Next Article in Journal
Hyperandrogenism and Its Possible Effects on Endometrial Receptivity: A Review
Next Article in Special Issue
The Proof-of-Concept of MBA121, a Tacrine–Ferulic Acid Hybrid, for Alzheimer’s Disease Therapy
Previous Article in Journal
DeepSATA: A Deep Learning-Based Sequence Analyzer Incorporating the Transcription Factor Binding Affinity to Dissect the Effects of Non-Coding Genetic Variants
Previous Article in Special Issue
Exploring the Potential of Sulfonamide-Dihydropyridine Hybrids as Multitargeted Ligands for Alzheimer’s Disease Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Sigma Receptors in Alzheimer’s Disease: New Potential Targets for Diagnosis and Therapy

1
Key Laboratory of Radiopharmaceuticals (Beijing Normal University), Ministry of Education, College of Chemistry, Beijing Normal University, Beijing 100875, China
2
Department of Nuclear Medicine, Xinqiao Hospital, Army Medical University, Chongqing 400037, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(15), 12025; https://doi.org/10.3390/ijms241512025
Submission received: 16 June 2023 / Revised: 14 July 2023 / Accepted: 16 July 2023 / Published: 27 July 2023
(This article belongs to the Special Issue Advances in Alzheimer’s Disease Drug Research and Development)

Abstract

:
Sigma (σ) receptors are a class of unique proteins with two subtypes: the sigma-1 (σ1) receptor which is situated at the mitochondria-associated endoplasmic reticulum (ER) membrane (MAM), and the sigma-2 (σ2) receptor, located in the ER-resident membrane. Increasing evidence indicates the involvement of both σ1 and σ2 receptors in the pathogenesis of Alzheimer’s disease (AD), and thus these receptors represent two potentially effective biomarkers for emerging AD therapies. The availability of optimal radioligands for positron emission tomography (PET) neuroimaging of the σ1 and σ2 receptors in humans will provide tools to monitor AD progression and treatment outcomes. In this review, we first summarize the significance of both receptors in the pathophysiology of AD and highlight AD therapeutic strategies related to the σ1 and σ2 receptors. We then survey the potential PET radioligands, with an emphasis on the requirements of optimal radioligands for imaging the σ1 or σ2 receptors in humans. Finally, we discuss current challenges in the development of PET radioligands for the σ1 or σ2 receptors, and the opportunities for neuroimaging to elucidate the σ1 and σ2 receptors as novel biomarkers for early AD diagnosis, and for monitoring of disease progression and AD drug efficacy.

1. Introduction

Alzheimer’s Disease (AD) is the most common form of dementia, accounting for 60–80% of all cases. According to the World Alzheimer Report 2022, about 55 million people lived with dementia worldwide in 2019. This number is expected to rise to 139 million in 2050 [1]. AD is characterized by two pathological hall markers—extracellular β-amyloid (Aβ) plaques and intraneuronal fibrillary tangles (NFTs) composed of hyperphosphorylated tau proteins. Based on the 2018 National Institute on Aging—Alzheimer’s Association (NIA-AA) research framework, biomarkers of AD are grouped into those for Aβ (A), pathologic tau (T), and neurodegeneration or neuronal injury (N) [2]. Currently, the ATN research framework, grounded on the above biomarker-based definition of AD, has been widely accepted for diagnosis of AD in clinical practice worldwide [2,3]. However, the exact cause of AD has not been fully elucidated, although many hypotheses on its etiology and pathogenesis have been put forward, including the Aβ cascade hypothesis [4], the misfolded tau protein hypothesis [5], the cholinergic hypothesis [6,7,8,9], oxidative stress [10], calcium dyshomeostasis [11,12], neuroinflammation [13,14] and the mitochondria cascade (or MAM) hypothesis [15,16].
Medications currently available on the market for AD treatment are listed in Table 1. Donepezil is a piperidine-based reversible inhibitor of acetylcholinesterase (AChE) with high inhibitory activity (IC50(AChE) = 5.7 nM) and high selectivity over butyrylcholinesterase (BuChE) (IC50(BuChE) = 7138 nM, 1252-fold) [17]. Interestingly, donepezil has also been shown to bind to the sigma-1 (σ1) receptors in the living human brain at therapeutic doses [18]. Rivastigmine is a carbamate-based pseudo-irreversible (slowly reversible) dual inhibitor of both AChE and BuChE with low inhibitory potency (IC50(AChE) = 32,100 nM, IC50(BuChE) = 390 nM, 82-fold) [19]. However, this drug has no affinity for muscarinic, α- or β-adrenergic, or dopamine (DA) receptors or opioid binding sites [20]. Galantamine is a reversible competitive inhibitor for AChE (IC50(AChE) = 350 nM) rather than BuChE (IC50(BuChE) = 18,600 nM, 53-fold) [21], and an allosteric modulator of nicotinic acetylcholine receptors [22]. The most common adverse events of these cholinesterase inhibitors are gastrointestinal (GI) and cardiovascular side effects. They are generally well-tolerated and all are still considered for first-line, symptomatic treatment of AD (for review, see [23,24,25,26,27]).
Memantine is a voltage-dependent, low affinity/fast off-rate and non-competitive N-methyl-D-aspartate receptor (NMDA) receptor antagonist [28]. It exerts its neuronal protective effects by inhibiting glutamate activity and is used for the treatment of moderate-to-severe AD alone or in combination with donepezil [29,30,31]. Adverse effects of memantine have been found to be comparable to those with a placebo, with the exception of an increased incidence of dizziness, headache, confusion, and constipation [32].
Both Aducanumab (Aduhelm) and Lecanemab (Leqembi) are anti-amyloid monoclonal antibodies (mAbs) and approved under the accelerated approval pathway for treatment of Alzheimer patients with mild cognitive impairment (MCI). On 6 July 2023, the US Food and Drug Administration (FDA) converted Leqembi to traditional approval. And thus, Leqembi represents the first Aβ-directed antibody fully approved for the treatment of AD without any restrictions. Aducanumab is a human IgG1 monoclonal antibody preferably targeting Aβ aggregates [33]. Lecanemab is a humanized monoclonal IgG1 of the mouse mAb158 selectively binding to soluble Aβ protofibrils [34]. Adverse effects from both medications include amyloid-related imaging abnormalities (ARIA) and infusion reactions [33,34,35,36]. Patients with apolipoprotein E ε4 (ApoE ε4) gene carriers, especially ApoE ε4 homozygotes, are at higher risk of ARIA such as brain swelling (edema or effusions) and bleeding (hemosiderin deposits) [36,37,38].
Sodium Oligomannate Capsules (GV-971) were approved by the National Medical Products Administration (NMPA) of China in 2019 for treatment of mild-to-moderate AD [39]. GV-971 was noted to cause an induced liver injury side effect [40]. The other side effects of GV-971 have not been extensively reported in the literature. However, international multicenter clinical trials of GV-971 (Phase III) were stopped in May 2022.
Recent evidence has pointed to the significance of sigma receptors in AD. The σ1 receptor, situated at the mitochondria-associated endoplasmic reticulum (ER) membrane (MAM), directly regulates Aβ generation in the MAM [41]. The sigma-2 (σ2) receptor has been positively identified as the ER-resident transmembrane protein 97 (TMEM97) [42], and the σ2 receptor/TMEM97, progesterone receptor membrane component 1 (PGRMC1) and low-density lipoprotein receptor (LDLR) have been found to form a trimeric complex and regulate the uptake of lipoproteins such as LDL and apolipoprotein E (ApoE) [43,44], whose E4 allele (ApoE ε4) is the greatest risk factor for AD development [45]. Further, this trimeric complex has been demonstrated to be a binding site for Aβ oligomers (AβOs). Inhibition of one of the three proteins results in disruption of the complex and decreased AβO uptake in neurons [44]. As a result, both the σ1 and σ2 receptors are increasingly viewed as playing critical roles in AD pathogenesis and progression, and are thus important targets for therapeutic intervention to inhibit Aβ neurotoxicity, neurodegeneration, and progression of AD [46,47,48,49,50,51,52,53].

2. Sigma Receptors in AD

2.1. The Sigma (σ) Receptors

The sigma (σ) receptors were initially identified in 1976 and thought to belong to the class of “opioid receptors” [54]. They were later found to possess binding sites distinct from those of opioid receptors [55], and divided into two subtypes termed σ1 and σ2 [56], based on the different binding sites of the radioligands (+)-[3H]pentazocine and [3H]1,3-di(2-tolyl) guanidine ([3H]DTG) [57]. Both σ1 and σ2 receptors are widely distributed in the central nervous system (CNS) [57,58] and peripheral tissues [59,60,61,62], acting as integral membrane proteins and playing crucial roles in a variety of human diseases [63]. However, the regional expression patterns of σ1 and σ2 receptors in the brain are clearly different [57,64,65,66]. Recent quantitative autoradiography studies with postmortem human brain tissues found higher concentrations of σ2 than σ1 receptor in all brain regions examined, except the red nucleus, as well as upregulation of σ2 receptors in the aged brains [67,68].
The σ1 receptor, consisting of 223 amino acids with molecular weight of 25.3 kDa [69], has been cloned from several tissues, including those from mice, rats and guinea pigs [70,71,72,73], and proved to be a unique “ligand-operated receptor chaperone” that is regulated by the agonist/antagonist activity of endogenous or synthetic ligands [74,75]. The crystal structure of the human σ1 receptor was elucidated in 2016, and found to have a trimeric structure with a single transmembrane domain in each protomer and a β-barrel cupin fold in the carboxy terminal domain [76]. Currently, there is no consensus on endogenous ligands for the σ1 receptor (for review, see [63]), even though some candidates such as the hallucinogen N, N-Dimethyltryptamine [77] have been proposed.
Compared with the σ1 receptor, the identification process for the σ2 receptor is more convoluted. In 1994, Bowen et al. employed [3H]azido-DTG to estimate the molecular weight of σ2 receptor isolated from rat liver membrane, at 21.5 kDa [60]. Subsequently, this enigmatic protein has been hypothesized to contain a histone binding site [78,79]. The σ2 receptor complex was also found to contain PGRMC-1 protein complex [80,81]. Then, in 2017, the σ2 receptor was positively identified as the four-domain TMEM97 (also known as meningioma-related protein 30, MAC30), residing in the ER membrane [42]. Finally, in 2021, Alon et al. successfully determined the crystal structure of the bovine σ2 receptor using high-affinity ligands [82]. The σ2 receptor/TMEM97 is now revealed as an intimately associated homodimer, with each of the two protomers having four kinked transmembrane helix sections, and both the N and C terminals facing the cytoplasm. The binding site of the σ2 receptor ligand is deeply embedded in the membrane, which suggests that a lipid may be the endogenous ligand [82]. The binding pocket opens laterally into the lipid bilayer, and its opening is lined with hydrophobic and aromatic residues [82]. Up to date, two putative endogenous ligands, histatin-1 [83] and 20(S)-hydroxycholesterol (20(S)-OHC) [84], have been reported.

2.2. Sigma-1 Receptor in AD

The σ1 receptor participates in various physiological and pathological processes, such as neurotransmission, neuroprotection and neuroinflammation, through interaction with diverse ion channels, ER proteins, neurotrophins and G protein-coupled transporters (GPCRs) [85]. Consequently, the σ1 receptor has been considered as a therapeutic target for a range of diseases [50,86] including amnesia and AD [46,85,87,88], Parkinson’s disease (PD) [89,90,91,92,93], Huntington’s disease (HD) [94,95], amyotrophic lateral sclerosis (ALS) [96,97], retinal disease [98,99,100,101], multiple sclerosis (MS) [102], major depressive disorder (MDD) [89,103], stroke [104,105,106], pain [107,108] and drug or alcohol addiction [109]. In cancers, the σ1 receptor is involved in tumor occurrence, development, metastasis and survival [110,111,112,113].
Increasing evidence has proved that the σ1 receptor holds great potential as a biomarker for early AD diagnosis and progression and the monitoring of AD drug efficacy [49]. Hallmarks of human AD include progressive cognitive decline that follows chronic neuroinflammation and the emergence of hyperphosphorylated tau protein aggregates and Aβ plaques [46], with all playing critical and perhaps interrelated roles in the progression of AD. In particular, Ca2+ plays a critical role in learning and memory processes [114,115,116,117,118], and Ca2+ dyshomeostasis is a pathological feature of AD and other neurodegenerative diseases [114,115,116,117,118]. The σ1 receptor forms a Ca2+-sensitive chaperone complex with the binding immunoglobulin protein/glucose-regulated protein 78 (BiP/GRP78), and prolongs Ca2+-signaling from ER into mitochondria by stabilizing inositol 1,4,5-trisphosphate receptor (IP3R) at the MAM [74]. The σ1 receptor participates in the regulation of intracellular Ca2+ migration and maintains homeostasis and protects cognitive function damage [119,120]. Research on the role of the σ1 receptor in mediating mitochondrial function has found that the σ1 receptor attenuates hippocampal dendrite formation through scavenging of free radicals, and protects cells from damage by mitochondria-derived reactive oxygen species (ROS) [121,122,123]. The ER stress sensor inositol-requiring enzyme 1 (IRE1) facilitates mitochondrion-ER-nucleus signaling for cellular survival via the σ1 receptor chaperone [75,124].
More importantly, the σ1 receptors regulate early Aβ generation in AD at the MAM [41], and thus could be considered as a bona fide MAM marker and responsible for neuroprotective regulatory functions [87]. It has been proposed that σ1 receptor comprises part of the endogenous cellular defense against toxic Aβ [85,87]. Growing evidence has demonstrated the neuroprotective activity of the σ1 receptor against Aβ neurotoxicity (for review, see [46,125]). Activation of σ1 receptor potentiates nerve growth factor (NGF)-induced neurite outgrowth through modulating the PLCγ-DAG-PKC, Ras-Raf-MEK-ERK-MAPK signaling pathways and protects Aβ25–35-impaired dentritic growth and survival of newborn neurons through a modulation of PI3K-Akt-mTOR-p70S6k signaling [119,126,127]. Moreover, activating the σ1 receptor increases vascular endothelial growth factor (VEGF) and low-density lipoprotein receptor-related protein 1 (LRP-1) expression levels and attenuates the blood–brain barrier (BBB) dysfunction caused by amyloid deposition in AD [48].
Additionally, the mixed muscarinic/σ1 ligand ANAVEX2-73 prevents tau hyperphosphorylation in Aβ25–35-injected mice [128]. Further, the σ1 receptor regulates proper tau phosphorylation and axon development by promoting p35 turnover via σ1 receptor–myristic acid interaction, thereby avoiding cyclin-dependent kinase 5 (CDK5)/P25 overactivity [129,130].
Studies with σ1 receptor knock-out mice showed that σ1 receptor deletion resulted in neurogenesis impairment [131] and cognitive dysfunction [132] including memory deficit, neurocyte susceptibility to Aβ-mediated toxicity and impairment to the intracellular lipid metabolism and immune response, and thus accelerated neural degeneration and oxidative stress-induced neural death [132]. Emerging evidence indicates that certain polymorphisms of the σ1 receptor gene, especially when present alongside the known AD risk factor ApoE ε4, are linked to the onset of AD neurodegeneration [133]. In a postmortem study, reduction in σ1 receptors was observed in the hippocampus of patients with AD [134]. Compounds with σ1 agonist activity have been shown to possess anti-amnestic and neuroprotective efficacy in both pharmacological and pathological AD models [119], including those resulting from cholinergic destruction [135,136], Aβ administration [135,137,138,139,140,141], glutamatergic/serotonergic [142] or calcium channel deficits [143] and normal aging [144], as well as senescence-accelerated mouse (SAM) model [145].
Taken together, it becomes increasingly evident that the σ1 receptor plays a key role in mediating AD pathology, and therefore presents as promising therapeutic target for AD.

2.3. Sigma-2 Receptor in AD

Compared to the σ1 receptor, there are only a few reports on the cellular and molecular biological roles of the σ2 receptor. It functions as a housekeeping protein under normal settings [146]. The σ2 receptor/TMEM97, a member of the expanded emopamil binding protein (EPR) superfamily, has sterol isomerase activity [147,148] and plays a critical role in cholesterol biology, with correlated expression genes taking part in lipid metabolism [147]. High expression of the σ2 receptor was found in a variety of tumor cells, with nearly 10-fold higher expression in a proliferating state tumor compared to a quiescent state [149,150,151,152]. The differential expression of σ2 receptors is associated with tumor stage, metastasis, and survival. As such, the σ2 receptor can act as a novel biomarker for tumor proliferation [152], and is thus a candidate target for the diagnosis and treatment of common hyperplastic tumors [153]. Further, results from recent studies have indicated the critical involvement of the σ2 receptor in the pathophysiology of many brain disorders. Thus, the σ2 receptor has been proposed as a novel therapeutic target for AD, HD, PD [147,154,155], depression [156], schizophrenia [157], neuropathic pain [158,159], and age-related macular degeneration [148].
In AD pathology, the σ2 receptor interacts with PGRMC1 and LDLR to block AβO from binding neuronal synapses and regulates cholesterol homeostasis [44], and acts as a novel biomarker for AD diagnose and drug development [160].
Recent consensus regards AβO as one of the most toxic and pathogenic forms of Aβ, and elevated AβO levels in the brain as the key causative factor in the formation of Aβ plaques [161]. Studies have shown that AβO-induced neurotoxicity subsequently caused synaptic injury and hampered synaptic plasticity, resulting in abnormalities in synaptic composition, structure and density [162]. The effects of AβO on receptors and signaling pathways are neurodegenerative changes, neuronal injury, synaptic dysfunction and neurofibrillary tangles (NFTs), which eventually lead to memory, learning and cognitive dysfunction. Compared with Aβ monomers and Aβ fibrils, soluble AβOs are more likely to induce neuronal loss and cognitive deficits in amyloid precursor protein (APP)/tau transgenic mice, and their concentrations correlated better with AD severity [163]. Hence, the prevention or reversal of AβO-induced neurotoxicity is thought to be key to AD treatment.
Several lines of evidence have pointed to the critical involvement of the σ2 receptor in mediating AβO neurotoxicity and thus the key role it plays in AD pathogenesis and progression [67]. Synthetic AβOs derived from the brains of AD patients were discovered to attach to nerve cells and display typical receptor–ligand pharmacological interaction [160,164]. AβOs specifically and saturably bound to hippocampal and cortical neurons both in vivo and in vitro. AβO treatment induced progressive upregulation of σ2 receptor expression in neurons, with more intense AβO binding associated with higher σ2 expression. Selective σ2 receptor modulators competitively inhibited/reversed AβO binding to neurons, and prevented synapse loss in a dose-dependent manner both in vitro, and in rat models of AD [164].
The σ2 receptor regulates the binding and signal transmission of AβO in CNS, and its antagonists can decrease AβO binding to nerve cells and disassociate the attached AβO from neurons [53,160]. Studies have found that the σ2 receptor/TMEM97, PGRMC1 and LDLR can form a ternary complex (σ2R-PGRMC1-LDLR) [43], which is a binding site for monomeric and oligomeric amyloid Aβ42, and plays an essential role in the uptake of fibers and oligomers via ApoE-dependent and independent mechanisms [44]. The knockout of the TMEM97/σ2 receptor, or PGRMC-1, or both, as well as inhibition of the TMEM97/σ2 receptor were all shown to reduce the uptake of Aβ1–42 and ApoE in primary neurons [44]. Moreover, the expression of the σ2 receptor is dramatically increased by approximately 1.5-fold in AD [165], and is localized to an increased area of synapses (approximately 1.8-fold) in brain tissue taken from people suffering from AD compared with healthy controls, suggesting a compensatory response to AD-related synaptic depression [148,165]. Neurons with knockout of the PGRMC-1 protein also displayed reduced capacity in binding to AβO [160]. The σ2 receptor/TMEM97 is present in synaptic fractions biochemically isolated from human temporal cortex, and its concentrations appeared to be higher in samples isolated from AD patient brains compared to those from healthy controls [165].
As a cholesterol-regulating protein [51], the malfunctions of the σ2 receptor/TMEM97 are involved in AD pathology [51,52]. The σ2 receptor ligands also potentially influence Aβ synthesis via cholesterol, which has been demonstrated to directly affect APP cleavage in neuronal cultures by boosting β- and γ-secretase activity [166]. In CNS, the neurons obtain cholesterol mostly via multiple ApoE receptors including LDLR, very-low-density lipoprotein receptor (VLDLR), and LDLR-related protein 1 (LRP1) [167]. High cholesterol levels are recognized to be a risk factor for AD [168]. The σ2 receptor ligands can potentially interrupt lipoprotein transport [167], decrease the level of cholesterol and exert anti-AD effects. Similarly, σ2 receptor ligands can influence tau phosphorylation via cholesterol [169]. Simultaneously, the hyperphosphorylated tau is found in lipid rafts, implying that cholesterol has the ability to control tau hyperphosphorylation [170]. More recently, a close physical colocalization of TMEM97 and TSPO was found in MP cells. The σ2 receptor ligands such as siramesine modulated TMEM97-TSPO association [171]. In addition, σ2 receptor ligands have been also reported to activate liver X receptors (LXRs) through oxysterols and inhibit the expression of inflammatory genes, thereby regulating neuroinflammation in AD [147,172]. Therefore, the σ2 receptor is a novel regulator of cholesterol homeostasis in the AD pathological process, and its ligands may target cholesterol homeostasis for AD treatment [147].
Similar to the σ1 receptor, the σ2 receptor is involved in the regulation of intracellular Ca2+ levels [173,174,175]. Binding of AβO to neurons upregulates the σ2 receptor in AD and triggers ER stress, to disrupt Ca2+ homeostasis. Antagonism of the σ2 receptor is believed to reduce ER stress, maintain Ca2+ homeostasis and protect neurons [176]. Small molecules acting at the σ2 receptor have also been shown to exert their neuroprotective activity via regulation of neuroinflammation and the nerve growth factor (NGF) [176,177,178].
Taken together, the intact σ2 receptor/TMEM97-PGRMC1-LDLR complex is a pathway for the cellular uptake of AβO via ApoE-dependent and independent mechanisms. The loss or pharmacological inhibition of one or both of these proteins results in the disruption of the complex leading to decreased uptake of AβO and ApoE in neurons. Targeting the σ2 receptor/TMEM97 represents a new strategy for inhibiting Aβ neurotoxicity and slowing neurodegeneration in AD [53,148].

3. Ligands Targeting σ1 or σ2 Receptors

3.1. Agonists/Antagonists of σ Receptors and Their Therapeutic Potential in Clinical Trials

Many ligands targeting σ1 and σ2 receptors have been investigated for their therapeutic potential [46,49,50,179,180,181,182,183,184,185]. Representative agents in clinical trials are provided in Figure 1 and Table 2. The progress of these ligands has been covered in some recent reviews [50,185]. Here, we describe only the ligands with therapeutic potential for AD.
Dextromethorphan (AVP-923), a σ1 receptor agonist, has been found to have multiple mechanisms of action that could be beneficial in AD, such as anti-inflammatory and antioxidant effects, modulation of neurotransmitters, and a neuroprotective effect by inhibiting Aβ aggregation and tau hyperphosphorylation in AD [186,187,188]. ANAVEX2-73 (blarcamesine), a σ1 receptor agonist, has been shown in Phase II clinical trials to provide significant and sustained improvement in cognitive function and reduce neurodegenerative pathology in mild-AD patients [189]. Endonerpic maleate (T-817MA), an orally available neurotropic drug with high affinity to the σ1 receptor [50,190], attenuates Aβ-induced neurotoxicity and memory deficits, promotes neurite outgrowth, and preserves hippocampal synapses, likely via σ1 receptor activation [191,192]. CT1812, a σ2 receptor allosteric antagonist for mild-to-moderate AD treatment [193], has proved to displace toxic AβO from the synaptic receptor, facilitate oligomer clearance into the CSF and restore cognitive function [193,194].
It is important to note that while these σ receptor ligands have shown promising results in preclinical studies, their efficacy and safety in human clinical trials are still being evaluated. It will require further research and development to determine their potential as therapeutic options for AD.
Table 2. Representative σ1 and σ2 ligands in clinical trials a.
Table 2. Representative σ1 and σ2 ligands in clinical trials a.
AgentsPropertyDiseaseClinical Trials IDPhase/Status
Pridopidineσ1 receptor agonistHDNCT03019289
NCT00724048
NCT04556656
NCT00665223
NCT02006472
NCT01306929
NCT02494778
I/Completed
II/III/Completed
III/Recruiting
III/Completed
II/Completed
II/Completed
II/Terminated
Levodopa-induced dyskinesia (PD)NCT03922711II/Terminated
ALSNCT04297683
NCT04615923
II/III/Recruiting
II/III/Active, not recruiting
Dextromethorphan (AVP-923) bσ1 receptor agonist
mu (μ) opioid agonist and NMDA receptor antagonist [195]
ADNCT00788047
NCT01584440
NCT01832350
NCT02446132
NCT02442778
NCT02442765
NCT00726726
NCT04947553
NCT05557409
NCT04797715
NCT00056524
I/Completed
II/Completed
IV/Terminated
III/Recruiting
III/Completed
III/Completed
I/Completed
III/Recruiting
III/Recruiting
III/Completed
III/Completed
SA4503σ1 receptor agonistIschemic strokeNCT00639249II/Completed
MDDNCT00551109II/Completed
MR309 (E-52862)σ1 receptor antagonistOxaliplatin-induced neuropathyRef. [196]IIa/Completed
ANAVEX2-73 (blarcamesine)σ1 receptor agonist
muscarinic receptor modulator
Moderate ADNCT04314934
NCT03790709
NCT02756858
NCT02244541
IIb/III/Recruiting
IIb/III/Completed
II/Completed
IIa/Completed
Rett syndromeNCT04304482
NCT03941444
NCT03758924
II/Recruiting
III/Completed
II/Completed
PDNCT03774459II/Completed
Edonerpic maleate (T-817MA)σ1 receptor activationMild-to-moderate ADNCT00663936
NCT04191486
II/Completed
II/Recruiting
Aβ inhibitorNCT02079909II/Completed
Hepatic impairmentNCT02693197I/Completed
Roluperidone (MIN-101)σ2 receptor antagonist and 5-HT2A receptor antagonistNegative symptoms of schizophreniaNCT03397134III/Completed
NCT02232529I/Completed
SchizophreniaNCT03038646I/Completed
Healthy subjectsNCT03072056I/Completed
CT1812σ2 receptor antagonistHealthy volunteersNCT03716427I/Completed
ADNCT05531656II/Not recruiting
NCT04735536II/Completed
NCT02907567I/II/Completed
NCT05248672I/Completed
NCT05225389I/Completed
NCT03507790II/Recruiting
NCT03493282I/II/Completed
NCT03522129I/Completed
Age-related macular degenerationNCT05893537II/Recruiting
Dementia with Lewy bodiesNCT05225415II/Recruiting
Cognitive impairmentNCT02570997I/Completed
a The clinical trials were obtained from https://www.clinicaltrials.gov (accessed on 14 July 2023). b There are 115 clinical trials for Dextromethorphan. Only AD-related trials are listed.

3.2. Development of Radioligands for Neuroimaging of σ Receptors

3.2.1. Characteristics of Optimal σ1 or σ2 Receptor Radioligands for PET Imaging in AD

Non-invasive radioligand-based molecular imaging technique such as positron emission tomography (PET) imaging is a powerful tool for the investigation of protein target expression and function in living subjects. It can visualize molecular biological processes in normal and disease states [197]. PET imaging of AD pathologic biomarkers such as Aβ and tau has been widely used for evaluating the pathologic features of AD, tracking AD progression, monitoring therapeutic interventions and facilitating drug development based on the ATN research framework [198]. Increasing evidence in recent years has proved that the σ1 and σ2 receptors play significantly distinct roles in AD pathology [46,147]. In vivo visualization of the σ1 and σ2 receptor changes in the progression of AD with PET radioligands will shed new light on the involvement of these receptors in the etiology and pathophysiology of AD, and provide a tool to monitor the treatment effect of σ1 and σ2 receptor-targeted therapeutic agents.
Similar to other neuroimaging radioligands, the development of suitable PET radioligands targeting the σ1 or σ2 receptor is a great challenge, due to the limited information of the target protein in the brain and presence of the blood–brain barrier (BBB). The optimal radioligands for imaging of the σ1 or σ2 receptors in the brain need to meet the following requirements: (1) appropriate affinity for the σ1 or σ2 receptors and high selectivity over other receptors, transporters and ion channels (> 50-fold); (2) an efficient method for radiosynthesis, with good radiochemical yield and high molar activity; (3) suitable physical–chemical properties, including desirable lipophilicity (log D = 1–3) and in vitro stability; (4) high brain uptake (SUV > 1) and high brain-to-blood ratios; (5) excellent in vivo stability without radioactive metabolites able to enter the brain; (6) appropriate pharmacological properties that reflect the regional expression of σ1 or σ2 receptors in the brain; (7) high specific binding to the σ1 or σ2 receptors in vivo in brain tissue; (8) suitable kinetic (reversible binding) in the human brain; and (9) acceptable toxicological properties, with no side effects in the range of injectable doses.

3.2.2. Radioligands Targeting the σ1 Receptor

Over the last two decades, many efforts have been devoted to the development of PET radioligands for the σ1 receptors. However, only a few radioligands have been investigated in non-human primates and humans, due to the difficulties in meeting the critical requirements outlined above [49,153,199,200,201]. They are depicted in Figure 2 and reviewed below.
[11C]SA4503 ([11C]1) was the first PET radioligand used for imaging the σ1 receptor in humans [202]. SA4503 was reported as a σ1 receptor agonist with high affinity and subtype selectivity over the σ2 receptor, and with low affinity for 36 other target proteins in the brain, except for the vesicular acetylcholine transporter (VAChT), with moderate affinity (Ki = 50.2 nM) [203,204,205]. Later, several groups reinvestigated the binding properties of SA4503 and reported slightly different affinities and subtype selectivity (Ki1) = 3.3–4.6 nM; Ki2) = 51–242 nM; Ki2)/Ki1) = 14–55) [203,206,207,208]. Density (Bmax) of the σ1 receptor was estimated to be 30–600 fmol/mg protein (approximately 3–60 nM) in the human brain [134,209,210]. Radioligands with nanomolar affinity (1–6 nM) appear to be suitable for σ1 receptor imaging in the brain, suggesting that SA4503 has a suitable range of affinity for quantitative in vivo imaging. Studies in rodents, cats, monkeys and humans indicated its potential to map σ1 receptors in the brain [58,202,211,212,213,214,215]. As a result, [11C]SA4503 has been used to investigate the σ1 receptor density in the brains of patients with AD [216,217] and PD [218], and σ1 receptor occupancy by fluvoxamine [219] and donepezil [18] at clinical doses. It should be noted that two studies with [11C]SA4503 to image σ1 receptor density in the brains of patients with AD have reported discrepant results. In an initial study, decreased accumulation of [11C]SA4503 was observed in the brains of AD patients, seemingly indicating downregulation of the σ1 receptor in AD [202,216]. However, a recent study using the same radioligand clearly demonstrated increased σ1 receptor expression (approximately 27% increase) in the brains of early-onset AD patients [217]. It was proposed that the lower σ1 receptor in AD observed in the previous study [216] may be attributed to the individuals treated with donepezil, which has high affinity for the σ1 receptor [220], and thus can act as a blocking agent to decrease the binding of [11C]SA4503 [217]. Despite these uses of [11C]SA4503 in clinical imaging research, it is not an optimal radioligand for neuroimaging, because of its relatively low subtype selectivity over the σ2 receptor (Ki1) = 3.3–4.6 nM; Ki2) = 51–63 nM; Ki2)/Ki1) = 14–15) [206,207,208] and relatively slow kinetics for a 11C-labeled radioligand in non-human primates and humans. Further, the use of [11C]SA4503 needs an on-site cyclotron, due to the short half-life of 11C (T1/2 = 20.4 min). Therefore, more recent efforts have been focused on the search for an optimal radioligand labeled with the longer-lived radionuclide 18F (T1/2 = 109.8 min), and thus are more amenable to translation into clinics for diagnostic applications and multicenter studies of drug-target occupancy and the dose–efficacy relationship.
Among the 18F-labelled radioligands shown in Figure 2, [18F]FTC-146 ([18F]2) [221,222] and [18F]FPS ([18F]3) [223] were found to display irreversible kinetics in the human brain and therefore are not suitable for neuroimaging applications.
Radioligands [18F]4, [18F]5 [224], (S)-(–)-[18F]fluspidine ([18F]6) and (R)-(+)-[18F]fluspidine ([18F]7) [225] are 18F-labeled spirocyclic piperidine ligands derived from the lead compound, 1′-benzyl-3-methoxy-3H-spiro [2-benzofuran-1,4′-piperidine], with nanomolar affinity for σ1 receptors and excellent selectivity over σ2 and more than 60 other receptors, transporters, and ion channels [226,227,228]. Their binding affinity, lipophilicity (Log D), the free fraction in monkey plasma (fP) and the regional non-displaceable binding potential (BPND) values are listed in Table 3. As expected, and consistent with results from rodent and pig studies [224,229,230,231,232], radioligands [18F]4 and (R)-(+)-[18F]fluspidine, with subnanomolar affinity for the σ1 receptors, exhibited irreversible binding kinetics in monkey brain regions, with no appreciable washout during the 4 h scan. And thus, they were judged as unsuitable for human neuroimaging. On the other hand, [18F]5 and (S)-(–)-[18F]fluspidine, with nanomolar affinity, displayed fast and reversible kinetics with good uptake and high specific binding signal in the monkey brain. These results again reinforce the principle in PET radioligand development: for a given target protein, higher affinity of the radioligand will lead to slower kinetics in the brain. Moreover, radioligand [18F]5 displayed on average > 2 times higher BPND values than (S)-(–)-[18F]fluspidine [233]. Note that (S)-(−)-[18F]fluspidine has been used to evaluate σ1 receptor changes in patients with major depression [234] and the σ1 receptor occupancy by pridopidine in the human brain of healthy volunteers and in patients with Huntington’s disease [235]. However, there has been no report on the use of this radioligand to investigate the σ1 receptor in AD.
In the past decades, most of the σ1 receptor ligands have been designed and synthesized based on Glennon’s pharmacophore model (two hydrophobic regions and a basic nitrogen atom) [236]. Encouraged by the results from the spirocyclic piperidine radioligands described above, we undertook a study to develop a radioligand constructed from a novel scaffold and with optimal lipophilicity. Wuensch considered the benzene ring of the O-heterocycle of the spirocyclic piperidine derivative as the “primary hydrophobic region” and the phenyl group of the N-substituent as the “secondary hydrophobic region” of the σ1 ligands in Glennon’s pharmacophore model [237]. We replaced the spirocyclic piperidine moiety in [18F]4 with a more hydrophilic group 1,4-dioxa-8-azaspiro [4.5]decane and simple piperidine [238]. The resulting ligands were found to maintain nanomolar affinity and subtype selectivity for σ1 receptors, indicating that removal of the benzene ring from the spiro(isobenzofuran piperidine) moiety still preserves the high affinity for the σ1 receptors [238]. Later, we replaced 1,4-dioxa-8-azaspiro [4.5]decane with 1,3-dioxane [239] or a tetrahydrofuran moiety [240], and found that these derivatives also maintained nanomolar affinity for the σ1 receptors. These findings demonstrated that smaller and less lipophilic moieties may serve as the “primary hydrophobic region” in the piperidine series of ligands. Thus the “primary hydrophobic region” in Glennon’s pharmacophore model appears to be more flexible, and can accommodate diverse structural moieties, not just those with an aromatic component.
Inspired by these new discoveries in Glennon’s pharmacophore model for σ1 receptor ligands, we designed and synthesized a novel radioligand [18F]FBFP with the smallest primary and secondary hydrophobic regions up to date for a σ1 receptor ligand. Gratifyingly, [18F]FBFP was found to have nanomolar affinity for the σ1 receptor, and high selectivity over the σ2 receptor, VAChT, and ten other receptors [240]. Studies in rodents and non-human primates indicated that this radioligand displayed fast, good brain uptake, favorable tissue kinetics, the highest plasma-free fraction and the highest specific binding signals in non-human primates among the σ1 receptor radioligands evaluated to date [240,241].
Similar to [18F]fluspidine, [18F]FBFP has a chiral center at the tetrahydrofuran moiety (denoted with an asterisk * in the structures shown in Figure 2), and thus is composed of two enantiomers. Enantiopure (S)-FBFP and (R)-FBFP were prepared from chiral synthesis with > 98% enantiomeric purity. In vitro evaluation demonstrated that (R)-FBFP with minus specific rotation behaved as an antagonist, while (S)-FBFP with plus specific rotation behaved as an agonist [242]. Both enantiomers possessed comparable low nanomolar affinity for the σ1 receptors and high selectivity over more than 40 other proteins. The enantiomerically pure radioligands (S)-(+)-[18F]FBFP and (R)-(−)-[18F]FBFP were obtained from their corresponding iodonium ylide precursors. Evaluation in rodents demonstrated excellent properties of both (S)-(+)-[18F]FBFP and (R)-(−)-[18F]FBFP with high brain uptake, high brain-to-blood ratios, high metabolic stability in the brain and high specific binding to the σ1 receptors [242]. In rhesus monkeys, both enantiomers display high brain uptake. Compared to (S)-(−)-[18F]fluspidine, both enantiomers exhibited much higher binding potential (BPND) in rhesus monkeys (ranging from 9.6 (thalamus) to 27.7 (frontal cortex) for (R)-(−)-[18F]FBFP vs. 6.3 (cerebellum) to 14.8 (cingulate cortex) for (S)-(+)-[18F]FBFP) [243]. Although additional validation is required to assess utility in humans, both (R)-(−)-[18F]FBFP and (S)-(+)-[18F]FBFP, with the highest BPND values among the current available σ1 receptor ligands, meet all the requirements mentioned above for an optimal radioligand, and thus hold great potential for PET imaging and quantification of the σ1 receptor changes in AD patients. Both radioligands are currently undergoing evaluation in humans.
In summary, it appears that subnanomolar affinity for the σ1 receptors (Ki < 1 nM) will result in near-irreversible binding kinetics in the non-human primate brain. Radioligands [18F]5, [18F]6, [18F]8 and [18F]9 are suitable candidates for imaging σ1 receptors in humans.
Table 3. Binding affinity (Ki, nM), Log D, fP and BPND of the σ1 receptor radioligands.
Table 3. Binding affinity (Ki, nM), Log D, fP and BPND of the σ1 receptor radioligands.
LigandKi1)Ki2)SelectivityLog DfPBPND
[18F]4 a0.792773512.558%-
[18F]5 a2.303271422.5018%2.78–5.21
[18F]6 a2.308973902.802%0.77–1.85
[18F]7 a0.57165028952.802%
[18F]8 b2.262991270.76 c73%6.3–14.8
[18F]9 b1.612461520.76 c67%9.6–27.7
a From Ref. [233]. b From Ref. [243]. c From Ref. [240].

3.2.3. Radioligands Targeting the σ2 Receptor

During the last decades, efforts in the development of σ2 receptor radioligands have been largely directed toward in vivo imaging of tumors in which upregulation of the σ2 receptor is found. Currently, [18F]ISO-1 ([18F]10) (Figure 3 and Table 4) is the only σ2 receptor radiotracer used in humans for tumor imaging [244,245]. However, it is not suitable for investigating neuronal σ2 receptors, due to its low brain uptake. There has been rekindled interest in the σ2 receptor as a therapeutic target for the treatment of neurologic and psychiatric diseases, especially AD. For example, the σ2 receptor antagonist CT1812 (Figure 1) is reported to prevent the binding of Aβ oligomers to neuronal receptors, and thus holds potential as a novel drug for the treatment of AD [193,194,246]. Hence, there remains an unmet clinical need to develop a suitable radioligand for neuroimaging of the σ2 receptor/TMEM97 in the human brain to investigate this target in AD progression, and to elucidate target engagement and the treatment mechanism of σ2 receptor-targeted drug candidates such as CT1812, in clinical trials.
Similar to what was found for σ1 receptor radioligands, the radiotracers for imaging σ2 receptors in the brain must meet the critical requirements outlined above. Due to undefined σ2 density in the brain of healthy human subjects and AD patients, the suitable affinity range required for effective imaging of σ2 receptors is yet to be defined. There has also been a paucity of ligands with high affinity and selectivity for the σ2 receptors. Although CT1812 is currently in Phase II clinical trial for treatment of mild-to-moderate AD, its affinity and subtype selectivity is only moderate [193]. Therefore, development of a suitable radioligand for neuroimaging of the σ2 receptors is even more challenging than the σ1 receptors. Nonetheless, there have been some recent activities in this endeavor, with several reports of brain-penetrant σ2 receptor radioligands, as depicted in Figure 3 [247,248,249,250].
In our search for highly selective radioligands for imaging σ2 receptors in the brain, we turned to the synthesis and evaluation of indole-based derivatives. Structure-activity relationship studies revealed that ligands with a four-carbon chain between the indole ring and the 6,7-dimethoxy-1,2,3,4-tetrahydroisoquinoline or 5,6-dimethoxyisoindoline pharmacophore displayed high σ2 receptor affinity and selectivity. Initial in vivo results indicated that radioligands with the 6,7-dimethoxy-1,2,3,4-tetrahydro-isoquinoline pharmacophore had lower brain uptake, brain-to-blood ratio, and σ2-specific binding [248,250]. Therefore, we focused on ligands with the 5,6-dimethoxyisoindoline pharmacophore to examine the influence of the fluoroethoxy group at different positions of the indole ring on tracer kinetics and specific binding. In biodistribution studies of three radioligands (named [18F]SYB4 ([18F]11), [18F]SYB5 ([18F]12), and [18F]SYB6 ([18F]13) (Table 4) in mice, they were found to readily enter the brain with good uptake (3.76–4.55% ID/g, 2 min) and high brain-to-blood ratios (10.6 for [18F]11, 30–60 min; 3.1 for [18F]12 and 4.5 for [18F]13, 15 min), and to bind specifically to the σ2 receptor, indicating a significant achievement as the first set of radioligands demonstrated to be suitable for brain imaging purposes [248,250].
Ex vivo autoradiography and blocking studies demonstrated a high level of regionally heterogeneous specific binding of [18F]11 in the mouse brain [248], with the distribution pattern clearly different from that we observed recently with a σ1 receptor radioligand [239]. Analysis results from a metabolism study in ICR mice indicated that the parent compound [18F]11 or [18F]13 was the predominant radioactive species (> 95%), indicating negligible entry of radiometabolites into the brain. Dynamic PET imaging and blocking studies in Sprague-Dawley rats confirmed regionally distinct distribution and high specific binding of radioligand [18F]11 to the σ2 receptors in the rat brain [248].
Table 4. Binding affinity (Ki, nM) and Log D of the σ2 receptor radioligands.
Table 4. Binding affinity (Ki, nM) and Log D of the σ2 receptor radioligands.
LigandKi1)Ki2)SelectivityLog D
[18F]SYB4 a3711.792072.43 b
[18F]SYB5 a1873.27572.29 b
[18F]SYB6 a3762.631432.17
[18F]ISO-1 c3306.95483.06
[18F]ISO-1 d
[18F]ISO-1 e
102
95.1
28.2
13.3
4
7
3.06
3.06
a From Ref. [250]. b From Ref. [248]. c From Ref. [244]. d From Ref. [251]. e From Ref. [252].
In evaluation in monkeys, [18F]SYB4 ([18F]11) [253] and [18F]SYB6 ([18F]13) [254] exhibited fast and reversible kinetics, with peak SUV of 2.2–4.5 and 2.5–3.6, respectively, within 30 min in grey matter regions. The highest uptake was in the cerebellum and putamen, followed by similar uptake values in the hippocampus and caudate. Pretreatment with CM398 (0.2 mg/kg) reduced tracer uptake significantly across all brain regions. Regional BPND values ranged from 0.56 (amygdala) to 2.59 (cerebellum) for [18F]11 and 0.92 (amygdala) to 2.11 (cerebellum) for [18F]13, indicating specific binding of both radioligands to the σ2 receptors [253,254]. These two radioligands represent the first generation of PET radiotracers demonstrated to be suitable for imaging and quantification of the σ2 receptor in the primate brain.
In addition to [18F]SYB4 and [18F]SYB6, several other putative σ2 receptor probes ([18F]14, [11C]15, [11C]16, and [18F]17, Figure 3) have been reported to have good brain uptake in mice. No further reports are available for their evaluation in non-human primates or humans.

4. Concluding Remarks

Hallmarks of human AD pathology (Aβ plaques and hyperphosphorylated tau protein tangles) have been proved to play critical and interrelated roles in AD pathogenesis and progression. However, increasing evidence has demonstrated the central role of the MAM dysfunctions in AD pathogenesis [255,256,257,258,259]. As a key chaperone situated at the MAM, the σ1 receptor is closely related to early Aβ generation, tau neurotoxicity, oxidative stress, and calcium dyshomeostasis [257]. Indeed, donepezil, the ‘gold standard’ acetylcholinesterase inhibitor (AChEI) in the symptomatic treatment of AD, has been found to have significant σ1 binding affinity (IC50 of 29.1 nM) [220], and to exert its anti-amnestic and neuroprotective activities against Aβ toxicity through activation of the σ1 receptors [139,220,260]. Rivastigmine, another AChEI used for AD treatment, is also found to derive its activity to enhance neuronal growth through interaction with the σ1 and σ2 receptors [178]. Finally, ANAVEX2-73, a σ1 receptor agonist, has been shown to provide significant and sustained improvement in cognitive function in mild-AD patients [189]. As a regulator of Aβ production and a surrogate biomarker for mitochondrial function, the σ1 receptor is increasingly viewed as playing a critical role in AD pathogenesis and progression, and thus holds great potential as an important target for therapeutic intervention and as a biomarker for early diagnosis, progression and monitoring of AD drug efficacy [49].
As a cholesterol-regulating gene, the σ2 receptor/TMEM97, PGRMC1 and LDLR form a trimeric complex (TMEM97/PGRMC1/LDLR) and behave as a binding site for monomeric and oligomeric amyloid β-peptide (1–42) (Aβ1–42) [44]. CT1812, a σ2 receptor antagonist in clinical trial for AD treatment, is reported to prevent the binding of Aβ oligomers to neuronal receptors [194,246], and to reduce the interaction between the σ2 receptor and Aβ oligomers in synapse in a dose-dependent manner [193,194,246]. The recently FDA-approved mAb for AD therapy, Leqembi, is shown to prevent the formation of Aβ oligomers which bind to the σ2 receptor/TMEM97-PGRMC1-LDLR complex [34]. Hence, the σ2 receptor/TMEM97 is considered a therapeutic target for AD.
For in vivo investigation of σ receptors, radioligands [18F]5, [18F]6, [18F]8 and [18F]9 have been proved to be suitable candidates for neuroimaging of σ1 receptors in non-human primates [233,241,243], with [18F]6, [18F]8 and [18F]9 in the clinical trials. Two radioligands, [18F]SYB4 and [18F]SYB6, have been found to be promising for neuroimaging of the σ2 receptors in rodents and non-human primates [248,250,253,254]. Their further investigation in clinical studies may finally afford us a radioligand suitable for imaging the σ2 receptor in humans. Advancement of these novel radioligands for imaging the σ1 and σ2 receptors in AD, especially in longitudinal studies, will visualize the changes in these receptors along the disease progression pathway, and thus help to elucidate the key roles of the σ1 and σ2 receptors in AD pathogenesis and progression, and to facilitate the development of effective therapeutic strategies for AD.

Author Contributions

Conceptualization, T.W. and H.J.; writing—original draft preparation, H.J.; writing—review and editing, T.W. and H.J.; supervision, project administration, and funding acquisition, H.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Beijing Natural Science Foundation (No. 7212203) and the National Natural Science Foundation of China (No. 22276016 and 21876013).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ADAlzheimer’s disease
AChEIAcetylcholinesterase inhibitors
APPAmyloid precursor protein
ApoEApolipoprotein E gene
ARIAAmyloid-related imaging abnormalities
β-amyloid
AβOsAβ oligomers
ALSAmyotrophic lateral sclerosis
BBBBlood–brain barrier
BuChEButyrylcholinesterase
BiPBinding immunoglobulin protein
CNSCentral nervous system
CDK5Cyclin-dependent kinase 5
DADopamine
EREndoplasmic reticulum
EPREmopamil binding protein
GIGastrointestinal
GPCRsG protein-coupled transporters
GRP78Glucose-regulated protein 78
HDHuntington’s Disease
IP3Rinositol 1,4,5-trisphosphate receptor
IRE1Inositol-requiring enzyme 1
LDLRLow-density lipoprotein receptor
LRP-1Low-density lipoprotein receptor-related protein 1
LXRsLiver X receptors
mAbMonoclonal antibody
MAMMitochondria-associated ER membrane
MAC30Meningioma-associated protein 30
MDDMajor depressive disorder
MSMultiple sclerosis
MCIMild cognitive impairment
NFTsNeurofibrillary tangles
NMDAN-methyl-D-aspartate receptor
NGFNerve growth factor
PDParkinson’s disease
PETPositron Emission Tomography
PGRMC1Progesterone receptor membrane component 1
ROSReactive oxygen species
SAMSenescence-accelerated mouse
TMEM97Transmembrane protein 97
VEGFVascular endothelial growth factor
VAChTVesicular acetylcholine transporter
VLDLRVery-low-density lipoprotein receptor

References

  1. Gauthier, S.; Webster, C.; Servaes, S.; Morais, J.A.; Rosa-Neto, P. World Alzheimer Report 2022: Life after Diagnosis: Navigating Treatment, Care and Support; Alzheimer’s Disease International: London, UK, 2022. [Google Scholar]
  2. Jack, C.R., Jr.; Bennett, D.A.; Blennow, K.; Carrillo, M.C.; Dunn, B.; Haeberlein, S.B.; Holtzman, D.M.; Jagust, W.; Jessen, F.; Karlawish, J.; et al. NIA-AA research framework: Toward a biological definition of Alzheimer’s disease. Alzheimers Dement. 2018, 14, 535–562. [Google Scholar] [CrossRef] [PubMed]
  3. Knopman, D.S.; Haeberlein, S.B.; Carrillo, M.C.; Hendrix, J.A.; Kerchner, G.; Margolin, R.; Maruff, P.; Miller, D.S.; Tong, G.; Tome, M.B.; et al. The national institute on aging and the Alzheimer’s association research framework for Alzheimer’s disease: Perspectives from the research roundtable. Alzheimers Dement. 2018, 14, 563–575. [Google Scholar] [CrossRef]
  4. Hardy, J.A.; Higgins, G.A. Alzheimer’s disease: The amyloid cascade hypothesis. Science 1992, 256, 184–185. [Google Scholar] [CrossRef]
  5. Grundke-Iqbal, I.; Iqbal, K.; Tung, Y.C.; Quinlan, M.; Wisniewski, H.M.; Binder, L.I. Abnormal phosphorylation of the microtubule-associated protein tau in Alzheimer cytoskeletal pathology. Proc. Natl. Acad. Sci. USA 1986, 83, 4913–4917. [Google Scholar] [CrossRef] [PubMed]
  6. Davies, P.; Maloney, A.J. Selective loss of central cholinergic neurons in Alzheimer’s disease. Lancet 1976, 2, 1403. [Google Scholar] [CrossRef] [PubMed]
  7. Coyle, J.T.; Price, D.L.; DeLong, M.R. Alzheimer’s disease: A disorder of cortical cholinergic innervation. Science 1983, 219, 1184–1190. [Google Scholar] [CrossRef] [PubMed]
  8. Aigner, T.G. Pharmacology of memory: Cholinergic-glutamatergic interactions. Curr. Opin. Neurobiol. 1995, 5, 155–160. [Google Scholar] [CrossRef]
  9. Bartus, R.T.; Dean, R.L., 3rd; Beer, B.; Lippa, A.S. The cholinergic hypothesis of geriatric memory dysfunction. Science 1982, 217, 408–414. [Google Scholar] [CrossRef]
  10. Bai, R.; Guo, J.; Ye, X.Y.; Xie, Y.; Xie, T. Oxidative stress: The core pathogenesis and mechanism of Alzheimer’s disease. Ageing Res. Rev. 2022, 77, 101619. [Google Scholar] [CrossRef]
  11. Khachaturian, Z.S. Calcium, membranes, aging, and Alzheimer’s disease. Introduction and overview. Ann. N. Y. Acad. Sci. 1989, 568, 1–4. [Google Scholar] [CrossRef]
  12. Cascella, R.; Cecchi, C. Calcium dyshomeostasis in Alzheimer’s disease pathogenesis. Int. J. Mol. Sci. 2021, 22, 4914. [Google Scholar] [CrossRef] [PubMed]
  13. McGeer, P.L.; McGeer, E.G. The amyloid cascade-inflammatory hypothesis of Alzheimer disease: Implications for therapy. Acta Neuropathol. 2013, 126, 479–497. [Google Scholar] [CrossRef] [PubMed]
  14. Wong-Guerra, M.; Calfio, C.; Maccioni, R.B.; Rojo, L.E. Revisiting the neuroinflammation hypothesis in Alzheimer’s disease: A focus on the druggability of current targets. Front. Pharmacol. 2023, 14, 1161850. [Google Scholar] [CrossRef] [PubMed]
  15. Area-Gomez, E.; Schon, E.A. On the pathogenesis of Alzheimer’s disease: The MAM hypothesis. FASEB J. 2017, 31, 864–867. [Google Scholar] [CrossRef] [Green Version]
  16. Swerdlow, R.H. Mitochondria and mitochondrial cascades in Alzheimer’s disease. J. Alzheimers Dis. 2018, 62, 1403–1416. [Google Scholar] [CrossRef] [Green Version]
  17. Sugimoto, H.; Iimura, Y.; Yamanishi, Y.; Yamatsu, K. Synthesis and structure-activity relationships of acetylcholinesterase inhibitors: 1-benzyl-4-[(5,6-dimethoxy-1-oxoindan-2-yl)methyl]piperidine hydrochloride and related compounds. J. Med. Chem. 1995, 38, 4821–4829. [Google Scholar] [CrossRef]
  18. Ishikawa, M.; Sakata, M.; Ishii, K.; Kimura, Y.; Oda, K.; Toyohara, J.; Wu, J.; Ishiwata, K.; Iyo, M.; Hashimoto, K. High occupancy of sigma-1 receptors in the human brain after single oral administration of donepezil: A positron emission tomography study using [11C]SA4503. Int. J. Neuropsychopharmacol. 2009, 12, 1127–1131. [Google Scholar] [CrossRef]
  19. Vicente-Zurdo, D.; Rosales-Conrado, N.; León-González, M.E.; Brunetti, L.; Piemontese, L.; Pereira-Santos, A.R.; Cardoso, S.M.; Madrid, Y.; Chaves, S.; Santos, M.A. Novel rivastigmine derivatives as promising multi-target compounds for potential treatment of Alzheimer’s disease. Biomedicines 2022, 10, 1510. [Google Scholar] [CrossRef]
  20. Enz, A.; Amstutz, R.; Boddeke, H.; Gmelin, G.; Malanowski, J. Brain selective inhibition of acetylcholinesterase: A novel approach to therapy for Alzheimer’s disease. Prog. Brain Res. 1993, 98, 431–438. [Google Scholar]
  21. Thomsen, T.; Kewitz, H. Selective inhibition of human acetylcholinesterase by galanthamine in vitro and in vivo. Life Sci. 1990, 46, 1553–1558. [Google Scholar] [CrossRef]
  22. Maelicke, A.; Albuquerque, E.X. Allosteric modulation of nicotinic acetylcholine receptors as a treatment strategy for Alzheimer’s disease. Eur. J. Pharmacol. 2000, 393, 165–170. [Google Scholar] [CrossRef] [PubMed]
  23. Brewster, J.T., 2nd; Dell’Acqua, S.; Thach, D.Q.; Sessler, J.L. Classics in chemical neuroscience: Donepezil. ACS Chem. Neurosci. 2019, 10, 155–167. [Google Scholar] [CrossRef] [PubMed]
  24. Sharma, K. Cholinesterase inhibitors as Alzheimer’s therapeutics (Review). Mol. Med. Rep. 2019, 20, 1479–1487. [Google Scholar] [CrossRef] [Green Version]
  25. Haake, A.; Nguyen, K.; Friedman, L.; Chakkamparambil, B.; Grossberg, G.T. An update on the utility and safety of cholinesterase inhibitors for the treatment of Alzheimer’s disease. Expert Opin. Drug Saf. 2020, 19, 147–157. [Google Scholar] [CrossRef]
  26. Vecchio, I.; Sorrentino, L.; Paoletti, A.; Marra, R.; Arbitrio, M. The state of the art on acetylcholinesterase inhibitors in the treatment of Alzheimer’s disease. J. Cent. Nerv. Syst. Dis. 2021, 13, 11795735211029113. [Google Scholar] [CrossRef] [PubMed]
  27. Giacobini, E.; Cuello, A.C.; Fisher, A. Reimagining cholinergic therapy for Alzheimer’s disease. Brain 2022, 145, 2250–2275. [Google Scholar] [CrossRef] [PubMed]
  28. Lipton, S.A. The molecular basis of memantine action in Alzheimer’s disease and other neurologic disorders: Low-affinity, uncompetitive antagonism. Curr. Alzheimer Res. 2005, 2, 155–165. [Google Scholar] [CrossRef]
  29. Guo, J.; Wang, Z.; Liu, R.; Huang, Y.; Zhang, N.; Zhang, R. Memantine, donepezil, or combination therapy-what is the best therapy for Alzheimer’s disease? A network meta-analysis. Brain Behav. 2020, 10, e01831. [Google Scholar] [CrossRef]
  30. Arvanitakis, Z.; Shah, R.C.; Bennett, D.A. Diagnosis and management of dementia: Review. JAMA 2019, 322, 1589–1599. [Google Scholar] [CrossRef]
  31. Pardo-Moreno, T.; González-Acedo, A.; Rivas-Domínguez, A.; García-Morales, V.; García-Cozar, F.J.; Ramos-Rodríguez, J.J.; Melguizo-Rodríguez, L. Therapeutic approach to Alzheimer’s disease: Current treatments and new perspectives. Pharmaceutics 2022, 14, 1117. [Google Scholar] [CrossRef]
  32. Rossom, R.; Adityanjee; Dysken, M. Efficacy and tolerability of memantine in the treatment of dementia. Am. J. Geriatr. Pharmacother. 2004, 2, 303–312. [Google Scholar] [CrossRef]
  33. Sevigny, J.; Chiao, P.; Bussière, T.; Weinreb, P.H.; Williams, L.; Maier, M.; Dunstan, R.; Salloway, S.; Chen, T.; Ling, Y.; et al. The antibody aducanumab reduces Aβ plaques in Alzheimer’s disease. Nature 2016, 537, 50–56. [Google Scholar] [CrossRef] [PubMed]
  34. van Dyck, C.H.; Swanson, C.J.; Aisen, P.; Bateman, R.J.; Chen, C.; Gee, M.; Kanekiyo, M.; Li, D.; Reyderman, L.; Cohen, S.; et al. Lecanemab in early Alzheimer’s disease. N. Engl. J. Med. 2023, 388, 9–21. [Google Scholar] [CrossRef] [PubMed]
  35. Cummings, J.; Aisen, P.; Apostolova, L.G.; Atri, A.; Salloway, S.; Weiner, M. Aducanumab: Appropriate use recommendations. J. Prev. Alzheimers Dis. 2021, 8, 398–410. [Google Scholar] [CrossRef]
  36. Cummings, J.; Apostolova, L.; Rabinovici, G.D.; Atri, A.; Aisen, P.; Greenberg, S.; Hendrix, S.; Selkoe, D.; Weiner, M.; Petersen, R.C.; et al. Lecanemab: Appropriate use recommendations. J. Prev. Alzheimers Dis. 2023, 10, 362–377. [Google Scholar] [CrossRef]
  37. Blasco, D.; Roberts, J.S. Editorial: Implications of emerging uses of genetic testing for Alzheimer’s disease. J. Prev. Alzheimers Dis. 2023, 10, 359–361. [Google Scholar]
  38. Vogt, A.S.; Jennings, G.T.; Mohsen, M.O.; Vogel, M.; Bachmann, M.F. Alzheimer’s disease: A brief history of immunotherapies targeting amyloid β. Int. J. Mol. Sci. 2023, 24, 3895. [Google Scholar] [CrossRef] [PubMed]
  39. Syed, Y.Y. Sodium oligomannate: First approval. Drugs 2020, 80, 441–444. [Google Scholar] [CrossRef]
  40. Cheng, H.-R.; Wen, C.-Y.; Zhang, C.; Kwapong, W. The use of GV-971 induces liver injury in an Alzheimer’s disease patient. Authorea 2020. preprints. [Google Scholar] [CrossRef]
  41. Schreiner, B.; Hedskog, L.; Wiehager, B.; Ankarcrona, M. Amyloid-β peptides are generated in mitochondria-associated endoplasmic reticulum membranes. J. Alzheimers Dis. 2015, 43, 369–374. [Google Scholar] [CrossRef]
  42. Alon, A.; Schmidt, H.R.; Wood, M.D.; Sahn, J.J.; Martin, S.F.; Kruse, A.C. Identification of the gene that codes for the σ2 receptor. Proc. Natl. Acad. Sci. USA 2017, 114, 7160–7165. [Google Scholar] [CrossRef]
  43. Riad, A.; Zeng, C.; Weng, C.C.; Winters, H.; Xu, K.; Makvandi, M.; Metz, T.; Carlin, S.; Mach, R.H. Sigma-2 receptor/TMEM97 and PGRMC-1 increase the rate of internalization of LDL by LDL receptor through the formation of a ternary complex. Sci. Rep. 2018, 8, 16845. [Google Scholar] [CrossRef]
  44. Riad, A.; Lengyel-Zhand, Z.; Zeng, C.; Weng, C.C.; Lee, V.M.; Trojanowski, J.Q.; Mach, R.H. The sigma-2 receptor/TMEM97, PGRMC1, and LDL receptor complex are responsible for the cellular uptake of Aβ42 and its protein aggregates. Mol. Neurobiol. 2020, 57, 3803–3813. [Google Scholar] [CrossRef] [PubMed]
  45. Fehér, Á.; Juhász, A.; László, A.; Kálmán, J., Jr.; Pákáski, M.; Kálmán, J.; Janka, Z. Association between a variant of the sigma-1 receptor gene and Alzheimer’s disease. Neurosci. Lett. 2012, 517, 136–139. [Google Scholar] [CrossRef] [PubMed]
  46. Jin, J.L.; Fang, M.; Zhao, Y.X.; Liu, X.Y. Roles of sigma-1 receptors in Alzheimer’s disease. Int. J. Clin. Exp. Med. 2015, 8, 4808–4820. [Google Scholar]
  47. Marrazzo, A.; Caraci, F.; Salinaro, E.T.; Su, T.P.; Copani, A.; Ronsisvalle, G. Neuroprotective effects of sigma-1 receptor agonists against beta-amyloid-induced toxicity. Neuroreport 2005, 16, 1223–1226. [Google Scholar] [CrossRef]
  48. An, Y.; Qi, Y.; Li, Y.; Li, Z.; Yang, C.; Jia, D. Activation of the sigma-1 receptor attenuates blood-brain barrier disruption by inhibiting amyloid deposition in Alzheimer’s disease mice. Neurosci. Lett. 2022, 774, 136528. [Google Scholar] [CrossRef]
  49. Jia, H.; Zhang, Y.; Huang, Y. Imaging sigma receptors in the brain: New opportunities for diagnosis of Alzheimer’s disease and therapeutic development. Neurosci. Lett. 2019, 691, 3–10. [Google Scholar] [CrossRef]
  50. Ye, N.; Qin, W.; Tian, S.; Xu, Q.; Wold, E.A.; Zhou, J.; Zhen, X.C. Small molecules selectively targeting sigma-1 receptor for the treatment of neurological diseases. J. Med. Chem. 2020, 63, 15187–15217. [Google Scholar] [CrossRef]
  51. Bartz, F.; Kern, L.; Erz, D.; Zhu, M.; Gilbert, D.; Meinhof, T.; Wirkner, U.; Erfle, H.; Muckenthaler, M.; Pepperkok, R.; et al. Identification of cholesterol-regulating genes by targeted RNAi screening. Cell Metab. 2009, 10, 63–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Shen, H.; Li, J.; Xie, X.; Yang, H.; Zhang, M.; Wang, B.; Kent, K.C.; Plutzky, J.; Guo, L.W. BRD2 regulation of sigma-2 receptor upon cholesterol deprivation. Life Sci. Alliance 2021, 4, e201900540. [Google Scholar] [CrossRef] [PubMed]
  53. Ma, W.H.; Chen, A.F.; Xie, X.Y.; Huang, Y.S. Sigma ligands as potent inhibitors of Aβ and AβOs in neurons and promising therapeutic agents of Alzheimer’s disease. Neuropharmacology 2021, 190, 108342. [Google Scholar] [CrossRef]
  54. Martin, W.R.; Eades, C.G.; Thompson, J.A.; Huppler, R.E.; Gilbert, P.E. The effects of morphine- and nalorphine-like drugs in the nondependent and morphine-dependent chronic spinal dog. J. Pharmacol. Exp. Ther. 1976, 197, 517–532. [Google Scholar] [PubMed]
  55. Vaupel, D.B. Naltrexone fails to antagonize the sigma effects of PCP and SKF 10,047 in the dog. Eur. J. Pharmacol. 1983, 92, 269–274. [Google Scholar] [CrossRef] [PubMed]
  56. Quirion, R.; Bowen, W.D.; Itzhak, Y.; Junien, J.L.; Musacchio, J.M.; Rothman, R.B.; Su, T.P.; Tam, S.W.; Taylor, D.P. A proposal for the classification of sigma binding sites. Trends Pharmacol. Sci. 1992, 13, 85–86. [Google Scholar] [CrossRef]
  57. Bouchard, P.; Quirion, R. [3H]1,3-di(2-tolyl)guanidine and [3H](+)pentazocine binding sites in the rat brain: Autoradiographic visualization of the putative sigma-1 and sigma-2 receptor subtypes. Neuroscience 1997, 76, 467–477. [Google Scholar] [CrossRef] [PubMed]
  58. Kawamura, K.; Ishiwata, K.; Tajima, H.; Ishii, S.; Matsuno, K.; Homma, Y.; Senda, M. In vivo evaluation of [11C]SA4503 as a PET ligand for mapping CNS sigma-1 receptors. Nucl. Med. Biol. 2000, 27, 255–261. [Google Scholar] [CrossRef] [PubMed]
  59. Novakova, M.; Ela, C.; Barg, J.; Vogel, Z.; Hasin, Y.; Eilam, Y. Inotropic action of sigma receptor ligands in isolated cardiac myocytes from adult rats. Eur. J. Pharmacol. 1995, 286, 19–30. [Google Scholar] [CrossRef]
  60. Hellewell, S.B.; Bruce, A.; Feinstein, G.; Orringer, J.; Williams, W.; Bowen, W.D. Rat liver and kidney contain high densities of sigma-1 and sigma-2 receptors: Characterization by ligand binding and photoaffinity labeling. Eur. J. Pharmacol. 1994, 268, 9–18. [Google Scholar] [CrossRef]
  61. Liu, Y.; Whitlock, B.B.; Pultz, J.A.; Wolfe, S.A., Jr. Sigma-1 receptors modulate functional activity of rat splenocytes. J. Neuroimmunol. 1995, 59, 143–154. [Google Scholar] [CrossRef]
  62. Wolfe, S.A., Jr.; Ha, B.K.; Whitlock, B.B.; Saini, P. Differential localization of three distinct binding sites for sigma receptor ligands in rat spleen. J. Neuroimmunol. 1997, 72, 45–58. [Google Scholar] [CrossRef]
  63. Schmidt, H.R.; Kruse, A.C. The molecular function of σ receptors: Past, present, and future. Trends Pharmacol. Sci. 2019, 40, 636–654. [Google Scholar] [CrossRef] [PubMed]
  64. Walker, J.M.; Bowen, W.D.; Goldstein, S.R.; Roberts, A.H.; Patrick, S.L.; Hohmann, A.G.; DeCosta, B. Autoradiographic distribution of [3H](+)-pentazocine and [3H]1,3-di-o-tolylguanidine (DTG) binding sites in guinea pig brain: A comparative study. Brain Res. 1992, 581, 33–38. [Google Scholar] [CrossRef] [PubMed]
  65. Søby, K.K.; Mikkelsen, J.D.; Meier, E.; Thomsen, C. Lu 28-179 labels a σ2-site in rat and human brain. Neuropharmacology 2002, 43, 95–100. [Google Scholar] [CrossRef] [PubMed]
  66. Leitner, M.L.; Hohmann, A.G.; Patrick, S.L.; Walker, J.M. Regional variation in the ratio of sigma-1 to sigma-2 binding in rat brain. Eur. J. Pharmacol. 1994, 259, 65–69. [Google Scholar] [CrossRef] [PubMed]
  67. Izzo, N.J.; Colom-Cadena, M.; Riad, A.A.; Xu, J.; Singh, M.; Abate, C.; Cahill, M.A.; Spires-Jones, T.L.; Bowen, W.D.; Mach, R.H.; et al. Proceedings from the fourth international symposium on σ2 receptors: Role in health and disease. eNeuro 2020, 7, ENEURO.0317-20.2020. [Google Scholar] [CrossRef] [PubMed]
  68. Riad, A.; Xu, J.; Mach, R.H. Sigma-2 receptors: An emerging target for CNS PET imaging studies. In PET and SPECT of Neurobiological Systems; Dierckx, R.A.J.O., Otte, A., de Vries, E.F.J., van Waarde, A., Lammertsma, A.A., Eds.; Springer International Publishing: Cham, Switzerland, 2021; pp. 973–991. [Google Scholar]
  69. Weber, E.; Sonders, M.; Quarum, M.; McLean, S.; Pou, S.; Keana, J.F. 1,3-Di(2-[5-3H]tolyl)guanidine: A selective ligand that labels sigma-type receptors for psychotomimetic opiates and antipsychotic drugs. Proc. Natl. Acad. Sci. USA 1986, 83, 8784–8788. [Google Scholar] [CrossRef]
  70. Seth, P.; Fei, Y.J.; Li, H.W.; Huang, W.; Leibach, F.H.; Ganapathy, V. Cloning and functional characterization of a sigma receptor from rat brain. J. Neurochem. 1998, 70, 922–931. [Google Scholar] [CrossRef]
  71. Hanner, M.; Moebius, F.F.; Flandorfer, A.; Knaus, H.G.; Striessnig, J.; Kempner, E.; Glossmann, H. Purification, molecular cloning, and expression of the mammalian sigma1-binding site. Proc. Natl. Acad. Sci. USA 1996, 93, 8072–8077. [Google Scholar] [CrossRef]
  72. Seth, P.; Leibach, F.H.; Ganapathy, V. Cloning and structural analysis of the cDNA and the gene encoding the murine type 1 sigma receptor. Biochem. Biophys. Res. Commun. 1997, 241, 535–540. [Google Scholar] [CrossRef]
  73. Kekuda, R.; Prasad, P.D.; Fei, Y.J.; Leibach, F.H.; Ganapathy, V. Cloning and functional expression of the human type 1 sigma receptor (hSigmaR1). Biochem. Biophys. Res. Commun. 1996, 229, 553–558. [Google Scholar] [CrossRef] [PubMed]
  74. Hayashi, T.; Su, T.P. Sigma-1 receptor chaperones at the ER-mitochondrion interface regulate Ca2+ signaling and cell survival. Cell 2007, 131, 596–610. [Google Scholar] [CrossRef] [Green Version]
  75. Su, T.P.; Su, T.C.; Nakamura, Y.; Tsai, S.Y. The sigma-1 receptor as a pluripotent modulator in living systems. Trends Pharmacol. Sci. 2016, 37, 262–278. [Google Scholar] [CrossRef] [Green Version]
  76. Schmidt, H.R.; Zheng, S.; Gurpinar, E.; Koehl, A.; Manglik, A.; Kruse, A.C. Crystal structure of the human sigma-1 receptor. Nature 2016, 532, 527–530. [Google Scholar] [CrossRef] [Green Version]
  77. Fontanilla, D.; Johannessen, M.; Hajipour, A.R.; Cozzi, N.V.; Jackson, M.B.; Ruoho, A.E. The hallucinogen N,N-dimethyltryptamine (DMT) is an endogenous sigma-1 receptor regulator. Science 2009, 323, 934–937. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Colabufo, N.A.; Berardi, F.; Abate, C.; Contino, M.; Niso, M.; Perrone, R. Is the sigma-2 receptor a histone binding protein? J. Med. Chem. 2006, 49, 4153–4158. [Google Scholar] [CrossRef] [PubMed]
  79. Abate, C.; Elenewski, J.; Niso, M.; Berardi, F.; Colabufo, N.A.; Azzariti, A.; Perrone, R.; Glennon, R.A. Interaction of the sigma-2 receptor ligand PB28 with the human nucleosome: Computational and experimental probes of interaction with the H2A/H2B dimer. ChemMedChem 2010, 5, 268–273. [Google Scholar] [CrossRef]
  80. Xu, J.; Zeng, C.; Chu, W.; Pan, F.; Rothfuss, J.M.; Zhang, F.; Tu, Z.; Zhou, D.; Zeng, D.; Vangveravong, S.; et al. Identification of the PGRMC1 protein complex as the putative sigma-2 receptor binding site. Nat. Commun. 2011, 2, 380. [Google Scholar] [CrossRef] [Green Version]
  81. Zeng, C.; Garg, N.; Mach, R.H. The PGRMC1 protein level correlates with the binding activity of a sigma-2 fluorescent probe (SW120) in rat brain cells. Mol. Imaging Biol. 2016, 18, 172–179. [Google Scholar] [CrossRef]
  82. Alon, A.; Lyu, J.; Braz, J.M.; Tummino, T.A.; Craik, V.; O’Meara, M.J.; Webb, C.M.; Radchenko, D.S.; Moroz, Y.S.; Huang, X.P.; et al. Structures of the σ2 receptor enable docking for bioactive ligand discovery. Nature 2021, 600, 759–764. [Google Scholar] [CrossRef]
  83. Son, K.N.; Lee, H.; Shah, D.; Kalmodia, S.; Miller, R.C.; Ali, M.; Balasubramaniam, A.; Cologna, S.M.; Kong, H.; Shukla, D.; et al. Histatin-1 is an endogenous ligand of the sigma-2 receptor. FEBS J. 2021, 288, 6815–6827. [Google Scholar] [CrossRef]
  84. Cheng, Y.S.; Zhang, T.; Ma, X.; Pratuangtham, S.; Zhang, G.C.; Ondrus, A.A.; Mafi, A.; Lomenick, B.; Jones, J.J.; Ondrus, A.E. A proteome-wide map of 20(S)-hydroxycholesterol interactors in cell membranes. Nat. Chem. Biol. 2021, 17, 1271–1280. [Google Scholar] [CrossRef]
  85. Maurice, T.; Goguadze, N. Sigma-1 (σ1) receptor in memory and neurodegenerative diseases. Handb. Exp. Pharmacol. 2017, 244, 81–108. [Google Scholar]
  86. Piechal, A.; Jakimiuk, A.; Mirowska-Guzel, D. Sigma receptors and neurological disorders. Pharmacol. Rep. 2021, 73, 1582–1594. [Google Scholar] [CrossRef]
  87. Maurice, T.; Goguadze, N. Role of σ1 receptors in learning and memory and Alzheimer’s disease-type dementia. Adv. Exp. Med. Biol. 2017, 964, 213–233. [Google Scholar]
  88. Nguyen, L.; Lucke-Wold, B.P.; Mookerjee, S.; Kaushal, N.; Matsumoto, R.R. Sigma-1 receptors and neurodegenerative diseases: Towards a hypothesis of sigma-1 receptors as amplifiers of neurodegeneration and neuroprotection. Adv. Exp. Med. Biol. 2017, 964, 133–152. [Google Scholar]
  89. Yang, K.; Wang, C.; Sun, T. The roles of intracellular chaperone proteins, sigma receptors, in Parkinson’s disease (PD) and major depressive disorder (MDD). Front. Pharmacol. 2019, 10, 528. [Google Scholar] [CrossRef] [Green Version]
  90. Wilson, H.; Pagano, G.; de Natale, E.R.; Mansur, A.; Caminiti, S.P.; Polychronis, S.; Middleton, L.T.; Price, G.; Schmidt, K.F.; Gunn, R.N.; et al. Mitochondrial complex 1, sigma-1, and synaptic vesicle 2A in early drug-naive Parkinson’s disease. Mov. Disord. 2020, 35, 1416–1427. [Google Scholar] [CrossRef]
  91. Francardo, V.; Geva, M.; Bez, F.; Denis, Q.; Steiner, L.; Hayden, M.R.; Cenci, M.A. Pridopidine induces functional neurorestoration via the sigma-1 receptor in a mouse model of Parkinson’s disease. Neurotherapeutics 2019, 16, 465–479. [Google Scholar] [CrossRef] [Green Version]
  92. Guo, C.H.; Cao, T.; Zheng, L.T.; Waddington, J.L.; Zhen, X.C. Development and characterization of an inducible Dicer conditional knockout mouse model of Parkinson’s disease: Validation of the antiparkinsonian effects of a sigma-1 receptor agonist and dihydromyricetin. Acta Pharmacol. Sin. 2020, 41, 499–507. [Google Scholar] [CrossRef]
  93. Siddiqui, T.; Bhatt, L.K. Targeting Sigma-1 receptor: A promising strategy in the treatment of Parkinson’s disease. Neurochem. Res. 2023, 1–11. [Google Scholar] [CrossRef]
  94. Eddings, C.R.; Arbez, N.; Akimov, S.; Geva, M.; Hayden, M.R.; Ross, C.A. Pridopidine protects neurons from mutant-huntingtin toxicity via the sigma-1 receptor. Neurobiol. Dis. 2019, 129, 118–129. [Google Scholar] [CrossRef]
  95. Kraskovskaya, N.A.; Bezprozvanny, I.B. Normalization of calcium balance in striatal neurons in Huntington’s disease: Sigma-1 receptor as a potential target for therapy. Biochemistry 2021, 86, 471–479. [Google Scholar] [CrossRef]
  96. Herrando-Grabulosa, M.; Gaja-Capdevila, N.; Vela, J.M.; Navarro, X. Sigma 1 receptor as a therapeutic target for amyotrophic lateral sclerosis. Br. J. Pharmacol. 2021, 178, 1336–1352. [Google Scholar] [CrossRef]
  97. Mavlyutov, T.A.; Guo, L.W.; Epstein, M.L.; Ruoho, A.E. Role of the sigma-1 receptor in amyotrophic lateral sclerosis (ALS). J. Pharmacol. Sci. 2015, 127, 10–16. [Google Scholar] [CrossRef] [Green Version]
  98. Ellis, D.Z.; Li, L.; Park, Y.; He, S.; Mueller, B.; Yorio, T. Sigma-1 receptor regulates mitochondrial function in glucose- and oxygen-deprived retinal ganglion cells. Invest. Ophthalmol. Vis. Sci. 2017, 58, 2755–2764. [Google Scholar] [CrossRef] [Green Version]
  99. Smith, S.B.; Wang, J.; Cui, X.; Mysona, B.A.; Zhao, J.; Bollinger, K.E. Sigma-1 receptor: A novel therapeutic target in retinal disease. Prog. Retin. Eye Res. 2018, 67, 130–149. [Google Scholar] [CrossRef]
  100. Jiang, G.; Mysona, B.; Dun, Y.; Gnana-Prakasam, J.P.; Pabla, N.; Li, W.; Dong, Z.; Ganapathy, V.; Smith, S.B. Expression, subcellular localization, and regulation of sigma receptor in retinal muller cells. Invest. Ophthalmol. Vis. Sci. 2006, 47, 5576–5582. [Google Scholar] [CrossRef] [Green Version]
  101. Mavlyutov, T.A.; Epstein, M.; Guo, L.W. Subcellular localization of the sigma-1 receptor in retinal neurons—An electron microscopy study. Sci. Rep. 2015, 5, 10689. [Google Scholar] [CrossRef] [Green Version]
  102. Brimson, J.M.; Brimson, S.; Chomchoei, C.; Tencomnao, T. Using sigma-ligands as part of a multi-receptor approach to target diseases of the brain. Expert Opin. Ther. Targets 2020, 24, 1009–1028. [Google Scholar] [CrossRef]
  103. Wang, Y.M.; Xia, C.Y.; Jia, H.M.; He, J.; Lian, W.W.; Yan, Y.; Wang, W.P.; Zhang, W.K.; Xu, J.K. Sigma-1 receptor: A potential target for the development of antidepressants. Neurochem. Int. 2022, 159, 105390. [Google Scholar] [CrossRef] [PubMed]
  104. Rodríguez-Muñoz, M.; Onetti, Y.; Cortés-Montero, E.; Garzón, J.; Sánchez-Blázquez, P. Cannabidiol enhances morphine antinociception, diminishes NMDA-mediated seizures and reduces stroke damage via the sigma-1 receptor. Mol. Brain 2018, 11, 51. [Google Scholar] [CrossRef] [PubMed]
  105. Zhang, Y.; Zhang, X.; Wei, Q.; Leng, S.; Li, C.; Han, B.; Bai, Y.; Zhang, H.; Yao, H. Activation of sigma-1 receptor enhanced pericyte survival via the interplay between apoptosis and autophagy: Implications for blood-brain barrier integrity in stroke. Transl. Stroke Res. 2020, 11, 267–287. [Google Scholar] [CrossRef]
  106. Zhang, X.; Wu, F.; Jiao, Y.; Tang, T.; Yang, L.; Lu, C.; Zhang, Y.; Zhang, Y.; Bai, Y.; Chao, J.; et al. An increase of sigma-1 receptor in the penumbra neuron after acute ischemic stroke. J. Stroke Cerebrovasc. Dis. 2017, 26, 1981–1987. [Google Scholar] [CrossRef]
  107. Romero, L.; Merlos, M.; Vela, J.M. Antinociception by sigma-1 receptor antagonists: Central and peripheral effects. Adv. Pharmacol. 2016, 75, 179–215. [Google Scholar]
  108. Merlos, M.; Burgueño, J.; Portillo-Salido, E.; Plata-Salamán, C.R.; Vela, J.M. Pharmacological modulation of the sigma-1 receptor and the treatment of pain. Adv. Exp. Med. Biol. 2017, 964, 85–107. [Google Scholar]
  109. Skuza, G. Ethanol withdrawal-induced depressive symptoms in animals and therapeutic potential of sigma1 receptor ligands. Pharmacol. Rep. 2013, 65, 1681–1687. [Google Scholar] [CrossRef]
  110. Bai, T.; Lei, P.; Zhou, H.; Liang, R.; Zhu, R.; Wang, W.; Zhou, L.; Sun, Y. Sigma-1 receptor protects against ferroptosis in hepatocellular carcinoma cells. J. Cell. Mol. Med. 2019, 23, 7349–7359. [Google Scholar] [CrossRef] [Green Version]
  111. Soriani, O.; Kourrich, S. The sigma-1 receptor: When adaptive regulation of cell electrical activity contributes to stimulant addiction and cancer. Front. Neurosci. 2019, 13, 1186. [Google Scholar] [CrossRef] [Green Version]
  112. Brimson, J.M.; Akula, K.K.; Abbas, H.; Ferry, D.R.; Kulkarni, S.K.; Russell, S.T.; Tisdale, M.J.; Tencomnao, T.; Safrany, S.T. Simple ammonium salts acting on sigma-1 receptors yield potential treatments for cancer and depression. Sci. Rep. 2020, 10, 9251. [Google Scholar] [CrossRef]
  113. Pontisso, I.; Combettes, L. Role of sigma-1 receptor in calcium modulation: Possible involvement in cancer. Genes 2021, 12, 139. [Google Scholar] [CrossRef] [PubMed]
  114. Ma, T.; Gong, K.; Yan, Y.; Song, B.; Zhang, X.; Gong, Y. Mitochondrial modulation of store-operated Ca2+ entry in model cells of Alzheimer’s disease. Biochem. Biophys. Res. Commun. 2012, 426, 196–202. [Google Scholar] [CrossRef]
  115. Hung, C.H.; Ho, Y.S.; Chang, R.C. Modulation of mitochondrial calcium as a pharmacological target for Alzheimer’s disease. Ageing Res. Rev. 2010, 9, 447–456. [Google Scholar] [CrossRef] [Green Version]
  116. Guan, P.P.; Cao, L.L.; Wang, P. Elevating the levels of calcium ions exacerbate Alzheimer’s disease via inducing the production and aggregation of β-Amyloid protein and phosphorylated tau. Int. J. Mol. Sci. 2021, 22, 5900. [Google Scholar] [CrossRef]
  117. Bezprozvanny, I. Calcium signaling and neurodegenerative diseases. Trends Mol. Med. 2009, 15, 89–100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Alzheimer’s Association Calcium Hypothesis Workgroup; Khachaturian, Z.S. Calcium hypothesis of Alzheimer’s disease and brain aging: A framework for integrating new evidence into a comprehensive theory of pathogenesis. Alzheimers Dement. 2017, 13, 178–182.e17. [Google Scholar] [CrossRef]
  119. van Waarde, A.; Ramakrishnan, N.K.; Rybczynska, A.A.; Elsinga, P.H.; Ishiwata, K.; Nijholt, I.M.; Luiten, P.G.; Dierckx, R.A. The cholinergic system, sigma-1 receptors and cognition. Behav. Brain Res. 2011, 221, 543–554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Callens, M.; Loncke, J.; Bultynck, G. Dysregulated Ca2+ homeostasis as a central theme in neurodegeneration: Lessons from Alzheimer’s disease and wolfram syndrome. Cells 2022, 11, 1963. [Google Scholar] [CrossRef]
  121. Resende, R.; Fernandes, T.; Pereira, A.C.; Marques, A.P.; Pereira, C.F. Endoplasmic reticulum-mitochondria contacts modulate reactive oxygen species-mediated signaling and oxidative stress in brain disorders: The key role of sigma-1 receptor. Antioxid. Redox Signal. 2022, 37, 758–780. [Google Scholar] [CrossRef]
  122. Tsai, S.Y.; Hayashi, T.; Harvey, B.K.; Wang, Y.; Wu, W.W.; Shen, R.F.; Zhang, Y.; Becker, K.G.; Hoffer, B.J.; Su, T.P. Sigma-1 receptors regulate hippocampal dendritic spine formation via a free radical-sensitive mechanism involving Rac1xGTP pathway. Proc. Natl. Acad. Sci. USA 2009, 106, 22468–22473. [Google Scholar] [CrossRef]
  123. Weng, T.Y.; Tsai, S.A.; Su, T.P. Roles of sigma-1 receptors on mitochondrial functions relevant to neurodegenerative diseases. J. Biomed. Sci. 2017, 24, 74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Mori, T.; Hayashi, T.; Hayashi, E.; Su, T.P. Sigma-1 receptor chaperone at the ER-mitochondrion interface mediates the mitochondrion-ER-nucleus signaling for cellular survival. PLoS One. 2013, 8, e76941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Voronin, M.V.; Abramova, E.V.; Verbovaya, E.R.; Vakhitova, Y.V.; Seredenin, S.B. Chaperone-dependent mechanisms as a pharmacological target for neuroprotection. Int. J. Mol. Sci. 2023, 24, 823. [Google Scholar] [CrossRef] [PubMed]
  126. Li, L.; Xu, B.; Zhu, Y.; Chen, L.; Sokabe, M.; Chen, L. DHEA prevents Aβ25-35-impaired survival of newborn neurons in the dentate gyrus through a modulation of PI3K-Akt-mTOR signaling. Neuropharmacology 2010, 59, 323–333. [Google Scholar] [CrossRef]
  127. Nishimura, T.; Ishima, T.; Iyo, M.; Hashimoto, K. Potentiation of nerve growth factor-induced neurite outgrowth by fluvoxamine: Role of sigma-1 receptors, IP3 receptors and cellular signaling pathways. PLoS One. 2008, 3, e2558. [Google Scholar] [CrossRef]
  128. Lahmy, V.; Meunier, J.; Malmström, S.; Naert, G.; Givalois, L.; Kim, S.H.; Villard, V.; Vamvakides, A.; Maurice, T. Blockade of Tau hyperphosphorylation and Aβ₁₋₄₂ generation by the aminotetrahydrofuran derivative ANAVEX2-73, a mixed muscarinic and σ₁ receptor agonist, in a nontransgenic mouse model of Alzheimer’s disease. Neuropsychopharmacology 2013, 38, 1706–1723. [Google Scholar] [CrossRef]
  129. Tsai, S.Y.; Pokrass, M.J.; Klauer, N.R.; Nohara, H.; Su, T.P. Sigma-1 receptor regulates tau phosphorylation and axon extension by shaping p35 turnover via myristic acid. Proc. Natl. Acad. Sci. USA 2015, 112, 6742–6747. [Google Scholar] [CrossRef]
  130. Ciesielski, J.; Su, T.P.; Tsai, S.Y. Myristic acid hitchhiking on sigma-1 receptor to fend off neurodegeneration. Recept. Clin. Investig. 2016, 3, e1114. [Google Scholar]
  131. Sha, S.; Qu, W.J.; Li, L.; Lu, Z.H.; Chen, L.; Yu, W.F.; Chen, L. Sigma-1 receptor knockout impairs neurogenesis in dentate gyrus of adult hippocampus via down-regulation of NMDA receptors. CNS Neurosci. Ther. 2013, 19, 705–713. [Google Scholar] [CrossRef]
  132. Couly, S.; Goguadze, N.; Yasui, Y.; Kimura, Y.; Wang, S.M.; Sharikadze, N.; Wu, H.E.; Su, T.P. Knocking out sigma-1 receptors reveals diverse health problems. Cell Mol. Neurobiol. 2022, 42, 597–620. [Google Scholar] [CrossRef]
  133. Raulin, A.C.; Doss, S.V.; Trottier, Z.A.; Ikezu, T.C.; Bu, G.; Liu, C.C. ApoE in Alzheimer’s disease: Pathophysiology and therapeutic strategies. Mol. Neurodegener. 2022, 17, 72. [Google Scholar] [CrossRef] [PubMed]
  134. Jansen, K.L.; Faull, R.L.; Storey, P.; Leslie, R.A. Loss of sigma binding sites in the CA1 area of the anterior hippocampus in Alzheimer’s disease correlates with CA1 pyramidal cell loss. Brain Res. 1993, 623, 299–302. [Google Scholar] [CrossRef]
  135. Antonini, V.; Marrazzo, A.; Kleiner, G.; Coradazzi, M.; Ronsisvalle, S.; Prezzavento, O.; Ronsisvalle, G.; Leanza, G. Anti-amnesic and neuroprotective actions of the sigma-1 receptor agonist (-)-MR22 in rats with selective cholinergic lesion and amyloid infusion. J. Alzheimers Dis. 2011, 24, 569–586. [Google Scholar] [CrossRef]
  136. Antonini, V.; Prezzavento, O.; Coradazzi, M.; Marrazzo, A.; Ronsisvalle, S.; Arena, E.; Leanza, G. Anti-amnesic properties of (+/−)-PPCC, a novel sigma receptor ligand, on cognitive dysfunction induced by selective cholinergic lesion in rats. J. Neurochem. 2009, 109, 744–754. [Google Scholar] [CrossRef] [PubMed]
  137. Yang, R.; Chen, L.; Wang, H.; Xu, B.; Tomimoto, H.; Chen, L. Anti-amnesic effect of neurosteroid PREGS in Aβ25–35-injected mice through σ1 receptor- and α7nAChR-mediated neuroprotection. Neuropharmacology 2012, 63, 1042–1050. [Google Scholar] [CrossRef] [PubMed]
  138. Maurice, T.; Su, T.P.; Privat, A. Sigma11) receptor agonists and neurosteroids attenuate β25–35-amyloid peptide-induced amnesia in mice through a common mechanism. Neuroscience 1998, 83, 413–428. [Google Scholar] [CrossRef]
  139. Meunier, J.; Ieni, J.; Maurice, T. The anti-amnesic and neuroprotective effects of donepezil against amyloid β25-35 peptide-induced toxicity in mice involve an interaction with the σ1 receptor. Br. J. Pharmacol. 2006, 149, 998–1012. [Google Scholar] [CrossRef] [Green Version]
  140. Villard, V.; Espallergues, J.; Keller, E.; Vamvakides, A.; Maurice, T. Anti-amnesic and neuroprotective potentials of the mixed muscarinic receptor/sigma 1 (σ1) ligand ANAVEX2-73, a novel aminotetrahydrofuran derivative. J. Psychopharmacol. 2011, 25, 1101–1117. [Google Scholar] [CrossRef]
  141. Villard, V.; Espallergues, J.; Keller, E.; Alkam, T.; Nitta, A.; Yamada, K.; Nabeshima, T.; Vamvakides, A.; Maurice, T. Antiamnesic and neuroprotective effects of the aminotetrahydrofuran derivative ANAVEX1-41 against amyloid beta25-35-induced toxicity in mice. Neuropsychopharmacology 2009, 34, 1552–1566. [Google Scholar] [CrossRef] [Green Version]
  142. Ukai, M.; Maeda, H.; Nanya, Y.; Kameyama, T.; Matsuno, K. Beneficial effects of acute and repeated administrations of sigma receptor agonists on behavioral despair in mice exposed to tail suspension. Pharmacol. Biochem. Behav. 1998, 61, 247–252. [Google Scholar] [CrossRef]
  143. Hayashi, T.; Maurice, T.; Su, T.P. Ca2+ signaling via sigma1-receptors: Novel regulatory mechanism affecting intracellular Ca2+ concentration. J. Pharmacol. Exp. Ther. 2000, 293, 788–798. [Google Scholar] [PubMed]
  144. Maurice, T. Beneficial effect of the sigma1 receptor agonist PRE-084 against the spatial learning deficits in aged rats. Eur. J. Pharmacol. 2001, 431, 223–227. [Google Scholar] [CrossRef] [PubMed]
  145. Maurice, T.; Roman, F.J.; Su, T.P.; Privat, A. Beneficial effects of sigma agonists on the age-related learning impairment in the senescence-accelerated mouse (SAM). Brain Res. 1996, 733, 219–230. [Google Scholar] [CrossRef] [PubMed]
  146. Chen, A.F.; Ma, W.H.; Xie, X.Y.; Huang, Y.S. Sigma-2 Receptor as a potential drug target. Curr. Med. Chem. 2021, 28, 4172–4189. [Google Scholar] [CrossRef]
  147. Yang, K.; Zeng, C.; Wang, C.; Sun, M.; Yin, D.; Sun, T. Sigma-2 receptor-A potential target for cancer/Alzheimer’s disease treatment via its regulation of cholesterol homeostasis. Molecules 2020, 25, 5439. [Google Scholar] [CrossRef]
  148. Lizama, B.N.; Kahle, J.; Catalano, S.M.; Caggiano, A.O.; Grundman, M.; Hamby, M.E. Sigma-2 receptors-from basic biology to therapeutic target: A focus on age-related degenerative diseases. Int. J. Mol. Sci. 2023, 24, 6251. [Google Scholar] [CrossRef]
  149. Mach, R.H.; Zeng, C.; Hawkins, W.G. The σ2 receptor: A novel protein for the imaging and treatment of cancer. J. Med. Chem. 2013, 56, 7137–7160. [Google Scholar] [CrossRef] [Green Version]
  150. Mach, R.H.; Smith, C.R.; al-Nabulsi, I.; Whirrett, B.R.; Childers, S.R.; Wheeler, K.T. Sigma-2 receptors as potential biomarkers of proliferation in breast cancer. Cancer Res. 1997, 57, 156–161. [Google Scholar]
  151. Al-Nabulsi, I.; Mach, R.H.; Wang, L.M.; Wallen, C.A.; Keng, P.C.; Sten, K.; Childers, S.R.; Wheeler, K.T. Effect of ploidy, recruitment, environmental factors, and tamoxifen treatment on the expression of sigma-2 receptors in proliferating and quiescent tumour cells. Br. J. Cancer 1999, 81, 925–933. [Google Scholar] [CrossRef] [Green Version]
  152. Wheeler, K.T.; Wang, L.M.; Wallen, C.A.; Childers, S.R.; Cline, J.M.; Keng, P.C.; Mach, R.H. Sigma-2 receptors as a biomarker of proliferation in solid tumours. Br. J. Cancer 2000, 82, 1223–1232. [Google Scholar] [CrossRef]
  153. Shaghaghi, Z.; Alvandi, M.; Ghanbarimasir, Z.; Farzipour, S.; Emami, S. Current development of sigma-2 receptor radioligands as potential tumor imaging agents. Bioorg. Chem. 2021, 115, 105163. [Google Scholar] [CrossRef] [PubMed]
  154. Jin, J.; Arbez, N.; Sahn, J.J.; Lu, Y.; Linkens, K.T.; Hodges, T.R.; Tang, A.; Wiseman, R.; Martin, S.F.; Ross, C.A. Neuroprotective effects of σ2R/TMEM97 receptor modulators in the neuronal model of Huntington’s disease. ACS Chem. Neurosci. 2022, 13, 2852–2862. [Google Scholar] [CrossRef] [PubMed]
  155. Limegrover, C.S.; Yurko, R.; Izzo, N.J.; LaBarbera, K.M.; Rehak, C.; Look, G.; Rishton, G.; Safferstein, H.; Catalano, S.M. Sigma-2 receptor antagonists rescue neuronal dysfunction induced by Parkinson’s patient brain-derived α-synuclein. J. Neurosci. Res. 2021, 99, 1161–1176. [Google Scholar] [CrossRef] [PubMed]
  156. Sánchez, C.; Papp, M. The selective sigma2 ligand Lu 28-179 has an antidepressant-like profile in the rat chronic mild stress model of depression. Behav. Pharmacol. 2000, 11, 117–124. [Google Scholar] [CrossRef]
  157. Davidson, M.; Saoud, J.; Staner, C.; Noel, N.; Luthringer, E.; Werner, S.; Reilly, J.; Schaffhauser, J.Y.; Rabinowitz, J.; Weiser, M.; et al. Efficacy and safety of MIN-101: A 12-week randomized, double-blind, placebo-controlled trial of a new drug in development for the treatment of negative symptoms in schizophrenia. Am. J. Psychiatry 2017, 174, 1195–1202. [Google Scholar] [CrossRef] [Green Version]
  158. Sahn, J.J.; Mejia, G.L.; Ray, P.R.; Martin, S.F.; Price, T.J. Sigma-2 receptor/Tmem97 agonists produce long lasting antineuropathic pain effects in mice. ACS Chem. Neurosci. 2017, 8, 1801–1811. [Google Scholar] [CrossRef] [Green Version]
  159. Wilson, L.L.; Alleyne, A.R.; Eans, S.O.; Cirino, T.J.; Stacy, H.M.; Mottinelli, M.; Intagliata, S.; McCurdy, C.R.; McLaughlin, J.P. Characterization of CM-398, a novel selective sigma-2 receptor ligand, as a potential therapeutic for neuropathic pain. Molecules 2022, 27, 3617. [Google Scholar] [CrossRef]
  160. Izzo, N.J.; Xu, J.; Zeng, C.; Kirk, M.J.; Mozzoni, K.; Silky, C.; Rehak, C.; Yurko, R.; Look, G.; Rishton, G.; et al. Alzheimer’s therapeutics targeting amyloid beta 1-42 oligomers II: Sigma-2/PGRMC1 receptors mediate abeta 42 oligomer binding and synaptotoxicity. PLoS ONE 2014, 9, e111899. [Google Scholar] [CrossRef]
  161. Esparza, T.J.; Zhao, H.; Cirrito, J.R.; Cairns, N.J.; Bateman, R.J.; Holtzman, D.M.; Brody, D.L. Amyloid-β oligomerization in Alzheimer dementia versus high-pathology controls. Ann. Neurol. 2013, 73, 104–119. [Google Scholar] [CrossRef]
  162. Lacor, P.N.; Buniel, M.C.; Furlow, P.W.; Clemente, A.S.; Velasco, P.T.; Wood, M.; Viola, K.L.; Klein, W.L. Abeta oligomer-induced aberrations in synapse composition, shape, and density provide a molecular basis for loss of connectivity in Alzheimer’s disease. J. Neurosci. 2007, 27, 796–807. [Google Scholar] [CrossRef] [Green Version]
  163. DaRocha-Souto, B.; Scotton, T.C.; Coma, M.; Serrano-Pozo, A.; Hashimoto, T.; Serenó, L.; Rodríguez, M.; Sánchez, B.; Hyman, B.T.; Gómez-Isla, T. Brain oligomeric β-amyloid but not total amyloid plaque burden correlates with neuronal loss and astrocyte inflammatory response in amyloid precursor protein/tau transgenic mice. J. Neuropathol. Exp. Neurol. 2011, 70, 360–376. [Google Scholar] [CrossRef] [Green Version]
  164. Izzo, N.J.; Staniszewski, A.; To, L.; Fa, M.; Teich, A.F.; Saeed, F.; Wostein, H.; Walko, T., 3rd; Vaswani, A.; Wardius, M.; et al. Alzheimer’s therapeutics targeting amyloid beta 1-42 oligomers I: Abeta 42 oligomer binding to specific neuronal receptors is displaced by drug candidates that improve cognitive deficits. PLoS One. 2014, 9, e111898. [Google Scholar] [CrossRef] [PubMed]
  165. Colom-Cadena, M.; Tulloch, J.; Jackson, R.J.; Catterson, J.H.; Rose, J.; Davies, C.; Hooley, M.; Anton-Fernandez, A.; Dunnett, S.; Tempelaar, R.; et al. TMEM97 increases in synapses and is a potential synaptic Aβ binding partner in human Alzheimer’s disease. bioRxiv 2021, 2021.02.01.428238. [Google Scholar] [CrossRef]
  166. Xiong, H.; Callaghan, D.; Jones, A.; Walker, D.G.; Lue, L.F.; Beach, T.G.; Sue, L.I.; Woulfe, J.; Xu, H.; Stanimirovic, D.B.; et al. Cholesterol retention in Alzheimer’s brain is responsible for high beta- and gamma-secretase activities and Abeta production. Neurobiol. Dis. 2008, 29, 422–437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Lane-Donovan, C.; Philips, G.T.; Herz, J. More than cholesterol transporters: Lipoprotein receptors in CNS function and neurodegeneration. Neuron 2014, 83, 771–787. [Google Scholar] [CrossRef] [Green Version]
  168. Pappolla, M.A.; Bryant-Thomas, T.K.; Herbert, D.; Pacheco, J.; Fabra Garcia, M.; Manjon, M.; Girones, X.; Henry, T.L.; Matsubara, E.; Zambon, D.; et al. Mild hypercholesterolemia is an early risk factor for the development of Alzheimer amyloid pathology. Neurology 2003, 61, 199–205. [Google Scholar] [CrossRef]
  169. Nicholson, A.M.; Ferreira, A. Increased membrane cholesterol might render mature hippocampal neurons more susceptible to beta-amyloid-induced calpain activation and tau toxicity. J. Neurosci. 2009, 29, 4640–4651. [Google Scholar] [CrossRef] [Green Version]
  170. Sawamura, N.; Gong, J.S.; Chang, T.Y.; Yanagisawa, K.; Michikawa, M. Promotion of tau phosphorylation by MAP kinase Erk1/2 is accompanied by reduced cholesterol level in detergent-insoluble membrane fraction in Niemann-Pick C1-deficient cells. J. Neurochem. 2003, 84, 1086–1096. [Google Scholar] [CrossRef]
  171. Thejer, B.M.; Infantino, V.; Santarsiero, A.; Pappalardo, I.; Abatematteo, F.S.; Teakel, S.; Van Oosterum, A.; Mach, R.H.; Denora, N.; Lee, B.C.; et al. Sigma-2 receptor ligand binding modulates association between TSPO and TMEM97. Int. J. Mol. Sci. 2023, 24, 6381. [Google Scholar] [CrossRef]
  172. Steffensen, K.R.; Jakobsson, T.; Gustafsson, J. Targeting liver X receptors in inflammation. Expert Opin. Ther. Targets 2013, 17, 977–990. [Google Scholar] [CrossRef]
  173. Vilner, B.J.; Bowen, W.D. Modulation of cellular calcium by sigma-2 receptors: Release from intracellular stores in human SK-N-SH neuroblastoma cells. J. Pharmacol. Exp. Ther. 2000, 292, 900–911. [Google Scholar] [PubMed]
  174. Cassano, G.; Gasparre, G.; Contino, M.; Niso, M.; Berardi, F.; Perrone, R.; Colabufo, N.A. The sigma-2 receptor agonist PB28 inhibits calcium release from the endoplasmic reticulum of SK-N-SH neuroblastoma cells. Cell Calcium 2006, 40, 23–28. [Google Scholar] [CrossRef] [PubMed]
  175. Cassano, G.; Gasparre, G.; Niso, M.; Contino, M.; Scalera, V.; Colabufo, N.A. F281, synthetic agonist of the sigma-2 receptor, induces Ca2+ efflux from the endoplasmic reticulum and mitochondria in SK-N-SH cells. Cell Calcium 2009, 45, 340–345. [Google Scholar] [CrossRef]
  176. Yi, B.; Sahn, J.J.; Ardestani, P.M.; Evans, A.K.; Scott, L.L.; Chan, J.Z.; Iyer, S.; Crisp, A.; Zuniga, G.; Pierce, J.T.; et al. Small molecule modulator of sigma-2 receptor is neuroprotective and reduces cognitive deficits and neuroinflammation in experimental models of Alzheimer’s disease. J. Neurochem. 2017, 140, 561–575. [Google Scholar] [CrossRef] [Green Version]
  177. Guo, L.; Zhen, X. Sigma-2 receptor ligands: Neurobiological effects. Curr. Med. Chem. 2015, 22, 989–1003. [Google Scholar] [CrossRef]
  178. Terada, K.; Migita, K.; Matsushima, Y.; Sugimoto, Y.; Kamei, C.; Matsumoto, T.; Mori, M.; Matsunaga, K.; Takata, J.; Karube, Y. Cholinesterase inhibitor rivastigmine enhances nerve growth factor-induced neurite outgrowth in PC12 cells via sigma-1 and sigma-2 receptors. PLoS One 2018, 13, e0209250. [Google Scholar] [CrossRef] [Green Version]
  179. Ju, Y.; Tam, K.Y. Pathological mechanisms and therapeutic strategies for Alzheimer’s disease. Neural Regen. Res. 2022, 17, 543–549. [Google Scholar]
  180. Terada, K.; Migita, K.; Matsushima, Y.; Kamei, C. Sigma-2 receptor as a potential therapeutic target for treating central nervous system disorders. Neural Regen. Res. 2019, 14, 1893–1894. [Google Scholar] [CrossRef]
  181. Kargbo, R.B. Sigma-1 and Sigma-2 receptor modulators as potential therapeutics for Alzheimer’s disease. ACS Med. Chem. Lett. 2021, 12, 178–179. [Google Scholar] [CrossRef] [PubMed]
  182. Prasanth, M.I.; Malar, D.S.; Tencomnao, T.; Brimson, J.M. The emerging role of the sigma-1 receptor in autophagy: Hand-in-hand targets for the treatment of Alzheimer’s. Expert Opin. Ther. Targets 2021, 25, 401–414. [Google Scholar] [CrossRef] [PubMed]
  183. Cummings, J.; Lee, G.; Ritter, A.; Sabbagh, M.; Zhong, K. Alzheimer’s disease drug development pipeline: 2020. Alzheimers Dement. 2020, 6, e12050. [Google Scholar] [CrossRef] [PubMed]
  184. Cummings, J.; Lee, G.; Nahed, P.; Kambar, M.; Zhong, K.; Fonseca, J.; Taghva, K. Alzheimer’s disease drug development pipeline: 2022. Alzheimers Dement. 2022, 8, e12295. [Google Scholar] [CrossRef]
  185. Malar, D.S.; Thitilertdecha, P.; Ruckvongacheep, K.S.; Brimson, S.; Tencomnao, T.; Brimson, J.M. Targeting sigma receptors for the treatment of neurodegenerative and neurodevelopmental disorders. CNS Drugs 2023, 37, 399–440. [Google Scholar] [CrossRef]
  186. McClure, E.W.; Daniels, R.N. Classics in chemical neuroscience: Dextromethorphan (DXM). ACS Chem. Neurosci. 2023, 14, 2256–2270. [Google Scholar] [CrossRef] [PubMed]
  187. Silva, A.R.; Dinis-Oliveira, R.J. Pharmacokinetics and pharmacodynamics of dextromethorphan: Clinical and forensic aspects. Drug Metab. Rev. 2020, 52, 258–282. [Google Scholar] [CrossRef] [PubMed]
  188. Khoury, R. Deuterated dextromethorphan/quinidine for agitation in Alzheimer’s disease. Neural Regen. Res. 2022, 17, 1013–1014. [Google Scholar] [CrossRef]
  189. Hampel, H.; Williams, C.; Etcheto, A.; Goodsaid, F.; Parmentier, F.; Sallantin, J.; Kaufmann, W.E.; Missling, C.U.; Afshar, M. A precision medicine framework using artificial intelligence for the identification and confirmation of genomic biomarkers of response to an Alzheimer’s disease therapy: Analysis of the blarcamesine (ANAVEX2-73) Phase 2a clinical study. Alzheimers Dement. 2020, 6, e12013. [Google Scholar] [CrossRef]
  190. Fukushima, T. Pharmacological properties of T-817MA, a novel neurotrophic agent, for treatment of Alzheimer’s disease. Nihon Yakurigaku Zasshi 2010, 136, 11–14. [Google Scholar] [CrossRef]
  191. Yano, T.; Tanabe, H.; Kobayashi, K.; Kobayashi, H.; Nabetani, A.; Sakai, Y.; Okuda, T.; Nakagawa, M.; Nakamura, T. P4-210: Sigma-1 receptor is a molecular target for novel neuroprotectant T-817MA. Alzheimers Dement. 2015, 11, P861. [Google Scholar] [CrossRef]
  192. Fukushima, T.; Nakamura, A.; Iwakami, N.; Nakada, Y.; Hattori, H.; Hoki, S.; Yamaguchi, H.; Nakagawa, M.; Terashima, N.; Narita, H. T-817MA, a neuroprotective agent, attenuates the motor and cognitive impairments associated with neuronal degeneration in P301L tau transgenic mice. Biochem. Biophys. Res. Commun. 2011, 407, 730–734. [Google Scholar] [CrossRef]
  193. Rishton, G.M.; Look, G.C.; Ni, Z.J.; Zhang, J.; Wang, Y.; Huang, Y.; Wu, X.; Izzo, N.J.; LaBarbera, K.M.; Limegrover, C.S.; et al. Discovery of investigational drug CT1812, an antagonist of the sigma-2 receptor complex for Alzheimer’s disease. ACS Med. Chem. Lett. 2021, 12, 1389–1395. [Google Scholar] [CrossRef]
  194. Izzo, N.J.; Yuede, C.M.; LaBarbera, K.M.; Limegrover, C.S.; Rehak, C.; Yurko, R.; Waybright, L.; Look, G.; Rishton, G.; Safferstein, H.; et al. Preclinical and clinical biomarker studies of CT1812: A novel approach to Alzheimer’s disease modification. Alzheimers Dement. 2021, 17, 1365–1382. [Google Scholar] [CrossRef]
  195. Lauterbach, E.C. Dextromethorphan as a potential rapid-acting antidepressant. Med. Hypotheses. 2011, 76, 717–719. [Google Scholar] [CrossRef] [PubMed]
  196. Bruna, J.; Videla, S.; Argyriou, A.A.; Velasco, R.; Villoria, J.; Santos, C.; Nadal, C.; Cavaletti, G.; Alberti, P.; Briani, C.; et al. Efficacy of a novel sigma-1 receptor antagonist for oxaliplatin-induced neuropathy: A randomized, double-blind, placebo-controlled phase IIa clinical trial. Neurotherapeutics. 2018, 15, 178–189. [Google Scholar] [CrossRef] [PubMed]
  197. Crișan, G.; Moldovean-Cioroianu, N.S.; Timaru, D.G.; Andrieș, G.; Căinap, C.; Chiș, V. Radiopharmaceuticals for PET and SPECT imaging: A literature review over the last decade. Int. J. Mol. Sci. 2022, 23, 5023. [Google Scholar] [CrossRef]
  198. Shojaie, M.; Tabarestani, S.; Cabrerizo, M.; DeKosky, S.T.; Vaillancourt, D.E.; Loewenstein, D.; Duara, R.; Adjouadi, M. PET imaging of tau pathology and amyloid-β, and MRI for Alzheimer’s disease feature fusion and multimodal classification. J. Alzheimers Dis. 2021, 84, 1497–1514. [Google Scholar] [CrossRef] [PubMed]
  199. Hashimoto, K.; Ishiwata, K. Sigma receptor ligands: Possible application as therapeutic drugs and as radiopharmaceuticals. Curr. Pharm. Des. 2006, 12, 3857–3876. [Google Scholar] [PubMed]
  200. Agha, H.; McCurdy, C.R. In vitro and in vivo sigma 1 receptor imaging studies in different disease states. RSC Med. Chem. 2021, 12, 154–177. [Google Scholar] [CrossRef]
  201. Mach, R.H.; Wheeler, K.T. Development of molecular probes for imaging sigma-2 receptors in vitro and in vivo. Cent. Nerv. Syst. Agents Med. Chem. 2009, 9, 230–245. [Google Scholar] [CrossRef] [Green Version]
  202. Sakata, M.; Kimura, Y.; Naganawa, M.; Oda, K.; Ishii, K.; Chihara, K.; Ishiwata, K. Mapping of human cerebral sigma-1 receptors using positron emission tomography and [11C]SA4503. Neuroimage 2007, 35, 1–8. [Google Scholar] [CrossRef]
  203. Shiba, K.; Ogawa, K.; Ishiwata, K.; Yajima, K.; Mori, H. Synthesis and binding affinities of methylvesamicol analogs for the acetylcholine transporter and sigma receptor. Bioorg. Med. Chem. 2006, 14, 2620–2626. [Google Scholar] [CrossRef] [PubMed]
  204. Matsuno, K.; Nakazawa, M.; Okamoto, K.; Kawashima, Y.; Mita, S. Binding properties of SA4503, a novel and selective sigma-1 receptor agonist. Eur. J. Pharmacol. 1996, 306, 271–279. [Google Scholar] [CrossRef] [PubMed]
  205. Matsuno, K.; Mita, S. SA4503: A novel sigma₁ receptor agonist. CNS Drug Rev. 1998, 4, 1–24. [Google Scholar] [CrossRef]
  206. Hirata, M.; Mori, T.; Soga, S.; Umeda, T.; Ohmomo, Y. Synthesis and in vitro evaluation of iodinated derivatives of piperazine as a new ligand for sigma receptor imaging by single photon emission computed tomography. Chem. Pharm. Bull. 2006, 54, 470–475. [Google Scholar] [CrossRef] [Green Version]
  207. Lever, J.R.; Gustafson, J.L.; Xu, R.; Allmon, R.L.; Lever, S.Z. Sigma-1 and sigma-2 receptor binding affinity and selectivity of SA4503 and fluoroethyl SA4503. Synapse 2006, 59, 350–358. [Google Scholar] [CrossRef]
  208. Wang, X.; Li, D.; Deuther-Conrad, W.; Lu, J.; Xie, Y.; Jia, B.; Cui, M.; Steinbach, J.; Brust, P.; Liu, B.; et al. Novel cyclopentadienyl tricarbonyl 99mTc complexes containing 1-piperonylpiperazine moiety: Potential imaging probes for sigma-1 receptors. J. Med. Chem. 2014, 57, 7113–7125. [Google Scholar] [CrossRef] [PubMed]
  209. Weissman, A.D.; Su, T.P.; Hedreen, J.C.; London, E.D. Sigma receptors in post-mortem human brains. J. Pharmacol. Exp. Ther. 1988, 247, 29–33. [Google Scholar]
  210. Kornhuber, J.; Schoppmeyer, K.; Bendig, C.; Riederer, P. Characterization of [3H]pentazocine binding sites in post-mortem human frontal cortex. J. Neural. Transm. 1996, 103, 45–53. [Google Scholar] [CrossRef]
  211. Kawamura, K.; Ishiwata, K.; Shimada, Y.; Kimura, Y.; Kobayashi, T.; Matsuno, K.; Homma, Y.; Senda, M. Preclinical evaluation of [11C]SA4503: Radiation dosimetry, in vivo selectivity and PET imaging of sigma1 receptors in the cat brain. Ann. Nucl. Med. 2000, 14, 285–292. [Google Scholar] [CrossRef]
  212. Ishiwata, K.; Tsukada, H.; Kawamura, K.; Kimura, Y.; Nishiyama, S.; Kobayashi, T.; Matsuno, K.; Senda, M. Mapping of CNS sigma1 receptors in the conscious monkey: Preliminary PET study with [11C]SA4503. Synapse 2001, 40, 235–237. [Google Scholar] [CrossRef]
  213. Kawamura, K.; Kimura, Y.; Tsukada, H.; Kobayashi, T.; Nishiyama, S.; Kakiuchi, T.; Ohba, H.; Harada, N.; Matsuno, K.; Ishii, K.; et al. An increase of sigma receptors in the aged monkey brain. Neurobiol. Aging 2003, 24, 745–752. [Google Scholar] [CrossRef] [PubMed]
  214. Toyohara, J.; Kobayashi, T.; Mita, S.; Ishiwata, K. Application of [¹¹C]SA4503 to selection of novel σ₁ selective agonists. Nucl. Med. Biol. 2012, 39, 1117–1121. [Google Scholar] [CrossRef] [PubMed]
  215. Toyohara, J.; Sakata, M.; Ishiwata, K. Imaging of sigma1 receptors in the human brain using PET and [11C]SA4503. Cent. Nerv. Syst. Agents Med. Chem. 2009, 9, 190–196. [Google Scholar] [CrossRef] [PubMed]
  216. Mishina, M.; Ohyama, M.; Ishii, K.; Kitamura, S.; Kimura, Y.; Oda, K.; Kawamura, K.; Sasaki, T.; Kobayashi, S.; Katayama, Y.; et al. Low density of sigma-1 receptors in early Alzheimer’s disease. Ann. Nucl. Med. 2008, 22, 151–156. [Google Scholar] [CrossRef] [PubMed]
  217. Venkataraman, A.V.; Mansur, A.; Rizzo, G.; Bishop, C.; Lewis, Y.; Kocagoncu, E.; Lingford-Hughes, A.; Huiban, M.; Passchier, J.; Rowe, J.B.; et al. Widespread cell stress and mitochondrial dysfunction occur in patients with early Alzheimer’s disease. Sci. Transl. Med. 2022, 14, eabk1051. [Google Scholar] [CrossRef]
  218. Mishina, M.; Ishiwata, K.; Ishii, K.; Kitamura, S.; Kimura, Y.; Kawamura, K.; Oda, K.; Sasaki, T.; Sakayori, O.; Hamamoto, M.; et al. Function of sigma1 receptors in Parkinson’s disease. Acta Neurol. Scand 2005, 112, 103–107. [Google Scholar] [CrossRef]
  219. Ishikawa, M.; Ishiwata, K.; Ishii, K.; Kimura, Y.; Sakata, M.; Naganawa, M.; Oda, K.; Miyatake, R.; Fujisaki, M.; Shimizu, E.; et al. High occupancy of sigma-1 receptors in the human brain after single oral administration of fluvoxamine: A positron emission tomography study using [11C]SA4503. Biol. Psychiatry 2007, 62, 878–883. [Google Scholar] [CrossRef]
  220. Kunitachi, S.; Fujita, Y.; Ishima, T.; Kohno, M.; Horio, M.; Tanibuchi, Y.; Shirayama, Y.; Iyo, M.; Hashimoto, K. Phencyclidine-induced cognitive deficits in mice are ameliorated by subsequent subchronic administration of donepezil: Role of sigma-1 receptors. Brain Res. 2009, 1279, 189–196. [Google Scholar] [CrossRef]
  221. Shen, B.; Park, J.H.; Hjørnevik, T.; Cipriano, P.W.; Yoon, D.; Gulaka, P.K.; Holly, D.; Behera, D.; Avery, B.A.; Gambhir, S.S.; et al. Radiosynthesis and first-in-human PET/MRI evaluation with clinical-grade [18F]FTC-146. Mol. Imaging Biol. 2017, 19, 779–786. [Google Scholar] [CrossRef]
  222. Hjørnevik, T.; Cipriano, P.W.; Shen, B.; Park, J.H.; Gulaka, P.; Holley, D.; Gandhi, H.; Yoon, D.; Mittra, E.S.; Zaharchuk, G.; et al. Biodistribution and radiation dosimetry of 18F-FTC-146 in humans. J. Nucl. Med. 2017, 58, 2004–2009. [Google Scholar] [CrossRef] [Green Version]
  223. Waterhouse, R.N.; Nobler, M.S.; Zhou, Y.; Chang, R.C.; Morales, O.; Kuwabawa, H.; Kumar, A.; VanHeertum, R.L.; Wong, D.F.; Sackeim, H.A. First evaluation of the sigma-1 receptor radioligand [18F]1-3-fluoropropyl-4-((4-cyanophenoxy)-methyl) piperidine ([18F]FPS) in healthy humans. Neuroimage 2004, 22, T29–T30. [Google Scholar]
  224. Li, Y.; Wang, X.; Zhang, J.; Deuther-Conrad, W.; Xie, F.; Zhang, X.; Liu, J.; Qiao, J.; Cui, M.; Steinbach, J.; et al. Synthesis and evaluation of novel 18F-labeled spirocyclic piperidine derivatives as σ1 receptor ligands for positron emission tomography imaging. J. Med. Chem. 2013, 56, 3478–3491. [Google Scholar] [CrossRef] [PubMed]
  225. Holl, K.; Falck, E.; Köhler, J.; Schepmann, D.; Humpf, H.U.; Brust, P.; Wünsch, B. Synthesis, characterization, and metabolism studies of fluspidine enantiomers. ChemMedChem 2013, 8, 2047–2056. [Google Scholar] [CrossRef] [PubMed]
  226. Maier, C.A.; Wünsch, B. Novel spiropiperidines as highly potent and subtype selective sigma-receptor ligands. Part 1. J. Med. Chem. 2002, 45, 438–448. [Google Scholar] [CrossRef]
  227. Maier, C.A.; Wünsch, B. Novel sigma receptor ligands. Part 2. SAR of spiro[[2]benzopyran-1,4′-piperidines] and spiro[[2]benzofuran-1,4′-piperidines] with carbon substituents in position 3. J. Med. Chem. 2002, 45, 4923–4930. [Google Scholar] [CrossRef]
  228. Wiese, C.; Grosse Maestrup, E.; Schepmann, D.; Vela, J.M.; Holenz, J.; Buschmann, H.; Wünsch, B. Pharmacological and metabolic characterisation of the potent sigma1 receptor ligand 1′-benzyl-3-methoxy-3H-spiro[[2]benzofuran-1,4′-piperidine]. J. Pharm. Pharmacol. 2009, 61, 631–640. [Google Scholar] [CrossRef]
  229. Fischer, S.; Wiese, C.; Maestrup, E.G.; Hiller, A.; Deuther-Conrad, W.; Scheunemann, M.; Schepmann, D.; Steinbach, J.; Wünsch, B.; Brust, P. Molecular imaging of σ receptors: Synthesis and evaluation of the potent σ1 selective radioligand [18F]fluspidine. Eur. J. Nucl. Med. Mol. Imaging 2011, 38, 540–551. [Google Scholar] [CrossRef]
  230. Brust, P.; Deuther-Conrad, W.; Becker, G.; Patt, M.; Donat, C.K.; Stittsworth, S.; Fischer, S.; Hiller, A.; Wenzel, B.; Dukic-Stefanovic, S.; et al. Distinctive in vivo kinetics of the new σ1 receptor ligands (R)-(+)- and (S)-(-)-18F-fluspidine in porcine brain. J. Nucl. Med. 2014, 55, 1730–1736. [Google Scholar] [CrossRef] [Green Version]
  231. Kranz, M.; Sattler, B.; Wüst, N.; Deuther-Conrad, W.; Patt, M.; Meyer, P.M.; Fischer, S.; Donat, C.K.; Wünsch, B.; Hesse, S.; et al. Evaluation of the enantiomer specific biokinetics and radiation doses of [18F]fluspidine-A new tracer in clinical translation for imaging of σ₁ receptors. Molecules 2016, 21, 1164. [Google Scholar] [CrossRef] [Green Version]
  232. Chen, Y.Y.; Wang, X.; Zhang, J.M.; Deuther-Conrad, W.; Zhang, X.J.; Huang, Y.; Li, Y.; Ye, J.J.; Cui, M.C.; Steinbach, J.; et al. Synthesis and evaluation of a 18F-labeled spirocyclic piperidine derivative as promising σ1 receptor imaging agent. Bioorg. Med. Chem. 2014, 22, 5270–5278. [Google Scholar] [CrossRef]
  233. Baum, E.; Cai, Z.; Bois, F.; Holden, D.; Lin, S.F.; Lara-Jaime, T.; Kapinos, M.; Chen, Y.; Deuther-Conrad, W.; Fischer, S.; et al. PET imaging evaluation of four σ1 radiotracers in nonhuman primates. J. Nucl. Med. 2017, 58, 982–988. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Meyer, P.; Strauss, M.; Becker, G.; Hesse, S.; Bednasch, K.; Ettrich, B.; Zientek, F.; Rullmann, M.; Wilke, S.; Luthardt, J.; et al. IIncreased sigma-1 receptor (Sig-1R) binding in the brain of unmedicated patients with acute major depressive disorder (MDD) using the novel Sig-1R-specific radioligand (-)-[18F]Fluspidine and PET. J. Nucl. Med. 2018, 59, 551. [Google Scholar]
  235. Grachev, I.D.; Meyer, P.M.; Becker, G.A.; Bronzel, M.; Marsteller, D.; Pastino, G.; Voges, O.; Rabinovich, L.; Knebel, H.; Zientek, F.; et al. Sigma-1 and dopamine D2/D3 receptor occupancy of pridopidine in healthy volunteers and patients with Huntington disease: A [18F] fluspidine and [18F] fallypride PET study. Eur. J. Nucl. Med. Mol. Imaging 2021, 48, 1103–1115. [Google Scholar] [CrossRef] [PubMed]
  236. Glennon, R.A. Pharmacophore identification for sigma-1 (σ1) receptor binding: Application of the “deconstruction-reconstruction-elaboration” approach. Mini Rev. Med. Chem. 2005, 5, 927–940. [Google Scholar] [CrossRef]
  237. Wunsch, B. Pharmacophore models and development of spirocyclic ligands for σ1 receptors. Curr. Pharm. Des. 2012, 18, 930–937. [Google Scholar] [CrossRef]
  238. Xie, F.; Bergmann, R.; Kniess, T.; Deuther-Conrad, W.; Mamat, C.; Neuber, C.; Liu, B.; Steinbach, J.; Brust, P.; Pietzsch, J.; et al. 18F-Labeled 1,4-dioxa-8-azaspiro [4.5]decane derivative: Synthesis and biological evaluation of a σ1 receptor radioligand with low lipophilicity as potent tumor imaging agent. J. Med. Chem. 2015, 58, 5395–5407. [Google Scholar] [CrossRef]
  239. Tian, J.; He, Y.; Deuther-Conrad, W.; Fu, H.; Xie, F.; Zhang, Y.; Wang, T.; Zhang, X.; Zhang, J.; Brust, P.; et al. Synthesis and evaluation of new 1-oxa-8-azaspiro [4.5]decane derivatives as candidate radioligands for sigma-1 receptors. Bioorg. Med. Chem. 2020, 28, 115560. [Google Scholar] [CrossRef]
  240. He, Y.; Xie, F.; Ye, J.; Deuther-Conrad, W.; Cui, B.; Wang, L.; Lu, J.; Steinbach, J.; Brust, P.; Huang, Y.; et al. 1-(4-[18F]Fluorobenzyl)-4-[(tetrahydrofuran-2-yl)methyl]piperazine: A novel suitable radioligand with low lipophilicity for imaging σ1 receptors in the brain. J. Med. Chem. 2017, 60, 4161–4172. [Google Scholar] [CrossRef]
  241. Jia, H.; Cai, Z.; Holden, D.; He, Y.; Lin, S.F.; Li, S.; Baum, E.; Shirali, A.; Kapinos, M.; Gao, H.; et al. Positron emission tomography imaging evaluation of a novel 18F-labeled sigma-1 receptor radioligand in cynomolgus monkeys. ACS Chem. Neurosci. 2020, 11, 1673–1681. [Google Scholar] [CrossRef]
  242. Wang, T.; Zhang, Y.; Zhang, X.; Chen, L.; Zheng, M.; Zhang, J.; Brust, P.; Deuther-Conrad, W.; Huang, Y.; Jia, H. Synthesis and characterization of the two enantiomers of a chiral sigma-1 receptor radioligand:(S)-(+)-and (R)-(-)-[18F]FBFP. Chin. Chem. Lett. 2022, 33, 3543–3548. [Google Scholar] [CrossRef]
  243. Zheng, M.; Holden, D.; Alluri, S.R.; Wang, T.; Felchner, Z.; Gao, H.; Zhang, L.; Labaree, D.; Ropchan, J.; Carson, R.; et al. Assessing the chiral selectivity of a sigma-1 receptor radiotracer: A PET imaging study of (R)- and (S)-18F-FBFP in non-human primates. J. Nucl. Med. 2022, 63 (Suppl. 2), 2324. [Google Scholar]
  244. Tu, Z.; Xu, J.; Jones, L.A.; Li, S.; Dumstorff, C.; Vangveravong, S.; Chen, D.L.; Wheeler, K.T.; Welch, M.J.; Mach, R.H. Fluorine-18-labeled benzamide analogues for imaging the σ2 receptor status of solid tumors with positron emission tomography. J. Med. Chem. 2007, 50, 3194–3204. [Google Scholar] [CrossRef]
  245. Dehdashti, F.; Laforest, R.; Gao, F.; Shoghi, K.I.; Aft, R.L.; Nussenbaum, B.; Kreisel, F.H.; Bartlett, N.L.; Cashen, A.; Wagner-Johnston, N.; et al. Assessment of cellular proliferation in tumors by PET using 18F-ISO-1. J. Nucl. Med. 2013, 54, 350–357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Grundman, M.; Morgan, R.; Lickliter, J.D.; Schneider, L.S.; DeKosky, S.; Izzo, N.J.; Guttendorf, R.; Higgin, M.; Pribyl, J.; Mozzoni, K.; et al. A phase 1 clinical trial of the sigma-2 receptor complex allosteric antagonist CT1812, a novel therapeutic candidate for Alzheimer’s disease. Alzheimers Dement. 2019, 5, 20–26. [Google Scholar] [CrossRef] [PubMed]
  247. Moldovan, R.P.; Gündel, D.; Teodoro, R.; Ludwig, F.A.; Fischer, S.; Toussaint, M.; Schepmann, D.; Wünsch, B.; Brust, P.; Deuther-Conrad, W. Design, radiosynthesis and preliminary biological evaluation in mice of a brain-penetrant 18F-labelled σ2 receptor ligand. Int. J. Mol. Sci. 2021, 22, 5447. [Google Scholar] [CrossRef]
  248. Zhang, Y.; Wang, T.; Zhang, X.; Deuther-Conrad, W.; Fu, H.; Cui, M.; Zhang, J.; Brust, P.; Huang, Y.; Jia, H. Discovery and development of brain-penetrant 18F-labeled radioligands for neuroimaging of the sigma-2 receptors. Acta Pharm. Sin. B 2022, 12, 1406–1415. [Google Scholar] [CrossRef]
  249. Kim, H.Y.; Lee, J.Y.; Hsieh, C.-J.; Riad, A.; Izzo, N.J.; Catalano, S.M.; Graham, T.J.A.; Mach, R.H. Screening of σ2 receptor ligands and in vivo evaluation of 11C-labeled 6,7-dimethoxy-2-[4-(4-methoxyphenyl)butan-2-yl]-1,2,3,4-tetrahydroisoquinoline for potential use as a σ2 receptor brain PET tracer. J. Med. Chem. 2022, 65, 6261–6272. [Google Scholar] [CrossRef]
  250. Wang, L.; Ye, J.; He, Y.; Deuther-Conrad, W.; Zhang, J.; Zhang, X.; Cui, M.; Steinbach, J.; Huang, Y.; Brust, P.; et al. 18F-Labeled indole-based analogs as highly selective radioligands for imaging sigma-2 receptors in the brain. Bioorg. Med. Chem. 2017, 25, 3792–3802. [Google Scholar] [CrossRef]
  251. Xie, F.; Kniess, T.; Neuber, C.; Deuther-Conrad, W.; Mamat, C.; Lieberman, B.P.; Liu, B.; Mach, R.H.; Brust, P.; Steinbach, J.; et al. Novel indole-based sigma-2 receptor ligands: Synthesis, structure–affinity relationship and antiproliferative activity. MedChemComm 2015, 6, 1093–1103. [Google Scholar] [CrossRef] [Green Version]
  252. Lee, I.; Lieberman, B.P.; Li, S.; Hou, C.; Makvandi, M.; Mach, R.H. Comparative evaluation of 4 and 6-carbon spacer conformationally flexible tetrahydroisoquinolinyl benzamide analogues for imaging the sigma-2 receptor status of solid tumors. Nucl. Med. Biol. 2016, 43, 721–731. [Google Scholar] [CrossRef] [Green Version]
  253. Alluri, S.R.; Zheng, M.; Holden, D.; Zhang, Y.; Li, S.; Felchner, Z.; Zhang, L.; Ropchan, J.; Carson, R.; Jia, H.; et al. Quantitative evaluation of a novel brain-penetrant sigma-2 receptor radioligand in non-human primates. J. Nucl. Med. 2022, 63 (Suppl. 2), 2845. [Google Scholar]
  254. Alluri, S.R.; Zheng, M.-Q.; Holden, D.; Zhang, Y.; Li, S.; Felchner, Z.; Kapinos, M.; Ropchan, J.; Carson, R.E.; Jia, H.; et al. Imaging brain sigma-2 receptor: Evaluation of 18F-radiotracers in nonhuman primates. In Proceedings of the Annual Meeting of the Society of Nuclear Medicine and Molecular Imaging, SNMMI, Chicago, IL, USA, 24–27 June 2023; Available online: https://s3.amazonaws.com/amz.xcdsystem.com/706224A1-90F0-EFDE-9D12FA836B3BDEDF_abstract_File1523/PresentationPoster_P655_0613015515.pdf (accessed on 21 July 2023).
  255. Klyucherev, T.O.; Olszewski, P.; Shalimova, A.A.; Chubarev, V.N.; Tarasov, V.V.; Attwood, M.M.; Syvänen, S.; Schiöth, H.B. Advances in the development of new biomarkers for Alzheimer’s disease. Transl. Neurodegener. 2022, 11, 25. [Google Scholar] [CrossRef] [PubMed]
  256. Eysert, F.; Kinoshita, P.F.; Mary, A.; Vaillant-Beuchot, L.; Checler, F.; Chami, M. Molecular dysfunctions of mitochondria-associated membranes (MAMs) in Alzheimer’s disease. Int. J. Mol. Sci. 2020, 21, 9521. [Google Scholar] [CrossRef] [PubMed]
  257. Li, Z.; Cao, Y.; Pei, H.; Ma, L.; Yang, Y.; Li, H. The contribution of mitochondria-associated endoplasmic reticulum membranes (MAMs) dysfunction in Alzheimer’s disease and the potential countermeasure. Front. Neurosci. 2023, 17, 1158204. [Google Scholar] [CrossRef] [PubMed]
  258. Yu, W.; Jin, H.; Huang, Y. Mitochondria-associated membranes (MAMs): A potential therapeutic target for treating Alzheimer’s disease. Clin. Sci. 2021, 135, 109–126. [Google Scholar] [CrossRef]
  259. Delprat, B.; Crouzier, L.; Su, T.P.; Maurice, T. At the crossing of ER stress and MAMs: A key role of sigma-1 receptor? Adv. Exp. Med. Biol. 2020, 1131, 699–718. [Google Scholar]
  260. Maurice, T.; Meunier, J.; Feng, B.; Ieni, J.; Monaghan, D.T. Interaction with sigma1 protein, but not N-methyl-D-aspartate receptor, is involved in the pharmacological activity of donepezil. J. Pharmacol. Exp. Ther. 2006, 317, 606–614. [Google Scholar] [CrossRef]
Figure 1. Chemical structures of representative σ1 and σ2 ligands with therapeutic potential in clinical trials.
Figure 1. Chemical structures of representative σ1 and σ2 ligands with therapeutic potential in clinical trials.
Ijms 24 12025 g001
Figure 2. Chemical structures of σ1 receptor radioligands investigated in primates (* chiral center of the compound).
Figure 2. Chemical structures of σ1 receptor radioligands investigated in primates (* chiral center of the compound).
Ijms 24 12025 g002
Figure 3. Chemical structures of potential radioligands for neuroimaging of σ2 receptors.
Figure 3. Chemical structures of potential radioligands for neuroimaging of σ2 receptors.
Ijms 24 12025 g003
Table 1. Drugs available on the market for AD treatment.
Table 1. Drugs available on the market for AD treatment.
AgentTarget aMechanism bChemical Structure
DonepezilAChEAChE reversible inhibitionIjms 24 12025 i001
GalantamineAChEAChE reversible inhibitionIjms 24 12025 i002
RivastigmineAChEAChE reversible inhibitionIjms 24 12025 i003
MemantineNMDA receptorsNMDA non-competitive antagonistIjms 24 12025 i004
Aducanumab (Aduhelm) cAβ plaquemAb immunotherapy against Aβ-
Lecanemab (Leqembi) dProtofibrillar and oligomeric forms of Aβ plaquemAb immunotherapy against Aβ-
GV-971 eGut microbiotayet to be fully elucidatedmarine-derived oligosaccharide
a AChE: acetylcholinesterase, NMDA: N-methyl-D-aspartate receptor, Aβ: β-amyloid. b mAb: monoclonal antibody. c Approved by the US Food and Drug Administration (FDA) for AD treatment on 7 June 2021 via the accelerated approval pathway. d Approved by FDA for AD treatment on 6 January 2023 via the accelerated approval pathway, and converted to traditional approval on 6 July 2023. e Approved by the National Medical Products Administration of China for AD treatment on 29 December 2019.
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

Wang, T.; Jia, H. The Sigma Receptors in Alzheimer’s Disease: New Potential Targets for Diagnosis and Therapy. Int. J. Mol. Sci. 2023, 24, 12025. https://doi.org/10.3390/ijms241512025

AMA Style

Wang T, Jia H. The Sigma Receptors in Alzheimer’s Disease: New Potential Targets for Diagnosis and Therapy. International Journal of Molecular Sciences. 2023; 24(15):12025. https://doi.org/10.3390/ijms241512025

Chicago/Turabian Style

Wang, Tao, and Hongmei Jia. 2023. "The Sigma Receptors in Alzheimer’s Disease: New Potential Targets for Diagnosis and Therapy" International Journal of Molecular Sciences 24, no. 15: 12025. https://doi.org/10.3390/ijms241512025

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