Next Article in Journal
Targeting Conserved Pathways in 3D Spheroid Formation of Diverse Cell Types for Translational Application: Enhanced Functional and Antioxidant Capacity
Next Article in Special Issue
A Systematic Review on Drugs Acting as Nicotinic Acetylcholine Receptor Agonists in the Treatment of Dementia
Previous Article in Journal
Autoimmune Limbic Encephalitis in Patients with Hematologic Malignancies after Haploidentical Hematopoietic Stem Cell Transplantation with Post-Transplant Cyclophosphamide
Previous Article in Special Issue
α7 Nicotinic Acetylcholine Receptors May Improve Schwann Cell Regenerating Potential via Metabotropic Signaling Pathways
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nicotinic Acetylcholine Receptor Dysfunction in Addiction and in Some Neurodegenerative and Neuropsychiatric Diseases

by
Ana Sofía Vallés
1 and
Francisco J. Barrantes
2,*
1
Bahía Blanca Institute of Biochemical Research (UNS-CONICET), Bahía Blanca 8000, Argentina
2
Biomedical Research Institute (BIOMED), Faculty of Medical Sciences, Pontifical Catholic University of Argentina—National Scientific and Technical Research Council, Av. Alicia Moreau de Justo 1600, Buenos Aires C1107AFF, Argentina
*
Author to whom correspondence should be addressed.
Cells 2023, 12(16), 2051; https://doi.org/10.3390/cells12162051
Submission received: 3 June 2023 / Revised: 20 July 2023 / Accepted: 10 August 2023 / Published: 11 August 2023
(This article belongs to the Special Issue Nicotinic Receptors in Health and Disease)

Abstract

:
The cholinergic system plays an essential role in brain development, physiology, and pathophysiology. Herein, we review how specific alterations in this system, through genetic mutations or abnormal receptor function, can lead to aberrant neural circuitry that triggers disease. The review focuses on the nicotinic acetylcholine receptor (nAChR) and its role in addiction and in neurodegenerative and neuropsychiatric diseases and epilepsy. Cholinergic dysfunction is associated with inflammatory processes mainly through the involvement of α7 nAChRs expressed in brain and in peripheral immune cells. Evidence suggests that these neuroinflammatory processes trigger and aggravate pathological states. We discuss the preclinical evidence demonstrating the therapeutic potential of nAChR ligands in Alzheimer disease, Parkinson disease, schizophrenia spectrum disorders, and in autosomal dominant sleep-related hypermotor epilepsy. PubMed and Google Scholar bibliographic databases were searched with the keywords indicated below.

1. Introduction

Nicotinic acetylcholine receptors (nAChRs) are members of the pentameric ligand-gated ion (cation) channel (pLGIC) superfamily, which includes neurotransmitter receptors in metazoa and other ion channels in prokaryota [1]. The various nAChR subunits (α1–α10, β1–β4, γ, ε, and δ) are encoded by 17 genes in vertebrates. nAChRs typically assemble in a combination of two pairs of αβ subunits and an accessory subunit, giving rise to a hetero-pentameric structure [2]. Homopentameric structures containing only α subunits are possible as well [2], allowing for a wide combinatorial diversity of neuronal and muscle-type nAChRs (fetal and adult) with distinctive pharmacological and biophysical properties [3,4,5,6,7,8].
nAChRs occur in multiple dynamic conformational states. In the absence of a ligand, the receptor rests with its ion channel in a closed state. In the presence of an agonist, the receptor protein rapidly shifts to an open state that allows the influx of small cations. The open channel conformer can either return to the closed state or transit to a desensitized (and closed) state. In the desensitized state, the nAChR is unable to be activated by ligand binding.
During brain development, neuronal nAChRs contribute to neurogenesis, neurite outgrowth, and synaptic maturation [9,10,11]. In the human brain, neuronal nAChRs are commonly found in the basal forebrain, hippocampus, cerebellum, and temporal cortex. These brain locations participate in learning, cognition, and memory [11]. Among neuronal nAChRs, heteromeric α4β2 nAChRs and homomeric α7 nAChRs are the most abundant subtypes, whereas combinations such as α3β4, α3β2, and α6β2β3 nAChRs are less common and are generally restricted to specific brain regions [12,13].
Cholinergic signaling is actively involved in the modulation of the finely tuned balance between excitatory and inhibitory neurotransmission in the brain. nAChR activation by the endogenous neurotransmitter acetylcholine (ACh) promotes the release of Ca2+ from intracellular stores and the induction of long-term potentiation (LTP) that favors a depolarized state of the neuron. Activation of nAChRs at pre-synaptic compartments favors the release of several neurotransmitters, including dopamine (DA), norepinephrine, γ-aminobutyric acid (GABA), and glutamate (Glu) [14]. Activation of α4β2 nAChRs allows the modulation of synaptic architecture by regulating the abundance of dendritic spines and heterologous synaptogenesis [15]. Post-synaptically located α7 nAChRs can regulate Glu receptors, thereby modulating synaptic plasticity and GABAergic interneuron activity [16,17]. Furthermore, α7 nAChRs can modulate network excitability, as either their activation or inhibition at the prelimbic cortex promotes the induction of LTP [18]. Thus, cholinergic receptors have a major role in the regulation of neural excitability and plasticity. Likewise, impaired depolarization of the postsynaptic membrane affects communication between neurons [19].
nAChR are expressed in non-neuronal cells as well; α7 nAChRs are present on peripheral immune cells, in astrocytes, in microglia, and in endothelial cells where they mediate neuroprotection and the inflammatory response to different insults [20,21,22].
α7 and α4β2 nAChRs have been associated with the pathogenesis of a range of neurological disorders, e.g., Alzheimer disease (AD) [23], schizophrenia spectrum disorders [24], Parkinson disease (PD) [25], nocturnal frontal lobe epilepsy [26], autism spectrum disorder [27], attention deficit hyperactivity disorder [28], and depression [29], to name some of the most relevant ones. Dysfunctional nAChRs have been implicated in non-neurological diseases as well, e.g., small-cell lung carcinomas and diabetes [30]. The predominant α7 and α4β2 nAChRs and the role of other nAChR subtypes in different diseases is currently the subject of basic and clinical research. One example is involvement of the α9 nAChR subunit in the pathophysiology of neuropathic pain [31,32] and participation of the α6 subunit in sensory processing and pain [31,33]. Muscle-type nAChR dysfunctions cover a wide clinical spectrum. Those due to inherited mutations can be associated with muscle weakness, myasthenia gravis, or congenital myasthenic syndromes [34,35]. Knowledge of nAChR involvement in neuromuscular, neurological, and psychiatric disorders makes these receptors critical targets for drug development [3,27,36,37,38,39].
Understanding the complex neurobiological mechanisms implicated in addiction is key to the development of a therapeutic protocol/strategy for the treatment of what is today considered a chronic neuropsychiatric disorder. Epidemiological studies have provided information on the association of certain neurological diseases, such as PD and schizophrenia, with heavy smoking habits and decreased brain expression of specific nAChRs. Therefore, information on the role of these receptors in such pathophysiological states is the first of many steps leading to therapeutic intervention.
Many if not all neurological diseases present an inflammatory component. nAChRs play a central role in the cholinergic anti-inflammatory pathway and in the regulation of immune functions in AD, PD, and schizophrenia spectrum disorders. Hence, activation of this pathway has emerged as a therapeutic tool in the amelioration of the neuroinflammatory component of these diseases. Neuroinflammatory comorbidities not only decrease the quality of life of the patient, they represent a social, emotional, and financial burden to society. Research on how nAChRs are altered in disease will undoubtedly contribute to the development of therapies to reverse or at least hamper the progression of these debilitating diseases.
The fundamental role played by nAChRs as modulators of neurotransmitter release determines that dysfunction or mutations of the genes encoding these receptor subunits affect many cognitive functions and favor the occurrence of a wide range of pathologies, such as addiction, neuropsychiatric and neurodegenerative diseases and epilepsy, as critically discussed in this review.

2. Addiction

Addiction can be defined as a chronic neuropsychological disorder in the form of an intense compulsive urge to seek immediate sensory rewards; it is characterized by functional alterations in brain circuits that participate in reward, memory, and self-control, leading to maladaptive behaviors [40]. The hallmark of addiction involves what are colloquially known as the four Cs: the appearance of craving, the compulsion to “use”, the loss of control, and adverse consequences [41]. An understanding of the neurobiological mechanisms and central actors involved in the development of the different stages of addiction is fundamental. Today, it is acknowledged that several neurotransmitters and neuromodulators regulate brain reward areas acting at the level of either the ventral tegmental area (VTA) or the nucleus accumbens (NAc). These areas of the brain balance and modulate emotions, stress, and interoception. The intake of any substance that disbalances the multiple neurotransmitter-specific neuroplasticity circuitries will induce neuroadaptations with different effects on the individual [41].
Nicotine (1-methyl-2-[3-pyridyl] pyrrolidine), mainly found in tobacco, is considered a highly addictive drug involving all of the criteria contained in the four Cs [42]. In its uncharged form it can permeate the plasma membrane and enter the brain, where it can change to its charged form to bind and trigger nAChRs [12,43]. By activating a variety of nAChRs located in DAergic neurons in the VTA that projects to the NAc, nicotine promotes the release of DA in the NAc, a key step in the initiation of addiction [43] (Figure 1).
In the VTA, nicotine activates α4β2 nAChRs, promoting GABA release in GABAergic neurons and thereby inhibiting dopamine release. However, because α4β2 nAChRs rapidly desensitize upon agonist binding, the GABAergic flux to the dopaminergic (DAergic) neurons is brief. In parallel, nicotine activates α7 nAChRs in presynaptic neurons of the VTA, promoting Glu release, which can in turn enhance DAergic release [42]. Because α7 nAChRs have lower affinity for nicotine than α4β2 nAChRs, they are less prone to desensitization in the presence of the nicotine concentrations found in the brain upon smoking [44,45,46]. Thus, nicotine addiction is promoted through the combinatory action of a reduction in the inhibitory GABAergic input to DAergic neurons and the potentiation of glutamatergic afferents to dopamine-releasing neurons. Therefore, nAChRs are central actors in the modulation of DA release, and consequently in the initiation of nicotine addiction [47,48,49,50].
Unlike the natural endogenous agonist ACh, which is rapidly hydrolyzed by acetylcholinesterase, nicotine cannot be removed from the synaptic cleft [6,43,51,52]; this constant exposure to nicotine triggers several neuroadaptations. Chronic nicotine exposure leads to nAChR upregulation, with modifications in receptor assembly, trafficking, and degradation that contribute to maintaining adequate brain homeostasis [5,53,54,55]. The neuroadaptations taking place in neurotransmitter systems as a consequence of nicotine exposure are considered to participate directly in nicotine addiction [52,56,57].
Through the mesolimbic pathway, nicotine–nAChR interactions mediate reward and reinforcement effects [42]. The habenula has been found to be involved in the regulation of feelings such as fear, anxiety and depression [43,58,59,60,61]. The aversive effects of nicotine withdrawal are mediated through the medial habenula-interpeduncular (MHb-IPN) [62,63].
The principal nAChR subunits expressed in the mesolimbic pathway are the α4, α6, α7, and β2 subunits, with the α3, α5, and β4 subunits being mostly expressed in the MHb-IPN [64]. It has been hypothesized that α5 subunits comprise ∼20% of functional nAChRs in rat MHb neurons that project to IPN [65]. Knockdown of α5 nAChRs in the MHb-IPN pathway further suggests that nicotine exerts stimulatory effects on α5-containing nAChRs [66]. Nicotine and other addictive substances have been reported to interact with the α3β4 nAChR expressed in the MHb-IPN circuit [67,68], and it has been proposed that they mediate drug- or psychostimulant-seeking behavior [69].
In addition to the many environmental and social factors affecting nicotine addiction, individual genetic factors play an important role [42]. The risk of developing nicotine addiction has been associated with genetic variations in genes that encode for nAChRs, particularly those located in the chromosomal region 15q25 (CHRNA5-CHRNA3-CHRNB4 gene cluster) [42], in chromosome 8 (CHRNB3–CHRNA6 gene cluster) [70], and in CHRNA2 [71]. The cited reports suggest a strong association between single-nucleotide polymorphism (SNPs) in nAChR genes and the number of cigarettes smoked per day, the age onset of daily smoking, and chronic smoking behaviors in adolescence and adulthood. Recent use of knockout/knock-in mice has contributed to our understanding of different behavioral phenotypes related to nicotine addiction. The rewarding DA-mediated effects along with the aversive consequences of nicotine withdrawal preclude active smokers from stopping the habit, and can often motivate relapse after periods of abstinence.
In line with studies showing nAChR involvement in cognition, inhibitory control, and decision-making mechanisms [72], nicotine activation of nAChRs has been shown to enhance attention in animal and human studies [73,74,75]. Nicotine-induced enhancement of cognition has been reported to be weaker in non-smokers than in smokers [76]. The higher incidence of smoking among individuals with psychiatric illnesses such as schizophrenia spectrum disorders may indicate that patients with these conditions smoke to ameliorate the attentional deficits associated with their disease condition [77,78]. In line with the self-medication hypothesis, it has been reported that schizophrenia patients have lower expression of nAChRs [79,80,81] and as such may smoke in order to up-regulate their nAChR and thereby augment their nicotine level [79,80,82].
Many epidemiological and longitudinal studies in recent years have revealed that most tobacco users consume cannabis as well [83,84,85]. Cannabis and nAChR receptors co-distribute in the same brain areas, suggesting that the two systems can engage in cross-talk [86]. Tobacco and cannabis are the most common drugs of abuse consumed by adolescents and young adults [87,88]. The co-use of these drugs has been suggested to produce mutually reinforcing effects and a decrease in adverse effects [89]. Several studies have indicated that consumption of Δ-9-tetrahydrocannabinol (THC), the main addictive component in Cannabis sativa, is associated with anxiogenic-like effects, working memory impairments, and ataxia [88,89]. These adverse THC effects appear to be reduced upon nicotine administration [90,91,92,93,94].
The CHRNA2 gene in chromosome 8 has recently been identified as one of the risk loci for both smoking behavior and nicotine dependence [71,95]. A recent study found that individuals with cannabinoid use disorder present reduced expression of the CHRNA2 gene in the cerebellum, suggesting that the gene that encodes for the α2 nAChR subunit may be involved in the susceptibility to developing this disorder. Furthermore, a negative correlation between the gene expressions of CHRNA2 and CNR1 (cannabinoid receptor 1) in the cerebellar cortex and cerebellar nuclei has been reported [87]. Participation of the homomeric α7 nAChR has been linked to the rewarding effects of cannabinoid use, while the α4β2 nAChR subtype has been associated with a reduction in cannabinoid-induced ataxia, and as such with a reduction in cannabinoid-induced motor impairment. In addition, the potential roles of the α5, α3, and β4 nAChR subunits in cannabinoid use disorder, particularly in tolerance- and withdrawal-associated symptoms, have been addressed [87].
To summarize, nicotine and/or cannabis addiction induces several neuroadaptations involving nAChRs in brain regions that modulate the mesolimbic reward system and the MHb-IPN withdrawal syndrome. Preclinical studies have provided a wealth of information on alterations to the neurocircuitry due to chronic consumption of these substances. The great diversity of nAChR subunits along with genetic differences in gene clusters that code for these subunits are important features of addiction and should be considered jointly in the design of therapeutic approaches.

3. Central and Peripheral Inflammation

The cholinergic system is involved in the modulation of inflammation in the central and peripheral nervous systems [11,31,96]. Cholinergic receptors are expressed in neurons, glial cells (microglia, astrocytes), and immune cells (e.g., macrophages) [11,31,96]. In the nervous system, neuroinflammation is a necessary process to restore the altered homeostasis caused by infections, trauma, and neurodegenerative diseases [97].
In the CNS, neuroinflammation comprises a dynamic multistage physiological response orchestrated by microglia and astrocytes [97]. Both types of cells are needed to support and sustain adequate neuronal function. In most neurodegenerative diseases a chronic inflammatory state is present. Under these circumstances, microglia remain activated for prolonged periods, with detrimental consequences for neuronal cells [97]. Activated microglial cells secrete various inflammatory molecules that may lead to neuronal dysfunction and degeneration [98]. Likewise, during sustained inflammation astrocytes release pro-inflammatory cytokines and prostaglandins that can alter neuronal function and the blood–brain barrier [99].
Both microglia and astroglia express α7 nAChRs, and several studies have demonstrated that this nAChR subtype exerts neuroprotective effects in the brain [20,100,101,102,103]. As such, glial α7 nAChRs are considered potential therapeutic targets in neurodegenerative diseases [104]; furthermore, α7 nAChR agonists have been reported to provide neuroprotection against various toxic insults including β-amyloid [105], MPTP (in vivo) and MPP+- or LPS (in vitro) [106]. The ionotropic activity of α7 nAChRs is neuron-specific. In non-neuronal cells, metabotropic activity is prevalent downstream of α7 nAChR activation [97]. Activation of α7 nAChR expressed in glial cells promotes the activation of phospholipase C (PLC), in turn inducing the enhanced production of inositol trisphosphate (IP3) [97]. This second messenger can bind to its receptor, located in the endoplasmic reticulum, and induce the release of Ca2+, while the cation mediates a decrease in phosphorylation, causing the activation of kinases involved in neuroinflammation. α7 nAChR activation in glial cells regulates the synthesis and release of inflammatory molecules such as TNFα, IL-6, and nitric oxide [98,107,108]. In addition, α7 nAChRs expressed in astrocytes have been shown to mediate anti-inflammatory effects by inhibiting the nuclear factor kappa-light-chain-enhancer of the activated B cell (NFkB) pathway and activation of the nuclear factor erythroid 2-related factor 2 (Nrf2) pathway [104]. Hence, α7 nAChR in glial cells appears to play a significant role in the modulation of neuroinflammation in the CNS (Figure 2).
Peripheral immune cells additionally express nAChRs [109], which can be activated by the endogenous neurotransmitter. Several studies have shown that specific ligands of the α7, α9, and α10-containing nAChRs can modulate the release of inflammatory cytokines from peripheral immune cells [109,110,111,112]. Although the precise mechanisms of signal transduction in these cells have not been fully elucidated, it is known that when ACh binds to α7 nAChRs on cytokine-producing cells, such as macrophages, activation of a signaling cascade via the Janus kinase 2 (JAK2) signal transducer and activator of transcription 3 (STAT3) takes place [100,113,114]. As a result, STAT3 translocates to the cell nucleus and interferes with the binding of NFkB to the DNA (Figure 2). The latter event prevents the transcription of genes that encode for inflammatory cytokines such as interleukin 1β, 6, and 8, TNF-α, or monocyte chemoattractant protein-1 (MCP1) (Figure 2).
An alternative anti-inflammatory mechanism initiated through the JAK2-STAT3 pathway [109,111,112,115] has been described. Activation of this pathway would promote the expression of interleukin-1 receptor-associated kinase M (IRAK-M), which can negatively regulate the innate Toll-like receptor (TLR)-mediated immune responses. The TLRs comprise a family of receptors that are necessary for the initiation of innate immune responses [115]. Thus, activation of α7 nAChRs in peripheral immune cells would suppress the production of pro-inflammatory molecules by inhibiting downstream inflammatory signals resulting from TLR activation [109].
nAChRs containing the α9/α10 subunits have been suggested to partake in these anti-inflammatory mechanisms [109]. Activation of α9/α10-containing nAChRs in human monocytes and whole blood cultures inhibit release of IL-1β, TNF-α, and IL-6 [116,117,118]. The mechanism by which these receptors relay this inflammatory protection appears to be the same as that mediated by α7 nAChR through activation of the JAK2/STAT3 pathway [109]. Indeed, the cholinergic anti-inflammatory pathway links the nervous system with the immune system to counteract inflammatory activation [110,119]. The neural circuit by which the vagus nerve interacts with the peripheral immune system to provide anti-inflammatory action involves activation of the splenic nerve. When splenic nerve fibers are activated, they release noradrenaline. Splenic T-cells release ACh in response to noradrenaline binding to β2 adrenergic receptors. In turn, released ACh can activate α7 nAChR in peripheral immune cells (Figure 2) and confer anti-inflammatory protection.
Recently, Simon and coworkers further described that in addition to the already known α7-mediated anti-inflammatory effects via vagus nerve stimulation, splenic nerve terminals that release noradrenaline can interact directly with noradrenergic receptors in splenic myeloid cells and exert anti-inflammatory effects [120]. Vagus nerve stimulation therapy was introduced in the 1980s to treat epilepsy [121]. The procedure is a non-invasive tool that has been applied more recently to AD [122], PD [123], and schizophrenic [124] patients. Furthermore, preclinical studies performed in rodents have shown that vagus nerve stimulation limits the accumulation of β-amyloid plaques, while clinical studies have shown promising results in the modulation of cognition.
In summary, neuroinflammation constitutes an ubiquitous pathology in the various CNS diseases discussed in this review. As the best characterized nicotinic receptor subtype in the immune system [96], stimulation of α7 nAChR is emerging as a promising target to counteract neuroinflammatory processes and a major contributor to the restoration of CNS homeostasis. A better understanding of the functionality of nAChR in both central and peripheral immune cells and their ability to abrogate inflammatory processes is of great clinical relevance.

4. Alzheimer Disease

AD is considered the most common form of dementia among elderly persons [125]. The development of amyloid senile plaques, containing amyloid peptides, and deposits of neurofibrillary tangles rich in tau protein are pathognomonic postmortem hallmarks of AD [126]. Reports contend that these plaques and deposits occur in the brain long before the clinical manifestations of AD become evident [125,127]. Moreover, it is these pathological alterations that have been suggested to induce neuronal dysfunction associated with clinical dementia, a strong decline in memory and cognitive functions, and a deterioration in the visual and motor coordination manifested in symptomatic AD dementia [128].
Cortical nAChRs are markedly reduced in the brains of AD patients, explaining the cholinergic deficits associated with AD [129]. In particular, altered expression levels and function of α7 nAChR have been described in AD [110,125]. Reduced α7 nAChR levels have been shown to correlate with β-amyloid (Aβ) plaque deposition and cognitive impairment [130,131,132], and there is strong evidence that α7 nAChR interacts directly with the Aβ peptide [125]. The Aβ peptide can bind to α7 nAChR at the surface of neurons with very high (pM) affinity [133]. Endocytic internalization of the α7 nAChR–Aβ complex and the ensuing Aβ aggregation then promotes the phosphorylation of microtubular tau protein, leading to the formation of neurofibrillar tangles [133].
In recent years, innate immune activation has been ascribed an important role in both the pathogenesis and progression of AD. While microglia can interact with Aβ and Aβ precursor protein (APP) through membrane receptors and clear Aβ from the brain through phagocytosis, it can release pro-inflammatory cytokines that result in damage to the surrounding neurons [134]. Therapeutic interventions targeting microglia in neurodegenerative diseases are currently in their infancy [134]. Future therapeutic approaches targeting microglial activation in AD should aim at specifically inhibiting the release of inflammatory factors without interfering with microglia’s beneficial effects on Aβ clearance.
Several genes encoding for immune receptors have been linked to AD development [135]. It is now recognized from preclinical and clinical evidence that systemic inflammation can affect the brain in many ways and can lead to the development of neurodegenerative diseases including AD [135]. Indeed, preclinical data have provided a large body of evidence on the association between peripheral inflammation and AD pathology [135]. Likewise, several clinical reports have described how systemic inflammation caused by specific environmental factors is associated with an increase in cognitive decline in AD [136,137,138]. Activation of α7 nAChRs expressed in rat hippocampal astrocytes was able to counteract this inflammatory scenario by reducing the Aβ protein load [139]. However, Aβ concentration changes as AD progresses; hence, α7 nAChRs may play different roles as the disease develops. At low picomolar concentrations, Aβ triggers the conversion of α7 nAChR to a desensitized conformation that is nevertheless able to respond to agonists and exert anti-inflammatory action, whereas at high nanomolar concentrations Aβ acts as a negative modulator of the receptor and possesses associated neurotoxicity [140,141,142]. Thus, the concentration of Aβ should be critically evaluated in terms of the benefits of α7 nAChR stimulation therapies. Furthermore, because of the strong affinity interaction between Aβ and α7 nAChR, the pharmacological selection of a competitive Aβ antagonist has been challenging [143]. Different drug candidates have emerged, including partial and allosteric modulators of the α7 nAChR. Many trials have been abandoned, however, either because of poor efficacy or high toxicity [143]. There are significant gaps to be filled between preclinical and clinical data in the interests of better AD therapeutic strategies.
The CHRFAM7A gene is exclusively found in humans [142,144,145,146]. This gene is the product of the partial duplication of exons 5 to 10 of the α7 nAChR-encoding gene CHRNA7 [147,148]. The CHRFAM7A human-specific gene that lacks the N-terminal domain of the CHRNA7 subunit codifies the dupα7 protein [142]. Thus, the agonist binding domain is absent in the dupα7 protein. The dupα7 protein cannot by itself assemble into functional nicotinic receptors. However, in combination with at least two α7 nAChR subunits it can form functional ion channels [142], albeit exerting a dominant negative effect on the latter [149]. Furthermore, polymorphism as well as a two base pair deletion of the CHRFAM7A on exon 6 has been described [147]. These genetic modifications translate into the dupΔα7 protein [147]. The number of CHRFAM7A copies varies [142], some individuals being non-carriers of the CHRNA7 duplication and others expressing one or two copies. The occurrence of the Δ2bp allele varies between different ethnic groups [142]. There is evidence that the number of CHRFAM7A copies carried by an individual may affect the response of α7 nAChR-positive allosteric modulators (PAM), agonists, and antagonists [142,150,151,152]. Interestingly, the expression of dupα7 has been associated with a protective role during the accumulating phase of Aβ [150]. The authors cited above describe the presence of CHRFAM7A as mitigating Aβ uptake in cells; thus, its expression could exert a protective role when Aβ concentrations are above physiological levels [150]. Previous reports using a transgenic mouse model of AD [153] on a knockout α7 nAChR mouse [154] showed that deletion of the α7 nAChR gene ameliorated cognitive deficiency and further improved synaptic physiology [155]. Altogether, these studies highlight the importance of α7 nAChR in the pathophysiology of the cognitive impairment associated with AD [133]. CHRFAM7A is expressed in non-neuronal cells as well; therefore, it can alter the anti-inflammatory effects mediated via α7 nAChR activation [149,156]. In one study, LPS-induced inflammatory responses were reported to down-regulate CHRFAM7A expression at the mRNA and protein levels. Conversely, CHRNA7 mRNA was upregulated [156].
Approximately 25% of the AD population are non-carriers of the CHRFAM7A gene. Considering that preclinical drug testing is carried out in animal models, any molecule screened to target the α7 nAChR will benefit a quarter of the population at most [150,151,152]. Therefore, future preclinical models examining the CHRFAM7A gene in greater detail should turn the focus to the rest of the population [151]. Non-carrier individuals should experience better outcomes based on preclinical data of AD drug trials, while more specific research is needed to understand the impact of CHRFAM7A on the pathogenesis of AD.

5. Parkinson Disease

The second most frequent neurodegenerative disease is PD. PD patients exhibit motor deficits, cognitive decline, and sleep and affective disturbances, and they progressively lose DAergic neurons from the substantia nigra (Figure 3).
The most common treatment for PD is replacement therapy with L-DOPA to enhance DA transmission, with beneficial effects on PD-associated motor dysfunction. However, this approach does not suffice to improve other PD-associated symptoms, nor does it help to prevent the progression of the disease. In addition, a common side effect of L-DOPA therapy is the induction of dyskinesias that are incapacitating.
Activation of nAChRs at presynaptic terminals can enhance DA release from DAergic neurons. Promising results for treating PD have been obtained in animal models and human studies involving activation of the nigrostriatal pathway by nAChR agonists [157,158,159,160]. The potential cognitive-enhancing properties of nAChR-targeting drugs may be of additional benefit to those patients suffering cognitive decline [25,159,161].
PD neuroprotection from toxin-induced DAergic cell loss by nAChR agonist administration has been described in animal models [162]. Indeed, the specific activation of α7 nAChRs in several PD animal models has proven beneficial for ameliorating PD-associated symptoms and for its antidyskinetic effects. In contrast, administration of methyllycaconitine, a specific α7 nAChR antagonist, hinders neuroprotection [106,162].
Activation of nAChRs containing the β2 subunit, as is the case with the abundant α4β2 subtype, has shown protection against 6-OHDA-induced nigrostriatal damage in rodents [25]. The nAChR β2 subunit has been suggested to mediate this neuroprotective effect, as nigrostriatal damage was not prevented in α4 subunit-knockout mice models lacking α4β2 nAChRs [163]. The β2 subunit has been found to modulate the expression of induced dyskinesia in several nonhuman primate models [164,165,166,167].
Nicotine-mediated α7 nAChR activation has been shown to produce an inhibitory effect on L-DOPA-induced dyskinetic side effects in nonhuman primate models [168,169]. However, α7 nAChR seems not to be the specific mediator of this effect, as mutant mice lacking the α7 subunit reduce the L-DOPA-induced abnormal movements when exposed to nicotine.
Neuroinflammatory processes are present in PD, and increased density of astrocytes and active microglia is observed as well. Microglial cells initiate the immune response, and astrocytes surround the area so as to localize the secretion of pro-inflammatory cytokines [170,171]. In consequence, when activated, microglia and astrocytes release pro-inflammatory cytokines in PD, and degeneration of dopaminergic neurons can occur [172]. Recent preclinical research has focused on preventing microglia activation to delay the progression of the disease [172,173], although the use of these drugs in clinical practice is far from being a reality. PD patients accumulate α-synuclein in the form of Lewy bodies [174]. This α-synuclein accumulation in PD contributes to neuroinflammation by promoting the release of pro-inflammatory molecules from glial cells, which has neurotoxic effects [174]. Although the origin of α-synuclein aggregation remains uncertain, two hypotheses have been proposed [175]. In the first hypothesis, α-synuclein accumulation is purported to arise in the brain and project into the peripheral autonomic nervous system. A second hypothesis postulates that α-synuclein pathology originates in the gastrointestinal tract and reaches the brain via the vagus nerve [175]. Because α7-nAChR is expressed in glial cells and in peripheral immune cells, this receptor subtype is envisaged as a possible therapeutic target to reduce neuroinflammation in PD. The expression of α7 nAChR in astrocytes is considered a novel therapeutic strategy for the treatment of PD [104]. Likewise, vagus nerve stimulation is becoming more accepted as a non-invasive therapeutic method to tackle neuroinflammation in PD [123].

6. Schizophrenia Spectrum Disorders

Schizophrenia and associated disorders have a profound negative impact on the quality of life of patients. These complex chronic neuropsychiatric disorders have an early onset [176]. The symptoms experienced by schizophrenic patients have been classified as positive symptoms (delusions, hallucinations), negative symptoms (social withdrawal, anhedonia), and cognitive deficits (learning and memory deficits, alogia) [177]. Altered neurotransmission has been proposed as the basic common pathophysiological mechanism in schizophrenia [178]. Augmented levels of pro-inflammatory cytokines have been found in schizophrenic patients as well [179].
Several forms of schizophrenia spectrum disorders have a heritable genetic component. However, no single causative gene has been reported to date. Trubetskoy and coworkers [180] recently studied the genomes of 76,755 individuals with schizophrenia as well as 243,649 healthy (control) participants. Their study demonstrated the occurrence of 342 common genetic variants that could increase the risk of developing schizophrenic disorders. Therefore, as with other neurological and neuropsychiatric diseases, it is likely that a combination of many genes and environmental factors contribute to the pathogenesis of this group of disorders.
Regarding alterations in cholinergic neurotransmission in schizophrenic patients, studies performed in postmortem hippocampus, cortex, thalamus, and striatum of schizophrenic patients have revealed that expression of α7 nAChR is lower than in control patients [24,181]. The lower number of receptors in schizophrenic patients does not appear to be region-specific.
The gene encoding for α7 nAChR is located in chromosome 15q14, which is linked to genetic transmission of schizophrenia spectrum disorders [177,182]. As is the case with AD, carriers of the CHRFAM7A gene mutation are associated with certain forms of this spectrum [142,177,183,184]. Additionally, the reduced expression of α7 nAChRs may result from increased expression and insertion of heteromeric dupα7/α7 nAChRs [177,185,186].
An endophenotype of schizophrenia spectrum disorders is the presence of P50 auditorily evoked response deficits. In these patients, the involvement of a nicotinic receptor is clearly apparent [187], as administration of nicotine transiently normalizes the P50 deficit. Furthermore, a strong link between α7 nAChRs and sensory gating deficits has been reported in some patients with schizophrenia [142,188,189]. Therefore, it is not surprising that these patients use smoking as a form of self-medication (see the above section on Addiction). Indeed, about 80% of patients with schizophrenia consume tobacco products, a figure that dramatically contrasts with 25% among the general population [190,191]. These statistics are underscored by the observation that therapeutic strategies targeting α7 nAChR show beneficial effects [142,192,193,194,195].
The presence of heteromeric dupα7/α7 nAChRs, increased CHRFAM7A, and/or reduced CHRNA7 expression in the prefrontal cortex has been reported in patients with bipolar disorder and schizophrenia spectrum disorders [142,196,197]. However, there are contradictory findings in the literature regarding the association of these variants with some forms of schizophrenia, probably due to differences in the ethnic groups and number of subjects under study and the phenotype under consideration [183,196,198]. Future preclinical trials should include these genetic variants in order to gain greater insight into the possible involvement of nAChR genes as risk factors in the development of schizophrenia spectrum disorders in order to improve therapeutic outcomes.

7. Epilepsy

Epilepsy comprises many syndromes characterized by the chronic occurrence of seizures [199]. The latter results from excessive neuronal activity in the brain [200]. Epileptic syndromes show heterogeneous origins, mechanisms, and clinical manifestations [201]. Because gliosis and microgliosis have been described in epilepsy, it is evident that this malady shares a neuroinflammatory component with other neuronal diseases [202]. Therefore, future preclinical and clinical studies should consider including agents to reduce neuroinflammation in their therapeutic approaches.
A high proportion of individuals presenting alterations in neurodevelopment suffer comorbid seizures. Again, nAChRs have been shown to play an important role in regulation of the excitatory microcircuitry that leads to seizures, and a vast body of evidence implicates nAChR dysregulation in epileptiform activity [199].
Both nAChRs and muscarinic receptors (mAChRs) have been associated with epilepsy, as their hyperstimulation can lead to the onset of seizures [203,204]. Different nAChR subtype mutations have been linked with genetic sleep-related epilepsy [26,199,205,206,207,208]. nAChRs can induce epileptogenic effects both during development and in adult stages due to their participation in synaptogenesis and the regulation of mature synaptic circuit excitability. Hyperactivation of M1 mAChRs upon application of the muscarinic agonist pilocarpine has been used as a model of temporal lobe epilepsy, to induce transient status epilepticus, and to generate chronic epileptic seizures [203,204,209].
In vivo studies have shown that nicotine doses over 2–3 mg/kg in rodents suffice to induce tonic–clonic convulsions [210,211]. Clinical data suggest that seizures can occur after multiple applications of transdermal nicotine patches [212]. In addition, repeated subconvulsive doses of nicotine in mice have been used as kindling agents [213]. Interestingly, it has been suggested that the sex dependency of nicotine-induced kindling is related to the lower availability of antioxidant defenses in females [214]. nAChR antagonists prevent the induction of pro-convulsive activity [215,216,217]. More difficult to explain is the fact that high concentrations of nAChR antagonists can have pro-convulsive activity as well [215,218]. Reports indicate that mutant mice deficient in α5 and/or β4 nAChR subunits are less prone to developing seizures [219,220]. One proposed mechanism of nicotine kindling and activity-dependent nAChR-induced seizures argues that the two pathologies are due to the induction of glutamatergic overactivation [221]. Autosomal dominant nocturnal frontal lobe epilepsy (ADNFLE), now renamed (AD)SHE, or (Autosomal dominant) sleep-related hypermotor epilepsy [222], was first linked to a missense mutation on CHRNA4 (α4S248F) [26,206,223]. More recently, CHRNA2 was found to be linked to ADSHE [224]. In these forms of epilepsy, hypermotor seizures or tonic–dystonic postures that last on average ∼30 s are observed. The new terminology, ADSHE, reinforces the concept that sleep-related seizures are not exclusively nocturnal and enhances the importance of hypermotor seizures as the central feature of the pathology. In addition to motor hyperactivity, ADSHE families that carry the CHRNA4 and CHRNB2 mutations present cognitive disabilities, mental retardation, and schizophrenia-like symptoms [199]. How nAChR-expressing mutant subunits are linked to the onset of the pathologies of ADSHE requires knowledge of the roles these receptors play in multiple neuronal circuits during development and in adulthood.
Deletion of CHRNA7 in mice does not alter the induction of seizures by nicotine administration [225,226]. Furthermore, human-based studies have revealed a genetic predisposition to developing an idiopathic form of generalized epilepsy [227,228] and to some of the neurodevelopmental disorders accompanied by seizures [227,229,230] when microdeletions of the chromosome region 15q13.3 that codes for CHRNA7 are present. These microdeletions have been associated with phenotypes that lead to schizophrenia- and epilepsy-related alterations in animal studies [231]. Additional studies are required in order to understand the possible implications of α7 nAChRs in epilepsy. Deeper knowledge of the molecular mechanisms by which nAChR mutations induce ADSHE in response to agonist/antagonist exposure is essential for the formulation of pharmacological strategies targeting these forms of epilepsy.

8. Concluding Remarks

By regulating neuronal excitability, immunity, inflammation, neuroprotection, and the release of other neurotransmitters or targeting the receptors for other neurotransmitters, nAChRs modulate multiple physiological, behavioral, and pathophysiological processes. Different levels of expression of nAChRs in specific brain regions and at different neurodevelopmental stages can be affected by dysfunction and lead to disease. The design of specific nAChR ligands, including PAMs, able to target specific diseases and be tailored to subtle variations in each pathophysiological scenario, calls for several translational gaps to be filled in preclinical and clinical trials. The inter-individual variability in genes that encode for nAChR subunits needs to be carefully considered in future personalized therapies, along with adequate genetic screening. In addition, nAChR-based or nAChR-targeted therapeutic strategies must include multiple genetic variants in order to improve the potential of these drugs in all populations.

Author Contributions

F.J.B. and A.S.V. conceived the work. A.S.V. searched the literature. F.J.B. and A.S.V. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

Figures were created using Servier Medical Art Commons Attribution 3.0 Unported License. http://smart.servier.com (accessed on 10 May 2023). Servier Medical Art by Servier is licensed under a Creative Commons Attribution 3.0 Unported License.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jaiteh, M.; Taly, A.; Hénin, J. Evolution of Pentameric Ligand-Gated Ion Channels: Pro-Loop Receptors. PLoS ONE 2016, 11, e0151934. [Google Scholar] [CrossRef] [Green Version]
  2. Jones, A.K.; Sattelle, D.B. Diversity of insect nicotinic acetylcholine receptor subunits. Adv. Exp. Med. Biol. 2010, 683, 25–43. [Google Scholar] [CrossRef] [PubMed]
  3. Caton, M.; Ochoa, E.L.M.; Barrantes, F.J. The role of nicotinic cholinergic neurotransmission in delusional thinking. NPJ Schizophr. 2020, 6, 16. [Google Scholar] [CrossRef] [PubMed]
  4. Wessler, I.K.; Kirkpatrick, C.J. The Non-neuronal cholinergic system: An emerging drug target in the airways. Pulm. Pharmacol. Ther. 2001, 14, 423–434. [Google Scholar] [CrossRef]
  5. Giniatullin, R.; Nistri, A.; Yakel, J.L. Desensitization of nicotinic ACh receptors: Shaping cholinergic signaling. Trends Neurosci. 2005, 28, 371–378. [Google Scholar] [CrossRef]
  6. Picciotto, M.R.; Addy, N.A.; Mineur, Y.S.; Brunzell, D.H. It is not “either/or”: Activation and desensitization of nicotinic acetylcholine receptors both contribute to behaviors related to nicotine addiction and mood. Prog. Neurobiol. 2008, 84, 329–342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Wang, J.; Lindstrom, J. Orthosteric and allosteric potentiation of heteromeric neuronal nicotinic acetylcholine receptors. Br. J. Pharmacol. 2018, 175, 1805–1821. [Google Scholar] [CrossRef] [Green Version]
  8. Zoli, M.; Pucci, S.; Vilella, A.; Gotti, C. Neuronal and Extraneuronal Nicotinic Acetylcholine Receptors. Curr. Neuropharmacol. 2018, 16, 338–349. [Google Scholar] [CrossRef] [PubMed]
  9. Veena, J.; Rao, B.S.S.; Srikumar, B.N. Regulation of adult neurogenesis in the hippocampus by stress, acetylcholine and dopamine. J. Nat. Sci. Biol. Med. 2011, 2, 26–37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Leonard, S.; Adams, C.; Breese, C.R.; Adler, L.E.; Bickford, P.; Byerley, W.; Coon, H.; Griffith, J.M.; Miller, C.; Myles-Worsley, M.; et al. Nicotinic receptor function in schizophrenia. Schizophr. Bull. 1996, 22, 431–445. [Google Scholar] [CrossRef]
  11. Dineley, K.T.; Pandya, A.A.; Yakel, J.L. Nicotinic ACh receptors as therapeutic targets in CNS disorders. Trends Pharmacol. Sci. 2015, 36, 96–108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Dani, J.A.; Bertrand, D. Nicotinic acetylcholine receptors and nicotinic cholinergic mechanisms of the central nervous system. Annu. Rev. Pharmacol. Toxicol. 2007, 47, 699–729. [Google Scholar] [CrossRef] [PubMed]
  13. Taly, A.; Corringer, P.-J.; Guedin, D.; Lestage, P.; Changeux, J.-P. Nicotinic receptors: Allosteric transitions and therapeutic targets in the nervous system. Nat. Rev. Drug Discov. 2009, 8, 733–750. [Google Scholar] [CrossRef] [PubMed]
  14. McQuiston, A.R. Acetylcholine release and inhibitory interneuron activity in hippocampal CA1. Front. Synaptic Neurosci. 2014, 6, 20. [Google Scholar] [CrossRef]
  15. Oda, A.; Yamagata, K.; Nakagomi, S.; Uejima, H.; Wiriyasermkul, P.; Ohgaki, R.; Nagamori, S.; Kanai, Y.; Tanaka, H. Nicotine induces dendritic spine remodeling in cultured hippocampal neurons. J. Neurochem. 2014, 128, 246–255. [Google Scholar] [CrossRef] [PubMed]
  16. Gu, Z.; Yakel, J.L. Timing-Dependent Septal Cholinergic Induction of Dynamic Hippocampal Synaptic Plasticity. Neuron 2011, 71, 155–165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Pidoplichko, V.I.; Prager, E.M.; Aroniadou-Anderjaska, V.; Braga, M.F.M. α7-Containing nicotinic acetylcholine receptors on interneurons of the basolateral amygdala and their role in the regulation of the network excitability. J. Neurophysiol. 2013, 110, 2358–2369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Udakis, M.; Wright, V.L.; Wonnacott, S.; Bailey, C.P. Integration of inhibitory and excitatory effects of α7 nicotinic acetylcholine receptor activation in the prelimbic cortex regulates network activity and plasticity. Neuropharmacology 2016, 105, 618–629. [Google Scholar] [CrossRef] [Green Version]
  19. Cohen, S.; Greenberg, M.E. Communication between the synapse and the nucleus in neuronal development, plasticity, and disease. Annu. Rev. Cell Dev. Biol. 2008, 24, 183–209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Bouzat, C.; Lasala, M.; Nielsen, B.E.; Corradi, J.; Esandi, M.D.C. Molecular function of α7 nicotinic receptors as drug targets. J. Physiol. 2018, 596, 1847–1861. [Google Scholar] [CrossRef] [Green Version]
  21. Wessler, I.; Kirkpatrick, C.J. Acetylcholine beyond neurons: The non-neuronal cholinergic system in humans. Br. J. Pharmacol. 2008, 154, 1558–1571. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Zanetti, S.R.; Ziblat, A.; Torres, N.I.; Zwirner, N.W.; Bouzat, C. Expression and Functional Role of α7 Nicotinic Receptor in Human Cytokine-stimulated Natural Killer (NK) Cells. J. Biol. Chem. 2016, 291, 16541–16552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. D’Andrea, M.R.; Nagele, R.G. Targeting the alpha 7 nicotinic acetylcholine receptor to reduce amyloid accumulation in Alzheimer’s disease pyramidal neurons. Curr. Pharm. Des. 2006, 12, 677–684. [Google Scholar] [CrossRef]
  24. Freedman, R.; Hall, M.; Adler, L.E.; Leonard, S. Evidence in postmortem brain tissue for decreased numbers of hippocampal nicotinic receptors in schizophrenia. Biol. Psychiatry 1995, 38, 22–33. [Google Scholar] [CrossRef]
  25. Quik, M.; Zhang, D.; McGregor, M.; Bordia, T. Alpha7 nicotinic receptors as therapeutic targets for Parkinson’s disease. Biochem. Pharmacol. 2015, 97, 399–407. [Google Scholar] [CrossRef] [Green Version]
  26. Steinlein, O.K.; Magnusson, A.; Stoodt, J.; Bertrand, S.; Weiland, S.; Berkovic, S.F.; Nakken, K.O.; Propping, P.; Bertrand, D. An insertion mutation of the CHRNA4 gene in a family with autosomal dominant nocturnal frontal lobe epilepsy. Hum. Mol. Genet. 1997, 6, 943–947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Vallés, A.S.; Barrantes, F.J. Dysregulation of Neuronal Nicotinic Acetylcholine Receptor–Cholesterol Crosstalk in Autism Spectrum Disorder. Front. Mol. Neurosci. 2021, 14, 232. [Google Scholar] [CrossRef]
  28. Wilens, T.E.; Decker, M.W. Neuronal nicotinic receptor agonists for the treatment of attention-deficit/hyperactivity disorder: Focus on cognition. Biochem. Pharmacol. 2007, 74, 1212–1223. [Google Scholar] [CrossRef] [Green Version]
  29. Philip, N.S.; Carpenter, L.L.; Tyrka, A.R.; Price, L.H. Nicotinic acetylcholine receptors and depression: A review of the preclinical and clinical literature. Psychopharmacology 2010, 212, 1–12. [Google Scholar] [CrossRef] [Green Version]
  30. Sciamanna, M.A.; Griesmann, G.E.; Williams, C.L.; Lennon, V.A. Nicotinic acetylcholine receptors of muscle and neuronal (alpha7) types coexpressed in a small cell lung carcinoma. J. Neurochem. 1997, 69, 2302–2311. [Google Scholar] [CrossRef] [PubMed]
  31. Hone, A.J.; McIntosh, J.M. Nicotinic acetylcholine receptors in neuropathic and inflammatory pain. FEBS Lett. 2018, 592, 1045–1062. [Google Scholar] [CrossRef] [PubMed]
  32. Hone, A.J.; Servent, D.; McIntosh, J.M. α9-containing nicotinic acetylcholine receptors and the modulation of pain. Br. J. Pharmacol. 2018, 175, 1915–1927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Wieskopf, J.S.; Mathur, J.; Limapichat, W.; Post, M.R.; Al-Qazzaz, M.; Sorge, R.E.; Martin, L.J.; Zaykin, D.V.; Smith, S.B.; Freitas, K.; et al. The nicotinic α6 subunit gene determines variability in chronic pain sensitivity via cross-inhibition of P2X2/3 receptors. Sci. Transl. Med. 2015, 7, 287ra72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Conti-Fine, B.M.; Milani, M.; Kaminski, H.J. Myasthenia gravis: Past, present, and future. J. Clin. Investig. 2006, 116, 2843–2854. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Engel, A.G.; Shen, X.-M.; Selcen, D.; Sine, S.M. Congenital myasthenic syndromes: Pathogenesis, diagnosis, and treatment. Lancet. Neurol. 2015, 14, 420–434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Lombardo, S.; Maskos, U. Role of the nicotinic acetylcholine receptor in Alzheimer’s disease pathology and treatment. Neuropharmacology 2015, 96, 255–262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Paz, M.L.; Barrantes, F.J. Autoimmune Attack of the Neuromuscular Junction in Myasthenia Gravis: Nicotinic Acetylcholine Receptors and Other Targets. ACS Chem. Neurosci. 2019, 10, 2186–2194. [Google Scholar] [CrossRef]
  38. Srinivasan, R.; Henderson, B.J.; Lester, H.A.; Richards, C.I. Pharmacological chaperoning of nAChRs: A therapeutic target for Parkinson’s disease. Pharmacol. Res. 2014, 83, 20–29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Perez-Lloret, S.; Barrantes, F.J. Deficits in cholinergic neurotransmission and their clinical correlates in Parkinson’s disease. NPJ Park. Dis. 2016, 2, 16001. [Google Scholar] [CrossRef] [Green Version]
  40. American Psychiatric Association. Diagnostic and Statistical Manual of Mental Disorders; American Psychiatric Association: Washington, DC, USA, 2013; ISBN 0-89042-555-8. [Google Scholar]
  41. Koob, G.F.; Volkow, N.D. Neurobiology of addiction: A neurocircuitry analysis. Lancet Psychiatry 2016, 3, 760–773. [Google Scholar] [CrossRef] [PubMed]
  42. Wittenberg, R.E.; Wolfman, S.L.; De Biasi, M.; Dani, J.A. Nicotinic acetylcholine receptors and nicotine addiction: A brief introduction. Neuropharmacology 2020, 177, 108256. [Google Scholar] [CrossRef] [PubMed]
  43. De Biasi, M.; Dani, J.A. Reward, addiction, withdrawal to nicotine. Annu. Rev. Neurosci. 2011, 34, 105–130. [Google Scholar] [CrossRef] [Green Version]
  44. Dani, J.A.; Radcliffe, K.A.; Pidoplichko, V.I. Variations in desensitization of nicotinic acetylcholine receptors from hippocampus and midbrain dopamine areas. Eur. J. Pharmacol. 2000, 393, 31–38. [Google Scholar] [CrossRef]
  45. Pidoplichko, V.I.; DeBiasi, M.; Williams, J.T.; Dani, J.A. Nicotine activates and desensitizes midbrain dopamine neurons. Nature 1997, 390, 401–404. [Google Scholar] [CrossRef] [PubMed]
  46. Wooltorton, J.R.A.; Pidoplichko, V.I.; Broide, R.S.; Dani, J.A. Differential desensitization and distribution of nicotinic acetylcholine receptor subtypes in midbrain dopamine areas. J. Neurosci. 2003, 23, 3176–3185. [Google Scholar] [CrossRef] [PubMed]
  47. Mansvelder, H.D.; McGehee, D.S. Long-term potentiation of excitatory inputs to brain reward areas by nicotine. Neuron 2000, 27, 349–357. [Google Scholar] [CrossRef] [Green Version]
  48. Mao, D.; Gallagher, K.; McGehee, D.S. Nicotine potentiation of excitatory inputs to ventral tegmental area dopamine neurons. J. Neurosci. 2011, 31, 6710–6720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Ostroumov, A.; Dani, J.A. Convergent Neuronal Plasticity and Metaplasticity Mechanisms of Stress, Nicotine, and Alcohol. Annu. Rev. Pharmacol. Toxicol. 2018, 58, 547–566. [Google Scholar] [CrossRef] [PubMed]
  50. Saal, D.; Dong, Y.; Bonci, A.; Malenka, R.C. Drugs of abuse and stress trigger a common synaptic adaptation in dopamine neurons. Neuron 2003, 37, 577–582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Baker, L.K.; Mao, D.; Chi, H.; Govind, A.P.; Vallejo, Y.F.; Iacoviello, M.; Herrera, S.; Cortright, J.J.; Green, W.N.; McGehee, D.S.; et al. Intermittent nicotine exposure upregulates nAChRs in VTA dopamine neurons and sensitises locomotor responding to the drug. Eur. J. Neurosci. 2013, 37, 1004–1011. [Google Scholar] [CrossRef] [PubMed]
  52. Nashmi, R.; Xiao, C.; Deshpande, P.; McKinney, S.; Grady, S.R.; Whiteaker, P.; Huang, Q.; McClure-Begley, T.; Lindstrom, J.M.; Labarca, C.; et al. Chronic Nicotine Cell Specifically Upregulates Functional α4* Nicotinic Receptors: Basis for Both Tolerance in Midbrain and Enhanced Long-Term Potentiation in Perforant Path. J. Neurosci. 2007, 27, 8202–8218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Feduccia, A.; Chatterjee, S.; Bartlett, S. Neuronal nicotinic acetylcholine receptors: Neuroplastic changes underlying alcohol and nicotine addictions. Front. Mol. Neurosci. 2012, 5, 83. [Google Scholar] [CrossRef] [Green Version]
  54. Fenster, C.P.; Hicks, J.H.; Beckman, M.L.; Covernton, P.J.; Quick, M.W.; Lester, R.A. Desensitization of nicotinic receptors in the central nervous system. Ann. N. Y. Acad. Sci. 1999, 868, 620–623. [Google Scholar] [CrossRef]
  55. Pirhaji, L.; Milani, P.; Dalin, S.; Wassie, B.T.; Dunn, D.E.; Fenster, R.J.; Avila-Pacheco, J.; Greengard, P.; Clish, C.B.; Heiman, M.; et al. Identifying therapeutic targets by combining transcriptional data with ordinal clinical measurements. Nat. Commun. 2017, 8, 623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Henderson, B.J.; Srinivasan, R.; Nichols, W.A.; Dilworth, C.N.; Gutierrez, D.F.; Mackey, E.D.W.; McKinney, S.; Drenan, R.M.; Richards, C.I.; Lester, H.A. Nicotine exploits a COPI-mediated process for chaperone-mediated up-regulation of its receptors. J. Gen. Physiol. 2014, 143, 51–66. [Google Scholar] [CrossRef] [PubMed]
  57. Matta, S.G.; Balfour, D.J.; Benowitz, N.L.; Boyd, R.T.; Buccafusco, J.J.; Caggiula, A.R.; Craig, C.R.; Collins, A.C.; Damaj, M.I.; Donny, E.C.; et al. Guidelines on nicotine dose selection for in vivo research. Psychopharmacology 2007, 190, 269–319. [Google Scholar] [CrossRef] [PubMed]
  58. Ikemoto, S. Brain reward circuitry beyond the mesolimbic dopamine system: A neurobiological theory. Neurosci. Biobehav. Rev. 2010, 35, 129–150. [Google Scholar] [CrossRef] [Green Version]
  59. Lecca, S.; Meye, F.J.; Mameli, M. The lateral habenula in addiction and depression: An anatomical, synaptic and behavioral overview. Eur. J. Neurosci. 2014, 39, 1170–1178. [Google Scholar] [CrossRef]
  60. Meye, F.J.; Trusel, M.; Soiza-Reilly, M.; Mameli, M. Neural circuit adaptations during drug withdrawal—Spotlight on the lateral habenula. Pharmacol. Biochem. Behav. 2017, 162, 87–93. [Google Scholar] [CrossRef] [PubMed]
  61. Winter, C.; Vollmayr, B.; Djodari-Irani, A.; Klein, J.; Sartorius, A. Pharmacological inhibition of the lateral habenula improves depressive-like behavior in an animal model of treatment resistant depression. Behav. Brain Res. 2011, 216, 463–465. [Google Scholar] [CrossRef] [PubMed]
  62. Kenny, P.J.; Markou, A. Neurobiology of the nicotine withdrawal syndrome. Pharmacol. Biochem. Behav. 2001, 70, 531–549. [Google Scholar] [CrossRef] [PubMed]
  63. Picciotto, M.R.; Kenny, P.J. Molecular mechanisms underlying behaviors related to nicotine addiction. Cold Spring Harb. Perspect. Med. 2013, 3, a012112. [Google Scholar] [CrossRef] [PubMed]
  64. Kanasuwan, A.; Deuther-Conrad, W.; Chongruchiroj, S.; Sarasamkan, J.; Chotipanich, C.; Vajragupta, O.; Arunrungvichian, K. Selective α(3)β(4) Nicotinic Acetylcholine Receptor Ligand as a Potential Tracer for Drug Addiction. Int. J. Mol. Sci. 2023, 24, 3614. [Google Scholar] [CrossRef] [PubMed]
  65. Grady, S.R.; Moretti, M.; Zoli, M.; Marks, M.J.; Zanardi, A.; Pucci, L.; Clementi, F.; Gotti, C. Rodent habenulo-interpeduncular pathway expresses a large variety of uncommon nAChR subtypes, but only the alpha3beta4* and alpha3beta3beta4* subtypes mediate acetylcholine release. J. Neurosci. 2009, 29, 2272–2282. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Fowler, C.D.; Lu, Q.; Johnson, P.M.; Marks, M.J.; Kenny, P.J. Habenular α5 nicotinic receptor subunit signalling controls nicotine intake. Nature 2011, 471, 597–601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. McLaughlin, I.; Dani, J.A.; De Biasi, M. Nicotine withdrawal. Curr. Top. Behav. Neurosci. 2015, 24, 99–123. [Google Scholar] [CrossRef] [Green Version]
  68. McLaughlin, I.; Dani, J.A.; De Biasi, M. The medial habenula and interpeduncular nucleus circuitry is critical in addiction, anxiety, and mood regulation. J. Neurochem. 2017, 142, 130–143. [Google Scholar] [CrossRef] [Green Version]
  69. Zaveri, N.; Jiang, F.; Olsen, C.; Polgar, W.; Toll, L. Novel α3β4 nicotinic acetylcholine receptor-selective ligands. Discovery, structure-activity studies, and pharmacological evaluation. J. Med. Chem. 2010, 53, 8187–8191. [Google Scholar] [CrossRef] [Green Version]
  70. Wen, L.; Yang, Z.; Cui, W.; Li, M.D. Crucial roles of the CHRNB3-CHRNA6 gene cluster on chromosome 8 in nicotine dependence: Update and subjects for future research. Transl. Psychiatry 2016, 6, e843. [Google Scholar] [CrossRef] [Green Version]
  71. Cannon, D.S.; Mermelstein, R.J.; Hedeker, D.; Coon, H.; Cook, E.H.; McMahon, W.M.; Hamil, C.; Dunn, D.; Weiss, R.B. Effect of neuronal nicotinic acetylcholine receptor genes (CHRN) on longitudinal cigarettes per day in adolescents and young adults. Nicotine Tob. Res. 2014, 16, 137–144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Counotte, D.S.; Smit, A.B.; Pattij, T.; Spijker, S. Development of the motivational system during adolescence, and its sensitivity to disruption by nicotine. Dev. Cogn. Neurosci. 2011, 1, 430–443. [Google Scholar] [CrossRef] [Green Version]
  73. Day, M.; Pan, J.B.; Buckley, M.J.; Cronin, E.; Hollingsworth, P.R.; Hirst, W.D.; Navarra, R.; Sullivan, J.P.; Decker, M.W.; Fox, G.B. Differential effects of ciproxifan and nicotine on impulsivity and attention measures in the 5-choice serial reaction time test. Biochem. Pharmacol. 2007, 73, 1123–1134. [Google Scholar] [CrossRef] [PubMed]
  74. Grottick, A.J.; Higgins, G.A. Effect of subtype selective nicotinic compounds on attention as assessed by the five-choice serial reaction time task. Behav. Brain Res. 2000, 117, 197–208. [Google Scholar] [CrossRef] [PubMed]
  75. Rycroft, N.; Rusted, J.M.; Hutton, S.B. Acute effects of nicotine on visual search tasks in young adult smokers. Psychopharmacology 2005, 181, 160–169. [Google Scholar] [CrossRef] [PubMed]
  76. Van Gaalen, M.M.; Brueggeman, R.J.; Bronius, P.F.C.; Schoffelmeer, A.N.M.; Vanderschuren, L.J.M.J. Behavioral disinhibition requires dopamine receptor activation. Psychopharmacology 2006, 187, 73–85. [Google Scholar] [CrossRef] [PubMed]
  77. Boksa, P. Smoking, psychiatric illness and the brain. J. Psychiatry Neurosci. 2017, 42, 147–149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Kalman, D.; Morissette, S.B.; George, T.P. Co-morbidity of smoking in patients with psychiatric and substance use disorders. Am. J. Addict. 2005, 14, 106–123. [Google Scholar] [CrossRef] [Green Version]
  79. Breese, C.R.; Lee, M.J.; Adams, C.E.; Sullivan, B.; Logel, J.; Gillen, K.M.; Marks, M.J.; Collins, A.C.; Leonard, S. Abnormal regulation of high affinity nicotinic receptors in subjects with schizophrenia. Neuropsychopharmacology 2000, 23, 351–364. [Google Scholar] [CrossRef] [Green Version]
  80. Durany, N.; Zöchling, R.; Boissl, K.W.; Paulus, W.; Ransmayr, G.; Tatschner, T.; Danielczyk, W.; Jellinger, K.; Deckert, J.; Riederer, P. Human post-mortem striatal alpha4beta2 nicotinic acetylcholine receptor density in schizophrenia and Parkinson’s syndrome. Neurosci. Lett. 2000, 287, 109–112. [Google Scholar] [CrossRef] [PubMed]
  81. D’Souza, D.C.; Esterlis, I.; Carbuto, M.; Krasenics, M.; Seibyl, J.; Bois, F.; Pittman, B.; Ranganathan, M.; Cosgrove, K.; Staley, J. Lower ß2*-nicotinic acetylcholine receptor availability in smokers with schizophrenia. Am. J. Psychiatry 2012, 169, 326–334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Nguyen, H.N.; Rasmussen, B.A.; Perry, D.C. Binding and functional activity of nicotinic cholinergic receptors in selected rat brain regions are increased following long-term but not short-term nicotine treatment. J. Neurochem. 2004, 90, 40–49. [Google Scholar] [CrossRef]
  83. Rabin, R.A.; George, T.P. A review of co-morbid tobacco and cannabis use disorders: Possible mechanisms to explain high rates of co-use. Am. J. Addict. 2015, 24, 105–116. [Google Scholar] [CrossRef] [PubMed]
  84. Amos, A.; Wiltshire, S.; Bostock, Y.; Haw, S.; McNeill, A. ’You can’t go without a fag...you need it for your hash’—A qualitative exploration of smoking, cannabis and young people. Addiction 2004, 99, 77–81. [Google Scholar] [CrossRef] [PubMed]
  85. Ream, G.L.; Benoit, E.; Johnson, B.D.; Dunlap, E. Smoking tobacco along with marijuana increases symptoms of cannabis dependence. Drug Alcohol Depend. 2008, 95, 199–208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Vallés, A.S.; Barrantes, F.J. Nanoscale Sub-Compartmentalization of the Dendritic Spine Compartment. Biomolecules 2021, 11, 1697. [Google Scholar] [CrossRef] [PubMed]
  87. Buzzi, B.; Koseli, E.; Moncayo, L.; Shoaib, M.; Damaj, M.I. Role of neuronal nicotinic acetylcholine receptors in cannabinoid dependence. Pharmacol. Res. 2023, 191, 106746. [Google Scholar] [CrossRef] [PubMed]
  88. Ramo, D.E.; Liu, H.; Prochaska, J.J. Tobacco and marijuana use among adolescents and young adults: A systematic review of their co-use. Clin. Psychol. Rev. 2012, 32, 105–121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Valjent, E.; Mitchell, J.M.; Besson, M.-J.; Caboche, J.; Maldonado, R. Behavioural and biochemical evidence for interactions between Delta 9-tetrahydrocannabinol and nicotine. Br. J. Pharmacol. 2002, 135, 564–578. [Google Scholar] [CrossRef] [Green Version]
  90. Lubman, D.I.; Cheetham, A.; Yücel, M. Cannabis and adolescent brain development. Pharmacol. Ther. 2015, 148, 1–16. [Google Scholar] [CrossRef]
  91. Manwell, L.A.; Miladinovic, T.; Raaphorst, E.; Rana, S.; Malecki, S.; Mallet, P.E. Chronic nicotine exposure attenuates the effects of Δ(9) -tetrahydrocannabinol on anxiety-related behavior and social interaction in adult male and female rats. Brain Behav. 2019, 9, e01375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Dellazizzo, L.; Potvin, S.; Giguère, S.; Dumais, A. Evidence on the acute and residual neurocognitive effects of cannabis use in adolescents and adults: A systematic meta-review of meta-analyses. Addiction 2022, 117, 1857–1870. [Google Scholar] [CrossRef] [PubMed]
  93. Smith, A.D.; Dar, M.S. Mouse cerebellar nicotinic-cholinergic receptor modulation of Delta9-THC ataxia: Role of the alpha4beta2 subtype. Brain Res. 2006, 1115, 16–25. [Google Scholar] [CrossRef] [PubMed]
  94. Justinova, Z.; Mascia, P.; Wu, H.-Q.; Secci, M.E.; Redhi, G.H.; Panlilio, L.V.; Scherma, M.; Barnes, C.; Parashos, A.; Zara, T.; et al. Reducing cannabinoid abuse and preventing relapse by enhancing endogenous brain levels of kynurenic acid. Nat. Neurosci. 2013, 16, 1652–1661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Liu, M.; Jiang, Y.; Wedow, R.; Li, Y.; Brazel, D.M.; Chen, F.; Datta, G.; Davila-Velderrain, J.; McGuire, D.; Tian, C.; et al. Association studies of up to 1.2 million individuals yield new insights into the genetic etiology of tobacco and alcohol use. Nat. Genet. 2019, 51, 237–244. [Google Scholar] [CrossRef]
  96. Gatta, V.; Mengod, G.; Reale, M.; Tata, A.M. Possible Correlation between Cholinergic System Alterations and Neuro/Inflammation in Multiple Sclerosis. Biomedicines 2020, 8, 153. [Google Scholar] [CrossRef] [PubMed]
  97. Piovesana, R.; Salazar Intriago, M.S.; Dini, L.; Tata, A.M. Cholinergic Modulation of Neuroinflammation: Focus on α7 Nicotinic Receptor. Int. J. Mol. Sci. 2021, 22, 4912. [Google Scholar] [CrossRef] [PubMed]
  98. Shytle, R.D.; Mori, T.; Townsend, K.; Vendrame, M.; Sun, N.; Zeng, J.; Ehrhart, J.; Silver, A.A.; Sanberg, P.R.; Tan, J. Cholinergic modulation of microglial activation by alpha 7 nicotinic receptors. J. Neurochem. 2004, 89, 337–343. [Google Scholar] [CrossRef]
  99. Varatharaj, A.; Galea, I. The blood-brain barrier in systemic inflammation. Brain. Behav. Immun. 2017, 60, 1–12. [Google Scholar] [CrossRef] [Green Version]
  100. Wang, H.; Yu, M.; Ochani, M.; Amelia, C.A.; Tanovic, M.; Susarla, S.; Li, J.H.; Wang, H.; Yang, N.; Ulloa, L.; et al. Nicotinic acetylcholine receptor α7 subunit is an essential regulator of inflammation. Nature 2003, 421, 384–388. [Google Scholar] [CrossRef]
  101. Lu, X.-X.; Hong, Z.-Q.; Tan, Z.; Sui, M.-H.; Zhuang, Z.-Q.; Liu, H.-H.; Zheng, X.-Y.; Yan, T.-B.; Geng, D.-F.; Jin, D.-M. Nicotinic Acetylcholine Receptor Alpha7 Subunit Mediates Vagus Nerve Stimulation-Induced Neuroprotection in Acute Permanent Cerebral Ischemia by a7nAchR/JAK2 Pathway. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2017, 23, 6072–6081. [Google Scholar] [CrossRef] [Green Version]
  102. Ito, T.; Inden, M.; Ueda, T.; Asaka, Y.; Kurita, H.; Hozumi, I. The neuroprotective effects of activated α7 nicotinic acetylcholine receptor against mutant copper-zinc superoxide dismutase 1-mediated toxicity. Sci. Rep. 2020, 10, 22157. [Google Scholar] [CrossRef]
  103. De Jonge, W.J.; Ulloa, L. The alpha7 nicotinic acetylcholine receptor as a pharmacological target for inflammation. Br. J. Pharmacol. 2007, 151, 915–929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Patel, H.; McIntire, J.; Ryan, S.; Dunah, A.; Loring, R. Anti-inflammatory effects of astroglial α7 nicotinic acetylcholine receptors are mediated by inhibition of the NF-κB pathway and activation of the Nrf2 pathway. J. Neuroinflamm. 2017, 14, 192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Kihara, T.; Sawada, H.; Nakamizo, T.; Kanki, R.; Yamashita, H.; Maelicke, A.; Shimohama, S. Galantamine modulates nicotinic receptor and blocks Aβ-enhanced glutamate toxicity. Biochem. Biophys. Res. Commun. 2004, 325, 976–982. [Google Scholar] [CrossRef]
  106. Liu, Y.; Hu, J.; Wu, J.; Zhu, C.; Hui, Y.; Han, Y.; Huang, Z.; Ellsworth, K.; Fan, W. α7 nicotinic acetylcholine receptor-mediated neuroprotection against dopaminergic neuron loss in an MPTP mouse model via inhibition of astrocyte activation. J. Neuroinflamm. 2012, 9, 98. [Google Scholar] [CrossRef] [Green Version]
  107. Lee, Y.B.; Schrader, J.W.; Kim, S.U. p38 map kinase regulates TNF-alpha production in human astrocytes and microglia by multiple mechanisms. Cytokine 2000, 12, 874–880. [Google Scholar] [CrossRef] [PubMed]
  108. Youssef, M.; Ibrahim, A.; Akashi, K.; Hossain, M.S. PUFA-Plasmalogens Attenuate the LPS-Induced Nitric Oxide Production by Inhibiting the NF-kB, p38 MAPK and JNK Pathways in Microglial Cells. Neuroscience 2019, 397, 18–30. [Google Scholar] [CrossRef]
  109. Hone, A.J.; McIntosh, J.M. Nicotinic acetylcholine receptors: Therapeutic targets for novel ligands to treat pain and inflammation. Pharmacol. Res. 2023, 190, 106715. [Google Scholar] [CrossRef]
  110. Costantini, E.; Carrarini, C.; Borrelli, P.; De Rosa, M.; Calisi, D.; Consoli, S.; D’Ardes, D.; Cipollone, F.; Di Nicola, M.; Onofrj, M.; et al. Different peripheral expression patterns of the nicotinic acetylcholine receptor in dementia with Lewy bodies and Alzheimer’s disease. Immun. Ageing 2023, 20, 3. [Google Scholar] [CrossRef]
  111. Chernyavsky, A.I.; Arredondo, J.; Skok, M.; Grando, S.A. Auto/paracrine control of inflammatory cytokines by acetylcholine in macrophage-like U937 cells through nicotinic receptors. Int. Immunopharmacol. 2010, 10, 308–315. [Google Scholar] [CrossRef] [Green Version]
  112. Maldifassi, M.C.; Atienza, G.; Arnalich, F.; López-Collazo, E.; Cedillo, J.L.; Martín-Sánchez, C.; Bordas, A.; Renart, J.; Montiel, C. A new IRAK-M-mediated mechanism implicated in the anti-inflammatory effect of nicotine via α7 nicotinic receptors in human macrophages. PLoS ONE 2014, 9, e108397. [Google Scholar] [CrossRef] [Green Version]
  113. Alen, N. V The cholinergic anti-inflammatory pathway in humans: State-of-the-art review and future directions. Neurosci. Biobehav. Rev. 2022, 136, 104622. [Google Scholar] [CrossRef]
  114. Pavlov, V.A.; Wang, H.; Czura, C.J.; Friedman, S.G.; Tracey, K.J. The cholinergic anti-inflammatory pathway: A missing link in neuroimmunomodulation. Mol. Med. 2003, 9, 125–134. [Google Scholar] [CrossRef] [Green Version]
  115. Vijay, K. Toll-like receptors in immunity and inflammatory diseases: Past, present, and future. Int. Immunopharmacol. 2018, 59, 391–412. [Google Scholar] [CrossRef]
  116. Richter, K.; Papke, R.L.; Stokes, C.; Roy, D.C.; Espinosa, E.S.; Wolf, P.M.K.; Hecker, A.; Liese, J.; Singh, V.K.; Padberg, W.; et al. Comparison of the Anti-inflammatory Properties of Two Nicotinic Acetylcholine Receptor Ligands, Phosphocholine and pCF3-diEPP. Front. Cell. Neurosci. 2022, 16, 779081. [Google Scholar] [CrossRef] [PubMed]
  117. Richter, K.; Mathes, V.; Fronius, M.; Althaus, M.; Hecker, A.; Krasteva-Christ, G.; Padberg, W.; Hone, A.J.; McIntosh, J.M.; Zakrzewicz, A.; et al. Phosphocholine – an agonist of metabotropic but not of ionotropic functions of α9-containing nicotinic acetylcholine receptors. Sci. Rep. 2016, 6, 28660. [Google Scholar] [CrossRef] [Green Version]
  118. Hecker, A.; Küllmar, M.; Wilker, S.; Richter, K.; Zakrzewicz, A.; Atanasova, S.; Mathes, V.; Timm, T.; Lerner, S.; Klein, J.; et al. Phosphocholine-Modified Macromolecules and Canonical Nicotinic Agonists Inhibit ATP-Induced IL-1β Release. J. Immunol. 2015, 195, 2325–2334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Courties, A.; Boussier, J.; Hadjadj, J.; Yatim, N.; Barnabei, L.; Péré, H.; Veyer, D.; Kernéis, S.; Carlier, N.; Pène, F.; et al. Regulation of the acetylcholine/α7nAChR anti-inflammatory pathway in COVID-19 patients. Sci. Rep. 2021, 11, 11886. [Google Scholar] [CrossRef]
  120. Simon, T.; Kirk, J.; Dolezalova, N.; Guyot, M.; Panzolini, C.; Bondue, A.; Lavergne, J.; Hugues, S.; Hypolite, N.; Saeb-Parsy, K.; et al. The cholinergic anti-inflammatory pathway inhibits inflammation without lymphocyte relay. Front. Neurosci. 2023, 17, 1125492. [Google Scholar] [CrossRef]
  121. Vonck, K.; Raedt, R.; Naulaerts, J.; De Vogelaere, F.; Thiery, E.; Van Roost, D.; Aldenkamp, B.; Miatton, M.; Boon, P. Vagus nerve stimulation…25 years later! What do we know about the effects on cognition? Neurosci. Biobehav. Rev. 2014, 45, 63–71. [Google Scholar] [CrossRef]
  122. Vargas-Caballero, M.; Warming, H.; Walker, R.; Holmes, C.; Cruickshank, G.; Patel, B. Vagus Nerve Stimulation as a Potential Therapy in Early Alzheimer’s Disease: A Review. Front. Hum. Neurosci. 2022, 16, 866434. [Google Scholar] [CrossRef]
  123. Sigurdsson, H.P.; Raw, R.; Hunter, H.; Baker, M.R.; Taylor, J.-P.; Rochester, L.; Yarnall, A.J. Noninvasive vagus nerve stimulation in Parkinson’s disease: Current status and future prospects. Expert Rev. Med. Devices 2021, 18, 971–984. [Google Scholar] [CrossRef]
  124. Smucny, J.; Visani, A.; Tregellas, J.R. Could vagus nerve stimulation target hippocampal hyperactivity to improve cognition in schizophrenia? Front. Psychiatry 2015, 6, 43. [Google Scholar] [CrossRef] [Green Version]
  125. Vallés, A.S.; Borroni, M.V.; Barrantes, F.J. Targeting brain α7 nicotinic acetylcholine receptors in alzheimer’s disease: Rationale and current status. CNS Drugs 2014, 28, 975–987. [Google Scholar] [CrossRef]
  126. Price, J.L.; Davis, P.B.; Morris, J.C.; White, D.L. The distribution of tangles, plaques and related immunohistochemical markers in healthy aging and Alzheimer’s disease. Neurobiol. Aging 1991, 12, 295–312. [Google Scholar] [CrossRef]
  127. Cummings, J.L. Alzheimer’s disease. N. Engl. J. Med. 2004, 351, 56–67. [Google Scholar] [CrossRef]
  128. Hoskin, J.L.; Al-Hasan, Y.; Sabbagh, M.N. Nicotinic Acetylcholine Receptor Agonists for the Treatment of Alzheimer’s Dementia: An Update. Nicotine Tob. Res. 2019, 21, 370–376. [Google Scholar] [CrossRef]
  129. Wu, J.; Ishikawa, M.; Zhang, J.; Hashimoto, K. Brain imaging of nicotinic receptors in Alzheimer’s disease. Int. J. Alzheimers. Dis. 2010, 2010, 548913. [Google Scholar] [CrossRef] [Green Version]
  130. Ma, K.-G.; Qian, Y.-H. Alpha 7 nicotinic acetylcholine receptor and its effects on Alzheimer’s disease. Neuropeptides 2019, 73, 96–106. [Google Scholar] [CrossRef] [PubMed]
  131. Schliebs, R.; Arendt, T. The cholinergic system in aging and neuronal degeneration. Behav. Brain Res. 2011, 221, 555–563. [Google Scholar] [CrossRef]
  132. Oddo, S.; Caccamo, A.; Green, K.N.; Liang, K.; Tran, L.; Chen, Y.; Leslie, F.M.; LaFerla, F.M. Chronic nicotine administration exacerbates tau pathology in a transgenic model of Alzheimer’s disease. Proc. Natl. Acad. Sci. USA 2005, 102, 3046–3051. [Google Scholar] [CrossRef] [PubMed]
  133. Barrantes, F.J.; Borroni, V.; Vallés, S. Neuronal nicotinic acetylcholine receptor-cholesterol crosstalk in Alzheimer’s disease. FEBS Lett. 2010, 584, 1856–1863. [Google Scholar] [CrossRef] [Green Version]
  134. Zhang, L.; Wang, Y.; Liu, T.; Mao, Y.; Peng, B. Novel Microglia-based Therapeutic Approaches to Neurodegenerative Disorders. Neurosci. Bull. 2023, 39, 491–502. [Google Scholar] [CrossRef]
  135. Xie, J.; Van Hoecke, L.; Vandenbroucke, R.E. The Impact of Systemic Inflammation on Alzheimer’s Disease Pathology. Front. Immunol. 2021, 12, 796867. [Google Scholar] [CrossRef]
  136. Holmes, C.; Cunningham, C.; Zotova, E.; Woolford, J.; Dean, C.; Kerr, S.; Culliford, D.; Perry, V.H. Systemic inflammation and disease progression in Alzheimer disease. Neurology 2009, 73, 768–774. [Google Scholar] [CrossRef] [Green Version]
  137. Leung, R.; Proitsi, P.; Simmons, A.; Lunnon, K.; Güntert, A.; Kronenberg, D.; Pritchard, M.; Tsolaki, M.; Mecocci, P.; Kloszewska, I.; et al. Inflammatory proteins in plasma are associated with severity of Alzheimer’s disease. PLoS ONE 2013, 8, e64971. [Google Scholar] [CrossRef] [Green Version]
  138. Motta, M.; Imbesi, R.; Di Rosa, M.; Stivala, F.; Malaguarnera, L. Altered plasma cytokine levels in Alzheimer’s disease: Correlation with the disease progression. Immunol. Lett. 2007, 114, 46–51. [Google Scholar] [CrossRef] [PubMed]
  139. Wang, Y.; Zhu, N.; Wang, K.; Zhang, Z.; Wang, Y. Identification of α7 nicotinic acetylcholine receptor on hippocampal astrocytes cultured in vitro and its role on inflammatory mediator secretion. Neural Regen. Res. 2012, 7, 1709–1714. [Google Scholar] [CrossRef]
  140. Ballinger, E.C.; Ananth, M.; Talmage, D.A.; Role, L.W. Basal Forebrain Cholinergic Circuits and Signaling in Cognition and Cognitive Decline. Neuron 2016, 91, 1199–1218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Oz, M.; Lorke, D.E.; Yang, K.-H.S.; Petroianu, G. On the interaction of β-amyloid peptides and α7-nicotinic acetylcholine receptors in Alzheimer’s disease. Curr. Alzheimer Res. 2013, 10, 618–630. [Google Scholar] [CrossRef] [PubMed]
  142. Di Lascio, S.; Fornasari, D.; Benfante, R. The Human-Restricted Isoform of the α7 nAChR, CHRFAM7A: A Double-Edged Sword in Neurological and Inflammatory Disorders. Int. J. Mol. Sci. 2022, 23, 3463. [Google Scholar] [CrossRef]
  143. Burns, L.H.; Pei, Z.; Wang, H.-Y. Targeting α7 nicotinic acetylcholine receptors and their protein interactions in Alzheimer’s disease drug development. Drug Dev. Res. 2023. early view. [Google Scholar] [CrossRef]
  144. Gault, J.; Robinson, M.; Berger, R.; Drebing, C.; Logel, J.; Hopkins, J.; Moore, T.; Jacobs, S.; Meriwether, J.; Choi, M.J.; et al. Genomic Organization and Partial Duplication of the Human α7 Neuronal Nicotinic Acetylcholine Receptor Gene (CHRNA7). Genomics 1998, 52, 173–185. [Google Scholar] [CrossRef]
  145. Riley, B.; Williamson, M.; Collier, D.; Wilkie, H.; Makoff, A. A 3-Mb map of a large Segmental duplication overlapping the alpha7-nicotinic acetylcholine receptor gene (CHRNA7) at human 15q13-q14. Genomics 2002, 79, 197–209. [Google Scholar] [CrossRef]
  146. O’Bleness, M.; Searles, V.B.; Varki, A.; Gagneux, P.; Sikela, J.M. Evolution of genetic and genomic features unique to the human lineage. Nat. Rev. Genet. 2012, 13, 853–866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Flomen, R.H.; Davies, A.F.; Di Forti, M.; La Cascia, C.; Mackie-Ogilvie, C.; Murray, R.; Makoff, A.J. The copy number variant involving part of the alpha7 nicotinic receptor gene contains a polymorphic inversion. Eur. J. Hum. Genet. 2008, 16, 1364–1371. [Google Scholar] [CrossRef] [PubMed]
  148. Szafranski, P.; Schaaf, C.P.; Person, R.E.; Gibson, I.B.; Xia, Z.; Mahadevan, S.; Wiszniewska, J.; Bacino, C.A.; Lalani, S.; Potocki, L.; et al. Structures and molecular mechanisms for common 15q13.3 microduplications involving CHRNA7: Benign or pathological? Hum. Mutat. 2010, 31, 840–850. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Araud, T.; Graw, S.; Berger, R.; Lee, M.; Neveu, E.; Bertrand, D.; Leonard, S. The chimeric gene CHRFAM7A, a partial duplication of the CHRNA7 gene, is a dominant negative regulator of α7*nAChR function. Biochem. Pharmacol. 2011, 82, 904–914. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Ihnatovych, I.; Nayak, T.K.; Ouf, A.; Sule, N.; Birkaya, B.; Chaves, L.; Auerbach, A.; Szigeti, K. iPSC model of CHRFAM7A effect on α7 nicotinic acetylcholine receptor function in the human context. Transl. Psychiatry 2019, 9, 59. [Google Scholar] [CrossRef] [Green Version]
  151. Szigeti, K.; Ihnatovych, I.; Birkaya, B.; Chen, Z.; Ouf, A.; Indurthi, D.C.; Bard, J.E.; Kann, J.; Adams, A.; Chaves, L.; et al. CHRFAM7A: A human specific fusion gene, accounts for the translational gap for cholinergic strategies in Alzheimer’s disease. EBioMedicine 2020, 59, 102892. [Google Scholar] [CrossRef]
  152. Terry, A.V.; Jones, K.; Bertrand, D. Nicotinic acetylcholine receptors in neurological and psychiatric diseases. Pharmacol. Res. 2023, 191, 106764. [Google Scholar] [CrossRef] [PubMed]
  153. Hsia, A.Y.; Masliah, E.; McConlogue, L.; Yu, G.Q.; Tatsuno, G.; Hu, K.; Kholodenko, D.; Malenka, R.C.; Nicoll, R.A.; Mucke, L. Plaque-independent disruption of neural circuits in Alzheimer’s disease mouse models. Proc. Natl. Acad. Sci. USA 1999, 96, 3228–3233. [Google Scholar] [CrossRef] [PubMed]
  154. Orr-Urtreger, A.; Göldner, F.M.; Saeki, M.; Lorenzo, I.; Goldberg, L.; De Biasi, M.; Dani, J.A.; Patrick, J.W.; Beaudet, A.L. Mice deficient in the alpha7 neuronal nicotinic acetylcholine receptor lack alpha-bungarotoxin binding sites and hippocampal fast nicotinic currents. J. Neurosci. 1997, 17, 9165–9171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Dziewczapolski, G.; Glogowski, C.M.; Masliah, E.; Heinemann, S.F. 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. 2009, 29, 8805–8815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Maroli, A.; Di Lascio, S.; Drufuca, L.; Cardani, S.; Setten, E.; Locati, M.; Fornasari, D.; Benfante, R. Effect of donepezil on the expression and responsiveness to LPS of CHRNA7 and CHRFAM7A in macrophages: A possible link to the cholinergic anti-inflammatory pathway. J. Neuroimmunol. 2019, 332, 155–166. [Google Scholar] [CrossRef]
  157. Meshul, C.K.; Kamel, D.; Moore, C.; Kay, T.S.; Krentz, L. Nicotine alters striatal glutamate function and decreases the apomorphine-induced contralateral rotations in 6-OHDA-lesioned rats. Exp. Neurol. 2002, 175, 257–274. [Google Scholar] [CrossRef]
  158. Tizabi, Y.; Getachew, B. Nicotinic Receptor Intervention in Parkinson’s Disease: Future Directions. Clin. Pharmacol. Transl. Med. 2017, 1, 14–19. [Google Scholar] [PubMed]
  159. Tizabi, Y.; Getachew, B.; Csoka, A.B.; Manaye, K.F.; Copeland, R.L. Novel targets for parkinsonism-depression comorbidity. Prog. Mol. Biol. Transl. Sci. 2019, 167, 1–24. [Google Scholar] [CrossRef]
  160. Lieberman, A.; Deep, A.; Olson, M.C.; Smith Hussain, V.; Frames, C.W.; McCauley, M.; Lockhart, T.E. Falls When Standing, Falls When Walking: Different Mechanisms, Different Outcomes in Parkinson Disease. Cureus 2019, 11, e5329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Tizabi, Y.; Getachew, B.; Aschner, M. Novel Pharmacotherapies in Parkinson’s Disease. Neurotox. Res. 2021, 39, 1381–1390. [Google Scholar] [CrossRef]
  162. Stuckenholz, V.; Bacher, M.; Balzer-Geldsetzer, M.; Alvarez-Fischer, D.; Oertel, W.H.; Dodel, R.C.; Noelker, C. The α7 nAChR agonist PNU-282987 reduces inflammation and MPTP-induced nigral dopaminergic cell loss in mice. J. Parkinsons. Dis. 2013, 3, 161–172. [Google Scholar] [CrossRef] [PubMed]
  163. Ryan, R.E.; Ross, S.A.; Drago, J.; Loiacono, R.E. Dose-related neuroprotective effects of chronic nicotine in 6-hydroxydopamine treated rats, and loss of neuroprotection in alpha4 nicotinic receptor subunit knockout mice. Br. J. Pharmacol. 2001, 132, 1650–1656. [Google Scholar] [CrossRef] [Green Version]
  164. Zhang, D.; Mallela, A.; Sohn, D.; Carroll, F.I.; Bencherif, M.; Letchworth, S.; Quik, M. Nicotinic receptor agonists reduce L-DOPA-induced dyskinesias in a monkey model of Parkinson’s disease. J. Pharmacol. Exp. Ther. 2013, 347, 225–234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Johnston, T.H.; Huot, P.; Fox, S.H.; Koprich, J.B.; Szeliga, K.T.; James, J.W.; Graef, J.D.; Letchworth, S.R.; Jordan, K.G.; Hill, M.P.; et al. TC-8831, a nicotinic acetylcholine receptor agonist, reduces L-DOPA-induced dyskinesia in the MPTP macaque. Neuropharmacology 2013, 73, 337–347. [Google Scholar] [CrossRef]
  166. Zhang, D.; Bordia, T.; McGregor, M.; McIntosh, J.M.; Decker, M.W.; Quik, M. ABT-089 and ABT-894 reduce levodopa-induced dyskinesias in a monkey model of Parkinson’s disease. Mov. Disord. 2014, 29, 508–517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Mather, J.; Burdette, C.; Posener, A.; Leventer, Y.; Poole, D.; Fox, R.; Johnston, H. Potential of AZD1446, a novel nicotinic agonist, for the treatment of L-DOPA-induced dyskinesia in Parkinson’s disease. Soc. Neurosci. Abstr 2014, 43, 137. [Google Scholar]
  168. Zhang, D.; McGregor, M.; Decker, M.W.; Quik, M. The α7 nicotinic receptor agonist ABT-107 decreases L-Dopa-induced dyskinesias in parkinsonian monkeys. J. Pharmacol. Exp. Ther. 2014, 351, 25–32. [Google Scholar] [CrossRef] [Green Version]
  169. Di Paolo, T.; Grégoire, L.; Feuerbach, D.; Elbast, W.; Weiss, M.; Gomez-Mancilla, B. AQW051, a novel and selective nicotinic acetylcholine receptor α7 partial agonist, reduces l-Dopa-induced dyskinesias and extends the duration of l-Dopa effects in parkinsonian monkeys. Park. Relat. Disord. 2014, 20, 1119–1123. [Google Scholar] [CrossRef]
  170. Booth, H.D.E.; Hirst, W.D.; Wade-Martins, R. The Role of Astrocyte Dysfunction in Parkinson’s Disease Pathogenesis. Trends Neurosci. 2017, 40, 358–370. [Google Scholar] [CrossRef] [Green Version]
  171. Wang, Q.; Liu, Y.; Zhou, J. Neuroinflammation in Parkinson’s disease and its potential as therapeutic target. Transl. Neurodegener. 2015, 4, 19. [Google Scholar] [CrossRef] [Green Version]
  172. Li, S.; Bi, G.; Han, S.; Huang, R. MicroRNAs Play a Role in Parkinson’s Disease by Regulating Microglia Function: From Pathogenetic Involvement to Therapeutic Potential. Front. Mol. Neurosci. 2021, 14, 744942. [Google Scholar] [CrossRef] [PubMed]
  173. Liu, B.; Du, L.; Hong, J.S. Naloxone protects rat dopaminergic neurons against inflammatory damage through inhibition of microglia activation and superoxide generation. J. Pharmacol. Exp. Ther. 2000, 293, 607–617. [Google Scholar] [PubMed]
  174. Vidović, M.; Rikalovic, M.G. Alpha-Synuclein Aggregation Pathway in Parkinson’s Disease: Current Status and Novel Therapeutic Approaches. Cells 2022, 11, 1732. [Google Scholar] [CrossRef]
  175. Kim, S.; Kwon, S.-H.; Kam, T.-I.; Panicker, N.; Karuppagounder, S.S.; Lee, S.; Lee, J.H.; Kim, W.R.; Kook, M.; Foss, C.A.; et al. Transneuronal Propagation of Pathologic α-Synuclein from the Gut to the Brain Models Parkinson’s Disease. Neuron 2019, 103, 627–641.e7. [Google Scholar] [CrossRef]
  176. Patel, K.R.; Cherian, J.; Gohil, K.; Atkinson, D. Schizophrenia: Overview and treatment options. Pharm. Ther. 2014, 39, 638–645. [Google Scholar]
  177. Kalkman, H.O.; Feuerbach, D. Modulatory effects of α7 nAChRs on the immune system and its relevance for CNS disorders. Cell. Mol. Life Sci. 2016, 73, 2511–2530. [Google Scholar] [CrossRef] [Green Version]
  178. Mueser, K.T.; Jeste, D.V. (Eds.) Clinical Handbook of Schizophrenia; The Guilford Press: New York, NY, USA, 2008; ISBN 978-1-59385-652-6. [Google Scholar]
  179. Müller, N. Inflammation in Schizophrenia: Pathogenetic Aspects and Therapeutic Considerations. Schizophr. Bull. 2018, 44, 973–982. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Trubetskoy, V.; Pardiñas, A.F.; Qi, T.; Panagiotaropoulou, G.; Awasthi, S.; Bigdeli, T.B.; Bryois, J.; Chen, C.-Y.; Dennison, C.A.; Hall, L.S.; et al. Mapping genomic loci implicates genes and synaptic biology in schizophrenia. Nature 2022, 604, 502–508. [Google Scholar] [CrossRef]
  181. Weiland, S.; Bertrand, D.; Leonard, S. Neuronal nicotinic acetylcholine receptors: From the gene to the disease. Behav. Brain Res. 2000, 113, 43–56. [Google Scholar] [CrossRef] [PubMed]
  182. Leonard, S.; Freedman, R. Genetics of chromosome 15q13-q14 in schizophrenia. Biol. Psychiatry 2006, 60, 115–122. [Google Scholar] [CrossRef]
  183. Sinkus, M.L.; Lee, M.J.; Gault, J.; Logel, J.; Short, M.; Freedman, R.; Christian, S.L.; Lyon, J.; Leonard, S. A 2-base pair deletion polymorphism in the partial duplication of the alpha7 nicotinic acetylcholine gene (CHRFAM7A) on chromosome 15q14 is associated with schizophrenia. Brain Res. 2009, 1291, 1–11. [Google Scholar] [CrossRef] [Green Version]
  184. Leonard, S.; Gault, J.; Hopkins, J.; Logel, J.; Vianzon, R.; Short, M.; Drebing, C.; Berger, R.; Venn, D.; Sirota, P.; et al. Association of promoter variants in the alpha7 nicotinic acetylcholine receptor subunit gene with an inhibitory deficit found in schizophrenia. Arch. Gen. Psychiatry 2002, 59, 1085–1096. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. de Lucas-Cerrillo, A.M.; Maldifassi, M.C.; Arnalich, F.; Renart, J.; Atienza, G.; Serantes, R.; Cruces, J.; Sánchez-Pacheco, A.; Andrés-Mateos, E.; Montiel, C. Function of partially duplicated human α77 nicotinic receptor subunit CHRFAM7A gene: Potential implications for the cholinergic anti-inflammatory response. J. Biol. Chem. 2011, 286, 594–606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Villiger, Y.; Szanto, I.; Jaconi, S.; Blanchet, C.; Buisson, B.; Krause, K.-H.; Bertrand, D.; Romand, J.-A. Expression of an alpha7 duplicate nicotinic acetylcholine receptor-related protein in human leukocytes. J. Neuroimmunol. 2002, 126, 86–98. [Google Scholar] [CrossRef] [PubMed]
  187. Leonard, S.; Adler, L.E.; Benhammou, K.; Berger, R.; Breese, C.R.; Drebing, C.; Gault, J.; Lee, M.J.; Logel, J.; Olincy, A.; et al. Smoking and mental illness. Pharmacol. Biochem. Behav. 2001, 70, 561–570. [Google Scholar] [CrossRef] [PubMed]
  188. Freedman, R.; Coon, H.; Myles-Worsley, M.; Orr-Urtreger, A.; Olincy, A.; Davis, A.; Polymeropoulos, M.; Holik, J.; Hopkins, J.; Hoff, M.; et al. Linkage of a neurophysiological deficit in schizophrenia to a chromosome 15 locus. Proc. Natl. Acad. Sci. USA 1997, 94, 587–592. [Google Scholar] [CrossRef]
  189. Freedman, R.; Olincy, A.; Ross, R.G.; Waldo, M.C.; Stevens, K.E.; Adler, L.E.; Leonard, S. The genetics of sensory gating deficits in schizophrenia. Curr. Psychiatry Rep. 2003, 5, 155–161. [Google Scholar] [CrossRef] [PubMed]
  190. Hughes, J.R.; Hatsukami, D.K.; Mitchell, J.E.; Dahlgren, L.A. Prevalence of smoking among psychiatric outpatients. Am. J. Psychiatry 1986, 143, 993–997. [Google Scholar] [CrossRef] [PubMed]
  191. Goff, D.C.; Henderson, D.C.; Amico, E. Cigarette smoking in schizophrenia: Relationship to psychopathology and medication side effects. Am. J. Psychiatry 1992, 149, 1189–1194. [Google Scholar] [CrossRef]
  192. Koike, K.; Hashimoto, K.; Takai, N.; Shimizu, E.; Komatsu, N.; Watanabe, H.; Nakazato, M.; Okamura, N.; Stevens, K.E.; Freedman, R.; et al. Tropisetron improves deficits in auditory P50 suppression in schizophrenia. Schizophr. Res. 2005, 76, 67–72. [Google Scholar] [CrossRef]
  193. Toyohara, J.; Hashimoto, K. α7 Nicotinic Receptor Agonists: Potential Therapeutic Drugs for Treatment of Cognitive Impairments in Schizophrenia and Alzheimer’s Disease. Open Med. Chem. J. 2010, 4, 37–56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Mackowick, K.M.; Barr, M.S.; Wing, V.C.; Rabin, R.A.; Ouellet-Plamondon, C.; George, T.P. Neurocognitive endophenotypes in schizophrenia: Modulation by nicotinic receptor systems. Prog. Neuropsychopharmacol. Biol. Psychiatry 2014, 52, 79–85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Parikh, V.; Kutlu, M.G.; Gould, T.J. nAChR dysfunction as a common substrate for schizophrenia and comorbid nicotine addiction: Current trends and perspectives. Schizophr. Res. 2016, 171, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Kunii, Y.; Zhang, W.; Xu, Q.; Hyde, T.M.; McFadden, W.; Shin, J.H.; Deep-Soboslay, A.; Ye, T.; Li, C.; Kleinman, J.E.; et al. CHRNA7 and CHRFAM7A mRNAs: Co-localized and their expression levels altered in the postmortem dorsolateral prefrontal cortex in major psychiatric disorders. Am. J. Psychiatry 2015, 172, 1122–1130. [Google Scholar] [CrossRef] [PubMed]
  197. De Luca, V.; Likhodi, O.; Van Tol, H.H.M.; Kennedy, J.L.; Wong, A.H.C. Regulation of alpha7-nicotinic receptor subunit and alpha7-like gene expression in the prefrontal cortex of patients with bipolar disorder and schizophrenia. Acta Psychiatr. Scand. 2006, 114, 211–215. [Google Scholar] [CrossRef] [PubMed]
  198. Gault, J.; Hopkins, J.; Berger, R.; Drebing, C.; Logel, J.; Walton, C.; Short, M.; Vianzon, R.; Olincy, A.; Ross, R.G.; et al. Comparison of polymorphisms in the alpha7 nicotinic receptor gene and its partial duplication in schizophrenic and control subjects. Am. J. Med. Genet. Part B Neuropsychiatr. Genet. 2003, 123, 39–49. [Google Scholar] [CrossRef]
  199. Becchetti, A.; Grandi, L.C.; Cerina, M.; Amadeo, A. Nicotinic acetylcholine receptors and epilepsy. Pharmacol. Res. 2023, 189, 106698. [Google Scholar] [CrossRef]
  200. Fisher, R.S.; van Emde Boas, W.; Blume, W.; Elger, C.; Genton, P.; Lee, P.; Engel, J.J. Epileptic seizures and epilepsy: Definitions proposed by the International League Against Epilepsy (ILAE) and the International Bureau for Epilepsy (IBE). Epilepsia 2005, 46, 470–472. [Google Scholar] [CrossRef]
  201. Pedley, T.A. Major advances in epilepsy in the last century: A personal perspective. Epilepsia 2009, 50, 358–363. [Google Scholar] [CrossRef] [PubMed]
  202. Alyu, F.; Dikmen, M. Inflammatory aspects of epileptogenesis: Contribution of molecular inflammatory mechanisms. Acta Neuropsychiatr. 2017, 29, 1–16. [Google Scholar] [CrossRef] [PubMed]
  203. Curia, G.; Longo, D.; Biagini, G.; Jones, R.S.G.; Avoli, M. The pilocarpine model of temporal lobe epilepsy. J. Neurosci. Methods 2008, 172, 143–157. [Google Scholar] [CrossRef]
  204. Hamilton, S.E.; Loose, M.D.; Qi, M.; Levey, A.I.; Hille, B.; McKnight, G.S.; Idzerda, R.L.; Nathanson, N.M. Disruption of the m1 receptor gene ablates muscarinic receptor-dependent M current regulation and seizure activity in mice. Proc. Natl. Acad. Sci. USA 1997, 94, 13311–13316. [Google Scholar] [CrossRef] [PubMed]
  205. Perucca, P.; Bahlo, M.; Berkovic, S.F. The Genetics of Epilepsy. Annu. Rev. Genomics Hum. Genet. 2020, 21, 205–230. [Google Scholar] [CrossRef] [PubMed]
  206. Steinlein, O.K.; Mulley, J.C.; Propping, P.; Wallace, R.H.; Phillips, H.A.; Sutherland, G.R.; Scheffer, I.E.; Berkovic, S.F. A missense mutation in the neuronal nicotinic acetylcholine receptor alpha 4 subunit is associated with autosomal dominant nocturnal frontal lobe epilepsy. Nat. Genet. 1995, 11, 201–203. [Google Scholar] [CrossRef]
  207. Becchetti, A.; Grandi, L.C.; Colombo, G.; Meneghini, S.; Amadeo, A. Nicotinic Receptors in Sleep-Related Hypermotor Epilepsy: Pathophysiology and Pharmacology. Brain Sci. 2020, 10, 907. [Google Scholar] [CrossRef] [PubMed]
  208. Becchetti, A.; Aracri, P.; Meneghini, S.; Brusco, S.; Amadeo, A. The role of nicotinic acetylcholine receptors in autosomal dominant nocturnal frontal lobe epilepsy. Front. Physiol. 2015, 6, 22. [Google Scholar] [CrossRef] [Green Version]
  209. Lévesque, M.; Biagini, G.; de Curtis, M.; Gnatkovsky, V.; Pitsch, J.; Wang, S.; Avoli, M. The pilocarpine model of mesial temporal lobe epilepsy: Over one decade later, with more rodent species and new investigative approaches. Neurosci. Biobehav. Rev. 2021, 130, 274–291. [Google Scholar] [CrossRef] [PubMed]
  210. Beleslin, D.B.; Krstić, S.K. Nicotine-induced convulsions in cats and central nicotinic receptors. Pharmacol. Biochem. Behav. 1986, 24, 1509–1511. [Google Scholar] [CrossRef] [PubMed]
  211. Damaj, M.I.; Glassco, W.; Dukat, M.; Martin, B.R. Pharmacological characterization of nicotine-induced seizures in mice. J. Pharmacol. Exp. Ther. 1999, 291, 1284–1291. [Google Scholar] [PubMed]
  212. Woolf, A.; Burkhart, K.; Caraccio, T.; Litovitz, T. Self-poisoning among adults using multiple transdermal nicotine patches. J. Toxicol. Clin. Toxicol. 1996, 34, 691–698. [Google Scholar] [CrossRef]
  213. Bastlund, J.F.; Berry, D.; Watson, W.P. Pharmacological and histological characterisation of nicotine-kindled seizures in mice. Neuropharmacology 2005, 48, 975–983. [Google Scholar] [CrossRef] [PubMed]
  214. Gomes, P.X.L.; de Oliveira, G.V.; de Araújo, F.Y.R.; de Barros Viana, G.S.; de Sousa, F.C.F.; Hyphantis, T.N.; Grunberg, N.E.; Carvalho, A.F.; Macêdo, D.S. Differences in vulnerability to nicotine-induced kindling between female and male periadolescent rats. Psychopharmacology 2013, 225, 115–126. [Google Scholar] [CrossRef] [PubMed]
  215. Dobelis, P.; Hutton, S.; Lu, Y.; Collins, A.C. GABAergic systems modulate nicotinic receptor-mediated seizures in mice. J. Pharmacol. Exp. Ther. 2003, 306, 1159–1166. [Google Scholar] [CrossRef] [PubMed]
  216. Newman, M.B.; Manresa, J.J.; Sanberg, P.R.; Shytle, R.D. Nicotine induced seizures blocked by mecamylamine and its stereoisomers. Life Sci. 2001, 69, 2583–2591. [Google Scholar] [CrossRef] [PubMed]
  217. Iha, H.A.; Kunisawa, N.; Shimizu, S.; Tokudome, K.; Mukai, T.; Kinboshi, M.; Ikeda, A.; Ito, H.; Serikawa, T.; Ohno, Y. Nicotine Elicits Convulsive Seizures by Activating Amygdalar Neurons. Front. Pharmacol. 2017, 8, 57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Cohen, S.L.; Morley, B.J.; Snead, O.C. An EEG analysis of convulsive activity produced by cholinergic agents. Prog. Neuropsychopharmacol. 1981, 5, 383–388. [Google Scholar] [CrossRef] [PubMed]
  219. Salas, R.; Orr-Urtreger, A.; Broide, R.S.; Beaudet, A.; Paylor, R.; De Biasi, M. The nicotinic acetylcholine receptor subunit alpha 5 mediates short-term effects of nicotine in vivo. Mol. Pharmacol. 2003, 63, 1059–1066. [Google Scholar] [CrossRef]
  220. Kedmi, M.; Beaudet, A.L.; Orr-Urtreger, A. Mice lacking neuronal nicotinic acetylcholine receptor beta4-subunit and mice lacking both alpha5- and beta4-subunits are highly resistant to nicotine-induced seizures. Physiol. Genom. 2004, 17, 221–229. [Google Scholar] [CrossRef]
  221. Wonnacott, S. Presynaptic nicotinic ACh receptors. Trends Neurosci. 1997, 20, 92–98. [Google Scholar] [CrossRef]
  222. Tinuper, P.; Bisulli, F.; Cross, J.H.; Hesdorffer, D.; Kahane, P.; Nobili, L.; Provini, F.; Scheffer, I.E.; Tassi, L.; Vignatelli, L.; et al. Definition and diagnostic criteria of sleep-related hypermotor epilepsy. Neurology 2016, 86, 1834–1842. [Google Scholar] [CrossRef]
  223. Hirose, S.; Iwata, H.; Akiyoshi, H.; Kobayashi, K.; Ito, M.; Wada, K.; Kaneko, S.; Mitsudome, A. A novel mutation of CHRNA4 responsible for autosomal dominant nocturnal frontal lobe epilepsy. Neurology 1999, 53, 1749–1753. [Google Scholar] [CrossRef] [PubMed]
  224. Aridon, P.; Marini, C.; Di Resta, C.; Brilli, E.; De Fusco, M.; Politi, F.; Parrini, E.; Manfredi, I.; Pisano, T.; Pruna, D.; et al. Increased sensitivity of the neuronal nicotinic receptor alpha 2 subunit causes familial epilepsy with nocturnal wandering and ictal fear. Am. J. Hum. Genet. 2006, 79, 342–350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Broide, R.S.; Salas, R.; Ji, D.; Paylor, R.; Patrick, J.W.; Dani, J.A.; De Biasi, M. Increased sensitivity to nicotine-induced seizures in mice expressing the L250T alpha 7 nicotinic acetylcholine receptor mutation. Mol. Pharmacol. 2002, 61, 695–705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Franceschini, D.; Paylor, R.; Broide, R.; Salas, R.; Bassetto, L.; Gotti, C.; De Biasi, M. Absence of alpha7-containing neuronal nicotinic acetylcholine receptors does not prevent nicotine-induced seizures. Brain Res. Mol. Brain Res. 2002, 98, 29–40. [Google Scholar] [CrossRef] [PubMed]
  227. Schaaf, C.P. Nicotinic acetylcholine receptors in human genetic disease. Genet. Med. 2014, 16, 649–656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Helbig, I.; Mefford, H.C.; Sharp, A.J.; Guipponi, M.; Fichera, M.; Franke, A.; Muhle, H.; de Kovel, C.; Baker, C.; von Spiczak, S.; et al. 15q13.3 microdeletions increase risk of idiopathic generalized epilepsy. Nat. Genet. 2009, 41, 160–162. [Google Scholar] [CrossRef]
  229. Shinawi, M.; Schaaf, C.P.; Bhatt, S.S.; Xia, Z.; Patel, A.; Cheung, S.W.; Lanpher, B.; Nagl, S.; Herding, H.S.; Nevinny-Stickel, C.; et al. A small recurrent deletion within 15q13.3 is associated with a range of neurodevelopmental phenotypes. Nat. Genet. 2009, 41, 1269–1271. [Google Scholar] [CrossRef] [Green Version]
  230. Sharp, A.J.; Mefford, H.C.; Li, K.; Baker, C.; Skinner, C.; Stevenson, R.E.; Schroer, R.J.; Novara, F.; De Gregori, M.; Ciccone, R.; et al. A recurrent 15q13.3 microdeletion syndrome associated with mental retardation and seizures. Nat. Genet. 2008, 40, 322–328. [Google Scholar] [CrossRef]
  231. Fejgin, K.; Nielsen, J.; Birknow, M.R.; Bastlund, J.F.; Nielsen, V.; Lauridsen, J.B.; Stefansson, H.; Steinberg, S.; Sorensen, H.B.D.; Mortensen, T.E.; et al. A mouse model that recapitulates cardinal features of the 15q13.3 microdeletion syndrome including schizophrenia- and epilepsy-related alterations. Biol. Psychiatry 2014, 76, 128–137. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic diagram showing key nuclei and pathways involved in addiction with strong participation of nAChRs. Nicotine binding to α7 and α4β2 nAChRs at the ventral tegmental area (VTA) promotes the initiation of addictive behavior by favoring the release of dopamine (DA) in the nucleus accumbens (NAc).
Figure 1. Schematic diagram showing key nuclei and pathways involved in addiction with strong participation of nAChRs. Nicotine binding to α7 and α4β2 nAChRs at the ventral tegmental area (VTA) promotes the initiation of addictive behavior by favoring the release of dopamine (DA) in the nucleus accumbens (NAc).
Cells 12 02051 g001
Figure 2. Schematic diagram of nodes and tracks of the cholinergic anti-inflammatory pathway. The connection between the vagus and the splenic nerve via the celiac ganglion promotes noradrenaline release (blue circles) and activation of splenic T-cells. T-cells release acetylcholine (ACh) that can bind to α7 nAChR on immune cells such as macrophages, inhibiting the release of pro-inflammatory cytokines. Activation of α7 nAChR inhibits the nuclear factor kappa-light-chain-enhancer of activated B cell (NFkB) translocation to the cell nucleus and activation of a Janus kinase 2 (JAK2)–signal transducer and activator of transcription 3 (STAT3)-mediated signaling pathway. In parallel, activation of α7 nAChRs may up-regulate the expression of interleukin-1 receptor-associated kinase M (IRAK-M), which can negatively regulate innate Toll-like receptor (TLR)-mediated immune responses, contributing to cholinergic anti-inflammatory effects.
Figure 2. Schematic diagram of nodes and tracks of the cholinergic anti-inflammatory pathway. The connection between the vagus and the splenic nerve via the celiac ganglion promotes noradrenaline release (blue circles) and activation of splenic T-cells. T-cells release acetylcholine (ACh) that can bind to α7 nAChR on immune cells such as macrophages, inhibiting the release of pro-inflammatory cytokines. Activation of α7 nAChR inhibits the nuclear factor kappa-light-chain-enhancer of activated B cell (NFkB) translocation to the cell nucleus and activation of a Janus kinase 2 (JAK2)–signal transducer and activator of transcription 3 (STAT3)-mediated signaling pathway. In parallel, activation of α7 nAChRs may up-regulate the expression of interleukin-1 receptor-associated kinase M (IRAK-M), which can negatively regulate innate Toll-like receptor (TLR)-mediated immune responses, contributing to cholinergic anti-inflammatory effects.
Cells 12 02051 g002
Figure 3. Dopaminergic neurons from the substantia nigra are substantially reduced in PD. Current treatments are aimed at counterbalancing the reduction of dopamine (DOPA) by exogenous administration of L-DOPA. nAChR agonists may offer amelioration of dyskinetic symptoms in PD by promoting the endogenous release of DOPA.
Figure 3. Dopaminergic neurons from the substantia nigra are substantially reduced in PD. Current treatments are aimed at counterbalancing the reduction of dopamine (DOPA) by exogenous administration of L-DOPA. nAChR agonists may offer amelioration of dyskinetic symptoms in PD by promoting the endogenous release of DOPA.
Cells 12 02051 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vallés, A.S.; Barrantes, F.J. Nicotinic Acetylcholine Receptor Dysfunction in Addiction and in Some Neurodegenerative and Neuropsychiatric Diseases. Cells 2023, 12, 2051. https://doi.org/10.3390/cells12162051

AMA Style

Vallés AS, Barrantes FJ. Nicotinic Acetylcholine Receptor Dysfunction in Addiction and in Some Neurodegenerative and Neuropsychiatric Diseases. Cells. 2023; 12(16):2051. https://doi.org/10.3390/cells12162051

Chicago/Turabian Style

Vallés, Ana Sofía, and Francisco J. Barrantes. 2023. "Nicotinic Acetylcholine Receptor Dysfunction in Addiction and in Some Neurodegenerative and Neuropsychiatric Diseases" Cells 12, no. 16: 2051. https://doi.org/10.3390/cells12162051

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

Article Metrics

Back to TopTop