinic α5 subunit deletion locally reduces high-affinity agonist activation without altering nicotinic receptor numbers, J. Neurochem 103 (2007) 204–215. [DOI] [PubMed] [Google Scholar]
  • [15].Castellani RJ, Zhu X, Moreira Lee H-g PI, Perry MA, Smith G, Neuropathology and treatment of Alzheimer disease: did we lose the forest for the trees? Expert Rev. Neurother 7 (2007) 473–485. [DOI] [PubMed] [Google Scholar]
  • [16].Chen C, Ma X, Wei J, Shakir N, Zhang JK, Zhang L, Nehme A, Cui Y, Ferguson D, Bai F, Qiu S, Early impairment of cortical circuit plasticity and connectivity in the 5XFAD Alzheimer’s disease mouse model, Transl. Psychiatry 12 (2022) 371. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [17].Chen L, Yamada K, Nabeshima T, Sokabe M, alpha7 Nicotinic acetylcholine receptor as a target to rescue deficit in hippocampal LTP induction in beta-amyloid infused rats, Neuropharmacology 50 (2006) 254–268. [DOI] [PubMed] [Google Scholar]
  • [18].Chin JH, Ma L, MacTavish D, Jhamandas JH, Amyloid beta protein modulates glutamate-mediated neurotransmission in the rat basal forebrain: involvement of presynaptic neuronal nicotinic acetylcholine and metabotropic glutamate receptors, J. Neurosci 27 (2007) 9262–9269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [19].Cirrito JR, May PC, O’Dell MA, Taylor JW, Parsadanian M, Cramer JW, Audia JE, Nissen JS, Bales KR, Paul SM, DeMattos RB, Holtzman DM, In vivo assessment of brain interstitial fluid with microdialysis reveals plaque-associated changes in amyloid-beta metabolism and half-life, J. Neurosci 23 (2003) 8844–8853. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [20].Clarke PBS, The fall and rise of neuronal α-bungarotoxin binding proteins, Trends Pharm. Sci 13 (1992) 407–413. [DOI] [PubMed] [Google Scholar]
  • [21].Cline EN, Bicca MA, Viola KL, Klein WL, The amyloid-β oligomer hypothesis: beginning of the third decade, J. Alzheimer’S. Dis 64 (2018) S567–S610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [22].Counts SE, He B, Che S, Ikonomovic MD, DeKosky ST, Ginsberg SD, Mufson EJ, Alpha7 nicotinic receptor up-regulation in cholinergic basal forebrain neurons in Alzheimer disease, Arch. Neurol 64 (2007) 1771–1776. [DOI] [PubMed] [Google Scholar]
  • [23].Court J, Martin-Ruiz C, Piggott M, Spurden D, Griffiths M, Perry E, Nicotinic receptor abnormalities in Alzheimer’s disease, Biol. Psychiatry 49 (2001) 175–184. [DOI] [PubMed] [Google Scholar]
  • [24].Couturier S, Bertrand D, Matter J-M, Hernandez M-C, Bertrand S, Millar N, Valera S, Barkas T, Ballivet M, A neuronal nicotinic acetylcholine receptor subunit (α7) is developmentally regulated and forms a homo-oligomeric channel blocked by α-BTX, Neuron 5 (1990) 847–856. [DOI] [PubMed] [Google Scholar]
  • [25].Cuevas J, Berg DK, Mammalian Nicotinic Receptors with α7 Subunits That Slowly Desensitize and Rapidly Recover from α-Bungarotoxin Blockade, J. Neurosci 18 (1998) 10335–10344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [26].Cuevas J, Roth AL, Berg DK, Two distinct classes of functional α7-containing nicotinic receptor on rat superior cervical ganglion neurons, J. Physiol 525 (2000) 735–746. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [27].Dani JA, Chapter One - Neuronal Nicotinic Acetylcholine Receptor Structure and Function and Response to Nicotine. international Review of Neurobiology (De Biasi M ed), Academic Press,, 2015, pp. 3–19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [28].Davies P, Maloney AJ, Selective loss of central cholinergic neurons in Alzheimer’s disease, Lancet 2 (1976) 1403. [DOI] [PubMed] [Google Scholar]
  • [29].Davis M, O`Connell T, Johnson S, Cline S, Merikle E, Martenyi F, Simpson K, Estimating Alzheimer’s Disease Progression Rates from Normal Cognition Through Mild Cognitive Impairment and Stages of Dementia, Curr. Alzheimer Res 15 (2018) 777–788. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [30].Dineley KT, Westerman M, Bui D, Bell K, Ashe KH, Sweatt JD, Beta-amyloid activates the mitogen-activated protein kinase cascade via hippocampal alpha7 nicotinic acetylcholine receptors: In vitro and in vivo mechanisms related to Alzheimer’s disease, J. Neurosci 21 (2001) 4125–4133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [31].Dineley KT, Beta-amyloid peptide–nicotinic acetylcholine receptor interaction: the two faces of health and disease, Front Biosci. 12 (2007) 5030–5038. [DOI] [PubMed] [Google Scholar]
  • [32].Dineley KT, Bell KA, Bui D, Sweatt JD, beta-Amyloid peptide activates alpha 7 nicotinic acetylcholine receptors expressed in Xenopus oocytes, J. Biol. Chem 277 (2002) 25056–25061. [DOI] [PubMed] [Google Scholar]
  • [33].Dougherty JJ, Wu J, Nichols RA, Beta-amyloid regulation of presynaptic nicotinic receptors in rat hippocampus and neocortex, J. Neurosci 23 (2003) 6740–6747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [34].Dziewczapolski G, Glogowski CM, Masliah E, Heinemann SF, Deletion of the alpha 7 nicotinic acetylcholine receptor gene improves cognitive deficits and synaptic pathology in a mouse model of Alzheimer’s disease, J. Neurosci 29 (2009) 8805–8815. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [35].Foley AM, Ammar ZM, Lee RH, Mitchell CS, Systematic Review of the Relationship between Amyloid-β Levels and Measures of Transgenic Mouse Cognitive Deficit in Alzheimer’s Disease, J. Alzheimer’S. Dis 44 (2015) 787–795. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [36].Forest KH, Nichols RA, Assessing neuroprotective agents for abeta-induced neurotoxicity, Trends Mol. Med 25 (2019) 685–695. [DOI] [PubMed] [Google Scholar]
  • [37].Forest KH, Alfulaij N, Arora K, Taketa R, Sherrin T, Todorovic C, Lawrence JLM, Yoshikawa GT, Ng HL, Hruby VJ, Nichols RA, Protection against beta-amyloid neurotoxicity by a non-toxic endogenous N-terminal beta-amyloid fragment and its active hexapeptide core sequence, J. Neurochem 144 (2018) 201–217. [DOI] [PubMed] [Google Scholar]
  • [38].Freund TF, Buzsáki G, Interneurons of the hippocampus, Hippocampus 6 (1996) 347–470. [DOI] [PubMed] [Google Scholar]
  • [39].Fu W, Jhamandas JH, Beta-amyloid peptide activates non-alpha7 nicotinic acetylcholine receptors in rat basal forebrain neurons, J. Neurophysiol 90 (2003) 3130–3136. [DOI] [PubMed] [Google Scholar]
  • [40].Garraway L (2022) [Ad hoc announcement pursuant to Art. 53 LR] Roche provides update on Phase III GRADUATE programme evaluating gantenerumab in early Alzheimer’s disease.
  • [41].George AA, Vieira JM, Xavier-Jackson C, Gee MT, Cirrito JR, Bimonte-Nelson HA, Picciotto MR, Lukas RJ, Whiteaker P, Implications of Oligomeric Amyloid-Beta (oAbeta42) Signaling through alpha7beta2-Nicotinic Acetylcholine Receptors (nAChRs) on Basal Forebrain Cholinergic Neuronal Intrinsic Excitability and Cognitive Decline, J. Neurosci 41 (2021) 555–575. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [42].Grassi F, Palma E, Tonini R, Amici M, Ballivet M, Eusebi F, Amyloid beta(1-42) peptide alters the gating of human and mouse alpha-bungarotoxin-sensitive nicotinic receptors, J. Physiol 547 (2003) 147–157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [43].Gravina SA, Ho L, Eckman CB, Long KE, Otvos L Jr., Younkin LH, Suzuki N, Younkin SG, Amyloid beta protein (A beta) in Alzheimer’s disease brain. Biochemical and immunocytochemical analysis with antibodies specific for forms ending at A beta 40 or A beta 42(43), J. Biol. Chem 270 (1995) 7013–7016. [DOI] [PubMed] [Google Scholar]
  • [44].Grossberg GT, Edwards KR, Zhao Q, Rationale for combination therapy with galantamine and memantine in Alzheimer’s disease, J. Clin. Pharm 46 (2006) 17S–26S. [DOI] [PubMed] [Google Scholar]
  • [45].Grossberg GT, Tong G, Burke AD, Tariot PN, Present Algorithms and Future Treatments for Alzheimer’s Disease, J. Alzheimers Dis 67 (2019) 1157–1171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [46].Guan ZZ, Yu WF, Nordberg A, Dual effects of nicotine on oxidative stress and neuroprotection in PC12 cells, Neurochem Int 43 (2003) 243–249. [DOI] [PubMed] [Google Scholar]
  • [47].Guillozet AL, Weintraub S, Mash DC, Mesulam MM, Neurofibrillary Tangles, Amyloid, and Memory in Aging and Mild Cognitive Impairment, Arch. Neurol 60 (2003) 729–736. [DOI] [PubMed] [Google Scholar]
  • [48].Gulisano W, Melone M, Li Puma DD, Tropea MR, Palmeri A, Arancio O, Grassi C, Conti F, Puzzo D, The effect of amyloid-beta peptide on synaptic plasticity and memory is influenced by different isoforms, concentrations, and aggregation status, Neurobiol. Aging 71 (2018) 51–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [49].Gulisano W, Maugeri D, Baltrons MA, Fa M, Amato A, Palmeri A, D’Adamio L, Grassi C, Devanand DP, Honig LS, Puzzo D, Arancio O, Role of amyloid-beta and tau proteins in Alzheimer’s disease: confuting the amyloid cascade, J. Alzheimers Dis 68 (2019) 415. [DOI] [PubMed] [Google Scholar]
  • [50].Gulisano W, Melone M, Ripoli C, Tropea MR, Li Puma DD, Giunta S, Cocco S, Marcotulli D, Origlia N, Palmeri A, Arancio O, Conti F, Grassi C, Puzzo D, Neuromodulatory Action of Picomolar Extracellular Abeta42 Oligomers on Presynaptic and Postsynaptic Mechanisms Underlying Synaptic Function and Memory, J. Neurosci 39 (2019) 5986–6000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [51].Gyure KA, Durham R, Stewart WF, Smialek JE, Troncoso JC, Intraneuronal Aβ-Amyloid Precedes Development of Amyloid Plaques in Down Syndrome, Arch. Pathol. Lab. Med 125 (2001) 489–492. [DOI] [PubMed] [Google Scholar]
  • [52].Hardy JA, Higgins GA, Alzheimer’s disease: the amyloid cascade hypothesis, Science 256 (1992) 184–185. [DOI] [PubMed] [Google Scholar]
  • [53].Hascup KN, Hascup ER, Soluble Amyloid-beta42 Stimulates Glutamate Release through Activation of the alpha7 Nicotinic Acetylcholine Receptor, J. Alzheimers Dis 53 (2016) 337–347. [DOI] [PubMed] [Google Scholar]
  • [54].Hazra A, Gu F, Aulakh A, Berridge C, Eriksen JL, Ziburkus J, Inhibitory neuron and hippocampal circuit dysfunction in an aged mouse model of Alzheimer’s disease, PLoS One 8 (2013), e64318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [55].Hellstrom-Lindahl E, Modulation of beta-amyloid precursor protein processing and tau phosphorylation by acetylcholine receptors, Eur. J. Pharm 393 (2000) 255–263. [DOI] [PubMed] [Google Scholar]
  • [56].Herber DL, Severance EG, Cuevas J, Morgan D, Gordon MN, Biochemical and histochemical evidence of nonspecific binding of alpha7nAChR antibodies to mouse brain tissue, J. Histochem Cytochem 52 (2004) 1367–1376. [DOI] [PubMed] [Google Scholar]
  • [57].Holtzman DM, Santucci D, Kilbridge J, Chua-Couzens J, Fontana DJ, Daniels SE, Johnson RM, Chen K, Sun Y, Carlson E, Alleva E, Epstein CJ, Mobley WC, Developmental abnormalities and age-related neurodegeneration in a mouse model of Down syndrome, Proc. Natl. Acad. Sci. USA 93 (1996) 13333–13338. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [58].Hoxha E, Boda E, Montarolo F, Parolisi R, Tempia F, Excitability and synaptic alterations in the cerebellum of APP/PS1 mice, PLoS One 7 (2012), e34726. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [59].Hu X, Teunissen CE, Spottke A, Heneka MT, Duzel E, Peters O, Li S, Priller J, Buerger K, Teipel S, Laske C, Verfaillie SCJ, Barkhof F, Coll-Padros N, Rami L, Molinuevo JL, van der Flier WM, Jessen F, Smaller medial temporal lobe volumes in individuals with subjective cognitive decline and biomarker evidence of Alzheimer’s disease-Data from three memory clinic studies, Alzheimers Dement 15 (2019) 185–193. [DOI] [PubMed] [Google Scholar]
  • [60].Huang Y-M, Shen J, Zhao H-L, Major clinical trials failed the amyloid hypothesis of Alzheimer’s disease, J. Am. Geriatr. Soc 67 (2019) 841–844. [DOI] [PubMed] [Google Scholar]
  • [61].Inayathullah M, Teplow DB, Structural dynamics of the ΔE22 (Osaka) familial Alzheimer’s disease-linked amyloid β-protein, Amyloid 18 (2011) 98–107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [62].Jack CR Jr, Bennett DA, Blennow K, Carrillo MC, Dunn B, Haeberlein SB, Holtzman DM, Jagust W, Jessen F, Karlawish J, Liu E, Molinuevo JL, Montine T, Phelps C, Rankin KP, Rowe CC, Scheltens P, Siemers E, Snyder HM, Sperling R and Contributors, NIA-AA Research Framework: Toward a biological definition of Alzheimer’s disease, Alzheimers Dement 14 (2018) 535–562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [63].Jensen AA, Frølund B, Liljefors T, Krogsgaard-Larsen P, Neuronal Nicotinic Acetylcholine Receptors: Structural Revelations, Target Identifications, and Therapeutic Inspirations, J. Med. Chem 48 (2005) 4705–4745. [DOI] [PubMed] [Google Scholar]
  • [64].Kaczorowski CC, Sametsky E, Shah S, Vassar R, Disterhoft JF, Mechanisms underlying basal and learning-related intrinsic excitability in a mouse model of Alzheimer’s disease, Neurobiol. Aging 32 (2011) 1452–1465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [65].Kerrigan TL, Brown JT, Randall AD, Characterization of altered intrinsic excitability in hippocampal CA1 pyramidal cells of the Abeta-overproducing PDAPP mouse, Neuropharmacology 79 (2014) 515–524. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [66].Khan GM, Tong M, Jhun M, Arora K, Nichols RA, beta-Amyloid activates presynaptic alpha7 nicotinic acetylcholine receptors reconstituted into a model nerve cell system: involvement of lipid rafts, Eur. J. Neurosci 31 (2010) 788–796. [DOI] [PubMed] [Google Scholar]
  • [67].Khiroug SS, Harkness PC, Lamb PW, Sudweeks SN, Khiroug L, Millar NS, Yakel JL, Rat nicotinic ACh receptor α7 and β2 subunits co-assemble to form functional heteromeric nicotinic receptor channels, J. Physiol 540 (2002) 425–434. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [68].Kihara T, Shimohama S, Sawada H, Kimura J, Kume T, Kochiyama H, Maeda T, Akaike A, Nicotinic receptor stimulation protects neurons against beta-amyloid toxicity, Ann. Neurol 42 (1997) 159–163. [DOI] [PubMed] [Google Scholar]
  • [69].Kihara T, Shimohama S, Urushitani M, Sawada H, Kimura J, Kume T, Maeda T, Akaike A, Stimulation of alpha4beta2 nicotinic acetylcholine receptors inhibits beta-amyloid toxicity, Brain Res 792 (1998) 331–334. [DOI] [PubMed] [Google Scholar]
  • [70].Kihara T, Shimohama S, Sawada H, Honda K, Nakamizo T, Shibasaki H, Kume T, Akaike A, alpha 7 nicotinic receptor transduces signals to phosphatidylinositol 3-kinase to block A beta-amyloid-induced neurotoxicity, J. Biol. Chem 276 (2001) 13541–13546. [DOI] [PubMed] [Google Scholar]
  • [71].King JR, Ullah A, Bak E, Jafri MS, Kabbani N, Ionotropic and metabotropic mechanisms of allosteric modulation of alpha7 nicotinic receptor intracellular calcium, Mol. Pharm 93 (2018) 601–611. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [72].Kutoku Y, Ohsawa Y, Kuwano R, Ikeuchi T, Inoue H, Ataka S, Shimada H, Mori H, Sunada Y, A Second Pedigree with Amyloid-less Familial Alzheimer’s Disease Harboring an Identical Mutation in the Amyloid Precursor Protein Gene (E693delta), Intern. Med 54 (2015) 205–208. [DOI] [PubMed] [Google Scholar]
  • [73].Lacor PN, Buniel MC, Chang L, Fernandez SJ, Gong Y, Viola KL, Lambert MP, Velasco PT, Bigio EH, Finch CE, Krafft GA, Klein WL, Synaptic Targeting by Alzheimer’s-Related Amyloid β Oligomers, J. Neurosci 24 (2004) 10191–10200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [74].Lasala M, Fabiani C, Corradi J, Antollini S, Bouzat C, Molecular Modulation of Human alpha7 Nicotinic Receptor by Amyloid-beta Peptides, Front Cell Neurosci. 13 (2019) 37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [75].Lesné SE, Sherman MA, Grant M, Kuskowski M, Schneider JA, Bennett DA, Ashe KH, Brain amyloid-β oligomers in ageing and Alzheimer’s disease, Brain 136 (2013) 1383–1398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [76].Liu Q, Kawai H, Berg DK, beta -Amyloid peptide blocks the response of alpha 7-containing nicotinic receptors on hippocampal neurons, Proc. Natl. Acad. Sci. USA 98 (2001) 4734–4739. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [77].Liu Q, Huang Y, Shen J, Steffensen S, Wu J, Functional alpha7beta2 nicotinic acetylcholine receptors expressed in hippocampal interneurons exhibit high sensitivity to pathological level of amyloid beta peptides, BMC Neurosci. 13 (2012) 155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [78].Liu Q, Huang Y, Xue F, Simard A, DeChon J, Li G, Zhang J-L, Lucero L, Wang M, Sierks M, Hu G, Chang Y.-c, Lukas RJ, Wu J, A novel nicotinic acetylcholine receptor subtype in basal forebrain cholinergic neurons with high sensitivity to amyloid peptides, J. Neurosci 29 (2009) 918–929. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [79].Luckhaus C, Janner M, Cohnen M, Fluss MO, Teipel SJ, Grothe M, Hampel H, Kornhuber J, Ruther E, Peters O, Supprian T, Gaebel W, Modder U, Wittsack HJ, A novel MRI-biomarker candidate for Alzheimer’s disease composed of regional brain volume and perfusion variables, Eur. J. Neurol 17 (2010) 1437–1444. [DOI] [PubMed] [Google Scholar]
  • [80].Luckhaus C, Cohnen M, Fluss MO, Janner M, Grass-Kapanke B, Teipel SJ, Grothe M, Hampel H, Peters O, Kornhuber J, Maier W, Supprian T, Gaebel W, Modder U, Wittsack HJ, The relation of regional cerebral perfusion and atrophy in mild cognitive impairment (MCI) and early Alzheimer’s dementia, Psychiatry Res 183 (2010) 44–51. [DOI] [PubMed] [Google Scholar]
  • [81].Luetje C, Patrick J, Both alpha- and beta-subunits contribute to the agonist sensitivity of neuronal nicotinic acetylcholine receptors, J. Neurosci 11 (1991) 837–845. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [82].Lukas RJ, Changeux J-P, le Novère N, Albuquerque EX, Balfour DJK, Berg DK, Bertrand D, Chiappinelli VA, Clarke PBS, Collins AC, Dani JA, Grady SR, Kellar KJ, Lindstrom JM, Marks MJ, Quik M, Taylor PW, Wonnacott S, International Union of Pharmacology. XX. Current Status of the Nomenclature for Nicotinic Acetylcholine Receptors and Their Subunits, Pharmacol. Rev 51 (1999) 397–401. [PubMed] [Google Scholar]
  • [83].Mann DM, The pathological association between Down syndrome and Alzheimer disease, Mech. Ageing Dev 43 (1988) 99–136. [DOI] [PubMed] [Google Scholar]
  • [84].Marchesi VT, Alzheimer’s Disease 2012: The Great Amyloid Gamble, Am. J. Pathol 180 (2012) 1762–1767. [DOI] [PubMed] [Google Scholar]
  • [85].Marks MJ, Laverty DS, Whiteaker P, Salminen O, Grady SR, McIntosh JM, Collins AC, John Daly’s compound, epibatidine, facilitates identification of nicotinic receptor subtypes, J. Mol. Neurosci 40 (2010) 96–104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [86].Martin SE, de Fiebre NE, de Fiebre CM, The alpha7 nicotinic acetylcholine receptor-selective antagonist, methyllycaconitine, partially protects against beta-amyloid1-42 toxicity in primary neuron-enriched cultures, Brain Res 1022 (2004) 254–256. [DOI] [PubMed] [Google Scholar]
  • [87].Masliah E, Mechanisms of synaptic dysfunction in Alzheimer’s disease, Histol. Histopathol 10 (1995) 509–519. [PubMed] [Google Scholar]
  • [88].McDade E, Why amyloid is still a target for Alzheimer disease clinical trials, J. Am. Geriatr. Soc 67 (2019) 845–847. [DOI] [PubMed] [Google Scholar]
  • [89].McDade E, Reply to: Major Clinical Trials Failed the Amyloid Hypothesis of Alzheimer’s Disease, J. Am. Geriatr. Soc 67 (2019) 848–849. [DOI] [PubMed] [Google Scholar]
  • [90].Mehta TK, Dougherty JJ, Wu J, Choi CH, Khan GM, Nichols RA, Defining presynaptic nicotinic receptors regulated by beta amyloid in mouse cortex and hippocampus with receptor null mutants, J. Neurochem 109 (2009) 1452–1458. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [91].Mesulam MM, Cholinergic circuitry of the human nucleus basalis and its fate in Alzheimer’s disease, J. Comp. Neurol 521 (2013) 4124–4144. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [92].Miller DL, Papayannopoulos IA, Styles J, Bobin SA, Lin YY, Biemann K, Iqbal K, Peptide compositions of the cerebrovascular and senile plaque core amyloid deposits of Alzheimer’s disease, Arch. Biochem Biophys 301 (1993) 41–52. [DOI] [PubMed] [Google Scholar]
  • [93].Minkeviciene R, Rheims S, Dobszay MB, Zilberter M, Hartikainen J, Fulop L, Penke A, Zilberter Y, Harkany T, Pitkanen A, Tanila H, Amyloid beta-induced neuronal hyperexcitability triggers progressive epilepsy, J. Neurosci 29 (2009) 3453–3462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [94].Moretti M, Zoli M, George AA, Lukas RJ, Pistillo F, Maskos U, Whiteaker P, Gotti C, The novel alpha7beta2-nicotinic acetylcholine receptor subtype is expressed in mouse and human basal forebrain: biochemical and pharmacological characterization, Mol. Pharm 86 (2014) 306–317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [95].Murray TA, Bertrand D, Papke RL, George AA, Pantoja R, Srinivasan R, Liu Q, Wu J, Whiteaker P, Lester HA, Lukas RJ, α7β2 Nicotinic Acetylcholine Receptors Assemble, Function, and Are Activated Primarily via Their α7-α7 Interfaces, Mol. Pharmacol 81 (2012) 175–188. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [96].Nelson LD, Scheibel KE, Ringman JM, Sayre JW, An experimental approach to detecting dementia in Down syndrome: a paradigm for Alzheimer’s disease, Brain Cogn. 64 (2007) 92–103. [DOI] [PubMed] [Google Scholar]
  • [97].Nelson PT, Braak H, Markesbery WR, Neuropathology and Cognitive Impairment in Alzheimer Disease: A Complex but Coherent Relationship, J. Neuropathol. Exp. Neurol 68 (2009) 1–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [98].Nelson PT, Jicha GA, Schmitt FA, Liu H, Davis DG, Mendiondo MS, Abner EL, Markesbery WR, Clinicopathologic Correlations in a Large Alzheimer Disease Center Autopsy Cohort: Neuritic Plaques and Neurofibrillary Tangles “Do Count” When Staging Disease Severity, J. Neuropathol. Exp. Neurol 66 (2007) 1136–1146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [99].Nichols RA, Gulisano W, Puzzo D, Editorial: Beta Amyloid: From Physiology to Pathogenesis, Front Mol. Neurosci 15 (2022), 876224. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [100].Nielsen BE, Minguez T, Bermudez I, Bouzat C, Molecular function of the novel α7β2 nicotinic receptor, Cell. Mol. Life Sci 75 (2018) 2457–2471. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [101].Nimmrich V, Grimm C, Draguhn A, Barghorn S, Lehmann A, Schoemaker H, Hillen H, Gross G, Ebert U, Bruehl C, Amyloid beta oligomers (A beta(1-42) globulomer) suppress spontaneous synaptic activity by inhibition of P/Q-type calcium currents, J. Neurosci 28 (2008) 788–797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [102].Nordberg A, Neuroreceptor changes in Alzheimer disease, Cereb. Brain Metab. Rev 4 (1992) 303–328. [PubMed] [Google Scholar]
  • [103].Palmert MR, Podlisny MB, Witker DS, Oltersdorf T, Younkin LH, Selkoe DJ, Younkin SG, The beta-amyloid protein precursor of Alzheimer disease has soluble derivatives found in human brain and cerebrospinal fluid, Proc. Natl. Acad. Sci. USA 86 (1989) 6338–6342. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [104].Palop JJ, Mucke L, Amyloid-beta-induced neuronal dysfunction in Alzheimer’s disease: from synapses toward neural networks, Nat. Neurosci 13 (2010) 812–818. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [105].Palop JJ, Chin J, Roberson ED, Wang J, Thwin MT, Bien-Ly N, Yoo J, Ho KO, Yu GQ, Kreitzer A, Finkbeiner S, Noebels JL, Mucke L, Aberrant excitatory neuronal activity and compensatory remodeling of inhibitory hippocampal circuits in mouse models of Alzheimer’s disease, Neuron 55 (2007) 697–711. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [106].Panza F, Lozupone M, Dibello V, Greco A, Daniele A, Seripa D, Logroscino G, Imbimbo BP, Are antibodies directed against amyloid-β (Aβ) oligomers the last call for the Aβ hypothesis of Alzheimer’s disease? Immunotherapy 11 (2019) 3–6. [DOI] [PubMed] [Google Scholar]
  • [107].Papke RL, Thinschmidt JS, The correction of alpha7 nicotinic acetylcholine receptor concentration-response relationships in Xenopus oocytes, Neurosci. Lett 256 (1998) 163–166. [DOI] [PubMed] [Google Scholar]
  • [108].Papke RL, Porter Papke JK, Comparative pharmacology of rat and human α7 nAChR conducted with net charge analysis, Br. J. Pharm 137 (2002) 49–61. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [109].Papke RL, Stokes C, Williams DK, Wang J, Horenstein NA, Cysteine accessibility analysis of the human alpha7 nicotinic acetylcholine receptor ligand-binding domain identifies L119 as a gatekeeper, Neuropharmacology 60 (2011) 159–171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [110].Perry EK, Gibson PH, Blessed G, Perry RH, Tomlinson BE, Neurotransmitter enzyme abnormalities in senile dementia: Choline acetyltransferase and glutamic acid decarboxylase activities in necropsy brain tissue, J. Neurol. Sci 34 (1977) 247–265. [DOI] [PubMed] [Google Scholar]
  • [111].Peters O and Nitschmann S (2023) [Lecanemab in the treatment of Alzheimer’s disease]. Inn Med (Heidelb) . [DOI] [PubMed] [Google Scholar]
  • [112].Pettit DL, Shao Z, Yakel JL, beta-Amyloid(1-42) peptide directly modulates nicotinic receptors in the rat hippocampal slice, J. Neurosci 21 (2001) Rc120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [113].Picciotto Marina R, Higley Michael J, Mineur Yann S, Acetylcholine as a Neuromodulator: Cholinergic Signaling Shapes Nervous System Function and Behavior, Neuron 76 (2012) 116–129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [114].Portelius E, Price E, Brinkmalm G, Stiteler M, Olsson M, Persson R, Westman-Brinkmalm A, Zetterberg H, Simon AJ, Blennow K, A novel pathway for amyloid precursor protein processing, Neurobiol. Aging 32 (2011) 1090–1098. [DOI] [PubMed] [Google Scholar]
  • [115].Prohovnik I, Perl DP, Davis KL, Libow L, Lesser G, Haroutunian V, Dissociation of neuropathology from severity of dementia in late-onset Alzheimer disease, Neurology 66 (2006) 49–55. [DOI] [PubMed] [Google Scholar]
  • [116].Puzzo D, Arancio O, Amyloid-beta peptide: Dr. Jekyll or Mr. Hyde? J. Alzheimers Dis 33 (Suppl 1:S) (2013) 111–120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [117].Puzzo D, Gulisano W, Palmeri A, Arancio O, Rodent models for Alzheimer’s disease drug discovery, Expert Opin. Drug Disco 10 (2015) 703–711. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [118].Puzzo D, Privitera L, Leznik E, Fa M, Staniszewski A, Palmeri A, Arancio O, Picomolar amyloid-beta positively modulates synaptic plasticity and memory in hippocampus, J. Neurosci 28 (2008) 14537–14545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [119].Puzzo D, Privitera L, Fa M, Staniszewski A, Hashimoto G, Aziz F, Sakurai M, Ribe M, Troy CM, Mercken M, Jung SS, Palmeri A, Arancio O, Endogenous amyloid-beta is necessary for hippocampal synaptic plasticity and memory, Ann. Neurol 69 (2011) 819–830. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [120].Puzzo D, Argyrousi EK, Staniszewski A, Zhang H, Calcagno E, Zuccarello E, Acquarone E, Fa M, Li Puma DD, Grassi C, D’Adamio L, Kanaan NM, Fraser PE, Arancio O, Tau is not necessary for amyloid-beta-induced synaptic and memory impairments, J. Clin. Invest 130 (2020) 4831–4844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [121].Pym L, Kemp M, Raymond-Delpech V, Buckingham S, Boyd CA, Sattelle D, Subtype-specific actions of beta-amyloid peptides on recombinant human neuronal nicotinic acetylcholine receptors (alpha7, alpha4beta2, alpha3beta4) expressed in Xenopus laevis oocytes, Br. J. Pharm 146 (2005) 964–971. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [122].Qi XL, Nordberg A, Xiu J, Guan ZZ, The consequences of reducing expression of the alpha7 nicotinic receptor by RNA interference and of stimulating its activity with an alpha7 agonist in SH-SY5Y cells indicate that this receptor plays a neuroprotective role in connection with the pathogenesis of Alzheimer’s disease, Neurochem Int 51 (2007) 377–383 [DOI] [PubMed] [Google Scholar]
  • [123].Scheltens P, Blennow K, Breteler MMB, de Strooper B, Frisoni GB, Salloway S, Van der Flier WM, Alzheimer’s disease, Lancet 388 (2016) 505–517. [DOI] [PubMed] [Google Scholar]
  • [124].Schoepfer R, Conroy WG, Whiting P, Gore M, Lindstrom J, Brain α-bungarotoxin binding protein cDNAs and MAbs reveal subtypes of this branch of the ligand-gated ion channel gene superfamily, Neuron 5 (1990) 35–48. [DOI] [PubMed] [Google Scholar]
  • [125].Selkoe DJ, Alzheimer’s disease is a synaptic failure, Science 298 (2002) 789–791. [DOI] [PubMed] [Google Scholar]
  • [126].Selkoe DJ, Hardy J, The amyloid hypothesis of Alzheimer’s disease at 25 years, EMBO Mol. Med 8 (2016) 595–608. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [127].Seo J, Kim S, Kim H, Park CH, Jeong S, Lee J, Choi SH, Chang K, Rah J, Koo J, Kim E, Suh Y, Effects of nicotine on APP secretion and Abeta- or CT(105)-induced toxicity, Biol. Psychiatry 49 (2001) 240–247. [DOI] [PubMed] [Google Scholar]
  • [128].Sevigny J, Chiao P, Bussière T, Weinreb PH, Williams L, Maier M, Dunstan R, Salloway S, Chen T, Ling Y, O’Gorman J, Qian F, Arastu M, Li M, Chollate S, Brennan MS, Quintero-Monzon O, Scannevin RH, Arnold HM, Engber T, Rhodes K, Ferrero J, Hang Y, Mikulskis A, Grimm J, Hock C, Nitsch RM, Sandrock A, The antibody aducanumab reduces Aβ plaques in Alzheimer’s disease, Nature 537 (2016) 50–56. [DOI] [PubMed] [Google Scholar]
  • [129].Shaw S, Bencherif M, Marrero MB, Janus kinase 2, an early target of alpha 7 nicotinic acetylcholine receptor-mediated neuroprotection against Abeta-(1-42) amyloid, J. Biol. Chem 277 (2002) 44920–44924. [DOI] [PubMed] [Google Scholar]
  • [130].Shimada H, Ataka S, Tomiyama T, Takechi H, Mori H, Miki T, Clinical Course of Patients with Familial Early-Onset Alzheimer’s Disease Potentially Lacking Senile Plaques Bearing the E693Δ Mutation in Amyloid Precursor Protein, Dement. Geriatr. Cogn. Disord 32 (2011) 45–54. [DOI] [PubMed] [Google Scholar]
  • [131].Singh IN, Sorrentino G, Sitar DS, Kanfer JN, -)Nicotine inhibits the activations of phospholipases A2 and D by amyloid beta peptide, Brain Res 800 (1998) 275–281. [DOI] [PubMed] [Google Scholar]
  • [132].Stine WB, Jungbauer L, Yu C, LaDu MJ, Preparing Synthetic Aβ in Different Aggregation States, in: Roberson ED (Ed.), Alzheimer’s Disease and Frontotemporal Dementia: Methods and Protocols, Humana Press, Totowa, NJ, 2011, pp. 13–32. [Google Scholar]
  • [133].Sudweeks SN, Yakel JL, Functional and molecular characterization of neuronal nicotinic ACh receptors in rat CA1 hippocampal neurons, J. Physiol 527 (2000) 515–528. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [134].Svensson AL, Nordberg A, Beta-estradiol attenuate amyloid beta-peptide toxicity via nicotinic receptors, Neuroreport 10 (1999) 3485–3489. [DOI] [PubMed] [Google Scholar]
  • [135].Svensson A-L, Nordberg A, Tacrine and donepezil attenuate the neurotoxic effect of Aβ(25-35) in rat PC12 cells, Neuroreport 9 (1998) 1519–1522. [DOI] [PubMed] [Google Scholar]
  • [136].Terry AV Jr, Buccafusco JJ, The cholinergic hypothesis of age and Alzheimer’s disease-related cognitive deficits: recent challenges and their implications for novel drug development, J. Pharm. Exp. Ther 306 (2003) 821–827. [DOI] [PubMed] [Google Scholar]
  • [137].Thomsen MS, Zwart R, Ursu D, Jensen MM, Pinborg LH, Gilmour G, Wu J, Sher E, Mikkelsen JD, alpha7 and beta2 Nicotinic Acetylcholine Receptor Subunits Form Heteromeric Receptor Complexes that Are Expressed in the Human Cortex and Display Distinct Pharmacological Properties, PLoS One 10 (2015), e0130572. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [138].Tomiyama T, Nagata T, Shimada H, Teraoka R, Fukushima A, Kanemitsu H, Takuma H, Kuwano R, Imagawa M, Ataka S, Wada Y, Yoshioka E, Nishizaki T, Watanabe Y, Mori H, A new amyloid β variant favoring oligomerization in Alzheimer’s-type dementia, Ann. Neurol 63 (2008) 377–387. [DOI] [PubMed] [Google Scholar]
  • [139].Tong M, Arora K, White MM, Nichols RA, Role of key aromatic residues in the ligand-binding domain of alpha7 nicotinic receptors in the agonist action of beta-amyloid, J. Biol. Chem 286 (2011) 34373–34381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [140].Tropea MR, Li Puma DD, Melone M, Gulisano W, Arancio O, Grassi C, Conti F, Puzzo D, Genetic deletion of alpha7 nicotinic acetylcholine receptors induces an age-dependent Alzheimer’s disease-like pathology, Prog. Neurobiol 206 (2021), 102154. [DOI] [PubMed] [Google Scholar]
  • [141].Ubhi K, Masliah E, Alzheimer’s disease: recent advances and future perspectives, J. Alzheimers Dis 33 (Suppl 1:S) (2013) 185–194. [DOI] [PubMed] [Google Scholar]
  • [142].Ulrich D, Amyloid-beta Impairs Synaptic Inhibition via GABA(A) Receptor Endocytosis, J. Neurosci 35 (2015) 9205–9210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [143].Uteshev VV, Meyer EM, Papke RL, Activation and inhibition of native neuronal alpha-bungarotoxin-sensitive nicotinic ACh receptors, Brain Res 948 (2002) 33–46. [DOI] [PubMed] [Google Scholar]
  • [144].Uteshev VV, Meyer EM, Papke RL, Regulation of neuronal function by choline and 4OH-GTS-21 through alpha 7 nicotinic receptors, J. Neurophysiol 89 (2003) 1797–1806. [DOI] [PubMed] [Google Scholar]
  • [145].Verret L, [Repairing rhythms in the brain of Alzheimer’s mouse models], Med Sci. (Paris) 28 (2012) 1044–1047. [DOI] [PubMed] [Google Scholar]
  • [146].Walsh DM, Klyubin I, Fadeeva JV, Cullen WK, Anwyl R, Wolfe MS, Rowan MJ, Selkoe DJ, Naturally secreted oligomers of amyloid beta protein potently inhibit hippocampal long-term potentiation in vivo, Nature 416 (2002) 535–539. [DOI] [PubMed] [Google Scholar]
  • [147].Wang HY, Lee DH, Davis CB, Shank RP, Amyloid peptide Abeta(1-42) binds selectively and with picomolar affinity to alpha7 nicotinic acetylcholine receptors, J. Neurochem 75 (2000) 1155–1161. [DOI] [PubMed] [Google Scholar]
  • [148].Wang HY, Lee DH, D’Andrea MR, Peterson PA, Shank RP, Reitz AB, beta-Amyloid(1-42) binds to alpha7 nicotinic acetylcholine receptor with high affinity. Implications for Alzheimer’s disease pathology, J. Biol. Chem 275 (2000) 5626–5632. [DOI] [PubMed] [Google Scholar]
  • [149].Whitehouse PJ, The cholinergic deficit in Alzheimer’s disease, J. Clin. Psychiatry 59 (Suppl 13) (1998) 19–22. [PubMed] [Google Scholar]
  • [150].Whitehouse PJ, Saini V, Making the case for the accelerated withdrawal of aducanumab, J. Alzheimer’S. Dis 87 (2022) 999–1001. [DOI] [PubMed] [Google Scholar]
  • [151].Whitehouse PJ, Price DL, Struble RG, Clark AW, Coyle JT, Delon MR, Alzheimer’s disease and senile dementia: loss of neurons in the basal forebrain, Science 215 (1982) 1237–1239. [DOI] [PubMed] [Google Scholar]
  • [152].Williams DK, Stokes C, Horenstein NA, Papke RL, The effective opening of nicotinic acetylcholine receptors with single agonist binding sites, J. Gen. Physiol 137 (2011) 369–384. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [153].Wong W, Economic burden of Alzheimer disease and managed care considerations, Am. J. Manag. Care 26 (2020) S177–sl83. [DOI] [PubMed] [Google Scholar]
  • [154].Yang SY, Chiu MJ, Chen TF, Horng HE, Detection of Plasma Biomarkers Using Immunomagnetic Reduction: A Promising Method for the Early Diagnosis of Alzheimer’s Disease, Neurol. Ther 6 (2017) 37–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [155].Yates CM, Simpson J, Maloney AF, Gordon A, Reid AH, Alzheimer-like cholinergic deficiency in Down syndrome, Lancet 2 (1980) 979. [DOI] [PubMed] [Google Scholar]
  • [156].Yokoyama M, Kobayashi H, Tatsumi L, Tomita T, Mouse Models of Alzheimer’s Disease, Front Mol. Neurosci 15 (2022), 912995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [157].Zaborszky L, Gombkoto P, Varsanyi P, Gielow MR, Poe G, Role LW, Ananth M, Rajebhosale P, Talmage DA, Hasselmo ME, Dannenberg H, Minces VH, Chiba AA, Specific Basal Forebrain-Cortical Cholinergic Circuits Coordinate Cognitive Operations, J. Neurosci 38 (2018) 9446–9458. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [158].Zhao XL, Wang WA, Tan JX, Huang JK, Zhang X, Zhang BZ, Wang YH, YangCheng HY, Zhu HL, Sun XJ, Huang FD, Expression of beta-amyloid induced age-dependent presynaptic and axonal changes in Drosophila, J. Neurosci 30 (2010) 1512–1522. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [159].Lee H Zhu G, Castellani X, Nunomura RJ, Perry A, Smith MAG, Amyloid-β in Alzheimer Disease: The Null versus the Alternate Hypotheses, J. Pharmacol. Exp. Ther 321 (2007) 823–829. [DOI] [PubMed] [Google Scholar]
  • [160].Zoli M, Pistillo F, Gotti C, Diversity of native nicotinic receptor subtypes in mammalian brain, Neuropharmacology 96 (2015) 302–311. [DOI] [PubMed] [Google Scholar]
  • [161].Zwart R, Strotton M, Ching J, Asdes PC, Sher E, Unique pharmacology of heteromeric α7β2 nicotinic acetylcholine receptors expressed in Xenopus laevis oocytes, Eur. J. Pharm 726 (2014) 77–86. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

Data will be made available on request.

RESOURCES